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.
Wu J  Zhu BB  Yu J  Zhu H  Qiu L  Kindy MS  Gu L  Seidel A  Li GM 《Nucleic acids research》2003,31(22):6428-6434
Benzo[c]phenanthrene dihydrodiol epoxide (B[c] PhDE) is well known as an important environmental chemical carcinogen that preferentially modifies DNA in adenine residues. However, the molecular mechanism by which B[c]PhDE induces tumorigenesis is not fully understood. In this report, we demonstrate that DNA mismatch repair (MMR), a genome maintenance system, plays an important role in B[c]PhDE-induced carcinogensis by promoting apoptosis in cells treated with B[c]PhDE. We show that purified human MMR recognition proteins, MutSα and MutSβ, specifically recognized B[c]PhDE-DNA adducts. Cell lines proficient in MMR exhibited several-fold more sensitivity to killing than cell lines defective in either MutSα or MutLα by B[c]PhDE; the nature of this sensitivity was shown to be due to increased apoptosis. Additionally, wild-type mice exposed to B[c]PhDE had intestinal crypt cells that underwent apoptosis significantly more often than intestinal crypt cells found in B[c]PhDE-treated Msh2–/– or Mlh1–/– mice. These findings, combined with previous studies, suggest that the MMR system may serve as a general sensor for chemical-caused DNA damage to prevent damaged cells from mutagenesis and carcinogenesis by promoting apoptosis.  相似文献   

2.
Eckert KA  Yan G 《Nucleic acids research》2000,28(14):2831-2838
Mutagenesis at [GT/CA]10, [TC/AG]11 and [TTCC/AAGG]9 microsatellite sequences inserted in the herpes simplex virus thymidine kinase (HSV-tk) gene was analyzed in isogenic mutL+ and mutL Escherichia coli. In both strains, significantly more expansion than deletion mutations were observed at the [TTCC/AAGG]9 motif relative to either dinucleo­tide motif. As the HSV-tk coding sequence contains an endogenous [G/C]7 mononucleotide repeat and ~1000 bp of unique sequence, we were able to compare mutagenesis among various sequence motifs. We observed that the relative risk of mutation in E.coli is: [TTCC/AAGG]9 > [GT/CA]10 ~ [TC/AG]11 > unique ~ [G/C]7. The mutation frequency varied 1400-fold in mutL+ cells between the tetranucleotide motif and the mononucleotide motif, but only 50-fold in mutL cells. The [G/C]7 sequence was destabilized the greatest and the tetranucleotide motif the least by loss of mismatch repair. These results demonstrate that the quantitative risk of mutation at various microsatellites greatly depends on the DNA sequence composition. We suggest alternative models for the production of expansion mutations during lagging strand replication of the [TTCC/AAGG]9 microsatellite.  相似文献   

3.
The interaction of hexamminecobalt(III), Co(NH3)63+, with 160 and 3000–8000 bp length calf thymus DNA has been investigated by circular dichroism, acoustic and densimetric techniques. The acoustic titration curves of 160 bp DNA revealed three stages of interaction: (i) Co(NH3)63+ binding up to the molar ratio [Co(NH3)63+]/[P] = 0.25, prior to DNA condensation; (ii) a condensation process between [Co(NH3)63+]/[P] = 0.25 and 0.30; and (iii) precipitation after [Co(NH3)63+]/[P] = 0.3. In the case of 3000–8000 bp DNA only two processes were observed: (i) binding up to [Co(NH3)63+]/[P] = 0.3; and (ii) precipitation after this point. In agreement with earlier observations, long DNA aggregates without changes in its B-form circular dichroism spectrum, while short DNA demonstrates a positive B→Ψ transition after [Co(NH3)63+]/[P] = 0.25. From ultrasonic and densimetric measurements the effects of Co(NH3)63+ binding on volume and compressibility have been obtained. The binding of Co(NH3)63+ to both short and long DNA is characterized by similar changes in volume and compressibility calculated per mole Co(NH3)63+: ΔV = 9 cm3 mol–1 and Δκ = 33 × 10–4 cm3 mol–1 bar–1. The positive sign of the parameters indicates dehydration, i.e. water release from Co(NH3)63+ and the atomic groups of DNA. This extent of water displacement would be consistent with the formation of two direct, hydrogen bonded contacts between the cation and the phosphates of DNA.  相似文献   

4.
Anoxia induces a rapid elevation of the cytosolic Ca2+ concentration ([Ca2+]cyt) in maize (Zea mays L.) cells, which is caused by the release of the ion from intracellular stores. This anoxic Ca2+ release is important for gene activation and survival in O2-deprived maize seedlings and cells. In this study we examined the contribution of mitochondrial Ca2+ to the anoxic [Ca2+]cyt elevation in maize cells. Imaging of intramitochondrial Ca2+ levels showed that a majority of mitochondria released their Ca2+ in response to anoxia and took up Ca2+ upon reoxygenation. We also investigated whether the mitochondrial Ca2+ release contributed to the increase in [Ca2+]cyt under anoxia. Analysis of the spatial association between anoxic [Ca2+]cyt changes and the distribution of mitochondrial and other intracellular Ca2+ stores revealed that the largest [Ca2+]cyt increases occurred close to mitochondria and away from the tonoplast. In addition, carbonylcyanide p-trifluoromethoxyphenyl hydrazone treatment depolarized mitochondria and caused a mild elevation of [Ca2+]cyt under aerobic conditions but prevented a [Ca2+]cyt increase in response to a subsequent anoxic pulse. These results suggest that mitochondria play an important role in the anoxic elevation of [Ca2+]cyt and participate in the signaling of O2 deprivation.  相似文献   

5.
Evolution depends on the manner in which genetic variation is translated into new phenotypes. There has been much debate about whether organisms might have specific mechanisms for “evolvability,” which would generate heritable phenotypic variation with adaptive value and could act to enhance the rate of evolution. Capacitor systems, which allow the accumulation of cryptic genetic variation and release it under stressful conditions, might provide such a mechanism. In yeast, the prion [PSI+] exposes a large array of previously hidden genetic variation, and the phenotypes it thereby produces are advantageous roughly 25% of the time. The notion that [PSI+] is a mechanism for evolvability would be strengthened if the frequency of its appearance increased with stress. That is, a system that mediates even the haphazard appearance of new phenotypes, which have a reasonable chance of adaptive value would be beneficial if it were deployed at times when the organism is not well adapted to its environment. In an unbiased, high-throughput, genome-wide screen for factors that modify the frequency of [PSI+] induction, signal transducers and stress response genes were particularly prominent. Furthermore, prion induction increased by as much as 60-fold when cells were exposed to various stressful conditions, such as oxidative stress (H2O2) or high salt concentrations. The severity of stress and the frequency of [PSI+] induction were highly correlated. These findings support the hypothesis that [PSI+] is a mechanism to increase survival in fluctuating environments and might function as a capacitor to promote evolvability.  相似文献   

6.
The divalent cation Sr2+ induced repetitive transient spikes of the cytosolic Ca2+ activity [Ca2+]cy and parallel repetitive transient hyperpolarizations of the plasma membrane in the unicellular green alga Eremosphaera viridis. [Ca2+]cy measurements, membrane potential measurements, and cation analysis of the cells were used to elucidate the mechanism of Sr2+-induced [Ca2+]cy oscillations. Sr2+ was effectively and rapidly compartmentalized within the cell, probably into the vacuole. The [Ca2+]cy oscillations cause membrane potential oscillations, and not the reverse. The endoplasmic reticulum (ER) Ca2+-ATPase blockers 2,5-di-tert-butylhydroquinone and cyclopiazonic acid inhibited Sr2+-induced repetitive [Ca2+]cy spikes, whereas the compartmentalization of Sr2+ was not influenced. A repetitive Ca2+ release and Ca2+ re-uptake by the ER probably generated repetitive [Ca2+]cy spikes in E. viridis in the presence of Sr2+. The inhibitory effect of ruthenium red and ryanodine indicated that the Sr2+-induced Ca2+ release from the ER was mediated by a ryanodine/cyclic ADP-ribose type of Ca2+ channel. The blockage of Sr2+-induced repetitive [Ca2+]cy spikes by La3+ or Gd3+ indicated the necessity of a certain influx of divalent cations for sustained [Ca2+]cy oscillations. Based on these data we present a mathematical model that describes the baseline spiking [Ca2+]cy oscillations in E. viridis.  相似文献   

7.
A key issue in the nucleotide excision repair (NER) of bulky carcinogen–DNA adducts is the ability of the NER machinery to recognize and repair certain adducts while failing to repair others. Unrepaired adducts can survive to cause mutations that initiate the carcinogenic process. Benzo[c]phenanthrene (B[c]Ph), a representative fjord region polycyclic aromatic hydrocarbon, can be metabolically activated to the enantiomeric benzo[c]phenanthrene diol epoxides (B[c]PhDEs), (+)-(1S,2R,3R,4S)-3,4- dihydroxy-1,2-epoxy-1,2,3,4-tetrahydrobenzo[c]phe nanthrene and the corresponding (–)-(1R,2S,3S,4R) isomer. These react predominantly with adenine residues in DNA to produce the stereoisomeric 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts. Duplexes containing the 1R (+) or 1S (–) B[c]Ph-dA adduct in codon 61 of the human N-ras mutational hotspot sequence CA*A, with B[c]Ph modification at A*, are not repaired by the human NER system. However, the analogous stereoisomeric DNA adducts of the bay region benzo[a]pyrene diol epoxide (B[a]PDE), 10S (+)- and 10R (–)-trans-anti-B[a]P-N6-dA, are repaired in the same base sequence. In order to elucidate structural and thermodynamic origins of this phenomenon, we have carried out a 2 ns molecular dynamics simulation for the 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts in an 11mer duplex containing the human N-ras codon 61 sequence, and compared these results with our previous study of the B[a]P-dA adducts in the same sequence. The molecular mechanics Poisson– Boltzmann surface area (MM-PBSA) method was applied to calculate the free energies of the pair of stereoisomeric B[c]Ph-dA adducts, and a detailed structural analysis was carried out. The different repair susceptibilities of the B[a]P-dA adducts and the B[c]Ph-dA adducts can be attributed to different degrees of distortion, stemming from combined effects of differences in the quality of Watson–Crick hydrogen bonding, unwinding, stretching and helix backbone perturbations. These differences are due to the different intrinsic topologies of the rigid, planar bay region adducts versus the twisted, sterically hindered fjord region adducts.  相似文献   

8.
In previous papers of this series the temperature-dependent Raman spectra of poly(dA)·poly(dT) and poly(dA–dT)·poly(dA–dT) were used to characterize structurally the melting and premelting transitions in DNAs containing consecutive A·T and alternating A·T/T·A base pairs. Here, we describe procedures for obtaining thermodynamic parameters from the Raman data. The method exploits base-specific and backbone-specific Raman markers to determine separate thermodynamic contributions of A, T and deoxyribosyl-phosphate moieties to premelting and melting transitions. Key findings include the following: (i) Both poly(dA)·poly(dT) and poly(dA–dT)· poly(dA–dT) exhibit robust premelting transitions, due predominantly to backbone conformational changes. (ii) The significant van’t Hoff premelting enthalpies of poly(dA)·poly(dT) [ΔHvHpm = 18.0 ± 1.6 kcal·mol–1 (kilocalories per mole cooperative unit)] and poly(dA–dT)·poly(dA–dT) (ΔHvHpm = 13.4 ± 2.5 kcal·mol–1) differ by an amount (~4.6 kcal·mol–1) estimated as the contribution from three-centered inter-base hydrogen bonding in (dA)n·(dT)n tracts. (iii) The overall stacking free energy of poly(dA)· poly(dT) [–6.88 kcal·molbp–1 (kilocalories per mole base pair)] is greater than that of poly(dA–dT)· poly(dA–dT) (–6.31 kcal·molbp–1). (iv) The difference between stacking free energies of A and T is significant in poly(dA)·poly(dT) (ΔΔGst = 0.8 ± 0.3 kcal· molbp–1), but marginal in poly(dA–dT)·poly(dA–dT) (ΔΔGst = 0.3 ± 0.3 kcal·molbp–1). (v) In poly(dA)· poly(dT), the van’t Hoff parameters for melting of A (ΔHvHA = 407 ± 23 kcal·mol–1, ΔSvHA = 1166 ± 67 cal·°K–1·mol–1, ΔGvH(25°C)A = 60.0 ± 3.2 kcal·mol–1) are clearly distinguished from those of T (ΔHvHT = 185 ± 38 kcal·mol–1, ΔSvHT = 516 ± 109 cal·°K–1·mol–1, ΔGvH(25°C)T = 27.1 ± 5.5 kcal·mol–1). (vi) Similar relative differences are observed in poly(dA–dT)· poly(dA–dT) (ΔHvHA = 333 ± 54 kcal·mol–1, ΔSvHA = 961 ± 157 cal·°K–1·mol–1, ΔGvH(25°C)A = 45.0 ± 7.6 kcal· mol–1; ΔHvHT = 213 ± 30 kcal·mol–1, ΔSvHT = 617 ± 86 cal·°K–1·mol–1, ΔGvH(25°C)T = 29.3 ± 4.9 kcal·mol–1). The methodology employed here distinguishes thermodynamic contributions of base stacking, base pairing and backbone conformational ordering in the molecular mechanism of double-helical B DNA formation.  相似文献   

9.
We report for the first time the in vitro characterization of a reverse tetracycline repressor (revTetR). The dimeric wild-type repressor (TetR) binds to tet operator tetO in the absence of the inducer anhydrotetracycline (atc) to confer tight repression. We have isolated the revTetR G96E L205S mutant, which, contrary to TetR, binds tetO only in the presence of atc. This reverse acting mutant was overproduced and purified. Effector and DNA binding properties were analyzed by EMSA and quantified by fluorescence titration and surface plasmon resonance. The association constant KA of revTetR for binding of [atcMg]+ is ~108 M–1, four orders of magnitude lower than that of TetR. The affinity of TetR for tetO is 5.6 ± 2 × 109 M–1 and that for revTetR in the presence of atc is 1 ± 0.2 × 108 M–1. Both induced forms, the atc-bound TetR and the free revTetR, have the same low affinity of 4 ± 1 × 105 M–1 for DNA. Therefore, atc does not act as a dimerization agent for revTetR. We discuss the structural differences between TetR and revTetR potentially underlying this reversal of activity.  相似文献   

10.
The [PSI+] prion causes widespread readthrough translation and is rare in natural populations of Saccharomyces, despite the fact that sex is expected to cause it to spread. Using the recently estimated rate of Saccharomyces outcrossing, we calculate the strength of selection necessary to maintain [PSI+] at levels low enough to be compatible with data. Using the best available parameter estimates, we find selection against [PSI+] to be significant. Inference regarding selection on modifiers of [PSI+] appearance depends on obtaining more precise and accurate estimates of the product of yeast effective population size Ne and the spontaneous rate of [PSI+] appearance m. The ability to form [PSI+] has persisted in yeast over a long period of evolutionary time, despite a diversity of modifiers that could abolish it. If mNe < 1, this may be explained by insufficiently strong selection. If mNe > 1, then selection should favor the spread of [PSI+] resistance modifiers. In this case, rare conditions where [PSI+] is adaptive may permit its persistence in the face of negative selection.  相似文献   

11.
Conjugative transfer of the Enterococcus faecalis plasmid pPD1 is activated by cPD1, one of several peptide sex pheromones secreted by plasmid-free recipient cells, and is blocked by a donor-produced peptide inhibitor, iPD1. Using a tritiated pheromone, [3H]cPD1, we investigated how pPD1-harboring donor cells receive these peptide signals. Donor cells rapidly incorporated [3H]cPD1. The cell extract but not the membrane fraction of the donor strain exhibited significant [3H]cPD1-binding activity. On the basis of these data and those of tracer studies, it was demonstrated that cPD1 was internalized, where it bound to a high-molecular-weight compound. The cell extract of a strain carrying the traA-bearing multicopy plasmid (pDLHH21) also exhibited high [3H]cPD1-binding activity. A recombinant TraA exhibited a dissociation constant of 0.49 ± 0.08 nM against [3H]cPD1. iPD1 competitively inhibited [3H]cPD1 binding to TraA, whereas pheromones and inhibitors relating to other plasmid systems did not. These results show that TraA is a specific intracellular receptor for cPD1 and that iPD1 acts as an antagonist for TraA. A strain carrying the traC-bearing multicopy plasmid (pDLES23) exhibited significant [3H]cPD1-binding activity. A strain carrying traC-disrupted pPD1 (pAM351CM) exhibited lower [3H]cPD1-binding activity as well as lower sensitivity to cPD1 than a wild-type donor strain. Some of the other pheromones and inhibitors inhibited [3H]cPD1 binding to the traC transformant like cPD1 and iPD1 did. These results show that TraC, as an extracellular less-specific pheromone-binding protein, supports donor cells to receive cPD1.  相似文献   

12.
The uptake and degradation of nanomolar levels of [methyl-14C]choline in estuarine water samples and in seawater filtrate cultures composed mainly of natural free-living bacteria was studied. Uptake of [14C]choline exhibited Michaelis-Menten kinetics, with Kt + Sn values of 1.7 to 2.9 nM in filtrate cultures and 1.7 to 4.1 nM in estuarine-water samples. Vmax values ranged from 0.5 to 3.3 nM · h−1. The uptake system for choline in natural microbial assemblages therefore displays very high affinity and appears able to scavenge this compound at the concentrations expected in seawater. Uptake of choline was inhibited by some natural structural analogs and p-chloromercuribenzoate, indicating that the transporter may be multifunctional and may involve a thiol binding site. When 11 nM [14C]choline was added to water samples, a significant fraction (>50%) of the methyl carbon was respired to CO2 in incubations lasting 10 to 53 h. Cells taking up [14C]choline produced [14C]glycine betaine ([14C]GBT), and up to 80% of the radioactivity retained by cells was in the form of GBT, a well-known osmolyte. Alteration of the salinity in filtrate cultures affected the relative proportion of [14C]choline degraded or converted to [14C]GBT, without substantially affecting the total metabolism of choline. Increasing the salinity from 14 to 25 or 35 ppt caused more [14C]GBT to be produced from choline but less 14CO2 to be produced than in the controls. Lowering the salinity to 7 ppt decreased [14C]GBT production and increased 14CO2 production slightly. Intracellular accumulations of [14C]GBT in the salt-stressed cultures were osmotically significant (34 mM). Choline may be used as an energy substrate by estuarine bacteria and may also serve as a precursor of the osmoprotectant GBT, particularly as bacteria are mixed into higher-salinity waters.  相似文献   

13.
Four genomic DNAs of differing GC content (Micrococcus luteus, 72% GC; Escherichia coli, 50% GC; calf thymus, 42% GC; Clostridium perfringens, 27% GC) have been employed as targets of interaction by the cationic polyamines spermidine {[H3N(CH2)3NH2(CH2)4NH3]3+} and spermine {[(CH2)4(NH2(CH2)3NH3)2]4+}. In solutions containing 60 mM DNA phosphate (~20 mg DNA/ml) and either 1, 5 or 60 mM polyamine, only Raman bands associated with the phosphates exhibit large spectral changes, demonstrating that B-DNA phosphates are the primary targets of interaction. Phosphate perturbations, which are independent of base composition, are consistent with a model of non-specific cation binding in which delocalized polyamines diffuse along DNA while confined by the strong electrostatic potential gradient perpendicular to the helix axis. This finding provides experimental support for models in which polyamine-induced DNA condensation is driven by non-specific electrostatic binding. The Raman spectra also demonstrate that major groove sites (guanine N7 and thymine C5H3) are less affected than phosphates by polyamine–DNA interactions. Modest dependence of polyamine binding on genome base composition suggests that sequence context plays only a secondary role in recognition. Importantly, the results demonstrate that polyamine binding has a negligible effect on the native B-form secondary structure. The capability of spermidine or spermine to bind and condense genomic B-DNA without disrupting the native structure must be taken into account when considering DNA organization within bacterial nucleoids or cell nuclei.  相似文献   

14.
High precision ultrasonic and densimetric techniques have been used to study the interaction of Ni2+ ions with right-handed poly[d(G-C)]·poly[d(G-C)], poly-[d(A-C)]·poly[d(G-T)] and poly[d(A-T)]·poly[d(A-T)] in 5 mM CsCl, 0.2 mM HEPES, pH 7.5 at 20°C. From these measurements the changes in the apparent molar volume and the apparent molar adiabatic compressibility due to the interaction have been obtained. The volume effects of the binding, calculated per mole of Ni2+ ions, range from 11.7 to 23.9 cm3 mol–1 and the compressibility effects range from 19.3 × 10–4 to 43.1 × 10–4 cm3 mol–1 bar–1. These data are interpreted in terms of dehydration of the polynucleotides and Ni2+ ions, i.e. the release of water molecules from the hydration shells of the molecules. An increase in G+C content gives an increase in volume and compressibility effects, indicating a rise in the extent of dehydration. The dehydration effects of Ni2+ binding to poly[d(G-C)]·poly[d(G-C)] are approximately twice those of poly[d(A-T)]·poly[d(A-T)]. The volume and compressibility effects of Ni2+–EDTA complex formation have also been measured and used as a model system for quantitative estimation. These values revealed that Ni2+ ions can coordinate two atomic groups of poly[d(G-C)]·poly[d(G-C)], while in the case of the Ni2+–poly[d(A-T)]·poly[d(A-T)] complex volume and compressibility effects correspond to one direct or two indirect (through water) contacts.  相似文献   

15.
We studied phosphopeptidomannans (PPMs) of two Saccharomyces cerevisiae NCYC 625 strains (S. diastaticus): a wild type strain grown aerobically, anaerobically, and in the presence of antimycin and a [rho0] mutant grown aerobically and anaerobically. The aerobic wild-type cultures were highly flocculent, but all others were weakly flocculent. Ligands implicated in flocculation of mutants or antimycin-treated cells were not aggregated as much by concanavalin A as were those of the wild type. The [rho0] mutants and antimycin-treated cells differ from the wild type in PPM composition and invertase, acid phosphatase, and glucoamylase activities. PPMs extracted from different cells differ in the protein but not in the glycosidic moiety. The PPMs were less stable in mitochondrion-deficient cells than in wild-type cells grown aerobically, and this difference may be attributable to defective mitochondrial function during cell wall synthesis. The reduced flocculation of cells grown in the presence of antimycin, under anaerobiosis, or carrying a [rho0] mutation may be the consequence of alterations of PPM structures which are the ligands of lectins, both involved in this cell-cell recognition phenomenon. These respiratory chain alterations also affect peripheral, biologically active glycoproteins such as extracellular enzymes and peripheral PPMs.  相似文献   

16.
Penetration of 3H-labeled water (3H2O) and the 14C-labeled organic acids benzoic acid ([14C]BA), salicylic acid ([14C]SA), and 2,4-dichlorophenoxyacetic acid ([14C]2,4-D) were measured simultaneously in isolated cuticular membranes of Prunus laurocerasus L., Ginkgo biloba L., and Juglans regia L. For each of the three pairs of compounds (3H2O/[14C]BA, 3H2O/[14C]SA, and 3H2O/[14C]2,4-D) rates of cuticular water penetration were highly correlated with the rates of penetration of the organic acids. Therefore, water and organic acids penetrated the cuticles by the same routes. With the combination 3H2O/[14C]BA, co-permeability was measured with isolated cuticles of nine other plant species. Permeances of 3H2O of all 12 investigated species were highly correlated with the permeances of [14C]BA (r2 = 0.95). Thus, cuticular transpiration can be predicted from BA permeance. The application of this experimental method, together with the established prediction equation, offers the opportunity to answer several important questions about cuticular transport physiology in future investigations.  相似文献   

17.
We recently showed that maize (Zea mays L.) leaves contain appreciable amounts of phosphoenolpyruvate carboxykinase (PEPCK; R.P. Walker, R.M. Acheson, L.I. Técsi, R.C. Leegood [1997] Aust J Plant Physiol 24: 459–468). In the present study, we investigated the role of PEPCK in C4 photosynthesis in maize. PEPCK activity and protein were enriched in extracts from bundle-sheath (BS) strands compared with whole-leaf extracts. Decarboxylation of [4-14C]aspartate (Asp) by BS strands was dependent on the presence of 2-oxoglutarate and Mn2+, was stimulated by ATP, was inhibited by the PEPCK-specific inhibitor 3-mercaptopicolinic acid, and was independent of illumination. The principal product of Asp metabolism was phosphoenolpyruvate, whereas pyruvate was a minor product. Decarboxylation of [4-14C]malate was stimulated severalfold by Asp and 3-phosphoglycerate, was only slightly reduced in the absence of Mn2+ or in the presence of 3-mercaptopicolinic acid, and was light dependent. Our data show that decarboxylation of Asp and malate in BS cells of maize occurs via two different pathways: Whereas malate is mainly decarboxylated by NADP-malic enzyme, decarboxylation of Asp is dependent on the activity of PEPCK.  相似文献   

18.
Beta‐2‐glycoprotein I (β2GPI) is the major antigen for the antiphospholipid antibodies in the antiphospholipid syndrome. The exposed epitope in domain I of β2GPI can be recognized by the anti‐β2GPI antibody. Here, we prepared the anionic di‐oleoyl‐phosphatidylserine (DOPS) and cardiolipin (CL) liposomes to interact with the β2GPI. The conformational changes of β2GPI upon binding with the liposomes were analyzed using hydrogen/deuterium exchange mass spectrometry. The exchange level of sequences 21–27 significantly increased after β2GPI had interacted with DOPS. This change indicated a reduced interaction between domain I and domain V, inferring to a protrusion of the sequences 21–27 from the ring conformation. After β2GPI had interacted with CL for 30 min, the exchange levels in 4 of the 5 domains increased significantly. The deuteration levels of sequences 1–20, 21–27, 196–205, 273–279 and 297–306 increased, suggesting that these regions had become more exposed, and the domain I was no longer in contact with domain V. The increasing deuteration levels in sequences 70–86, 153–162, 191–198, 196–205 and 273–279 indicated β2GPI undergoing conformational changes to expose these inner regions, suggesting a structural transition. Overall, DOPS and CL induced minor conformational changes of β2GPI at sequences 21–27 and forms an intermediate conformation after 10 min of interaction. After a complete protein–lipid interaction, high negatively charged CL membrane induced a major conformation transition of β2GPI.  相似文献   

19.
Addition of membrane-permeable cyclic GMP (cGMP) and cyclic AMP (cAMP) were shown to cause elevation of cytosolic Ca2+ concentration ([Ca2+]cyt) in tobacco (Nicotiana plumbaginofolia) protoplasts. Under the same conditions these cyclic nucleotides were shown to provoke a physiological swelling response in the protoplasts. Nonmembrane-permeable cAMP and cGMP were unable to trigger a detectable [Ca2+]cyt response. Cyclic-nucleotide-mediated elevations in [Ca2+]cyt involved both internal and external Ca2+ stores. Both cAMP- and cGMP-mediated [Ca2+]cyt elevations could be inhibited by the Ca2+-channel blocker verapamil. Addition of inhibitors of phosphodiesterases (isobutylmethylxanthine and zaprinast) and the adenylate cyclase agonist forskolin to the protoplasts (predicted to elevate in vivo cyclic-nucleotide concentrations) caused elevations in [Ca2+]cyt. Addition of the adenylate cyclase inhibitor 2′,5′-dideoxyadenosine before forskolin significantly inhibited the forskolin-induced [Ca2+]cyt elevation. Taken together, these data suggest that a potential communication point for cross-talk between signal transduction pathways using cyclic nucleotides in plants is at the level of Ca2+ signaling.  相似文献   

20.
Recent genomic analyses revealed many kinds of tandem repeats of specific sequences. Some of them are related to genetic diseases, but their biological functions and structures are still unknown. Two X-ray structures of a short DNA fragment d(gcGA[G]1Agc) show that four base-intercalated duplexes are assembled to form an octaplex at a low K+ concentration, in which the eight G5 residues form a stacked double G-quartet in the central part. At a higher K+ concentration, however, the octaplex is split into just two halves. These structural features suggest a folding process of eight tandem repeats of d(ccGA[G]4Agg), according to a double Greek-key motif. Such a packaging of the repeats could facilitate slippage of a certain sequence during DNA replication, to induce increase or decrease of the repeats.  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