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 46 毫秒
1.
A derivative of H5ttda (=3,6,10‐tris(carboxymethyl)‐3,6,10‐triazadodecanedioic acid=N‐{2‐[bis(carboxymethyl)amino]ethyl}‐N‐{3‐[bis(carboxymethyl)amino]propyl}glycine), H5[(S)‐4‐Bz‐ttda] (=(4S)‐4‐benzyl‐3,6,10‐tris(carboxymethyl)‐3,6,10‐triazadodecanedioic acid=N‐{(2S)‐2‐[bis(carboxymethyl)amino]‐3‐phenylpropyl}‐N‐{3‐[bis(carboxymethyl)amino]propyl}glycine; 1 ) carrying a benzyl group was synthesized and characterized. The stability constants of the complexes formed with Ca2+, Zn2+, Cu2+, and Gd3+ were determined by potentiometric methods at 25.0±0.1° and 0.1M ionic strength in Me4NNO3. The observed water proton relaxivity value of [Gd{(S)‐4‐Bz‐ttda}]2− was constant with respect to pH changes over the range pH 4.5–12.0. From the 17O‐NMR chemical shift of H2O induced by [Dy{(S)‐4‐Bz‐ttda}]2− at pH 6.80, the presence of 0.9 inner‐sphere water molecules was deduced. The water proton spin‐lattice relaxation rate for [Gd{(S)‐4‐Bz‐ttda}]2− at 37.0±0.1° and 20 MHz was 4.90±0.05 mM −1 s−1. The EPR transverse electronic relaxation rate and 17O‐NMR transverse‐relaxation time for the exchange lifetime of the coordinated H2O molecule (τM), and 2H‐NMR longitudinal‐relaxation rate of the deuterated diamagnetic lanthanum complex for the rotational correlation time (τR) were thoroughly investigated, and the results were compared with those previously reported for the other lanthanide(III) complexes. The exchange lifetime (τM) for [Gd{(S)‐4‐Bz‐ttda}]2− (2.3±1.3 ns) was significantly shorter than that of the [Gd(dtpa)(H2O)]2− complex (dtpa=diethylenetriaminepentaacetic acid). The rotational correlation time τR for [Gd{(S)‐4‐Bz‐ttda}]2− (70±6 ps) was slightly longer than that of the [Gd(dtpa)(H2O)]2− complex. The marked increase of relaxivity of [Gd{(S)‐4‐Bz‐ttda}]2− mainly resulted from its longer rotational time rather than from its fast water‐exchange rate. The noncovalent interaction between human serum albumin (HSA) and the [Gd{(S)‐4‐Bz‐ttda}]2− complex containing the hydrophobic substituent was investigated by measuring the solvent proton relaxation rate of the aqueous solutions. The association constant (KA) was less than 100 M −1, indicating a weaker interaction of [Gd{(S)‐4‐Bz‐ttda}]2− with HSA.  相似文献   

2.
To confirm the observation that [Gd(ttda)] derivatives have a significantly shorter residence time τM of the coordinated H2O molecule than [Gd(dtpa)], four new C‐functionalized [Gd(ttda)] complexes, [Gd(4‐Me‐ttda)] ( 1 ), [Gd(4‐Ph‐ttda)] ( 2 ), [Gd(9‐Me‐ttda)] ( 3 ), and [Gd(9‐Ph‐ttda)] ( 4 ), were prepared and characterized (H5ttda=3,6,10‐tris(carboxymethyl)‐3,6,10‐triazadodecanedioic acid; H5dtpa=3,6,9‐tris(carboxymethyl)‐3,6,9‐triazaundecanedioic acid). The temperature dependence of the proton relaxivity for these complexes at 0.47 T and of the 17O transverse relaxation rate of H217O at 7.05 T confirm that the proton relaxivity is not limited by the H2O‐exchange rate. The residence time of the H2O molecules in the first coordination sphere of the gadolinium complexes at 310 K, as calculated from 17O‐NMR data, is 13, 43, 2.9, and 56 ns for 1, 2, 3 , and 4 , respectively. At 310 K, the longitudinal relaxivity of 2 is higher than for the parent compound [Gd(ttda)] and the other complexes of the series. The stability of the new compounds was studied by transmetallation with Zn2+ ions. All the new complexes are more stable than the parent compound [Gd(ttda)].  相似文献   

3.
Two N‐2‐hydroxy‐1‐phenylethyl and N‐2‐hydroxy‐2‐phenylethyl derivatives of DTPA (3,6,9‐tri(carboxymethyl)‐3,6,9‐triazaundecanedioic acid), DTPA‐H1P = 3,9‐di(carboxymethyl)‐6‐2‐hydroxy‐1‐phenylethyl‐3,6,9‐triazaundecanedioic acid, and DTPA‐H2P = 3,9‐di(carboxymethyl)‐6‐2‐hydroxy‐2‐phenylethyl‐3,6,9‐triazaundecanedioic acid were synthesized. Their protonation constants were determined by Potentiometric titration in 0.10 M Me4NNO3 and by NMR pH titration at 25.0 ± 0.1°C. The formations of lanthanide(III), copper(II), zinc(II) and calcium(II) complexes were investigated quantitatively by potentiometry. The stability constant for Gd(III) complex is larger than those for Ca(II), Zn(II) and Cu(II) complexes with these two ligands. The selectivity constants and modified selectivity constants of the DTPA‐H1P and DTPA‐H2P for Gd(III) over endogenously available metal ions were calculated. Comparing pM values at physiological pH 7.4 assesses effectiveness of these two ligands in binding divalent and trivalent metal ions in biological media. The observed water proton relaxivity values of [Gd(DTPA‐H1P)]? and [Gd(DTPA‐H2P)]? became constant with respect to pH changes over the range of 4‐10. 17O NMR shifts showed that the [Dy(DTPA‐H1P)]? and [Dy(DTPA‐H2P)]? complexes at pH 6.30 had 1.91 and 2.28 inner‐sphere water molecules, respectively. Water proton spin‐lattice relaxation rates of [Gd(DTPA‐H1P)]? and [Gd(DTPA‐H2P)]? complexes were also consistent with the inner‐sphere Gd(III) coordination.  相似文献   

4.
《Polyhedron》1999,18(8-9):1147-1152
A new bis(amide) derivative of diethylenetriamine-N,N,N′,N″,N″-pentaacetic acid (H5dtpa), diethylenetriamine-N,N′,N″-triacetic-N,N″-bis(2-methoxyphenethylamide) (H3L), has been synthesized. The crystal structure of gadolinium(III) complex of H3L ([GdL]) has been determined by X-ray crystallography. The coordination sphere of Gd(III) comprises three amine nitrogens, two amide oxygens, three carboxylic acid oxygens, and one water molecule. 17O NMR shifts showed that the [DyL] complex had one inner-sphere water molecule. Relaxivity studies with the gadolinium(III) complex showed that they associate strongly with proteins leading to a significant relaxivity enhancement.  相似文献   

5.
The analysis of 17O NMR transverse relaxation rates and EPR transverse electronic relaxation rates for aqueous solutions of the four DTPA‐like (DTPA = diethylenetriamine‐N,N,N,N″,N″‐pentaacetic acid) complexes, [Gd(DTPA‐PY)(H2O)]? (DTPA‐PY = N′‐(2‐pyridylmethyl)), [Gd(DTPA‐HP)(H2O)2]? (DTPA‐HP = N′‐(2‐hydroxypropyl)), [Gd(DTPA‐H1P)(H2O)2]? (DTPA‐H1P = N′‐(2‐hydroxy‐1‐phenylethyl)) and [Gd(DTPA‐H2P)(H2O)2] (DTPA‐H2P = N′‐(2‐hydroxy‐2‐phenylethyl)), at various temperatures allows us to understand the water exchange dynamics of these four complexes. The water‐exchange lifetime (τM) parameters for [Gd(DTPA‐PY)(H2O)]?, [Gd(DTPA‐HP)(H2O)2]?, [Gd(DTPA‐H1P)(H2O)2]? and [Gd(DTPA‐H2P)(H2O)2] are of 585, 98, 163, and 69 ns, respectively. Compared with [Gd(DTPA)(H2O)]2? (τM = 303 ns), the τM value of [Gd(DTPA‐PY)(H2O)]? is slightly higher, but the other three complexes values are significantly lower than those of [Gd(DTPA)(H2O)]2?. This difference is explained by the fact that the gadolinium(III) complexes of DTPA‐HP, DTPA‐H1P, and DTPA‐H2P have two inner‐sphere waters. The 2H longitudinal relaxation rates of the labeled diamagnetic lanthanum complex allow the calculation of its rotational correlation time (τR). The τR values calculated for DTPA‐PY, DTPA‐HP, DTPA‐H1P, and DTPA‐H2P are of 127, 110, 142 and 147 ps, respectively. These four values are higher than the value of [La(DTPA)]2? (τR = 103 ps), because the rotational correlation time is related to the magnitude of its molecular weight.  相似文献   

6.
Three cyano‐bridged aqua(N,N‐dimethylacetamide)(cyanoiron)lanthanide complexes were synthesized by the reaction of K3Fe(CN)6, Ln(NO3)3⋅6 H2O (Ln=Sm, Gd, Ho), and N,N‐dimethylacetamide (DMA). The obtained complexes 1 – 3 exhibit different coordination geometries and crystal structures. The polymeric {[Sm(DMA)2(H2O)4Fe(CN)6⋅5 H2O}n ([SmFe]n; 1 ) has a one‐dimensional chain structure with approximately parallel trans‐positioned bridging CN ligands between the Sm‐ and Fe‐atoms. [(Gd(DMA)3(H2O)4)2Fe(CN)6]⋅[Fe(CN)6]⋅3 H2O (Gd2Fe; 2 ) is an isolated trinuclear Gd(1)−Fe−Gd(2) complex with two approximately perpendicular cis‐positioned bridging CN ligands between the two Gd‐atoms and the Fe‐atom. [Ho(DMA)3(H2O)3Fe(CN)6]⋅3 H2O (HoFe; 3 ) adopts a single dinuclear crystal structure with only one bridging CN between the Ho‐ and Fe‐atom. Magnetochemistry experiments establish weak antiferromagnetic interactions between GdIII (and HoIII) and FeIII atoms. Especially the [SmFe]n complex 1 exhibits long‐range magnetic ordering, Tc=3.5 K, and a stronger coercive force, Hc=1400 Oe.  相似文献   

7.
The kinetics of oxidation of [CrIIIcdta(H2O)]? and [CrIIIdtpa(H2O)]2? (where cdta = trans‐1,2‐diaminocyclohexane‐N,N,N′,N′‐tetraacetate and dtpa = diethylenetriaminepentaacetate) by periodate ion has been studied in aqueous solutions. The oxidation of these complexes was carried out in the pH range 5.52–7.44 for the [CrIIIcdta(H2O)]? complex and the pH range 5.56–8.56 for the [CrIIIdtpa(H2O)]2? complex. The reaction exhibited an uncommon second‐order dependence on [CrIIIL(H2O)]n (L = cdta or dtpa and n=?1 or ?2, respectively) and a first‐order dependence on [IO?4]. At fixed reaction conditions, the reaction rate is described by Eq. (i). The third‐order rate constant, k3, varied with [H+] according to Eq. (ii). (i) (ii) A mechanism in which simultaneous one‐electron transfer from two [CrIIIL(OH)]n?1 ions to I(VII) is proposed. The two [CrIIIL(OH)]n?1 ions are bridged to I(VII) via the hydroxo group. Periodate ion is known to undergo rapid substitution or expansion of its coordination number from four to six. The activation parameters ΔH* and ΔS* were calculated using the Eyring equation. The relatively high negative values of ΔS* are consistent with an associative process preceding electron transfer. © 2012 Wiley Periodicals, Inc. Int J Chem Kinet 44: 729–735, 2012  相似文献   

8.
The new bis(amide) derivatives of DTPA (diethylenetriamine-N,N,N′,N″,N″-pentaacetic acid), diethylenetriamine-N,N′,N″-triacetic-N,N″-bis(benzylamide) (DTPA-BBA) have been synthesized. The crystal structure of gadolinium(III) complex of DTPA-BBA ([Gd(DTPA-BBA)]) has been determined by X-ray crystallography: C28H52GdN5O17, Mw = 889 g mol?1, space group $ {\rm P}\bar 1 $ (#2) (triclinic), a = 12.645(4), b=14.125(8), c = 12.623(4) Å, α = 111.60(3), β = 114.79(3), γ = 88.39(4)°, V = 1881(1) Å3, Z = 2, Dx = 1.569 g/cm3, λ(Mo Kα) =0.71069 Å, μ = 18.44 cm?1, final R = 0.047, Rw = 0.046 for 3755 independent observed reflections at 23 °C. The coordination sphere of Gd(III) comprises three amine nitrogens, two amide oxygens, three carboxylic acid oxygens, and one water. The relaxivity of Gd(III) complex was determined to be R1 = 4.08(4) and R2 = 6.06(5) dm3 mmol?1 s?1 at pH = 7.0, 20 MHz, and 37(1) °C. Additionally, the R1 relaxivity for Gd(III) chelate was found to be invariant with respect to pH changes over the range of 2-10. This indicates that a constant inner-sphere hydration number is associated with the [Gd(DTPA-BBA)] complex. Hence the high stability of the complex is demonstrated.  相似文献   

9.
Two rare-earth metal coordination compounds, (NH4)4[SmIII2(Httha)2]·16H2O (1) (H6ttha?=?triethylenetetramine-N,N,N,N′′,N′′′,N′′′-hexaacetic acid) and (NH4)4[SmIII2(dtpa)2]·10H2O (2) (H5dtpa?=?diethylenetriamine-N,N,N,N′′,N′′-pentaacetic acid), have been synthesized through reflux and characterized by FT-IR spectroscopy, thermal analysis, and single-crystal X-ray diffraction techniques. SmIII of (NH4)4[SmIII2(Httha)2]·16H2O (1) is nine-coordinate, forming tricapped trigonal prismatic coordination with three amine nitrogens and six oxygens, in which four oxygens are from one ttha and two from the other ttha. (NH4)4[SmIII2(Httha)2]·16H2O (1) crystallizes in the monoclinic crystal system with P2(1)/c space group. The crystal data are: a?=?13.9340(13) Å, b?=?22.890(3) Å, c?=?20.708(2) (14) Å, β?=?99.521(2)°, and V?=?6513.7(13) Å3. There are two –NH+– groups in the [SmIII2(Httha)2]4?. The polymeric (NH4)4[SmIII2(dtpa)2]·10H2O (2) also is nine-coordinate with tricapped trigonal prismatic conformation and crystallizes in the triclinic crystal system with P–1 space group. The cell dimensions are: a?=?9.8240(8) Å, b?=?10.0329(9) Å, c?=?13.0941(11) Å, β?=?77.1640(10)°, and V?=?1227.30(18) Å3. In (NH4)4[SmIII2(dtpa)2]·10H2O, there are two types of ammonium cations, which connect [SmIII2(dtpa)2]4? and lattice water through hydrogen bonds, leading to a 2-D ladder-like layer structure.  相似文献   

10.
The two isomorphous title compounds, [M(C5H7N6)2(C9H6O4)2(H2O)2]·4H2O or M2+(Hdap+)2(hpt2−)2(H2O)2·4H2O {where dap is 2,6‐diaminopurine, H2hpt is homophthalic acid [2‐(2‐carboxyphenyl)acetic acid] and M is NiII or CoII}, consist of neutral M2+(Hdap+)2(hpt2−)2(H2O)2 monomers, where the MII cation lies on an inversion centre and its MN2O4 octahedral environment is defined by one N atom (from Hdap+), two O atoms (from one hpt2− dianion and one water molecule) and their inversion images. The structures are unusual in that the Hdap+ cation occurs in an uncommon protonated state (as 2,6‐diamino‐7H‐purin‐1‐ium) and both ligands bind in an unprecedented monodentate fashion. The existence of a large number of donors and acceptors for hydrogen bonding, together with π–π interactions, leads to a rather complex three‐dimensional structure.  相似文献   

11.
In the title compound, [Mn(C5H2N2O4)(C12H9N3)2]·H2O, the MnII centre is surrounded by three bidentate chelating ligands, namely, one 6‐oxido‐2‐oxo‐1,2‐dihydropyrimidine‐5‐carboxylate (or uracil‐5‐carboxylate, Huca2−) ligand [Mn—O = 2.136 (2) and 2.156 (3) Å] and two 2‐(2‐pyridyl)‐1H‐benzimidazole (Hpybim) ligands [Mn—N = 2.213 (3)–2.331 (3) Å], and it displays a severely distorted octahedral geometry, with cis angles ranging from 73.05 (10) to 105.77 (10)°. Intermolecular N—H...O hydrogen bonds both between the Hpybim and the Huca2− ligands and between the Huca2− ligands link the molecules into infinite chains. The lattice water molecule acts as a hydrogen‐bond donor to form double O...H—O—H...O hydrogen bonds with the Huca2− O atoms, crosslinking the chains to afford an infinite two‐dimensional sheet; a third hydrogen bond (N—H...O) formed by the water molecule as a hydrogen‐bond acceptor and a Hpybim N atom further links these sheets to yield a three‐dimensional supramolecular framework. Possible partial π–π stacking interactions involving the Hpybim rings are also observed in the crystal structure.  相似文献   

12.
Coordination polymers are a thriving class of functional solid‐state materials and there have been noticeable efforts and progress toward designing periodic functional structures with desired geometrical attributes and chemical properties for targeted applications. Self‐assembly of metal ions and organic ligands is one of the most efficient and widely utilized methods for the construction of CPs under hydro(solvo)thermal conditions. 2‐(Pyridin‐3‐yl)‐1H‐imidazole‐4,5‐dicarboxylate (HPIDC2−) has been proven to be an excellent multidentate ligand due to its multiple deprotonation and coordination modes. Crystals of poly[aquabis[μ3‐5‐carboxy‐2‐(pyridin‐3‐yl)‐1H‐imidazole‐4‐carboxylato‐κ5N1,O5:N3,O4:N2]copper(II)dicopper(I)], [CuIICuI2(C10H5N3O4)2(H2O)]n, (I), were obtained from 2‐(pyridin‐3‐yl)‐1H‐imidazole‐4,5‐dicarboxylic acid (H3PIDC) and copper(II) chloride under hydrothermal conditions. The asymmetric unit consists of one independent CuII ion, two CuI ions, two HPIDC2− ligands and one coordinated water molecule. The CuII centre displays a square‐pyramidal geometry (CuN2O3), with two N,O‐chelating HPIDC2− ligands occupying the basal plane in a trans geometry and one O atom from a coordinated water molecule in the axial position. The CuI atoms adopt three‐coordinated Y‐shaped coordinations. In each [CuN2O] unit, deprotonated HPIDC2− acts as an N,O‐chelating ligand, and a symmetry‐equivalent HPIDC2− ligand acts as an N‐atom donor via the pyridine group. The HPIDC2− ligands in the polymer serve as T‐shaped 3‐connectors and adopt a μ3‐κ2N,O2N′,O′:κN′′‐coordination mode, linking one CuII and two CuI cations. The Cu cations are arranged in one‐dimensional –Cu1–Cu2–Cu3– chains along the [001] direction. Further crosslinking of these chains by HPIDC2− ligands along the b axis in a –Cu2–HPIDC2−–Cu3–HPIDC2−–Cu1– sequence results in a two‐dimensional polymer in the (100) plane. The resulting (2,3)‐connected net has a (123)2(12)3 topology. Powder X‐ray diffraction confirmed the phase purity for (I), and susceptibilty measurements indicated a very weak ferromagnetic behaviour. A thermogravimetric analysis shows the loss of the apical aqua ligand before decomposition of the title compound.  相似文献   

13.
The synthesis and structural characterization of 2‐(furan‐2‐yl)‐1‐(furan‐2‐ylmethyl)‐1H‐benzimidazole [C16H12N2O2, (I)], 2‐(furan‐2‐yl)‐1‐(furan‐2‐ylmethyl)‐1H‐benzimidazol‐3‐ium chloride monohydrate [C16H13N2O2+·Cl·H2O, (II)] and the hydrobromide salt 5,6‐dimethyl‐2‐(furan‐2‐yl)‐1‐(furan‐2‐ylmethyl)‐1H‐benzimidazol‐3‐ium bromide [C18H17N2O2+·Br, (III)] are described. Benzimidazole (I) displays two sets of aromatic interactions, each of which involves pairs of molecules in a head‐to‐tail arrangement. The first, denoted set (Ia), exhibits both intermolecular C—H...π interactions between the 2‐(furan‐2‐yl) (abbreviated as Fn) and 1‐(furan‐2‐ylmethyl) (abbreviated as MeFn) substituents, and π–π interactions involving the Fn substituents between inversion‐center‐related molecules. The second, denoted set (Ib), involves π–π interactions involving both the benzene ring (Bz) and the imidazole ring (Im) of benzimidazole. Hydrated salt (II) exhibits N—H...OH2...Cl hydrogen bonding that results in chains of molecules parallel to the a axis. There is also a head‐to‐head aromatic stacking of the protonated benzimidazole cations in which the Bz and Im rings of one molecule interact with the Im and Fn rings of adjacent molecules in the chain. Salt (III) displays N—H...Br hydrogen bonding and π–π interactions involving inversion‐center‐related benzimidazole rings in a head‐to‐tail arrangement. In all of the π–π interactions observed, the interacting moieties are shifted with respect to each other along the major molecular axis. Basis set superposition energy‐corrected (counterpoise method) interaction energies were calculated for each interaction [DFT, M06‐2X/6‐31+G(d)] employing atomic coordinates obtained in the crystallographic analyses for heavy atoms and optimized H‐atom coordinates. The calculated interaction energies are −43.0, −39.8, −48.5, and −55.0 kJ mol−1 for (Ia), (Ib), (II), and (III), respectively. For (Ia), the analysis was used to partition the interaction energies into the C—H...π and π–π components, which are 9.4 and 24.1 kJ mol−1, respectively. Energy‐minimized structures were used to determine the optimal interplanar spacing, the slip distance along the major molecular axis, and the slip distance along the minor molecular axis for 2‐(furan‐2‐yl)‐1H‐benzimidazole.  相似文献   

14.
Three potassium edta (edta is ethylenediaminetetraacetic acid, H4Y) salts which have different degrees of ionization of the edta anion, namely dipotassium 2‐({2‐[bis(carboxylatomethyl)azaniumyl]ethyl}(carboxylatomethyl)azaniumyl)acetate dihydrate, 2K+·C10H14N2O82−·2H2O, (I), tripotassium 2,2′‐({2‐[bis(carboxylatomethyl)amino]ethyl}ammonio)diacetate dihydrate, 3K+·C10H13N2O83−·2H2O, (II), and tetrapotassium 2,2′,2′′,2′′′‐(ethane‐1,2‐diyldinitrilo)tetraacetate 3.92‐hydrate, 4K+·C10H12N2O84−·3.92H2O, (III), were obtained in crystalline form from water solutions after mixing edta with potassium hydroxide in different molar ratios. In (II), a new mode of coordination of the edta anion to the metal is observed. The HY3− anion contains one deprotonated N atom coordinated to K+ and the second N atom is involved in intramolecular bifurcated N—H...O and N—H...N hydrogen bonds. The overall conformation of the HY3− anions is very similar to that of the Y4− anions in (III), although a slightly different spatial arrangement of the –CH2COO groups in relation to (III) is observed, whereas the H2Y2− anions in (I) adopt a distinctly different geometry. The preferred synclinal conformation of the –NCH2CH2N– moiety was found for all edta anions. In all three crystals, the anions and water molecules are arranged in three‐dimensional networks linked via O—H...O and C—H...O [and N—H...O in (I) and (II)] hydrogen bonds. K...O interactions also contribute to the three‐dimensional polymeric architecture of the salts.  相似文献   

15.
16.
A multinuclear NMR study on [Ln(ttha)]3? and [Ln{ttha(NHR)2}]? complexes (R=Et, CH2(CHOH)4CH2OH) shows that coordinating groups of the organic ligands in these complexes are occupying all coordination sites of the metal ions, leaving no space for coordination of H2O molecules (H6ttha=triethylenetetramine‐N,N,N′,N″,N′′′,N′′′‐hexaacetic acid). The lanthanides of the first half of the series bind the ttha‐type ligands in a decadentate fashion, while the complexes formed with the smaller ions of the second half of the lanthanide series are nonadentate. One carboxylate group of the ligand remains unbound in the latter complexes. In principle, the ttha complexes can exist in six enantiomeric forms. Only one of the pair of diastereoisomers can interconvert without decoordination of the ligand. This pair of isomers seems to be predominant in solution. For the [Ln{ttha(NHR)2}]? complexes, the number of chiral centers is larger, resulting in 32 possible enantiomeric forms of the complexes. The NMR spectra of [Nd{ttha(NHEt)2}]? indicate that two dynamic processes occur between the isomers in solution. The NMRD curves of [Gd(ttha)]3?, [Gd{ttha(NHEt)2}]?, and [Gd{ttha(NHgluca)2}]? (NHgluca=D ‐glucamine) show significant differences with the previously determined outer‐sphere contributions to the NMRD profiles of the corresponding [Gd{dtpa(NHR)2}]? complexes, which can be ascribed to differences in the parameters determining the electronic relaxation.  相似文献   

17.
The novel title hybrid isomorphous organic–inorganic mixed‐metal dichromates, [Ni(Cr2O7)(C10H8N2)2] and [Cu(Cr2O7)(C10H8N2)2], have been synthesized. A non‐centrosymmetric three‐dimensional (4,6)‐net is formed from a linear chain of vertex‐linked [Cr2O]2− and [MN4O]2+ (M = Ni and Cu) units, which in turn are linked by the planar bidentate 4,4′‐­bipyridine ligand through the four remaining vertices of the [MN4O]2+ octahedra. There are two such three‐dimensional nets that interpenetrate with inversion symmetry.  相似文献   

18.
In the title salt, (C6H8N4)[Mn(C14H8O4)2(C6H6N4)2]·6H2O, the MnII atom lies on an inversion centre and is coordinated by four N atoms from two 2,2′‐biimidazole (biim) ligands and two O atoms from two biphenyl‐2,4′‐dicarboxylate (bpdc) anions to give a slightly distorted octahedral coordination, while the cation lies about another inversion centre. Adjacent [Mn(bpdc)2(biim)2]2− anions are linked via two pairs of N—H...O hydrogen bonds, leading to an infinite chain along the [100] direction. The protonated [H2biim]2+ moiety acts as a charge‐compensating cation and space‐filling structural subunit. It bridges two [Mn(bpdc)2(biim)2]2− anions through two pairs of N—H...O hydrogen bonds, constructing two R22(9) rings, leading to a zigzag chain in the [2] direction, which gives rise to a ruffled set of [H2biim]2+[Mn(bpdc)2(biim)2]2− moieties in the [01] plane. The water molecules give rise to a chain structure in which O—H...O hydrogen bonds generate a chain of alternating four‐ and six‐membered water–oxygen R42(8) and R66(12) rings, each lying about independent inversion centres giving rise to a chain along the [100] direction. Within the water chain, the (H2O)6 water rings are hydrogen bonded to two O atoms from two [Mn(bpdc)2(biim)2]2− anions, giving rise to a three‐dimensional framework.  相似文献   

19.
Two new NiII complexes involving the ancillary ligand bis[(pyridin‐2‐yl)methyl]amine (bpma) and two different carboxylate ligands, i.e. homophthalate [hph; systematic name: 2‐(2‐carboxylatophenyl)acetate] and benzene‐1,2,4,5‐tetracarboxylate (btc), namely catena‐poly[[aqua{bis[(pyridin‐2‐yl)methyl]amine‐κ3N,N′,N′′}nickel(II)]‐μ‐2‐(2‐carboxylatophenyl)aceteto‐κ2O:O′], [Ni(C9H6O4)(C12H13N3)(H2O)]n, and (μ‐benzene‐1,2,4,5‐tetracarboxylato‐κ4O1,O2:O4,O5)bis(aqua{bis[(pyridin‐2‐yl)methyl]amine‐κ3N,N′,N′′}nickel(II)) bis(triaqua{bis[(pyridin‐2‐yl)methyl]amine‐κ3N,N′,N′′}nickel(II)) benzene‐1,2,4,5‐tetracarboxylate hexahydrate, [Ni2(C10H2O8)(C12H13N3)2(H2O)2]·[Ni(C12H13N3)(H2O)3]2(C10H2O8)·6H2O, (II), are presented. Compound (I) is a one‐dimensional polymer with hph acting as a bridging ligand and with the chains linked by weak C—H...O interactions. The structure of compound (II) is much more complex, with two independent NiII centres having different environments, one of them as part of centrosymmetric [Ni(bpma)(H2O)]2(btc) dinuclear complexes and the other in mononuclear [Ni(bpma)(H2O)3]2+ cations which (in a 2:1 ratio) provide charge balance for btc4− anions. A profuse hydrogen‐bonding scheme, where both coordinated and crystal water molecules play a crucial role, provides the supramolecular linkage of the different groups.  相似文献   

20.
In 2‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐1,2,4‐triazine‐3,5(2H,4H)‐dione (6‐aza‐2′‐deoxy­uridine), C8H11N3O5, (I), the conformation of the glycosylic bond is between anti and high‐anti [χ = −94.0 (3)°], whereas the derivative 2‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐N4‐(2‐methoxy­benzoyl)‐1,2,4‐triazine‐3,5(2H,4H)‐dione (N3‐anisoyl‐6‐aza‐2′‐deoxy­uridine), C16H17N3O7, (II), displays a high‐anti conformation [χ = −86.4 (3)°]. The furanosyl moiety in (I) adopts the S‐type sugar pucker (2T3), with P = 188.1 (2)° and τm = 40.3 (2)°, while the sugar pucker in (II) is N (3T4), with P = 36.1 (3)° and τm = 33.5 (2)°. The crystal structures of (I) and (II) are stabilized by inter­molecular N—H⋯O and O—H⋯O inter­actions.  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