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.
Activation of precursor 25‐hydroxyvitamin D3 (25D) to hormonal 1,25‐dihydroxyvitamin D3 (1,25D) is a pivotal step in vitamin D physiology, catalysed by the enzyme 25‐hydroxyvitamin D‐1α‐hydroxylase (1α‐hydroxylase). To establish new models for assessing the physiological importance of the 1α‐hydroxylase‐25D‐axis, we used Danio rerio (zebrafish) to characterize expression and biological activity of the gene for 1α‐hydroxylase (cyp27b1). Treatment of day 5 zebrafish larvae with inactive 25D (5–150 nM) or active 1,25D (0.1–10 nM) induced dose responsive expression (15–95‐fold) of the vitamin D‐target gene cyp24a1 relative to larvae treated with vehicle, suggesting the presence of Cyp27b1 activity. A full‐length zebrafish cyp27b1 cDNA was then generated using RACE and RT‐PCR methods. Sequencing of the resulting clone revealed an open reading frame encoding a protein of 505 amino acids with 54% identity to human CYP27B1. Transfection of a cyp27b1 expression vector into HKC‐8, a human kidney proximal tubular epithelial cell line, enhanced intracrine metabolism of 25D to 1,25D resulting in greater than twofold induction of CYP24A1 mRNA expression and a 25‐fold increase in 1,25D production compared to empty vector. These data indicate that we have cloned a functional zebrafish CYP27B1, representing a phylogenetically distant branch from mammals of this key enzyme in vitamin D metabolism. Further analysis of cyp27b1 expression and activity in zebrafish may provide new perspectives on the biological importance of 25D metabolism. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

2.
Short‐term temperature effects on photosynthesis were investigated by measuring O2 production, PSII‐fluorescence kinetics, and 14C‐incorporation rates in monocultures of the marine phytoplankton species Prorocentrum minimum (Pavill.) J. Schiller (Dinophyceae), Prymnesium parvum f. patelliferum (J. C. Green, D. J. Hibberd et Pienaar) A. Larsen (Coccolithophyceae), and Phaeodactylum tricornutum Bohlin (Bacillariophyceae), grown at 15°C and 80 μmol photons · m?2 · s?1. Photosynthesis versus irradiance curves were measured at seven temperatures (0°C–30°C) by all three approaches. The maximum photosynthetic rate (PCmax) was strongly stimulated by temperature, reached an optimum for Pro. minimum only (20°C–25°C), and showed a similar relative temperature response for the three applied methods, with Q10 ranging from 1.7 to 3.5. The maximum light utilization coefficient (αC) was insensitive or decreased slightly with increasing temperature. Absolute rates of O2 production were calculated from pulse‐amplitude‐modulated (PAM) fluorometry measurements in combination with biooptical determination of absorbed quanta in PSII. The relationship between PAM‐based O2 production and measured O2 production and 14C assimilation showed a species‐specific correlation, with 1.2–3.3 times higher absolute values of PCmax and αC when calculated from PAM data for Pry. parvum and Ph. tricornutum but equivalent for Pro. minimum. The offset seemed to be temperature insensitive and could be explained by a lower quantum yield for O2 production than the theoretical maximum (due to Mehler‐type reactions). Conclusively, the PAM technique can be used to study temperature responses of photosynthesis in microalgae when paying attention to the absorption properties in PSII.  相似文献   

3.
We established clonal cultures of Dinophysis acuminata Clap. et Lachm. and D. fortii Pavill. isolated from western Japan and examined toxin production in them, focusing on intracellular production and extracellular excretion. At the end of incubations, the total amounts of pectenotoxin‐2 (PTX‐2), dinophysistoxin‐1 (DTX‐1), and okadaic acid (OA) in the D. acuminata cultures reached up to 672.7 ± 14.7 (mean ± SD), 88.1 ± 2.8, and 539.3 ± 39.7 ng · mL?1, respectively, and the excreted extracellular amounts were equivalent to 5.1, 79.5, and 79.5% of the total amounts, respectively. Similarly, at the end of incubations, the total amounts of PTX‐2, DTX‐1, and OA in the D. fortii cultures reached up to 526.6 ± 52.6 (mean ±SD), 4.4 ± 0.4, and 135.9 ± 3.9 ng · mL?1, respectively, and the excreted extracellular amounts were equivalent to 1.8, 80.1, and 86.6% of the total amounts, respectively. Further, we tested the availability of cell debris and dissolved organic substances that originated from the ciliate prey Myrionecta rubra for growth and toxin production in D. acuminata. Although no significant growth was observed in D. acuminata in the medium containing the cell debris and organic substances originated from M. rubra, the toxicity was significantly greater than that in the control (P < 0.05–0.001); this finding suggested the availability of organic substances for toxin production. However, toxin productivity was remarkably lower than that of Dinophysis species feeding on living M. rubra.  相似文献   

4.
2D metal organic frameworks (MOF) have received tremendous attention due to their organic–inorganic hybrid nature, large surface area, highly exposed active sites, and ultrathin thickness. However, the application of 2D MOF in light‐to‐hydrogen (H2) conversion is rarely reported. Here, a novel 2D MOF [Ni(phen)(oba)]n·0.5nH2O (phen = 1,10‐phenanthroline, oba = 4,4′‐oxybis(benzoate)) is for the first time employed as a general, high‐performance, and earth‐abundant platform to support CdS or Zn0.8Cd0.2S for achieving tremendously improved visible‐light‐induced H2‐production activity. Particularly, the CdS‐loaded 2D MOF exhibits an excellent H2‐production activity of 45 201 µmol h?1 g?1, even exceeding that of Pt‐loaded CdS by 185%. Advanced characterizations, e.g., synchrotron‐based X‐ray absorption near edge structure, and theoretical calculations disclose that the interactive nature between 2D MOF and CdS, combined with the high surface area, abundant reactive centers, and favorable band structure of 2D MOFs, synergistically contribute to this distinguished photocatalytic performance. The work not only demonstrates that the earth‐abundant 2D MOF can serve as a versatile and effective platform supporting metal sulfides to boost their photocatalytic H2‐production performance without noble‐metal co‐catalysts, but also paves avenues to the design and synthesis of 2D‐MOF‐based heterostructures for catalysis and electronics applications.  相似文献   

5.
The plastidic glyceraldehyde‐3‐phosphate dehydrogenase (GAPDH) catalyzes the only reductive step in the Calvin cycle and exists as different forms of which GapC1 enzyme is present in chromalveolates, such as diatoms. Biochemical studies on diatoms are still fragmentary, and, thus, in this report, GAPDH from the freshwater diatom Asterionella formosa Hassall has been purified and kinetically characterized. It is a homotetrameric enzyme with a molecular mass of ~150 ± 15 kDa. The enzyme showed Michaelis–Menten kinetics with respect to both cofactors, NADPH and NADH, with a 16‐fold greater catalytic constant for NADPH. The Km for NADPH was 140 μM, the lowest affinity reported, while the catalytic constant, 815 s?1, is the highest reported. The Km for NADH was 93 μM, and the catalytic constant was 50 s?1, both are similar to reported values for other types of GAPDH. The GapC1 enzyme, like the Chlamydomonas reinhardtii A4 GAPDH, exhibits a cooperative behavior toward the substrate, 1,3‐bisphosphoglyceric acid (BPGA), with both cofactors. Mass spectrometry analysis showed that when GapC1 enzyme was purified without reducing agents, it copurified with a small protein with a mass of 8.2 kDa. This protein was recognized by antibodies against CP12. When associated with this protein, GAPDH displayed a lag that disappeared upon incubation with reducing agent in the presence of either BPGA or NADPH as a consequence of dissociation of the GAPDH/CP12 complex. Thus, as in other species of algae and higher plants, regulation of GapC1 enzyme in A. formosa may occur through association‐dissociation processes linked to dark‐light transitions.  相似文献   

6.
The rdpA gene of strains Delftia acidovorans MC1, Rhodoferax sp. P230, and Sphingobium herbicidovorans MH proved to be identical. However, when RdpA [(R)‐2‐(2,4‐dichlorophenoxy)propionate/α‐ketoglutarate dioxygenase] was investigated after purification from the various strains, significant differences in the kinetics and some chemical properties of the enzymes were observed. The preference for substrates ranged in the order (R)‐2‐(2,4‐dichlorophenoxy)propionate (2,4‐DP) > (R)‐2‐(4‐chloro‐2‐methylphenoxy)propionate (MCPP) >> 2,4‐dichlorophenoxyacetate (2,4‐D) ~ 4‐chloro‐2‐methylphenoxyacetate (MCPA), but detailed kinetic investigations revealed significant strain‐dependent differences in the kcat and KM values. While the KM values of RdpA from the various strains were low and their range rather narrow with 2,4‐DP (19–60 μM) and MCPP (35–64 μM), larger differences were observed with phenoxyacetates which were distinctly higher and spanned a wider range with 2,4‐D (237–935 μM) and MCPA (164–510 μM). The lowest KM values with 2,4‐D and MCPA were found for RdpA originating from strain P230. Investigation of the enzymes from the various sources by 2D gel electrophoresis revealed up to three monomeric enzyme forms which differed in the pI value. The 2D‐patterns were similar with RdpA from strains MC1 and MH, and after heterologous expression of the enzyme in Escherichia coli, but differed significantly from that of strain P230. The presence of enzyme forms and their different composition coincided apparently with the differences observed in the kinetic properties of RdpA in the various strains. The effects are discussed in terms of posttranslational modification of RdpA which appears to be different in extent and kind in the various strains.  相似文献   

7.
Hydrogen and methane production from desugared molasses by a two‐stage thermophilic anaerobic process was investigated in a series of two up‐flow anaerobic sludge blanket (UASB) reactors. The first reactor that was dominated with hydrogen‐producing bacteria of Thermoanaerobacterium thermosaccharolyticum and Thermoanaerobacterium aciditolerans could generate a high hydrogen production rate of 5600 mL H2/day/L, corresponding to a yield of 132 mL H2/g volatile solid (VS). The effluent from the hydrogen reactor was further converted to methane in the second reactor with the optimal production rate of 3380 mL CH4/day/L, corresponding to a yield of 239 mL CH4/g VS. Aceticlastic Methanosarcina mazei was the dominant methanogen in the methanogenesis stage. This work demonstrates that biohydrogen production can be very efficiently coupled with a subsequent step of methane production using desugared molasses. Furthermore, the mixed gas with a volumetric content of 16.5% H2, 38.7% CO2, and 44.8% CH4, containing approximately 15% energy by hydrogen is viable to be bio‐hythane.  相似文献   

8.
Four new 13,14‐seco‐withanolides, minisecolides A – D ( 1  –  4 ), together with three known analogues 5  –  7 , were isolated from the whole plants of Physalis minima. The structures of new compounds were determined on the basis of spectroscopic analysis, including 1H‐, 13C‐NMR, 2D‐NMR (HMBC, HSQC, ROESY), and HR‐ESI‐MS. Evaluation of all isolates for their inhibitory effects on nitric oxide (NO) production was conducted on lipopolysaccaride‐activated RAW264.7 macrophages. Compounds 2 , 3 , 5 , and 6 showed inhibitory activities, especially for compound 5 with IC50 value of 3.87 μm .  相似文献   

9.
Diatoms are frequently exposed to high light (HL) levels, which can result in photoinhibition and damage to PSII. Many microalgae can photoreduce oxygen using the Mehler reaction driven by PSI, which could protect PSII. The ability of Nitzschia epithemioides Grunow and Thalassiosira pseudonana Hasle et Heimdal grown at 50 and 300 μmol photons · m?2 · s?1 to photoreduce oxygen was examined by mass spectrometric measurements of 18O2. Both species exhibited significant rates of oxygen photoreduction at saturating light levels, with cells grown in HL exhibiting higher rates. HL‐grown T. pseudonana had maximum rates of oxygen photoreduction five times greater than N. epithemoides, with 49% of electrons transported through PSII being used to reduce oxygen. Exposure to excess light (1,000 μmol photons · m?2 · s?1) produced similar decreases in the operating quantum efficiency of PSII (Fq′/Fm′) of low light (LL)‐ and HL‐grown N. epithemoides, whereas HL‐grown T. pseudonana exhibited much smaller decreases in Fq′/Fm′ than LL‐grown cells. HL‐grown T. pseudonana and N. epithemioides exhibited greater superoxide and hydrogen peroxide production, higher activities (in T. pseudonana) of superoxide dismutase (SOD) and ascorbate peroxidase (APX), and increased expression of three SOD‐ and one APX‐encoding genes after 60 min of excess light compared to LL‐grown cells. These responses provide a mechanism that contributes to the photoprotection of PSII against photodamage.  相似文献   

10.
A gene encoding an sn‐glycerol‐1‐phosphate dehydrogenase (G1PDH) was identified in the hyperthermophilic archaeon Pyrobaculum calidifontis. The gene was overexpressed in Escherichia coli, and its product was purified and characterized. In contrast to conventional G1PDHs, the expressed enzyme showed strong preference for NADH: the reaction rate (Vmax) with NADPH was only 2.4% of that with NADH. The crystal structure of the enzyme was determined at a resolution of 2.45 Å. The asymmetric unit consisted of one homohexamer. Refinement of the structure and HPLC analysis showed the presence of the bound cofactor NADPH in subunits D, E, and F, even though it was not added in the crystallization procedure. The phosphate group at C2’ of the adenine ribose of NADPH is tightly held through the five biased hydrogen bonds with Ser40 and Thr42. In comparison with the known G1PDH structure, the NADPH molecule was observed to be pushed away from the normal coenzyme binding site. Interestingly, the S40A/T42A double mutant enzyme acquired much higher reactivity than the wild‐type enzyme with NADPH, which suggests that the biased interactions around the C2’‐phosphate group make NADPH binding insufficient for catalysis. Our results provide a unique structural basis for coenzyme preference in NAD(P)‐dependent dehydrogenases. Proteins 2016; 84:1786–1796. © 2016 Wiley Periodicals, Inc.  相似文献   

11.
Phosphofructokinase‐1 from Saccharomyces cerevisiae is composed of two types of subunits, α and β. Subunit‐specific monoclonal antibodies were raised to elucidate structural and functional properties of both subunits. One monoclonal antibody, α‐F3, binds to an epitope either at the C‐terminal or at the N‐terminal part of the α‐polypeptide chain. By screening a heptapeptide library with this monoclonal antibody, a set of heptapeptides was selected, which contained the consensus sequences D–A–F and D–S–F. Two heptapeptides with these motifs were synthesized in order assess their capacity to inhibit the binding of antibody α‐F3 to native phosphofructokinase‐1. The peptide G–I–K–D–A–F–L inhibited the binding more strongly (IC50 = 1.5 µM) than the peptide A–P–W–H–D–S–F (IC50 = 33.3 µM). Sequence matching revealed the presence of the D–A–F motif in the polypeptide chain of phosphofructokinase‐1 at amino acid position 172–174. As a control, the nonapeptide A–P–T–S–K–D–A–F–L which corresponds to the sequence of the putative epitope was tested in the inhibition assay. In view of the high inhibitory capacity (IC50 = 0.3 µM) it was concluded that this nonapeptide represents the continuous epitope of phosphofructokinase‐1 that is recognized by antibody α‐F3. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

12.
Epilepsy, one of the most frequent neurological disorders, is still insufficiently treated in about 30% of patients. As a consequence, identification of novel anticonvulsant agents is an important issue in medicinal chemistry. In the present article we report synthesis, physicochemical, and pharmacological evaluation of N‐trans‐cinnamoyl derivatives of R and S‐2‐aminopropan‐1‐ol, as well as R and S‐2‐aminobutan‐1‐ol. The structures were confirmed by spectroscopy and for derivatives of 2‐aminopropan‐1‐ols the configuration was evaluated by means of crystallography. The investigated compounds were tested in rodent models of seizures: maximal electroshock (MES) and subcutaneous pentetrazol test (scPTZ), and also in a rodent model of epileptogenesis: pilocarpine‐induced status prevention. Additionally, derivatives of 2‐aminopropan‐1‐ols were tested in benzodiazepine‐resistant electrographic status epilepticus rat model as well as in vitro for inhibition of isoenzymes of cytochrome P450. All of the tested compounds showed promising anticonvulsant activity in MES. For R(–)‐(2E)‐N‐(1‐hydroxypropan‐2‐yl)‐3‐phenylprop‐2‐enamide pharmacological parameters were found as follows: ED50 = 76.7 (68.2–81.3) mg/kg (MES, mice i.p., time = 0.5 h), ED50 = 127.2 (102.1–157.9) mg/kg (scPTZ, mice i.p., time = 0.25 h), TD50 = 208.3 (151.4–230.6) mg/kg (rotarod, mice i.p., time = 0.25 h). Evaluation in pilocarpine status prevention proved that all of the reported compounds reduced spontaneous seizure activity and act as antiepileptogenic agents. Both enantiomers of 2‐aminopropan‐1‐ols did not influence cytochrome P450 isoenzymes activity in vitro and are likely not to interact with CYP substrates in vivo. Chirality 28:482–488, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

13.
A high degree of pseudo‐cryptic diversity was reported in the well‐studied diatom genus Pseudo‐nitzschia. Studies off the coast of Washington State revealed the presence of hitherto undescribed diversity of Pseudo‐nitzschia. Forty‐one clonal strains, representing six different taxa of the P. pseudodelicatissima complex, were studied morphologically using LM and EM, and genetically using genes from three different cellular compartments: the nucleus (D1–D3 of the LSU of rDNA and internal transcribed spacers [ITSs] of rDNA), the mitochondria (cytochrome c oxidase 1), and the plastids (LSU of RUBISCO). Strains in culture at the same time were used in mating studies to study reproductive isolation of species, and selected strains were examined for the production of the neurotoxin domoic acid (DA). Two new species, P. hasleana sp. nov. and P. fryxelliana sp. nov., are described based on morphological and molecular data. In all phylogenetic analyses, P. hasleana appeared as sister taxa to a clade comprising P. calliantha and P. mannii, whereas the position of P. fryxelliana was more uncertain. In the phylogenies of ITS, P. fryxelliana appeared to be most closely related to P. cf. turgidula. Morphologically, P. hasleana differed from most other species of the complex because of a lower density of fibulae, whereas P. fryxelliana had fewer sectors in the poroids and a higher poroid density than most of the other species. P. hasleana did not produce detectable levels of DA; P. fryxelliana was unfortunately not tested. In P. cuspidata, production of DA in offspring cultures varied from higher than the parent cultures to undetectable.  相似文献   

14.
An efficient synthetic strategy to 3‐methylidene‐2,3‐dihydroquinolin‐4(1H)‐ones variously substituted in position 2 has been developed. The title compounds were synthesized in the reaction sequence involving reaction of diethyl methylphosphonate with methyl 2‐(tosylamino)benzoate, condensation of thus formed diethyl 2‐oxo‐2‐(2‐N‐tosylphenyl)ethylphosphonate with various aldehydes followed by successful application of the obtained 3‐(diethoxyphosphoryl)‐1,2‐dihydroquinolin‐4‐ols as Horner–Wadsworth–Emmons reagents for the olefination of formaldehyde. Also, enantioselective approach to the target compounds has been evaluated using 3‐dimenthoxyphosphoryl group as a chiral auxiliary. Single X‐ray crystal analysis of (2S)‐3‐(dimenthoxyphosphoryl)‐2‐phenyl‐1‐tosyldihydroquinolin‐4‐ol revealed the presence of strong resonance‐assisted hydrogen bond (RAHB). The obtained 3‐methylidene‐2,3‐dihydroquinolin‐4(1H)‐ones were then tested for their cytotoxic activity against two leukemia cell lines NALM‐6 and HL‐60 and a breast cancer MCF‐7 cell line. All compounds showed very high cytotoxic activity with the IC50 values mostly below 1 μm in all three cancer cell lines. The selected analogs were also tested on human umbilical vein endothelial cells (HUVEC) and on human mammary gland/breast cells (MCF‐10A) to evaluate their influence on normal cells. Since one of the most serious problems in cancer chemotherapy is the development of drug resistance, the mRNA levels and activity of ABCB1 transporter considered to be the most important factor engaged in drug resistance, were evaluated in MCF‐7 cells treated with two selected analogs. Both compounds were strong ABCB1 transporter inhibitors that could prevent efflux of anticancer drugs from cancer cells.  相似文献   

15.
Human leukemic THP‐1 promonocytes are widely used as a model for peripheral blood monocytes. However, superoxide production during respiratory burst (RB) of non‐differentiated THP‐1 (nd‐THP‐1) cells is very low. Here we present a rapid and low‐cost method for measuring the chemiluminescence (CL) of opsonized zymosan (OZ) induced RB which allows detection of Escherichia coli lipopolysaccharide (LPS) induced priming of nd‐THP‐1 cells on the basis of CL reaction kinetics. Maximum CL intensity obtained was 2.20 ± 0.25 and 1.30 ± 0.11 relative light units, while CL peak time was achieved at 18.1 ± 2.6 and 28.7 ± 1.3 min in primed and non‐primed cells, respectively. The priming of nd‐THP‐1 cells with LPS evoked typical TNF‐α and IL‐6 production. We tested the effects of bovine lactoferrin and protein fractions from Lactobacillus helveticus BGRA43 fermented milk for potential anti‐inflammatory effects on LPS primed nd‐THP‐1 cells. Four fractions were found to inhibit the OZ‐induced CL in a dose‐dependent manner (IC50 3–30 µg/mL), while lactoferrin inhibited CL to a lesser extent (IC50 270 µg/mL). These results suggest that measuring CL response of nd‐THP‐1 cells can serve as a method for screening anti‐inflammatory compounds which could be used in reducing the risk of phagocyte‐mediated inflammatory diseases. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

16.
This study assessed the implication of oxidative stress in the mortality of cells of Microcystis aeruginosa Kütz. Cultures grown at 25°C were exposed to 32°C, darkness, and hydrogen peroxide (0.5 mM) for 96 h. The cellular abundance, chl a concentration and content, maximum photochemical efficiency of PSII (Fv/Fm ratio), intracellular oxidative stress (determined with dihydrorhodamine 123 [DHR]), cell mortality (revealed by SYTOX‐labeling of DNA), and activation of caspase 3–like proteins were assessed every 24 h. The presence of DNA degradation in cells of M. aeruginosa was also assessed using a terminal deoxynucletidyl transferase‐mediated dUTP nick end labeling (TUNEL) assay at 96 h. Transferring cultures from 25°C to 32°C was generally beneficial to the cells. The cellular abundance and chl a concentration increased, and the mortality remained low (except for a transient burst at 72 h) as did the oxidative stress. In darkness, cells did not divide, and the Fv/Fm continuously decreased with time. The slow increase in intracellular oxidative stress coincided with the activation of caspase 3–like proteins and a 15% and 17% increase in mortality and TUNEL‐positive cells, respectively. Exposure to hydrogen peroxide had the most detrimental effect on cells as growth ceased and the Fv/Fm declined to near zero in less than 24 h. The 2‐fold increase in oxidative stress matched the activation of caspase 3–like proteins and a 40% and 37% increase in mortality and TUNEL‐positive cells, respectively. These results demonstrate the implication of oxidative stress in the stress response and mortality of M. aeruginosa.  相似文献   

17.
Stratospheric ozone depletion increases the amount of ultraviolet‐B radiation (UVBR) (280–320 nm) reaching the surface of the earth, potentially affecting phytoplankton. In this work, Anabaena sp. PCC 7120, a typically nitrogen (N)‐fixing filamentous bloom‐forming cyanobacterium in freshwater, was individually cultured in N‐deficient and N‐enriched media for long‐term acclimation before being subjected to ultraviolet‐B (UVB) exposure experiments. Results suggested that the extent of breakage in the filaments induced by UVBR increases with increasing intensity of UVB stress. In general, except for the 0.1 W · m?2 treatment, which showed a mild increase, UVB exposure inhibits photosynthesis as evidenced by the decrease in the chl fluorescence parameters maximum photochemical efficiency of PSII (Fv/Fm) and maximum relative electron transport rate. Complementary chromatic acclimation was also observed in Anabaena under different intensities of UVB stress. Increased total carbohydrate and soluble protein may provide some protection for the culture against damaging UVB exposure. In addition, N‐deficient cultures with higher recovery capacity showed overcompensatory growth under low UVB (0.1 W · m?2) exposure during the recovery period. Significantly increased (~830%) ATPase activity may provide enough energy to repair the damage caused by exposure to UVB.  相似文献   

18.
Brassinosteroids (BRs) are essential for plant growth and development; however, whether and how they promote stomatal closure is not fully clear. In this study, we report that 24‐epibrassinolide (EBR), a bioactive BR, induces stomatal closure in Arabidopsis (Arabidopsis thaliana) by triggering a signal transduction pathway including ethylene synthesis, the activation of Gα protein, and hydrogen peroxide (H2O2) and nitric oxide (NO) production. EBR initiated a marked rise in ethylene, H2O2 and NO levels, necessary for stomatal closure in the wild type. These effects were abolished in mutant bri1‐301, and EBR failed to close the stomata of gpa1 mutants. Next, we found that both ethylene and Gα mediate the inductive effects of EBR on H2O2 and NO production. EBR‐triggered H2O2 and NO accumulation were canceled in the etr1 and gpa1 mutants, but were strengthened in the eto1‐1 mutant and the cGα line (constitutively overexpressing the G protein α‐subunit AtGPA1). Exogenously applied H2O2 or sodium nitroprusside (SNP) rescued the defects of etr1‐3 and gpa1 or etr1 and gpa1 mutants in EBR‐induced stomatal closure, whereas the stomata of eto1‐1/AtrbohF and cGα/AtrbohF or eto1‐1/nia1‐2 and cGα/nia1‐2 constructs had an analogous response to H2O2 or SNP as those of AtrbohF or Nia1‐2 mutants. Moreover, we provided evidence that Gα plays an important role in the responses of guard cells to ethylene. Gα activator CTX largely restored the lesion of the etr1‐3 mutant, but ethylene precursor ACC failed to rescue the defects of gpa1 mutants in EBR‐induced stomatal closure. Lastly, we demonstrated that Gα‐activated H2O2 production is required for NO synthesis. EBR failed to induce NO synthesis in mutant AtrbohF, but it led to H2O2 production in mutant Nia1‐2. Exogenously applied SNP rescued the defect of AtrbohF in EBR‐induced stomatal closure, but H2O2 did not reverse the lesion of EBR‐induced stomatal closure in Nia1‐2. Together, our results strongly suggest a signaling pathway in which EBR induces ethylene synthesis, thereby activating Gα, and then promotes AtrbohF‐dependent H2O2 production and subsequent Nia1‐catalyzed NO accumulation, and finally closes stomata.  相似文献   

19.
Mimicry of structural motifs is a common feature in proteins. The 10‐membered hydrogen‐bonded ring involving the main‐chain C?O in a β‐turn can be formed using a side‐chain carbonyl group leading to Asx‐turn. We show that the N? H component of hydrogen bond can be replaced by a Cγ‐H group in the side chain, culminating in a nonconventional C? H···O interaction. Because of its shape this β‐turn mimic is designated as ω‐turn, which is found to occur ~three times per 100 residues. Three residues (i to i + 2) constitute the turn with the C? H···O interaction occurring between the terminal residues, constraining the torsion angles ?i + 1, ψi + 1, ?i + 2 and χ1(i + 2) (using the interacting Cγ atom). Based on these angles there are two types of ω‐turns, each of which can be further divided into two groups. Cβ‐branched side‐chains, and Met and Gln have high propensities to occur at i + 2; for the last two residues the carbonyl oxygen may participate in an additional interaction involving the S and amino group, respectively. With Cys occupying the i + 1 position, such turns are found in the metal‐binding sites. N‐linked glycosylation occurs at the consensus pattern Asn‐Xaa‐Ser/Thr; with Thr at i + 2, the sequence can adopt the secondary structure of a ω‐turn, which may be the recognition site for protein modification. Location between two β‐strands is the most common occurrence in protein tertiary structure, and being generally exposed ω‐turn may constitute the antigenic determinant site. It is a stable scaffold and may be used in protein engineering and peptide design. Proteins 2015; 83:203–214. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
We investigated copper (Cu) acquisition mechanisms and uptake kinetics of the marine diatoms Thalassiosira oceanica Hasle, an oceanic strain, and Thalassiosira pseudonana Hasle et Heimdal, a coastal strain, grown under replete and limiting iron (Fe) and Cu availabilities. The Cu‐uptake kinetics of these two diatoms followed classical Michaelis–Menten kinetics. Biphasic uptake kinetics as a function of Cu concentration were observed, suggesting the presence of both high‐ and low‐affinity Cu‐transport systems. The half‐saturation constants (Km) and the maximum Cu‐uptake rates (Vmax) of the high‐affinity Cu‐transport systems (~7–350 nM and 1.5–17 zmol · μm?2 · h?1, respectively) were significantly lower than those of the low‐affinity systems (>800 nM and 30–250 zmol · μm?2 · h?1, respectively). The two Cu‐transport systems were controlled differently by low Fe and/or Cu. The high‐affinity Cu‐transport system of both diatoms was down‐regulated under Fe limitation. Under optimal‐Fe and low‐Cu growth conditions, the Km of the high‐affinity transport system of T. oceanica was lower (7.3 nM) than that of T. pseudonana (373 nM), indicating that T. oceanica had a better ability to acquire Cu at subsaturating concentrations. When Fe was sufficient, the low‐affinity Cu‐transport system of T. oceanica saturated at 2,000 nM Cu, while that of T. pseudonana did not saturate, indicating different Cu‐transport regulation by these two diatoms. Using CuEDTA as a model organic complex, our results also suggest that diatoms might be able to access Cu bound within organic Cu complexes.  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