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.
《Inorganica chimica acta》2006,359(5):1421-1426
Synthesis, structural characterization, and spectroscopic and magnetic properties of three new cyano-bridged 3d–4f bimetallic complexes, LnIII(DMF)4(H2O)3CrIII (CN)6 · nH2O (Ln = Nd, Sm, Gd), have been described. The Nd–Cr complex crystallizes in the monoclinic P21/n space group with the following unit cell parameters: a = 20.063(7) Å, b = 8.967(4) Å, c = 18.023(6) Å, b = 96.12(3)°, V = 3224(2) Å3, and Z = 4. The neodymium (III) ion, which adopts anti-prism eight-coordination environment, is linked to the [CrIII(CN)6]3− moiety through a bridging cyanide ligand with Nd–N = 2.550(4) Å and Nd–N–C = 164.4(4)°. The variable-temperature (0.5 T at 2–300 K) and variable-field (0–5 T at 2 and 5 K) magnetic measurements reveal that the weak interaction of Gd–Cr complexes differs from that of Nd–Cr and Sm–Cr ones mainly because of the lack of orbital angular momentum. The XPS and diffuse reflectance electronic spectra were also measured to discuss charge transfer transitions concerning π-backdonation from the viewpoint of magneto-optical functions.  相似文献   

2.
The adhesive domain of SdrD from Staphylococcus aureus was solubly expressed in Escherichia coli in high yield. After a series of purification steps, the purified protein was >95% pure, which was SdrD from S. aureus identified by SDS–PAGE and MALDI-TOF MS. Crystals were grown at 18 °C using 25% polyethylene glycol 3350 as precipitant. Diffraction by the crystal extends to 1.65 Å resolution, and the crystal belongs to the space group C2, with the unit cell parameters a = 133.3, b = 58.3, c = 112.3 Å, α = 90.00, β = 111.14, γ = 90.00.  相似文献   

3.
《Inorganica chimica acta》2006,359(5):1351-1356
Energy-transfer rate-constants from photo-excited [Ru(N–N)3]2+ (N–N = 2,2′-bipyridine (bpy), 4,4′-dimethyl-2,2′-bipyridine (4dmb), 5,5′-dimethyl-2,2′-bipyridine (5dmb)) to [Cr(O–O)3]3− (O–O2− = ox2− ((COO)2), mal2− (CH2(COO)2)) and [Cr(CN)6]3− in encounter complexes were evaluated in aqueous solutions containing alkali metal ion. The rate constant depends on the molecular size of the ruthenium(II) complex: 1.8 × 108 s−1 for [Ru(bpy)3]2+ (molecular radius, r = 5.8 Å), 1.4 × 108 s−1 for [Ru(5dmb)3]2+ (r = 6.1 Å) and 0.96 × 108 s−1 for [Ru(4dmb)3]2+ (r = 6.7 Å) in the system of [Ru(N–N)3]2+–[Cr(ox)3]3− in aqueous solution. However, the rate constant is much more sensitive to the chromate(III) complex than to ruthenium(II) complex; 1.8 × 108 s−1 and 0.43 × 108 s−1 for [Cr(ox)3]3− (r = 4.0 Å) and [Cr(mal)3]3− (r = 4.2 Å) in the [Ru(bpy)3]2+–[Cr(O–O)3]3− systems, respectively. We conclude that the congeniality between the donor’s and acceptor’s ligands in encounter complex plays an important role in energy transfer in aqueous solution.  相似文献   

4.
5.
Samarcandin 1, a natural sesquiterpene-coumarin, was isolated as well as elucidated from F. assa-foetida which has significant effect in Iranian traditional medicine because of its medicinal attitudes. The crystal structure of samarcandin was determined by single-crystal X-ray structure analysis. It is orthorhombic, with unit cell parameters a = 10.8204 (5) Å, b = 12.9894 (7) Å, c = 15.2467 (9) Å, V = 2142.9 (2) Å3, space group P212121 and four symmetry equivalent molecules in the unit cell. Samarcandin was isolated in order to study for its theoretical studies as well as its cellular toxicity as anti-cancer drug against two cancerous cells. In comparison with controls, our microscopic and MTT assay data showed that samarcandin suppresses cancer cell proliferation in a dose-dependent manner with IC50 = 11 μM and 13 for AGS and WEHI-164 cell lines, respectively. Density functional theory (DFT) and time-dependent density functional theory (TD-DFT) of the structure was computed by three functional methods and 6-311++G7 standard basis set. The optimized molecular geometry and theoretical analysis agree closely to that obtained from the single crystal X-ray crystallography. To sum up, the good correlations between experimental and theoretical studies by UV, NMR, and IR spectra were found.  相似文献   

6.
Diphytanoylphosphatidylcholine (DPhyPC) is a branched chain lipid often used for model membrane studies, including peptide/lipid interactions, ion channels and lipid rafts. This work reports results of volume measurements, water permeability measurements Pf, X-ray scattering from oriented samples, and X-ray and neutron scattering from unilamellar vesicles at T = 30 °C. We measured the volume/lipid VL = 1426 ± 1 Å3. The area/lipid was found to be 80.5 ± 1.5 Å2 when both X-ray and neutron data were combined with the SDP model analysis (Ku?erka, N., Nagle, J.F., Sachs, J.N., Feller, S.E., Pencer, J., Jackson, A., Katsaras, J., 2008. Lipid bilayer structure determined by the simultaneous analysis of neutron and X-ray scattering data. Biophys. J. 95, 2356–2367); this is substantially larger than the area of DOPC which has the largest area of the common linear chain lipids. Pf was measured to be (7.0 ± 1.0) × 10?3 cm/s; this is considerably smaller than predicted by the recently proposed 3-slab model (Nagle, J.F., Mathai, J.C., Zeidel, M.L., Tristram-Nagle, S., 2008. Theory of passive permeability through lipid bilayers. J. Gen. Physiol. 131, 77–85). This disagreement can be understood if there is a diminished diffusion coefficient in the hydrocarbon core of DPhyPC and that is supported by previous molecular dynamics simulations (Shinoda, W., Mikami, M., Baba, T., Hato, M., 2004. Molecular dynamics study on the effects of chain branching on the physical properties of lipid bilayers. 2. Permeability. J. Phys. Chem. B 108, 9346–9356). While the DPhyPC head–head thickness (DHH = 36.4 Å), and Hamaker parameter (H = 4.5 × 10?21 J) were similar to the linear chain lipid DOPC, the bending modulus (KC = 5.2 ± 0.5 × 10?21 J) was 30% smaller. Our results suggest that, from the biophysical perspective, DPhyPC belongs to a different family of lipids than phosphatidylcholines that have linear chain hydrocarbon chains.  相似文献   

7.
《Inorganica chimica acta》2001,312(1-2):188-196
The reaction of MoO3 and 2,4,6-tripyridyltriazine (tptz) in water at 180°C for 48 h and pH 5.5 produces (H2tptz)2[Mo8O26]·2H2O in 70% yield. The structure is constructed from δ-Mo8O26 4− clusters, H2tptz2+ and H3O+ cations linked through hydrogen bonding into a network. Crystal data: C18H16Mo4N6O14; monoclinic P21/n; a=10.2225(5) Å, b=14.0072(6) Å, c=18.1154(8) Å, β=93.896(1)°, V=2587.9(2) Å3, Z=4, Dcalc=2.372 g cm−3; R1=0.0271 based on 3212 reflections.  相似文献   

8.
《Inorganica chimica acta》2006,359(7):2029-2040
Two μ-alkoxo-μ-carboxylato bridged dinuclear copper(II) complexes, [Cu2(L1)(μ-HCO2)] (1) ((H3L1 = 1,3-bis(5-bromosalicylideneamino)-2-propanol)), [Cu2(L2)(μ-HCO2)] · dmf (2) (H3L2 = 1,3-bis(3,5-chlorosalicylideneamino-2-propanol)), and two μ-alkoxo-μ-dicarboxylato doubly bridged tetranuclear copper(II) complexes, [{Cu2(L3)}2(μ-O2C–C(CH3)2–CO2)] · 5H2O · 3CH3OH (3) ((H3L3 = 1,3-bis(salicylid-deneamino)-2-propanol)) and [{Cu2(L3)}2(μ- O2CCH2–C6H4–CH2CO2)] · 2H2O (4) have been prepared and characterized. The single crystal X-ray analysis shows that the structures of complexes 1 and 2 are dimeric with two adjacent copper(II) atoms bridged by μ-alkoxo-μ-carboxylato ligands with the Cu⋯Cu distances and Cu–O(alkoxo)–Cu angles are 3.511 Å and 132.85° for 1, 3.517 Å and 131.7° for 2, respectively. Complexes 3 and 4 consist of μ-alkoxo-μ-dicarboxylato doubly bridged tetranuclear Cu(II) complexes with mean Cu–Cu distances and Cu–O–Cu angles of 3.158 Å and 108.05° for 3 and 3.081 Å and 104.76° for 4, respectively. Magnetic measurements reveal that 1 and 2 are strong antiferromagnetically coupled with 2J = −156 and −152 cm−1, respectively, while 3 and 4 exhibit ferromagnetic coupling with 2J = 86 and 155.2 cm−1, respectively. The 2J values of 14 are linearly correlated to the Cu–O–Cu angles. Dependence of the pH at 25 °C on the reaction rate of oxidation of 3,4-di-tert-butylcatechol (3,5-dtbc) to the corresponding quinone catalyzed by 14 was studied. Complexes 14 exhibit high catecholase-like activity at pH 9.0 and 25 °C for oxidation of 3,5-di-tert-butylcatechol.  相似文献   

9.
The purpose of the work was to provide a crystallographic demonstration of the venerable idea that CO photolyzed from ferrous heme-a3 moves to the nearby cuprous ion in the cytochrome c oxidases. Crystal structures of CO-bound cytochrome ba3-oxidase from Thermus thermophilus, determined at ~ 2.8–3.2 Å resolution, reveal a Fe–C distance of ~ 2.0 Å, a Cu–O distance of 2.4 Å and a Fe–C–O angle of ~ 126°. Upon photodissociation at 100 K, X-ray structures indicate loss of Fea3–CO and appearance of CuB–CO having a Cu–C distance of ~ 1.9 Å and an O–Fe distance of ~ 2.3 Å. Absolute FTIR spectra recorded from single crystals of reduced ba3–CO that had not been exposed to X-ray radiation, showed several peaks around 1975 cm? 1; after photolysis at 100 K, the absolute FTIR spectra also showed a significant peak at 2050 cm? 1. Analysis of the ‘light’ minus ‘dark’ difference spectra showed four very sharp CO stretching bands at 1970 cm? 1, 1977 cm? 1, 1981 cm? 1, and 1985 cm? 1, previously assigned to the Fea3–CO complex, and a significantly broader CO stretching band centered at ~ 2050 cm? 1, previously assigned to the CO stretching frequency of CuB bound CO. As expected for light propagating along the tetragonal axis of the P43212 space group, the single crystal spectra exhibit negligible dichroism. Absolute FTIR spectrometry of a CO-laden ba3 crystal, exposed to an amount of X-ray radiation required to obtain structural data sets before FTIR characterization, showed a significant signal due to photogenerated CO2 at 2337 cm? 1 and one from traces of CO at 2133 cm? 1; while bands associated with CO bound to either Fea3 or to CuB in “light” minus “dark” FTIR difference spectra shifted and broadened in response to X-ray exposure. In spite of considerable radiation damage to the crystals, both X-ray analysis at 2.8 and 3.2 Å and FTIR spectra support the long-held position that photolysis of Fea3–CO in cytochrome c oxidases leads to significant trapping of the CO on the CuB atom; Fea3 and CuB ligation, at the resolutions reported here, are otherwise unaltered. This article is part of a Special Issue entitled: Respiratory Oxidases.  相似文献   

10.
Modified 3D-SDAR fingerprints combining 13C and 15N NMR chemical shifts augmented with inter-atomic distances were used to model the potential of chemicals to induce phospholipidosis (PLD). A curated dataset of 328 compounds (some of which were cationic amphiphilic drugs) was used to generate 3D-QSDAR models based on tessellations of the 3D-SDAR space with grids of different density. Composite PLS models averaging the aggregated predictions from 100 fully randomized individual models were generated. On each of the 100 runs, the activities of an external blind test set comprised of 294 proprietary chemicals were predicted and averaged to provide composite estimates of their PLD-inducing potentials (PLD+ if PLD is observed, otherwise PLD−). The best performing 3D-QSDAR model utilized a grid with a density of 8 ppm × 8 ppm in the C–C region, 8 ppm × 20 ppm in the C–N region and 20 ppm × 20 ppm in the N–N region. The classification predictive performance parameters of this model evaluated on the basis of the external test set were as follows: accuracy = 0.70, sensitivity = 0.73 and specificity = 0.66. A projection of the most frequently occurring bins on the standard coordinate space suggested a toxicophore composed of an aromatic ring with a centroid 3.5–7.5 Å distant from an amino-group. The presence of a second aromatic ring separated by a 4–5 Å spacer from the first ring and at a distance of between 5.5 Å and 7 Å from the amino-group was also associated with a PLD+ effect. These models provide comparable predictive performance to previously reported models for PLD with the added benefit of being based entirely on non-confidential, publicly available training data and with good predictive performance when tested in a rigorous, external validation exercise.  相似文献   

11.
Background: The human SWItch/Sucrose Non-Fermentable (SWI/SNF) chromatin remodeling complex plays essential roles in a variety of cellular processes and has been implicated in human cancer. However, the role of germline genetic variants in this complex in relation to cancer risk is not well studied. Methods: We assessed the association of 16 variants in the catalytic subunits (SMARCA2 and SMARCA4) of the SWI/SNF complex with the risk of glioma subtypes (lower grade astrocytoma, oligodendroglioma and glioblastoma [GBM]) and with mortality from high-grade tumors (GBM) in a multicenter US case–control study that included 561 cases and 574 controls. Associations were estimated with odds ratios (OR, for risk) or hazards ratios (HR, for mortality) with 95% confidence intervals (CI). False discovery rate (FDR-q) was used to control for multiple testing in risk associations. Results: None of the investigated SNPs was associated with overall glioma risk. However, analyses according to histological subtypes revealed a statistically significant increased risk of oligodendroglioma in association with SMARCA2 rs2296212 (OR = 4.05, 95%CI = 1.11–14.80, P = 0.030, q = 0.08) and rs4741651 (OR = 4.68, 95%CI = 1.43–15.30, P = 0.011, q = 0.08) and SMARCA4 rs11672232 (OR = 1.90, 95%CI = 1.01–3.58, P = 0.048, q = 0.08) and rs12232780 (OR = 2.14, 95%CI = 1.06–4.33, P = 0.035, q = 0.08). No significant risk associations were observed for GBM or lower grade astrocytoma. Suggestive associations with GBM mortality were not validated in the Cancer Genome Atlas. Conclusion: Our findings suggest that genetic variants in SMARCA2 and SMARCA4 influence the risk of oligodendroglioma. Further research is warranted on the SWI/SNF complex genes and epigenetic mechanisms more generally in the development of glioma in adults.  相似文献   

12.
《Inorganica chimica acta》2001,312(1-2):215-220
The reaction of [M(H2O)3(CO)3]+ (M=Tc, Re) with Na[CpCo[PO(OR)2]3] (NaLOR; R=Me, Et) in water produced the compounds M(CO)3(LOR), all of which were yellow solids, in yields varying from 55 to 89%. The two compounds M(CO)3(LOEt) were structurally characterized by single crystal X-ray crystallography. In both cases, the ligand LOEt was bound to the metal center in a tridentate fashion utilizing an {OOO} donor set. The ligands LOR can be used as models for facially coordinated triaqua groups owing to their position in the spectrochemical series. Therefore, these four compounds, M(CO)3(LOR), can be considered structural models for [M(H2O)3(CO)3]+. Crystal data for Tc(CO)3(LOEt) are as follows: molecular formula C20H35CoO12P3Tc, MW=717.32, monoclinic, a=11.5661(11) Å, b=18.671(2) Å, c=13.7852(13) Å, β=92.770(2)°, V=2973.5(5) Å3, space group P21/n, Z=4, final R1=0.0669, wR2=0.1361. Crystal data for Re(CO)3(LOEt) are as follows: molecular formula C20H35CoO12P3Re, MW=805.52, monoclinic, a=11.5113(7) Å, b=18.6022(12) Å, c=13.7397(8) Å, β=92.7580(10)°, V=2938.7(3) Å3, space group P21/n, Z=4, final R1=0.0384, wR2=0.0760.  相似文献   

13.
DNA and chromosome damages in peripheral blood lymphocytes were evaluated in 151 workers occupationally exposed to formaldehyde (FA) and 112 non-FA exposed controls. The effects of polymorphisms in three glutathione-S-transferase (GSTs) genes on the DNA and chromosome damages were assessed as well. Alkaline comet assay and cytokinesis-block micronucleus (CBMN) assay were used to determine DNA and chromosome damages, respectively. The genotypes of GSTP1 (Ile105Val), GSTT1, and GSTM1 were assayed. The mean 8-h time-weighted average (TWA) concentrations of FA in two plywood factories were 0.83 ppm (range: 0.08–6.30 ppm). FA-exposed workers had higher olive tail moment (TM) and CBMN frequency compared with controls (Olive TM, 3.54, 95%CI = 3.19–3.93 vs. 0.93, 95%CI = 0.78–1.10, P < 0.01; CBMN frequency, 5.51 ± 3.37 vs. 2.67 ± 1.32, P < 0.01). Olive TM and the CBMN frequency also had a dose-dependent relation with the personal FA exposure. Significant association between FA exposure history and olive TM and CBMN frequency were also identified. The level of olive TM was slightly higher in FA-exposed workers with GSTM1 null genotype than those with non-null genotype (3.86, 95%CI = 3.31–4.50 vs. 3.27, 95%CI = 2.83–3.78, P = 0.07) with adjustment of covariates. We also found that FA-exposed workers carrying GSTP1 Val allele had a slightly higher CBMN frequency compared with workers carrying only the wild-type allele (6.32 ± 3.78 vs. 5.01 ± 2.98, P = 0.05). Our results suggest that the FA exposure in this occupational population increased DNA and chromosome damages and polymorphisms in GSTs genes may modulate the genotoxic effects of FA exposure.  相似文献   

14.
Two copper(II) complexes, [Cu(sac)2(4-cypy)2(H2O)], 1 and [Cu(sac)2(4-Ampy)2(H2O)], 2 (4-cypy: 4-cyanopyridine; 4-Ampy: 4-aminopyridine) were prepared. Physicochemical properties of the complexes were studied by spectroscopic (solution UV–vis, diffuse reflectance and IR) techniques. Structural X-ray diffraction data could be obtained only for [Cu(sac)2(4-cypy)2(H2O)] that it crystallized in the tetragonal space group P4cc with a=b=15.313(1), c=13.240(1) Å, and Z=4 molecules per unit cell. The complex was cited on a crystallographic C2-axis with the Cu(II) ion in a square–pyramidal environment, coordinated at the pyramid basis to the nitrogen atom of two saccharine anions [d(Cu–N)=2.011(3) Å] and the pyridine N-atom of two 4-cyanopyridine ligands [d(Cu–N)=2.038(4) Å]. The coordination was completed by a water molecule at the pyramid apex [d(Cu–Ow)=2.189(5) Å]. Elemental and spectroscopic analyses revealed an O-saccharinate coordination mode for complex 2 and a square–pyramidal structure. Only complex 2 retained its structure in methanolic solution. However, both complexes were able to catalyze the dismutation of superoxide anion (O2?) (pH 7.5) at micromolar concentrations. Therefore, these complexes behaved as useful SOD-mimetic compounds.  相似文献   

15.
Obstructive sleep apnea (OSA) is characterized by recurrent apnea during sleep that may unbalance oxidative stress, increasing atherosclerosis. Among oxidative stress markers, 15-F2t-isoprostane is considered one of the most sensitive and specific metabolites of lipid peroxidation. To explore the relationship between urinary 15-F2t-isoprostane with sleep apnea severity and carotid modifications in nonobese OSA patients, 31 nonobese sleep apnea patients were studied, along with 10 lean subjects without OSA. Patients were assessed by polysomnography, blood pressure measurement, and ultrasonography to determine the carotid intima–media thickness (IMT). Urinary 15-F2t-isoprostanes were measured by liquid chromatography–tandem mass spectrometry. Urinary 15-F2t-isoprostane concentrations were increased in severe OSA patients compared to control subjects (20.2 ± 7.3 vs 12.3 ± 2.8 ng/mmol creatinine; P = 0.020). Mean carotid IMT was correlated with 15-F2t-isoprostane (r = 0.532; P < 0.001) and with the apnea–hypopnea index (r = 0.345; P = 0.029). 15-F2t-Isoprostane level was related to the night time spent at SaO2 < 90% (r = 0.478; P = 0.002), the apnea–hypopnea index (r = 0.465; P = 0.003), and the mean nocturnal SaO2 (r = ? 0.424; P = 0.007). These results showed a relationship between lipid peroxidation, carotid intima–media thickness, and intermittent hypoxia in nonobese OSA patients, thus reinforcing the hypothesis that oxidative stress could be involved in the early atherosclerotic process.  相似文献   

16.
Recent evidences suggested a possible relationship between zinc deficiency and leptin levels in pathogenesis of anorexia in chronic kidney disease. The present study addressed the relationship between zinc and leptin in hemodialysis (HD) patients.MethodsFifty HD patients (54.3 ± 12.7 years old, 62% men) were studied and compared to 21 healthy volunteers (50.7 ± 15.7 years old, 43% men). Biochemical data, serum zinc, plasma leptin, IL-6, TNF-α and C-Reactive Protein levels were determined. Anthropometric parameters, food intake and appetite score were also assessed.ResultsThe leptin levels were higher in HD patients (16.1 μg/mL (0.21–118.25) vs 6.0 μg/mL (0.50–23.10)) in healthy volunteers (p = 0.04), whereas serum zinc levels were lower (54.5 ± 16.3 μg/dL) compared to healthy volunteers (78.4 ± 9.4 μg/dL) (p = 0.0001). The plasma leptin was correlated negatively with plasma zinc (r = ?0.33; p = 0.007), energy (r = ?0.38; p = 0.002) and protein intake (r = ?0.34; p = 0.006) and, positively correlated with BMI (r = 0.54; p = 0.0001), % body fat (r = 0.70; p = 0.0001) and conicity index (r = 0.46; p = 0.001). Plasma zinc was associated with hemoglobin (r = 0.30; p = 0.04) and negatively associated with TNF-α (r = ?0.37; p = 0.002) and C-Reactive Protein (r = ?0.37; p = 0.004). There was no correlation among Zn, leptin and appetite score in these patients.ConclusionThis study showed that low plasma zinc levels are negatively associated with high leptin levels in HD patients.  相似文献   

17.
BackgroundTo assess the existence of association between neutrophil to lymphocyte ratio (NLR) and the risk of sarcopenia in COVID-19 patients.MethodsA retrospective cross-sectional study was conducted in a university hospital with patients with an active COVID-19 infection admitted to the nursing ward or intensive care unit (ICU) between September to December 2020. Sarcopenia risk was assessed using the Strength, Assistance for walking, Rise from a chair, Climb stairs and Falls (SARC-F). Biochemical analyses were assessed by circulating of C-reactive protein, D-dimer, neutrophils, lymphocytes count and NLR. Sixty-eight patients were evaluated and divided into tertiles of NLR values and the association between NLR and sarcopenia risk were tested using the linear regression analyses and p < 0.05 were considered as significant.ResultsSixty-eight patients were evaluated and divided in NLR tertiles being the 1st (men = 52.2%; 71.1 ± 9.0 y; NLR: 1.1–3.85), 2nd (women = 78.3%; 73.2 ± 9.1 y; NLR: 3.9–6.0) and 3rd (men = 72.7%; 71.7 ± 10.4 y; NLR: 6.5–20.0). There was a difference between the tertiles in relation to the first to the biochemical parameters of total neutrophils count (p = 0.001), C-reactive protein (p = 0.012), and D-dimer (p = 0.012). However, no difference was found in linear regression analysis between tertiles of NLR and SARC-F, if in total sample (p = 0.054) or divided by sex, if men (p = 0.369) or women (p = 0.064).ConclusionIn elderly patients hospitalized with COVID-19, we do not find an association between the risk of sarcopenia and NLR.  相似文献   

18.
Mutations in the second EF-hand (D61N, D63N, D65N, and E72A) of S100B were used to study its Ca2 + binding and dynamic properties in the absence and presence of a bound target, TRTK-12. With D63NS100B as an exception (D63NKD = 50 ± 9 μM), Ca2 + binding to EF2-hand mutants were reduced by more than 8-fold in the absence of TRTK-12 (D61NKD = 412 ± 67 μM, D65NKD = 968 ± 171 μM, and E72AKD = 471 ± 133 μM), when compared to wild-type protein (WTKD = 56 ± 9 μM). For the TRTK-12 complexes, the Ca2 +-binding affinity to wild type (WT + TRTKKD = 12 ± 10 μM) and the EF2 mutants was increased by 5- to 14-fold versus in the absence of target (D61N + TRTKKD = 29 ± 1.2 μM, D63N + TRTKKD = 10 ± 2.2 μM, D65N + TRTKKD = 73 ± 4.4 μM, and E72A + TRTKKD = 18 ± 3.7 μM). In addition, Rex, as measured using relaxation dispersion for side‐chain 15N resonances of Asn63 (D63NS100B), was reduced upon TRTK-12 binding when measured by NMR. Likewise, backbone motions on multiple timescales (picoseconds to milliseconds) throughout wild type, D61NS100B, D63NS100B, and D65NS100B were lowered upon binding TRTK-12. However, the X-ray structures of Ca2 +-bound (2.0 Å) and TRTK-bound (1.2 Å) D63NS100B showed no change in Ca2 + coordination; thus, these and analogous structural data for the wild-type protein could not be used to explain how target binding increased Ca2 +-binding affinity in solution. Therefore, a model for how S100B–TRTK‐12 complex formation increases Ca2 + binding is discussed, which considers changes in protein dynamics upon binding the target TRTK-12.  相似文献   

19.
《BBA》2013,1827(10):1165-1173
Proton matrix ENDOR was performed to investigate the protons close to the manganese cluster in oriented samples of photosystem II (PS II). Eight pairs of ENDOR signals were detected in oriented PS II membranes. At an angle of θ = 0° between the membrane normal vector n and the external field H0, five pairs of ENDOR signals were exchangeable in D2O medium and three pairs were not exchangeable in D2O medium. The hyperfine splitting of 3.60 MHz at θ = 0° increased to 3.80 MHz at θ = 90°. The non-exchangeable signals with 1.73 MHz hyperfine splitting at θ = 0°, which were assigned to a proton in an amino acid residue, were not detected at θ = 90° in oriented PS II or in non-oriented PS II. Highly resolved spectra show that only limited numbers of protons were detected by CW-ENDOR spectra, although many protons were located near the CaMn4O5 cluster. The detected exchangeable protons were proposed to arise from the protons belonging to the water molecules, labeled W1-W4 in the 1.9 Å crystal structure, directly ligated to the CaMn4O5 cluster, and nearby amino-acid residue.  相似文献   

20.
A group of cyclic imides (110) was designed for evaluation as a selective COX-2 inhibitors and investigated in vivo for their anti-inflammatory activity. Compounds 6a, 6b, 8a, 8b, 9a, 9b, 10a and 10b were proved to be potent COX-2 inhibitors with IC50 range of 0.1–4.0 μM. In vitro COX-1/COX-2 inhibition structure–activity studies identified compound 8a as a highly potent (IC50 = 0.1 μM), and an extremely selective [COX-2 (SI) > 1000] comparable to celecoxib [COX-2 (SI) > 384], COX-2 inhibitor that showed superior anti-inflammatory activity (ED50 = 72.4 mg/kg) relative to diclofenac (ED50 = 114 mg/kg). Molecular modeling was carried out through docking the designed compounds into the COX-2 binding site to predict if these compounds have analogous binding mode to the COX-2 inhibitors. The study showed that the homosulfonamide fragment of 8a inserted deep inside the 2°-pocket of the COX-2 active site, where the SO2NH2 group underwent H-bonding interaction with Gln192(2.95 Å), Phe518(2.82 Å) and Arg513(2.63 and 2.73 Å). Docking study of the synthesized compound 8a into the active site of COX-2 revealed a similar binding mode to SC-558, a selective COX-2 inhibitor.  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