To see the other types of publications on this topic, follow the link: Nitrosyl group.

Journal articles on the topic 'Nitrosyl group'

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 'Nitrosyl group.'

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

Kostin, Gennadiy A., Ruslan Kozlov, Artem Bogomyakov, Svyatoslav Tolstikov, Dmitriy Sheven, and Sergey Korenev. "New Ruthenium Nitrosyl Complexes Combining Potentially Photoactive Nitrosyl Group with the Magnetic Nitroxide Radicals as Ligands." International Journal of Molecular Sciences 24, no. 17 (2023): 13371. http://dx.doi.org/10.3390/ijms241713371.

Full text
Abstract:
Two ruthenium nitrosyl complexes of Na[RuNOCl4L] with nitronyl nitroxide radicals coordinated to ruthenium with N-donor pyridine rings were prepared and described. The crystal structure of both complexes is 1D or 2D polymeric, due to the additional coordination of sodium cation by bridging the chloride ligands or oxygen atoms of nitroxides. Partially, the oligomeric forms remain in the solutions of the complexes in acetonitrile. The magnetic measurements in the solid state demonstrate the presence of antiferromagnetic interactions through the exchange channels, with the distance between paramagnetic centers equal to 3.1–3.9 Å. The electrochemical behavior of the prepared complexes was investigated in acetonitrile solutions.
APA, Harvard, Vancouver, ISO, and other styles
2

Li, Guoliang, Limei Wen, and R. Bruce King. "Heterobimetallic Chromium Manganese Carbonyl Nitrosyls: Comparison with Isoelectronic Homometallic Binuclear Chromium Carbonyl Nitrosyls and Manganese Carbonyls." Inorganics 7, no. 10 (2019): 127. http://dx.doi.org/10.3390/inorganics7100127.

Full text
Abstract:
The heterometallic chromium-manganese carbonyl nitrosyls CrMn(NO)(CO)n (n = 9, 8) have been investigated by density functional theory. The lowest energy CrMn(NO)(CO)9 structures have unbridged staggered conformations with a ~2.99 Å Cr–Mn single bond similar to the experimental and lowest energy structures of the isoelectronic Mn2(CO)10 and Cr2(NO)2(CO)8. A significantly higher energy CrMn(NO)(CO)9 isomer has a nearly symmetrical bridging nitrosyl group and a very weakly semibridging carbonyl group. The two lowest energy structures of the unsaturated CrMn(NO)(CO)8 have a five-electron donor bridging η2-µ-NO nitrosyl group or a four-electron donor bridging η2-µ-CO group, as well as a Cr–Mn single bond of length ~2.94 Å. The next higher energy CrMn(NO)(CO)8 structure has exclusively terminal CO and NO ligands and a shorter Cr–Mn single bond of ~2.85 Å, suggesting an 18-electron configuration for the manganese atom and a 16-electron configuration for the chromium atom indicated by a vacant coordination site nearly perpendicular to the Cr–Mn bond.
APA, Harvard, Vancouver, ISO, and other styles
3

Chin, Teen Teen, Peter Legzdins, James Trotter, and Vivien C. Yee. "Organometallic nitrosyl chemistry. 51. New organometallic nitrosyl cations containing the Group 6 elements." Organometallics 11, no. 2 (1992): 913–22. http://dx.doi.org/10.1021/om00038a062.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Kia, Reza. "Non-covalent sulfoxide⋯(nitrosyl group) interactions involving coordinated nitrosyl in a Ru(ii) nitrosyl complex with an α-diimine ligand: structural and computational studies". CrystEngComm 22, № 44 (2020): 7532–37. http://dx.doi.org/10.1039/d0ce01063d.

Full text
Abstract:
Inspection of the X-ray structure of the newly prepared Ru–nitrosyl complex bearing an α-diimine ligand revealed for the first time the π-hole interaction involving the coordinated nitrosyl group with DMSO.
APA, Harvard, Vancouver, ISO, and other styles
5

Deb, Alok K., and Sreebrata Goswami. "Oxidation reactions by the coordinated nitrosyl group." Polyhedron 12, no. 11 (1993): 1419–21. http://dx.doi.org/10.1016/s0277-5387(00)84337-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Ambach, Eberhard, та Wolfgang Beck. "Metallkomplexe mit biologisch wichtigen Liganden XXVI : Reaktionen von Bis(α-aminoacidato)platin(II)-Komplexen mit Nitrosylsalzen: Nitrosylplatin-Komplexe[xxx] + X- (X = BF4 , PF6 , SbF6 )/Metal Complexes with Biologically Important Ligands XXXVI : Reaction of Bis(α-aminoacidato) Platinum(II) Complexes with Nitrosyl Salts: Nitrosyl Platinum Complexes[xxx]+X-(X = BF4 , PF6 , SbF 6)". Zeitschrift für Naturforschung B 40, № 2 (1985): 288–91. http://dx.doi.org/10.1515/znb-1985-0224.

Full text
Abstract:
AbstractThe reactions of the glycinato, alaninato, or cyclo-leucinato chelate platinum complexes trans -[xxx] with nitrosyl salts NO+X- (X = BF4, PF6, SbF6) in acetonitrile at -20 °C give the blue nitrosyl compounds {(0N)Pt [xxx] X·nCH3CN. Nitrosation at the amino group is not observed. The spectroscopic data of the nitrosyl complexes (IR, XPE) are reported. [xxx] + PF6- reacts with lithium halides or halogens in DMF to give the platinum (IV) complexes X2Pt(NH2CH2CO2)2 (X = CI, Br, I).
APA, Harvard, Vancouver, ISO, and other styles
7

Legzdins, Peter, and Michelle A. Young. "Nitrosyl N–O Bond Cleavage During Reactions of Organometallic Nitrosyl Complexes of the Group 6 Elements." Comments on Inorganic Chemistry 17, no. 4 (1995): 239–54. http://dx.doi.org/10.1080/02603599508033859.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Philipp, Gunnar, Sigrid Wocadlo, Werner Massa, et al. "Die Kristallstrukturen von [MoCl2(NO)2(OPEt3)]2, [MoCl3(NO)(OPPh3)2•MoCl4(OPPh3)2], [MoCl2(NO)(PPh3)2(CH3CN)] und [MoCl4(NPPh3)(OPPh3)] / The Crystal Structures of [MoCl2(NO)2(OPEt3)]2, [MoCl3(NO)(OPPh3)2 • MoCl4(OPPh3)2], [MoCl2(NO)(PPh3)2(CH3CN)], and [MoCl4(NPPh3)(OPPh3)]." Zeitschrift für Naturforschung B 50, no. 1 (1995): 1–10. http://dx.doi.org/10.1515/znb-1995-0102.

Full text
Abstract:
The title compounds have been prepared by reactions of MoCl2(NO)2 with PPh3, OPPh3, or Me3SiNPR3 (R = Et, Ph) in dichloromethane and acetonitrile suspension, respectively. All complexes were characterized by IR spectroscopy and by crystal structure determinations. [MoCl2(NO)2(OPEt3)]2: Space group C2/c, Z = 4, 2950 observed unique reflections, R = 0.029. Lattice dimensions at —70 °C: a = 2390.5(11), b = 875.5(5), c = 1399.1(6) pm, β = 113.18(2)°. The complex forms a centrosymmetric dimer with MoCl2Mo bridges, the nitrosyl groups being in a cis-arrangement. The OPEt3 ligand is coordinated in trans position to one of the nitrosyl ligands.[MoCl3(NO)(OPPh3)2 • MoCl4(OPPh3)2]: Space group P21/c, Z = 4, 10243 observed unique reflections, R = 0.060. Lattice dimensions at —60 °C: a = 1900.4(3), b = 1689.1(5), c = 2209.3(7) pm, β = 95.92(2)°. The structure consists of the two independent complexes [MoCl3(NO)(OPPh3)2] and [MoCl4(OPPh3)2]. In both complexes the OPPh3 groups are in a cis-arrangement at the octahedrally coordinated Mo atoms; in the nitrosyl complex one of the OPPh3 molecules is in trans-position to the nitrosyl ligand.[MoCl2(NO)(PPh3)2(CH3CN)]: Space group P21/n, Z = 4, 5107 observed unique reflections, R = 0.028. Lattice dimensions at 20 °C: a = 1006.5(2), b = 1527.2(2), c = 2342.3(2) pm, β = 90.97(1)°. The PPh3 molecules are in trans-positions to one another at the octahedrally coordinated Mo atom, whereas the acetonitrile molecule is in trans-position to the nitrosyl ligand.[MoCl4(NPPh3)(OPPhi)]: Space group P21, Z = 2, 3323 observed unique reflections, R = 0.057. Lattice dimensions at —70 °C: a = 985.7(8), b = 1471.2(9), c = 1215.9(11) pm, β = 100.50(3)°. The OPPh3 molecule coordinates in trans-position to the phosphorane iminato ligand at the octahedrally coordinated Mo atom.
APA, Harvard, Vancouver, ISO, and other styles
9

Chen, Shengchun, Xuejun Feng, Yaoming Xie, R. Bruce King, and Henry F. Schaefer. "Trinuclear and Tetranuclear Ruthenium Carbonyl Nitrosyls: Oxidation of a Carbonyl Ligand by an Adjacent Nitrosyl Ligand." Molecules 29, no. 17 (2024): 4165. http://dx.doi.org/10.3390/molecules29174165.

Full text
Abstract:
Trinuclear and tetranuclear ruthenium carbonyls of the types Ru3(CO)n(NO)2, Ru3(N)(CO)n(NO), Ru3(N)2(CO)n, Ru3(N)(CO)n(NCO), Ru3(CO)n(NCO)(NO), Ru4(N)(CO)n(NO), Ru4(N)(CO)n(NCO), and Ru4(N)2(CO)n related to species observed experimentally in the chemistry of Ru3(CO)10(µ-NO)2 have been investigated using density functional theory. In all cases, the experimentally observed structures have been found to be low-energy structures. The low-energy trinuclear structures typically have a central strongly bent Ru–Ru–Ru chain with terminal CO groups and bridging nitrosyl, isocyanate, and/or nitride ligands across the end of the chain. The low-energy tetranuclear structures typically have a central Ru4N unit with terminal CO groups and a non-bonded pair of ruthenium atoms bridged by a nitrosyl or isocyanate group.
APA, Harvard, Vancouver, ISO, and other styles
10

Kundel, P., and H. Berke. "Chemische Effekte der Nitrosylsubstitution in hexakoordinierten Carbonylwolfram-Verbindungen/ Chemical Effects of Nitrosyl Substitution in Hexacoordinated Carbonyl Tungsten Complexes." Zeitschrift für Naturforschung B 42, no. 8 (1987): 993–99. http://dx.doi.org/10.1515/znb-1987-0811.

Full text
Abstract:
AbstractThe reaction of mer-tricarbonyl-nitrosyl-bis(triisopropylphosphite)tungsten hexafluorophosphate (1a) with MeMgl or Bu4NI affords tricarbonyl-iodonitrosyl-(triisopropylphosphite)tungsten (2). Similarly, a carbonyl group of la can be replaced by a fluoride ion resulting in transdicarbonyl- fluoronitrosyl-bis(triisopropylphosphite)tungsten (3). Oxygen nucleophiles like OPh- , OCH3- or OH- either substitute or attack a carbonyl group in la and cause formation of dicarbonyl-nitrosyl-bis(triisopropylphosphite)tungsten phenoxide. methoxide or formate derivatives (4, 5 or 7). respectively. Nitrite ions react with la under CO substitution and bind as nitro ligands to the tungsten center. Addition of equivalent amounts of MeLi, PhLi and NaBH4 to la leads to the formation of tricarbonyl-nitrosyl-(triisopropylphosphite)tungsten methyl and hydrido compounds (8a, 11a) as well as to dicarbonylnitrosyl-bis(triisopropylphosphite)tungsten methyl, phenyl and hydrido complexes (8b, 10 and 11b ) . In the presence of an excess of MeLi., la is transformed to a β-diketonate derivative 9.
APA, Harvard, Vancouver, ISO, and other styles
11

Steinberg. "Red Meat-Derived Nitroso Compounds, Lipid Peroxidation Products and Colorectal Cancer." Foods 8, no. 7 (2019): 252. http://dx.doi.org/10.3390/foods8070252.

Full text
Abstract:
About 20 years ago, the research group of Sheila Anne Bingham in Cambridge, UK, showed for the first time that volunteers consuming large amounts of red meat excrete high amounts of nitroso compounds via feces. In the meantime, it has been demonstrated that heme leads to the enhanced formation of nitroso compounds in the gastrointestinal tract and that the main nitroso compounds formed in the gastrointestinal tract are S-nitrosothiols and the nitrosyl heme. Moreover, it has been postulated that these endogenously formed nitroso compounds may alkylate guanine at the O6-position, resulting in the formation of the promutagenic DNA lesions O6-methylguanine and O6-carboxymethylguanine, which, if not repaired (in time), could lead to gene mutations and, subsequently to the development of colorectal cancer. Alternatively, it has been postulated that heme iron could contribute to colorectal carcinogenesis by inducing lipid peroxidation. In the present review, the evidence supporting the above-mentioned hypotheses will be presented.
APA, Harvard, Vancouver, ISO, and other styles
12

Arashiba, Kazuya, Hidetaka Iizuka, Shoji Matsukawa, et al. "Synthesis, Structures, and Properties of Group 9− and Group 10−Group 6 Heterodinuclear Nitrosyl Complexes." Inorganic Chemistry 47, no. 10 (2008): 4264–74. http://dx.doi.org/10.1021/ic702309h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

LEGZDINS, P., and M. A. YOUNG. "ChemInform Abstract: Nitrosyl N-O Bond Cleavage During Reactions of Organometallic Nitrosyl Complexes of the Group 6 Elements." ChemInform 27, no. 17 (2010): no. http://dx.doi.org/10.1002/chin.199617269.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Lampeka, R. D., A. I. Zubenko, and V. V. Skopenko. "Type of coordination of ?-acid ligands containing a nitrosyl group." Theoretical and Experimental Chemistry 20, no. 5 (1985): 488–94. http://dx.doi.org/10.1007/bf00522439.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Hockley, Rose, Hira Irshad, Tippu S. Sheriff, Majid Motevalli та Sarantos Marinakis. "Crystal structure of bromidonitrosylbis(triphenylphosphane-κP)nickel(II)". Acta Crystallographica Section E Crystallographic Communications 71, № 4 (2015): m87—m88. http://dx.doi.org/10.1107/s2056989015004703.

Full text
Abstract:
The asymmetric unit of the title complex, [NiBr(NO){P(C6H5)3}2], comprises two independent molecules each with a similar configuration. The NiIIcation is coordinated by one bromide anion, one nitrosyl anion and two triphenylphosphane molecules in a distorted BrNP2tetrahedral coordination geometry. The coordination of the nitrosyl group is non-linear, the Ni—N—O angles being 150.2 (5) and 151.2 (5)° in the two independent molecules. In the crystal, molecules are linked by weak C—H...Br hydrogen bonds and weak C—H...π interactions into a three-dimensional supramolecular architecture.
APA, Harvard, Vancouver, ISO, and other styles
16

Baillie, Rhett A., and Peter Legzdins. "The rich and varied chemistry of group 6 cyclopentadienyl nitrosyl complexes." Coordination Chemistry Reviews 309 (February 2016): 1–20. http://dx.doi.org/10.1016/j.ccr.2015.11.003.

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

Kurtikyan, Tigran S., Hayk A. Harutyunyan, Robert K. Ghazaryan, and John A. Goodwin. "Spectral study of the nitrogen monoxide interaction with sublimed layers of meso-mono-4-pyridyl-tri-phenyl- and meso-mono-3-pyridyl-tri-phenyl-porphyrinatocobalt(II)." Journal of Porphyrins and Phthalocyanines 10, no. 07 (2006): 971–77. http://dx.doi.org/10.1142/s1088424606000302.

Full text
Abstract:
The low-temperature interaction of NO(15NO ) with sublimed layers of meso-mono-4-pyridyl-tri-phenyl- and meso-mono-3-pyridyl-tri-phenyl-porphyrinatocobalt(II) ( CoM4PyTPP (I) and CoM3PyTPP (II), respectively) has been investigated by means of FTIR and UV-visible spectroscopy. In addition to the stable five-coordinate nitrosyl complexes that are similar to the closely-related meso-tetraphenylporphyrinatocobalt(II)-nitrosyl Co(TPP)(NO) complex, a new type of complex with coordinated NO (15 NO ) has been found for the layers that were maintained at room temperature overnight before addition of nitric oxide at low temperature. The ν{ NO (15 NO )} in this species are more than 20 cm−1 lower than in five-coordinate compounds. These adducts are assigned to six-coordinate nitrosyl complexes, in which the fifth coordination site is occupied by the pyridyl group of the adjacent I (II) molecules. Warming the layers containing six-coordinate nitrosyl complexes of I almost completely transforms them to stable five-coordinate nitrosyl species indicating oligomers' disruption rather than loss of nitric oxide. In the case of II, however, a noticeable fraction of the six-coordinate species is left upon warming. Introducing new portions of NO to these layers at low temperature leads to formation of additional quantities of the six-coordinate species. Hence, part of six-coordinate complexes in II decomposed upon warming by releasing NO instead of by breaking Co -pyridyl bonds, therefore leaving free the sixth coordination sites in these layers. This result suggests a good possibility for creating solid state NO storage devices in which nitrogen monoxide can be kept and easily released by warming of the system.
APA, Harvard, Vancouver, ISO, and other styles
18

Smékal, Zdenek, Zdenek Trávnícek, Jaromír Marek, and Milan Nádvornik. "Cyano-Bridged Bimetallic Complexes of Copper(II) with Nitroprusside. Crystal Structure of [Cu(H2NCH2CH(NH2)CH3)2Fe(CN)5NO] . H2O." Australian Journal of Chemistry 53, no. 3 (2000): 225. http://dx.doi.org/10.1071/ch99131.

Full text
Abstract:
Five new complexes of compositions [Cu(1,2-pn)2Fe(CN)5NO]·H2O (1,2-pn = propane-1,2-diamine) and [Cu(L)Fe(CN)5NO]·xH2O (L = tmen (N,N,N′,N′-tetramethylethane-1,2-diamine), x = 0.5; L = trimeen (N,N,N′-trimethylethane-1,2-diamine), x = 1; L = dien (N-(2-aminoethyl)ethane-1,2-diamine), x = 0; L = medpt (N-(3-aminopropyl)-N-methylpropane-1,3-diamine), x = 2) have been isolated from the reaction mixture of Cu(ClO4)2·6H2O (or CuCl2·2H2O), the amine and Na2 [Fe(CN)5NO]·2H2O in water. The complexes have been characterized by infrared and ultraviolet–visible spectroscopies, and magnetic measurements. Single-crystal X-ray structural analysis revealed that the [Cu(1,2-pn)2Fe(CN)5NO]·H2O complex assumes a cyanide-bridged binuclear structure in which iron(II) is six-coordinated by five cyanide ligands and one nitrosyl group (the nitrosyl group lies cis to the bridging cyanide group), while copper(II) is five-coordinated by two propane-1,2-diamine ligands and a bridging cyanide ligand in a distorted tetragonal pyramidal arrangement.
APA, Harvard, Vancouver, ISO, and other styles
19

Herberhold, Max, Guo-Xin Jin, Walter Kremnitz, Arnold L. Rheingold та Brian S. Haggerty. "Halfsandwich cyclo-Pentasulfido and cyclo-Pentaselenido Complexes, (η5-C5Me5)M(NO)(E5) (E = S, Se; M = Cr, Mo, W)". Zeitschrift für Naturforschung B 46, № 4 (1991): 500–506. http://dx.doi.org/10.1515/znb-1991-0413.

Full text
Abstract:
The reactions of the pentamethylcyclopentadienyl halfsandwich nitrosyl complexes Cp*Cr(NO)2I and Cp*M(NO)I2 (M = Mo, W) with either methanolic ammonium polysulfide solutions or hydrogen selenide solutions (generated by hydrolysis of Al2Se3) can be used to prepare cyc/o-pentachalcogenido compounds such as Cp*M(NO)(S5) (M = Cr (la), Mo (2a), W (3a)) and Cp*M(NO)(Se5) (M = Cr (lb), W (3b)). The chromium complex Cp*Cr(NO)(S5) (la) is also formed in high yield by photodecarbonylation of Cp*Cr(CO)2NO in acetonitrile solution in the presence of excess sulfur. An X-ray structure analysis of Cp*W(NO)(S5) (3 a) revealed a cyclo-S5 chelate ligand in the chair conformation and an almost linear nitrosyl group.
APA, Harvard, Vancouver, ISO, and other styles
20

Lopes, Luiz G. F., Maria G. Gomes, Simone S. S. Borges, and Douglas W. Franco. "Correlation Between the Lever Parameter and Electronic Properties of Nitrosyl Ruthenium(II) Complexes." Australian Journal of Chemistry 51, no. 9 (1998): 865. http://dx.doi.org/10.1071/c97216.

Full text
Abstract:
The Lever parameter (EL), for a series of ruthenium nitrosyl complexes of the type [Ru(NH3)4NOL] n+, where ligand L = nicotinamide (nic), isonicotinamide (isn), pyrazine (pz), pyridine (py), imidazole (imN), L-histidine (L-hist), NH3, trimethyl phosphite [P(OMe)3] and triethyl phosphite [P(OEt)3], is correlated to Epc1 (the potential for the reduction RuNO++e- → RuNO 0) and v(NO) data. A correlation is observed between ∑EL and Epc1, and between ∑EL and v(NO), since these parameters are dependent on electronic characteristics of the nitrosyl group. For ∑EL v. Epc1, the slope, SL, and intercept, IL, of the linear correlation are related to interligand coupling through the metal centre and to the extent of the Ru → NO back bonding.
APA, Harvard, Vancouver, ISO, and other styles
21

Nagatsugi, F. "Selective nitrosyl group transfer reaction to cytidine using oligonucleotides bearing S-nitrosothioguanosine." Nucleic Acids Symposium Series 48, no. 1 (2004): 23–24. http://dx.doi.org/10.1093/nass/48.1.23.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

LEGZDINS, P. "ChemInform Abstract: Aspects of Organometallic Nitrosyl Chemistry of the Group 6 Elements." ChemInform 22, no. 32 (2010): no. http://dx.doi.org/10.1002/chin.199132279.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Boyko, Alexandr M., Nikolay O. Kamenshchikov, Yuriy K. Podoksenov, et al. "EFFECT OF PERIOPERATIVE NITRIC OXIDE DELIVERY ON NITROSYL STRESS AND LOCAL INFLAMMATION-MEDIATED TUBULAR KIDNEY INJURY DURING HEMIARCH SURGERY." Complex Issues of Cardiovascular Diseases 14, no. 3 (2025): 40–50. https://doi.org/10.17802/2306-1278-2025-14-3-40-50.

Full text
Abstract:
HighlightsThe organoprotective effect of nitric oxide during cardiac surgery, including reconstructive interventions on the aorta, is widely studied, however, its use may be limited due to the likely development of side effects, including kidney damage. This paper presents data on the study of the possible negative effects of nitric oxide on the kidneys in patients undergoing Hemiarch surgery. AbstractAim. To evaluate the effect of perioperative nitric oxide delivery on the severity of nitrosyl stress and renal tubular injury mediated by local inflammation activation during Hemiarch surgery under conditions of cardiopulmonary bypass and hypothermic circulatory arrest.Methods. The work presents the data of a single-center, single-blind, prospective, randomized controlled trial. The study included 80 patients over 18 years of age who underwent Hemiarch surgeries under artificial circulation and hypothermic circulatory arrest for non-syndromic ascending aortic aneurysms in the period 2020–2023. All patients were randomized into two groups in a 1:1 ratio: the NO group (main group), which received perioperative delivery of nitric oxide at a concentration of 80 ppm, and the standard perioperative support group (control group, NO delivery was not performed). To assess the severity of NO-mediated nitrosyl stress, the concentration of nitrotyrosine in the blood serum was measured. Blood was sampled immediately after placement of the central venous catheter and 4 hours after the end of the surgery. To assess the severity of renal tubular injury mediated by local inflammation activation, the concentration of IL-18 in urine was determined. Urine was collected after bladder catheterization and 4 hours after the end of surgery.Results. The concentration of nitrotyrosine 4 hours after surgery was 10.67 [8.99; 12.50] ng/ml in the NO group and 6.74 [5.89; 10.50] ng/ml in the no-NO group (p = 0.13). The concentration of IL-18 in urine 4 hours after surgery was 5.01 [4.06; 5.98] ng/ml in the NO group and 5.82 [3.60; 29.40] ng/ml in the no-NO group (p = 0.50).Conclusion. Perioperative delivery of 80 ppm NO during Hemiarch surgery under CPB and hypothermic circulatory arrest does not induce NO-mediated nitrosyl stress and does not affect renal tubular injury mediated by local inflammatory activation.
APA, Harvard, Vancouver, ISO, and other styles
24

Fernando, Veani, Xunzhen Zheng, Yashna Walia, Vandana Sharma, Joshua Letson, and Saori Furuta. "S-Nitrosylation: An Emerging Paradigm of Redox Signaling." Antioxidants 8, no. 9 (2019): 404. http://dx.doi.org/10.3390/antiox8090404.

Full text
Abstract:
Nitric oxide (NO) is a highly reactive molecule, generated through metabolism of L-arginine by NO synthase (NOS). Abnormal NO levels in mammalian cells are associated with multiple human diseases, including cancer. Recent studies have uncovered that the NO signaling is compartmentalized, owing to the localization of NOS and the nature of biochemical reactions of NO, including S-nitrosylation. S-nitrosylation is a selective covalent post-translational modification adding a nitrosyl group to the reactive thiol group of a cysteine to form S-nitrosothiol (SNO), which is a key mechanism in transferring NO-mediated signals. While S-nitrosylation occurs only at select cysteine thiols, such a spatial constraint is partially resolved by transnitrosylation, where the nitrosyl moiety is transferred between two interacting proteins to successively transfer the NO signal to a distant location. As NOS is present in various subcellular locales, a stress could trigger concerted S-nitrosylation and transnitrosylation of a large number of proteins involved in divergent signaling cascades. S-nitrosylation is an emerging paradigm of redox signaling by which cells confer protection against oxidative stress.
APA, Harvard, Vancouver, ISO, and other styles
25

Kosmachevskaya, Olga V., Elvira I. Nasybullina, Olesya V. Pokidova, Natalia A. Sanina, and Alexey F. Topunov. "Effects of Nitrosyl Iron Complexes with Thiol, Phosphate, and Thiosulfate Ligands on Hemoglobin." International Journal of Molecular Sciences 25, no. 13 (2024): 7194. http://dx.doi.org/10.3390/ijms25137194.

Full text
Abstract:
Nitrosyl iron complexes are remarkably multifactorial pharmacological agents. These compounds have been proven to be particularly effective in treating cardiovascular and oncological diseases. We evaluated and compared the antioxidant activity of tetranitrosyl iron complexes (TNICs) with thiosulfate ligands and dinitrosyl iron complexes (DNICs) with glutathione (DNIC-GS) or phosphate (DNIC-PO4−) ligands in hemoglobin-containing systems. The studied effects included the production of free radical intermediates during hemoglobin (Hb) oxidation by tert-butyl hydroperoxide, oxidative modification of Hb, and antioxidant properties of nitrosyl iron complexes. Measuring luminol chemiluminescence revealed that the antioxidant effect of TNICs was higher compared to DNIC-PO4−. DNIC-GS either did not exhibit antioxidant activity or exerted prooxidant effects at certain concentrations, which might have resulted from thiyl radical formation. TNICs and DNIC-PO4− efficiently protected the Hb heme group from decomposition by organic hydroperoxides. DNIC-GS did not exert any protective effects on the heme group; however, it abolished oxoferrylHb generation. TNICs inhibited the formation of Hb multimeric forms more efficiently than DNICs. Thus, TNICs had more pronounced antioxidant activity than DNICs in Hb-containing systems.
APA, Harvard, Vancouver, ISO, and other styles
26

Sharp, W. Brett, Peter Legzdins та Brian O. Patrick. "O-Protonation of a Terminal Nitrosyl Group To Form an η1-Hydroxylimido Ligand". Journal of the American Chemical Society 123, № 33 (2001): 8143–44. http://dx.doi.org/10.1021/ja002738m.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Baillie, Rhett A., and Peter Legzdins. "ChemInform Abstract: The Rich and Varied Chemistry of Group 6 Cyclopentadienyl Nitrosyl Complexes." ChemInform 47, no. 10 (2016): no. http://dx.doi.org/10.1002/chin.201610234.

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

Kou, Hui-Zhong, Hong-Mei Wang, Dai-Zheng Liao, et al. "A New One-Dimensional Bimetallic Complex: Cu(en)2Fe(CN)5(NO). Synthesis, Crystal Structure and Magnetic Behaviour." Australian Journal of Chemistry 51, no. 8 (1998): 661. http://dx.doi.org/10.1071/c97162.

Full text
Abstract:
The nitrosyl cyanide Cu(en)2Fe(CN)5(NO) was prepared by the reaction of Cu(en)2Cl2 and Na2 [Fe(CN)5(NO)].2H2O in aqueous solution. Single-crystal X-ray structure analysis revealed that the complex assumes a cyanide-bridged chain structure in which iron(II) is coordinated by five cyanide ligands and a nitrosyl group, and copper(II) is coordinated by two ethylenediamine ligands and two bridging cyanide ligands. The copper centres display a typical Jahn–Teller distortion characteristic of octahedral copper complexes with bond distance deviations in lattice symmetry from octahedral ideality: Cu-N(equatorial, mean) 2·010(3) Å and Cu-N(axial, mean) 2·569(3) Å. Infrared and ultraviolet studies shed light on the above molecular structure. A magnetic investigation showed the presence of a weak antiferromagnetic interaction (J = -1·06 cm-1) between the copper atoms within each chain through the diamagnetic Fe(CN)5(NO)2- ions.
APA, Harvard, Vancouver, ISO, and other styles
29

Olabe, José. "Coordination Chemistry of Nitric Oxide and Biological Signaling." Science Reviews - from the end of the world 2, no. 1 (2020): 64–99. http://dx.doi.org/10.52712/sciencereviews.v2i1.33.

Full text
Abstract:
Nitric Oxide (NO) is a key intermediate in the nitrogen redox cycles that operate in soils, water and biological fluids, affording reversible interconversions between nitrates to ammonia and vice-versa. The discovery of its biosynthesis in mammals for signaling purposes generated a research explosion on the ongoing chemistry occurring in specific cellular compartments, centered on NO reactivity toward O2, thiols, amines, and transition metals, as well as derivatives thereof. The present review deals with the coordination chemistry of NO toward selected iron and ruthenium centers. We place specific attention to the three redox states of the nitrosyl group: NO+, NO and NO–/HNO, describing changes in structure and reactivity as coordination takes place. Noteworthy are the results with the most reduced nitroxyl-species that allow establishing the changes in the measurable pKa values for the HNO-bound complexes, also revealing the abrupt decrease in reducing power and trans-releasing abilities of the protonated species over the unprotonated ones, NO–. Comparative results using non-heme and heme proteins and models prove useful for suggesting further improvements in the current research status of complex enzymatic behavior.
APA, Harvard, Vancouver, ISO, and other styles
30

Bastos, Tanira M., Marília I. F. Barbosa, Monize M. da Silva, et al. "Nitro/Nitrosyl-Ruthenium Complexes Are Potent and Selective Anti-Trypanosoma cruzi Agents Causing Autophagy and Necrotic Parasite Death." Antimicrobial Agents and Chemotherapy 58, no. 10 (2014): 6044–55. http://dx.doi.org/10.1128/aac.02765-14.

Full text
Abstract:
ABSTRACTcis-[RuCl(NO2)(dppb)(5,5′-mebipy)] (complex 1),cis-[Ru(NO2)2(dppb)(5,5′-mebipy)] (complex 2),ct-[RuCl(NO)(dppb)(5,5′-mebipy)](PF6)2(complex 3), andcc-[RuCl(NO)(dppb)(5,5′-mebipy)](PF6)2(complex 4), where 5,5′-mebipy is 5,5′-dimethyl-2,2′-bipyridine and dppb is 1,4-bis(diphenylphosphino)butane, were synthesized and characterized. The structure of complex 2 was determined by X-ray crystallography. These complexes exhibited a higher anti-Trypanosoma cruziactivity than benznidazole, the current antiparasitic drug. Complex 3 was the most potent, displaying a 50% effective concentration (EC50) of 2.1 ± 0.6 μM against trypomastigotes and a 50% inhibitory concentration (IC50) of 1.3 ± 0.2 μM against amastigotes, while it displayed a 50% cytotoxic concentration (CC50) of 51.4 ± 0.2 μM in macrophages. It was observed that the nitrosyl complex 3, but not its analog lacking the nitrosyl group, releases nitric oxide into parasite cells. This release has a diminished effect on the trypanosomal protease cruzain but induces substantial parasite autophagy, which is followed by a series of irreversible morphological impairments to the parasites and finally results in cell death by necrosis. In infected mice, orally administered complex 3 (five times at a dose of 75 μmol/kg of body weight) reduced blood parasitemia and increased the survival rate of the mice. Combination index analysis of complex 3 indicated that itsin vitroactivity against trypomastigotes is synergic with benznidazole. In addition, drug combination enhanced efficacy in infected mice, suggesting that ruthenium-nitrosyl complexes are potential constituents for drug combinations.
APA, Harvard, Vancouver, ISO, and other styles
31

Masters, A. P., M. Parvez, T. S. Sorensen та F. Sun. "Organometallic products from the reaction of the isoelectronic Mn(CO)5− and Cr(CO)4NO− metallate anions with bis-α-bromocyclopropyl ketone". Canadian Journal of Chemistry 71, № 2 (1993): 230–38. http://dx.doi.org/10.1139/v93-034.

Full text
Abstract:
Mn(CO)5− and Cr(CO)4NO− react with the title ketone to give organometallic products. In the chromium case, a single metallofuran product is produced. In the manganese reaction, one can isolate a series of four complexes, two of which have a structure closely related to the chromium complex. The other two complexes are assigned an acyl manganese structure. The structures of the chromium complex and one of the acyl manganese complexes have been determined by X-ray methods. One finds a distorted octahedral bonding about the metal atom in each case. The chromium complex has bond lengths very similar to those reported for a closely related manganese structure, implying a delocalization of electrons within a metallofuran ring. As expected for this mixed carbonyl nitrosyl complex, the nitrosyl group is positioned trans to the oxide bond, representing the weakest and strongest "trans-effect" substituents. The acyl manganese structures contain an Mn(CO)4 unit in contrast to the usual Mn(CO)5 acyl complexes, with the 18e− count being provided by an internal chelation with the n-electrons of the dicyclopropyl ketone group. The reaction mechanism for the formation of the complexes is postulated to involve initial bromine abstraction (two-electron reduction) by the metallate anion, with the Cr(CO)4NO− reaction being much faster compared to Mn(CO)5−. The resulting organic enolate then "back-reacts" with a carbonyl group of the just-formed metal carbonyl bromide to form a transient anionic Fischer carbene complex. Ultimately this intermediate loses the metal bromide ion with the formation of the neutral complexes.
APA, Harvard, Vancouver, ISO, and other styles
32

Brouwer, M., W. Chamulitrat, G. Ferruzzi, DL Sauls, and JB Weinberg. "Nitric oxide interactions with cobalamins: biochemical and functional consequences." Blood 88, no. 5 (1996): 1857–64. http://dx.doi.org/10.1182/blood.v88.5.1857.1857.

Full text
Abstract:
Abstract Nitric oxide (NO) is a paramagnetic gas that has been implicated in a wide range of biologic functions. The common pathway to evoke the functional response frequently involves the formation of an iron- nitrosyl complex in a target (heme) protein. In this study, we report on the interactions between NO and cobalt-containing vitamin B12 derivatives. Absorption spectroscopy showed that of the four Co(III) derivatives (cyanocobalamin [CN-Cbl], aquocobalamin [H2O-Cbl], adenosylcobalamin [Ado-Cbl], and methylcobalamin [MeCbl]), only the H2O- Cbl combined with NO. In addition, electron paramagnetic resonance spectroscopy of H2O-Cbl preparations showed the presence of a small amount of Cob-(II)alamin that was capable of combining with NO. The Co(III)-NO complex was very stable, but could transfer its NO moiety to hemoglobin (Hb). The transfer was accompanied by a reduction of the Co(III) to Co(II), indicating that NO+ (nitrosonium) was the leaving group. In accordance with this, the NO did not combine with the Hb Fe(II)-heme, but most likely with the Hb cysteine-thiolate. Similarly, the Co(III)-NO complex was capable of transferring its NO to glutathione. Ado-Cbl and Me-Cbl were susceptible to photolysis, but CN- Cbl and H2O-Cbl were not. The homolytic cleavage of the Co(III)-Ado or Co(III)-Me bond resulted in the reduction of the metal. When photolysis was performed in the presence of NO, formation of NO-Co(II) was observed. Co(II)-nitrosyl oxidized slowly to form Co(III)-nitrosyl. The capability of aquocobalamin to combine with NO had functional consequences. We found that nitrosylcobalamin had diminished ability to serve as a cofactor for the enzyme methionine synthase, and that aquocobalamin could quench NO-mediated inhibition of cell proliferation. Our in vitro studies therefore suggest that interactions between NO and cobalamins may have important consequences in vivo.
APA, Harvard, Vancouver, ISO, and other styles
33

Brouwer, M., W. Chamulitrat, G. Ferruzzi, DL Sauls, and JB Weinberg. "Nitric oxide interactions with cobalamins: biochemical and functional consequences." Blood 88, no. 5 (1996): 1857–64. http://dx.doi.org/10.1182/blood.v88.5.1857.bloodjournal8851857.

Full text
Abstract:
Nitric oxide (NO) is a paramagnetic gas that has been implicated in a wide range of biologic functions. The common pathway to evoke the functional response frequently involves the formation of an iron- nitrosyl complex in a target (heme) protein. In this study, we report on the interactions between NO and cobalt-containing vitamin B12 derivatives. Absorption spectroscopy showed that of the four Co(III) derivatives (cyanocobalamin [CN-Cbl], aquocobalamin [H2O-Cbl], adenosylcobalamin [Ado-Cbl], and methylcobalamin [MeCbl]), only the H2O- Cbl combined with NO. In addition, electron paramagnetic resonance spectroscopy of H2O-Cbl preparations showed the presence of a small amount of Cob-(II)alamin that was capable of combining with NO. The Co(III)-NO complex was very stable, but could transfer its NO moiety to hemoglobin (Hb). The transfer was accompanied by a reduction of the Co(III) to Co(II), indicating that NO+ (nitrosonium) was the leaving group. In accordance with this, the NO did not combine with the Hb Fe(II)-heme, but most likely with the Hb cysteine-thiolate. Similarly, the Co(III)-NO complex was capable of transferring its NO to glutathione. Ado-Cbl and Me-Cbl were susceptible to photolysis, but CN- Cbl and H2O-Cbl were not. The homolytic cleavage of the Co(III)-Ado or Co(III)-Me bond resulted in the reduction of the metal. When photolysis was performed in the presence of NO, formation of NO-Co(II) was observed. Co(II)-nitrosyl oxidized slowly to form Co(III)-nitrosyl. The capability of aquocobalamin to combine with NO had functional consequences. We found that nitrosylcobalamin had diminished ability to serve as a cofactor for the enzyme methionine synthase, and that aquocobalamin could quench NO-mediated inhibition of cell proliferation. Our in vitro studies therefore suggest that interactions between NO and cobalamins may have important consequences in vivo.
APA, Harvard, Vancouver, ISO, and other styles
34

Crespo, Paula M., Oscar F. Odio, and Edilso Reguera. "Photochemistry of Metal Nitroprussides: State-of-the-Art and Perspectives." Photochem 2, no. 2 (2022): 390–404. http://dx.doi.org/10.3390/photochem2020027.

Full text
Abstract:
This contribution summarizes the current state in the photochemistry of metal nitroprussides, which is dominated by the electronic structure of the nitrosyl group. From the combination of p orbitals of the nitrogen and oxygen atoms in the NO+ ligand, a π*NO molecular orbital of relatively low energy is formed, which has π*2px and π*2py character. This is a double degenerate orbital. When the nitrosyl group is found coordinated to the iron atom in the nitroprusside ion, the availability of that low energy π*NO orbital results in light-induced electronic transitions from the iron atom dxy, dxz and dyz orbitals, 2b2 (xy) → 7e (π*NO) and 6e (xz,yz) → 7e (π*NO), which are observed at 498 and 394 nm, respectively. These light-induced transitions and the possibility of NO isomer formation dominate the photochemistry of metal nitroprussides. In this feature paper, we discuss the implications of such transitions in the stability of coordination compounds based on the nitroprusside ion in the presence of water molecules for both 3D and 2D structures, including the involved degradation mechanisms. These photo-induced electronic transitions modify the physical and functional properties of solids where the nitroprusside ion forms part of their structure and appear as an opportunity for tuning their magnetic, electrical, optical and as energy-applied materials, for instance. This contribution illustrates these opportunities with results from some recently reported studies, and possible research subjects, even some not explored, are mentioned.
APA, Harvard, Vancouver, ISO, and other styles
35

Romashev, Nikolai F., Ivan V. Bakaev, Veronika I. Komlyagina, et al. "Iridium Complexes with BIAN-Type Ligands: Synthesis, Structure and Redox Chemistry." International Journal of Molecular Sciences 24, no. 13 (2023): 10457. http://dx.doi.org/10.3390/ijms241310457.

Full text
Abstract:
A series of iridium complexes with bis(diisopropylphenyl)iminoacenaphtene (dpp-bian) ligands, [Ir(cod)(dpp-bian)Cl] (1), [Ir(cod)(NO)(dpp-bian)](BF4)2 (2) and [Ir(cod)(dpp-bian)](BF4) (3), were prepared and characterized by spectroscopic techniques, elemental analysis, X-ray diffraction analysis and cyclic voltammetry (CV). The structures of 1–3 feature a square planar backbone consisting of two C = C π-bonds of 1,5-cyclooctadiene (cod) and two nitrogen atoms of dpp-bian supplemented with a chloride ion (for 1) or a NO group (for 2) to complete a square-pyramidal geometry. In the nitrosyl complex 2, the Ir-N-O group has a bent geometry (the angle is 125°). The CV data for 1 and 3 show two reversible waves between 0 and -1.6 V (vs. Ag/AgCl). Reversible oxidation was also found at E1/2 = 0.60 V for 1. Magnetochemical measurements for 2 in a range from 1.77 to 300 K revealed an increase in the magnetic moment with increasing temperature up to 1.2 μB (at 300 K). Nitrosyl complex 2 is unstable in solution and loses its NO group to yield [Ir(cod)(dpp-bian)](BF4) (3). A paramagnetic complex, [Ir(cod)(dpp-bian)](BF4)2 (4), was also detected in the solution of 2 as a result of its decomposition. The EPR spectrum of 4 in CH2Cl2 is described by the spin Hamiltonian Ĥ = gβHŜ with S = 1/2 and gxx = gyy = 2.393 and gzz = 1.88, which are characteristic of the low-spin 5d7-Ir(II) state. DFT calculations were carried out in order to rationalize the experimental results.
APA, Harvard, Vancouver, ISO, and other styles
36

Richter-Addo, George B., Ralph A. Wheeler, Christopher Adam Hixson та ін. "Unexpected Nitrosyl-Group Bending in Six-Coordinate {M(NO)}6σ-Bonded Aryl(iron) and -(ruthenium) Porphyrins". Journal of the American Chemical Society 123, № 26 (2001): 6314–26. http://dx.doi.org/10.1021/ja010276m.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

SIZOVA, OL'GA V., VICTOR I. BARANOVSKI, NINA V. IVANOVA, and VLADIMIR V. SIZOV. "Valence bond analysis of the bonding in transition metal compounds: the RuNO group in nitrosyl complexes." Molecular Physics 101, no. 6 (2003): 715–20. http://dx.doi.org/10.1080/0026897031000075606.

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

Hayton, Trevor W., Brian O. Patrick, Peter Legzdins, and W. Stephen McNeil. "The solid-state molecular structure of W(NO)3Cl3 and the nature of its W—NO bonding." Canadian Journal of Chemistry 82, no. 2 (2004): 285–92. http://dx.doi.org/10.1139/v03-206.

Full text
Abstract:
The monomeric trinitrosyl complex, W(NO)3Cl3, can be prepared by the treatment of WCl6 in CH2Cl2 with NO gas, and its identity has been unambiguously confirmed by a single-crystal X-ray diffraction analysis. The complex crystallizes in the space group Pmn21 as a three-component twin (a = 10.4280(4) Å, b = 6.3289(2) Å, c = 5.6854(2) Å, Z = 2, R1 = 0.065, wR2 = 0.176). Its solid-state molecular structure consists of a tungsten centre bound to three chloride ligands and three linear nitrosyl ligands in a fac-octahedral stereochemistry. In addition, the structure contains a crystallographically imposed mirror plane. The two independent W—N linkages are 1.88(2) and 1.92(1) Å long, while the two corresponding N—O bond lengths are 1.13(2) and 1.16(2) Å. DFT calculations on fac-W(NO)3Cl3 at the B3LYP/LANL2DZ level of theory afford optimized intramolecular metrical parameters that match the X-ray crystallographically determined bond lengths and bond angles quite well. In addition, they provide a rationale for the nearly linear W-N-O linkages extant in the complex. Solutions of fac-W(NO)3Cl3 in CH2Cl2 lose ClNO under ambient conditions and deposit the well-known [W(NO)2Cl2]n polymer, and this conversion is fully reversible.Key words: nitrosyl, tungsten, structure, bonding.
APA, Harvard, Vancouver, ISO, and other styles
39

Rahman, Md Hafizur, and Michael D. Ryan. "Ligand noninnocence in an Fe(NO) porphyrin dianion: Infrared spectroelectrochemistry and DFT calculations." Journal of Porphyrins and Phthalocyanines 29, no. 05n06 (2025): 759–66. https://doi.org/10.1142/s1088424625500658.

Full text
Abstract:
Iron(II) octaethylporphyrin nitrosyl, [Fe(OEP)(NO)], was reduced in two well-separated electron transfer steps. The first reduction to form the anion, [Fe(OEP)(NO)][Formula: see text], has been studied using voltammetry and spectroelectrochemistry (visible and infrared). The results of these works showed that the {FeNO}7 moiety was reduced to a {FeNO}8 complex, with the reduction mostly occurring in the nitrosyl group. Voltammetry and visible spectroelectrochemistry have shown that the second reduction forms a stable dianion complex, [Fe(OEP)(NO)][Formula: see text]. This second reduction was studied using infrared spectroelectrochemistry and DFT calculations to characterize the dianionic species. The results showed that the [Formula: see text] for the anion downshifted from 1440 cm[Formula: see text] to 1380 cm[Formula: see text], a downshift of 60 cm[Formula: see text]. The identity of this vibration was confirmed by substituting [Formula: see text]NO for [Formula: see text]NO. DFT calculations showed that the porphyrin ring of [Fe(OEP)(NO)][Formula: see text] was non-innocent with the formation of a porphyrin radical anion and little change in the {FeNO}8 moiety. As with the first reduction, the experimental downshift was larger than the DFT calculated values.
APA, Harvard, Vancouver, ISO, and other styles
40

Sellmann, Dieter, Helmut Schillinger, and Falk Knoch. "Übergangsmetallkomplexe mit Schwefelliganden, LXXXVI. Säure-Base- und Redox-Reaktionen von [Ni(′MeS2′)2] mit H+, PMe3, NO+ und NO. Röntgenstrukturanalyse von [Ni(PMe3)(′MeS2′)2] (′MeS2′ = o-(Methylthio)thiophenolat(1–)) / Transition-Metal Complexes with Sulfur Ligands, LXXXVI. Acid-Base and Redox Reactions of [Ni(′MeS2')2] with H+, PMe3, NO+ and NO. X-Ray Structure Determination of [Ni(PMe3)(′MeS2′)2] (′MeS2′ = o-(methylthio)thiophenolate(1–))." Zeitschrift für Naturforschung B 47, no. 5 (1992): 748–54. http://dx.doi.org/10.1515/znb-1992-0523.

Full text
Abstract:
In order to elucidate specific properties of nickel-sulfur complexes, addition and substitution reactions of [Ni(′MeS2′)2] (1) were investigated. 1 is rapidly hydrolyzed by aqueous HCl yielding ′MeS2′–H and Ni(II) ions. 1 coordinates phosphines as coligands, thioether donors decoordinate, however, simultaneously. The monophosphine complex [Ni(PMe3)(′MeS2′ )2] (2) was characterized by X-ray structure determination. It contains a square-planar NiS3P unit and one decoordinated thioether group. Redox reactions of 1 occur with NO+ and NO, yielding the binuclear nitrosyl complexes cis- and trans-[Ni(NO)(′MeS2′)]2 (3) and the disulfide [′MeS2']2•
APA, Harvard, Vancouver, ISO, and other styles
41

Pierloot, Kristine, Quan Manh Phung, and Abhik Ghosh. "Electronic Structure of Neutral and Anionic Iron–Nitrosyl Corrole. A Multiconfigurational and Density Matrix Renormalization Group Investigation." Inorganic Chemistry 59, no. 16 (2020): 11493–502. http://dx.doi.org/10.1021/acs.inorgchem.0c01312.

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

Mikhailov, Artem, Vedran Vuković, Christian Kijatkin, et al. "Combining photoinduced linkage isomerism and nonlinear optical properties in ruthenium nitrosyl complexes." Acta Crystallographica Section B Structural Science, Crystal Engineering and Materials 75, no. 6 (2019): 1152–63. http://dx.doi.org/10.1107/s205252061901357x.

Full text
Abstract:
The complex trans-[RuNO(NH3)4F]SiF6 was synthesized in quantitative yield and the structure was characterized by X-ray diffraction and spectroscopic methods. The complex crystallizes in the non-centrosymmetric space group Pn. Hirshfeld surface analysis revealed that the dominant intermolecular interactions are of types H...F and F...O, which are likely to be responsible for the packing of the molecules in a non-centrosymmetric structure. Irradiation with blue light leads to the formation of Ru–ON (metastable state MS1) and Ru–η2-(NO) (metastable state MS2) bond isomers, as shown by IR and UV–Vis spectroscopy. The structural features of the MS1 isomer were elucidated by photocrystallography. The complex exhibits exceptionally good thermal stability of the metastable state MS1, such that it can be populated by light at 290–300 K, which is important for potential applications. The second harmonic (SH) emission can be generated by femtosecond-pulsed irradiation of the complex. The generated SH is rather efficient and stable under long-term exposure. Finally, since both metastable states and harmonic generation can be generated at room temperature, an attempt to drive the SH response by photoisomerization of the nitrosyl ligand was made and is discussed.
APA, Harvard, Vancouver, ISO, and other styles
43

Beck, Johannes, and Joachim Strähle. "Synthese und Kristallstruktur von Mo(NO)(N3)3terpy, einem Nitrosyl-azido-Komplex des Molybdäns mit siebenfacher Koordination." Zeitschrift für Naturforschung B 40, no. 7 (1985): 891–94. http://dx.doi.org/10.1515/znb-1985-0707.

Full text
Abstract:
Abstract Mo(NO)(N3)3terpy was obtained in form of yellow platelets from Mo(CO)3terpy and (CH3)3SiN3 in nitroethane as solvent. It crystallizes in the triclinic space group P1̄ with the lattice parameters a = 817.3, b = 866.7, c = 1416.2 pm, a = 71.97°, β = 102.15°, γ = 107.34°. The unit cell contains two monomeric complexes in which the central Mo atom exhibits the coordination number seven in a slightly distorted pentagonal bipyramid. One azido group and the nitrosyl ligand are in axial positions. NO+ coordinates in a linear arrangement (Mo-N-O = 176.2°) and acts as a three electron donor completing the noble gas configuration for the molybdenum atom. A weak trans effect of the NO-ligand causes different Mo-N distances to the axial (Mo-N31 = 218.4 pm) and the equatorial (Mo-N11 = 211.3, Mo-N21 = 212.9 pm) azido groups.
APA, Harvard, Vancouver, ISO, and other styles
44

Marcinkowska-Lesiak, Monika, Kazem Alirezalu, Adrian Stelmasiak, et al. "Physicochemical Characteristics of Pork Liver Pâtés Containing Nonthermal Air Plasma-Treated Egg White as an Alternative Source of Nitrite." Applied Sciences 13, no. 7 (2023): 4464. http://dx.doi.org/10.3390/app13074464.

Full text
Abstract:
The use of nonthermal air plasma is rapidly becoming a novel technology as an alternative source of nitrites in the meat industry. As egg white is a versatile and cost-effective ingredient commonly used to improve the texture of meat products, the effect of its addition after plasma treatment (PTEW) on the yield, pH, residual nitrite, nitrosyl hemochrome, TBARS, color, texture parameters, and aroma profile of pork liver pâtés was studied. The nitrite ion content of plasma-activated egg whites was adjusted to the positive controls containing 60 ppm (PC1) and 120 ppm (PC2) sodium nitrite by modifying the duration of their plasma treatment (PTEW1 and PTEW2, respectively). A group without the addition of nitrites was also manufactured (NC). Each treatment (NC, PC1, PC2, PTEW1, PTEW2) was analyzed on days 1, 3, 5, and 7 of storage at 4 °C. The results showed that liver pâtés containing plasma-treated egg whites had a similar nitrite and nitrosyl hemochrome content compared to samples containing the same amount of nitrite ions derived from sodium nitrite (p ≥ 0.05). In addition, 40 ppm nitrite ions, regardless of the source, was sufficient to achieve the desired reddish-pink color of the product over the entire storage period. Both nitrites from sodium nitrite and plasma-treated egg whites also significantly reduced lipid oxidation compared to the NC group (between 10% and 23% reduction on the last day), but had no significant effect on yield, pH, and texture parameters of the products. Based on the principal component analysis (PCA), the aroma profile of pâtés differed significantly between the groups with and without nitrites, with the largest differences observed on the first day (approx. 88%). Importantly, PTEW1 and PTEW2 aroma after production was similar to group PC2. The results of our study suggest that plasma-activated egg whites can be used as a potential source of nitrite in liver pâté production without adversely affecting the technological properties and shelf life of the final product.
APA, Harvard, Vancouver, ISO, and other styles
45

Field, Leslie D., Trevor W. Hambley, Taian He,, Anthony F. Masters, and Peter Turner. "The Structures of the Decaphenylmetallocenium Cations of Chromium and Cobalt." Australian Journal of Chemistry 50, no. 11 (1997): 1035. http://dx.doi.org/10.1071/c97075.

Full text
Abstract:
Decaphenylchromocenium and decaphenylcobaltocenium cations [M(η5-C5Ph5)2]+, M = Cr and Co, were synthesized by oxidation of the corresponding neutral decaphenylchromocene and decaphenylcobaltocene respectively with nitrosyl tetrafluoroborate. The complexes are air-stable and were fully characterized; decaphenylchromocenium tetrafluoroborate (1) and decaphenylcobaltocenium tetrafluoroborate (2) were structurally characterized by X-ray crystallography. Crystals of (1) (as a water/methylene chloride solvate), C70·5H50BClCrF4O0·5, M 1079·42, are triclinic, space group P -1 (No. 2), a 13·634(5), b 17·424(5), c 13·298(4) Å, α 106·45(2), β 101·83(3), γ 74·56(2)°, Z 2. Crystals of (2) (as a water/methylene chloride solvate), C70·5H50BClCoF4O0·5, M 1086·36, are triclinic, space group P-1 (No. 2), a 13·633(2), b 17·683(3), c 13·255(3) Å, α 107·95(1), β 102·71(1), γ 73·23(1)°, Z2.
APA, Harvard, Vancouver, ISO, and other styles
46

Batista, Alzir A., Salete L. Queiroz, Peter C. Healy, et al. "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 (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
47

Ali, Md Monsur, Fumi Nagatsugi, Shizuka Nakayama, Md Rowshon Alam, Takeshi Kawasaki, and Shigeki Sasaki. "DESIGN OF HIGHLY EFFICIENT AND SELECTIVE TRANSFER REACTION OF NITROSYL GROUP TO dC AND dMC RESULTING IN SPECIFIC DEAMINATION." Nucleosides, Nucleotides & Nucleic Acids 24, no. 5-7 (2005): 721–24. http://dx.doi.org/10.1081/ncn-200060312.

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

Legzdins, Peter, W. Stephen McNeil, Edward G. Vessey, Raymond J. Batchelor, and Frederick W. B. Einstein. "Organometallic nitrosyl chemistry. 55. Regioselective synthesis of 2-pyrones mediated by organometallic dinitrosyl cations of the Group 6 metals." Organometallics 11, no. 7 (1992): 2718–23. http://dx.doi.org/10.1021/om00043a071.

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

Tanifuji, Naoki, Kenji Matsuda, and Masahiro Irie. "Effect of Imino Nitroxyl and Nitronyl Nitroxyl Groups on the Photochromic Reactivity of Diarylethenes." Organic Letters 7, no. 17 (2005): 3777–80. http://dx.doi.org/10.1021/ol051463i.

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

Fenske, Dieter, Udo Demant, and Kurt Dehnicke. "Chloronitrosylkomplexe von Ruthenium(II) Die Kristallstruktur von (PPh3Me)2[Ru(NO)Cl4]2-2 CH2Cl2 / Chloro Nitrosyl Complexes of Ruthenium (II) The Crystal Structure of (PPh3Me)2 [Ru(NO)Cl4]2 • 2 CH2Cl2." Zeitschrift für Naturforschung B 40, no. 12 (1985): 1672–76. http://dx.doi.org/10.1515/znb-1985-1213.

Full text
Abstract:
Abstract Ruthenium trichloride, obtained from its hydrate with thionyl chloride, reacts with excess trichloronitrom ethane yielding polymer [Ru(NO)Cl3]; by addition of triphenylmethylphosphonium chloride in dichlorom ethane (PPh3Me)2[Ru(NO)Cl4]2 · 2 CH2Cl2 is obtained, the IR spectrum of which is reported and assigned. Its crystal structure was determined with X-ray diffraction data (6404 independent observed reflexions, R = 0.068). Crystal data at -90 °C: a = 1145, b = 1591, c = 1406 pm, β = 96,0°, Z = 2, space group P21/C. The structure consists of PPh3Me⊕ cations, centrosymmetric anions [Ru(NO)Cl4]22⊖ nearly fulfilling C2h symmetry, and CH2Cl2 molecules. In the anions the Ru atoms are linked via chloro bridges; the nitrosyl groups occupy axial positions with bond distances RuN of 175 and NO of 113 pm, bond angle RuNO 172.7°.
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!