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 93 毫秒
1.
N-(Morpholinothiocarbonyl) benzamide (C(12)H(14)N(2)O(2)S) and N-(piperidylthiocarbonyl) benzamide (C(13)H(16)N(2)OS) and their Co(III) complexes have been synthesized and characterized by elemental analysis, FTIR and NMR methods. The complex Co(C(12)H(14)N(2)O(2)S)(3), crystallizes in the triclinic space group P1, with Z=2, and unit cell parameters, a=12.080(7)A, b=12.195(7)A, c=13.025(6)A, alpha=90.198(7) degrees, beta=95.721(7) degrees, gamma=106.426(9) degrees, V=1830.4(17)A(3). The antifungal activity against the major pathogens responsible for important plant diseases (Botrytis cinerea, Myrothecium and Verticillium dahliae dleb) of N-(piperidylthiocarbonyl) benzamide and its complex with cobalt (III) are studied and compared with N-(morpholinothiocarbonyl) benzamide.  相似文献   

2.
Ex vivo ?(13)C, (2)H? NMR spectroscopy allowed to estimate the relative sizes of neuronal and glial glutamate pools and the relative contributions of (1-(13)C) glucose and (2-(13)C, 2-(2)H(3)) acetate to the neuronal and glial tricarboxylic acid cycles of the adult rat brain. Rats were infused during 60 min in the right jugular vein with solutions containing (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose or (2-(13)C, 2-(2)H(3)) acetate only. At the end of the infusion the brains were frozen in situ and perchloric acid extracts were prepared and analyzed by high resolution (13)C NMR spectroscopy (90.5 MHz). The relative sizes of the neuronal and glial glutamate pools and the contributions of acetyl-CoA molecules derived from (2-(13)C, (2)H(3)) acetate or (1-(13)C) glucose entering the tricarboxylic acid cycles of both compartments, could be determined by the analysis of (2)H-(13)C multiplets and (2)H induced isotopic shifts observed in the C4 carbon resonances of glutamate and glutamine. During the infusions with (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose, the glial glutamate pool contributed 9% of total cerebral glutamate being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (4%), (2-(13)C) acetyl-CoA (3%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (2%). The neuronal glutamate pool accounted for 91% of the total cerebral glutamate being mainly originated from (2-(13)C) acetyl-CoA (86%) and (2-(13)C, 2-(2)H) acetyl-CoA (5%). During the infusions of (2-(13)C, 2-(2)H(3)) acetate only, the glial glutamate pool contributed 73% of the cerebral glutamate, being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (36%), (2-(13)C, 2-(2)H) acetyl-CoA (27%) and (2-(13)C) acetyl-CoA (10%). The neuronal pool contributed 27% of cerebral glutamate being formed from (2-(13)C) acetyl-CoA (11%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (16%). These results illustrate the potential of ?(13)C, (2)H? NMR spectroscopy as a novel approach to investigate substrate selection and metabolic compartmentation in the adult mammalian brain.  相似文献   

3.
The carboxyl-terminal domain of RNA polymerase II, which is rich in phosphorylation sites, contains 17--52 tandem repeats with the consensus sequence of the heptapeptide, YSPTSPS. The repeat unit of the heptapeptide has two SPXX motifs showing potential beta-turns, SPTS and SPSY. NMR studies were performed in water at pH 4.0 for two cyclic peptides containing one and two repeat units, cyclo-[C(1)R(2)D(3)Y(4)S(5)P(6)T(7)S(8)P(9)S(10)Y(11)S(12)R(13)D(14)C(15)] (peptide 1) and cyclo-[C(1)R(2)D(3)Y(4)S(5)P(6)T(7)S(8)P(9)S(10)Y(11)S(12)P(13)T(14)S(15)P(16)N(17)Y(18)S(19)R(20)D(21)C(22)] (peptide 2), which are cyclized with a disulfide bridge of two Cys residues at the N- and C-termini. SP in 1 and 2 are predominantly in trans form. The following NMR parameters were detected: (1) lower temperature coefficients of amide proton chemical shifts of T7 and S8 in 1, and Tx (T7 or T14), Sx (S8 or S15), Tz (T14 or T7) and Sz (S15 or S8) in 2, (2) significantly large deviation of H(alpha) chemical shifts from its random coil value (Delta H(alpha)) of Pro preceding the Thr (P6 in 1, and Px and Pz in 2), (3) relatively large (3)J(HNH alpha) coupling constants (>8.7 Hz) of T7 in 1 and Tx and Tz in 2, and (4) NOE (d(NN) (i, i+1)) connectivities between the amide protons of T7-S8 and S10-Y11 in 1, and Tx-Sx, S10-Y11, Tz-Sz, and N17-Y18 in 2, although two Pro-Thr-Ser segments in 2 (each of these are annotated by 'x' and 'z') in the first and second repeat units were not distinguishable. Comparison of the NMR parameters between the cyclic peptides and the corresponding linear peptides indicates that cyclization promotes structural stabilization in water. The present NMR data were consistent with the presence of a beta-turn at both SPTS and SPSY: S(5)P(6)T(7)S(8) and S(8)P(9)S(10)Y(11) in 1, and SPxTxSx, SPzTzSz, SP(9)S(10)Y(11), SP(16)N(17)Y(18) in 2. However, the structure of the SPTS segment is more stable than that of the SPSY segment. Conformations consistent with NMR parameters including NOE distances were obtained through molecular dynamics and energy minimization methods. These calculations yielded two stable conformers for the SPTS segment. One of the two corresponds to a type I beta-turn.  相似文献   

4.
Nan YH  Shin SY 《BMB reports》2011,44(11):747-752
To investigate the effects of disulphide bond position on the salt resistance and lipopolysaccharide (LPS)-neutralizing activity of α-helical homo-dimeric antimicrobial peptides (AMPs), we synthesized an α-helical model peptide (K6L4W1) and its homo-dimeric peptides (di-K(6)L(4)W(1)-N, di-K(6)L(4)W(1)-M, and di-K(6)L(4)W(1)-C) with a disulphide bond at the N-terminus, the central position, and the C-terminus of the molecules, respectively. Unlike (6)L(4)W(1) and di-K(6)L(4)W(1)-M, the antimicrobial activity of di-K(6)L(4)W(1)-N and di-K(6)L(4)W(1)-C was unaffected by 150 mM NaCl. Both di-K(6)L(4)W(1)-N and di-K(6)L(4)W(1)-C caused much greater inhibitory effects on nitric oxide (NO) release in LPS-induced mouse macrophage RAW 264.7 cells, compared to di-K(6)L(4)W(1)-M. Taken together, our results indicate that the presence of a disulphide bond at the N- or C-terminus of the molecule, rather than at the central position, is more effective when designing salt-resistant α-helical homo-dimeric AMPs with potent antimicrobial and LPS-neutralizing activities. [BMB reports 2011; 44(11): 747-752].  相似文献   

5.
Complexes [Au(2)(H(2)Gy3DH)(2)]Cl(2) (1), [Au(H(2)Gy3Me)]Cl(3) (2) and [Au(H(2)Gy3Et)]Cl(3) (3) were obtained with glyoxaldehyde bis(thiosemicarbazone) (H(2)Gy3DH) and its N(3)-methyl (H(2)Gy3Me) and N(3)-ethyl (H(2)Gy3Et) derivatives. The bis(thiosemicarbazones) and their gold(I) and gold(III) complexes exhibited anti-proliferative activity against HL-60, Jurkat (leukemia) and MCF-7 (breast cancer) cells at 10 μmol L(-1). Complex (2) was able to in vitro inhibit thioredoxin reductase (TrxR) activity, which suggests that inhibition of TrxR could be part of its mechanism of action.  相似文献   

6.
The family of gamma-aminobutyric acid type A receptors (GABA(A)Rs) mediates two types of inhibition in the mammalian brain. Phasic inhibition is mediated by synaptic GABA(A)Rs that are mainly comprised of alpha(1), beta(2), and gamma(2) subunits, whereas tonic inhibition is mediated by extrasynaptic GABA(A)Rs comprised of alpha(4/6), beta(2), and delta subunits. We investigated the activation properties of recombinant alpha(4)beta(2)delta and alpha(1)beta(2)gamma(2S) GABA(A)Rs in response to GABA and 4,5,6,7-tetrahydroisoxazolo[5,4-c]pyridin-3(2H)-one (THIP) using electrophysiological recordings from outside-out membrane patches. Rapid agonist application experiments indicated that THIP produced faster opening rates at alpha(4)beta(2)delta GABA(A)Rs (beta approximately 1600 s(-1)) than at alpha(1)beta(2)gamma(2S) GABA(A)Rs (beta approximately 460 s(-1)), whereas GABA activated alpha(1)beta(2)gamma(2S) GABA(A)Rs more rapidly (beta approximately 1800 s(-1)) than alpha(4)beta(2)delta GABA(A)Rs (beta < 440 s(-1)). Single channel recordings of alpha(1)beta(2)gamma(2S) and alpha(4)beta(2)delta GABA(A)Rs showed that both channels open to a main conductance state of approximately 25 pS at -70 mV when activated by GABA and low concentrations of THIP, whereas saturating concentrations of THIP elicited approximately 36 pS openings at both channels. Saturating concentrations of GABA elicited brief (<10 ms) openings with low intraburst open probability (P(O) approximately 0.3) at alpha(4)beta(2)delta GABA(A)Rs and at least two "modes" of single channel bursting activity, lasting approximately 100 ms at alpha(1)beta(2)gamma(2S) GABA(A)Rs. The most prevalent bursting mode had a P(O) of approximately 0.7 and was described by a reaction scheme with three open and three shut states, whereas the "high" P(O) mode ( approximately 0.9) was characterized by two shut and three open states. Single channel activity elicited by THIP in alpha(4)beta(2)delta and alpha(1)beta(2)gamma(2S) GABA(A)Rs occurred as a single population of bursts (P(O) approximately 0.4-0.5) of moderate duration (approximately 33 ms) that could be described by schemes containing two shut and two open states for both GABA(A)Rs. Our data identify kinetic properties that are receptor-subtype specific and others that are agonist specific, including unitary conductance.  相似文献   

7.
One-electron oxidation of (6R)-5,6,7,8-tetrahydrobiopterin (H(4)B) by the azide radical generates the radical cation (H(4)B(*)(+)) which rapidly deprotonates at physiological pH to give the neutral trihydrobiopterin radical (H(3)B(*)); pK(a) (H(4)B(*)(+) <==> H(3)B(*) + H(+)) = (5.2 +/- 0.1). In the absence of ascorbate both the H(4)B(*)(+) and H(3)B(*) radicals undergo disproportionation to form quinonoid dihydrobiopterin (qH(2)B) and the parent H(4)B with rate constants k(H(4)B(*)(+) + H(4)B(*)(+)) = 6.5 x 10(3) M(-1) s(-1) and k(H(3)B(*) + H(3)B(*)) = 9.3 x 10(4) M(-1) s(-1), respectively. The H(3)B(*) radical is scavenged by ascorbate (AscH(-)) with an estimated rate constant of k(H(3)B(*) + AscH(-)) similar 1.7 x 10(5) M(-1) s(-1). At physiological pH the pterin rapidly scavenges a range of biological oxidants often associated with cellular oxidative stress and nitric oxide synthase (NOS) dysfunction including hydroxyl ((*)OH), nitrogen dioxide (NO(2)(*)), glutathione thiyl (GS(*)), and carbonate (CO(3)(*-)) radicals. Without exception these radicals react appreciably faster with H(4)B than with AscH(-) with k(*OH + H(4)B) = 8.8 x 10(9) M(-1) s(-1), k(NO(2)(*) + H(4)B) = 9.4 x 10(8) M(-1) s(-1), k(CO(3)(*-) + H(4)B) = 4.6 x 10(9) M(-1) s(-1), and k(GS(*) + H(4)B) = 1.1 x 10(9) M(-1) s(-1), respectively. The glutathione disulfide radical anion (GSSG(*-)) rapidly reduces the pterin to the tetrahydrobiopterin radical anion (H(4)B(*-)) with a rate constant of k(GSSG(*-) + H(4)B) similar 4.5 x 10(8) M(-1) s(-1). The results are discussed in the context of the general antioxidant properties of the pterin and the redox role played by H(4)B in NOS catalysis.  相似文献   

8.
Yttrium and indium complexes of 1,4,7,10-tetraaza-4,7,10-tris(carboxymethyl)-1-cyclododecylacetylbenzylamine (DOTA-BA) and 1,4,7,10-tetraaza-4,7,10-tris(carboxymethyl)-1-cyclododecylacetyl-R-(+)-alpha-methylbenzylamine (DOTA-MBA) were prepared in order to study solution structures of (90)Y- and (111)In-labeled DOTA-biomolecule conjugates. (90)Y and (111)In complexes M(L) (M = (90)Y and (111)In; L = DOTA-BA and DOTA-MBA) were prepared from the reaction of MCl(3) with DOTA-BA and DOTA-MBA, respectively, in ammonium acetate buffer. A reverse phase HPLC method revealed that both (90)Y and (111)In complexes show only one radiometric peak in their radio-HPLC chromatograms. It was also found that (111)In(DOTA-BA) and (111)In(DOTA-MBA) are more hydrophilic than their corresponding (90)Y analogues, suggesting different coordination spheres in (111)In and (90)Y complexes of the same DOTA conjugate. Complexes M(L) (M = Y and In; L = DOTA-BA and DOTA-MBA) were prepared and characterized by HPLC, LC-MS, and NMR ((1)H and (13)C) methods. The HPLC concordance experiments for (90)Y(DOTA-MBA)/Y(DOTA-MBA) and (111)In(DOTA-MBA)/In(DOTA-MBA) show that the same complex is prepared at both tracer and macroscopic levels. The NMR data ((1)H and (13)C) clearly demonstrates that Y(DOTA-BA) and Y(DOTA-MBA) exist in solution as one predominant isomer. VT NMR data ((1)H and (13)C) show that In(DOTA-BA) and In(DOTA-MBA) are fluxional at room temperature while Y(DOTA-BA) and Y(DOTA-MBA) become fluxional only at elevated temperatures. The fluxionality of these complexes is due to rapid rotation of acetate/acetamide chelating arms and inversion of ethylenic groups of the macrocyclic ring.  相似文献   

9.
Shuxing Z  Ying WS  Siahaan TJ  Jois SD 《Peptides》2003,24(6):827-835
Cell-adhesion molecules are critical for immune response. It is well known that the inhibition of adhesion is very effective in immunotherapy and that the peptides derived from leukocyte function associated antigen (LFA-1) and intercellular adhesion molecule (ICAM-1) modulate cell-adhesion interaction. The three-dimensional structure of a cyclic peptide, Cyclo(1,12)Pen(1)-Asp(2)-Leu(3)-Ser(4)-Tyr(5)-Ser(6)-Leu(7)-Asp(8)-Asp(9)-Leu(10)-Arg(11)-Cys(12) (cLBEL) derived from the beta subunit of LFA-1 which is known to modulate homotypic T-cell-adhesion process has been studied using NMR, CD and molecular dynamics (MD) simulation. The peptide exhibits two possible conformations in solution. Structure I has a conformation with two consecutive beta-turns involving residues Tyr(5)-Ser(6)-Leu(7)-Asp(8) and Asp(9)-Leu(10)-Arg(11)-Cys(12). Structure II has a beta-turn at Tyr(5)-Ser(6)-Leu(7)-Asp(8) and forms a beta-hairpin type of conformation.  相似文献   

10.
The binding of carbon dioxide by horse haemoglobin   总被引:15,自引:7,他引:8  
1. Three modified horse haemoglobins have been prepared: (i) alpha(c) (2)beta(c) (2), in which both the alpha-amino groups of the alpha- and beta-chains have reacted with cyanate, (ii) alpha(c) (2)beta(2), in which the alpha-amino groups of the alpha-chains have reacted with cyanate, and (iii) alpha(2)beta(c) (2), in which the two alpha-amino groups of the beta-chain have reacted with cyanate. 2. The values of n (the Hill constant) for alpha(c) (2)beta(c) (2), alpha(2)beta(c) (2) and alpha(c) (2)beta(2) were (respectively) 2.5, 2.0 and 2.6, indicating the presence of co-operative interactions between the haem groups for all derivatives. 3. In the alkaline pH range (about pH8.0) all the derivatives show the same charge as normal haemoglobin whereas in the acid pH range (about pH6.0) alpha(c) (2)beta(c) (2) differs by four protonic charges and alpha(c) (2)beta(2), alpha(2)beta(c) (2) by two protonic charges from normal haemoglobin, indicating that the expected number of ionizing groups have been removed. 4. alpha(c) (2)beta(2) and alpha(c) (2)beta(c) (2) show a 25% decrease in the alkaline Bohr effect, in contrast with alpha(2)beta(c) (2), which has the same Bohr effect as normal haemoglobin. 5. The deoxy form of alpha(c) (2)beta(c) (2) does not bind more CO(2) than the oxy form of alpha(c) (2)beta(c) (2), whereas alpha(c) (2)beta(2) and alpha(2)beta(c) (2) show intermediate binding. 6. The results reported confirm the hypothesis that, under physiological conditions, haemoglobin binds CO(2) through the four terminal alpha-amino groups and that the two terminal alpha-amino groups of alpha-chains are involved in the Bohr effect.  相似文献   

11.
The binding properties of a spacer-linked synthetic Sd(a) tetrasaccharide beta-D-GalpNAc-(1-->4)-alpha-Neu5Ac-(2-->3)]-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (1), two tetrasaccharide mimics beta-D-Galp-(1-->4)-alpha-Neu5Ac-(2-->3)]-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (2) and beta-D-GlcpNAc-(1-->4)-alpha-Neu5Ac-(2-->3)]-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (3), and two trisaccharide mimics beta-D-GalpNAc-(1-->4)-3-O-(SO(3)H)-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (4) and beta-D-GalpNAc-(1-->4)-3-O-(CH(2)COOH)-beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->O)-(CH(2))(5)-NH(2) (5) with lectins from Dolichos biflorus (DBL), Maackia amurensis (MAL), Phaseolus limensis (PLL), Ptilota plumosa (PPL), Ricinus communis 120 (RCL120) and Triticum vulgaris (wheat germ agglutinin, WGA) have been investigated by surface plasmon resonance (SPR) detection. MAL, PPL, RCL120 and WGA did not display any binding activity with compounds 1-5. However, DBL and PLL, both exhibiting GalNAc-specificity, showed strong binding activity with compounds 1, 4 and 5, and 1, 3, 4 and 5, respectively. The results demonstrate that SPR is a very useful analysis system for identifying biologically relevant oligosaccharide mimics of the Sd(a) determinant.  相似文献   

12.
Complexes [Bi(2Fo4Ph)Cl(2) ] (1), [Bi(2Ac4Ph)Cl(2) ] (2), [Bi(2Bz4Ph)Cl(2) ] (3), [Bi(H(2) Gy3DH)Cl(3) ] (4), [Bi(H(2) Gy4Et)(OH)(2) Cl] (5), and [Bi(H(2) Gy4Ph)Cl(3) ] (6) were prepared with pyridine-2-carbaldehyde 4-phenylthiosemicarbazone (H2Fo4Ph), 1-(pyridin-2-yl)ethanone 4-phenylthiosemicarbazone (H2Ac4Ph), phenyl(pyridin-2-yl)methanone 4-phenylthiosemicarbazone (H2Bz4Ph), as well as with glyoxaldehyde bis(thiosemicarbazone) (H(2) Gy4DH) and its 4-Et (H(2) Gy4Et) and 4-Ph (H(2) Gy4Ph) derivatives. The complexes exhibited antibacterial activities against Staphylococcus aureus, Staphylococcus epidermidis, Enterococcus faecalis, and Pseudomonas aeruginosa. Coordination to Bi(III) proved to be an effective strategy to increase the antibacterial activity of the thiosemicarbazones and bis(thiosemicarbazones).  相似文献   

13.
Five new organotin(IV) molecules with the heterocyclic thioamides; 2-mercaptobenzothiazole (Hmbzt), 5-chloro-2-mercaptobenzothiazole (Hcmbzt), 3-methyl-2-mercaptobenzothiazole (mmbzt) and 2-mercaptonicotinic acid (H(2)mna) of formulae [(n-C(4)H(9))(2)Sn(mbzt)(2)] (1), [(C(6)H(5))(2)Sn(mbzt)(2)] (2), [(CH(3))(2)Sn(cmbzt)(2)].1.7(H(2)O)] (3), [(n-C(4)H(9))(2)SnCl(2)(mmbzt)(2).(CH(2)Cl(2))] (4) and [[(C(6)H(5))(3)Sn](2)(mna).[(CH(3))(2)CO]] (5) have been synthesized and characterized by elemental analysis, 1H-, 13C-NMR, FT-IR and M?ssbauer spectroscopic techniques. Crystal structures of molecules 1, 3 and 5 have been determined by X-ray diffraction at 173(1) K (1 and 5) and 293(2) K (3). Compound 1 C(22)H(26)N(2)S(4)Sn, is monoclinic, space group C2/c, a=44.018(2), b=8.8864(5), c=12.8633(7) A, beta=104.195(5) degrees, Z=8. Compound 3 is also monoclinic, space group P2(1)/c and a=17.128(2) A, b=17.919(2) A, c=7.3580(10) A, beta=98.290(10) degrees, Z=4. In both molecules 1 and 3, two carbon atoms from aryl groups, two sulfur and two nitrogen atoms from thione ligands form a distorted octahedral geometry around tin(IV) with trans-C(2), cis-N(2), cis-S(2) configurations. Compound 5 C(45)H(39)NO(3)SSn(2) is monoclinic, space group P2(1)/n, a=9.1148(2) A, b=29.2819(6), c=15.5556(4) A, beta=106.2851(9) degrees, Z=4. Complex 5 contains two [(C(6)H(5))(3)Sn(IV)] moieties linked by a double deprotonated 2-mercaptonicotinic acid (H(2)mna). Both tin(IV) ions are five coordinated. This complex is the an example of a pentacoordinated Ph(3)SnXY system with an axial-equatorial arrangement of the phenyl groups at Sn(1) atom. Compounds 1, 3 and 5 were tested for in vitro cytotoxicity against the cancer cell line of sarcoma cells (mesenchymal tissue) from the Wistar rat, polycyclic aromatic hydrocarbons (benzo[a]pyrene) carcinogenesis. Compound 5 exhibits strong cytotoxic activity, while complexes 1 and 3 show less cytotoxic activity.  相似文献   

14.
Togrul M  Turgut Y  Hoşgören H 《Chirality》2004,16(6):351-355
Chiral monoaza-15-crown-5 ethers (1, 2) were prepared from (R)-(-)-2-amino-1-butanol in high yield. The chiral monoaza-15-crown-5 ethers were purified directly as NaClO(4) complexes. Molecular recognition by these chiral monoaza-crown ethers of (R)- and (S)-PhEtHClO(4) and (R)- and (S)-NapEtHClO(4) as characterized by UV-vis spectroscopy. The order of enantiomeric selectivity is (R)- > (S)- PhEtHClO(4) and (S)- > (R)-NapEtHClO(4) for 1. In the case of 2 it was (R)- > (S)-PhEtHClO(4) and (R)- > (S)- NapEtHClO(4). The cavity of macrocycle and steric hindrance of the benzene units appears to play an important role in recognition.  相似文献   

15.
The G-protein G(salpha) exists in three isoforms, the G(salpha) splice variants G(salphashort) (G(salphaS)) and G(salphalong) (G(salphaL)), and the G-protein G(alphaolf) that is not only involved in olfactory signaling but also in extrapyramidal motor regulation. Studies with beta(2)-adrenoceptor (beta(2)AR)-G(salpha) fusion proteins showed that G(salpha) proteins activate adenylyl cyclase (AC) in the order of efficacy G(salphaS)>G(salphaL) approximately G(alphaolf) and that G(salpha) proteins confer the hallmarks of constitutive activity to the beta(2)AR in the order of efficacy G(salphaL)>G(alphaolf)>G(salphaS). However, it is unclear whether such differences between G(salpha) proteins also exist in the nonfused state. In the present study, we co-expressed the beta(2)AR and dopamine D(1)-receptor (D(1)R) with G(salpha) proteins at different ratios in Sf9 insect cells. In agreement with the fusion protein studies, nonfused G(alphaolf) was less efficient than nonfused G(salphaS) and G(salphaL) at activating AC, but otherwise, we did not observe differences between the three G(salpha) isoforms. Thus, it is much easier to dissect differences between G(salpha) isoforms using beta(2)AR-G(salpha) fusion proteins than nonfused G(salpha) isoforms.  相似文献   

16.
Wei C  Jia G  Yuan J  Feng Z  Li C 《Biochemistry》2006,45(21):6681-6691
Free-base porphyrin (5,10,15,20-tetrakis(1-methyl-4-pyridyl)-21H,23H-porphine) (H(2)TMPyP4) has been shown to be an effective telomerase inhibitor by an in vitro assay. Here, we examined the interactions of the H(2)TMPyP4 with three distinct G-quadruplex DNAs, the parallel-stranded (TG(4)T)4, dimer-hairpin-folded (G(4)T(4)G(4))2, and monomer-folded AG(3)(T(2)AG(3))(3), by ultraviolet resonance Raman spectroscopy (UVRR), UV-vis absorption spectroscopy, fluorescence spectroscopy, and surface-enhanced Raman spectroscopy (SERS). The data obtained by the continuous variation titration method show that the binding stoichiometry of H(2)TMPyP4/G-quadruplex is 2:1 for (TG(4)T)4 and 4:1 for (G(4)T(4)G(4))2 or AG(3)(T(2)AG(3))(3). The results of SERS spectra, UV-vis absorption titration, and fluorescence emission spectra together with the binding stoichiometries reveal that two H(2)TMPyP4 molecules are externally stacked at two ends of the parallel (TG(4)T)4 G-quadruplex, whereas H(2)TMPyP4 molecules can intercalate within their diagonal or lateral loop regions and intervals between two G-tetrads for (G(4)T(4)G(4))2 and AG(3)(T(2)AG(3))(3) G-quadruplexes. The binding of H(2)TMPyP4 to (TG(4)T)4 G-quadruplex results in the hypochromicity of the UV Raman signal of (TG(4)T)4, indicating that the stacking effects between H(2)TMPyP4 and DNA bases are significant. The Raman hyperchromicities and shifts are observed after the binding of H(2)TMPyP4 to both (G(4)T(4)G(4))2 and AG(3)(T(2)AG(3))(3) G-quadruplexes. This indicates that the intercalative H(2)TMPyP4 can lengthen the vertical distance between adjacent G-tetrads of (G(4)T(4)G(4))2 and AG(3)(T(2)AG(3))(3) and change their conformations. The present study provides new insights into the effect of H(2)TMPyP4 binding on the structures of G-quadruplexes and also demonstrates that Raman spectroscopy is an ideal method for examining the interaction between drugs and G-quadruplexes.  相似文献   

17.
The synthesis and characterization of complexes [(η(6)-arene)Ru(N,N')X][PF(6)], where arene is para-cymene (p-cym), biphenyl (bip), ethyl benzoate (etb), hexamethylbenzene (hmb), indane (ind) or 1,2,3,4-tetrahydronaphthalene (thn), N,N' is 2,2'-bipyrimidine (bpm) and X is Cl, Br or I, are reported, including the X-ray crystal structures of [(η(6)-p-cym)Ru(bpm)I][PF(6)], [(η(6)-bip)Ru(bpm)Cl][PF(6)], [(η(6)-bip)Ru(bpm)I][PF(6)] and [(η(6)-etb)Ru(bpm)Cl][PF(6)]. Complexes in which N,N' is 1,10-phenanthroline (phen), 1,10-phenanthroline-5,6-dione or 4,7-diphenyl-1,10-phenanthroline (bathophen) were studied for comparison. The Ru(II) arene complexes undergo ligand-exchange reactions in aqueous solution at 310?K; their half-lives for hydrolysis range from 14 to 715?min. Density functional theory calculations on [(η(6)-p-cym)Ru(bpm)Cl][PF(6)], [(η(6)-p-cym)Ru(bpm)Br][PF(6)], [(η(6)-p-cym)Ru(bpm)I][PF(6)], [(η(6)-bip)Ru(bpm)Cl][PF(6)], [(η(6)-bip)Ru(bpm)Br][PF(6)] and [(η(6)-bip)Ru(bpm)I][PF(6)] suggest that aquation occurs via an associative pathway and that the reaction is thermodynamically favourable when the leaving ligand is I?>?Br?≈?Cl. pK (a)* values for the aqua adducts of the complexes range from 6.9 to 7.32. A binding preference for 9-ethylguanine (9-EtG) compared with 9-ethyladenine (9-EtA) was observed for [(η(6)-p-cym)Ru(bpm)Cl][PF(6)], [(η(6)-hmb)Ru(bpm)Cl](+), [(η(6)-ind)Ru(bpm)Cl](+), [(η(6)-thn)Ru(bpm)Cl](+), [(η(6)-p-cym)Ru(phen)Cl](+) and [(η(6)-p-cym)Ru(bathophen)Cl](+) in aqueous solution at 310?K. The X-ray crystal structure of the guanine complex [(η(6)-p-cym)Ru(bpm)(9-EtG-N7)][PF(6)](2) shows multiple hydrogen bonding. Density functional theory calculations show that the 9-EtG adducts of all complexes are thermodynamically preferred compared with those of 9-EtA. However, the bmp complexes are inactive towards A2780 human ovarian cancer cells. Calf thymus DNA interactions for [(η(6)-p-cym)Ru(bpm)Cl][PF(6)] and [(η(6)-p-cym)Ru(phen)Cl][PF(6)] consist of weak coordinative, intercalative and monofunctional coordination. Binding to biomolecules such as glutathione may play a role in deactivating the bpm complexes.  相似文献   

18.
A kinetic study of CO(2) hydration was carried out using the water-soluble zinc model complex with water-soluble nitrilotris(2-benzimidazolylmethyl-6-sulfonate) L1S, [L1SZn(OH(2))](-), mimicking the active site of carbonic anhydrase, in the presence and absence of anion inhibitors NCS(-) and Cl(-). The obtained rate constants k(cat) for CO(2) hydration were 5.9x10(2), 1. 7x10(3), and 3.1x10(3) M(-1) s(-1) at 5, 10, and 15 degrees C, respectively: the k(cat)=ca. 10(4) M(-1) s(-1) extrapolated towards 25 degrees C has been the largest among the reported k(cat) using zinc model complexes for carbonic anhydrase. It was also revealed that NCS(-), Cl(-) and acetazolamide play a role of inhibitors by the decrease of k(cat): 7x10(2) and 2x10(3) M(-1) s(-1) for NCS(-) and Cl(-) at 15 degrees C, respectively. The sequence of their magnitudes in k(cat) is Cl(-) approximately acetazolamide>NCS(-), where the sequence Cl(-)>NCS(-) is confirmed for native carbonic anhydrase. The difference of k(cat) or k(obs) between NCS(-) and Cl(-) resulted from that between the stability constants K(st)=2x10(3) for [L1SZn(NCS)](2-) and 1x10(2) M(-1) for [L1SZnCl](2-) in D(2)O: for water-insoluble tris(2-benzimidazolylmethyl)amine L1, K(st)=1.8x10(4) for [L1Zn(NCS)](2-) and 1.5x10(3) M(-1) for [L1ZnCl](2-)in CD(3)CN/D(2)O (50% v/v). The crystal structure of anion-binding zinc model complexes [L1Zn(OH(2))](0.5)[L1ZnCl](0.5) (ClO(4))(1.5) 1(0.5)2(0.5)(ClO(4))(1.5) was revealed by X-ray crystallography. The geometry around Zn(2+) in 1 and 2 was tetrahedrally coordinated by three benzimidazolyl nitrogen atoms and one oxygen atom of H(2)O, or Cl(-).  相似文献   

19.
Shi YB  Fang JL  Liu XY  Du L  Tang WX 《Biopolymers》2002,65(2):81-88
The secondary structures of porcine brain Cu(4)Zn(3)-metallothionein (MT)-III and Cd(5)Zn(2)MT-I, Cd(5)Zn(2)MT-II, and Zn(7)MT-I from rabbit livers in the solid state are investigated by Fourier transform IR spectroscopy (FTIR) and Fourier transform Raman spectroscopy (FT-Raman). The Cu(4)Zn(3)MT-III contains 26-28% beta-turns and half-turns, 13-14% 3(10)-helices, 47-49% random coils, and 11-12% beta-extended chains. The structural comparison of porcine brain Cu(4)Zn(3)MT-III with rabbit liver Cd(5)Zn(2)MT-I (II) and Zn(7)MT-I shows that the contents of the random coil structure are obviously increased. The results indicate that the insert of an acidic hexapeptide in the alpha domain of Cu(4)Zn(3)MT-III possibly forms an alpha helix. However, because the bands assigned to the alpha-helix and random coil structures are overlapped in the spectra, the content of random coil structures in Cu(4)Zn(3)MT-III is therefore higher than those in Cd(5)Zn(2)MT-I, Cd(5)Zn(2)MT-II, and Zn(7)MT-I.  相似文献   

20.
Brearley CA  Hanke DE 《Plant physiology》2000,122(4):1209-1216
Using a permeabilization strategy to introduce Ins(3,4,5,6) P(4) into mesophyll protoplasts of Commelina communis, we have identified Ins(3,4,5,6) P(4) 1-kinase activity in mesophyll cells. Multiple InsP(3) isomers were identified in Spirodela polyrhiza and Arabidopsis. Only two of these, Ins(1,2,3) P(3) and Ins(3,4,6) P(3), have previously been identified in plants and only in monocots. The isomers detected in S. polyrhiza included D- and/or L-Ins(3,4,5) P(3), D- and/or L-Ins(3,5,6) P(3), and D- and/or L-Ins(2,4,5) P(3). Ins(1,4,5) P(3), if present, was only a tiny fraction of total InsP(3) species. We have also identified inositol polyphosphate phosphatase activities, Ins(3,4,5,6) P(4) 6-phosphatase and Ins(3,4, 5, 6) P(4) 4-phosphatase, whose action on endogenous inositol polyphosphates explains the presence of D- and/or L-Ins(3,4,5) P(3) and D- and/or L-Ins(3,5,6) P(3) in mesophyll cells. Inositol trisphosphates identified in Arabidopsis include Ins(1,2,3) P(3) and D- and/or L-Ins(3,4,6) P(3), suggesting that dicots may share pathways of InsP(6) biosynthesis and breakdown in common with monocots.  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