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 968 毫秒
1.
Photosynthesis of washed cells of Synechococcus UTEX 625 grown on 5% CO2 was markedly stimulated (647 ± 50%) at pH 8.0 by the addition of low concentrations of NaCl (concentration required for half-maximal response, K½, = 18 micromolar). Studies with KCl and Na2SO4 showed that the stimulation was due to Na+. Photosynthesis at pH 6.1 was only slightly stimulated by Na+. The response of photosynthesis at pH 8.0 to [Na+] was strongly sigmoidal for dissolved inorganic carbon ([DIC] ≤ 500 micromolar). Cells grown with high total [DIC], but air-levels of CO2, at pH 9.6 showed the same response to low [Na+]. The absence of Na+ could be partially, but not completely overcome, by higher [DIC]. Various methods for examining CO2 or HCO3 use (K½CO2 determination; isotopic disequilibrium; and consideration of HCO3 dehydration rate) were consistent with CO2 use by the cells, but HCO3 use could not be ruled out. Isotopic disequilibrium studies showed that CO2 use was stimulated by Na+. Cells grown on 5% CO2 accumulated DIC against a concentration gradient by a process (or processes) dependent on Na+. No evidence for uptake of Na+ concomitant with DIC uptake could be found. The lack of O2 evolution during the initial and most rapid period of DIC accumulation suggested that the required energy was obtained from cyclic photophosphorylation.  相似文献   

2.
The Na+ requirement for photosynthesis and its relationship to dissolved inorganic carbon (DIC) concentration and Li+ concentration was examined in air-grown cells of the cyanobacterium Synechococcus leopoliensis UTEX 625 at pH 8. Analysis of the rate of photosynthesis (O2 evolution) as a function of Na+ concentration, at fixed DIC concentration, revealed two distinct regions to the response curve, for which half-saturation values for Na+ (K½[Na+]) were calculated. The value of both the low and the high K½(Na+) was dependent upon extracellular DIC concentration. The low K½(Na+) decreased from 1000 micromolar at 5 micromolar DIC to 200 micromolar at 140 micromolar DIC whereas over the same DIC concentration range the high K½(Na+) decreased from 10 millimolar to 1 millimolar. The most significant increases in photosynthesis occurred in the 1 to 20 millimolar range. A fraction of total photosynthesis, however, was independent of added Na+ and this fraction increased with increased DIC concentration. A number of factors were identified as contributing to the complexity of interaction between Na+ and DIC concentration in the photosynthesis of Synechococcus. First, as revealed by transport studies and mass spectrometry, both CO2 and HCO3 transport contributed to the intracellular supply of DIC and hence to photosynthesis. Second, both the CO2 and HCO3 transport systems required Na+, directly or indirectly, for full activity. However, micromolar levels of Na+ were required for CO2 transport while millimolar levels were required for HCO3 transport. These levels corresponded to those found for the low and high K½(Na+) for photosynthesis. Third, the contribution of each transport system to intracellular DIC was dependent on extracellular DIC concentration, where the contribution from CO2 transport increased with increased DIC concentration relative to HCO3 transport. This change was reflected in a decrease in the Na+ concentration required for maximum photosynthesis, in accord with the lower Na+-requirement for CO2 transport. Lithium competitively inhibited Na+-stimulated photosynthesis by blocking the cells' ability to form an intracellular DIC pool through Na+-dependent HCO3 transport. Lithium had little effect on CO2 transport and only a small effect on the size of the pool it generated. Thus, CO2 transport did not require a functional HCO3 transport system for full activity. Based on these observations and the differential requirement for Na+ in the CO2 and HCO3 transport system, it was proposed that CO2 and HCO3 were transported across the membrane by different transport systems.  相似文献   

3.
The role of external carbonic anhydrase in inorganic carbon acquisition and photosynthesis by Chlamydomonas reinhardii at alkaline pH (8.0) was studied. Acetazolamide (50 micromolar) completely inhibited external carbonic anhydrase (CA) activity as determined from isotopic disequilibrium experiments. Under these conditions, photosynthetic rates at low dissolved inorganic carbon (DIC) were far greater than could be maintained by CO2 supplied from the spontaneous dehydration of HCO3 thereby showing that C. reinhardii has the ability to utilize exogenous HCO3. Acetazolamide increased the concentration of DIC required to half-saturate photosynthesis from 38 to 80 micromolar, while it did not affect the maximum photosynthetic rate. External CA activity was also removed from the cell-wall-less mutant (CW-15) by washing. This had no effect on the photosynthetic kinetics of the algae while the addition of acetazolamide to washed cells (CW-15) increased the K½DIC from 38 to 80 micromolar. Acetazolamide also caused a buildup of the inorganic carbon pool upon NaHCO3 addition, indicating that this compound partially inhibited internal CA activity. The effects of acetazolamide on the photosynthetic kinetics of C. reinhardii are likely due to the inhibition of internal rather than a consequence of the inhibition of external CA. Further analysis of the isotopic disequilibrium experiments at saturating concentration of DIC provided evidence consistent with active CO2 transport by C. reinhardii. The observation that C. reinhardii has the ability to take up both CO2 and bicarbonate throws into question the role of external CA in the accumulation of DIC in this alga.  相似文献   

4.
At low levels of dissolved inorganic carbon (DIC) and alkaline pH the rate of photosynthesis by air-grown cells of Synechococcus leopoliensis (UTEX 625) was enhanced 7- to 10-fold by 20 millimolar Na+. The rate of photosynthesis greatly exceeded the CO2 supply rate and indicated that HCO3 was taken up by a Na+-dependent mechanism. In contrast, photosynthesis by Synechococcus grown in standing culture proceeded rapidly in the absence of Na+ and exceeded the CO2 supply rate by 8 to 45 times. The apparent photosynthetic affinity (K½) for DIC was high (6-40 micromolar) and was not markedly affected by Na+ concentration, whereas with air-grown cells K½ (DIC) decreased by more than an order of magnitude in the presence of Na+. Lithium, which inhibited Na+-dependent HCO3 uptake in air-grown cells, had little effect on Na+-independent HCO3 uptake by standing culture cells. A component of total HCO3 uptake in standing culture cells was also Na+-dependent with a K½ (Na+) of 4.8 millimolar and was inhibited by lithium. Analysis of 14C-fixation during isotopic disequilibrium indicated that standing culture cells also possessed a Na+-independent CO2 transport system. The conversion from Na+-independent to Na+-dependent HCO3 uptake was readily accomplished by transferring cells grown in standing to growth in cultures bubbled with air. These results demonstrated that the conditions experienced during growth influenced the mode by which Ssynechococcus acquired HCO3 for subsequent photosynthetic fixation.  相似文献   

5.
An O2 electrode system with a specially designed chamber for `whorl' cell complexes of Chara corallina was used to study the combined effects of inorganic carbon and O2 concentrations on photosynthetic O2 evolution. At pH = 5.5 and 20% O2, cells grown in HCO3 medium (low CO2, pH ≥ 9.0) exhibited a higher affinity for external CO2 (K½(CO2) = 40 ± 6 micromolar) than the cells grown for at least 24 hours in high-CO2 medium (pH = 6.5), (K½(CO2) = 94 ± 16 micromolar). With O2 ≤ 2% in contrast, both types of cells showed a high apparent affinity (K½(CO2) = 50 − 52 micromolar). A Warburg effect was detectable only in the low affinity cells previously cultivated in high-CO2 medium (pH = 6.5). The high-pH, HCO3-grown cells, when exposed to low pH (5.5) conditions, exhibited a response indicating an ability to fix CO2 which exceeded the CO2 externally supplied, and the reverse situation has been observed in high-CO2-grown cells. At pH 8.2, the apparent photosynthetic affinity for external HCO3 (K½[HCO3]) was 0.6 ± 0.2 millimolar, at 20% O2. But under low O2 concentrations (≤2%), surprisingly, an inhibition of net O2 evolution was elicited, which was maximal at low HCO3 concentrations. These results indicate that: (a) photorespiration occurs in this alga and can be revealed by cultivation in high-CO2 medium, (b) Chara cells are able to accumulate CO2 internally by means of a process apparently independent of the plasmalemma HCO3 transport system, (c) molecular oxygen appears to be required for photosynthetic utilization of exogenous HCO3: pseudocyclic electron flow, sustained by O2 photoreduction, may produce the additional ATP needed for the HCO3 transport.  相似文献   

6.
To examine the factors which limit photosynthesis and their role in photosynthetic adaptation to growth at low dissolved inorganic carbon (DIC), Synechococcus leopoliensis was grown at three concentrations (as signified by brackets) of DIC, high (1000-1800 micromolar), intermediate (200-300 micromolar), and low (10-20 micromolar). In all cell types photosynthesis varied from being ribulose bisphosphate (RuBP)-saturated at low external [DIC] to RuBP-limited at high external [DIC]. The maximum rate of photosynthesis (Pmax) was achieved when the internal concentration of RuBP fell below the active site density of RuBP carboxylase/oxygenase (Rubisco). At rates of photosynthesis below Pmax, photosynthetic capacity was limited by the ability of the cell to transport inorganic carbon and to supply CO2 to Rubisco. Adaptation to low DIC was reflected by a decrease in the [DIC] required to half-saturate photosynthesis. Simultaneous mass-spectrometric measurement of rates of photosynthesis and DIC transport showed that the initial slope of the photosynthesis versus [DIC] curve is identical to the initial slope of the DIC transport versus [DIC] curve. This provided evidence that the enhanced capacity for DIC transport which occurs upon adaptation to low [DIC] was responsible for the increase in the initial slope of the photosynthesis versus [DIC] curve and therefore the decrease in the half saturation constant of photosynthesis with respect to DIC. Levels of RuBP and in vitro Rubisco activity varied only slightly between high and intermediate [DIC] grown cells but fell significantly (65-70%) in low [DIC] grown cells. Maximum rates of photosynthesis followed a similar pattern with Pmax only slightly lower in intermediate [DIC] grown cells than in high [DIC] grown cells, but much lower in low [DIC] grown cells. The changing response of photosynthesis to [DIC] during adaptation to low DIC, may be explained by the interaction between DIC-transport limited and [RuBP]-limited photosynthesis.  相似文献   

7.
Climate change is expected to bring about alterations in the marine physical and chemical environment that will induce changes in the concentration of dissolved CO2 and in nutrient availability. These in turn are expected to affect the physiological performance of phytoplankton. In order to learn how phytoplankton respond to the predicted scenario of increased CO2 and decreased nitrogen in the surface mixed layer, we investigated the diatom Phaeodactylum tricornutum as a model organism. The cells were cultured in both low CO2 (390 μatm) and high CO2 (1000 μatm) conditions at limiting (10 μmol L−1) or enriched (110 μmol L−1) nitrate concentrations. Our study shows that nitrogen limitation resulted in significant decreases in cell size, pigmentation, growth rate and effective quantum yield of Phaeodactylum tricornutum, but these parameters were not affected by enhanced dissolved CO2 and lowered pH. However, increased CO2 concentration induced higher rETRmax and higher dark respiration rates and decreased the CO2 or dissolved inorganic carbon (DIC) affinity for electron transfer (shown by higher values for K1/2 DIC or K1/2 CO2). Furthermore, the elemental stoichiometry (carbon to nitrogen ratio) was raised under high CO2 conditions in both nitrogen limited and nitrogen replete conditions, with the ratio in the high CO2 and low nitrate grown cells being higher by 45% compared to that in the low CO2 and nitrate replete grown ones. Our results suggest that while nitrogen limitation had a greater effect than ocean acidification, the combined effects of both factors could act synergistically to affect marine diatoms and related biogeochemical cycles in future oceans.  相似文献   

8.
The blue-green alga (Cyanobacterium) Synechococcus leopoliensis (Racib.) Komarek was grown in dissolved inorganic carbon [DIC]-limited chemostats over the entire range of growth rates. At each growth rate, the kinetics of photosynthesis with respect to [DIC] and the maximal rate of photosynthesis (Pmax) were determined. The half-saturation constant for [DIC]-limited photosynthesis (K1/2DIC) for cells growing below 1.7 d?1 was constant (4.7 μM) whereas for growth rates between 1.7 d?1 and 2.1 d?1max) the kinetics of photosynthesis were multiphasic with an apparent K1/2DIC between 1.5–2.0 mM. Pmax increased in a linear fashion with growth rate for growth rates below 1.7 d?1. No trend in Pmax was apparent for growth rates greater than 1.7 d?1. These kinetic parameters were used to predict a growth rate versus [DIC] relationship. Results show that the Monod relationship is a physiologically valid expression of growth as a function of [DIC] provided (K1/2DIC) remains constant. The major change in (K1/2DIC) as μ approaches μmax results in the conclusion that two separate and distinct Monod equations must be used to describe growth as a function of DIC over the entire growth range. These results point to a major discontinuity in the μ vs. [DIC] curve at 1.7 d?1 which corresponds to the change from high to low affinity photosynthetic kinetics. We believe these results account for the previously described deficiencies of the Monod equation in describing [DIC]-limited algal growth.  相似文献   

9.
Goyal A  Tolbert NE 《Plant physiology》1989,89(4):1264-1269
Neither Dunaliella cells grown with 5% CO2 nor their isolated chloroplasts had a CO2 concentrating mechanism. These cells primarily utilized CO2 from the medium because the K(0.5) (HCO3) increase from 57 micromolar at pH 7.0 to 1489 micromolar at pH 8.5, where as the K(0.5) CO2 was about 12 micromolar over the pH range. After air adaptation for 24 hours in light, a CO2 concentrating mechanism was present that decreased the K0.5 (CO2) to about 0.5 micromolar and K0.5 (HCO3) to 11 micromolar at pH 8. These K0.5 values suggest that air-adapted cells preferentially concentrated CO2 but could also use HCO3 from the medium. Chloroplasts isolated from air-adapted cells had a K(0.5) for total inorganic carbon of less than 10 micromolar compared to 130 micromolar for chloroplasts from cells grown on high CO2. Chloroplasts from air-adapted cells, but not CO2-grown cells, concentrate inorganic carbon internally to 1 millimolar in 60 seconds from 240 micromolar in the medium. Maximum uptake rates occurred after preillumination of 45 seconds to 3 minutes. The CO2 concentrating mechanism by chloroplasts from air-adapted cells was light dependent and inhibited by 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) or flurocarbonyl-cyamidephenylhydrazone (FCCP). Phenazine-methosulfate at 10 micromolar to provide cyclic phosphorylation partially reversed the inhibition by DCMU but not by FCCP. One to 0.1 millimolar vanadate, an inhibitor of plasma membrane ATPase, inhibited inorganic carbon accumulation by isolated chloroplasts. Vanadate had no effect on CO2 concentration by whole cells, as it did not readily cross the cell plasmalemma. Addition of external ATP to the isolated chloroplast only slightly stimulated inorganic carbon uptake and did not reverse vanadate inhibition by more than 25%. These results are consistent with a CO2 concentrating mechanism in Dunaliella cells which consists in part of an inorganic carbon transporter at the chloroplast envelope that is energized by ATP from photosynthetic electron transport.  相似文献   

10.
Tomany MJ  Kent SS 《Plant physiology》1986,80(4):1055-1058
When ribulose-1,5-bisphosphate carboxylase is assayed under N2 using [3H]ribulose 1,5-bisphosphate and 14CO2, [3H]3-phosphoglycerate and [14C]3-phosphoglycerate are produced in nonstoichiometric amounts in a ratio which approaches 7 at low concentrations of CO2 (2 micromolar) assuming a 1:1 ratio at Vmax (280 micromolar). The log of the molar ratio varies as a linear function of log[CO2]. Nonstoichiometry could be explained by CO2 contaminatio of the reactants or tritium contamination of the products. However, the magnitude of CO2 contamination required (18 ± 4 micromolar) is far in excess of controlled CO2 (<0.1 micromolar), and the required tritium contaminant would have to vary from 30 to 85% of the purified 3-phosphoglycerate at the 58 and 2 micromolar CO2 assay levels, respectively. This contrasts with detectable tritium contamination which is only 1 to 4% and correctable. Nonstoichiometry is evident using either 1 or 5 labeled [3H]ribulose 1,5-bisphosphate. When 3-phosphoglycerate is reisolated as glycerate the 3H/14C ratio remains unchanged.  相似文献   

11.
Scenedesmus cells grown on high CO2, when adapted to air levels of CO2 for 4 to 6 hours in the light, formed two concentrating processes for dissolved inorganic carbon: one for utilizing CO2 from medium of pH 5 to 8 and one for bicarbonate accumulation from medium of pH 7 to 11. Similar results were obtained with assays by photosynthetic O2 evolution or by accumulation of dissolved inorganic carbon inside the cells. The CO2 pump with K0.5 for O2 evolution of less than 5 micromolar CO2 was similar to that previously studied with other green algae such as Chlamydomonas and was accompanied by plasmalemma carbonic anhydrase formation. The HCO3 concentrating process between pH 8 to 10 lowered the K0.5 (DIC) from 7300 micromolar HCO3 in high CO2 grown Scenedesmus to 10 micromolar in air-adapted cells. The HCO3 pump was inhibited by vanadate (Ki of 150 micromolar), as if it involved an ATPase linked HCO3 transporter. The CO2 pump was formed on low CO2 by high-CO2 grown cells in growth medium within 4 to 6 hours in the light. The alkaline HCO3 pump was partially activated on low CO2 within 2 hours in the light or after 8 hours in the dark. Full activation of the HCO3 pump at pH 9 had requirements similar to the activation of the CO2 pump. Air-grown or air-adapted cells at pH 7.2 or 9 accumulated in one minute 1 to 2 millimolar inorganic carbon in the light or 0.44 millimolar in the dark from 150 micromolar in the media, whereas CO2-grown cells did not accumulate inorganic carbon. A general scheme for concentrating dissolved inorganic carbon by unicellular green algae utilizes a vanadate-sensitive transporter at the chloroplast envelope for the CO2 pump and in some algae an additional vanadate-sensitive plasmalemma HCO3 transporter for a HCO3 pump.  相似文献   

12.
The enzymic properties of ribulose 1,5-bisphosphate (RuBP) carboxylase/oxygenase purified from rice (Oryza sativa L.) leaves were studied. Rice RuBPcarboxylase, activated by preincubation with CO2 and Mg2+ like other higher plant carboxylases, had an activation equilibrium constant (KcKMg) of 1.90 × 105 to 2.41 × 105 micromolar2 (pH 8.2 and 25°C). Kinetic parameters of carboxylation and oxygenation catalyzed by the completely activated enzyme were examined at 25°C and the respective optimal pHs. The Km(CO2), Km(RuBP), and Vmax values for carboxylation were 8 micromolar, 31 micromolar, and 1.79 units milligram−1, respectively. The Km(O2), Km(RuBP), and Vmax values for oxygenation were 370 micromolar, 29 micromolar, and 0.60 units milligram−1, respectively.

Comparison of rice leaf RuBP carboxylase with other C3 plant carboxylases showed that it had a relatively high affinity for CO2 but the lowest catalytic turnover number (Vmax) among the species examined.

  相似文献   

13.
The human pathogen Shigella flexneri subverts host function and defenses by deploying a cohort of effector proteins via a type III secretion system. The IpaH family of 10 such effectors mimics ubiquitin ligases but bears no sequence or structural homology to their eukaryotic counterpoints. Using rates of 125I-polyubiquitin chain formation as a functional read out, IpaH9.8 displays V-type positive cooperativity with respect to varying concentrations of its Ubc5B∼125I-ubiquitin thioester co-substrate in the nanomolar range ([S]½ = 140 ± 32 nm; n = 1.8 ± 0.1) and cooperative substrate inhibition at micromolar concentrations ([S]½ = 740 ± 240 nm; n = 1.7 ± 0.2), requiring ordered binding to two functionally distinct sites per subunit. The isosteric substrate analog Ubc5BC85S-ubiquitin oxyester acts as a competitive inhibitor of wild-type Ubc5B∼125I-ubiquitin thioester (Ki = 117 ± 29 nm), whereas a Ubc5BC85A product analog shows noncompetitive inhibition (Ki = 2.2 ± 0.5 μm), consistent with the two-site model. Re-evaluation of a related IpaH3 crystal structure (PDB entry 3CVR) identifies a symmetric dimer consistent with the observed cooperativity. Genetic disruption of the predicted IpaH9.8 dimer interface reduces the solution molecular weight and significantly ablates the kcat but not [S]½ for polyubiquitin chain formation. Other studies demonstrate that cooperativity requires the N-terminal leucine-rich repeat-targeting domain and is transduced through Phe395. Additionally, these mechanistic features are conserved in a distantly related SspH2 Salmonella enterica ligase. Kinetic parallels between IpaH9.8 and the recently revised mechanism for E6AP/UBE3A (Ronchi, V. P., Klein, J. M., and Haas, A. L. (2013) E6AP/UBE3A ubiquitin ligase harbors two E2∼ubiquitin binding sites. J. Biol. Chem. 288, 10349–10360) suggest convergent evolution of the catalytic mechanisms for prokaryotic and eukaryotic ligases.  相似文献   

14.
Data on calcification rate of coral and crustose coralline algae were used to test the proton flux model of calcification. There was a significant correlation between calcification (G) and the ratio of dissolved inorganic carbon (DIC) to proton concentration ([DIC] : [H+] ratio). The ratio is tightly correlated with [CO32−] and with aragonite saturation state (Ωa). An argument is presented that correlation does not prove cause and effect, and that Ωa and [CO32−] have no basic physiological meaning on coral reefs other than a correlation with [DIC] : [H+] ratio, which is the driver of G.  相似文献   

15.
A new method is presented for measurement of the CO2/O2 specificity factor of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). The [14C]3-phosphoglycerate (PGA) from the Rubisco carboxylase reaction and its dilution by the Rubisco oxygenase reaction was monitored by directly measuring the specific radioactivity of PGA. 14CO2 fixation with Rubisco occurred under two reaction conditions: carboxylase with oxygenase with 40 micromolar CO2 in O2-saturated water and carboxylase only with 160 micromolar CO2 under N2. Detection of the specific radioactivity used the amount of PGA as obtained from the peak area, which was determined by pulsed amperometry following separation by high-performance anion exchange chromatography and the radioactive counts of the [14C]PGA in the same peak. The specificity factor of Rubisco from spinach (Spinacia oleracea L.) (93 ± 4), from the green alga Chlamydomonas reinhardtii (66 ± 1), and from the photosynthetic bacterium Rhodospirillum rubrum (13) were comparable with the published values measured by different methods.  相似文献   

16.
While increasing atmospheric carbon dioxide (CO2) concentration alters global water chemistry (Ocean Acidification; OA), the degree of changes vary on local and regional spatial scales. Inshore fringing coral reefs of the Great Barrier Reef (GBR) are subjected to a variety of local pressures, and some sites may already be marginal habitats for corals. The spatial and temporal variation in directly measured parameters: Total Alkalinity (TA) and dissolved inorganic carbon (DIC) concentration, and derived parameters: partial pressure of CO2 (pCO2); pH and aragonite saturation state (Ωar) were measured at 14 inshore reefs over a two year period in the GBR region. Total Alkalinity varied between 2069 and 2364 µmol kg−1 and DIC concentrations ranged from 1846 to 2099 µmol kg−1. This resulted in pCO2 concentrations from 340 to 554 µatm, with higher values during the wet seasons and pCO2 on inshore reefs distinctly above atmospheric values. However, due to temperature effects, Ωar was not further reduced in the wet season. Aragonite saturation on inshore reefs was consistently lower and pCO2 higher than on GBR reefs further offshore. Thermodynamic effects contribute to this, and anthropogenic runoff may also contribute by altering productivity (P), respiration (R) and P/R ratios. Compared to surveys 18 and 30 years ago, pCO2 on GBR mid- and outer-shelf reefs has risen at the same rate as atmospheric values (∼1.7 µatm yr−1) over 30 years. By contrast, values on inshore reefs have increased at 2.5 to 3 times higher rates. Thus, pCO2 levels on inshore reefs have disproportionately increased compared to atmospheric levels. Our study suggests that inshore GBR reefs are more vulnerable to OA and have less buffering capacity compared to offshore reefs. This may be caused by anthropogenically induced trophic changes in the water column and benthos of inshore reefs subjected to land runoff.  相似文献   

17.
The activities of three enzymes and the concentration of intermediates involved in the synthesis of N,N-dimethyltryptamine (DMT) from endogenous tryptophan (TRP) have been measured in vitro in seedlings of Phalaris aquatica L. cv Australian Commercial over 16 days after planting. The activities of tryptophan decarboxylase and the two N-methyl-transferases increased rapidly to maximal rates of substrate conversion at day 5 of 95, 1000, and 2200 micromoles per hour per milliliter, respectively. After these maximal rates, the activities decreased rapidly. The concentration of intermediates increased rapidly from zero in the seeds to maximal values of 25 and 53 micromolar at day 5 for tryptamine (T) and N-methyltryptamine (MT), respectively, 1000 micromolar at day 6 for TRP, and 650 micromolar at day 8 for DMT. The concentration of DMT and of all the intermediates in its synthesis declined rapidly after the maximal value had been reached. A mathematical model of the pathway from TRP to DMT using these enzymes correctly predicts the concentrations of T and MT, intermediates whose concentration is determined only by the pathway, and confirms that these three enzymes are responsible for the in vivo synthesis of DMT. Kinetic studies are reported for these enzymes. Tryptophan decarboxylase uses pyridoxal phosphate (PALP) as a coenzyme and has the following kinetic constants: KmPALP = 2.5 micromolar, KmTRP = 200 micromolar, KiMT = 5 millimolar, and KiDMT = 4 millimolar. The N-methyltransferases use S-adenosylmethionine (SAM) as substrate; S-adenosylhomocysteine (SAH) is assumed to be the product. The mechanism of secondary indolethylamine-N-methyltransferase, determined by initial velocity studies, is rapid equilibrium random with formation of both dead end complexes. Secondary indolethylamine-N-methyltransferase methylates both MT and 5-methoxy-N-methyltryptamine (5MeOMT). The kinetic constants for the methylation of MT are: KMT = 40 ± 6, KSAM = 55 ± 15, KDMT = 60, KSAH = 4.3 ± 0.4 micromolar with unity interaction factors. The kinetic constants for the conversion of 5MeOMT to 5-methoxy-N,N-dimethyltryptamine (5MeODMT) are K5MeOMT = 40 ± 10, KSAM = 90 ± 40, and KSAH = 2.9 ± 0.3 micromolar with unity interaction factors, except for SAM-5MeODMT = 2.0 ± 0.9 and SAH-5MeOMT = 0.45 ± 0.25. The kinetic constants for primary indolethylamine N-methyltransferase are KmT = 20, KmSAM = 40, KiDMT = 450 micromolar with the substrates binding independently.  相似文献   

18.
Physiological properties of photosynthesis were determined in the marine diatom, Phaeodactylum tricornutum UTEX640, during acclimation from 5% CO2 to air and related to H2CO3 dissociation kinetics and equilibria in artificial seawater. The concentration of dissolved inorganic carbon at half maximum rate of photosynthesis (K0·5[DIC]) value in high CO2‐grown cells was 1009 mmol m ? 3 but was reduced three‐fold by the addition of bovine carbonic anhydrase (CA), whereas in air‐grown cells K0·5[DIC] was 71 mmol m ? 3, irrespective of the presence of CA. The maximum rate of photosynthesis (Pmax) values varied between 300 and 500 μ mol O2 mg Chl ? 1 h ? 1 regardless of growth pCO2. Bicarbonate dehydration kinetics in artificial seawater were re‐examined to evaluate the direct HCO3 ? uptake as a substrate for photosynthesis. The uncatalysed CO2 formation rate in artificial seawater of 31·65°/oo of salinity at pH 8·2 and 25 °C was found to be 0·6 mmol m ? 3 min ? 1 at 100 mmol m ? 3 DIC, which is 53·5 and 7·3 times slower than the rates of photosynthesis exhibited in air‐ and high CO2‐grown cells, respectively. These data indicate that even high CO2‐grown cells of P. tricornutum can take up both CO2 and HCO3 ? as substrates for photosynthesis and HCO3 ? use improves dramatically when the cells are grown in air. Detailed time courses were obtained of changes in affinity for DIC during the acclimation of high CO2‐grown cells to air. The development of high‐affinity photosynthesis started after a 2–5 h lag period, followed by a steady increase over the next 15 h. This acclimation time course is the slowest to be described so far. High CO2‐grown cells were transferred to controlled DIC conditions, at which the concentrations of each DIC species could be defined, and were allowed to acclimate for more than 36 h. The K0·5[DIC] values in acclimated cells appeared to be correlated only with [CO2(aq)] in the medium but not to HCO3 ? , CO32 ? , total [DIC] or the pH of the medium and indicate that the critical signal regulating the affinity of cells for DIC in the marine diatom, P. tricornutum, is [CO2(aq)] in the medium.  相似文献   

19.
Inorganic Carbon Uptake by Chlamydomonas reinhardtii   总被引:15,自引:12,他引:3  
The rates of CO2-dependent O2 evolution by Chlamydomonas reinhardtii, grown with either air levels of CO2 or air with 5% CO2, were measured at varying external pH. Over a pH range of 4.5 to 8.5, the external concentration of CO2 required for half-maximal rates of photosynthesis was constant, averaging 25 micromolar for cells grown with 5% CO2. This is consistent with the hypothesis that these cells take up CO2 but not HCO3 from the medium and that their CO2 requirement for photosynthesis reflects the Km(CO2) of ribulose bisphosphate carboxylase. Over a pH range of 4.5 to 9.5, cells grown with air required an external CO2 concentration of only 0.4 to 3 micromolar for half-maximal rates of photosynthesis, consistent with a mechanism to accumulate external inorganic carbon in these cells. Air-grown cells can utilize external inorganic carbon efficiently even at pH 4.5 where the HCO3 concentration is very low (40 nanomolar). However, at high external pH, where HCO3 predominates, these cells cannot accumulate inorganic carbon as efficiently and require higher concentrations of NaHCO3 to maintain their photosynthetic activity. These results imply that, at the plasma membrane, CO2 is the permeant inorganic carbon species in air-grown cells as well as in cells grown on 5% CO2. If active HCO3 accumulation is a step in CO2 concentration by air-grown Chlamydomonas, it probably takes place in internal compartments of the cell and not at the plasmalemma.  相似文献   

20.
Light-induced acidification by the cyanobacterium Anabaena variabilis is biphasic (a fast phase I and slow phase II) and shown to be sodium-dependent with an optimum concentration of 40 to 60 millimolar Na+. Cells grown under low CO2 concentrations at pH 9 (i.e. mainly HCO3 present in the medium) exhibited the slow phase II of proton efflux only, while cells grown under low CO2 concentrations at pH 6.3 (i.e. CO2 and HCO3 present) exhibited both phases. Light-induced proton release of phase I was dependent on inorganic carbon available in the bathing medium with an apparent Km for CO2 of 20 to 70 micromolar. As was concluded from the CO2 dependence of acidification measured at different pH of the bathing medium, bicarbonate inhibited phase-I acidification noncompetetively. Acidification was inhibited by acetazolamide, an inhibitor of carbonic anhydrase. Apparently, acidification of phase I is due to a light-dependent uptake of CO2 being converted to HCO3 by a carbonic anhydrase-like function of the HCO3-transport system (M Volokita, D Zenvirth, A Kaplan, L Reinhold 1984 Plant Physiol 76: 599-602) before or during entering the cell, thus releasing one proton per CO2 converted to HCO3.  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