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.
Bacterial persistence is a feature that allows susceptible bacteria to survive extreme concentrations of antibiotics and it has been verified in a number of species, such as Escherichia coli, Pseudomonas aeruginosa, Staphylococcus spp., Mycobacterium spp. However, even though Acinetobacter baumannii is an important nosocomial pathogen, data regarding its persistence phenotype are still lacking. Therefore, the aim of this study was to evaluate the persistence phenotype in A. baumannii strains, as well as its variation among strains after treatment with polymyxin B and tobramycin. Stationary cultures of 37 polymyxin B-susceptible clinical strains of A. baumannii were analyzed for surviving cells after exposure to 15 µg/mL of polymyxin B for 6 h, by serial dilutions and colony counting. Among these, the 30 tobramycin-susceptible isolates also underwent tobramycin treatment at a concentration of 160 µg/mL and persister cells occurrence was evaluated equally. A high heterogeneity of persister cells formation patterns among isolates was observed. Polymyxin B-treated cultures presented persister cells corresponding from 0.0007% to 10.1% of the initial population and two isolates failed to produce detectable persister cells under this condition. A high variability could also be observed when cells were treated with tobramycin: the persister fraction corresponded to 0.0003%–11.84% of the pre-treatment population. Moreover, no correlation was found between persister subpopulations comparing both antibiotics among isolates, indicating that different mechanisms underlie the internal control of this phenotype. This is the first report of persister cells occurrence in A. baumannii. Our data suggest that distinct factors regulate the tolerance for unrelated antibiotics in this species, contrasting the multi-drug tolerance observed in other species (eg. dormancy-mediated tolerance). Supporting this observation, polymyxin B – an antibiotic that is believed to act on non-dividing cells as well – failed to eradicate persister cells in the majority of the isolates, possibly reflecting a disconnection between persistence and dormancy.  相似文献   

2.
Bacterial persistence: a model of survival in changing environments   总被引:16,自引:0,他引:16       下载免费PDF全文
Kussell E  Kishony R  Balaban NQ  Leibler S 《Genetics》2005,169(4):1807-1814
The persistence phenotype is an epigenetic trait exhibited by a subpopulation of bacteria, characterized by slow growth coupled with an ability to survive antibiotic treatment. The phenotype is acquired via a spontaneous, reversible switch between normal and persister cells. These observations suggest that clonal bacterial populations may use persister cells, whose slow division rate under growth conditions leads to lower population fitness, as an "insurance policy" against antibiotic encounters. We present a model of Escherichia coli persistence, and using experimentally derived parameters for both wild type and a mutant strain (hipQ) with markedly different switching rates, we show how fitness loss due to slow persister growth pays off as a risk-reducing strategy. We demonstrate that wild-type persistence is suited for environments in which antibiotic stress is a rare event. The optimal rate of switching between normal and persister cells is found to depend strongly on the frequency of environmental changes and only weakly on the selective pressures of any given environment. In contrast to typical examples of adaptations to features of a single environment, persistence appears to constitute an adaptation that is tuned to the distribution of environmental change.  相似文献   

3.
崔鹏  许涛  张文宏  张颖 《遗传》2016,38(10):859-871
持留菌是细菌群体中一小部分具有表型耐药的细菌。自1944年被发现后,近几十年来因其在慢性持续性感染和生物膜感染中的重要作用而得到越来越多的重视。已有的研究结果表明,细菌持留的机理复杂,涉及的相关信号通路有毒素-抗毒素系统、细胞能量代谢及蛋白核酸合成等生理状态的降低、DNA保护修复系统、蛋白酶系统、反式翻译、外排泵系统等。虽然不同细菌的持留机理有一定的相似性和保守性,但不同细菌的持留机制也存在差异,如毒素-抗毒素系统在大肠埃希菌(Escherichia coli)中的过度激活可导致持留菌增加,但在金黄色葡萄球菌(Staphylococcus aureus)中却并无相同作用。本文从持留菌的研究历史出发,综述了当前对革兰氏阴性菌和阳性菌的持留机制方面的研究进展,同时探讨了在持留菌相关感染疾病方面的治疗策略,以期为更好地解决持留菌带来的问题,缩短治疗时间提供新的思路。  相似文献   

4.
Bacterial persister cells are a small population of dormant cells that are tolerant to essentially all antibiotics. Recently, we reported that a quorum sensing (QS) inhibitor, (Z)-4-bromo-5-(bromomethylene)-3-methylfuran-2(5H)-one (BF8), can revert antibiotic tolerance of Pseudomonas aeruginosa persister cells. To better understand this phenomenon, several synthetic brominated furanones with similar structures were compared for their activities in persister control and inhibition of acyl-homoserine lactone (AHL) mediated QS. The results show that some other furanones in addition to BF8 are also AHL QS inhibitors and can revert antibiotic tolerance of P. aeruginosa PAO1 persister cells. However, not all QS inhibiting BFs can revert persistence at growth non-inhibitory concentrations, suggesting that QS inhibition itself is not sufficient for persister control.  相似文献   

5.
Bacterial persisters (defined as dormant, non-dividing cells with globally reduced metabolism) are the major cause of recurrent infections. As they neither grow nor die in presence of antibiotics, it is difficult to eradicate these cells using antibiotics, even at higher concentrations. Reports of metabolites (which help in waking up of these inactive cells) enabled eradication of bacterial persistence by aminoglycosides, suggest the new potential strategy to improve antibiotic therapy. Here we propose, mannitol enabled elimination of Salmonella persister cells by the nisin–antibiotic combination. For this, persister cells were developed and characterized for their typical properties such as non-replicative state and metabolic dormancy. Different carbon sources viz. glucose, glycerol, and mannitol were used, each as an adjunct to ampicillin for the eradication of persister cells. The maximum (but not complete) killing was observed with mannitol–ampicillin, out of all the combinations used. However, significant elimination (about 78%) could be observed, when nisin (an antimicrobial peptide) was used with ampicillin in presence of mannitol, which might have mediated the transfer of antibiotic–nisin combination at the same time when the cells tried to grab the carbon molecule. Further, the effectiveness of the trio was confirmed by flow cytometry. Overall, our findings highlight the potential of this trio-combination for developing it as an option for tackling Salmonella persister cells.  相似文献   

6.
Microbial populations often contain a fraction of slow‐growing persister cells that withstand antibiotics and other stress factors. Current theoretical models predict that persistence levels should reflect a stable state in which the survival advantage of persisters under adverse conditions is balanced with the direct growth cost impaired under favourable growth conditions, caused by the nonreplication of persister cells. Based on this direct growth cost alone, however, it remains challenging to explain the observed low levels of persistence (<<1%) seen in the populations of many species. Here, we present data from the opportunistic human pathogen Pseudomonas aeruginosa that can explain this discrepancy by revealing various previously unknown costs of persistence. In particular, we show that in the absence of antibiotic stress, increased persistence is traded off against a lengthened lag phase as well as a reduced survival ability during stationary phase. We argue that these pleiotropic costs contribute to the very low proportions of persister cells observed among natural P. aeruginosa isolates (3 × 10?8–3 × 10?4) and that they can explain why strains with higher proportions of persister cells lose out very quickly in competition assays under favourable growth conditions, despite a negligible difference in maximal growth rate. We discuss how incorporating these trade‐offs could lead to models that can better explain the evolution of persistence in nature and facilitate the rational design of alternative therapeutic strategies for treating infectious diseases.  相似文献   

7.
Persister cells and tolerance to antimicrobials   总被引:26,自引:0,他引:26  
Bacterial populations produce persister cells that neither grow nor die in the presence of microbicidal antibiotics. Persisters are largely responsible for high levels of biofilm tolerance to antimicrobials, but virtually nothing was known about their biology. Tolerance of Escherichia coli to ampicillin and ofloxacin was tested at different growth stages to gain insight into the nature of persisters. The number of persisters did not change in lag or early exponential phase, and increased dramatically in mid-exponential phase. Similar dynamics were observed with Pseudomonas aeruginosa (ofloxacin) and Staphylococcus aureus (ciprofloxacin and penicillin). This shows that production of persisters depends on growth stage. Maintaining a culture of E. coli at early exponential phase by reinoculation eliminated persisters. This suggests that persisters are not at a particular stage in the cell cycle, neither are they defective cells nor cells created in response to antibiotics. Our data indicate that persisters are specialized survivor cells.  相似文献   

8.
The ability of a high frequency (10(-2)) of Escherichia coli to survive prolonged exposure to penicillin antibiotics, called high persistence, is associated with mutations in the hipA gene. The hip operon is located in the chromosomal terminus near dif and consists of two genes, hipA and hipB. The wild-type hipA gene encodes a toxin, whereas hipB encodes a DNA-binding protein that autoregulates expression of the hip operon and binds to HipA to nullify its toxic effects. We have characterized the hipA7 allele, which confers high persistence, and established that HipA7 is non-toxic, contains two mutations (G22S and D291A) and that both mutations are required for the full range of phenotypes associated with hip mutants. Furthermore, expression of hipA7 in the absence of hipB is sufficient to establish the high persistent phenotype, indicating that hipB is not required. There is a strong correlation between the frequency of persister cells generated by hipA7 strains and cell density, with hipA7 strains generating a 20-fold higher frequency of persisters as cultures approach stationary phase. It is also demonstrated that relA knock-outs diminish the high persistent phenotype in hipA7 mutants and that relA spoT knock-outs eliminate high persistence altogether, suggesting that hipA7 facilitates the establishment of the persister state by inducing (p)ppGpp synthesis. Consistent with this proposal, ectopic expression of relA' from a plasmid was shown to increase the number of persistent cells produced by hipA7 relA double mutants by 100-fold or more. A model is presented that postulates that hipA7 increases the basal level of (p)ppGpp synthesis, allowing a significantly greater percentage of cells in a population to assume a persistent, antibiotic-insensitive state by potentiating a rapid transition to a dormant state upon application of stress.  相似文献   

9.
Pseudomonas aeruginosa is a major pathogen causing chronic pulmonary infections; for example, 80% of cystic fibrosis patients get infected by this bacterium as the disease progresses. Such chronic infections are challenging because P. aeruginosa exhibits high-level tolerance to antibiotics by forming biofilms (multicellular structures attached to surfaces), by entering dormancy and forming antibiotic tolerant persister cells, and by conversion to the mucoid phenotype. Recently, we reported that a synthetic quorum sensing inhibitor, (Z)-4-bromo-5-(bromomethylene)-3-methylfuran-2(5H)-one (BF8), can sensitize both planktonic and biofilm-associated persister cells of P. aeruginosa PAO1 to antibiotics at the concentrations non-inhibitory to its growth. In this study, we further characterized the effects of this compound on the mucoid strain P. aeruginosa PDO300. BF8 was found to reduce persistence during the growth of PDO300 and effectively kill the persister cells isolated from PDO300 cultures. In addition to planktonic cells, BF8 was also found to inhibit biofilm formation of PDO300 and reduce associated persistence. These findings broaden the activities of this class of compounds and indicate that BF8 also has other targets in P. aeruginosa in addition to quorum sensing.  相似文献   

10.
Bacterial cells may escape the effects of antibiotics without undergoing genetic change; these cells are known as persisters. Unlike resistant cells that grow in the presence of antibiotics, persister cells do not grow in the presence of antibiotics. These persister cells are a small fraction of exponentially growing cells (due to carryover from the inoculum) but become a significant fraction in the stationary phase and in biofilms (up to 1%). Critically, persister cells may be a major cause of chronic infections. The mechanism of persister cell formation is not well understood, and even the metabolic state of these cells is debated. Here, we review studies relevant to the formation of persister cells and their metabolic state and conclude that the best model for persister cells is still dormancy, with the latest mechanistic studies shedding light on how cells reach this dormant state.  相似文献   

11.
12.
Burkholderia pseudomallei is the causative agent of melioidosis, a disease with a mortality rate of up to 40% even with treatment. Despite the ability of certain antibiotics to control initial infection, relapse occurs in treated patients. The inability of antibiotics to clear this bacterial infection is in part due to persistence, an evasion mechanism against antibiotics and the effect of host defenses. Evaluation of antibiotic efficacy against B. pseudomallei revealed that up to 48% of in vitro grown populations can survive in a persister state. Toxin–antitoxin (TA) systems have been previously implicated in modulating bacterial persistence. We generated three isogenic TA mutants and found that loss of each toxin gene did not alter antibiotic persistence or macrophage survival. In response to macrophage‐induced persistence, all three toxin mutants demonstrated increased intracellular susceptibility to levofloxacin which in part was due to the inability of the mutants to induce persistence after nitric oxide or nutrient starvation. In an inhalational model of murine melioidosis, both ΔBPSS0395 and ΔBPSS1584 strains were attenuated, and treatment with levofloxacin led to significant reduction in lung colonisation and reduced splenic colonisation by ΔBPSS0395. Based on our findings, these toxins deserve additional evaluation as putative therapeutic targets.  相似文献   

13.
Similar to persister bacterial cells that survive antibiotic treatments, some cancer cells can evade drug treatments. This Commentary discusses the different classes of persister cells and their implications for developing more efficient cancer treatments. Subject Categories: Cancer

Similar to persister bacterial cells that survive antibiotic treatments, small populations of cancer cells can evade drug treatments and cause recurrent disease. This Commentary discusses the different classes of persister cells and their implications for developing more efficient cancer treatments.In 1944, Joseph Bigger, a lieutenant‐colonel in the British Royal Army Medical Corps, reported a peculiar population of bacteria that could survive very high concentrations of penicillin (Bigger, 1944). He termed these hard‐to‐kill cells “persisters” and argued they might explain the limited success of penicillin in curing infections. At the time, 16 years after antibiotics revolutionized bacterial infection treatment, this was a groundbreaking hypothesis as it was largely believed that partial killing was mostly due to inadequate blood supply or tissue barriers. Later on, the understanding that cell‐intrinsic properties may contribute to transient drug tolerance sparked research aimed at targeting microbial persister cells. In a seminal paper, Sherma and colleagues (Sharma et al, 2010) showed that reversible cell‐intrinsic resistance can also be observed in cancer cells in response to therapy. Similar to bacterial persisters, these cancer persister cells gave rise to a drug‐sensitive cell progeny following a short “drug‐holiday” and did not harbor any known resistance‐mediating alteration mutation. However, in contrast to microbial persisters that are largely dormant, a small fraction of cancer persister cells were able to resume proliferation even under continued drug treatment. Understanding the similarities and differences between cancer and microbial persister cells is pivotal to devise approaches to eliminate them (Fig 1).Open in a separate windowFigure 1Different persister classes(A) Classic persisters, (B) targeted‐persisters, and (C) immune‐persisters. The mechanism of escape is dependent on the mode of action of the drug. While classical persisters are common to both bacteria and cancer cells, other persister classes are cancer‐specific and are associated with the ability of cancer cells to probe a wide range of cells states and lineage trajectories.So why can some bacteria persist in the face of therapy? The answer largely lays in the mode of action of antimicrobial drugs. Penicillin and newer generation antibiotics target bacterial cell division. As such, if the bacteria are dormant or reside in a low metabolic state, they are unafflicted by the drug. Dormant bacteria are frequently resistant to multiple stressors and drugs making them difficult to eradicate even with a very aggressive treatment. Unsurprisingly, similar phenomena are observed in the context of chemotherapy treatments in cancer. Like antibiotics, early cancer therapies were largely based on drugs that target highly proliferative cells. Sustained proliferation in the absence of external stimuli is one of the hallmarks of cancer. Because cancer cells divide more frequently than most normal cells, they are more likely to be killed by chemotherapy treatment. As both antibiotics and chemotherapy treatments target proliferating cells, it is not surprising that cell dormancy was linked to cell persistence in both cases. “Classical” nondividing persister cells have been implicated in treatment failure both in cancer and in microbial infections and are thought to provide a reservoir for subsequent relapse events.In the last 20 years, a new class of cancer drugs, called targeted therapies, have emerged and revolutionized patient care. Unlike chemotherapies or antibiotics, these drugs do not target proliferating cell per se but rather act on specific molecular targets associated with cancer. For example, some targeted therapies target proteins that are more abundant on the surface of cancer cells compared with that of normal cells. While slow proliferation has also been implicated in tolerance in the context of targeted therapy, multiple additional mechanisms are at play, which are not characteristic of microbial persister cells. For instance, oncogene‐targeted therapies are taking advantage of the acquired dependence of a cancer cell on the activity of a single oncogenic gene product. As many oncogenes control cell metabolism (Levine & Puzio‐Kuter, 2010), for example by regulating glucose uptake, drugs that target oncogene addiction can have profound effects on metabolism. In line with this, oncogenic‐persisters, for example, persisters that escape killing by oncogene‐targeted therapies show higher levels of fatty acid oxidation (Oren et al, 2021). This shift away from the “Warburg” glycolytic state into a more mitochondrially active energy production state, which resembles non‐transformed cells, might indicate the release from oncogenic addiction. Importantly, this shift does not lead to overall lower metabolic activity and in some cases might even allow persisters that were arrested to resume cell cycle in the presence of a drug. This high modularity is possible in cancer cells as they can, under certain conditions, tap into a vast space of cellular states that reflect different tissues and developmental trajectories. Cancer persister cell plasticity is perhaps best exemplified by phenotypic transformation from non‐small‐cell lung adenocarcinoma to small‐cell lung cancer upon prolonged treatment with EGFR inhibitors (Shaurova et al, 2020). Such lineage switching accounts for up to 14% of acquired resistance to EGFR‐targeted therapy. Clinical data of relapsed patients strongly support the hypothesis that this transformation happens via persister cells that were able to withstand EGFR therapy. Taken together, these observations show that cancer persister cells can circumvent oncogenic withdrawal by adopting alternative cell states. Notably, these changes do not necessarily require any genetic alteration and in theory can be reversible and potentially mediated by microenvironment signaling.The most recent addition to the cancer‐fighting arsenal are immunotherapies designed to boost immune responses. Immune‐persisters, cells that can evade immune response, have been reported in multiple cancer types and are thought to underlie the late relapse frequently observed in patients (Shen et al, 2020). While tumor dormancy might play a role in this context as well, it is interesting to note that immune evasion can be achieved by modulating immune checkpoint molecules without any need to suppress cell proliferation. Furthermore, in the case of CAR T‐cell therapy, a class of immunotherapy that is based on revamped T cells, persistence might be viewed as a dynamic cell‐to‐cell communication process. It was shown that to elicit killing a cancer cell has to have multiple interactions with a T cell (Weigelin et al, 2021). This multihit sequential process that can take more than an hour in vivo may allow cancer cells to modulate the cytotoxic T cell in a way that would favor their persistence. Hence, understanding what underlies T‐cell phenotypes might as be as important as studying the cancer persister cells they are targeting.The holy grail of the persister filed is finding ways to target these drug‐tolerant cells in a manner that would prevent disease recurrence. However, given at least three classes of persisters have been already reported, and more are expected to arise as we continue to expand our therapeutic toolbox, would it even be possible to implement a single approach to eliminate them? Studies that searched for a magic bullet that could eliminate persister cells were largely based on the hope that persister cells would be less heterogenous than the drug‐naïve cell population they were derived from (Cabanos & Hata, 2021; Hangauer et al, 2017). If such convergence on similar cell states exists upon treatment, it simplifies the need to combine multiple drugs to eliminate the entire cell population. Unfortunately, it seems that persister cells can come in multiple forms and that distinct persister phenotypes may coexist in a single tumor. The major drivers of this heterogeneity currently remain unclear and may include tumor lineage, treatment type, or a combination of both. Moreover, it is unknown if the heterogeneity in persister phenotypes can be predicated based on the drug‐naïve population and how these diverse persister fates are associated with clinical outcomes. Understanding persister heterogeneity is critical as the simplistic approach of trying to eliminate as many persister cells as possible, assumes that all cells are equally pathogenic, which might not be the case if only a subset of them are able to contribute to relapse. Furthermore, persister cells might differ in their aptitude to give rise to cells that harbor a resistance‐mediating mutation. Such differences in evolvability must be considered when weighing possible treatments. Answering these questions would be key to devising effective therapeutic approaches to eliminate persister cells. In the last century, the study of microbial persistence had provided important insights into how to fight infections. Hopefully, in the years to come, we will build upon this valuable knowledge foundation and expend it to devise better ways to fight cancer.  相似文献   

14.
目的:探讨猪大肠杆菌的耐药质粒图谱、耐药性及耐药基因之间的关系。方法:从湖南省株洲、益阳的四个猪场分离出9株大肠杆菌,进行质粒电泳图谱分析、用PCR法检测耐喹诺酮类耐药基因Gyr A、Par C和耐四环素类耐药基因Tet A、Tet B,并采用Kirby-bauer法对这9株大肠杆菌进行药敏(18种抗生素)试验。结果:其中9株大肠杆菌含有三条或者三条以上的质粒条带,且其质粒谱型均不相同;9株大肠杆菌均检测出4种耐药基因Gyr A、Par C、Tet A和Tet B;9株大肠杆菌对所选用的抗生素存在不同程度的耐药性,其中7株大肠杆菌对10种或10种以上的抗生素耐药,最高对13种抗生素耐药,氨苄西林、青霉素、阿莫西林、红霉素的耐药率达100%,对四环素、多西环素的耐药率达到88.9%,而多粘菌素B、阿奇霉素、大观霉素耐药率较低。结论:耐药性与质粒条带数、耐药基因之间并无明显的相关性;猪大肠杆菌呈多重耐药之势,在治疗大肠杆菌病时最好根据药敏实验结果选用合适的抗生素。  相似文献   

15.
Signaling-mediated bacterial persister formation   总被引:1,自引:0,他引:1  
Here we show that bacterial communication through indole signaling induces persistence, a phenomenon in which a subset of an isogenic bacterial population tolerates antibiotic treatment. We monitor indole-induced persister formation using microfluidics and identify the role of oxidative-stress and phage-shock pathways in this phenomenon. We propose a model in which indole signaling 'inoculates' a bacterial subpopulation against antibiotics by activating stress responses, leading to persister formation.  相似文献   

16.
In a phenomenon called persistence, small numbers of bacterial cells survive even after exposure to antibiotics. Recently, bactericidal antibiotics have been demonstrated to kill bacteria by increasing the levels of hydroxyl radicals inside cells. In the present study, we report a direct correlation between intracellular hydroxyl radical formation and bacterial persistence. By conducting flow cytometric analysis in a three-dimensional space, we resolved distinct bacterial populations in terms of intracellular hydroxyl radical levels, morphology and viability. We determined that, upon antibiotic treatment, a small sub-population of Escherichia coli survivors do not overproduce hydroxyl radicals and maintain normal morphology, whereas most bacterial cells were killed by accumulating hydroxyl radicals and displayed filamentous morphology. Our results suggest that bacterial persisters can be formed once they have transient defects in mediating reactions involved in the hydroxyl radical formation pathway. Thus, it is highly probable that persisters do not share a common mechanism but each persister cell respond to antibiotics in different ways, while they all commonly show lowered hydroxyl radical formation and enhanced tolerance to antibiotics.  相似文献   

17.
Persister cells are dormant phenotypic variants inherent in a bacterial population. They play important roles in chronic infections and present great challenges to therapy due to extremely enhanced tolerance to antibiotics compared to that of normal cells of the same genotype. In this study, we report that cationic membrane-penetrating peptides containing various numbers of arginine and tryptophan repeats are effective in killing persister cells of Escherichia coli HM22, a hyper-persister producer. The activities of three linear peptides [(RW)(n)-NH(2), where n is 2, 3, or 4] and a dendrimeric peptide, (RW)(4D), in killing bacterial persisters were compared. Although the dendrimeric peptide (RW)(4D) requires a lower threshold to kill planktonic persisters, octameric peptide (RW)(4)-NH(2) is the most effective against planktonic persister cells at high concentrations. For example, treatment with 80 μM (RW)(4)-NH(2) for 60 min led to a 99.7% reduction in the number of viable persister cells. The viability of persister cells residing in surface-attached biofilms was also significantly reduced by (RW)(4)-NH(2) and (RW)(4D). These two peptides were also found to significantly enhance the susceptibility of biofilm cells to ofloxacin. The potency of (RW)(4)-NH(2) was further marked by its ability to disperse and kill preformed biofilms harboring high percentages of persister cells. Interestingly, approximately 70% of the dispersed cells were found to have lost their intrinsic tolerance and become susceptible to ampicillin if not killed directly by this peptide. These results are helpful for better understanding the activities of these peptides and may aid in future development of more effective therapies of chronic infections.  相似文献   

18.
19.
The majority of cells transferred from stationary-phase culture into fresh medium resume growth quickly, while a few remain in a nongrowing state for longer. These temporarily nonproliferating bacteria are tolerant of several bactericidal antibiotics and constitute a main source of persisters. Several genes have been shown to influence the frequency of persisters in Escherichia coli, although the exact mechanism underlying persister formation is unknown. This study demonstrates that the frequency of persisters is highly dependent on the age of the inoculum and the medium in which it has been grown. The hipA7 mutant had 1,000 times more persisters than the wild type when inocula were sampled from younger stationary-phase cultures. When started after a long stationary phase, the two displayed equal and elevated persister frequencies. The lower persister frequencies of glpD, dnaJ, and surA knockout strains were increased to the level of the wild type when inocula aged. The mqsR and phoU deletions showed decreased persister levels only when the inocula were from aged cultures, while sucB and ygfA deletions had decreased persister levels irrespective of the age of the inocula. A dependency on culture conditions underlines the notion that during screening for mutants with altered persister frequencies, the exact experimental details are of great importance. Unlike ampicillin and norfloxacin, which always leave a fraction of bacteria alive, amikacin killed all cells in the growth resumption experiment. It was concluded that the frequency of persisters depends on the conditions of inoculum cultivation, particularly its age, and the choice of antibiotic.  相似文献   

20.
Persistent bacterial infections do not respond to current antibiotic treatments and thus present a great medical challenge. These conditions have been linked to the formation of dormant subpopulations of bacteria, known as persister cells, that are growth-arrested and highly tolerant to conventional antibiotics. Here, we report a new strategy of persister control and demonstrate that minocycline, an amphiphilic antibiotic that does not require active transport to penetrate bacterial membranes, is effective in killing Escherichia coli persister cells [by 70.8 ± 5.9% (0.53 log) at 100 μg/mL], while being ineffective in killing normal cells. Further mechanistic studies revealed that persister cells have reduced drug efflux and accumulate more minocycline than normal cells, leading to effective killing of this dormant subpopulation upon wake-up. Consistently, eravacycline, which also targets the ribosome but has a stronger binding affinity than minocycline, kills persister cells by 3 logs when treated at 100 μg/mL. In summary, the findings of this study reveal that while dormancy is a well-known cause of antibiotic tolerance, it also provides an Achilles’ heel for controlling persister cells by leveraging dormancy associated reduction of drug efflux.  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