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.
Dissimilatory reduction of NO(2) to N(2)O and NH(4) by a soil Citrobacter sp. was studied in an attempt to elucidate the physiological and ecological significance of N(2)O production by this mechanism. In batch cultures with defined media, NO(2) reduction to NH(4) was favored by high glucose and low NO(3) concentrations. Nitrous oxide production was greatest at high glucose and intermediate NO(3) concentrations. With succinate as the energy source, little or no NO(2) was reduced to NH(4) but N(2)O was produced. Resting cell suspensions reduced NO(2) simultaneously to N(2)O and free extracellular NH(4). Chloramphenicol prevented the induction of N(2)O-producing activity. The K(m) for NO(2) reduction to N(2)O was estimated to be 0.9 mM NO(2), yet the apparent K(m) for overall NO(2) reduction was considerably lower, no greater than 0.04 mM NO(2). Activities for N(2)O and NH(4) production increased markedly after depletion of NO(3) from the media. Amendment with NO(3) inhibited N(2)O and NH(4) production by molybdate-grown cells but not by tungstate-grown cells. Sulfite inhibited production of NH(4) but not of N(2)O. In a related experiment, three Escherichia coli mutants lacking NADH-dependent nitrite reductase produced N(2)O at rates equal to the wild type. These observations suggest that N(2)O is produced enzymatically but not by the same enzyme system responsible for dissimilatory reduction of NO(2) to NH(4).  相似文献   

2.
The functional regulation of integrins is a major determinant of cell adhesion, migration and tissue maintenance. The binding of cytoskeletal proteins to various sites of integrin cytoplasmic domains is a key mechanism of this functional regulation. Expression of recombinant integrin alpha(IIb)beta(3) and alpha(M)beta(2) lacking the GFFKR-region in CHO cells results in constitutively activated integrins. In contrast, CHO cells stably expressing either a GFFKR-deleted alpha(V(del))beta(3) or a FF to AA-substituted alpha(V(AA))beta(3) do not reveal a constitutively activated integrin. Adhesion to immobilized fibrinogen is strongly impaired in alpha(V(del))beta(3) or alpha(V(AA))beta(3)-expressing cells, whereas it is not impaired in alpha(IIb)beta(3) and alpha(M)beta(2), both lacking the GFFKR-region. In a parallel plate flow chamber assay, alpha(V)beta(3)-expressing cells adhere firmly to fibrinogen and spread even at shear rates of 15 to 20 dyn/cm(2), whereas alpha(V(del))beta(3) or alpha(V(AA))beta(3) cells are detached at 15 dyn/cm(2). Actin stress fiber formation and focal adhesion plaques containing alpha(V)beta(3) are observed in alpha(V)beta(3) cells but not in alpha(V(del))beta(3) or alpha(V(AA))beta(3)-expressing cells. As an additional manifestation of impaired outside-in signaling, phosphorylation of pp125(FAK) was reduced in these cells. In summary, we report that the GFFKR-region of the alpha(V)-cytoplasmic domain and in particular two phenylalanines are essential for integrin alpha(V)beta(3) function, especially for outside-in signaling. Our results suggest that the two beta(3)-integrins alpha(IIb)beta(3) and alpha(V)beta(3) are differentially regulated via their GFFKR-region.  相似文献   

3.
Two integrin-type collagen receptors, alpha(1)beta(1) and alpha(2)beta(1), are structurally very similar. However, cells can concomitantly express the both receptors and they might have independent functions. Here, Chinese hamster ovary (CHO) cells, which lack endogenous collagen receptors, were transfected with either alpha(1) or alpha(2) integrin cDNA. Cells were allowed to adhere to various collagen types and their integrin function was tested by observing the progression of cell spreading. The cells expressing alpha(1)beta(1) integrin could spread on collagen types I, III, IV, and V but not on type II, while alpha(2)beta(1) integrin could mediate cell spreading on collagen types I-V. Type XIII is a transmembrane collagen and its interaction with the integrins has not been previously studied. CHO-alpha1beta1 cells could spread on human recombinant type XIII collagen, unlike CHO-alpha2beta1 cells. Integrins alpha(1)beta(1) and alpha(2)beta(1) recognize collagens with the specific alphaI domains. The alpha(1)I and alpha(2)I domains were produced as recombinant proteins, labeled with europium and used in a sensitive solid-phase binding assay based on time-resolved fluorescence. alpha(1)I domain, unlike the alpha(2)I domain, could attach to type XIII collagen. The results indicate, that alpha(1)beta(1) and alpha(2)beta(1) have different ligand binding specificity. Distinct recognition of different collagen subtypes by the alphaI domains can partially explain the differences seen in cell spreading. However, despite the fact that CHO-alpha1beta1 cells could not spread on type II collagen alpha(1)I domain could bind to this collagen type. Thus, the cell spreading on collagens may also be regulated by factors other than the integrins.  相似文献   

4.
The alpha(L) I (inserted or interactive) domain of integrin alpha(L)beta(2) undergoes conformational changes upon activation. Recent studies show that the isolated, activated alpha(L) I domain is sufficient for strong ligand binding, suggesting the beta(2) subunit to be only indirectly involved. It has been unclear whether the activity of the alpha(L) I domain is regulated by the beta(2) subunit. In this study, we demonstrate that swapping the disulfide-linked CPNKEKEC sequence (residues 169-176) in the beta(2) I domain with a corresponding beta(3) sequence, or mutating Lys(174) to Thr, constitutively activates alpha(L)beta(2) binding to ICAM-1. These mutants do not require Mn(2+) for ICAM-1 binding and are insensitive to the inhibitory effect of Ca(2+). We have also localized a component of the mAb 24 epitope (a reporter of beta(2) integrin activation) in the CPNKEKEC sequence. Glu(173) and Glu(175) of the beta(2) I domain are identified as critical for mAb 24 binding. Because the epitope is highly expressed upon beta(2) integrin activation, it is likely that the CPNKEKEC sequence is exposed or undergoes conformational changes upon activation. Deletion of the alpha(L) I domain did not eliminate the mAb 24 epitope. This confirms that the alpha(L) I domain is not critical for mAb 24 binding, and indicates that mAb 24 detects a change expressed in part in the beta(2) subunit I domain. These results suggest that the CPNKEKEC sequence of the beta(2) I domain is involved in regulating the alpha(L) I domain.  相似文献   

5.
The alpha(1)beta(1) and alpha(2)beta(1) integrins are cell surface collagen receptors. Cells expressing the alpha(1)beta(1) integrin preferentially adhere to collagen IV, whereas cells expressing the alpha(2)beta(1) integrin preferentially adhere to collagen I. Recombinant alpha(1) and alpha(2) integrin I domains exhibit the same collagen type preferences as the intact integrins. In addition, the alpha(2) integrin I domain binds echovirus 1; the alpha(1) I domain does not. To identify the structural components of the I domains responsible for the varying ligand specificities, we have engineered several alpha(1)/alpha(2) integrin I domain chimeras and evaluated their virus and collagen binding activities. Initially, large secondary structural components of the alpha(2) I domain were replaced with corresponding regions of the alpha(1) I domain. Following analysis in echovirus 1 and collagen binding assays, chimeras with successively smaller regions of alpha(1) I were constructed and analyzed. The chimeras were analyzed by ELISA with several different alpha(2) integrin monoclonal antibodies to assess their proper folding. Three different regions of the alpha(1) I domain, when present in the alpha(2) I domain, conferred enhanced collagen IV binding activity upon the alpha(2) I domain. These include the alpha3 and alpha5 helices and a portion of the alpha6 helix. Echovirus 1 binding was lost in a chimera containing the alphaC-alpha6 loop; higher resolution mapping identified Asn(289) as playing a critical role in echovirus 1 binding. Asn(289) had not been implicated in previous echovirus 1 binding studies. Taken together, these data reveal the existence of multiple determinants of ligand binding specificities within the alpha(1) and alpha(2) integrin I domains.  相似文献   

6.
7.
We investigated the effect of transforming factor factor-beta(1) (TGF-beta(1)) on thromboxane B(2) (TXB(2)) and prostaglandin E(2) (PGE(2)) production in in vitro silica dust-exposed rat alveolar macrophages (AM). In the presence of 5 mug of anti-TGF-beta(1) antibodies, TXB(2) production decreased, but PGE(2) production increased. Addition of 2 ng of TGF-beta(1) to the culture medium potentiated TXB(2) production, but PGE(2) production apparently did not change. At 50 ng of TGF-beta(1), TXB(2) production decreased, and PGE(2) production varied. Our data suggest that in rat AM: (1) both endogenous and exogenous TGF-beta(1) regulate TXB(2) production; and (2) in the absence of endogenous TGF-beta(1) the liberation of PGE(2) increases; however, exogenous TGF-beta(1) does not have a regulatory effect on PGE(2).  相似文献   

8.
L-Histidine (L-His) enhances the clastogenic effects of hydrogen peroxide (H(2)O(2)). We previously suggested the involvement of active transport in the efficient influx of an L-His--H(2)O(2) adduct into cells (Oya-Ohta et al. [1]). In this study, we detected intracellular H(2)O(2) by monitoring formation of 2',7'-dichlorofluorescein (DCF) from its precursor. More fluoroproduct accumulated dose-dependently in cells treated with a mixture of L-His and H(2)O(2) (mixture) than with H(2)O(2) alone. This observation supports our hypothesis that active transport is involved in the enhanced incorporation of H(2)O(2) into cells. Moreover, both mixture and the L-His--H(2)O(2) adduct were less active in the generation of hydroxyl radicals (*OH) upon addition of FeCl(2) than was H(2)O(2) alone in a cell-free system. This result suggests that the Fenton reaction might occur more effectively around the nucleus in cells. An immunohistochemical assay using 8-oxodG-specific monoclonal antibodies did not reveal whether the accumulation of H(2)O(2) generates 8-oxodeoxyguanosine (8-oxodG). No 8-oxodG was evident in cells treated with mixture or with H(2)O(2) alone, or even in cells treated with H(2)O(2) at high doses up to 20 mM and, in some cases, pre-treated with catalase inhibitors. It appears, therefore, that *OH and, specifically, *OH derived from intracellular Fenton reactions, might not play a role in the formation of 8-oxodG. However, exposure to UV-C of cells treated with H(2)O(2) yielded more 8-oxodG in the presence of L-His than in the absence of L-His. Thus, the previously observed enhancing effects of L-His were also noted during the induction of formation of 8-oxodG by UV-C plus H(2)O(2). The formation of 8-oxodG in response to UV-C alone was very limited and, hence, H(2)O(2) seemed to be an effective source of *OH only in the presence of UV-C. It is suggested that the *OH that induces formation of 8-oxodG is not *OH formed via intracellular Fenton reactions but is *OH formed via the dissociation of H(2)O(2) under UV-C.  相似文献   

9.
In contrast to beta(1)-adrenoreceptor (beta(1)-AR) signaling, beta(2)-AR stimulation in cardiomyocytes augments L-type Ca(2+) current in a cAMP-dependent protein kinase (PKA)-dependent manner but fails to phosphorylate phospholamban, indicating that the beta(2)-AR-induced cAMP/PKA signaling is highly localized. Here we show that inhibition of G(i) proteins with pertussis toxin (PTX) permits a full phospholamban phosphorylation and a de novo relaxant effect following beta(2)-AR stimulation, converting the localized beta(2)-AR signaling to a global signaling mode similar to that of beta(1)-AR. Thus, beta(2)-AR-mediated G(i) activation constricts the cAMP signaling to the sarcolemma. PTX treatment did not significantly affect the beta(2)-AR-stimulated PKA activation. Similar to G(i) inhibition, a protein phosphatase inhibitor, calyculin A (3 x 10(-8) M), selectively enhanced the beta(2)-AR but not beta(1)-AR-mediated contractile response. Furthermore, PTX and calyculin A treatment had a non-additive potentiating effect on the beta(2)-AR-mediated positive inotropic response. These results suggest that the interaction of the beta(2)-AR-coupled G(i) and G(s) signaling affects the local balance of protein kinase and phosphatase activities. Thus, the additional coupling of beta(2)-AR to G(i) proteins is a key factor causing the compartmentalization of beta(2)-AR-induced cAMP signaling.  相似文献   

10.
A more sensitive analytical method for NO(3) was developed based on the conversion of NO(3) to N(2)O by a denitrifier that could not reduce N(2)O further. The improved detectability resulted from the high sensitivity of the Ni electron capture gas chromatographic detector for N(2)O and the purification of the nitrogen afforded by the transformation of the N to a gaseous product with a low atmospheric background. The selected denitrifier quantitatively converted NO(3) to N(2)O within 10 min. The optimum measurement range was from 0.5 to 50 ppb (50 mug/liter) of NO(3) N, and the detection limit was 0.2 ppb of N. The values measured by the denitrifier method compared well with those measured by the high-pressure liquid chromatographic UV method above 2 ppb of N, which is the detection limit of the latter method. It should be possible to analyze all types of samples for nitrate, except those with inhibiting substances, by this method. To illustrate the use of the denitrifier method, NO(3) concentrations of <2 ppb of NO(3) N were measured in distilled and deionized purified water samples and in anaerobic lake water samples, but were not detected at the surface of the sediment. The denitrifier method was also used to measure the atom% of N in NO(3). This method avoids the incomplete reduction and contamination of the NO(3) -N by the NH(4) and N(2) pools which can occur by the conventional method of NO(3) analysis. N(2)O-producing denitrifier strains were also used to measure the apparent K(m) values for NO(3) use by these organisms. Analysis of N(2)O production by use of a progress curve yielded K(m) values of 1.7 and 1.8 muM NO(3) for the two denitrifier strains studied.  相似文献   

11.
We have observed an exchange of (18)O in the reactions of CO(2) with peroxynitrite using membrane-inlet mass spectrometry and HPLC negative electrospray ionization mass spectrometry. The exchange appeared on addition of peroxynitrite to a solution containing (18)O-labeled CO(2) in equilibrium with bicarbonate. It was observed as a temporarily enhanced rate of depletion of (18)O from CO(2), a rate that was greater than the rate of (18)O depletion caused by the hydration/dehydration cycle of CO(2). In addition, we detected the appearance of mass peaks attributed to (18)O in product NO(3)(-).As a further measure of the (18)O exchange, there was a redistribution of (18)O such that the ratio of doubly to singly labeled CO(2) could not be described by the binomial expansion. This is not due to the hydration/dehydration cycle of CO(2) but most likely to recycling of CO(2) in the reaction with peroxynitrite. This (18)O exchange associated with the reactions of CO(2) and peroxynitrite may open a new methodology for studying this significant process.  相似文献   

12.
Hydroxylation of proline residues in the Yaa position of the Gly-Xaa-Yaa repeated sequence to 4(R)-hydroxyproline is essential for the formation of the collagen triple helix. A small number of 3(S)-hydroxyproline residues are present in most collagens in the Xaa position. Neither the structural nor a biological role is known for 3(S)-hydroxyproline. To characterize the structural role of 3(S)-hydroxyproline, the peptide Ac-(Gly-3(S)Hyp-4(R)Hyp)10-NH2 was synthesized and analyzed by circular dichroism spectroscopy, analytical ultracentrifugation, and 1H nuclear magnetic resonance spectroscopy. At 4 degrees C in water the circular dichroism spectrum indicates that this peptide was in a polyproline-II-like secondary structure with a positive peak at 225 nm similar to Ac-(Gly-Pro-4(R)Hyp)10-NH2. The positive peak at 225 nm almost linearly decreases with increasing temperature to 95 degrees C without an obvious transition. Although the peptide Ac-(Gly-Pro-4(R)Hyp)10-NH2 forms a trimer at 10 degrees C, sedimentation equilibrium experiments indicate that Ac-(Gly-3(S)Hyp-4(R)Hyp)10-NH2 is a monomer in water at 7 degrees C. To study the role of 3(S)-hydroxyproline in the Yaa position, we synthesized Ac-(Gly-Pro-3(S)Hyp)10-NH2. This peptide also does not form a triple helix in water. 1H Nuclear magnetic resonance spectroscopy data (including line widths and nuclear Overhauser effects) are entirely consistent, with neither Ac-(Gly-3(S)Hyp-4(R)Hyp)10-NH2 nor Ac-(Gly-Pro-3(S)Hyp)10-NH2 forming a triple helix in water. Therefore 3(S)-hydroxyproline destabilizes the collagen triple helix in either position. In contrast, when 3(S)-hydroxyproline is inserted as a guest in the highly stable -Gly-Pro-4(R)Hyperepeated host sequence, Ac-(Gly-Pro-4(R)Hyp)3-Gly-3(S)Hyp-4(R)Hyp-(Gly-Pro-4(R)Hyp)4-Gly-Gly-NH2 forms as stable a trimer (Tm=49.6 degrees C) as Ac-(Gly-Pro-4(R)Hyp)8-Gly-Gly-NH2 (Tm=48.9 degrees C). Given that Ac-(Gly-Pro-4(R)Hyp)3-Gly-4(R)Hyp-Pro-(Gly-Pro-4(R)Hyp)4-Gly-Gly-NH2 forms a triple helix nearly as stable as the above two peptides (Tm=45.0 degrees C) and the knowledge that Ac-(Gly-4(R)Hyp-Pro)10-NH2 does not form a triple helix, we conclude that the host environment dominates the structure of host-guest peptides and that these peptides are not necessarily accurate predictors of triple helical stability.  相似文献   

13.
We are attempting to supply a new insight on interaction between Na(+)/K(+)-ATPase and H(2)O(2). We demonstrate that in vitro the Na(+)/K(+)-ATPase, a non heme-protein, is able to disproportionate H(2)O(2) catalatically into dioxygen and water, as well as C(40) catalase. By polarography, we quantify O(2) production and by Raman spectroscopy H(2)O(2) consumption. A comparative analysis of kinetics parameters relative to O(2) production shows that for Na(+)/K(+)-ATPase the affinity of the catalytic site able to transform H(2)O(2) into O(2) is twice weaker than that for C(40) catalase. It also shows that the molar activity for O(2) production is 300-fold weaker for ATPase than for catalase. Inhibitors, pH and GSH studies highlight the differences between the heme- and nonheme-proteins. Indeed, for C(40), NaN(3) is strongly inhibiting, but much less for ATPase. The pH range for the catalatic activity of ATPase is wide (6.5 to 8.5), while it is not for C(40) catalase (optimum at pH 8). The Na(+)/K(+)-ATPase catalatic activity is reduced in presence of glutathione, while it is not the case with C(40) catalase.  相似文献   

14.
The integrin alpha(4)beta(7) is the cell adhesion receptor for the mucosal vascular addressin MAdCAM-1, and this interaction is dominant in lymphocyte homing to Peyer's patch high endothelial venules, and plays key roles in lymphocyte recruitment at sites of inflammation. To identify alpha(4) subunit amino acids important for alpha(4)beta(7)/MAdCAM-1 interaction, we expressed mutant alpha(4) and wild type beta(7) chains in K562 cells and analyzed the effect of the mutations on cell adhesion to a soluble MAdCAM-1 (sMAdCAM-1-Ig). Transfectants expressing mutated alpha(4) at Tyr(187) displayed a substantial decrease in adhesion to this ligand, which was associated with a reduced alpha(4)beta(7)/sMAdCAM-1-Ig interaction, as determined by soluble binding assays. Addition of Mn(2+) to the adhesion assays did not restore the impaired adhesion. Mutations at alpha(4) Gln(152)Asp(153) also affected transfectant adhesion to sMAdCAM-1-Ig, but did not involve an alteration of alpha(4)beta(7)/MAdCAM-1 binding, and adhesion was restored by Mn(2+). Instead, mutations at alpha(4) Asn(123)Glu(124) did not affect this adhesion. Mutation of alpha(4) Tyr(187) abolished alpha(4)beta(7)-mediated cell adhesion to CS-1/fibronectin, an additional ligand for alpha(4)beta(7), while alpha(4) Gln(152)Asp(153) transfectant mutants showed a reduced adhesion. These results identify alpha(4) Tyr(187) as a key residue during receptor alpha(4)beta(7)/ligand interactions, indicating that it plays important roles in alpha(4)beta(7)-mediated leukocyte adhesion, and provide a potential target for therapeutic intervention in several inflammatory pathologies.  相似文献   

15.
Oxidant stress plays a significant role in the pathogenesis of periventricular leukomalacia (PVL). Isoprostanes (IsoPs) are bioactive products of lipid peroxidation abundantly generated during hypoxic-ischemic injuries. Because loss of oligodendrocytes (OLs) occurs early in PVL, we hypothesized that IsoPs could induce progenitor OL death. 15-E(2t)-IsoP but not 15-F(2t)-IsoP elicited a concentration-dependent death of progenitor OLs by oncosis and not by apoptosis, but exerted minimal effects on mature OLs. 15-E(2t)-IsoP-induced cytotoxicity could not be explained by its conversion into cyclopentenones, because PGA(2) was hardly cytotoxic. On the other hand, thromboxane A(2) (TxA(2)) synthase inhibitor CGS12970 and cyclooxygenase inhibitor ibuprofen attenuated 15-E(2t)-IsoP-induced cytotoxicity. Susceptibility of progenitor OLs was independent of TxA(2) receptor (TP) expression, which was far less in progenitor than in mature OLs. However, TxA(2) synthase was detected in precursor but not in mature OLs, and TxA(2) mimetic U46619 induced hydroperoxides generation and progenitor OL death. The glutathione synthesis enhancer N-acetylcysteine prevented 15-E(2t)-IsoP-induced progenitor cell death. Depletion of glutathione in mature OLs with buthionine sulfoximine rendered them susceptible to cytotoxicity of 15-E(2t)-IsoP. These novel data implicate 15-E(2t)-IsoP as a product of oxidative stress that may contribute in the genesis of PVL.  相似文献   

16.
The P2Y(2) nucleotide receptor (P2Y(2)R) contains the integrin-binding domain arginine-glycine-aspartic acid (RGD) in its first extracellular loop, raising the possibility that this G protein-coupled receptor interacts directly with an integrin. Binding of a peptide corresponding to the first extracellular loop of the P2Y(2)R to K562 erythroleukemia cells was inhibited by antibodies against alpha(V)beta(3)/beta(5) integrins and the integrin-associated thrombospondin receptor, CD47. Immunofluorescence of cells transfected with epitope-tagged P2Y(2)Rs indicated that alpha(V) integrins colocalized 10-fold better with the wild-type P2Y(2)R than with a mutant P2Y(2)R in which the RGD sequence was replaced with RGE. Compared with the wild-type P2Y(2)R, the RGE mutant required 1,000-fold higher agonist concentrations to phosphorylate focal adhesion kinase, activate extracellular signal-regulated kinases, and initiate the PLC-dependent mobilization of intracellular Ca(2+). Furthermore, an anti-alpha(V) integrin antibody partially inhibited these signaling events mediated by the wild-type P2Y(2)R. Pertussis toxin, an inhibitor of G(i/o) proteins, partially inhibited Ca(2+) mobilization mediated by the wild-type P2Y(2)R, but not by the RGE mutant, suggesting that the RGD sequence is required for P2Y(2)R-mediated activation of G(o), but not G(q). Since CD47 has been shown to associate directly with G(i/o) family proteins, these results suggest that interactions between P2Y(2)Rs, integrins, and CD47 may be important for coupling the P2Y(2)R to G(o).  相似文献   

17.
In vivo labeling experiments of Forsythia intermedia plant tissue with [8-14C]- and [9,9-2H2,OC2H3]coniferyl alcohols revealed that the lignans, (-)-secoisolariciresinol and (-)-matairesinol, were derived from two coniferyl alcohol molecules; no evidence for the formation of the corresponding (+)-enantiomers was found. Administration of (+-)-[Ar-3H]secoisolariciresinols to excised shoots of F. intermedia resulted in a significant conversion into (-)-matairesinol; again, the (+)-antipode was not detected. Experiments using cell-free extracts of F. intermedia confirmed and extended these findings. In the presence of NAD(P)H and H2O2, the cell-free extracts catalyzed the formation of (-)-secoisolariciresinol, with either [8-14C]- or [9,9-2H2,OC2H3]coniferyl alcohols as substrates. The (+)-enantiomer was not formed. Finally, when either (-)-[Ar-3H] or (+-)-[Ar-2H]secoisolariciresinols were used as substrates, in the presence of NAD(P), only (-)- and not (+)-matairesinol formation occurred. The other antipode, (+)-secoisolariciresinol, did not serve as a substrate for the formation of either (+)- or (-)-matairesinol. Thus, in F. intermedia, the formation of the lignan, (-)-secoisolariciresinol, occurs under strict stereochemical control, in a reaction or reactions requiring NAD(P)H and H2O2 as cofactors. This stereoselectivity is retained in the subsequent conversion into (-)-matairesinol, since (+)-secoisolariciresinol is not a substrate. These are the first two enzymes to be discovered in lignan formation.  相似文献   

18.
The covalent binding of cis-Pt(NH3)2Cl2 on the double stranded poly(I) . poly(C) induced an irreversible dissociation of the two strands. This dissociation was evidenced mainly by poly(I)-Agarose affinity chromatography which allowed to recover free strands of cis-Pt(NH3)2Cl2-poly(I) from a cis-Pt(NH3)2Cl2-poly(I) . poly(C) complex, by density equilibrium centrifugation where free poly(C) could be isolated, and by acid titrations of the metal-poly(I) . poly(C) complexes. The separation of the two strands of the polyribonucleotide upon cis-Pt(NH3)2Cl2 fixation was shown not to exceed 90--95%. A dissociation curve of the polynucleotide double helix as a function of the amount of bound cis-Pt(NH3)2Cl2 was determined and was shown to be of a characteristic cooperative effect. The fixation of the paltinum compound to poly(I) . poly(C) seemed also to be cooperative.  相似文献   

19.
A 1H-NMR investigation was carried out on the tetranucleotides U-m6(2)A-U-m6(2)A and m6(2)A-m6(2)A-U-m6(2)A (m6(2) = N6-dimethyladenosine) as well as on the hybrid trinucleotide dA-r(U-A). An extensive comparison with m6(2)A-U-m6(2)A and other relevant compounds is made. Previous proton NMR studies on trinucleotides have shown that purine-pyrimidine-purine sequences prefer to adopt a mixture of states which have as a common feature that the interior pyrimidine residue bulges out, whereas the flanking purine residues stack upon each other. A stacking interaction on the 3' side of the bulge is known to have no measurable effect on the bulge population. Chemical-shift data, ribose ring conformational analysis and information from NOE experiments now show unambiguously that the moderate U(1)-m6(2)A(2) stack in U-m6(2)A-U-m6(2)A diminishes the population of bulged-out structures in favour of a regular stack. This tendency towards conformational transmission in the downstream 5'----3' direction is fully confirmed by the fact that the strong m6(2)A(1)-m6(2)A(2) stack in the tetranucleotide m6(2)A-m6(2)A-U-m6(2)A virtually precludes the formation of bulged-out structures. The conformational characteristics of dA-r(U-A) appear comparable with those of m6(2)A-U-m6(2)A, which indicates that the presence of a 2'-hydroxyl group in the first purine residue is not a necessary prerequisite for the formation of a bulge.  相似文献   

20.
Platelets represent a target of reactive oxygen species produced under oxidative stress conditions. Controversial data on the effect of these species on platelet functions have been reported so far. In this study we evaluated the effect of a wide range of H(2)O(2) concentrations on platelet adhesion to immobilized fibrinogen and on pp72(syk) and pp125(FAK) tyrosine phosphorylation. Our results demonstrate that: (1) H(2)O(2) does not affect the adhesion of unstimulated or apyrase-treated platelets to immobilized fibrinogen; (2) H(2)O(2) does not affect pp72(syk) phosphorylation induced by platelet adhesion to fibrinogen-coated dishes; (3) H(2)O(2) reduces, in a dose-dependent fashion, pp125(FAK) phosphorylation of fibrinogen-adherent platelets; (4) concentrations of H(2)O(2) near to physiological values (10-12 microM) are able to strengthen the subthreshold activation of pp125(FAK) induced by epinephrine in apyrase-treated platelets; (5) H(2)O(2) doses higher than 0.1 mM inhibit ADP-induced platelet aggregation and dense granule secretion. The ability of H(2)O(2) to modulate pp125(FAK) phosphorylation suggests a role of this molecule in physiological hemostasis as well as in thrombus generation.  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