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 31 毫秒
1.
A Δsud deletion mutant of Wolinella succinogenes that lacked the periplasmic sulfide dehydrogenase (Sud) was constructed using homologous recombination. The mutant grew with sulfide and fumarate, indicating that Sud was not a component of the electron transport chain that catalyzed fumarate respiration with sulfide as an electron donor. Likewise, growth with formate and either polysulfide or sulfur was not affected by the deletion. Removal of Sud from wild-type W. succinogenes by spheroplast formation did not decrease the activity of electron transport to polysulfide. The Δpsr deletion mutant that lacks polysulfide reductase (Psr) grew by fumarate respiration with sulfide as an electron donor, indicating that Psr is not required for this activity. Received: 31 August 1995 / Accepted: 25 October 1995  相似文献   

2.
During growth with fumarate as the terminal electron transport acceptor and either formate or sulfide as the electron donor, Wolinella succinogenes induced a peri-plasmic protein (54 kDa) that reacted with an antiserum raised against the periplasmic fumarate reductase (Fcc) of Shewanella putrefaciens. However, the periplasmic cell fraction of W. succinogenes did not catalyze fumarate reduction with viologen radicals. W. succinogenes grown with polysulfide instead of fumarate contained much less (< 10%) of the 54-kDa antigen, and the antigen was not detectable in nitrate-grown bacteria. The antigen was most likely encoded by the fccA gene of W. succinogenes. The antigen was absent from a ΔfccABC mutant, and its size is close to that of the protein predicted by fccA. The fccA gene probably encodes a pre-protein carrying an N-terminal signal peptide. The sequence of the mature FccA (481 residues, 52.4 kDa) is similar (31% identity) to that of the C-terminal part (450 residues) of S. putrefaciens fumarate reductase. As indicated by Northern blot analysis, fccA is cotranscribed with fccB and fccC. The proteins predicted from the fccB and fccC gene sequences represent tetraheme cytochromes c. FccB is similar to the N-terminal part (150 residues) of S. putrefaciens fumarate reductase, while FccC resembles the tetraheme cytochromes c of the NirT/NapC family. The ΔfccABC mutant of W. succinogenes grew with fumarate and formate or sulfide, suggesting that the deleted proteins were not required for fumarate respiration with either electron donor. Received: 26 September 1997 / Accepted: 8 December  相似文献   

3.
Wolinella succinogenes can grow by anaerobic respiration with fumarate or polysulfide as the terminal electron acceptor, and H2 or formate as the electron donor. A ΔhydABC mutant lacking the hydrogenase structural genes did not grow with H2 and either fumarate or polysulfide. In contrast to the wild-type strain, the mutant grown with fumarate and with formate instead of H2 did not catalyze the reduction of fumarate, polysulfide, dimethylnaphthoquinone, or benzyl viologen by H2. Growth and enzymic activities were restored upon integration of a plasmid carrying hydABC into the genome of the ΔhydABC mutant. The ΔhydABC mutant was complemented with hydABC operons modified by artificial stop codons in hydA (StopA) or at the 5′-end of hydC (StopC). The StopC mutant lacked HydC, and the hydrophobic C-terminus of HydA was missing in the hydrogenase of the StopA mutant. The two mutants catalyzed benzyl viologen reduction by H2. The enzyme activity was located in the membrane of the mutants. A mutant with both modifications (StopAC) contained the activity in the periplasm. The three mutants did not grow with H2 and either fumarate or polysulfide, and did not catalyze dimethylnaphthoquinone reduction by H2. We conclude that the same hydrogenase serves in the anaerobic respiration with fumarate and with polysulfide. HydC and the C-terminus of HydA appear to be required for both routes of electron transport and for dimethylnaphthoquinone reduction by H2. The hydrogenase is anchored in the membrane by HydC and by the C-terminus of HydA. The catalytic subunit HydB is oriented towards the periplasmic side of the membrane. Received: 29 December 1997 / Accepted: 6 March 1998  相似文献   

4.
 Reactions (25  °C) of galactose oxidase, GOaseox from Fusarium NRRL 2903 with five different primary-alcohol-containing substrates RCH2OH:- D-galactose (I) and 2-deoxy-d-galactose (II) (monosaccharides); methyl-β-d-galactopyranoside (III) (glycoside);d-raffinose (IV) (trisaccharide); and dihydroxyacetone (V) have been studied in the presence of O2. The GOaseox state has a tyrosyl radical coordinated at a square-pyramidal CuII active site, and is a two-equivalent oxidant. Reactant concentrations were [GOaseox] (0.8–10 μM), RCH2OH (1.0–6.0 mM), and O2 (0.14–0.29 mM), with I=0.100 M (NaCl). The reactions, monitored at 450 nm by stopped-flow spectrophotometry, terminated with depletion of the O2. Each trace was fitted to the competing reactions GOaseox+RCH2 OH → GOaseredH2+RCHO (k 1), and GOaseredH2+O2→ GOaseox+H2O2 (k 2), with GOaseredH2 written as the doubly protonated two-electron-reduced CuI product. It was necessary to avoid auto-redox interconversion of GOaseox and GOasesemi . Information obtained at pH 7.5 indicates a 5 : 95 (ox : semi) "native" mix equilibration complete in ∼3 h. At pH >7.5, rate constants 10–4k 1 / M–1 s–1 for the reactions of GOaseox with (I) (1.19), (II) (1.07), (III) (1.29), (IV) (1.81), (V) (2.94) were determined. On decreasing the pH to 5.5, k 1 values decreased by factors of up to a half, and acid dissociation pK as in the range 6.6–6.9 were obtained. UV-Vis spectrophotometric studies on GOaseox gave an independently determined pK a of 6.7. No corresponding reactions of the Tyr495Phe variant were observed, and there are no similar UV-Vis absorbance changes for this variant. The pK a is therefore assigned to protonation of Tyr-495 which is a ligand to the Cu. The rate constant k 2 (1.01×107 M–1 s–1) is independent of pH in the range 5.5–9.0 investigated, suggesting that H+ (or H-atoms) for the O2 → H2O2 change are provided by the active site of GOasered . The CuI of GOasered is less extensively complexed, and a coordination number of three is likely. Received: 4 February 1997 / Accepted: 16 May 1997  相似文献   

5.
In [FeFe]-hydrogenases, the H cluster (hydrogen-activating cluster) contains a di-iron centre ([2Fe]H subcluster, a (L)(CO)(CN)Fe(μ-RS2)(μ-CO)Fe(CysS)(CO)(CN) group) covalently attached to a cubane iron-sulphur cluster ([4Fe-4S]H subcluster). The Cys-thiol functions as the link between one iron (called Fe1) of the [2Fe]H subcluster and one iron of the cubane subcluster. The other iron in the [2Fe]H subcluster is called Fe2. The light sensitivity of the Desulfovibrio desulfuricans enzyme in a variety of states has been studied with infrared (IR) spectroscopy. The aerobic inactive enzyme (Hinact state) and the CO-inhibited active form (Hox–CO state) were stable in light. Illumination of the Hox state led to a kind of cannibalization; in some enzyme molecules the H cluster was destroyed and the released CO was captured by the H clusters in other molecules to form the light-stable Hox–CO state. Illumination of active enzyme under 13CO resulted in the complete exchange of the two intrinsic COs bound to Fe2. At cryogenic temperatures, light induced the photodissociation of the extrinsic CO and the bridging CO of the enzyme in the Hox–CO state. Electrochemical redox titrations showed that the enzyme in the Hinact state converts to the transition state (Htrans) in a reversible one-electron redox step (E m, pH 7=–75 mV). IR spectra demonstrate that the added redox equivalent not only affects the [4Fe-4S]H subcluster, but also the di-iron centre. Enzyme in the Htrans state reacts with extrinsic CO, which binds to Fe2. The Htrans state converts irreversibly into the Hox state in a redox-dependent reaction most likely involving two electrons (E m, pH 7=–261 mV). These electrons do not end up on any of the six Fe atoms of the H cluster; the possible destiny of the two redox equivalents is discussed. An additional reversible one-electron redox reaction leads to the Hred state (E m, pH 7=–354 mV), where both Fe atoms of the [2Fe]H subcluster have the same formal oxidation state. The possible oxidation states of Fe1 and Fe2 in the various enzyme states are discussed. Low redox potentials (below –500 mV) lead to destruction of the [2Fe]H subcluster.  相似文献   

6.
Elevated values of molar growth yield (Yx/s = 14–26 g mol–1) were obtained during exponential growth (μ > 0.4 h–1) of Zymomonas mobilis ATCC 29191 by using reduced concentrations of glucose (6.25–100 mM) and increased oxygen supply (E h > 300 mV) in the growth medium, as compared to the Yx/s of anaerobic exponential growth (8–10 g mol–1). Aerobically grown cells showed an increased maximum growth rate (μmax), and a reduced specific glucose consumption rate (qs), and specific ethanol formation rate (qp), thus demonstrating a more pronounced energy-coupling growth under oxic conditions. These results can be neither explained by the concept of a solely operating Entner-Doudoroff pathway as an ATP source in aerobically growing cultures of Z. mobilis nor considered to be consistent with existing data on the lack of the Pasteur effect in this bacterium. Therefore, the results rather give evidence for the essential contribution of aerobic ATP generation under the reported conditions. Received: 24 September 1996 / Accepted: 9 December 1996  相似文献   

7.
Wolinella succinogenes HAP-1 is a Gram-negative microaerophile which reduces perchlorate to chloride by the proposed pathway ClO4 to ClO3 to ClO2 to Cl + O2. A cost-effective perchlorate treatment process has been established using a consortium of facultative anaerobic organisms and W. succinogenes HAP-1. The mixed anaerobic bacterial culture containing W. succinogenes HAP-1 was examined for the ability to form a biofilm capable of perchlorate reduction. An up-flow anaerobic fixed bed reactor (UAFBR) was packed with diatomaceous earth pellets and operated at residence times of 1.17 and 0.46 h to insure a viable biofilm had attached to the diatomaceous earth pellets. Reduction rates of perchlorate to chloride in the UAFBR could be maintained at 1 g of perchlorate reduced h−1 L−1. Studies with the same bacterial consortium in continuously stirred tank reactors (CSTR) generally reduced 0.5–0.7 g of perchlorate h−1. Viable cell counts were performed periodically on the diatomaceous earth pellets and demonstrated that the W. succinogenes HAP-1 population was maintained at 28–47% of the total microbial population. Scanning electron micrographs demonstrated that the external and internal surfaces of the diatomaceous pellets were densely colonized with microbial cells of multiple cell types. This is the first report of an anaerobic mixed culture forming a biofilm capable of perchlorate reduction. Received 22 May 1997/ Accepted in revised form 07 January 1998  相似文献   

8.
 The synthesis, spectroscopic, and electrochemical properties of trans-[L(Pyr)(NH3)4RuII/III] (Pyr=py, 3-phpy, 4-phpy, 3-bnpy, or 4-bnpy; L=H2O, Guo, dGuo, 1MeGuo, Gua, Ino, or G7-DNA) are reported. As expected, the Pyr ligand slows DNA binding by trans-[(H2O)(Pyr)(NH3)4RuII]2+ relative to [(H2O)(NH3)5RuII]2+ and favors reduction of RuIII by about 150 mV. The pyridine ligand also promotes the disproportionation of RuIII to afford the corresponding complexes of RuII and, presumably, RuIV. For L=Ino, disproportionation follows the rate law: d[RuII]/dt=k 0[RuIII]+k 1[OH][RuIII], k 0=(2.7±0.7)×10–4 s–1 and k 1=70±1 M–1 s–1. Received: 11 May 1998 / Accepted: 3 March 1999  相似文献   

9.
The hydrogen-activating cluster (H cluster) in [FeFe]-hydrogenases consists of two moieties. The [2Fe]H subcluster is a (L)(CO)(CN)Fe(μ-RS2)(μ-CO)Fe(CysS)(CO)(CN) centre. The Cys-bound Fe is called Fe1, the other iron Fe2. The Cys-thiol forms a bridge to a [4Fe–4S] cluster, the [4Fe–4S]H subcluster. We report that electron paramagnetic resonance (EPR) spectra of the 57Fe-enriched enzyme from Desulfovibrio desulfuricans in the Hox–CO state are consistent with a magnetic hyperfine interaction of the unpaired spin with all six Fe atoms of the H cluster. In contrast to the inactive aerobic enzyme, the active enzyme is easily destroyed by light. The [2Fe]H subcluster in some enzyme molecules loses CO by photolysis, whereupon other molecules firmly bind the released CO to form the Hox–CO state giving rise to the so-called axial 2.06 EPR signal. Though not destroyed by light, the Hox–CO state is affected by it. As demonstrated in the accompanying paper [49] two of the intrinsic COs, both bound to Fe2, can be exchanged by extrinsic 13CO during illumination at 2 °C. We found that only one of the three 13COs, the one at the extrinsic position, gives an EPR-detectable isotropic superhyperfine interaction of 0.6 mT. At 30 K both the inhibiting extrinsic CO bound to Fe2 and one more CO can be photolysed. EPR spectra of the photolysed products are consistent with a 3d 7 system of Fe with the formal oxidation state +1. The damaged enzyme shows a light-sensitive g=5 signal which is ascribed to an S=3/2 form of the [2Fe]H subcluster. The light sensitivity of the enzyme explains the occurrence of the g=5 signal and the axial 2.06 signal in published EPR spectra of nearly all preparations studied thus far.  相似文献   

10.
The ability of spring barley (Hordeum vulgare cv. Akcent) to adjust the composition and function of the photosynthetic apparatus to growth irradiances of 25–1200 μmol m−2 s−1 was studied by gas exchange and chlorophyll a fluorescence measurements and high-performance liquid chromatography. The increased growth irradiance stimulated light- and CO2-saturated rates of CO2 assimilation expressed on a leaf area basis up to 730 μmol m−2 s−1 (HL730), whereas at an irradiance of 1200 μmol m−2 s−1 (EHL1200) both rates decreased significantly. Further, the acclimation to EHL1200 was associated with an extremely high chlorophyll a/b ratio (3.97), a more than doubled xanthophyll cycle pool (VAZ) and a six-fold higher de-epoxidation state of the xanthophyll cycle pigments as compared to barley grown under 25 μmol m−2 s−1 (LL25). EHL1200 plants also exhibited a long-term inhibition of Photosystem II (PS II) photochemical efficiency (F v/F m). Photosynthetic capacity, chlorophyll a/b and VAZ revealed a linear trend of dependence on PS II excitation pressure in a certain range of growth irradiances (100–730 μmol m−2 s−1). The deviation from linearity of these relationships for EHL1200 barley is discussed. In addition, the role of increased VAZ and/or accumulation of zeaxanthin and antheraxanthin in acclimation of barley to high irradiance is studied with respect to regulation of non-radiative dissipation and/or photochemical efficiency within PS II. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

11.
Direct comparisons between photosynthetic O2 evolution rate and electron transport rate (ETR) were made in situ over 24 h using the benthic macroalga Ulva lactuca (Chlorophyta), growing and measured at a depth of 1.8 m, where the midday irradiance rose to 400–600 μmol photons m−2 s−1. O2 exchange was measured with a 5-chamber data-logging apparatus and ETR with a submersible pulse amplitude modulated (PAM) fluorometer (Diving-PAM). Steady-state quantum yield ((Fm′−Ft)/Fm′) decreased from 0.7 during the morning to 0.45 at midday, followed by some recovery in the late afternoon. At low to medium irradiances (0–300 μmol photons m−2 s−1), there was a significant correlation between O2 evolution and ETR, but at higher irradiances, ETR continued to increase steadily, while O2 evolution tended towards an asymptote. However at high irradiance levels (600–1200 μmol photons m−2 s−1) ETR was significantly lowered. Two methods of measuring ETR, based on either diel ambient light levels and fluorescence yields or rapid light curves, gave similar results at low to moderate irradiance levels. Nutrient enrichment (increases in [NO3 ], [NH4 +] and [HPO4 2-] of 5- to 15-fold over ambient concentrations) resulted in an increase, within hours, in photosynthetic rates measured by both ETR and O2 evolution techniques. At low irradiances, approximately 6.5 to 8.2 electrons passed through PS II during the evolution of one molecule of O2, i.e., up to twice the theoretical minimum number of four. However, in nutrient-enriched treatments this ratio dropped to 5.1. The results indicate that PAM fluorescence can be used as a good indication of the photosynthetic rate only at low to medium irradiances. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

12.
Photosynthetic Response of Carrots to Varying Irradiances   总被引:7,自引:3,他引:4  
Kyei-Boahen  S.  Lada  R.  Astatkie  T.  Gordon  R.  Caldwell  C. 《Photosynthetica》2003,41(2):301-305
Response to irradiance of leaf net photosynthetic rates (P N) of four carrot cultivars: Cascade, Caro Choice (CC), Oranza, and Red Core Chantenay (RCC) were examined in a controlled environment. Gas exchange measurements were conducted at photosynthetic active radiation (PAR) from 100 to 1 000 μmol m−2 s−1 at 20 °C and 350 μmol (CO2) mol−1(air). The values of P N were fitted to a rectangular hyperbolic nonlinear regression model. P N for all cultivars increased similarly with increasing PAR but Cascade and Oranza generally had higher P N than CC. None of the cultivars reached saturation at 1 000 μmol m−2 s−1. The predicted P N at saturation (P Nmax) for Cascade, CC, Oranza, and RCC were 19.78, 16.40, 19.79, and 18.11 μmol (CO2) m−2 s−1, respectively. The compensation irradiance (I c) occurred at 54 μmol m−2 s−1 for Cascade, 36 μmol m−2 s−1 for CC, 45 μmol m−2 s−1 for Oranza, and 25 μmol m−2 s−1 for RCC. The quantum yield among the cultivars ranged between 0.057–0.033 mol(CO2) mol−1(PAR) and did not differ. Dark respiration varied from 2.66 μmol m−2 s−1 for Cascade to 0.85 μmol m−2 s−1 for RCC. As P N increased with PAR, intercellular CO2 decreased in a non-linear manner. Increasing PAR increased stomatal conductance and transpiration rate to a peak between 600 and 800 μmol m−2 s−1 followed by a steep decline resulting in sharp increases in water use efficiency. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

13.
The differences in pigment levels, photosynthetic activity and the chlorophyll fluorescence decrease ratio R Fd (as indicator of photosynthetic rates) of green sun and shade leaves of three broadleaf trees (Platanus acerifolia Willd., Populus alba L., Tilia cordata Mill.) were compared. Sun leaves were characterized by higher levels of total chlorophylls a + b and total carotenoids x + c as well as higher values for the weight ratio chlorophyll (Chl) a/b (sun leaves 3.23–3.45; shade leaves: 2.74–2.81), and lower values for the ratio chlorophylls to carotenoids (a + b)/(x + c) (with 4.44–4.70 in sun leaves and 5.04–5.72 in shade leaves). Sun leaves exhibited higher photosynthetic rates P N on a leaf area basis (mean of 9.1–10.1 μmol CO2 m−2 s−1) and Chl basis, which correlated well with the higher values of stomatal conductance G s (range 105–180 mmol m−2 s−1), as compared to shade leaves (G s range 25–77 mmol m−2 s−1; P N: 3.2–3.7 μmol CO2 m−2 s−1). The higher photosynthetic rates could also be detected via imaging the Chl fluorescence decrease ratio R Fd, which possessed higher values in sun leaves (2.8–3.0) as compared to shade leaves (1.4–1.8). In addition, via R Fd images it was shown that the photosynthetic activity of the leaves of all trees exhibits a large heterogeneity across the leaf area, and in general to a higher extent in sun leaves than in shade leaves.  相似文献   

14.
Photosynthetic parameters of phytoplankton and sea ice algae from landfast sea ice of the Chukchi Sea off Point Barrow, Alaska, were assessed in spring 2005 and winter through spring 2006 using Pulse Amplitude Modulated (PAM) fluorometry including estimates of maximum quantum efficiency (F v/F m), maximum relative electron transport rate (rETRmax), photosynthetic efficiency (α), and the photoadaptive index (E k). The use of centrifuged brine samples allowed to document vertical gradients in ice algal acclimation with 5 cm vertical resolution for the first time. Bottom ice algae (0–5 cm from ice–water interface) expressed low F v/F m (0.331–0.426) and low α (0.098–0.130 (μmol photons m−2s−1)−1) in December. F v/F m and α increased in March and May (0.468–0.588 and 0.141–0.438 (μmol photons m−2s−1)−1, respectively) indicating increased photosynthetic activity. In addition, increases in rETRmax (3.3–16.4 a.u.) and E k (20–88 μmol photons m−2 s−1) from December to May illustrates a higher potential for primary productivity as communities become better acclimated to under-ice light conditions. In conclusion, photosynthetic performance by ice algae (as assessed by PAM fluorometry) was tightly linked to sea ice salinity, temperature, and inorganic nutrient concentrations (mainly nitrogen).  相似文献   

15.
In caulonemal filaments of the moss, Physcomitrella patens, which had been incubated in darkness, 3 s irradiation with blue light (λmax 450 nm) at fluence rates of 100 μmol m−2 s−1 and above caused a transient␣increase in cytosolic calcium ion concentration, [Ca2+]cyt, which was both intensity- and time-dependent. Measurements of [Ca2+]cyt were made using moss transformed with the cDNA for apoaequorin and reconstituting the Ca2+-dependent photoprotein aequorin in the cytosol by incubation in coelenterazine.␣In response to blue light at fluence rates of 100–1000 μmol photons m−2 s−1, [Ca2+]cyt increased transiently from a basal level of approximately 50 nM to between 200 and 700 nM. Irradiation with red light did not evoke any measurable change in [Ca2+]cyt. The presence of calcium in the incubating medium was not required for the increase in [Ca2+]cyt to occur. A mutant strain, gad-139, was identified which required an irradiance of only 1 s to evoke a response. The kinetics showed a delay of approximately 6 s from the beginning of illumination before the beginning of the increase in [Ca2+]cyt. The data suggest that the activation of a photoreceptor rather than the direct opening of calcium channels is involved in this blue-light response. Received: 4 December 1997 / Accepted: 4 May 1998  相似文献   

16.
In an experimental site for reforestation of degraded area, three-year-old plants of Bertholletia excelsa Humb. & Bonpl. were subjected to different fertilization treatments: T0 = unfertilized control, T1 = green fertilization (branches and leaves) and T2 = chemical fertilization. Higher net photosynthetic rates (P N) were observed in T1 [13.2±1.0 μmol(CO2) m−2 s−1] compared to T2 [8.0±1.8 μmol(CO2) m−2 s−1] and T0 [4.8±1.3 μmol(CO2) m−2 s−1]. Stomatal conductance (g s), transpiration rate (E) and water use efficiency (WUE) of individuals of T1 and T2 did not differ significantly, however, they were by 88, 55 and 63%, respectively, higher in T1 than in the control. The mean values of variable fluorescence (Fv), performance index (P.I.) and total chlorophyll [Chl (a+b)] were higher in T1. Our results indicate that green fertilization improves photosynthetic structure and function in plants of B. excelsa in young phase.  相似文献   

17.
N-alkylated polyamine analogues have potential as anticancer and antiparasitic drugs. However, their metabolism in the host has remained incompletely defined thus potentially limiting their utility. Here, we have studied the degradation of three different spermine analogues N,N′-bis-(3-ethylaminopropyl)butane-1,4-diamine (DESPM), N-(3-benzyl-aminopropyl)-N′-(3-ethylaminopropyl)butane-1,4-diamine (BnEtSPM) and N,N′-bis-(3-benzylaminopropyl)butane-1,4-diamine (DBSPM) and related mono-alkylated derivatives as substrates of recombinant human polyamine oxidase (APAO) and spermine oxidase (SMO). APAO and SMO metabolized DESPM to EtSPD [K m(APAO) = 10 μM, k cat(APAO) = 1.1 s−1 and K m(SMO) = 28 μM, k cat(SMO) = 0.8 s−1, respectively], metabolized BnEtSPM to EtSPD [K m(APAO) = 0.9 μM, k cat(APAO) = 1.1 s−1 and K m(SMO) = 51 μM, k cat(SMO) = 0.4 s−1, respectively], and metabolized DBSPM to BnSPD [K m(APAO) = 5.4 μM, k cat(APAO) = 2.0 s−1 and K m(SMO) = 33 μM, k cat(SMO) = 0.3 s−1, respectively]. Interestingly, mono-alkylated spermine derivatives were metabolized by APAO and SMO to SPD [EtSPM K m(APAO) = 16 μM, k cat(APAO) = 1.5 s−1; K m(SMO) = 25 μM, k cat(SMO) = 8.2 s−1; BnSPM K m(APAO) = 6.0 μM, k cat(APAO) = 2.8 s−1; K m(SMO) = 19 μM, k cat(SMO) = 0.8 s−1, respectively]. Surprisingly, EtSPD [K m(APAO) = 37 μM, k cat(APAO) = 0.1 s−1; K m(SMO) = 48 μM, k cat(SMO) = 0.05 s−1] and BnSPD [K m(APAO) = 2.5 μM, k cat(APAO) = 3.5 s−1; K m(SMO) = 60 μM, k cat(SMO) = 0.54 s−1] were metabolized to SPD by both the oxidases. Furthermore, we studied the degradation of DESPM, BnEtSPM or DBSPM in the DU145 prostate carcinoma cell line. The same major metabolites EtSPD and/or BnSPD were detected both in the culture medium and intracellularly after 48 h of culture. Moreover, EtSPM and BnSPM were detected from cell samples. Present data shows that inducible SMO parallel with APAO could play an important role in polyamine based drug action, i.e. degradation of parent drug and its metabolites, having significant impact on efficiency of these drugs, and hence for the development of novel N-alkylated polyamine analogues.  相似文献   

18.
Light acclimatisation capabilities of Elodea nuttallii at nearly ambient DIC conditions were investigated by determining growth characteristics, main photosynthetic parameters and pigmentation of plants incubated at 5 different irradiances (10–146 μmol photons m−2 s−1). Positive net growth was observed under all light treatments tested. Maximum ratio root versus shoot (r:s) of 1.86 was achieved at medium irradiances (72–94 μmol photons m−2 s−1), whereas at low (10 μmol photons m−2 s−1) and high irradiances (146 μmol photons m−2 s−1) r:s was significantly lower (0.39 and 1.05, respectively). With respect to main photosynthetic parameters, an increase of light compensation points (E c), attended by decreasing ratios of light saturation points of photosynthesis (E k)/irradiance were observed. E c values were comparable to other low-light adapted macrophytes, which indicate that E. nuttallii can be regarded as a low-light adapted plant, under photorespiratory conditions. This was also confirmed by maximum E k values of just 73 μmol photons m−2 s−1. Further support was achieved from pigmentation and non-photochemical quenching (NPQ) data, both indicating rather limited acclimatisation ability at light treatments above 90 μmol photons m−2 s−1. These results are discussed with respect to the competitive abilities of E. nuttallii under nearly ambient (photorespiratory) DIC conditions, especially in dense stands and turbid phytoplankton-dominated waters.  相似文献   

19.
The neonicotinoid insecticide imidacloprid is an agonist on insect nicotinic acetylcholine receptors (nAChRs). We utilised fura-2-based calcium imaging to investigate the actions of imidacloprid on cultured GFP-tagged cholinergic neurons from the third instar larvae of the genetic model organism Drosophila melanogaster. We demonstrate dose-dependent increases in intracellular calcium ([Ca2+]i) in cholinergic neurons upon application of imidacloprid (10 nM–100 μM) that are blocked by nAChR antagonists mecamylamine (10 μM) and α-bungarotoxin (α-BTX, 1 μM). When compared to other (untagged) neurons, cholinergic neurons respond to lower concentrations of imidacloprid (10–100 nM) and exhibit larger amplitude responses to higher (1–100 μM) concentrations of imidacloprid. Although imidacloprid acts via nAChRs, increases in [Ca2+]i also involve voltage-gated calcium channels (VGCCs) in both groups of neurons. Thus, we demonstrate that cholinergic neurons express nAChRs that are highly sensitive to imidacloprid, and demonstrate a role for VGCCs in amplifying imidacloprid-induced increases in [Ca2+]i.  相似文献   

20.
Increased cytosolic calcium ([Ca2+] i ) and nitric oxide (NO) are suggested to be associated with apoptosis that is a main feature of many liver diseases and is characterized by biochemical and morphological features. We sought to investigate the events of increase in [Ca2+] i and endoplasmic reticulum (ER) calcium depletion by thapsigargin (TG), a selective inhibitor of sarco-ER-Ca2+-ATPases, in relation to NO production and apoptotic and necrotic markers of cell death in primary rat hepatocyte culture. Cultured hepatocytes were treated with TG (1 and 5 μmol/L) for 0–24 or 24–48 h. NO production and inducible NO synthase (iNOS) expression were determined as nitrite levels and by iNOS-specific antibody, respectively. Hepatocyte apoptosis was estimated by caspase-3 activity, cytosolic cytochrome c content and DNA fragmentation, and morphologically using Annexin-V/propidium iodide staining. Hepatocyte viability and mitochondrial activity were evaluated by ALT leakage and MTT test. Increasing basal [Ca2+] i by TG, NO production and apoptotic/necrotic parameters were altered in different ways, depending on TG concentration and incubation time. During 0–24 h, TG dose-dependently decreased iNOS-mediated spontaneous NO production and simultaneously enhanced hepatocyte apoptosis. In addition, TG 5 μmol/L produced secondary necrosis. During 24–48 h, TG dose-dependently enhanced basal NO production and rate of necrosis. TG 5 μmol/L further promoted mitochondrial damage as demonstrated by cytochrome c release. A selective iNOS inhibitor, aminoguanidine, suppressed TG-stimulated NO production and ALT leakage from hepatocytes after 24–48 h. Our data suggest that the extent of the [Ca2+] i increase and the modulation of NO production due to TG treatment contribute to hepatocyte apoptotic and/or necrotic events.  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