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 156 毫秒
1.
We have compared the characteristics of nitrate uptake by Aphanothece halophytica grown under non-stress and salt-stress conditions. Both cell types showed essentially similar patterns of nitrate uptake toward ammonium, nitrite, and DL-glyceraldehyde. Although the affinities of nitrate to non-stress cells and salt-stress cells were not significantly different, i.e., Ks = 416 and 450 microM, respectively, the V(max) value for non-stress cells was about twofold of that for salt-stress cells (9.1 vs 5.3 micromol min(-1) mg(-1) Chl). Nitrate uptake by A. halophytica was found to be dependent on Na+. Ammonium inhibited nitrate uptake, and the presence of methionine sulfoximine could not release the inhibition by ammonium. Nitrite appeared to competitively inhibit nitrate uptake with a K(i) value of 84 microM. Both chloride and phosphate anions did not affect nitrate uptake. DL-Glyceraldehyde, an inhibitor of CO2 fixation, caused a reduction in the uptake of nitrate.  相似文献   

2.
The function of HKT1 in roots is controversial. We tackled this controversy by studying Na+ uptake in barley (Hordeum vulgare) roots, cloning the HvHKT1 gene, and expressing the HvHKT1 cDNA in yeast (Saccharomyces cerevisiae) cells. High-affinity Na+ uptake was not detected in plants growing at high K+ but appeared soon after exposing the plants to a K(+)-free medium. It was a uniport, insensitive to external K+ at the beginning of K+ starvation and inhibitable by K+ several hours later. The expression of HvHKT1 in yeast was Na+ (or K+) uniport, Na(+)-K+ symport, or a mix of both, depending on the construct from which the transporter was expressed. The Na+ uniport function was insensitive to external K+ and mimicked the Na+ uptake carried out by the roots at the beginning of K+ starvation. The K+ uniport function only took place in yeast cells that were completely K+ starved and disappeared when internal K+ increased, which makes it unlikely that HvHKT1 mediates K+ uptake in roots. Mutation of the first in-frame AUG codon of HvHKT1 to CUC changed the uniport function into symport. The expression of the symport from either mutants or constructs keeping the first in-frame AUG took place only in K(+)-starved cells, while the uniport was expressed in all conditions. We discuss here that the symport occurs only in heterologous expression. It is most likely related to the K+ inhibitable Na+ uptake process of roots that heterologous systems fail to reproduce.  相似文献   

3.
Yeast biomass enriched with iron could represent a new and safer solution for prevention from anaemia development. Such an iron source is less toxic and has better absorbability in organisms. The purpose of our research was the determination of the most suitable iron source in the cultivation medium for the yeast Saccharomyces cerevisiae, regarding good growth and iron accumulation in cells. Iron(III) citrate, iron(III) chloride, iron(III) nitrate and Fe-EDTA complex were used. The uptake of the chosen iron compound, Fe(III) citrate, by the yeasts Candida intermedia and Kluyveromyces marxianus was also investigated. Different growth behaviour of the three yeast strains in the presence of Fe(III) citrate was observed. The highest amounts of accumulated iron in S. cerevisiae, C. intermedia and K. marxianus biomass were about 13, 20 and 34mgFeg(-1)dry wt., respectively. To optimise the accumulation of iron in K. marxianus and to characterise iron enriched yeast biomass, further experiments are needed.  相似文献   

4.
1. When yeast N.C.Y.C. 240 was grown with maltose in a complex medium based on yeast extract and peptone, washed cell preparations fermented alpha-methyl glucoside much more slowly than maltose. 2. The yeast absorbed alpha-methyl[14C]glucoside from a 10mM solution in the presence of antimycin and iodoacetamide, producing [14C]glucose, which accumulated outside the cells. The yeast itself contained hexose phosphates, trehalose, alpha-methyl glucoside and other products labelled with 14C, but no alpha-methyl glucoside phosphate. 3. About 1 equiv. of protons was absorbed with each equivalent of alpha-methylglucoside, and 1 equiv. of K+ ions left the yeast. 4. alpha-Thioethyl glucoside was also absorbed along with protons. Studies by g.l.c. showed that the yeast concentrated the compound without metabolizing it. 5. The presence of trehalose, sucrose, maltose, L-sorbose, glucose or alpha-phenyl glucoside in each case immediately stimulated proton uptake, whereas fructose, 3-O-methylglucose and 2-deoxyglucose failed to do so. 6. The observations support the conclusion that alpha-thioethyl glucoside, alpha-methyl glucoside and maltose are substrates of one or more proton symports, whereas they seem inconsistent with the notion that the absorption of alpha-methyl glucoside involves the phosphorylation of the carbohydrate [Van Stevenick (1970) Biochim. Biophys. Acta 203, 376-384].  相似文献   

5.
CHL1 (AtNRT1.1) is a dual-affinity nitrate transporter of Arabidopsis thaliana, in which phosphorylation at Thr 101 switches CHL1 from low to high nitrate affinity. CHL1 expressed in a Hansenula polymorpha high-affinity nitrate-transporter deficient mutant (Deltaynt1) restores nitrate uptake and growth. These events take place at nitrate concentrations as low as 500 muM, suggesting that CHL1 has a high-affinity for nitrate in yeast. Accordingly, CHL1 expressed in H. polymorpha presents a K (m) for nitrate of about 125 muM. The absence of nitrate, the CHL1 gene inducer, showed the high turnover rate of CHL1 expressed in yeast, which is counteracted by nitrate CHL1 induction. Furthermore, H. polymorpha strains expressing CHL1 become sensitive to 250 muM chlorate, as expected for CHL1 high-affinity behaviour. Given that CHL1 presented high affinity by nitrate, we study the role of CHL1 Thr101 in yeast. Strains producing CHL1Thr101Ala, unable to undergo phosphorylation, and CHL1Thr101Asp, where CHL1 phosphorylation is constitutively mimicked, were used. Yeast strains expressing CHL1Thr101Ala, CHL1Thr101Asp and CHL1 at the same rate showed that Deltaynt1CHL1Thr101Ala is strikingly unable to transport nitrate and contains a very low amount of CHL1 protein; however, Deltaynt1CHL1Thr101Asp restores nitrate uptake and growth, although no significant changes in nitrate affinity were observed. Our results show that CHL1-Thr101 is involved in regulating the levels of CHL1 expressed in yeast and suggest that the phosphorylation of this residue could be involved in targeting this nitrate transporter to the plasma membrane. The functional expression of CHL1 in H. polymorpha reveals that this yeast is a suitable tool for evaluating the real nitrate transport capacity of plant putative nitrate transporters belonging to different families and study their regulation and structure function relationship.  相似文献   

6.
1. At pH 4.5 and 30degreesC, yeast preparations depleted of ATP in the presence of antimycin and deoxyglucose spontaneously lost K+, gaining roughly an equivalent amount of H+. 2. Five proton conductors including azide and 2,4-dinitrophenol accelerated this process, as did [14C]glycine, which was absorbed with two extra equivalents of H+. 3. The rate of glycine uptake at pH 4.5 diminished fourfold when cellular K+ fell by 20%. 4. The distribution of [14C]propionate indicated that the intracellular pH fell from 6.2 to 5.7 when the cellular content of K+ fell by 30%. 5. Glycine uptake from a 5 muM solution was about 400 times faster at pH 4.5 than it was at pH 7.4 with 100mM-KC1 present ostensibly to lower the membrane potential. 6. Yeast preparations containing 2mM-[14C]glycine absorbed a further amount from a 0.1 muM solution at pH 4.5. After about 10 min a net movement of [14C]glycine out of the yeast occurred. The ratio of the cellular [14Ia1glycine concentration to the concentration outside the yeast reached 4 X 10(4) in these assays, whereas at pH 7.4 in the presence of 100mM-KC1 it did not exceed 15 in 3h. Dimitrophenol lowered the accumulation ratio at pH 4.5, apparently by causing proton conduction. 7. The observations are consistent with the notion that glycine uptake is driven by a proton symport mechanism. 8. Possible factors governing the strikingly low rate of glycine efflux as opposed to its optimum rate of influx are discussed.  相似文献   

7.
The hyperthermophilic marine archaeon Thermococcus litoralis exhibits high-affinity transport activity for maltose and trehalose at 85 degrees C. The K(m) for maltose transport was 22 nM, and that for trehalose was 17 nM. In cells that had been grown on peptone plus yeast extract, the Vmax for maltose uptake ranged from 3.2 to 7.5 nmol/min/mg of protein in different cell cultures. Cells grown in peptone without yeast extract did not show significant maltose or trehalose uptake. We found that the compound in yeast extract responsible for the induction of the maltose and trehalose transport system was trehalose. [14C]maltose uptake at 100 nM was not significantly inhibited by glucose, sucrose, or maltotriose at a 100 microM concentration but was completely inhibited by trehalose and maltose. The inhibitor constant, Ki, of trehalose for inhibiting maltose uptake was 21 nM. In contrast, the ability of maltose to inhibit the uptake of trehalose was not equally strong. With 20 nM [14C]trehalose as the substrate, a 10-fold excess of maltose was necessary to inhibit uptake to 50%. However, full inhibition was observed at 2 microM maltose. The detergent-solubilized membranes of trehalose-induced cells contained a high-affinity binding protein for maltose and trehalose, with an M(r) of 48,000, that exhibited the same substrate specificity as the transport system found in whole cells. We conclude that maltose and trehalose are transported by the same high-affinity membrane-associated system. This represents the first report on sugar transport in any hyperthermophilic archaeon.  相似文献   

8.
Confluent monolayer cultures of the Madin-Darby canine kidney (MDCK) cell line have been shown to possess a furosemide and bumetanide-sensitive (Na+,K+)-cotransport system. We have studied the effect of anion substitutions on (Na+,K+)-cotransport. In Na+-depleted cells, bumetanide-sensitive uptake of 22Na+ or 86Rb+ exhibited an absolute requirement for extracellular Cl-. Chloride could be replaced in the buffers by Br-, but not by F-, I-, acetate, nitrate, thiocyanate, sulfate, or gluconate. The effect of Cl- was saturating, and Na+-stimulated 86RB+ uptake as well as K+-stimulated 22Na+ uptake was shown to be dependent on the square of the Cl- concentration. The concentration of Cl- which gave half-maximal stimulation of cation cotransport varied between 58 and 70 mM. There was a small degree of cooperativity between the binding affinities for Cl- and K+ at constant Na+ concentrations. Bumetanide-sensitive 36Cl- uptake could be demonstrated when extracellular Na+ and K+ were present simultaneously. Uptake through this system was unaffected by changes in the membrane potential or by the imposition of pH gradients. Together these data strongly suggest that the bumetanide-sensitive transport system in Madin-Darby canine kidney cells co-transports Na+, K+, and Cl- in a ratio of 1:1:2.  相似文献   

9.
We have previously shown that fission yeast encodes a PPZ-like phosphatase, designated Pzhl, which is an important determinant of cation homeostasis. pzh1 delta mutants display increased tolerance to Na+ ions, but they are hypersensitive to KC1 [Balcells, L., Gómez, N., Casamayor, A., Clotet, J. & Ari?o, J. (1997) Eur. J. Biochem. 250, 476-483]. We have immunodetected Pzh1 in yeast extracts and found that this phosphatase is largely associated with particulate fractions. Cells defective in Pzh1 do not show altered efflux of Na+ or Li+ ions, but they accumulate these cations more slowly than wild-type cells. K+ ion content of pzh1 delta cells is about twice that of wild-type cells, and this can be explained by decreased efflux of K+. Therefore, Pzh1 may regulate both Na+ influx and K+ efflux in fission yeast. To test the possible relationship between K+ uptake, Na+ tolerance and Pzh1 function, we deleted the trk1+ gene, which encodes a putative high-affinity transporter of K+ ions. trkl delta mutants grew well even at relatively low concentrations of KCl and did not show significantly altered content or influx of K+ ions. However, they showed a Na(+)-sensitive phenotype which was greatly intensified by deletion of the sod2+ gene (which encodes the major determinant for efflux of Na+ ions), and clearly ameliorated by deletion of the pzh1 phosphatase, as well as by moderate concentrations of KCl in the medium. These results suggest that Trk1 does not mediate the effect of Pzh1 on NaCl tolerance and that fission yeast contains efficient systems, other than Trk1, for uptake of K+ ions.  相似文献   

10.
The wild-type and an R215E mutant of the rat renal cortex sodium/phosphate cotransporter type 2 (NaPi-2) were functionally expressed in the yeast Saccharomyces cerevisiae strain MB192, a cell line lacking the high-affinity endogenous H+/P(i) cotransporter. The expression of the mRNA molecules and corresponding proteins was confirmed by Northern and Western blot analysis, respectively. As detected by indirect immunofluorescence and antibody capture assay, both wild-type and mutant NaPi-2 proteins are expressed in the yeast plasma membrane in comparable amounts. In the presence of 5 microM phosphate, Na+ promotes phosphate uptake into yeast cells expressing the wild-type NaPi-2 with a K(0.5) of 5.6 +/- 1.1 mM. The maximum uptake of phosphate (649 +/- 30 pmol/10 min) is approximately 8-fold higher than the uptake obtained with nontransformed cells (76.8 +/- 8 pmol/10 min). Yeast cells expressing the R215E mutant of NaPi-2 accumulate 213 +/- 9 pmol of phosphate/10 min under the same conditions. The K(0.5) for the stimulation of phosphate uptake by Na+ is 4.2 +/- 0.8 mM for the R215E mutant and thus not significantly different from the value obtained with cells expressing the wild-type cotransporter. The reduced level of accumulation of phosphate in yeast cells expressing the R215E mutant is probably due to a reduction of the first-order rate constant k for phosphate uptake: while cells expressing wild-type NaPi-2 accumulate phosphate with a k of 0.06 min(-1), the rate for phosphate uptake into cells expressing the R215E mutant (k) is 0.016 min(-1) and therefore about 4-fold lower. In comparison, the rate for phosphate uptake into nontransformed cells (k) is 0.0075 min(-1). Phosphate uptake into yeast cells that express the wild-type NaPi-2 in the presence of 150 mM NaCl is promoted by extracellular phosphate with a K(0.5) of 45 +/- 4 microM. A phosphate-dependent phosphate accumulation is also observed with cells expressing the R215E mutant, but the K(0.5) is twice as high (86 +/- 5 microM) as that obtained with the wild-type cotransporter. We conclude that the yeast expression system is a useful tool for the investigation of structure-function relationships of the renal sodium/phosphate cotransporter and that (215)R, although not involved in Na+ recognition, is a part of the structure involved in phosphate recognition and considerably influences the rate of phosphate uptake by the NaPi-2 cotransporter.  相似文献   

11.
1. Proton uptake into the yeast Saccharomyces carlsbergensis, was studied at pH4.5-5.5 in the presence of both antimycin and 2-deoxyglucose to inhibit energy metabolism. Previous work had shown that the cells then absorbed about 20nmol of glycine or l-phenylalanine against a considerable amino acid concentration gradient. The addition of the amino acid immediately stimulated the rate of uptake of protons two- to three-fold. About 2 extra equivalents of H(+) accompanied a given amount of the amino acids into the yeast preparations exposed to the metabolic inhibitors for 2-4min and about 1.2 equivalents after 20min exposure. 2. Analogous observations were made during serial additions of glycine, l-phenylalanine, l-leucine and l-lysine to preparations lacking the metabolic inhibitors and deficient in substrates needed for energy metabolism. In fresh cellular preparations the influx of glycine was then closely coupled to a stimulated flow of 2.1 equiv. of H(+) into the yeast. A similar number of K(+) ions left the cells. About 30% of the extra protons was subsequently ejected from the yeast. Deoxyglucose and antimycin together inhibited the ejection of protons. When the yeast had been fed with glucose energy metabolism was stimulated and almost as many protons as were absorbed with the amino acid were apparently ejected again. 3. Yeast preparations containing Na(+), instead of K(+), as the principal cation absorbed about 1 extra equivalent of H(+) after the addition of phenylalanine, glycine or leucine. This response was not observed in the presence of both deoxyglucose and antimycin. 4. The observations show that H(+) and, in certain circumstances, K(+) are co-substrates in the transport of the amino acids into the yeast. An analogy is drawn with the roles of Na(+) and K(+) as co-substrates in certain mammalian systems. The results lead to various models relating the physical flow of the co-substrate ions on the amino acid carrier to the transduction of chemical energy in an associated ion pump forming part of the mechanism for transporting amino acids into the yeast.  相似文献   

12.
Identification of a high affinity NH4+ transporter from plants.   总被引:17,自引:1,他引:17       下载免费PDF全文
  相似文献   

13.
Cadmium (Cd(2+)) or copper (Cu(2+)) ions are toxic for Chlamydomonas reinhardtii growth, at 300 microM, and the alga may accumulate about 0.90+/-0.02 and 0.64+/-0.02% of its dry weight, respectively. Metal contamination changes the elemental composition of dried alga biomass, which indicates the possibility to use C. reinhardtii as biosensor and bioremediator of the aquatic contamination by heavy metals. Either, Cd(2+) or Cu(2+), inhibits about 20% of the nitrate consumption rate by the cells, while only Cd(2+) increases about 40% the sulfate consumption rate. The presence of 1 mM calcium (Ca(2+)) in the culture medium increases the C. reinhardtii productivity (about 50%), the nitrate uptake rate (about 20%) and the sulfate uptake rate (about 30%). In addition, Ca(2+) overcomes the Cd(2+) (300 microM) toxicity by decreasing (about 35%) the intracellular accumulation of metal. Sulfur-starvation induces in C. reinhardtii the expression of serine acetyltransferase and O-acetylserine(thiol)lyase activities, but decreases 50% the consumption rate of nitrate by the cells. Sulfate is also required for the full expression of the nitrate reductase (NR), nitrite reductase (NiR) and glutamate synthase activities.  相似文献   

14.
Plant uptake rates of nitrate and potassium (K) are under feedback control from plant concentrations of nitrogen (N) and K, respectively. However, there is uncertainty concerning the interactions between nitrate and K uptake. We tested the hypothesis that plant concentrations of N affect K uptake and plant concentrations of K affect uptake rates of nitrate. Two experiments were carried out with wheat. Each consisted of three phases. In Phase 1, the plants were grown in complete nutrient solution for 17–18 days. In Phase 2, nitrate and K treatments were imposed – the plants were either starved of the nutrient or provided excess for 2.5 or 5 days. This generated plants of similar size but with different N and K concentrations. In Phase 3, complete nutrient solutions were restored for all plants and nitrate and K uptake was followed for 8 h. Uptake till the end of Phases 2 and 3 was examined by analysis of variance. Uptake during Phase 3 was also examined using a simple model that included Michaelis‐Menten uptake kinetics and feedback control due to whole‐plant N and K concentrations, but assumed no interaction between nitrate and K uptake. The model was fitted using Phase 3 measurements of nitrate uptake following nitrate starvation or excess and K uptake following K starvation or excess. There was little influence of plant N on K uptake or of plant K on nitrate uptake for periods that started with adequate plant concentrations of N and K (Phases 1 and 2). However, in Phase 3 when N starvation was broken, the plants took up more K than controls and model predictions, so feedback control of K uptake had been lessened. Similarly, when K starvation was broken, feedback control of nitrate uptake was lessened. If plants had endured both N and K starvation then on release of that they took up more K but less nitrate than predicted by the model. These results support the hypothesis under test, and suggest interactions between the mechanisms that regulate uptake of nitrate and K.  相似文献   

15.
An ammonium transporter cDNA, named TbAMT1, was isolated from the ectomycorrhizal ascomycetous truffle Tuber borchii. The polypeptide encoded by TbAMT1 (52 kDa) functionally complements ammonium uptake-defective yeast mutants and shares sequence similarity with previously characterized ammonium transporters from Saccharomyces (Mep) and Arabidopsis (AtAMT1). Structural characteristics common to the Mep/Amt family and peculiar features of the Tuber transporter have been evidenced by a detailed topological model of the TbAMT1 protein, which predicts 11 transmembrane helices with an N terminus(OUT)/C terminus(IN) orientation. As revealed by uptake/competition experiments conducted in yeast, TbAMT1 is a high-affinity transporter with an apparent K(m) for ammonium of 2 microM. The TbAMT1 mRNA was very slowly, yet specifically upregulated in nitrogen-deprived T. borchii mycelia. Instead, a much faster return to basal expression levels was observed upon resupplementation of either ammonium or nitrate, which thus appear to be utilized as equally effective nitrogen sources by Tuber mycelia.  相似文献   

16.
In DCMU‐poisoned wild‐type and in non‐photosynthetic pigment mutant cells of Chlorella kessleri , grown heterotrophically with glucose as a carbon source and with nitrate as sole nitrogen source, the known blue light‐enhanced uptake of oxygen and breakdown of starch were reduced by staurosporine and K252a, both potent inhibitors of protein kinase C. This corresponded to sensitivity to these inhibitors of blue light‐induced uptake of nitrate of such organisms. Cells grown with ammonia as sole nitrogen source responded to short wavelength visible irradiation with an increase in oxygen uptake, and this, too, was inhibited by staurosporine and K252a. However, these cells did not show any blue light‐enhanced uptake of nitrate. From these results, enhanced consumption of oxygen under blue light cannot be a consequence of blue light‐induced protein phosphorylation involved in the light‐dependent uptake of nitrate. However, existence of a specific protein phosphorylation within the process of enhancement of oxygen uptake under blue light is not yet proven by the data. There might be a master reaction that induces both processes independently, or there may be influences of other light‐induced processes which lead to enhanced starch breakdown, thereby supplying the glucose for oxidative degradation.  相似文献   

17.
A study has been made of the effects of both varying the pH and extracellular [K(+)] on the initial rate of uptake of glycine (v) by a strain of Saccharomyces carlsbergensis that concentrated the amino acid, with respect to the extracellular phase, by up to 1400 times. When no other substrate than glycine was provided and [glycine] was relatively small (相似文献   

18.
The Arabidopsis CHL1 (AtNRT1) gene confers sensitivity to the herbicide chlorate and encodes a nitrate-regulated nitrate transporter. However, how CHL1 participates in nitrate uptake in plants is not yet clear. In this study, we examined the in vivo function of CHL1 with in vivo uptake measurements and in situ hybridization experiments. Under most conditions tested, the amount of nitrate uptake by a chl1 deletion mutant was found to be significantly less than that of the wild type. This uptake deficiency was reversed when a CHL1 cDNA clone driven by the cauliflower mosaic virus 35S promoter was expressed in transgenic chl1 plants. Furthermore, tissue-specific expression patterns showed that near the root tip, CHL1 mRNA is found primarily in the epidermis, but further from the root tip, the mRNA is found in the cortex or endodermis. These results are consistent with the involvement of CHL1 in nitrate uptake at different stages of root cell development. A functional analysis in Xenopus oocytes indicated that CHL1 is a low-affinity nitrate transporter with a K(m) value of approximately 8.5 mM for nitrate. This finding is consistent with the chlorate resistance phenotype of chl1 mutants. However, these results do not fit the current model of a single, constitutive component for the low-affinity uptake system. To reconcile this discrepancy and the complex uptake behavior observed, we propose a "two-gene" model for the low-affinity nitrate uptake system of Arabidopsis.  相似文献   

19.
AtKuP1: a dual-affinity K+ transporter from Arabidopsis.   总被引:19,自引:0,他引:19       下载免费PDF全文
H H Fu  S Luan 《The Plant cell》1998,10(1):63-73
Plant roots contain both high- and low-affinity transport systems for uptake of K+ from the soil. In this study, we characterize a K+ transporter that functions in both high- and low-affinity uptake. Using yeast complementation analysis, we isolated a cDNA for a functional K+ transporter from Arabidopsis (referred to as AtKUP1 for Arabidopsis thaliana K+ uptake). When expressed in a yeast mutant, AtKUP1 dramatically increased K+ uptake capacity at both a low and high [K+] range. Kinetic analyses showed that AtKUP1-mediated K+ uptake displays a "biphasic" pattern similar to that observed in plant roots. The transition from the high-affinity phase (K(m) of 44 microM) to the low-affinity phase (K(m) of 11 mM) occurred at 100 to 200 microM external K+. Both low- and high-affinity K+ uptake via AtKUP1 were inhibited by 5 mM or higher concentrations of NaCl. In addition, AtKUP1-mediated K+ uptake was inhibited by K+ channel blockers, including tetraethylammonium, Cs+, and Ba2+. Consistent with a possible function in K+ uptake from the soil, the AtKUP1 gene is primarily expressed in roots. We conclude that the AtKUP1 gene product may function as a K+ transporter in Arabidopsis roots over a broad range of [K+] in the soil.  相似文献   

20.
The nitrate-tolerant organism Klebsiella oxytoca CECT 4460 tolerates nitrate at concentrations up to 1 M and is used to treat wastewater with high nitrate loads in industrial wastewater treatment plants. We studied the influence of the C source (glycerol or sucrose or both) on the growth rate and the efficiency of nitrate removal under laboratory conditions. With sucrose as the sole C source the maximum specific growth rate was 0.3 h−1, whereas with glycerol it was 0.45 h−1. In batch cultures K. oxytoca cells grown on sucrose or glycerol were able to immediately use sucrose as a sole C source, suggesting that sucrose uptake and metabolism were constitutive. In contrast, glycerol uptake occurred preferentially in glycerol-grown cells. Independent of the preculture conditions, when sucrose and glycerol were added simultaneously to batch cultures, the sucrose was used first, and once the supply of sucrose was exhausted, the glycerol was consumed. Utilization of nitrate as an N source occurred without nitrite or ammonium accumulation when glycerol was used, but nitrite accumulated when sucrose was used. In chemostat cultures K. oxytoca CECT 4460 efficiently removed nitrate without accumulation of nitrate or ammonium when sucrose, glycerol, or mixtures of these two C sources were used. The growth yields and the efficiencies of C and N utilization were determined at different growth rates in chemostat cultures. Regardless of the C source, yield carbon (YC) ranged between 1.3 and 1.0 g (dry weight) per g of sucrose C or glycerol C consumed. Regardless of the specific growth rate and the C source, yield nitrogen (YN) ranged from 17.2 to 12.5 g (dry weight) per g of nitrate N consumed. In contrast to batch cultures, in continuous cultures glycerol and sucrose were utilized simultaneously, although the specific rate of sucrose consumption was higher than the specific rate of glycerol consumption. In continuous cultures double-nutrient-limited growth appeared with respect to the C/N ratio of the feed medium and the dilution rate, so that for a C/N ratio between 10 and 30 and a growth rate of 0.1 h−1 the process led to simultaneous and efficient removal of the C and N sources used. At a growth rate of 0.2 h−1 the zone of double limitation was between 8 and 11. This suggests that the regimen of double limitation is influenced by the C/N ratio and the growth rate. The results of these experiments were validated by pulse assays.  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