首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
The labeling of (bio)molecules with metallic radionuclides such as 99mTc demands conjugated, multidentate chelators. However, this is not always necessary since phenyl rings can directly serve as integrated, organometallic ligands. Bis-arene sandwich complexes are generally prepared by the Fischer–Hafner reaction. In extension of this, we show that [99mTc(η6-C6R6)2]+-type complexes are directly accessible from water and [99mTcO4], even using arenes incompatible with Fischer–Hafner conditions. To unambiguously confirm the nature of these unprecedented 99mTc complexes, their rhenium homologous have been prepared by substituting naphthalene ligands in [Re(η6-C10H8)2]+ with the corresponding phenyl groups. The ease with which highly stable [99mTc(η6-C6R6)2]+ complexes are formed under standard labeling conditions enables a multitude of new potential imaging agents based on commercial pharmaceuticals or lead structures.  相似文献   

2.
Reported herein is a new concept for the labelling of biomolecules with small [99 mTcO3]+ complexes through a [3+2] cycloaddition with alkenes for radiopharmaceutical applications. We developed convenient reactions for the synthesis of small, water stable fac‐[TcO3(tacn‐R)]+ complexes (99Tc and 99mTc, tacn=1,4,7‐triazacyclononane, R=H, ‐CH2‐C6H5, ‐CH2‐C6H4COOH). With alkenes, these high valent [99mTcO3]+ complexes undergo [3+2] cycloaddition with formation of the corresponding TcV–glycolato complexes. The 99mTcV and 99mTcVII complexes are stable at 37 °C in water and in the presence of serum proteins. Therefore, new opportunities in technetium chemistry are enabled with a high potential for medicinal and biological applications. In contrast to classical labelling, the presented strategy is ligand and not metal‐centred.  相似文献   

3.
[99mTc(I)]+ and [99mTc(I)(CO)3]+ complexes with isocyanide exhibit high stability, which makes them suitable platforms to develop novel 99mTc radiopharmaceuticals. To develop novel 99mTc radiotracers for imaging hypoxia, in this study, a novel L ligand (4-nitroimidazole isocyanide derivative) was synthesized and labelled using [99mTc(I)]+ core and [99mTc(I)(CO)3]+ core to produce [99mTc(L)6]+ and [99mTc(CO)3(L)3]+ with high yields. To verify the structure of the 99mTc complexes, corresponding rhenium analogues were synthesized and characterized. Both of the 99mTc complexes were stable and hydrophilic. in vitro cellular uptake results showed they could exhibit good hypoxic selectivity. The evaluation of biodistribution in mice bearing S180 tumors indicated both of them could accumulate in tumor. Between them, [99mTc(L)6]+ exhibited higher tumor uptake and tumor/non-target ratio than [99mTc(CO)3(L)3]+. Further, single photon emission computed tomography (SPECT) imaging studies of [99mTc(L)6]+ indicated an obvious accumulation in tumor and the value of the region-of-interest (ROI) ratio of the uptake for the tumor site to the corresponding non-tumor region was 5.64 ± 0.52. The above results suggested [99mTc(L)6]+ would be a potential tracer for imaging tumor hypoxia.  相似文献   

4.
Syntheses and Crystal Structures of (η6‐Diarene)TiII‐bis(tetrachloroaluminate) Complexes, Diarene = Biphenyl or 3,5,3′,5′‐Tetramethyl‐biphenyl Syntheses of (η6‐diarene)TiII(AlCl4)2 complexes were performed by the Fischer‐Hafner method. The diarenes employed were biphenyl and 3,5,3′,5′‐tetramethyl‐biphenyl. In each of the resulting complexes, (η6‐C12H10)TiII(AlCl4)2 ( 1 ) and (η6‐C16H18)TiII(AlCl4)2 ( 2 ), only one C6‐ring of a diarene is coordinatively active. 1 : Space group Pbca, Z = 8, lattice constants at 20 °C: a = 16.864(3), b = 13.931(3), c = 18.807(3) Å; R1 = 0.048. 2 : Space group P21/n, Z = 4, lattice constants at 20 °C: a = 9.775(1), b = 13.720(1), c = 20.214(1) Å; β = 95.50(1)°; R1 = 0.050.  相似文献   

5.
Diimido, Imido Oxo, Dioxo, and Imido Alkylidene Halfsandwich Compounds via Selective Hydrolysis and α—H Abstraction in Molybdenum(VI) and Tungsten(VI) Organyl Complexes Organometal imides [(η5‐C5R5)M(NR′)2Ph] (M = Mo, W, R = H, Me, R′ = Mes, tBu) 4 — 8 can be prepared by reaction of halfsandwich complexes [(η5‐C5R5)M(NR′)2Cl] with phenyl lithium in good yields. Starting from phenyl complexes 4 — 8 as well as from previously described methyl compounds [(η5‐C5Me5)M(NtBu)2Me] (M = Mo, W), reactions with aqueous HCl lead to imido(oxo) methyl and phenyl complexes [(η5‐C5Me5)M(NtBu)(O)(R)] M = Mo, R = Me ( 9 ), Ph ( 10 ); M = W, R = Ph ( 11 ) and dioxo complexes [(η5‐C5Me5)M(O)2(CH3)] M = Mo ( 12 ), M = W ( 13 ). Hydrolysis of organometal imides with conservation of M‐C σ and π bonds is in fact an attractive synthetic alternative for the synthesis of organometal oxides with respect to known strategies based on the oxidative decarbonylation of low valent alkyl CO and NO complexes. In a similar manner, protolysis of [(η5‐C5H5)W(NtBu)2(CH3)] and [(η5‐C5Me5)Mo(NtBu)2(CH3)] by HCl gas leads to [(η5‐C5H5)W(NtBu)Cl2(CH3)] 14 und [(η5‐C5Me5)Mo(NtBu)Cl2(CH3)] 15 with conservation of the M‐C bonds. The inert character of the relatively non‐polar M‐C σ bonds with respect to protolysis offers a strategy for the synthesis of methyl chloro complexes not accessible by partial methylation of [(η5‐C5R5)M(NR′)Cl3] with MeLi. As pure substances only trimethyl compounds [(η5‐C5R5)M(NtBu)(CH3)3] 16 ‐ 18 , M = Mo, W, R = H, Me, are isolated. Imido(benzylidene) complexes [(η5‐C5Me5)M(NtBu)(CHPh)(CH2Ph)] M = Mo ( 19 ), W ( 20 ) are generated by alkylation of [(η5‐C5Me5)M(NtBu)Cl3] with PhCH2MgCl via α‐H abstraction. Based on nmr data a trend of decreasing donor capability of the ligands [NtBu]2— > [O]2— > [CHR]2— ? 2 [CH3] > 2 [Cl] emerges.  相似文献   

6.
A novel C3′‐functionalized thymidine dithiocarbamate derivative (3’DTC‐TdR) was successfully synthesized and labelled using [99mTcO]3+ core and [99mTc(CO)3(H2O)3]+ core with high yields. The structures of the 99mTc complexes were verified by preparation and characterization of the corresponding stable rhenium complexes. Both of the complexes were lipophilic and stable in vitro. Cell internalization experiments indicated that the uptakes of 99mTcO‐3’DTC‐TdR were related to nucleoside transporters. Biodistribution of these complexes in mice bearing tumor showed that they had high tumor uptakes, good tumor/muscle ratios and tumor/blood ratios. Especially for 99mTcO‐3’DTC‐TdR, it exhibited the highest tumor/muscle ratio and tumor/blood ratio at 4 h post‐injection. SPECT/CT imaging studies indicated clear accumulation in tumor, suggesting 99mTcO‐3’DTC‐TdR would be a promising candidate for tumor imaging.  相似文献   

7.
Isocyanide is a strong coordination ligand that can coordinate with [99mTc(I)(CO)3]+ core and [99mTc(I)]+ core to produce stable 99mTc complexes, therefore developing a 99mTc-labeled isocyanide complex for single-photon emission computed tomography (SPECT) imaging is considered to be of great interest. In order to develop potential tumor imaging agents with satisfied tumor uptake and suitable pharmacokinetic properties in vivo, a novel d -glucosamine isocyanide derivative, 4-isocyano-N-(2,4,5-trihydroxy-6-(hydroxymethyl)tetrahydro-2H-pyran-3-yl)butanamide (CN3DG), was synthesized and radiolabeled with [99mTc(I)]+ and [99mTc(CO)3]+ cores to obtain [99mTc(CN3DG)6]+ and [99mTc(CO)3(CN3DG)3]+ in high radiolabeling yields (>95%). Both of the complexes showed good hydrophilicity and great stability in vitro. Cell uptake studies performed in S180 cells demonstrated they were transported into cells by glucose transporters. Biodistribution studies of the two complexes in mice bearing S180 tumor showed they had high tumor uptakes and rapid clearance from muscle and blood so that the tumor/blood and tumor/muscle ratios were high. By comparison, [99mTc(CN3DG)6]+ was superior to [99mTc(CO)3(CN3DG)3]+ in regard to tumor uptake, tumor/blood and tumor/liver ratios. S180 tumors could be seen clearly from the SPECT/CT images with [99mTc(CN3DG)6]+. Considering its favorable properties, [99mTc(CN3DG)6]+ would be a promising tumor imaging agent and needs to be further studied.  相似文献   

8.
Bis-arene sandwich complexes are generally prepared by the Fischer-Hafner reaction, which conditions are incompatible with most O- and N- functional groups. We report a new way for the synthesis of sandwich type complexes [Re(η6-arene)2]+ and [Re(η6-arene)(η6-benzene)]+ from [Re(η6-napht)2]+ and [Re(η6-napht)(η6-benzene)]+, with functionalized arenes and pharmaceuticals. N-methylpyrrolidine (NMP) facilitates the substitution of naphthalene with the incoming arene. A series of fully characterized rhenium sandwich complexes with simple arenes, such as aniline, as well as with active compounds like lidocaine and melatonin are presented. With these rhenium compounds in hand, the radioactive sandwich complexes [99mTc(η6-pharm)2]+ (pharm=pharmaceutical) can be unambiguously confirmed. The direct labelling of pharmaceuticals with 99mTc through η6-coordination to phenyl rings and the confirmation of the structures with the rhenium homologues opens a path into molecular theranostics.  相似文献   

9.
Chiral Half‐sandwich Pentamethylcyclopentadienyl Rhodium(III) and Iridium(III) Complexes with Schiff Bases from Salicylaldehyde and α‐Amino Acid Esters [1] A series of diastereoisomeric half‐sandwich complexes with Schiff bases from salicylaldehyde and L‐α‐amino acid esters including chiral metal atoms, [(η5‐C5H5)(Cl)M(N,O‐Schiff base)], has been obtained from chloro bridged complexes [(η5‐C5Me5)(Cl)M(μ‐Cl)]2 (M = Rh, Ir). Abstraction of chloride from these complexes with Ag[BF4] or Ag[SO3CF3] affords the highly sensitive compounds [(η5‐C5Me5)M(N,O‐Schiff base]+X? (M = Rh, Ir; X = BF4, CF3SO3) to which PPh3 can be added under formation of [(η5‐C5Me5)M(PPh3)(N,O‐Schiff base)]+X?. The diastereoisomeric ratio of the complexes ( 1 ‐ 7 and 11 ‐ 12 ) has been determined from NMR spectra.  相似文献   

10.
Monocationic bis‐allyl complexes [Ln(η3‐C3H5)2(thf)3]+[B(C6X5)4]? (Ln=Y, La, Nd; X=H, F) and dicationic mono‐allyl complexes of yttrium and the early lanthanides [Ln(η3‐C3H5)(thf)6]2+[BPh4]2? (Ln=La, Nd) were prepared by protonolysis of the tris‐allyl complexes [Ln(η3‐C3H5)3(diox)] (Ln=Y, La, Ce, Pr, Nd, Sm; diox=1,4‐dioxane) isolated as a 1,4‐dioxane‐bridged dimer (Ln=Ce) or THF adducts [Ln(η3‐C3H5)3(thf)2] (Ln=Ce, Pr). Allyl abstraction from the neutral tris‐allyl complex by a Lewis acid, ER3 (Al(CH2SiMe3)3, BPh3) gave the ion pair [Ln(η3‐C3H5)2(thf)3]+[ER31‐CH2CH?CH2)]? (Ln=Y, La; ER3=Al(CH2SiMe3)3, BPh3). Benzophenone inserts into the La? Callyl bond of [La(η3‐C3H5)2(thf)3]+[BPh4]? to form the alkoxy complex [La{OCPh2(CH2CH?CH2)}2(thf)3]+[BPh4]?. The monocationic half‐sandwich complexes [Ln(η5‐C5Me4SiMe3)(η3‐C3H5)(thf)2]+[B(C6X5)4]? (Ln=Y, La; X=H, F) were synthesized from the neutral precursors [Ln(η5‐C5Me4SiMe3)(η3‐C3H5)2(thf)] by protonolysis. For 1,3‐butadiene polymerization catalysis, the yttrium‐based systems were more active than the corresponding lanthanum or neodymium homologues, giving polybutadiene with approximately 90 % 1,4‐cis stereoselectivity.  相似文献   

11.
The title compounds, [Cr(C12H10)(CO)3] and [Cr2(C12H10)(CO)6], serve as a fundamental standard of comparison for other mono‐ and polysubstituted (η6‐bi­phenyl)­tri­carbonyl­chromium compounds. (η6‐Bi­phenyl)­tri­carbonyl­chromium has a typical piano‐stool coordination about the Cr center, and the dihedral angle between the planes of the phenyl rings is 23.55 (5)°. The corresponding angle in μ‐(η66)‐bi­phenyl‐bis­(tri­carbonyl­chromium) is 0° because the mol­ecule occupies a crystallographic inversion center; the Cr atoms reside on opposite sides of the bi­phenyl ligand. Density functional theory and natural bonding orbital theory analyses were used to scrutinize the geometry of these and closely related compounds to explain important structural features.  相似文献   

12.
Chloride abstraction from the half‐sandwich complexes [RuCl2(η6p‐cymene)(P*‐κP)] ( 2a : P* = (Sa,R,R)‐ 1a = (1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl bis[(1R)‐1‐phenylethyl)]phosphoramidite; 2b : P* = (Sa,R,R)‐ 1b = (1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl bis[(1R)‐(1‐(1‐naphthalen‐1‐yl)ethyl]phosphoramidite) with (Et3O)[PF6] or Tl[PF6] gives the cationic, 18‐electron complexes dichloro(η6p‐cymene){(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl {(1R)‐1‐[(1,2‐η)‐phenyl]ethyl}[(1R)‐1‐phenylethyl]phosphoramidite‐κP}ruthenium(II) hexafluorophosphate ( 3a ) and [Ru(S)]‐dichloro(η6p‐cymene){(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl {(1R)‐1‐[(1,2‐η)‐naphthalen‐1‐yl]ethyl}[(1R)‐1‐(naphthalen‐1‐yl)ethyl]phosphoramidite‐κP)ruthenium(II) hexafluorophosphate ( 3b ), which feature the η2‐coordination of one aryl substituent of the phosphoramidite ligand, as indicated by 1H‐, 13C‐, and 31P‐NMR spectroscopy and confirmed by an X‐ray study of 3b . Additionally, the dissociation of p‐cymene from 2a and 3a gives dichloro{(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl [(1R)‐(1‐(η6‐phenyl)ethyl][(1R)‐1‐phenylethyl]phosphoramidite‐κP)ruthenium(II) ( 4a ) and di‐μ‐chlorobis{(1Sa)‐[1,1′‐binaphthalene]‐2,2′‐diyl [(1R)‐1‐(η6‐phenyl)ethyl][(1R)‐1‐phenylethyl]phosphoramidite‐κP}diruthenium(II) bis(hexafluorophosphate) ( 5a ), respectively, in which one phenyl group of the N‐substituents is η6‐coordinated to the Ru‐center. Complexes 3a and 3b catalyze the asymmetric cyclopropanation of α‐methylstyrene with ethyl diazoacetate with up to 86 and 87% ee for the cis‐ and the trans‐isomers, respectively.  相似文献   

13.
Novel ruthenium (II) complexes were prepared containing 2‐phenyl‐1,8‐naphthyridine derivatives. The coordination modes of these ligands were modified by addition of coordinating solvents such as water into the ethanolic reaction media. Under these conditions 1,8‐naphthyridine (napy) moieties act as monodentade ligands forming unusual [Ru(CO)2Cl21‐2‐phenyl‐1,8‐naphthyridine‐ kN )(η1‐2‐phenyl‐1,8‐naphthyridine‐kN′)] complexes. The reaction was reproducible when different 2‐phenyl‐1,8‐naphthyridine derivatives were used. On the other hand, when dry ethanol was used as the solvent we obtained complexes with napy moieties acting as a chelating ligand. The structures proposed for these complexes were supported by NMR spectra, and the presence of two ligands in the [Ru(CO)2Cl21‐2‐phenyl‐1,8‐naphthyridine‐ kN )(η1‐2‐phenyl‐1,8‐naphthyridine‐kN′)] type complexes was confirmed using elemental analysis. All complexes were tested as catalysts in the hydroformylation of styrene showing moderate activity in N,N′‐dimethylformamide. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

14.
We report the synthesis of [n]manganoarenophanes (n=1, 2) featuring boron, silicon, germanium, and tin as ansa‐bridging elements. Their preparation was achieved by salt‐elimination reactions of the dilithiated precursor [Mn(η5‐C5H4Li)(η6‐C6H5Li)]?pmdta (pmdta=N,N,N′,N′,N′′‐pentamethyldiethylenetriamine) with corresponding element dichlorides. Besides characterization by multinuclear NMR spectroscopy and elemental analysis, the identity of two single‐atom‐bridged derivatives, [Mn(η5‐C5H4)(η6‐C6H5)SntBu2] and [Mn(η5‐C5H4)(η6‐C6H5)SiPh2], could also be determined by X‐ray structural analysis. We investigated for the first time the reactivity of these ansa‐cyclopentadienyl–benzene manganese compounds. The reaction of the distannyl‐bridged complex [Mn(η5‐C5H4)(η6‐C6H5)Sn2tBu4] with elemental sulfur was shown to proceed through the expected oxidative addition of the Sn?Sn bond to give a triatomic ansa‐bridge. The investigation of the ring‐opening polymerization (ROP) capability of [Mn(η5‐C5H4)(η6‐C6H5)SntBu2] with [Pt(PEt3)3] showed that an unexpected, unselective insertion into the Cipso?Sn bonds of [Mn(η5‐C5H4)(η6‐C6H5)SntBu2] had occurred.  相似文献   

15.
The extensive development of radiopharmaceuticals towards early tumour detection and treatment has increased the demand for new ligands with higher tumour selectivity. Research has been done on the potential of the novel O,O′‐diethylethylenediamine‐N,N′‐di‐3‐propanoate ( L ) ligand as a radionuclide vehicle for tumour targeting. Under alkaline conditions, L hydrolyses and produces half ester ligand ( L' ) and diacid ligand ( L'' ), with characteristic donor atom array N,N,O. Ligand L was successfully labelled with 99mTc at pH = 9 by coordination with the octahedral fac‐[99mTc(CO)3(H2O)3]+ intermediate, forming the main radioproduct fac‐[99mTcL′(CO)3] (Tc1). The 99mTc complex showed a low lipophilic character (log P = 0.48) and low binding affinity to human serum albumin (2.51 ± 0.48%). In vitro stability studies in saline and human plasma, as well as challenge studies with cysteine and histidine, revealed high stability of the complex during 24 h. Biodistribution studies of Tc1 in female C57BL/6 mice bearing B16/F1 melanoma metastases showed significant tumour uptake: 9.81 ± 1.19%ID g?1 in the liver, 5.87 ± 0.54%ID g?1 in the lungs and 3.17 ± 0.33%ID g?1 in the ovary at 30 min post‐injection. Favourable physicochemical properties, satisfactory in vitro/in vivo stability and biodistribution profile in the experimental metastatic melanoma model indicate the possible application of the radiolabelled ligand in tumour diagnosis. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

16.
The confirmation that N-substituted imidodiacetic acids, as small and simple ligand systems containing amines and carboxylic acids, could be coordinated to the tricarbonyl core and form inert complexes with [99mTc (CO)3(H2O)3]+, is demonstrated. The HPLC quality control results of 99mTc-carbonyl tagged IDA molecules, performed by gradient HPLC, have shown that HIDA, EHIDA and p-butyl-IDA form complexes with [99mTc(CO)3(H2O)3]+, with a labeling yield of ~90% for each of 99mTc(CO)3 IDA derivatives. However, the changes in the structure of labeled compounds, e.g., EHIDA, influence the changes in the biological behavior. In comparison with 99mTc-EHIDA, the biliary excretion of 99mTc(CO)3 EHIDA was lower, but the urinary excretion higher. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

17.
The synthesis of ansa complexes has been studied intensively owing to their importance as homogeneous catalysts and as precursors of metal‐containing polymers. However, paramagnetic non‐metallocene derivatives are rare and have been limited to examples with vanadium and titanium. Herein, we report an efficient procedure for the selective dilithiation of paramagnetic sandwich complex [Cr(η5‐C5H5)(η6‐C6H6)], which allows the preparation of a series of [n]chromoarenophanes (n=1, 2, 3) that feature silicon, germanium, and tin atoms at the bridging positions. The electronic and structural properties of these complexes were probed by X‐ray diffraction analysis, cyclic voltammetry, and by UV/Vis and EPR spectroscopy. The spectroscopic parameters for the strained and less strained complexes (i.e., with multiple‐atom linkers) indicate that the unpaired electron resides primarily in a d orbital on chromium(I); this result was also supported by density functional theory (DFT) calculations. We did not observe a correlation between the experimental UV/Vis and EPR data and the degree of molecular distortion in these ansa complexes. The treatment of tin‐bridged complex [Cr(η5‐C5H4)(η6‐C6H5)SntBu2] with [Pt(PEt3)3] results in the non‐regioselective insertion of the low‐valent Pt0 fragment into the Cipso? Sn bonds in both the five‐ and six‐membered rings, thereby furnishing a bimetallic complex. This observed reactivity suggests that ansa complexes of this type are promising starting materials for the synthesis of bimetallic complexes in general and also underline their potential to undergo ring‐opening processes to yield new metal‐containing polymers.  相似文献   

18.
Crystallization experiments with the dinuclear chelate ring complex di‐μ‐chlorido‐bis[(η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)platinum(II)], [Pt2(C15H19O4)2Cl2], containing a derivative of the natural compound eugenol as ligand, have been performed. Using five different sets of crystallization conditions resulted in four different complexes which can be further used as starting compounds for the synthesis of Pt complexes with promising anticancer activities. In the case of vapour diffusion with the binary chloroform–diethyl ether or methylene chloride–diethyl ether systems, no change of the molecular structure was observed. Using evaporation from acetonitrile (at room temperature), dimethylformamide (DMF, at 313 K) or dimethyl sulfoxide (DMSO, at 313 K), however, resulted in the displacement of a chloride ligand by the solvent, giving, respectively, the mononuclear complexes (acetonitrile‐κN)(η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)chloridoplatinum(II) monohydrate, [Pt(C15H19O4)Cl(CH3CN)]·H2O, (η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)chlorido(dimethylformamide‐κO)platinum(II), [Pt(C15H19O4)Cl(C2H7NO)], and (η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)chlorido(dimethyl sulfoxide‐κS)platinum(II), determined as the analogue {η2‐2‐allyl‐4‐methoxy‐5‐[(ethoxycarbonyl)methoxy]phenyl‐κC1}chlorido(dimethyl sulfoxide‐κS)platinum(II), [Pt(C14H17O4)Cl(C2H6OS)]. The crystal structures confirm that acetonitrile interacts with the PtII atom via its N atom, while for DMSO, the S atom is the coordinating atom. For the replacement, the longest of the two Pt—Cl bonds is cleaved, leading to a cis position of the solvent ligand with respect to the allyl group. The crystal packing of the complexes is characterized by dimer formation via C—H…O and C—H…π interactions, but no π–π interactions are observed despite the presence of the aromatic ring.  相似文献   

19.
The reaction of 99mTc of different oxidation states (+7, +4) with 2-thiouracil and 5-nitrobarbituric acid have been studied at different temperatures, pH and concentrations. The reaction mixtures have been analyzed at different times using thin layer chromatography (TLC) and a radio detector to show the peaks at the plates. 99mTc is obtained from the Mo generators with oxidation state (+7). The use of SnCl2 as a reducing agent gave 99mTc with oxidation state (+4). It is very difficult to separate the complexes formed from the reactions in very small concentration. The percentage of 99mTc and its oxidation state involved in the complexes can be determined. The labeling efficiencies (percentage of complex) in the reaction of 99mTc+7 with 5-nitro-barbituric-acid increases mostly at pH  10. Both oxidation states of 99mTc(+7, +4) can be detected at pH’s 4 and 10, but at pH  4, the reduced form 99mTCO2, is more pronounced. At pH  7 no complexes were detected and most of 99mTc remains as 99mTCO4 . By increasing the ligand concentration, the labeling efficiencies of the complex increases. For the reaction of 99mTc of oxidation states (+4, +7) with 2-thiouracil at different temperatures and analytical times it is concluded that several complexes with different Rf values were observed in equilibrium and most of these complexes were unstable.  相似文献   

20.
This study focuses on a series of cationic complexes of iridium that contain aminopyridinate (Ap) ligands bound to an (η5‐C5Me5)IrIII fragment. The new complexes have the chemical composition [Ir(Ap)(η5‐C5Me5)]+, exist in the form of two isomers ( 1+ and 2+ ) and were isolated as salts of the BArF? anion (BArF=B[3,5‐(CF3)2C6H3]4). Four Ap ligands that differ in the nature of their bulky aryl substituents at the amido nitrogen atom and pyridinic ring were employed. In the presence of H2, the electrophilicity of the IrIII centre of these complexes allows for a reversible prototropic rearrangement that changes the nature and coordination mode of the aminopyridinate ligand between the well‐known κ2‐N,N′‐bidentate binding in 1+ and the unprecedented κ‐N3‐pseudo‐allyl‐coordination mode in isomers 2+ through activation of a benzylic C?H bond and formal proton transfer to the amido nitrogen atom. Experimental and computational studies evidence that the overall rearrangement, which entails reversible formation and cleavage of H?H, C?H and N?H bonds, is catalysed by dihydrogen under homogeneous conditions.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号