Monday 18 June 2018

MICROBE'S NUTRIENT UPTAKE WHILE DORMANCY


MICROBES NUTRIENT UP TAKE WHILE DORMANCY
Common Features of Quiescent Cells
Carbon Storage An almost universal property of quiescent cells is the accumulation of carbon stores, although the chemical structure of the storage form can differ. During low growth states, the yeast Saccharomyces cerevisiae accumulates glycogen, trehalose, and triglycerides as the main forms of metabolizable carbon (Gray et al., 2004).
The bacterial pathogen Vibrio cholerae accumulates glycogen in preparation for survival in nutrient-poor environments (Bourassa and Camilli, 2009). Additionally, many bacteria store fatty acids in the form of triglycerides (Daniel et al., 2004; Kalscheuer et al., 2007) and wax esters (Sirakova et al., 2012). Both triglycerides and wax esters also accumulate in plant seeds (Radunz and Schmid, 2000), indicating that this mode of storage is advantageous for organisms that represent vastly separated domains of life.
In addition, linear plastic polymers like polyhydroxyalkanoates and poly-b-hydroxybutyric acid can serve as a carbon repository in a variety of bacteria living in the soil and the rhizosphere (Kadouri et al., 2005).
What is the purpose of carbon storage? The most intuitive answer is that these cells are simply ‘‘storing nuts for winter,’’ and these nutritional stores can be rapidly mobilized to fuel growth when environmental conditions improve. This role has been most clearly demonstrated in the S. cerevisiae cell, where the trehalose stores that accumulate in stationary cultures are immediately consumed upon addition of fresh media to fuel rapid regrowth (Shi et al., 2010).
Glycogen may serve a similar role in V. cholerae, a bacterium whose life cycle relies on periodic switches from the nutrient-replete mammalian gut to nutrient poor aquatic environments (Bourassa and Camilli, 2009). Carbon storage has also been found to play an important role in remodeling cellular carbon fluxes and facilitating entry into the quiescent state.
 Diverse stresses, such as low oxygen, low pH, or low iron, all induce a storage response in M. tuberculosis through the activation of a common sensor-kinase system, DosRST (Bacon et al., 2007; Baek et al., 2011; Daniel et al., 2011). The DosS sensor likely responds to alterations in cellular redox state in these contexts (Honaker et al., 2010), and triggers the synthesis of triglycerides that are stored in large cytosolic inclusions (Garton et al., 2002). The impact of this response appears to extend beyond the generation of nutrient stores. That is, disruption of the triglyceride biosynthesis pathway in M. tuberculosis reverses the growth arrest that is normally caused by these stresses, but has little effect on the subsequent recovery of growth when the stress is relieved (Baek et al., 2011). This inverse relationship between growth and triglyceride production appears to result from the redirection of acetyl-CoA from the TCA cycle, where it is used to generate energy during aerobic respiration, into lipid synthesis, where acetyl CoA serves as a building block for fatty acids. The growth-limiting effect of carbon storage is unlikely to be restricted to mycobacteria. For example, S. cerevisiae mutants that are unable to produce glycogen or trehalose consume more CO2 than the wild-type strain during slow growth (Sillje´ et al., 1999), indicating higher TCA flux in the absence of carbon storage. The almost universal propensity of microorganisms to accumulate acetyl CoA-derived carbon stores under growth-limiting stresses suggests that this may represent a common strategy for reducing growth and metabolic rate.
Cell Wall Modification
 Virtually all bacteria are surrounded by an elastic meshwork of peptidoglycan that maintains cellular integrity under changing environmental conditions. This structure is composed of glycan chains, consisting of N-acetylglucosamine (NAG) and N-acetylmuramic acid (NAM), crosslinked through short peptide moieties. Not surprisingly, the long-term survival of both spores and quiescent cells depends on specific alterations in the composition of this structure. For example, in stationary phase cultures, the Gram-positive bacteria Staphylococcus aureus generates a cell wall that is structurally different from the peptidoglycan found during exponential phase growth, in that it contains fewer pentaglycine bridges, which crosslink the glycan chains, and is significantly thicker (Zhou and Cegelski, 2012).
Similarly, the level and gradient of crosslinking are important for the formation of bacterial spores. In the spore peptidoglycan layer of the soil-dwelling bacteria Bacillus subtilis, the peptide side chains serving as crosslinkers are completely or partially removed from the NAM residues and replaced by muramic-dlactam, a specificity determinant for germination autolytic enzymes, at every second NAM position in the cortex glycan strands. As a consequence, overall levels of crosslinking are markedly decreased in the spore cortex as compared to the vegetative cell wall (Atrih et al., 1996). Thus, common features of the peptidoglycan in both quiescent cells and spores are reduced crosslinks and increased peptidoglycan mass.
The regulation of these modifications is likely complex, but recent observations suggest that extracellular D-amino acids, such as D-methionine and D-leucine, could play an important role. D-amino acids accumulate to millimolar levels in the supernatants of stationary phase bacterial culture, where they regulate cell wall synthetic enzymes and are incorporated into the peptidoglycan polymer. The increased abundance of D-amino acids in cultures of nongrowing cells and their ability to alter the osmotic sensitivity of V. cholerae (Lam et al., 2009) suggests a likely role in remodeling the cell wall for quiescence. During exponential growth, M. tuberculosis peptidoglycan is crosslinked largely via linkages between the third and fourth amino acids in the stem peptide, the chain of amino acids in peptidoglycan that crosslinks adjacent strands (i.e., 4/3 linkages).
In addition to its structural roles, cell wall metabolism also appears to play an important role in generating signals that regulate the germination of spores and the exit from quiescence.
It may seem intuitive that RNA and protein synthesis will proceed at negligible rates in the quiescent cell. Protein turnover increases 5-fold in famished E. coli cells due to proteases that are produced in early stationary phase.
Indeed, quiescence in S. cerevisiae is accompanied by a 3- to 5-fold decrease in overall transcription rate (Choder, 1991), and a 20-fold decrease in protein synthesis (Fuge et al., 1994). The same analysis has not been performed on nonreplicating cells, and it remains likely that both initiation and elongation rate slow.
Energetics and Metabolism during Quiescence
Maintenance of membrane potential and ATP synthesis is not required for sustaining the viability of spores, even though a repertoire of ATPases and ATP-dependent regulatory proteins is utilized during the initiation of germination (Errington, 2003). In contrast, quiescent bacteria maintain their membrane potential (Pernthaler and Amann, 2004; Rao et al., 2008), and energy homeostasis appears to be critical for survival. In nonreplicating M. tuberculosis cells starved for oxygen or nutrients, ATP levels are maintained at a steady level, which is only 5-fold lower than replicating cells (Gengenbacher et al., 2010; Rao et al., 2008). This maintenance of ATP homeostasis is clearly important, as disruption of the proton motive force or chemical inhibition of the F0F1 ATP synthase involved in ATP synthesis induces cell death in nutrient-starved or hypoxic cultures (Rao et al., 2008; Sala et al., 2010). Diverse strategies can be used to maintain energy homeostasis.
Preservation of Genome Integrity
Maintaining genome fidelity when little or no metabolic capacity is available for canonical DNA repair mechanisms is a challenge faced by both quiescent cells and dormant spores. One strategy common to both types of cells is altering chromosomal structure to a more chemically stable form. The chromosome of stationary phase E. coli assumes an extremely compact structure. A nucleoid-associated protein called Dps, which is expressed only in stationary phase, mediates biocrystallization of the nucleoid and protects DNA from damage (Martinez and Kolter, 1997). This compaction of DNA can be very dynamic as bacteria enter and exit different growth states.
In the photosynthetic cyanobacterium, Synechococcus elongates, a circadian clock controlled mechanism induces periodic chromosome compaction during the night (Smith and Williams, 2006), and the resulting alterations in DNA supercoiling control global gene expression patterns (Vijayan et al., 2009). M. tuberculosis might use a similar mechanism to protect its chromosome.
Truly dormant spores are not able to actively maintain their chromosome but depend on the induction of DNA repair systems upon exit from the dormant state. M. tuberculosis, exit the cell cycle with two chromosomal copies (Wayne, 1977). Thus, high-fidelity recombinational repair mechanisms, which often dominate in growing cells, are only available to a subset of quiescent organisms. Despite the apparent presence of a recombinational template in nonreplicating M. tuberculosis, this organism still appears to utilize more error-prone repair systems. For example, error-prone translation polymerases, which replicate past DNA damage lesions, are important for the survival of slowly growing M. tuberculosis in chronically infected animals (Boshoff et al., 2003).
NOTE:Content compiled from Web content

No comments:

Post a Comment

 https://onedrive.live.com/edit.aspx?resid=6BF90612664526FF!520&cid=6bf90612664526ff&CT=1698476756796&OR=ItemsView https://onedr...