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 109 毫秒
1.
Two major indicators were used to access the degree of photorespiration in various photosynthetic types of Flaveria species (C3, C3-C4, C4-like, and C4): the O2 inhibition of photosynthesis measured above the O2 partial pressure which gives a maximum rate, and O2- and light-dependent whole-chain electron flow measured at the CO2 compensation point (). The optimum level of O2 for maximum photosynthetic rates under atmospheric levels of CO2 (34 Pa) was lower in C3 and C3-C4 species (ca. 2 kPa) than in C4-like and C4 species (ca. 9 kPa). Increasing O2 partial pressures from the optimum for photosynthesis up to normal atmospheric levels (ca. 20 kPa) caused an inhibition of photosynthesis which was more severe under lower CO2. This inhibition was calculated as the O2 inhibition index (A, the percentage inhibition of photosynthesis per kPa increase in O2). From measurements of 18 Flaveria species at atmospheric CO2, the A values decreased from C3 (1.9–2.1) to C3-C4 (1.2–1.6), C4-like (0.6–0.8) and C4 species (0.3–0.4), indicating a progressive decrease in apparent photorespiration in this series. With increasing irradiance at under atmospheric levels of O2, and increasing O2 partial pressure at 300 mol quanta·m–2·s–1, there was a similar increase in the rate of O2 evolution associated with whole-chain electron flow (Jo 2, calculated from chlorophyll fluorescence analysis) in the C3 and C3-C4 species compared to a much lower rate in the C4-like and C4 species. The results indicate that there is substantial O2-dependent electron flow in C3 and C3-C4 species, reflecting a high level of photorespiration compared to that in C4-like and C4 species. Consistent with these results, there was a significant decrease in from C3 (6–6.2 Pa) to C3-C4 (1.0–3.0 Pa), to C4-like and C4 species (0.3–0.8 Pa), indicating a progressive decrease in apparent photorespiration. However, C3 and C3-C4 species examined had high intrinsic levels of photorespiration with the latter maintaining low apparent rates of photorespiration and lower values, primarily by refixing photorespired CO2. The C4-like and C4 Flaveria species had low, but measurable, levels of photorespiration via selective localization of ribulose-1,5-bisphosphate carboxylase in bundle sheath cells and operation of a CO2 pump via the C4 pathway.Abbreviations and Symbols A CO2 assimilation rate - CE carboxylation efficiency - Ci intercellular CO2 partial pressure - Ia absorbed PPFD - Jo 2 oxygen evolution from PSII - PPFD photosynthetic photon flux density (mol · m–2· s–1) - Rubisco ribulose-1,5-bisphosphate carboxylase/oxygenase - RuBP ribulose-1,5-bisphosphate - VPD water-vapor pressure difference between the leaf and atmospheric air - CO2 compensation point - CO 2 quantum yield of CO2 assimilation - PSII quantum yield of photosystem II - A O2 inhibition index for photosynthesis (percentage inhibition of photosynthesis per kPa increase in O2) This research was supported by the National Science Foundation Grant IBN 9317756 and Equipment (Grant DMB-8515521 and DOE/USDA/NSF Triagency Plnat Biochemistry Research Training Grant Program.  相似文献   

2.
Summary Acyl-CoA: lysophosphatidylcholine acyltransferase (LPCAT) (EC 2.3.1.23) activity was assayed in liver microsomes from rainbow trout,Salmo gairdneri, acclimated to 5°C and 20°C to assess its contribution to the temperature-induced restructuring of phospholipid acyl chain composition. The synthesis of phosphatidylcholine (PC) (from lyso-PC) was threefold the synthesis of phosphatidylethanolamine (PE) (from lyso-PE) under similar assay conditions. LPCAT activity (i) displayed an absolute requirement for lysophosphatidylcholine (LPC) and was enhanced by the presence of ATP, MgCl2 and CoA (which reduced the impact of endogenous acyl-CoA hydrolase activity by regenerating the acyl-CoA substrate) in the assay medium; (ii) remained linear with time up to 30 min; and (iii) increased linearly with microsomal protein concentration up to 0.2 mg/ml for the 20°C assay and 0.4 mg/ml for the 5°C assay. There was no difference in Km or Vmax values due to the acclimation history of the fish, but there were obvious differences due to assay temperature. The apparent Km values for LPC were 58.54±7.24 M and 12.26±2.14 M when assayed at 5°C and 20°C respectively; values for oleoyl-CoA were 9.11±0.78 M and 1.23±0.25 M under the same assay conditions. Activity was 1.99±0.31 nmol min–1 mg protein–1 when assayed at 5°C, and 3.8±0.45 nmol min–1 mg protein–1 when assayed at 20°C. These findings indicate that adjustments in the activity of LPCAT play no significant role in the temperature-induced restructuring of PC molecular species composition. However, the marked temperature dependence of the Km values for LPC and oleoyl CoA suggest that patterns of fatty acid incorporation (i.e. substrate preference) may vary with assay temperature, and in this way LPCAT could contribute to the restructuring response.Abbreviations PC phosphatidylcholine - PE phosphatidylethanolamine - LPCAT acyl-CoA: lysophosphatidylcholine acyltransferase - LPEAT acyl-CoA: lysophosphatidylethanolamine acyltransferase - LPC 1-palmitoyl,2-lysophosphatidylcholine  相似文献   

3.
Summary Net annual productivity and annual carbon budgets were determined for populations of Littorella uniflora var. americana and Isoetes macrospora in a mesotrophic and oligotrophic lake in northern Wisconsin, to assess the contribution of Crassulacean Acid Metabolism (CAM) to annual productivity of the species in their natural environment. Nocturnal carbon accumulation (CAM), daytime uptake of external CO2 via the C3 mechanism, and refixation of endogenously generated CO2 from daytime respiration were the sources of carbon income. CAM activity as diurnal acid rhythms reached maxima of 89 to 182 eq·g-1 leaf fresh weight for the various populations.Maximum rates of daytime 14C uptake ranged from 0.56 to 1.46 mg C·g-1 leaf dry wt.·h-1 for the study populations. Refixation of daytime respired CO2 averaged 37% for the four populations. Carbon loss was due largely to dark respiration, during the day and night. Nocturnal carbon accumulation, daytime CO2 uptake and 24-h dark respiration were of similar magnitude, indicating dark respiration was equivalent to 50% of gross photosynthesis.Net annual production was measured for each population by following leaf turnover. Turnover rates for the Littorella populations were 1.56 and 1.72·yr-1, and for the Isoetes populations, 0.85 and 1.00·yr-1. Measured net annual productivity and calculated net annual productivity (based on carbon exchange) agreed within an average of 12% for the four populations. While CAM activity was greater for the more productive population of each species, the results suggest that the contribution of CAM to annual productivity is greater for the less productive population of each species. CAM contributed 45 to 55% of the annual carbon gain for the study populations.  相似文献   

4.
Short- and medium-chain acyl coenzyme A (acyl-CoA) synthetases catalyze the formation of acyl-CoA from an acyl substrate, ATP, and CoA. These enzymes catalyze mechanistically similar two-step reactions that proceed through an enzyme-bound acyl-AMP intermediate. Here we describe the characterization of a member of this enzyme family from the methane-producing archaeon Methanosarcina acetivorans. This enzyme, a medium-chain acyl-CoA synthetase designated MacsMa, utilizes 2-methylbutyrate as its preferred substrate for acyl-CoA synthesis but cannot utilize acetate and thus cannot catalyze the first step of acetoclastic methanogenesis in M. acetivorans. When propionate or other less favorable acyl substrates, such as butyrate, 2-methylpropionate, or 2-methylvalerate, were utilized, the acyl-CoA was not produced or was produced at reduced levels. Instead, acyl-AMP and PPi were released in the absence of CoA, whereas in the presence of CoA, the intermediate was broken down into AMP and the acyl substrate, which were released along with PPi. These results suggest that although acyl-CoA synthetases may have the ability to utilize a broad range of substrates for the acyl-adenylate-forming first step of the reaction, the intermediate may not be suitable for the thioester-forming second step. The MacsMa structure has revealed the putative acyl substrate- and CoA-binding pockets. Six residues proposed to form the acyl substrate-binding pocket, Lys256, Cys298, Gly351, Trp259, Trp237, and Trp254, were targeted for alteration. Characterization of the enzyme variants indicates that these six residues are critical in acyl substrate binding and catalysis, and even conservative alterations significantly reduced the catalytic ability of the enzyme.AMP-forming acetyl coenzyme A (acetyl-CoA) synthetase (Acs; acetate:CoA ligase [AMP forming], EC 6.2.1.1), which catalyzes the activation of acetate to acetyl-CoA, is a member of the acyl-adenylate-forming enzyme superfamily (8), which consists of acyl- and aryl-CoA ligases, nonribosomal peptide synthetases that mediate the synthesis of peptide and polyketide secondary metabolites, such as gramicidin and tyrocidine, and the enzymes firefly luciferase and α-aminoadipate reductase. Although these enzymes share the property of forming an acyl-adenylate intermediate and are structurally related, they share limited sequence homology and catalyze unrelated reactions in which the intermediate serves different functions for different members of this enzyme family.A two-step mechanism for Acs (equations 1 and 2) in which the reaction proceeds through an acetyl-AMP intermediate has been proposed based on evidence including detection of an enzyme-bound acetyl-AMP (2-4, 38): (1) (2)In the CoA-dependent first step of the reaction, an enzyme-bound acetyl-AMP intermediate is formed from acetate and ATP, and inorganic pyrophosphate (PPi) is released. In the second step, the acetyl group is transferred to the sulfhydryl group of CoA and AMP is released. Other short- (Sacs) and medium-chain acyl-CoA synthetases (Macs) follow a similar reaction mechanism using acyl substrates other than acetate (8, 15).In the 2.3-Å structure of trimeric Saccharomyces cerevisiae Acs1 in a binary complex with AMP (19), the C-terminal domain is positioned away from the N-terminal domain in a conformation for catalysis of the first step of the reaction (equation 1). The 1.75-Å structure of the monomeric Salmonella enterica Acs (AcsSe) (13) in complex with both CoA and adenosine-5′-propylphosphate, an inhibitor of the related propionyl-CoA synthetase (12, 15), which mimics the acetyl-adenylate intermediate, reveals that the C-terminal domain of Acs is rotated approximately 140° toward the N-terminal domain to form the complete active site for catalysis of the second half-reaction (equation 2). In this orientation, the CoA thiol is properly positioned for nucleophilic attack on the acetyl group. In structure/function studies of 4-chlorobenzoate:CoA ligase (CBAL), a distant member of the acyl- and aryl-CoA synthetase subfamily of the acyl-adenylate-forming enzyme superfamily, Wu et al. (39) and Reger et al. (28) provide evidence that PPi produced in the first step of the reaction dissociates from the enzyme before the switch from the first conformation to the second conformation required for CoA binding and catalysis of the second step of the reaction.Acs and Sacs/Macs are widespread in all three domains of life and play a key role in archaea, as suggested by the finding that several thermophilic archaea have multiple open reading frames (ORFs) (up to seven) that encode putative Sacs or Macs (33). The chemolithoautotrophic methanoarchaeon Methanothermobacter thermautotrophicus has two ORFs with high identity to Acs and a third ORF that is likely to encode a Macs. M. thermautotrophicus Acs1 (Acs1Mt) has been biochemically and kinetically characterized, has been shown to have a strong preference for acetate as the acyl substrate, and can also utilize propionate but not butyrate (16, 17).Methanosarcina and Methanosaeta are the only two methanoarchaea isolated that are able to utilize acetate as substrate for methane production. Unlike Methanosaeta species, which utilize Acs for catalyzing the first step of methanogenesis (18, 34), Methanosarcina species employ the acetate kinase-phosphotransacetylase pathway for activation of acetate to acetyl-CoA, and an Acs activity has not been observed in Methanosarcina (1, 23, 30, 32). Surprisingly, an Acs-related sequence was identified in the Methanosarcina acetivorans genome. Here we describe the kinetic characterization this enzyme, designated MacsMa, and show that it utilizes longer acyl substrates than Acs. The preferred acyl substrate was shown to be 2-methylbutyrate, and 2-methylbutyryl-CoA, AMP, and PPi were the products of the reaction, as expected. However, when propionate was used as the acyl substrate, propionyl-CoA was not produced. Instead, in the absence of CoA, propionyl-AMP and PPi were released, whereas in the presence of CoA, the propionyl-AMP intermediate was broken down into AMP and propionate and released along with PPi. Intermediate results were obtained with other acyl substrates, with both acyl-CoA and acyl-AMP production observed.The 2.1-Å crystal structure of MacsMa (31), determined in the absence of any substrate, revealed the enzyme to be in a conformation similar to that of the S. enterica Acs (13) with respect to the position of the C-terminal domain. Through inspection of the MacsMa structure and alignment of Acs, Sacs, and Macs sequences, we identified six residues that form the putative acyl substrate-binding pocket. Individual alterations at these residues dramatically diminished enzyme activity and indicate that the acyl substrate-binding pocket of MacsMa has a very precise architecture that cannot be perturbed.  相似文献   

5.
After inhibiting ion and water transport with 10-6 mol·l-1 serotonin and 10-6 mol·l-1 methacholine, a muscarinic agonist of acetylcholine, 10-5 mol·l-1 (±)noradrenaline restored the serosa-negative transepithelial potential difference and short-circuit current in a step-like manner, accompanied by an increase in water absorption across the seawater eel intestine. Such recovery by noradrenalin was not obtained after pretreatment with 10-7 mol·l-1 eel atrial natriuretic peptide. This means that the inhibitory mechanisms of serotonin and acetylcholine are different from those of atrial natriuretic peptide. Similarly, 10-7 mol·l-1 clonidine and guanabenz (2-agonists) also reversed the inhibitory action of serotonin and methacholine, but 10-7 mol·l-1 phenylephrine (1-agonists) and 10-7 mol·l-1 isoproterenol (-agonist) did not antagonize serotonin and methacholine actions. Further, the enhancement by 10-5 mol·l-1 noradrenalin was blocked by 10-4 mol·l-1 yohimbine (2-agonists) and 10-4 mol·l-1 prazosin (1-agonists), but not by 10-4 mol·l-1 propranolol (-antagonist). Although relatively high dosage is required to obtain a significant effect, and discrimination between 1- and 2- is not successful in the present study, these results suggest that noradrenalin acts on an -type receptor. The -type receptor may exist on the enterocytes, since the effects of noradrenalin are observed even in the presence of 10-6 mol·l-1 tetrodotoxin. Interestingly, the tissue resistance also increased in parallel with increase in the short-circuit current after treatment with noradrenalin in the posterior part of the seawater eel intestine.Abbreviations ACh acetylcholine - 5-HT serotonin - eANP eel atrial natriuretic peptide - I sc short-circuit current - MCh methacholine - NA noradrenalin - PD transepithelial potential difference - R t tissue resistance - TTX tetrodotoxin - VIP vasoactive intestinal peptide  相似文献   

6.
Summary Penicillin G recovery is investigated in a continuous flotation column in the presence of different collectors which form a complex with penicillin. The performance of the penicillin recovery was investigated as a function of the mole ratio () of collector-to-penicillin and the aliphatic chain length of the collector. At =1 and low penicillin concentrations (e.g., 20 mg·1-1), high foam liquid concentrations (680 mg·l-1), low residue concentrations (12 mg·l-1) and high penicillin separation (56) can be attained. At =4 the separation increases to 150, and 95% of the penicillin can be recovered.Symbols Cp penicillin concentration in feed (mg·l-1) - CR penicillin concentration in outlet liquid (mg·l-1) - CS penicillin concentration in foam liquid (mg·l-1) - CS/CP penicillin enrichment (-) - CS/CR penicillin separation (-) - % Pen in S penicillin yield in foam liquid (%) - VV}S foam liquid volume flow (ml·min-1) - VV}P feed (ml·min-1) - VVN 2 nitrogen flow rate (ml·s-1) - temperature  相似文献   

7.
8.
Summary Regulation of the paracellular pathway in rabbit distal colon by the hormone aldosterone was investigated in vitro in Ussing chambers by means of transepithelial and microelectrode techniques. To evaluate the cellular and paracellular resistances an equivalent circuit analysis was used. For the analysis the apical membrane resistance was altered using the antibiotic nystatin. Under control conditions two groups of epithelia were found, each clearly dependent on the light: dark regime. Low-transporting epithelia (LT) were observed in the morning and high-transporting epithelia (HT) in the afternoon. Na+ transport was about 3-fold higher in HT than in LT epithelia. Incubating epithelia of both groups with 0.1 mol·1-1 aldosterone on the serosal side nearly doubled in LT epithelia the short circuit current and transepithelial voltage but the transepithelial resistance was not influenced. Maximal values were reached after 4–5 h of aldosterone treatment. In HT epithelia due to the effect of aldosterone all three transepithelial parameters remained constant over time. Evaluation of the paracellular resistance revealed a significant increase after aldosterone stimulation in both epithelial groups. This increase suggests that tight junctions might have been regulated by aldosterone. The hormonal effect on electrolyte transport was also dependent on the physiological state of the rabbit colon. Since net Na+ absorption in distal colon is, in addition to transcellular absorption capacity, also dependent on the permeability of the paracellular pathway, the regulation of tight junctions by aldosterone may be a potent mechanism for improving Na+ absorption during hormone-stimulated ion transport.Abbreviations V t transepithelial potential difference (mV) - R t transepithelial resistance (·cm2) - G t transepithelial conductance (mS·cm-2) - Isc calculated short circuit current (A·cm-2) - V a apical membrane potential difference (mV) - V bl basolateral membrane potential difference (mV) - voltage divider ratio - R a apical membrane resistance (·cm2) - R bl basolateral membrane resistance (·cm2) - R c cellular resistance ( of apical and basolateral resistance) (·cm2) - R p resistance of the paracellular pathway (·cm2) - G a apical membrane conductance (mS·cm-2) - G bl basolateral membrane conductance (mS·cm-2) - G p paracellular conductance (mS·cm-2) - G t transepithelial conductance (mS·cm-2) - HT contr high transporting control epithelia - LT contr low transporting control epithelia - HT aldo aldosterone incubated high transporting epithelia - LT aldo aldosterone incubated low transporting epithelia  相似文献   

9.
The dorsal skin of the leech Hirudo medicinalis was used for electrophysiological measurements performed in Ussing chambers. The leech skin is a tight epithelium (transepithelial resistance = 10.5±0.5 k· cm-2) with an initial short-circuit current of 29.0±2.9 A·cm-2. Removal of Na+ from the apical bath medium reduced short-circuit current about 55%. Ouabain (50mol·l-1) added to the basolateral solution, depressed the short-circuit current completely. The Na+ current saturated at a concentration of 90 mmol Na+·l-1 in the apical solution (K M=11.2±1.8 mmol·l-1). Amiloride (100 mol·l-1) on the apical side inhibited ca. 40% of the Na+ current and indicated the presence of Na+ channels. The dependence of Na+ current on the amiloride concentration followed Michaclis-Menten kinetics (K i=2.9±0.4 mol·l-1). The amiloride analogue benzamil had a higher affinity to the Na+ channel (K i=0.7±0.2 mol·l-1). Thus, Na+ channels in leech integument are less sensitive to amiloride than channels known from vertebrate epithelia. With 20 mmol Na+·l-1 in the mucosal solution the tissue showed an optimum amiloride-inhibitable current, and the amiloride-sensitive current under this condition was 86.8±2.3% of total short-circuit current. Higher Na+ concentrations lead to a decrease in amiloride-blockade short-circuit current. Sitmulation of the tissue with cyclic adenosine monophosphate (100 mol·l-1) and isobutylmethylxanthine (1 mmol·l-1) nearly doubled short-circuit current and increased amiloride-sensitive Na+ currents by 50%. By current fluctuation analysis we estimated single Na+ channel current (2.7±0.9 pA) and Na+ channel density (3.6±0.6 channels·m-2) under control conditions. After cyclic adenosine monophosphate stimulation Na+ channel density increased to 5.4±1.1 channels·m-2, whereas single Na+ channel current showed no significant change (1.9±0.2 pA). These data present a detailed investigation of an invertebrate epithelial Na+ channel, and show the similarities and differences to vertebrate Na+ channels. Whereas the channel properties are different from the classical vertebrate Na+ channel, the regulation by cyclic adenosine monophosphate seems similar. Stimulation of Na+ uptake by cyclic adenosine monophosphate is mediated by an increasing number of Na+ channels.Abbreviations slope of the background noise component - ADH antidiuretic hormone - cAMP cyclic adenosine monophosphate - f frequency - f c coner frequency of the Lorentzian noise component - Hepes N-hydroxyethylpiperazine-N-ethanesulphonic acid - BMX isobutyl-methylxanthine - i Na single Na+ channel current - I Na max, maximal inhibitable Na+ current - I SC short circuit current - K i half maximal blocker concentration - K M Michaelis constandard error of the mean - S (f) power density of the Lorentzian noise component - S 0 plateau value of the Lorentzian noise component - TMA tetramethylammonium - Trizma TRIS-hydroxymethyl-amino-methane - V max maximal reaction velocity - V T transepithelial potential - K half maximal blocker concentration  相似文献   

10.
The activities of three acylation systems for 1-alkenylglycerophosphoethanolamine (1-alkenyl-GPE), 1-acyl-GPE and 1-acylglycerophosphocholine (1-acyl-GPC) were compared in rat brain microsomes and the acyl selectivity of each system was clarified. The rate of CoA-independent transacylation of 1-[3H]alkenyl-GPE (approx. 4.5 nmol/10 min per mg protein) was about twice as high as in the case of 1-[3H]acyl-GPE and 1-[14C]acyl-GPC. On the other hand, the rates of CoA-dependent transacylation and CoA + ATP-dependent acylation (acylation of free fatty acids by acyl-CoA synthetase and acyl-CoA acyltransferase) of lysophospholipids were in the order 1-acyl-GPC greater than 1-acyl-GPE much greater than 1-alkenyl-GPE. HPLC analysis of newly synthesized molecular species revealed that the CoA-independent transacylation system exclusively esterified docosahexaenoate and arachidonate, regardless of the lysophospholipid class. The CoA-dependent transacylation and CoA + ATP-dependent acylation systems were almost the same with respect to the selectivities for unsaturated fatty acids when the same acceptor lysophospholipid was used, but some distinctive acyl selectivities were observed with different acceptor lysophospholipids. 1-Alkenyl-GPE selectively acquired only oleate in these two systems. 1-Acyl-GPE and 1-acyl-GPC showed selectivities for both arachidonate and oleate. In addition, an appreciable amount of palmitate was transferred to 1-acyl-GPC, not to 1-acyl-GPE, in CoA- or CoA + ATP-dependent manner. The acylation of exogenously added acyl-CoA revealed that the acyl selectivities of the CoA-dependent transacylation and CoA + ATP-dependent acylation systems may be mainly governed through the selective action of acyl-CoA acyltransferase. The preferential utilization of oleoyl-CoA by all acceptors and the different utilization of arachidonoyl-CoA between alkenyl and acyllysophospholipids indicated that there might be two distinct acyl-CoA:lysophospholipid acyltransferases that discriminate between oleoyl-CoA and arachidonoyl-CoA, respectively. Our present results clearly show that all three microsomal acylation systems can be active in the reacylation of three major brain glycerophospholipids and that the higher contribution of the CoA-independent system in the reacylation of ethanolamine glycerophospholipids, especially alkenylacyl-GPE, may tend to enrich docosahexaenoate in these phospholipids, as compared with in the case of diacyl-GPC.  相似文献   

11.
Summary Zymomonas mobilis is able to convert glucose and fructose to gluconic acid and sorbitol. The enzyme, glucose-fructose oxidoreductase, catalysing the intermolecular oxidation-reduction of glucose and fructose to gluconolactone and sorbitol, was formed in high amounts [1.4 units (U)·mg-1] when Z. mobilis was grown in chemostats with glucose as the only carbon source under non-carbon-limiting conditions. The activity of a gluconolactone-hydrolysing lactonase was constant at 0.2 U·mg-1. Using glucose-grown cells for the conversion of equimolar fructose and glucose mixtures up to 60% (w/v), a maximum product concentration of only 240 g·1-1 of sorbitol was found. The gluconic acid accumulated was further metabolized to ethanol. After permeabilizing the cells using cationic detergents, maximum sorbitol and gluconic acid concentrations of 295 g·1-1 each were reached; no ethanol production occurred. In a continuous process with -carrageenan-immobilized and polyethylenimin-hardened, permeabilized cells no significant decrease in the conversion yield was observed after 75 days. The specific production rates for a high yield conversion ( > 98%) in a continuous two-stage process were 0.19 g·g-1·h-1 for sorbitol and 0.21 g·g-1·h-1 for gluconic acid, respectively. For the sugar conversion of cetyltrimethylammonium bromide-treated -carrageenan-immobilized cells a V max of 1.7 g·g-1·h-1 for sorbitol production and a K m of 77.2 g·1-1 were determinedOffprint requests to: B. Rehr  相似文献   

12.
[14C]Methylamine influx intoPisum sativum L. cv. Feltham First seedlings showed Michaelis-Menten-type kinetics with apparentV max=49.2 mol·g-1 FW·h-1 and apparentK m=0.51 mM. The competitive interactions between ammonium and methylamine were most obvious when biphasic kinetics were assumed with saturation of the first phase at 0.05 mM. The inhibitor constant for ammonium (K i)=0.027 mM. When [14C]methylamine was used in trace amounts with ammonium added as substrate, the influx of tracer showed Michaelis-Menten-type kinetics with apparentV max=3.46 mol·g-1 FW·h-1 and apparentK m=0.15 mM. The initial rate of net ammonium uptake corresponded with that found when [14C]methylamine was used to trace ammonium influx. The latter was also stimulated by high pHo and inhibited by nitrate. Ammonium pretreatment±methionine sulphoximine or glutamine pretreatment of the seedlings inhibited subsequent [14C]methylamine influx, while methylamine or asparagine pretreatment stimulated [14C]methylamine influx. There was also a stimulatory effect of prior inoculation withRhizobium. The results are discussed in terms of current models for the regulation of ammonium uptake in plants.  相似文献   

13.
Desulfovibrio vulgaris (Marburg) was grown on H2 plus sulfate and H2 plus thiosulfate as the sole energy sources and acetate plus CO2 as the sole carbon sources. Conditions are described under which the bacteria grew exponentially. Specific growth rates () and molar growth yields (Y) at different pH were determined. and Y were found to be strongly dependent on the pH. Highest growth rates and molar growth yields were observed for growth on H2 plus sulfate at pH 6.5 (=0.15h-1; Y SO 4 2- =8.3g·mol-1) and for growth on H2 plus thiosulfate at pH 6.8 (=0.21h-1; Y S 2O 3 2 =16.9g·mol-1).The growth yields were found to increase with increasing growth rates: plots of 1/Y versus 1/ were linear. Via extrapolation to infinite growth rates a Y SO4 2- /max of 12.2g·mol-1 and a YS2O 3 2- /max of 33.5g·mol-1 was obtained.The growth yield data are interpred to indicate that dissimilatory sulfate reduction to sulfide is associated with a net synthesis of 1 mol of ATP and that near to 3 mol of ATP are formed during dissimilatory sulfite reduction to sulfide.  相似文献   

14.
The cyanobacterium Oscillatoria agardhii was subjected to changes in irradiance and to changes in light period. During transient states parameters as growth rate, pigment contents, photosynthetic activities and pool sizes of carbohydrate and proteins were followed. The changes in pigments and photosynthesis were similar for irradiance transitions and transitions in light period length. Carbohydrates served for the supply of carbon and energy during adaptation to low light conditions until a basal level of 125 g · mg dry wt-1 was reached. After transfer to high light conditions excess carbon fixation led to the storage of carbohydrate reserve polymers up to 600 g · mg dry wt-1. During adaptation to longer light periods cells showed an overcapacity for carbohydrate accumulation even in the presence of a high carbohydrate content at the start of the light period. A model for the feed back repression of photosynthesis related to carbohydrate accumulation was presented. In all cases protein synthesis was directly maximized under the given conditions. Growth rate defined as specific rate of change in carbon showed the fastest response after a shift in light conditions. It was concluded that adaptation of O. agardhii to changes in light conditions was directed to the optimization of growth. The observation that carbohydrate is used to supply carbon and/or energy during adaptation leads to the conclusion that changes on survival in low light depend on carbohydrate level, the efficiency of its conversion in cell material and the maintenance requirements. Such a survival strategy enables cyanobacteria to cope succesfully with light limiting conditions.Abbreviations HL high irradiance Em-2s-1 - LL low irradiance Em-2s-1 - L/D light-dark cycle h - specific growth rate h-1 - e specific maintenance coefficient h-1 - max maximal specific growth rate h-1 - c specific rate of change of carbon h-1 - protein specific rate of change of protein h-1 - Chl a chlorophyll a - CPC C-phycocyanin - dry wt dry weight mg - light utilization efficiency nmol O2 · mg dry wt-1 · min-1/Em-1 s-1 - P max photosynthetic capacity nmol O2 · mg dry wt-1 · min-1 - PSI photosystem I - PS II photosystem II - RC II reaction center of PS II - PQ plastoquinone  相似文献   

15.
In C4 grasses belonging to the NADP-malic enzyme-type subgroup, malate is considered to be the predominant C4 acid metabolized during C4 photosynthesis, and the bundle sheath cell chloroplasts contain very little photosystem-II (PSII) activity. The present studies showed that Flaveria bidentis (L.), an NADP-malic enzyme-type C4 dicotyledon, had substantial PSII activity in bundle sheath cells and that malate and aspartate apparently contributed about equally to the transfer of CO2 to bundle sheath cells. Preparations of bundle sheath cells and chloroplasts isolated from these cells evolved O2 at rates between 1.5 and 2 mol · min–1 · mg–1 chlorophyll (Chl) in the light in response to adding either 3-phosphoglycerate plus HCO 3 or aspartate plus 2-oxoglutarate. Rates of more than 2 mol O2 · min–1 · mg–1 Chl were recorded for cells provided with both sets of these substrates. With bundle sheath cell preparations the maximum rates of light-dependent CO2 fixation and malate decarboxylation to pyruvate recorded were about 1.7 mol · min–1 · mg–1 Chl. Compared with NADP-malic enzyme-type grass species, F. bidentis bundle sheath cells contained much higher activities of NADP-malate dehydrogenase and of aspartate and alanine aminotransferases. Time-course and pulse-chase studies following the kinetics of radiolabelling of the C-4 carboxyl of C4 acids from 14CO2 indicated that the photosynthetically active pool of malate was about twice the size of the aspartate pool. However, there was strong evidence for a rapid flux of carbon through both these pools. Possible routes of aspartate metabolism and the relationship between this metabolism and PSII activity in bundle sheath cells are considered.Abbreviations DHAP dihydroxyacetone phosphate - NADP-ME(-type) NADP-malic enzyme (type) - NADP-MDH NADP-malate dehydrogenase - OAA oxaloacetic acid - 2-OG 2-oxoglutarate - PEP phosphoenolpyruvate - PGA 3-phosphoglycerate - Pi orthophosphate - Ru5P ribulose 5-phosphate  相似文献   

16.
Summary A new, fast method is described to determine kLa either off-line, or on-line during animal-cell cultivation. Since it does not need the equilibrium concentration of oxygen in the liquid phase (C*), it is not required to await a new steady state. Furthermore, the results do not depend on the calibration value of the dissolved-oxygen probe. The method yielded accurate values for kLa, both for an oxygen-consuming and a non-consuming system.Nomenclature C L Dissolved-oxygen concentration [mol·m-3] - C * C L in equilibrium with the oxygen concentration in the gas phase [mol·m-3] - C L, Equilibrium oxygen concentration at stationary conditions [mol·m-3] - kLa Volumetric oxygen transfer coefficient [s-1] - r Specific oxygen consumption of biomass [mol·cell-1·s-1] - X Cell concentration [cells·m-3] - t Time [s] - Noise of dissolved-oxygen probe [mol·m-3] - Absolute error of kLa-measurement [s-1]  相似文献   

17.
The closo- and nido-carborane-diphenylphosphine complexes [Hg2{1,2-(PPh2)2-1,2-C2B10H10}2(μ-Cl2)2(μ-HgCl2)3]·2CH2Cl2 (1) and [HgCl(PPh3){7,8-(PPh2)2-7,8-C2B9H10}] (2) have been synthesized and characterized by elemental analysis, FT-IR and X-ray structure determination. The X-ray structure analysis for these two complexes showed that the carborane cage ligand was coordinated bidentately to the Hg(II) center through its two phosphorus atoms. The coordination geometry of the mercury atom complexed by P2Cl2 unit in complex 1 or P3Cl unit in complex 2 was a distorted tetrahedron, while the mercury atom in complex 2 coordinated to six Cl atoms was a slightly distorted octahedron. X-ray analysis reveals that the complex 1 forms a 1D chain coordination polymer via bridged Hg-Cl bonds. For complex 2, it displays a 3D network constructed by the C-H···Cl hydrogen bonds and C-H···H-B dihydrogen bonds.  相似文献   

18.
The fluorescence of the voltage sensitive dye, diS-C3-(5), has been analyzed by means of synchronous excitation spectroscopy. Using this rather rare fluorescence technique we have been able to distinguish between the slightly shifted spectra of diS-C3-(5) fluorescence from cells and from the supernatant. It has been found that diS-C3-(5) fluorescence in the supernatant can be selectively monitored at exc = 630 nm and em= 650 nm, while the cell associated fluorescence can be observed at exc= 690 nm and em = 710 nm. A modified theory for the diSC3-(5) fluorescence response to the membrane potential is presented, according to which a linear relationship exists between the logarithmic increment of the dye fluorescence intensity in the supernatant, In I/I°, and the underlying change in the plasma membrane potential, p=pp. The theory has been tested on human myeloid leukemia cells (line ML-1) in which membrane potential changes were induced by valinomycin clamping in various K+ gradients. It has been demonstrated that the membrane potential change, p,can be measured on an absolute scale. Offprint requests to: J. Plasek  相似文献   

19.
Desulfotomaculum acetoxidans has been proposed to oxidize acetate to CO2 via an oxidative acetyl-CoA/carbon monoxide dehydrogenase pathway rather than via the citric acid cycle. We report here the presence of the enzyme activities required for the operation of the novel pathway. In cell extracts the following activities were found (values in brackets=specific activities and apparent K m; 1 U·mg-1=1 mol·min-1·mg protein-1 at 37°C): Acetate kinase (6.3 U·mg-1; 2 mM acetate; 2.4 mM ATP); phosphate acetyltransferase (60 U·mg-1, 0.4 mM acetylphosphate; 0.1 mM CoA); carbon monoxide dehydrogenase (29 U·mg-1; 13% carbon monoxide; 1.3 mM methyl viologen); 5,10-methylenetetrahydrofolate reductase (3 U·mg-1, 0.06 mM CH3–FH4); methylenetetrahydrofolate dehydrogenase (3.6 U·mg-1, 0.9 mM NAD, 0.1 mM CH2=FH4); methenyltetrahydrofolate cyclohydrolase (0.3 U·mg-1); formyltetrahydrofolate synthetase (3 U·mg-1, 1.4 mM FH4, 0.4 mM ATP, 13 mM formate); and formate dehydrogenase (10 U·mg-1, 0.4 mM formate, 0.5 mM NAD). The specific activities are sufficient to account for the in vivo acetate oxidation rate of 0.26 U·mg-1.Non-standard abbreviations FH4 Tetrahydrofolate - CHO-FH4 N10-formyltetrahydrofolate - CHFH4 N5,N10-methenyltetrahydrofolate - CH2=FH4 N5,N10-methylenetetrahydrofolate - CH3–FH4 N5-methyltetrahydrofolate - MOPS morpholinopropane sulfonic acid - DTT d,l-1,4-dithiothreitol - TRIS tris-(hydroxymethyl)-aminomethane - Ap5A p1,P5-di(adenosine-5)pentaphosphate - MV methyl viologen  相似文献   

20.
The native lipid composition and the capacity of cell-free extracts to biosynthesize acyl lipids in vitro were determined for the first time using the recently reported microspore-derived (MD) embryo system from the Brassica campestris low erucic acid line BC-2 (Baillie et al. 1992). The total lipid fraction isolated from midcotyledonary stage MD embryos (21 days in culture) was composed primarily of triacylglycerol (76%) with an acyl composition quite similar to that of mature BC-2 seed. When incubated in the presence of glycerol-3-phosphate, 14C 181-CoA, and reducing equivalents, homogenates prepared from 21-day cultured MD embryos were able to biosynthesize glycerolipids via the Kennedy pathway. The maximum in vitro rate of triacylglycerol biosynthesis could more than account for the known rate of lipid accumulation in vivo. The homogenate catalyzed the desaturation of 181 to 182 and to a lesser extent, 183. The newly-synthesized polyunsaturated fatty acids initially accumulated in the polar lipid fraction (primarily phosphatidic acid and phosphatidylcholine) but began to appear in the triacylglycerol fraction after longer incubation periods. As expected for a low erucic acid cultivar, homogenates of MD embryos from the BC-2 line were incapable of biosynthesizing very long chain monounsaturated fatty acyl moieties (201 and 221) from 181-CoA in vitro. Nonetheless, embryo extracts were still capable of incorporating these fatty acyl moieties into triacylglycerols when supplied with 14C 201-CoA or 14C 221-CoA. Collectively, the data suggest that developing BC-2 MD embryos constitute an excellent experimental system for studying pathways for glycerolipid bioassembly and the manipulation of this process in B. campestris.Abbreviations CPT sn-1,2-diacylglycerol cholinephosphotransferase - DAG diacylglycerol - DGAT diacylglycerol acyltransferase - DGDG digalactosyldiacylglycerol - G-3-P glycerol-3-phosphate - G-3-PAT glycerol-3-phosphate acyltransferase - LPA lyso-phosphatidic acid - LPAT lyso-phosphatidic acid acyltransferase - LPC lyso-phosphatidylcholine - LPCAT acyl-CoA: lyso-phosphatidylcholine acyltransferase - LPE lyso-phosphatidylethanolamine - MGDG monogalactosyldiacylglycerol - PA phosphatidic acid - PA Phosphatase, phosphatidic acid phosphatase - PC phosphatidylcholine - PE phosphatidylethanolamine - PG phosphatidylglycerol - TAG triacylglycerol - 181-CoA oleoyl-Coenzyme A - 181 oleic acid, cis-9-octadecenoic acid - 182 linoleic acid, cis-9,12-octadecadienoic acid - 183 -linolenic acid, cis-9,12,15-octadecatrienoic acid - 201 cis-11-eicosenoic acid - 221 erucic acid, cis-13-docosenoic acid; all other fatty acids are designated by number of carbon atoms: number of double bonds National Research Council of Canada Publication No. 35896  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