首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
This paper describes the metabolic adaptation of Lactococcus lactis during the transition from a growing to a non‐growing state using retentostat cultivation. Under retentostat cultivation, the specific growth rate decreased from 0.025 h?1 to 0.0001 h?1 in 42 days, while doubling time increased to more than 260 days. Viability of the overall culture was maintained above 90% but included approximately 20% damaged cells, which had lost their colony forming capacity on solid media. Although culture biomass and viability had reached a steady‐state after 14 days of retentostat cultivation, the morphology of the cells changed from coccus‐to‐rod shape at later stages of retentostat cultivation, by which the cell's surface to volume ratio was estimated to increase 2.4‐fold. Furthermore, the metabolic patterns switched between homolactic and mixed‐acid fermentation during the retentostat cultivation. Retentostat cultivation enabled the calculation of accurate substrate‐ and energy‐related maintenance coefficients and biomass yields under non‐growing conditions, which were in good agreement with those calculated by extrapolation from chemostat cultivations at high dilution rates. In this study, we illustrate how retentostat cultivation allows decoupling of growth and non‐growth associated processes in L. lactis, enabling the analysis of quantitative physiological responses of this bacterium to near zero‐specific growth rates.  相似文献   

2.
3.
4.
Fourteen genes encoding putative secondary amino acid transporters were identified in the genomes of Lactococcus lactis subsp. cremoris strains MG1363 and SK11 and L. lactis subsp. lactis strains IL1403 and KF147, 12 of which were common to all four strains. Amino acid uptake in L. lactis cells overexpressing the genes revealed transporters specific for histidine, lysine, arginine, agmatine, putrescine, aromatic amino acids, acidic amino acids, serine, and branched-chain amino acids. Substrate specificities were demonstrated by inhibition profiles determined in the presence of excesses of the other amino acids. Four knockout mutants, lacking the lysine transporter LysP, the histidine transporter HisP (formerly LysQ), the acidic amino acid transporter AcaP (YlcA), or the aromatic amino acid transporter FywP (YsjA), were constructed. The LysP, HisP, and FywP deletion mutants showed drastically decreased rates of uptake of the corresponding substrates at low concentrations. The same was observed for the AcaP mutant with aspartate but not with glutamate. In rich M17 medium, the deletion of none of the transporters affected growth. In contrast, the deletion of the HisP, AcaP, and FywP transporters did affect growth in a defined medium with free amino acids as the sole amino acid source. HisP was essential at low histidine concentrations, and AcaP was essential in the absence of glutamine. FywP appeared to play a role in retaining intracellularly synthesized aromatic amino acids when these were not added to the medium. Finally, HisP, AcaP, and FywP did not play a role in the excretion of accumulated histidine, glutamate, or phenylalanine, respectively, indicating the involvement of other transporters.  相似文献   

5.
In Gouda and Cheddar type cheeses the amino acid conversion to aroma compounds, which is a major process for aroma formation, is essentially due to lactic acid bacteria (LAB). In order to evaluate the respective role of starter and nonstarter LAB and their interactions in cheese flavor formation, we compared the catabolism of phenylalanine, leucine, and methionine by single strains and strain mixtures of Lactococcus lactis subsp. cremoris NCDO763 and three mesophilic lactobacilli. Amino acid catabolism was studied in vitro at pH 5.5, by using radiolabeled amino acids as tracers. In the presence of α-ketoglutarate, which is essential for amino acid transamination, the lactobacillus strains degraded less amino acids than L. lactis subsp. cremoris NCDO763, and produced mainly nonaromatic metabolites. L. lactis subsp. cremoris NCDO763 produced mainly the carboxylic acids, which are important compounds for cheese aroma. However, in the reaction mixture containing glutamate, only two lactobacillus strains degraded amino acids significantly. This was due to their glutamate dehydrogenase (GDH) activity, which produced α-ketoglutarate from glutamate. The combination of each of the GDH-positive lactobacilli with L. lactis subsp. cremoris NCDO763 had a beneficial effect on the aroma formation. Lactobacilli initiated the conversion of amino acids by transforming them mainly to keto and hydroxy acids, which subsequently were converted to carboxylic acids by the Lactococcus strain. Therefore, we think that such cooperation between starter L. lactis and GDH-positive lactobacilli can stimulate flavor development in cheese.  相似文献   

6.
Aminotransferases, which catalyze the last step of biosynthesis of most amino acids and the first step of their catabolism, may be involved in the growth of Lactococcus lactis in milk. Previously, we isolated two aminotransferases from L. lactis, AraT and BcaT, which are responsible for the transamination of aromatic amino acids, branched-chain amino acids, and methionine. In this study, we demonstrated that double inactivation of AraT and BcaT strongly reduced the growth of L. lactis in milk. Supplementation of milk with amino acids and keto acids that are substrates of both aminotransferases did not improve the growth of the double mutant. On the contrary, supplementation of milk with isoleucine or a dipeptide containing isoleucine almost totally inhibited the growth of the double mutant, while it did not affect or only slightly affected the growth of the wild-type strain. These results suggest that AraT and BcaT play a major role in the growth of L. lactis in milk by degrading the intracellular excess isoleucine, which is responsible for the growth inhibition. The growth inhibition by isoleucine is likely to be due to CodY repression of the proteolytic system, which is necessary for maximal growth of L. lactis in milk, since the growth of the CodY mutant was not affected by addition of isoleucine to milk. Moreover, we demonstrated that AraT and BcaT are part of the CodY regulon and therefore are regulated by nutritional factors, such as the carbohydrate and nitrogen sources.  相似文献   

7.
Growth at near-zero specific growth rates is a largely unexplored area of yeast physiology. To investigate the physiology of Saccharomyces cerevisiae under these conditions, the effluent removal pipe of anaerobic, glucose-limited chemostat culture (dilution rate, 0.025 h−1) was fitted with a 0.22-μm-pore-size polypropylene filter unit. This setup enabled prolonged cultivation with complete cell retention. After 22 days of cultivation, specific growth rates had decreased below 0.001 h−1 (doubling time of >700 h). Over this period, viability of the retentostat cultures decreased to ca. 80%. The viable biomass concentration in the retentostats could be accurately predicted by a maintenance coefficient of 0.50 mmol of glucose g−1 of biomass h−1 calculated from anaerobic, glucose-limited chemostat cultures grown at dilution rates of 0.025 to 0.20 h−1. This indicated that, in contrast to the situation in several prokaryotes, maintenance energy requirements in S. cerevisiae do not substantially change at near-zero specific growth rates. After 22 days of retentostat cultivation, glucose metabolism was predominantly geared toward alcoholic fermentation to meet maintenance energy requirements. The strict correlation between glycerol production and biomass formation observed at higher specific growth rates was not maintained at the near-zero growth rates reached in the retentostat cultures. In addition to glycerol, the organic acids acetate, d-lactate, and succinate were produced at low rates during prolonged retentostat cultivation. This study identifies robustness and by-product formation as key issues in attempts to uncouple growth and product formation in S. cerevisiae.Laboratory studies on microorganisms are often performed in batch cultures. During the initial phase of batch cultivation, all nutrients are usually present in excess. As a consequence, the initial specific growth rate, μ, of the microorganism in such cultures equals the maximum specific growth rate, μmax. In natural environments, the specific growth rate of microorganisms is likely to be constrained by the limited availability of one or more growth-limiting nutrients, resulting in specific growth rates far below μmax (8, 24). In chemostat cultures fed with a medium containing a single growth-limiting nutrient, the dilution rate determines the specific growth rate. Chemostat cultivation therefore offers the possibility to study microbial physiology at carefully controlled, submaximal specific growth rates and to investigate the effect of specific growth rate on cellular physiology (20). Chemostat cultivation of the yeast Saccharomyces cerevisiae has demonstrated strong effects of specific growth rate on biomass composition (26, 51), product formation (5, 37), and cell size (23). Moreover, during energy-limited growth at low specific growth rates, a relatively large fraction of the energy substrate has to be dissimilated for maintenance-related processes such as maintenance of chemi-osmotic gradients and turnover of cellular components (34). Not surprisingly, recent genome-wide studies have shown strong effects of specific growth rate on levels of mRNAs and proteins (9, 14, 38).In chemostat studies on S. cerevisiae, the steady-state specific growth rate is usually between 0.03 h−1 and 0.40 h−1. While this range is relevant for many industrial applications, there are several incentives to study growth of this yeast at even lower specific growth rates. In many natural environments, growth at a μ of 0.03 h−1, corresponding to a doubling time of 23.1 h, probably still represents extremely fast growth. Furthermore, in industrial applications, S. cerevisiae and other microorganisms can be considered as self-replicating catalysts, and, unless biomass is the desired product, growth can be considered as undesirable by-product formation leading to nonproductive substrate consumption. This problem is further augmented when the excess yeast biomass cannot be valorized because it is genetically modified or has been used for the production of compounds that are not compatible with use as, for example, cattle feed. A third incentive for exploring the physiology of S. cerevisiae at near-zero growth rates is related to the increasing interest in this yeast as a systems biology model for human cells (16, 27, 33). At near-zero growth rates, the age of individual yeast cells becomes much higher than can be achieved in conventional batch or chemostat cultures. Studies on extremely slow growth of S. cerevisiae under defined conditions may therefore provide an interesting model for ageing of human cells.Retentostat cultivation, first proposed by Herbert (18), is a modification of chemostat cultivation that has been specifically designed to study microbial physiology at near-zero specific growth rates. In a retentostat, sometimes referred to as recycling fermentor or recyclostat, the growth-limiting energy substrate is fed at a constant rate, and biomass is retained in the fermentor by an internal filter probe connected to the effluent line or by an external filter module. Prolonged retentostat cultivation should, in theory, result in a situation where the specific growth rate becomes zero and where the specific rate of substrate consumption equals the maintenance energy requirement. This situation is fundamentally different from starvation, which involves deterioration of physiological processes, and from resting states typified by spores, which have little or no metabolic activity. Retentostat cultivation has been applied to several bacterial systems including Escherichia coli (11), Paracoccus denitrificans, and Bacillus licheniformis (49) and the autotrophs Nitrosomonas europaea and Nitrobacter winogradskyi (46, 47). These studies demonstrated that the physiology of these prokaryotes at extremely low specific growth rates could not be accurately predicted by a simple extrapolation of results obtained at higher specific growth rates. In particular, near-zero specific growth rates coincided with increased levels of ppGpp (2), which induces the stringent response, a regulatory program that diverts cellular resources from growth to amino acid biosynthesis (10, 21). Furthermore, it was concluded that extremely slow growth led to a reduction of the maintenance energy requirement of prokaryotes. A recent quantitative analysis on cell retention cultures of S. cerevisiae was performed under severely nitrogen-limited growth conditions and used incomplete cell retention (7), which precluded a quantitative comparison with maintenance energy requirements calculated from energy-limited chemostat cultures.The goal of the present study was to quantitatively analyze the physiology of S. cerevisiae at extremely low specific growth rates in glucose-limited retentostat cultures. To this end, an internal filter probe was introduced in the effluent line of standard laboratory chemostat fermentors and used in long-term cultivation runs with complete cell retention. Anaerobic conditions were chosen to facilitate quantification of catabolic fluxes and growth energetics.  相似文献   

8.
9.
The current knowledge of the physiology and gene expression of industrially relevant microorganisms is largely based on laboratory studies under conditions of rapid growth and high metabolic activity. However, in natural ecosystems and industrial processes, microbes frequently encounter severe calorie restriction. As a consequence, microbial growth rates in such settings can be extremely slow and even approach zero. Furthermore, uncoupling microbial growth from product formation, while cellular integrity and activity are maintained, offers perspectives that are economically highly interesting. Retentostat cultures have been employed to investigate microbial physiology at (near-)zero growth rates. This minireview compares information from recent physiological and gene expression studies on retentostat cultures of the industrially relevant microorganisms Lactobacillus plantarum, Lactococcus lactis, Bacillus subtilis, Saccharomyces cerevisiae, and Aspergillus niger. Shared responses of these organisms to (near-)zero growth rates include increased stress tolerance and a downregulation of genes involved in protein synthesis. Other adaptations, such as changes in morphology and (secondary) metabolite production, were species specific. This comparison underlines the industrial and scientific significance of further research on microbial (near-)zero growth physiology.  相似文献   

10.
11.
Two strains of Lactococcus lactis subsp. lactis were used to determine the influence of lactose and arginine on viability and amino acid use during carbohydrate starvation. Lactose provided energy for logarithmic-phase growth, and amino acids such as arginine provided energy after carbohydrate exhaustion. Survival time, cell numbers, and ATP concentrations increased with the addition of arginine to the basal medium. By the onset of lactose exhaustion, the concentrations of glycine-valine and glutamate had decreased by as much as 67% in L. lactis ML3, whereas the serine concentration increased by 97% during the same period. When no lactose was added, the concentrations of these amino acids remained constant. Similar trends were observed for L. lactis 11454. Without lactose or arginine, L. lactis ML3 was nonculturable on agar but was viable after 2 days, as measured by fluorescent viability stains and intracellular ATP levels. However, L. lactis 11454 without lactose or arginine remained culturable for at least 14 days. These data suggest that lactococci become viable but nonculturable in response to carbohydrate depletion. Additionally, these data indicate that amino acids other than arginine facilitate the survival of L. lactis during carbohydrate starvation.  相似文献   

12.
13.
Specificity of Milk Peptide Utilization by Lactococcus lactis   总被引:1,自引:0,他引:1       下载免费PDF全文
To study the substrate specificity of the oligopeptide transport system of Lactococcus lactis for its natural substrates, the growth of L. lactis MG1363 was studied in a chemically defined medium containing milk peptides or a tryptic digest of αs2-casein as the source of amino acids. Peptides were separated into acidic, neutral, and basic pools by solid-phase extraction or by cation-exchange liquid chromatography. Their ability to sustain growth and the time course of their utilization demonstrated the preferential use of hydrophobic basic peptides with molecular masses ranging between 600 and 1,100 Da by L. lactis MG1363 and the inability to use large, acidic peptides. These peptide utilization preferences reflect the substrate specificity of the oligopeptide transport system of the strain, since no significant cell lysis was inferred. Considering the free amino acid content of milk and these findings on peptide utilization, it was demonstrated that the cessation of growth of L. lactis MG1363 in milk was due to deprivation of leucine and methionine.  相似文献   

14.
Minimal Requirements for Exponential Growth of Lactococcus lactis   总被引:7,自引:0,他引:7       下载免费PDF全文
A minimal growth medium containing glucose, acetate, vitamins, and eight amino acids allowed for growth of Lactococcus lactis subsp. lactis, with a specific growth rate in batch culture of μ = 0.3 h-1. With 19 amino acids added, the growth rate increased to μ = 0.7 h-1 and the exponential growth phase proceeded until high cell concentrations were reached. We show that morpholinepropanesulfonic acid (MOPS) is a suitable buffer for L. lactis and may be applied in high concentrations.  相似文献   

15.
Polymerase chain reaction (PCR) primers designed from a multiple alignment of predicted amino acid sequences from bacterial aroA genes were used to amplify a fragment of Lactococcus lactis DNA. An 8 kb fragment was then cloned from a lambda library and the DNA sequence of a 4.4 kb region determined. This region was found to contain the genes tyrA, aroA, aroK, and pheA, which are involved in aromatic amino acid biosynthesis and folate metabolism. TyrA has been shown to be secreted and AroK also has a signal sequence, suggesting that these proteins have a secondary function, possibly in the transport of amino acids. The aroA gene from L. lactis has been shown to complement an E. coli mutant strain deficient in this gene. The arrangement of genes involved in aromatic amino acid biosynthesis in L. lactis appears to differ from that in other organisms.  相似文献   

16.
Lactococcus lactis subsp. lactis IL1403 was grown in medium containing unlabelled free amino acids and 15N-labelled yeast hydrolysate to gain insight into the role of peptides as a source of amino acids under conditions where free amino acids are abundant. A mathematical model was composed to estimate the fluxes of free and peptide-derived amino acids into and out of the intracellular amino acid pool. We observed co-consumption of peptides and free amino acids and a considerable efflux of most free amino acids during growth. We did not observe significant differences between the peptide consumption patterns of essential and non-essential amino acids, which suggests that the incorporation of a particular amino acid is more dependent on its availability in a readily assimilated form than the organism’s auxotrophy for it. For most amino acids the contribution of peptide-bound forms to the formation of biomass was initially between 30 and 60 % with the remainder originating from free amino acids. During the later stages of fermentation we observed a decrease in the utilization of peptide-bound amino acids, thus indicating that the more readily assimilated peptides are gradually exhausted from the medium during growth.  相似文献   

17.
Efficient microbial cell factory for the production of homoserine from glucose has been developed by iterative and rational engineering of Escherichia coli W3110. The whole pathway from glucose to homoserine was divided into three groups, namely, glucose transport and glycolysis (‘up-stream’), TCA and glyoxylate cycles (‘mid-stream’), and homoserine module (conversion of aspartate to homoserine and its secretion; ‘down-stream’), and the carbon flux in each group as well as between the groups were accelerated and balanced. Altogether, ∼18 genes were modified for active and consistent production of homoserine during both the actively-growing and non-growing stages of cultivation. Finally, fed-batch, two-stage bioreactor experiments, separating the growth from the production stage, were conducted for 61 h, which gave the high titer of 110.8 g/L, yield of 0.64 g/g glucose and volumetric productivity of 1.82 g/L/h, with an insignificant amount of acetate (<0.5 g/L) as the only noticeable byproduct. The metabolic engineering strategy employed in this study should be applicable for the biosynthesis of other amino acids or chemicals derived from aspartic acid.  相似文献   

18.
Amino acid conversion to aroma compounds by Lactococcus lactis is limited by the low production of α-ketoglutarate that is necessary for the first step of conversion. Recently, glutamate dehydrogenase (GDH) activity that catalyzes the reversible glutamate deamination to α-ketoglutarate was detected in L. lactis strains isolated from a vegetal source, and the gene responsible for the activity in L. lactis NCDO1867 was identified and characterized. The gene is located on a 70-kb plasmid also encoding cadmium resistance. In this study, gdh gene inactivation and overexpression confirmed the direct impact of GDH activity of L. lactis on amino acid catabolism in a reaction medium at pH 5.5, the pH of cheese. By using cadmium resistance as a selectable marker, the plasmid carrying gdh was naturally transmitted to another L. lactis strain by a mating procedure. The transfer conferred to the host strain GDH activity and the ability to catabolize amino acids in the presence of glutamate in the reaction medium. However, the plasmid appeared unstable in a strain also containing the protease lactose plasmid pLP712, indicating an incompatibility between these two plasmids.  相似文献   

19.
Lactic acid bacteria are extensively used in food technology and for the production of various compounds, but they are fastidious in nutrient requirements. In order to elucidate the role of each component precisely, defined multicomponent media are required. This study focuses on determining nutrient auxotrophies and minimizing media components (amino acids, vitamins, metal ions, buffers and additional compounds) for the cultivation of Lactococcus lactis subsp. lactis IL1403, using microtitre plates and test tubes. It was shown that glutamine and asparagine were the most important media components for achieving higher biomass yields while the branched-chain amino acids were necessary to increase specific growth rate. The amino acid and glucose ratio was reduced to achieve minimal residual concentration of amino acids in the medium after the growth of cells, whereas the specific growth rate and biomass yield of cells were not considerably affected. As the percentage of each consumed amino acid compared to initial amount is larger than measurement error, these optimized media are important for achieving more precise data about amino acid utilization and metabolism.  相似文献   

20.
Lactococcus lactis is a primary constituent of many starter cultures used for the manufacturing of fermented dairy products, but the species also occurs in various nondairy niches such as (fermented) plant material. Three genome sequences of L. lactis dairy strains (IL-1403, SK11, and MG1363) are publicly available. An extensive molecular and phenotypic diversity analysis was now performed on two L. lactis plant isolates. Diagnostic sequencing of their genomes resulted in over 2.5 Mb of sequence for each strain. A high synteny was found with the genome of L. lactis IL-1403, which was used as a template for contig mapping and locating deletions and insertions in the plant L. lactis genomes. Numerous genes were identified that do not have homologs in the published genome sequences of dairy L. lactis strains. Adaptation to growth on substrates derived from plant cell walls is evident from the presence of gene sets for the degradation of complex plant polymers such as xylan, arabinan, glucans, and fructans but also for the uptake and conversion of typical plant cell wall degradation products such as α-galactosides, β-glucosides, arabinose, xylose, galacturonate, glucuronate, and gluconate. Further niche-specific differences are found in genes for defense (nisin biosynthesis), stress response (nonribosomal peptide synthesis and various transporters), and exopolysaccharide biosynthesis, as well as the expected differences in various mobile elements such as prophages, plasmids, restriction-modification systems, and insertion sequence elements. Many of these genes were identified for the first time in Lactococcus lactis. In most cases good correspondence was found with the phenotypic characteristics of these two strains.  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