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 15 毫秒
1.
The reaction of 2,6-diformyl-4-methylphenol (DFMF) with 1-amino-2-propanol (AP) and tris(hydroxymethyl)aminomethane (THMAM) was investigated in the presence of Cobalt(II) salts, (X = ClO4, CH3CO2, Cl, NO3), sodium azide (NaN3), and triethylamine (TEA). In one pot, the variation in Cobalt(II) salt results in the self-assembly of dinuclear, tetranuclear, and H-bonding-directed polynuclear coordination complexes of Cobalt(III), Cobalt(II), and mixed-valence CoIICoIII: [Co2III(H2L1)2(AP1)(N3)](ClO4)2 (1), [Co4(H2L1)23-1,1,1-N3)2(µ-1,1-N3)2Cl2(CH3OH)2]·4CH3OH (2), [Co2IICo2III(HL2)2(µ-CH3CO2)23-OH)2](NO3)2·2CH3CH2OH (3), and [Co2IICo2III (H2L12)2(THMAM−1)2](NO3)4 (4). In 1, two cobalt(III) ions are connected via three single atom bridges; two from deprotonated ethanolic oxygen atoms in the side arms of the ligands and one from the1-amino-2-propanol moiety forming a dinuclear unit with a very short (2.5430(11) Å) Co-Co intermetallic separation with a coordination number of 7, a rare feature for cobalt(III). In 2, two cobalt(II) ions in a dinuclear unit are bridged through phenoxide O and μ3-1,1,1-N3 azido bridges, and the two dinuclear units are interconnected by two μ-1,1-N3 and two μ3-1,1,1-N3 azido bridges generating tetranuclear cationic [Co4(H2L1)23-1,1,1-N3)2(µ-1,1-N3)2Cl2(CH3OH)2]2+ units with an incomplete double cubane core, which grow into polynuclear 1D-single chains along the a-axis through H-bonding. In 3, HL2− holds mixed-valent Co(II)/Co(III) ions in a dinuclear unit bridged via phenoxide O, μ-1,3-CH3CO2, and μ3-OH bridges, and the dinuclear units are interconnected through two deprotonated ethanolic O in the side arms of the ligands and two μ3-OH bridges generating cationic tetranuclear [Co2IICo2III(HL2)2(µ-CH3CO2)23-OH)2]2+ units with an incomplete double cubane core. In 4, H2L1−2 holds mixed-valent Co(II)/Co(III) ions in dinuclear units which dimerize through two ethanolic O (μ-RO) in the side arms of the ligands and two ethanolic O (μ3-RO) of THMAM bridges producing centrosymmetric cationic tetranuclear [Co2IICo2III (H2L12)2(THMAM−1)2]4+ units which grow into 2D-sheets along the bc-axis through a network of H-bonding. Bulk magnetization measurements on 2 demonstrate that the magnetic interactions are completely dominated by an overall ferromagnetic coupling occurring between Co(II) ions.  相似文献   

2.
The Mg(II) and heterometallic Mn(II)/Na(I) complexes of isoferulic acid (3-hydroxy-4-methoxycinnamic acid, IFA) were synthesized and characterized by infrared spectroscopy FT-IR, FT-Raman, electronic absorption spectroscopy UV/VIS, and single-crystal X-ray diffraction. The reaction of MgCl2 with isoferulic acid in the aqueous solutions of NaOH resulted in synthesis of the complex salt of the general formula of [Mg(H2O)6]⋅(C10H9O4)2⋅6H2O. The crystal structure of this compound consists of discrete octahedral [Mg(H2O)6]2+ cations, isoferulic acid anions and solvent water molecules. The hydrated metal cations are arranged among the organic layers. The multiple hydrogen-bonding interactions established between the coordinated and lattice water molecules and the functional groups of the ligand stabilize the 3D architecture of the crystal. The use of MnCl2 instead of MgCl2 led to the formation of the Mn(II)/Na(I) complex of the general formula [Mn3Na2(C10H7O4)8(H2O)8]. The compound is a 3D coordination polymer composed of centrosymmetric pentanuclear subunits. The antioxidant activity of these compounds was evaluated by assays based on different antioxidant mechanisms of action, i.e., with OH, DPPH and ABTS•+ radicals as well as CUPRAC (cupric ions reducing power) and lipid peroxidation inhibition assays. The pro-oxidant property of compounds was measured as the rate of oxidation of Trolox. The Mg(II) and Mn(II)/Na(I) complexes with isoferulic acid showed higher antioxidant activity than ligand alone in DPPH (IFA, IC50 = 365.27 μM, Mg(II) IFA IC50 = 153.50 μM, Mn(II)/Na(I) IFA IC50 = 149.00 μM) and CUPRAC assays (IFA 40.92 μM of Trolox, Mg(II) IFA 87.93 μM and Mn(II)/Na(I) IFA 105.85 μM of Trolox; for compounds’ concentration 10 μM). Mg(II) IFA is a better scavenger of OH than IFA and Mn(II)/Na(I) IFA complex. There was no distinct difference in ABTS•+ and lipid peroxidation assays between isoferulic acid and its Mg(II) complex, while Mn(II)/Na(I) complex showed lower activity than these compounds. The tested complexes displayed only slight antiproliferative activity tested in HaCaT human immortalized keratinocyte cell line within the solubility range. The Mn(II)/Na(I) IFA (16 μM in medium) caused an 87% (±5%) decrease in cell viability, the Mg salt caused a comparable, i.e., 87% (±4%) viability decrease in a concentration of 45 μM, while IFA caused this level of cell activity attenuation (87% ± 5%) at the concentration of 1582 μM (significant at α = 0.05).  相似文献   

3.
In this paper, we prepared two phosphorescent Ir complexes with ligands of 2-phenyl pyridine (ppy), and two phosphorous ligands with large steric hindrance, hoping to allow enough time for the transformation of the highly phosphorescent 3MLLCT (metal-to-ligand-ligand-charge-transfer) excited state. Their large steric hindrance minimized the π-π interaction between complex molecules, so that the aggregation-induced phosphorescence emission (AIPE) influence could be minimized. Their single crystals indicated that they took a distorted octahedral coordination mode. Photophysical comparison between these Ir complexes in solution, in the solid state and in electrospun fibers was performed to confirm the realization of limited aggregation-caused quenching (ACQ). The potential surface crossing and energy transfer from 3MLBPECT/3MLBPELppyCT to 3MLppyCT in these Ir complexes were revealed by density functional theory calculation and temperature-dependent emission. It was confirmed that these two phosphorous ligands offered large steric hindrance, which decreased the ACQ effect, allowing the efficient emissive decay of the 3MLppyCT excited state. This is one of the several luminescent Ir complexes with a high emission yield (Φ = 0.27) and long emission lifetime (0.43 μs) in the solid state.  相似文献   

4.
A [NiFe] hydrogenase model compound having a distorted trigonal-pyramidal nickel center, (CO)3Fe(μ-StBu)3Ni(SDmp), 1 (Dmp = C6H3-2,6-(mesityl)2), was synthesized from the reaction of the tetranuclear Fe-Ni-Ni-Fe complex [(CO)3Fe(μ-StBu)3Ni]2(μ-Br)2, 2 with NaSDmp at -40 °C. The nickel site of complex 1 was found to add CO or CNtBu at -40 °C to give (CO)3Fe(StBu)(μ-StBu)2Ni(CO)(SDmp), 3, or (CO)3Fe(StBu)(μ-StBu)2Ni(CNtBu)(SDmp), 4, respectively. One of the CO bands of 3, appearing at 2055 cm-1 in the infrared spectrum, was assigned as the Ni-CO band, and this frequency is comparable to those observed for the CO-inhibited forms of [NiFe] hydrogenase. Like the CO-inhibited forms of [NiFe] hydrogenase, the coordination of CO at the nickel site of 1 is reversible, while the CNtBu adduct 4 is more robust.  相似文献   

5.
Let ρn denote the standard n-dimensional representation of GL(n,C) and ρn2 its symmetric square. For each automorphic cuspidal representation π of GL(2,A) we introduce an Euler product L(s,π,ρ22) of degree 3 which we prove is entire. We also prove that there exists an automorphic representation II of GL(3)—“the lift of π”—with the property that L(s,II,ρ3) = L(s,π,ρ22). Our results confirm conjectures described in a more general context by R. P. Langlands [(1970) Lecture Notes in Mathematics, no. 170 (Springer-Verlag, Berlin-Heidelberg-New York)].  相似文献   

6.
Xishi Tai  Jinhe Jiang 《Materials》2012,5(9):1626-1634
A new trinuclear Cd (II) complex [Cd3(L)6(2,2-bipyridine)3] [L = N-phenylsulfonyl-L-leucinato] has been synthesized and characterized by elemental analysis, IR and X-ray single crystal diffraction analysis. The results show that the complex belongs to the orthorhombic, space group P212121 with a = 16.877(3) Å, b = 22.875(5) Å, c = 29.495(6) Å, α = β = γ = 90°, V = 11387(4) Å3, Z = 4, Dc= 1.416 μg·m−3, μ = 0.737 mm−1, F (000) = 4992, and final R1 = 0.0390, ωR2 = 0.0989. The complex comprises two seven-coordinated Cd (II) atoms, with a N2O5 distorted pengonal bipyramidal coordination environment and a six-coordinated Cd (II) atom, with a N2O4 distorted octahedral coordination environment. The molecules form one dimensional chain structure by the interaction of bridged carboxylato groups, hydrogen bonds and π-π interaction of 2,2-bipyridine. The luminescent properties of the Cd (II) complex and N-Benzenesulphonyl-L-leucine in solid and in CH3OH solution also have been investigated.  相似文献   

7.
Bicarbonate transporters are regulated by signaling molecules/ions such as protein kinases, ATP, and Ca2+. While phospholipids such as PIP2 can stimulate Na-H exchanger activity, little is known about phospholipid regulation of bicarbonate transporters. We used the patch-clamp technique to study the function and regulation of heterologously expressed rat NBCe1-A in excised macropatches from Xenopus laevis oocytes. Exposing the cytosolic side of inside-out macropatches to a 5% CO2/33 mM HCO3 solution elicited a mean inward current of 14 pA in 74% of macropatches attached to pipettes (−Vp = −60 mV) containing a low-Na+, nominally HCO3-free solution. The current was 80–90% smaller in the absence of Na+, approximately 75% smaller in the presence of 200 μM DIDS, and absent in macropatches from H2O-injected oocytes. NBCe1-A currents exhibited time-dependent rundown that was inhibited by removing Mg2+ in the presence or absence of vanadate and F to reduce general phosphatase activity. Applying 5 or 10 μM PIP2 (diC8) in the presence of HCO3 induced an inward current in 54% of macropatches from NBC-expressing, but not H2O-injected oocytes. PIP2-induced currents were HCO3-dependent and somewhat larger following more NBCe1-A rundown, 62% smaller in the absence of Na+, and 90% smaller in the presence of 200 μM DIDS. The polycation neomycin (250–500 μM) reduced the PIP2-induced inward current by 69%; spermine (100 μM) reduced the current by 97%. Spermine, poly-D-lysine, and neomycin all reduced the baseline HCO3-induced inward currents by as much as 85%. In summary, PIP2 stimulates NBCe1-A activity, and phosphoinositides are regulators of bicarbonate transporters.  相似文献   

8.
Xishi Tai  Na Wei  Donghao Wang 《Materials》2012,5(4):558-565
A new complex [Mg(L)2(phen)(H2O)2](phen)(H2O)2 [L= N-benzenesulphonyl-L-leucine] was synthesized by the reaction of magnesium chloride hexahydrate with N-benzenesulphonyl-L-leucine and 1,10-phenanthroline in the CH3CH2OH/H2O (v:v = 5:1). It was characterized by elemental analysis, IR and X-ray single crystal diffraction analysis. The crystal of the title complex [Mg(L)2(phen)(H2O)2](phen)(H2O)2 belongs to triclinic, space group P-1 with a = 0.72772(15) nm, b = 1.4279(3) nm, c = 1.4418(3) nm, α = 63.53(3)°, β = 79.75(3)°, γ = 81.83(3)°, V = 1.3163(5) nm3, Z =1, Dc= 1.258 μg·m−3, μ = 0.177 mm−1, F(000) = 526, and final R1 = 0.0506, ωR2 = 0.1328. The complex comprises a six-coordinated magnesium(II) center, with a N2O4 distorted octahedron coordination environment. The molecules are connected by hydrogen bonds and π-π stacking to form one dimensional chain structure. The luminescent property of the Mg(II) complex has been investigated in solid.  相似文献   

9.
(E)-β-Farnesene is a sesquiterpene semiochemical that is used extensively by both plants and insects for communication. This acyclic olefin is found in the essential oil of peppermint (Mentha x piperita) and can be synthesized from farnesyl diphosphate by a cell-free extract of peppermint secretory gland cells. A cDNA from peppermint encoding (E)-β-farnesene synthase was cloned by random sequencing of an oil gland library and was expressed in Escherichia coli. The corresponding synthase has a deduced size of 63.8 kDa and requires a divalent cation for catalysis (Km for Mg2+ ≈ 150 μM; Km for Mn2+ ≈ 7 μM). The sesquiterpenoids produced by the recombinant enzyme, as determined by radio-GC and GC-MS, are (E)-β-farnesene (85%), (Z)-β-farnesene (8%), and δ-cadinene (5%) with the native C15 substrate farnesyl diphosphate (Km ≈ 0.6 μM; Vrel = 100) and Mg2+ as cofactor, and (E)-β-farnesene (98%) and (Z)-β-farnesene (2%) with Mn2+ as cofactor (Vrel = 80). With the C10 analog, GDP, as substrate (Km = 1.5 μM; Vrel = 3 with Mg2+ as cofactor), the monoterpenes limonene (48%), terpinolene (15%), and myrcene (15%) are produced.  相似文献   

10.
Theorem 1. For α, β on the range 1,..., μ, let Q(z) = *aαβzαzβ be a real valued, nonsingular, symmetric quadratic form. For positive integers r and s such that μ = r + s set (z1,..., zμ) = (u1,..., ur:S1,..., Sn), Q(z) = P(u, s) and [Formula: see text] Let B = (z(1),..., z(r)) be a base “over R” for points z ε πr. For an arbitrary r-tuple ω1,..., ωr set [Formula: see text] index HB(ω) = κ and nullity HB(ω) = ν. Then [Formula: see text]  相似文献   

11.
This paper focuses on the synthesis, structural characterization, and study of the optical, magnetic, and thermal properties of novel architectures combining metal ions as magnetoactive centers and photoactive blocks formed by carbazole units. For this purpose, a series of azomethine complexes of the composition [Fe(L)2]X (L = 3,6-bis[(3′,6′-di-tert-butyl-9-carbazol)-9-carbazol]benzoyloxy-4-salicylidene-N′-ethyl-N-ethylenediamine, X = NO3, Cl, PF6) were synthesized by the reaction of metal salts with Schiff bases in a mixture of solvents. The UV–Vis absorption properties were studied in dichloromethane and rationalized via time-dependent density functional theory (DFT) calculations. Upon excitation at 350 nm, the compounds exhibited an intense dual fluorescence with two emission bands centered at ~445 and ~485 nm, which were assigned to πcarbπ* intraligand and πcarb–dFe ligand-to-metal charge-transfer excited states. EPR spectroscopy and SQUID magnetometry revealed solid-state partial spin crossover in some compounds, and antiferromagnetic interactions between the neighboring Fe(III) ions.  相似文献   

12.
Methyllycaconitine (MLA), α-conotoxin ImI, and α-bungarotoxin inhibited the release of catecholamines triggered by brief pulses of acetylcholine (ACh) (100 μM, 5 s) applied to fast-superfused bovine adrenal chromaffin cells, with IC50s of 100 nM for MLA and 300 nM for α-conotoxin ImI and α-bungarotoxin. MLA (100 nM), α-conotoxin ImI (1 μM), and α-bungarotoxin (1 μM) halved the entry of 45Ca2+ stimulated by 5-s pulses of 300 μM ACh applied to incubated cells. These supramaximal concentrations of α7 nicotinic receptor blockers depressed by 30% (MLA), 25% (α-bungarotoxin), and 50% (α-conotoxin ImI) the inward current generated by 1-s pulses of 100 μM ACh, applied to voltage-clamped chromaffin cells. In Xenopus oocytes expressing rat brain α7 neuronal nicotinic receptor for acetylcholine nAChR, the current generated by 1-s pulses of ACh was blocked by MLA, α-conotoxin ImI, and α-bungarotoxin with IC50s of 0.1 nM, 100 nM, and 1.6 nM, respectively; the current through α3β4 nAChR was unaffected by α-conotoxin ImI and α-bungarotoxin, and weakly blocked by MLA (IC50 = 1 μM). The functions of controlling the electrical activity, the entry of Ca2+, and the ensuing exocytotic response of chromaffin cells were until now exclusively attributed to α3β4 nAChR; the present results constitute the first evidence to support a prominent role of α7 nAChR in controlling such functions, specially under the more physiological conditions used here to stimulate chromaffin cells with brief pulses of ACh.  相似文献   

13.
Studies have shown that fish oils, containing n-3 fatty acids, have protective effects against ischemia-induced, fatal cardiac arrhythmias in animals and perhaps in humans. In this study we used the whole-cell voltage-clamp technique to assess the effects of dietary, free long-chain fatty acids on the Na+ current (INa,α) in human embryonic kidney (HEK293t) cells transfected with the α-subunit of the human cardiac Na+ channel (hH1α). Extracellular application of 0.01 to 30 μM eicosapentaenoic acid (EPA, C20:5n-3) significantly reduced INa,α with an IC50 of 0.51 ± 0.06 μM. The EPA-induced suppression of INa,α was concentration- and voltage-dependent. EPA at 5 μM significantly shifted the steady-state inactivation relationship by −27.8 ± 1.2 mV (n = 6, P < 0.0001) at the V1/2 point. In addition, EPA blocked INa,α with a higher “binding affinity” to hH1α channels in the inactivated state than in the resting state. The transition from the resting state to the inactivated state was markedly accelerated in the presence of 5 μM EPA. The time for 50% recovery from the inactivation state was significantly slower in the presence of 5 μM EPA, from 2.1 ± 0.8 ms for control to 34.8 ± 2.1 ms (n = 5, P < 0.001). The effects of EPA on INa,α were reversible. Furthermore, docosahexaenoic acid (C22:6n-3), α-linolenic acid (C18:3n-3), conjugated linoleic acid (C18:2n-7), and oleic acid (C18:1n-9) at 5 μM and all-trans-retinoic acid at 10 μM had similar effects on INa,α as EPA. Even 5 μM of stearic acid (C18:0) or palmitic acid (C16:0) also significantly inhibited INa,α. In contrast, 5 μM EPA ethyl ester did not alter INa,α (8 ± 4%, n = 8, P > 0.05). The present data demonstrate that free fatty acids suppress INa,α with high “binding affinity” to hH1α channels in the inactivated state and prolong the duration of recovery from inactivation.  相似文献   

14.
Specific binding of phorbol ester tumor promoters   总被引:16,自引:12,他引:16       下载免费PDF全文
[20-3H]Phorbol 12,13-dibutyrate bound to particulate preparations from chicken embryo fibroblasts in a specific, saturable, reversible fashion. Equilibrium binding occurred with a Kd of 25 nM; this value is very close to the 50% effective dose (ED50), 50 nM, previously determined for the biological response (induction of fibronectin loss) in growing chicken embryo fibroblasts. At saturation, 1.4 pmol of [20-3H]phorbol 12,13-dibutyrate was bound per mg of protein (approximately 7 × 104 molecules per cell). Binding was inhibited by phorbol 12-myristate 13-acetate (Ki = 2 nM), mezerein (Ki = 180 nM), phorbol 12,13-dibenzoate (Ki = 180 nM), phorbol 12,13-diacetate (Ki = 1.7 μM), phorbol 12,13,20-triacetate (Ki = 39 μM), and phorbol 13-acetate (Ki = 120 μM). The measured Ki values are all within a factor of 3.5 of the ED50 values of these derivatives for inducing loss of fibronectin in intact cells. Binding was not inhibited by the inactive compounds phorbol (10 μg/ml) and 4α-phorbol 12,13-didecanoate (10 μg/ml) or by the inflammatory but nonpromoting phorbol-related diterpene esters resiniferatoxin (100 ng/ml) and 12-deoxyphorbol 13-isobutyrate 20-acetate (100 ng/ml). These data suggest that biological responses to the phorbol esters in chicken embryo fibroblasts are mediated by this binding activity and that the binding activity corresponds to the phorbol ester target in mouse skin involved in tumor promotion. Binding was not inhibited by the nonphorbol promoters anthralin (1 μM), phenol (1 mM), iodoacetic acid (1.7 μM), and cantharidin (75 μM), or by epidermal growth factor (100 ng/ml), dexamethasone acetate (2 μM), retinoic acid (10 μM), or prostaglandin E2 (1 μM). These agents thus appear to act at a target distinct from that of the phorbol esters.  相似文献   

15.

Background and objectives

The generation of key uremic nephrovascular toxins, indoxyl sulfate (IS), and p-cresyl sulfate (PCS), is attributed to the dysbiotic gut microbiota in CKD. The aim of our study was to evaluate whether synbiotic (pre- and probiotic) therapy alters the gut microbiota and reduces serum concentrations of microbiome–generated uremic toxins, IS and PCS, in patients with CKD.

Design, setting, participants, & measurements

Predialysis adult participants with CKD (eGFR=10–30 ml/min per 1.73 m2) were recruited between January 5, 2013 and November 12, 2013 to a randomized, double–blind, placebo–controlled, crossover trial of synbiotic therapy over 6 weeks (4-week washout). The primary outcome was serum IS. Secondary outcomes included serum PCS, stool microbiota profile, eGFR, proteinuria-albuminuria, urinary kidney injury molecule-1, serum inflammatory biomarkers (IL-1β, IL-6, IL-10, and TNF-α), serum oxidative stress biomarkers (F2-isoprostanes and glutathione peroxidase), serum LPS, patient-reported health, Gastrointestinal Symptom Score, and dietary intake. A prespecified subgroup analysis explored the effect of antibiotic use on treatment effect.

Results

Of 37 individuals randomized (age =69±10 years old; 57% men; eGFR=24±8 ml/min per 1.73 m2), 31 completed the study. Synbiotic therapy did not significantly reduce serum IS (−2 μmol/L; 95% confidence interval [95% CI], −5 to 1 μmol/L) but did significantly reduce serum PCS (−14 μmol/L; 95% CI, −27 to −2 μmol/L). Decreases in both PCS and IS concentrations were more pronounced in patients who did not receive antibiotics during the study (n=21; serum PCS, −25 μmol/L; 95% CI, −38 to −12 μmol/L; serum IS, −5 μmol/L; 95% CI, −8 to −1 μmol/L). Synbiotics also altered the stool microbiome, particularly with enrichment of Bifidobacterium and depletion of Ruminococcaceae. Except for an increase in albuminuria of 38 mg/24 h (P=0.03) in the synbiotic arm, no changes were observed in the other secondary outcomes.

Conclusion

In patients with CKD, synbiotics did not significantly reduce serum IS but did decrease serum PCS and favorably modified the stool microbiome. Large–scale clinical trials are justified.  相似文献   

16.
W D Rees  L C Gibbons  L A Turnberg 《Gut》1983,24(9):784-789
The effects of non-steroidal anti-inflammatory drugs and prostaglandins E2 and F on the secretory and electrical activity of isolated rabbit fundic mucosa have been studied. Spontaneous acid secretion was inhibited by serosal side application of sodium thiocyanate (6×10−2M) and the resulting alkali secretion measured by pH stat tiration. Serosal side application of indomethacin (10−5M) or aspirin (3×10−3M) inhibited alkali secretion (0·55±0·06 to 0·12±0·06 μmol/cm2/h, n=6, p<0·01 and 0·28±0·06 to 0·11±0·03 μmol/cm2/h, n=7, p<0·02 respectively). Mucosal or serosal side prostaglandin E2 (10−5 to 10−10M) and F (10−4 to 10−10M) failed to alter the rate of alkalinisation but secretion was significantly increased by serosal side 16,16-dimethyl-prostaglandin E2 (10−6M) (0·90±0·20 to 1·50±0·30 μmol/cm2/h, n=6, p<0·01). Serosal side application of 10−6M prostaglandin E2 to fundic mucosae pretreated with either aspirin (5×10−3M) or indomethacin (10−5M), to reduce endogenous E2 formation, also failed to alter alkali secretion. Pretreatment of the mucosa with 16,16-dimethyl-E2 (10−6M) abolished the inhibitory effect of indomethacin (10−5M) on alkali secretion (n=6) but did not modify the secretory response to aspirin (3×10−3M) (fall in alkali secretion with aspirin = 81±11% and with aspirin plus 16,16-dimethyl-E2 = 72±10%, n=7). In the doses used, none of the prostaglandins or non-steroidal anti-inflammatory drugs altered transmucosal potential difference or electrical resistance. These results show that the damaging agents, aspirin and indomethacin, both inhibit gastric alkali secretion but that modes of action may differ. The observation that prostaglandins, E2 and F failed to increase alkali production suggests that their protective activity against a variety of damaging agents as shown by others, may be mediated by another mechanism.  相似文献   

17.

BACKGROUND:

Despite the widespread clinical use of cyclooxygenase (COX) inhibitors, dilemmas regarding the potential impact of these drugs on the cardiovascular system persist.

OBJECTIVE:

To estimate the effects of different COX inhibitors (meloxicam, acetylsalicylic acid [ASA] and SC-560) on cardiac function and coronary flow in isolated rat hearts, with special focus on the L-arginine/nitric oxide system.

METHODS:

The hearts of eight-week-old male Wistar albino rats (n=72; 12 rats per group; body mass 180 g to 200 g) were retrogradely perfused according to the Langendorff technique at gradually increased perfusion pressure (40 cmH2O to 120 cmH2O). After control experiments, the hearts were perfused with the following drugs: 100 μM ASA, alone or in combination with 30 μM N(ω)-nitro-L-arginine monomethyl ester (L-NAME), 0.3 μM meloxicam with or without 30 μM L-NAME, 3 μM meloxicam with or without 30 μM L-NAME, 30 μM L-NAME and 0.25 μM SC-560. In the control and experimental groups, the following parameters of heart function were continuously recorded: maximum rate of left ventricular pressure development, minimum rate of left ventricular pressure development, systolic left ventricular pressure, diastolic left ventricular pressure, heart rate and mean blood pressure. Coronary flow was measured flowmetrically. The amount of released NO2 was determined spectrophotometrically in coronary venous effluent.

RESULTS:

While meloxicam and SC-560 were found to have an adverse influence on cardiac function and coronary perfusion, ASA did not negatively affect the intact model of the heart.

CONCLUSION:

It appeared that interaction between COX and the L-arginine/nitric oxide system truly exists in coronary circulation and may explain the causes of the observed effects.  相似文献   

18.
The effects of lobeline and tubocurarine on the voltage-clamped endplates of frog sartorius and cutaneous pectoris muscles were examined at room temperature (20-23°C). Like tubocurarine, lobeline causes nondepolarizing neuromuscular blockade. The half-time of decay (t½) of endplate currents (e.p.c.s) recorded at a holding potential (Vm) of -90 mV was significantly shorter in endplates treated with lobeline (50 μM; mean t½ ± SEM = 0.41 ± 0.02 ms) or tubocurarine (11.4 μM; t½ = 0.64 ± 0.04 ms) than in those treated with Mg2+ (13 mM; t½ = 1.39 ± 0.11 ms) or a low concentration of tubocurarine (3 μM; t½ = 0.87 ± 0.05 ms). Similarly, lobeline (10 μM) shortened the t½ of untreated miniature e.p.c.s by 35%; tubocurarine, however, abolished miniature e.p.c.s at the concentration required to observe its actions on e.p.c. decay kinetics. The t½ of e.p.c.s recorded from preparations treated with Mg2+ (13 mM), tubocurarine at low concentrations (3 μM), or untreated miniature e.p.c.s was logarithmically related to Vm, being slower at more hyperpolarized values. By contrast, the t½s of e.p.c.s recorded in either lobeline (50 μM) or tubocurarine (11.4 μM) were independent of voltage in the range -150 to -80 mV. The ability of lobeline to shorten t½ and to remove the voltage dependence of t½ was partially antagonized by Mg2+ (13 mM). As expected, when lobeline or tubocurarine was removed from the bath or when acetylcholine release from the motor nerve terminals was increased by 4-aminopyridine (20 μM) and Ca2+ (10 mM) (in the presence of lobeline or tubocurarine), the amplitude of e.p.c.s increased as a function of time. However, the t½ of the decay phase of the e.p.c.s remained shortened (i.e., unaltered from the earlier treatment). These results suggest that both tubocurarine and lobeline have at least two distinct postjunctional actions including: (i) a block of the acetylcholine receptor and (ii) a block of the ionic channel associated with the acetylcholine receptor.  相似文献   

19.
Fast time-resolved infrared spectroscopic measurements have allowed precise determination of the rates of activation of alkanes by Cp′Rh(CO) (Cp = η5-C5H5 or η5-C5Me5). We have monitored the kinetics of C─H activation in solution at room temperature and determined how the change in rate of oxidative cleavage varies from methane to decane. The lifetime of CpRh(CO)(alkane) shows a nearly linear behavior with respect to the length of the alkane chain, whereas the related Cp*Rh(CO)(alkane) has clear oscillatory behavior upon changing the alkane. Coupled cluster and density functional theory calculations on these complexes, transition states, and intermediates provide the insight into the mechanism and barriers in order to develop a kinetic simulation of the experimental results. The observed behavior is a subtle interplay between the rates of activation and migration. Unexpectedly, the calculations predict that the most rapid process in these Cp′Rh(CO)(alkane) systems is the 1,3-migration along the alkane chain. The linear behavior in the observed lifetime of CpRh(CO)(alkane) results from a mechanism in which the next most rapid process is the activation of primary C─H bonds (─CH3 groups), while the third key step in this system is 1,2-migration with a slightly slower rate. The oscillatory behavior in the lifetime of Cp*Rh(CO)(alkane) with respect to the alkane’s chain length follows from subtle interplay between more rapid migrations and less rapid primary C─H activation, with respect to CpRh(CO)(alkane), especially when the CH3 group is near a gauche turn. This interplay results in the activation being controlled by the percentage of alkane conformers.  相似文献   

20.
We set out to study the use of a series of ruthenocenes as possible and promising sources for ruthenium and/or ruthenium oxide film formation.The thermal stability of a series of ruthenocenes, including (η5-C5H4R)(η5-C5H4R´)Ru (1), R = R´ = H (3), R = H, R´ = CH2NMe2 (5), R = H, R´= C(O)Me (6), R = R´ = C(O)Me (7), R = H, R´ = C(O)(CH2)3CO2H (8), R = H, R´ = C(O)(CH2)2CO2H (9), R = H, R´ = C(O)(CH2)3CO2Me (10), R = H, R´= C(O)(CH2)2CO2Me (11), R = R´ = SiMe3), (η5-C4H3O-2,4-Me2)2Ru (2), and (η5-C5H5-2,4-Me2)2Ru (4) was studied by thermogravimetry. From these studies, it could be concluded that 1–4, 6 and 9–11 are the most thermally stable molecules. The sublimation pressure of these sandwich compounds was measured using a Knudsen cell. Among these, the compound 11 shows the highest vapor pressure.  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