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 0 毫秒
1.
2.
陈尚武 《生命的化学》2001,21(5):379-381
免疫和造血细胞的生长、分化及其他功能受到细胞因子网络的控制。由于大多数细胞因子受体缺乏胞浆段的激酶结构域 ,配体依赖的酪氨酸磷酸化由非受体酪氨酸激酶来中介。细胞因子刺激后早期激活的主要酪氨酸激酶是Januskinase(JAK)家族。事实上 ,JAK STAT途径是许多细胞因子激活基因转录最重要机制之一。当细胞因子结合到细胞表面的受体 ,引起受体的二聚化 ,进而活化JAK激酶 ,活化的JAK激酶反过来磷酸化细胞因子受体 ,导致其他的信号分子如STAT家族蛋白的介入并被激活 ,活化的STAT转入细胞核 ,激活大量细…  相似文献   

3.

Background

In Xenopus early embryogenesis, various genes are involved with mesoderm formation. In particular, dorsal mesoderm contains the organizer region and induces neural tissues through the inhibition of bone morphogenetic protein (BMP) signaling. In our initial study to identify novel genes necessary for maintaining the undifferentiated state, we unexpectedly revealed mesoderm-inducing activity for mNanog in Xenopus.

Methodology/Principal Findings

The present series of experiments investigated the effect of mNanog gene expression on Xenopus embryo. Ectopic expression of mNanog induced dorsal mesoderm gene activity, secondary axis formation, and weakly upregulated Activin/nodal signaling. The injection of mNanog also effectively inhibited the target genes of BMP signaling, while Xvent2 injection downregulated the dorsal mesoderm gene expression induced by mNanog injection.

Conclusions/Significance

These results suggested that mNanog expression induces dorsal mesoderm by regulating both Activin/nodal signaling and BMP signaling in Xenopus. This finding highlights the possibly novel function for mNanog in stimulating the endogenous gene network in Xenopus mesoderm formation.  相似文献   

4.
The secreted protein Noggin1 is an embryonic inducer that can sequester TGFβ cytokines of the BMP family with extremely high affinity. Owing to this function, ectopic Noggin1 can induce formation of the headless secondary body axis in Xenopus embryos. Here, we show that Noggin1 and its homolog Noggin2 can also bind, albeit less effectively, to ActivinB, Nodal/Xnrs and XWnt8, inactivation of which, together with BMP, is essential for the head induction. In support of this, we show that both Noggin proteins, if ectopically produced in sufficient concentrations in Xenopus embryo, can induce a secondary head, including the forebrain. During normal development, however, Noggin1 mRNA is translated in the presumptive forebrain with low efficiency, which provides the sufficient protein concentration for only its BMP-antagonizing function. By contrast, Noggin2, which is produced in cells of the anterior margin of the neural plate at a higher concentration, also protects the developing forebrain from inhibition by ActivinB and XWnt8 signaling. Thus, besides revealing of novel functions of Noggin proteins, our findings demonstrate that specification of the forebrain requires isolation of its cells from BMP, Activin/Nodal and Wnt signaling not only during gastrulation but also at post-gastrulation stages.  相似文献   

5.
《Autophagy》2013,9(3):235-237
Ubiquitously expressed mu- and m-calpain proteases consist of 80-kDa catalytic subunits encoded by the Capn1 and Capn2 genes, respectively, and a common 28-kDa regulatory subunit encoded by the calpain small 1 (Capns1) gene.

The mu- and m-calpain proteases have been implicated in both pro- or anti-apoptotic functions. We have found that Capns1 depletion is coupled to increased sensitivity to apoptosis triggered by a number of autophagy-inducing stimuli in mammalian cells. Therefore we investigated the involvement of calpains in autophagy using MEFs derived from Capns1 knockout mice and Capns1 depleted human cells as model systems.

We found that autophagy is impaired in Capns1-deficient cells by immunostaining of the endogenous autophagosome marker LC3 and electron microscopy experiments. Accordingly, the enhancement of lysosomal activity and long-lived proteins degradation, normally occurring upon starvation, are also reduced.

In Capns1-depleted cells ectopic LC3 accumulates in early endosome-like vesicles that might represent a salvage pathway for protein degradation when autophagy is defective.

Addendum to:

Calpain is Required for Macroautophagy in Mammalian Cells

Francesca Demarchi, Cosetta Bertoli, Tamara Copetti, Isei Tanida, Claudio Brancolini, Eeva-Liisa Eskelinen and Claudio Schneider

J Cell Biol 2006; 175:595-605  相似文献   

6.
7.
8.
9.
Sli-1, a Negative Regulator of Let-23-Mediated Signaling in C. Elegans   总被引:10,自引:0,他引:10       下载免费PDF全文
By screening for suppressors of hypomorphic mutations of let-23, a receptor tyrosine kinase necessary for vulval induction in Caenorhabditis elegans, we recovered >/=12 mutations defining the sli-1 (suppressor of lineage defect) locus. sli-1 mutations suppress four of five phenotypes associated with hypomorphic alleles of let-23 but do not suppress let-23 null alleles. Thus, a sli-1 mutation does not bypass the requirement for functional let-23 but rather allows more potent LET-23-dependent signaling. Mutations at the sli-1 locus are otherwise silent with respect to vulval differentiation and cause only a low-penetrance abnormal head phenotype. Mutations at sli-1 also suppress the vulval defects but not other defects associated with mutations of sem-5, whose product likely interacts with LET-23 protein during vulval induction. Mutations at sli-1 suppress lin-2, lin-7 and lin-10 mutations but only partially suppress lin-3 and let-60 mutations and do not suppress a lin-45 mutation. The sli-1 locus displays dosage sensitivity: severe reduction of function alleles of sli-1 are semidominant suppressors; a duplication of the sli-1 (+) region enhances the vulvaless phenotype of hypomorphic mutations of let-23. We propose that sli-1 is a negative regulator that acts at or near the LET-23-mediated step of the vulval induction pathway. Our analysis suggests that let-23 can activate distinct signaling pathways in different tissues: one pathway is required for vulval induction; another pathway is involved in hermaphrodite fertilty and is not regulated by sli-1.  相似文献   

10.

Background

Endocytosis is a key regulatory step of diverse signalling pathways, including receptor tyrosine kinase (RTK) signalling. Hrs and Stam constitute the ESCRT-0 complex that controls the initial selection of ubiquitinated proteins, which will subsequently be degraded in lysosomes. It has been well established ex vivo and during Drosophila embryogenesis that Hrs promotes EGFR down regulation. We have recently isolated the first mutations of stam in flies and shown that Stam is required for air sac morphogenesis, a larval respiratory structure whose formation critically depends on finely tuned levels of FGFR activity. This suggest that Stam, putatively within the ESCRT-0 complex, modulates FGF signalling, a possibility that has not been examined in Drosophila yet.

Principal Findings

Here, we assessed the role of the Hrs/Stam complex in the regulation of signalling activity during Drosophila development. We show that stam and hrs are required for efficient FGFR signalling in the tracheal system, both during cell migration in the air sac primordium and during the formation of fine cytoplasmic extensions in terminal cells. We find that stam and hrs mutant cells display altered FGFR/Btl localisation, likely contributing to impaired signalling levels. Electron microscopy analyses indicate that endosome maturation is impaired at distinct steps by hrs and stam mutations. These somewhat unexpected results prompted us to further explore the function of stam and hrs in EGFR signalling. We show that while stam and hrs together downregulate EGFR signalling in the embryo, they are required for full activation of EGFR signalling during wing development.

Conclusions/Significance

Our study shows that the ESCRT-0 complex differentially regulates RTK signalling, either positively or negatively depending on tissues and developmental stages, further highlighting the importance of endocytosis in modulating signalling pathways during development.  相似文献   

11.
Activin as a morphogen in Xenopus mesoderm induction.   总被引:3,自引:0,他引:3  
Activin, a member of the Transforming Growth Factor beta (TGF-beta) superfamily, can behave as a morphogen in cells of the early Xenopus embryo by inducing a range of mesodermal genes in a concentration-dependent manner. This review examines the behaviour of activin as it forms a morphogen gradient. It also discusses how a cell can perceive its position in a concentration gradient in order to activate appropriate mesodermal gene responses.  相似文献   

12.
Targeting of cancer stem cells (CSCs) has the potential to address the recalcitrance of pancreatic cancer to chemotherapy. In this issue of Cell Stem Cell, Lonardo et al. (2011) demonstrate that Nodal/Activin signaling is crucial for the maintenance and tumor-initiating capacity of pancreatic CSCs.  相似文献   

13.
Kinase cascades, in which enzymes are sequentially activated by phosphorylation, are quintessential signaling pathways. Signal transduction is not always achieved by direct activation, however. Often, kinases activate pathways by deactivation of a negative regulator; this indirect mechanism, pervasive in Akt signaling, has yet to be systematically explored. Here, we show that the indirect mechanism has properties that are distinct from direct activation. With comparable parameters, the indirect mechanism yields a broader range of sensitivity to the input, beyond saturation of regulator phosphorylation, and kinetics that become progressively slower, not faster, with increasing input strength. These properties can be integrated in network motifs to produce desired responses, as in the case of feedforward loops.Phosphorylation of proteins and lipids, catalyzed by specific kinase enzymes, is ubiquitous in intracellular signal transduction. A classic example in eukaryotes is the canonical structure of the mitogen-activated protein kinase cascades, in which three kinases are sequentially activated by phosphorylation (1). Another example is the PI3K (phosphoinositide 3-kinase)/Akt pathway, which (like the mammalian mitogen-activated protein kinases) is prominently dysregulated in human cancers (2). Type-I PI3Ks phosphorylate a lipid substrate to produce the lipid second messenger, PIP3, which recruits the protein kinase Akt and mediates its activation by phosphorylation (3,4). In no small part because of these important pathways, we typically think of phosphorylation as a direct means of activating molecular interactions and reactions in signal transduction. This is not the only way to increase the flux through a signaling pathway, however. Consider signaling downstream of Akt, which phosphorylates a host of protein substrates to affect diverse functions. A survey of the Akt signaling hub shows that many of these reactions result in a decrease, rather than an increase, in activity/function of the substrates (3). And, among those substrates, the four listed in Fig. S1 in the Supporting Material). Whereas negative regulators are appreciated for their roles in feedback adaptation of signaling, the implications of deactivating a negative regulator as an indirect mechanism of pathway activation has yet to be explored.

Table 1

Survey of Akt substrates and downstream signaling
Substrate (site)Effect on substrateOutcome
TSC2 (T1462)GAP activity ↓Rheb, mTOR ↑
PRAS40 (T246)mTOR binding ↓mTOR ↑
GSK3α/β (S21/S9)kinase activity ↓β-catenin ↑
BAD (S136)Bcl-2/xL binding ↓Bcl-2/xL ↑
Open in a separate windowHere, we use simple kinetic models to elucidate the basic properties of pathway activation by deactivation of a negative regulator (hereafter referred to as mechanism II), as compared with the standard activation of a positive regulator (mechanism I). The analysis is presented in the context of protein phosphorylation, but the conclusions may be generalized to other reversible modifications or to allosteric binding interactions. The common first step is phosphorylation of the regulatory molecule by the kinase. The activity of the upstream kinase such as Akt may be represented by a dimensionless, time (t)-dependent input signal function, s(t). We assume that the total amount of regulator is constant and define its phosphorylated fraction as ϕ(t). Neglecting concentration gradients and saturation of the upstream kinase and of the opposing (constitutively active) phosphatase(s), the conservation of phosphorylated regulator is expressed as follows (see Text S1 in the Supporting Material):dϕdt=kp[s(1ϕ)ϕ];ϕ(0)=0.(1)The parameter kp is the pseudo-first-order rate constant of protein dephosphorylation. In the case of s = constant (i.e., subject to a step change at t = 0), the properties of this simplified kinetic equation are well known (5) and may be summarized as follows. As the magnitude of the signal strength s increases, the steady-state value of ϕ, ϕss, increases in a saturable fashion; when s >> 1, ϕss approaches its maximum value of 1 and is insensitive to further increases in s. The kinetics of ϕ(t) approaching ϕss become progressively faster as s increases, however.Next, we model the influence of the regulator on a downstream response. Defining the fractional response as ρ and following analogous assumptions as above, we formulate equations for mechanisms I and II as follows:dρdt={[ka,0+(ka,maxka,0)ϕ](1ρ)kd,0ρ(I)ka,0(1ρ)[kd,0(kd,0kd,min)ϕ]ρ(II).(2)In each equation, the first term on the right-hand side describes activation, and the second, deactivation. In mechanism I, the effective rate constant of activation increases linearly with ϕ, from a minimum value of ka,0 when ϕ = 0 up to a maximum value of ka,max when ϕ = 1; the deactivation rate constant is fixed at kd,0. Conversely, in mechanism II, the effective rate constant of deactivation decreases linearly with ϕ, from a maximum value of kd,0 when ϕ = 0 down to a minimum value of kd,min when ϕ = 1; in this mechanism, the activation rate constant is fixed at ka,0. The initial condition is assigned so that ρ is stationary when ϕ = 0. To further set the two mechanisms on a common basis, we define dimensionless parameters such that the maximum steady-state value of ρ (with ϕss = 1) is the same for both mechanisms I and II,gka,max/ka,0kd,0/kd,minKka,0/kd,0.(3)With these definitions, each conservation equation is reduced to the following dimensionless form:1kd,0dρdt={K[1+(g1)ϕ](1ρ)ρ(I)K(1ρ)[1(1g1)ϕ]ρ(II).(4)Mechanisms I and II (Fig. 1 a) are compared first at the level of their steady-state solutions, ρss, for stationary s. Equation 1 yields the familiar hyperbolic dependence of ϕss on s, and ρss(s) has the same shape for both mechanisms. However, whereas ρss of mechanism I shows saturation at a lower value of s than ϕss, the opposite is true of mechanism II (Fig. 1 b). Thus, mechanism II retains sensitivity to the input even while phosphorylation of the upstream regulator shows saturation. This is perhaps more readily seen when ϕss(s) is replaced with a sigmoidal Hill function (i.e., with s replaced by sn in Eq. 1) (Fig. 1 c). The key parameter that affects the relative sensitivities of mechanisms I and II and the disparity between them is the gain constant, g (see Text S1 in the Supporting Material). As this parameter is increased, ρss of mechanism I becomes increasingly saturable with respect to ϕss (Fig. 1 d), whereas ρss of mechanism II gains sensitivity as ϕss approaches 1 (Fig. 1 e). As an illustrative example, consider that when ϕss is increased from 0.90 to 0.95, or from 0.98 to 0.99, the amount of the negative regulator in the active state is reduced by a factor of 2 (see Fig. S2).Open in a separate windowFigure 1Steady-state properties of mechanisms I and II. (a) Schematics of direct (I) and indirect (II) activation. (b) Steady-state dose responses, ρss(s), of mechanisms I and II along with phosphorylation of the upstream regulator, ϕss(s) (Eq. 1 at steady state); K = 0.05, g = 100. (c) Same as panel b, except with a sigmoidal ϕss(s) (Hill function with n = 4). (d) Steady-state output, ρss, of mechanism I vs. ϕss for K = 0.05 and indicated values of the gain constant, g. (e) Same as panel d, but for mechanism II. To see this figure in color, go online.The two mechanisms also show distinct temporal responses. In the response of mechanism I to a step increase in s, ρ(t) approaches ρss with a timescale that generally becomes faster as s increases. Unless the kinetics of ϕ(t) are rate-limiting, the timescale is ∼kd,0–1(1–ρss) (Fig. 2 a; see also Text S1 and Fig. S3 in the Supporting Material). Conversely, the response of mechanism II generally becomes slower as s increases, inasmuch as the frequency of deactivation decreases whereas that of activation is constant, with a timescale of ∼ka,0–1ρss (Fig. 2 b). To approximate a transient input, we model s(t) as a step increase followed by a decay. For mechanism I, the response ρ(t) is such that the variation in the time of the peak, as a function of the step size, is modest. The subsequent decay is prolonged when ϕ(t) hovers close to saturation (Fig. 2 c). Such kinetic schemes have been analyzed in some detail previously (6,7). In contrast, the response of mechanism II to the transient input is such that the system retains sensitivity and consistent decay kinetics beyond the saturation of ϕ(t). The distinctive feature is that ρ(t) peaks noticeably later in time as the magnitude of the peak increases (Fig. 2 d).Open in a separate windowFigure 2Kinetic properties of mechanisms I and II. (a) Response of mechanism I to a step change in s from zero to the indicated s(0). Time is given in units of kpt; parameters are K = 0.05, g = 10, and kd,0 = 0.1kp. (b) Same as panel a, but for mechanism II. (c) Same as panel a, but for a transient input, s(t) = s(0)exp(–0.03kpt). d) Same as panel c, but for mechanism II. To see this figure in color, go online.Having established the basic steady state and kinetic properties of mechanism II as compared with the canonical mechanism I, we considered what outcomes could be achieved by linking these motifs in series or in parallel. Such schemes are identified in the Akt/mTOR signaling network, for example (see Fig. S4). In a standard kinase activation cascade, it is understood that the properties of saturation and sensitivity are compounded with each step of the cascade (8). Thus, two sequential steps of mechanism I yield progressive saturation of the steady-state output at lower s (Fig. 3 a), and the desaturating effect of mechanism II is likewise compounded (Fig. 3 b). By corollary it follows that a sequence of mechanisms I and II will show an intermediate dose response; that is, the mechanism II step offsets the saturation effect of mechanism I.Open in a separate windowFigure 3Serial and parallel schemes incorporating mechanism I or/and II. (a) Steady-state outputs of two response elements, ρ1 and ρ2, activated by mechanism I in series. At each level, K = 0.05, g = 100. (b) Same as panel a, but for mechanism II in series. (c) Incoherent feedforward loop (FFL) in which mechanisms I and II are activated in parallel to activate and inhibit, respectively, the terminal output. For both mechanisms I and II, K = 0.05, g = 100. The parameters for Eq. 5 are α = 2.5, β = 50. To see this figure in color, go online.A more complex scheme is to combine the two mechanisms in parallel, as in an incoherent feedforward loop (FFL) connected to an “AND NOT” output as follows:Output = αρI/(1 + αρIβρII).(5)Given the differential saturation properties of mechanisms I and II, this scheme readily yields the expected biphasic dose response (9) without the need for disparate values of the parameters (Fig. 3 c). Regarding the kinetics, the analysis shown in Fig. 2 makes it clear that mechanism II naturally introduces time delays in cascades or network motifs. Thus, for the incoherent FFL at high, constant s, activation of inhibition by mechanism II would tend to yield a dynamic response marked by a peak followed by adaptation (see Fig. S5). Analogous calculations were carried out for a coherent FFL as well (see Fig. S6).To summarize our conclusions and their implications for signaling downstream of Akt and other kinases, we have described a distinct, indirect signal transduction mechanism characterized by deactivation of a negative regulator. This motif shows steady-state sensitivity beyond saturation, and therefore the activity of the upstream kinase, such as Akt, can be relatively high. By comparison, the direct activation of signaling by phosphorylation requires that activity of the kinase be regulated, or specifically countered by high phosphatase activity, to maintain sensitivity and avoid saturation of the response. The mechanism described here also introduces relatively slow kinetics (for comparable parameter values). This property, together with its extended range of sensitivity, would allow the motif to be incorporated in signaling networks to yield desired steady and unsteady responses in a robust manner. Considering that key signaling processes mediated by Akt (notably activation of the mammalian target of rapamycin (mTOR) pathway) are achieved by deactivation of negative regulators, we assert that greater recognition of this mechanism and of its distinct properties is warranted.  相似文献   

14.
15.
Lung cancer is currently the most deadly malignancy in industrialized countries and accounts for 18% of all cancer-related deaths worldwide. Over 70% of patients with non-small cell lung cancer (NSCLC) are diagnosed at a late stage, with a 5-year survival below 10%. KRAS and the EGFR are frequently mutated in NSCLC and while targeted therapies for patients with EGFR mutations exist, oncogenic KRAS is thus far not druggable. KRAS activates multiple signalling pathways, including the PI3K/Akt pathway, the Raf-Mek-Erk pathway and the RalGDS/Ral pathway. Lung-specific expression of BrafV600E, the most prevalent BRAF mutation found in human tumors, results in Raf-Mek-Erk pathway activation and in the formation of benign adenomas that undergo widespread senescence in a Cre-activated Braf mouse model (BrafCA). However, oncogenic KRAS expression in mice induces adenocarcinomas, suggesting additional KRAS-activated pathways cooperate with sustained RAF-MEK-ERK signalling to bypass the oncogene-induced senescence proliferation arrest.To determine which KRAS effectors were responsible for tumor progression, we created four effector domain mutants (S35, G37, E38 and C40) in G12V-activated KRAS and expressed these alone or with BrafV600E in mouse lungs… The S35 and E38 mutants bind to Raf proteins but not PI3K or RalGDS; the G37 mutant binds to RalGDS and not Raf or PI3K and the C40 mutant is specific to PI3K. We designed lentiviral vectors to code for Cre recombinase along with KRAS mutants (V12, V12/S35, V12/G37, V12/E38 or V12/C40) or EGFP as a negative control.. These lentiviruses were used to infect BrafCA and wild-type mice. Surprisingly there was a significant decrease in tumor number and penetrance with each KRAS effector domain mutant relative to controls, suggesting that KRAS directly activates effectors with tumor suppressive functions.  相似文献   

16.
The Jun N-terminal kinase and p38 pathways, also known as stress-activated protein kinase (SAPK) pathways, are signaling conduits reiteratively used throughout the development and adult life of metazoans where they play central roles in the control of apoptosis, immune function, and environmental stress responses. We recently identified a Drosophila Ser/Thr phosphatase of the PP2C family, named Alphabet (Alph), which acts as a negative regulator of the Ras/ERK pathway. Here we show that Alph also plays an inhibitory role with respect to Drosophila SAPK signaling during development as well as under stress conditions such as oxidative or genotoxic stresses. Epistasis experiments suggest that Alph acts at a step upstream of the MAPKKs Hep and Lic. Consistent with this interpretation, biochemical experiments identify the upstream MAPKKKs Slpr, Tak1, and Wnd as putative substrates. Together with previous findings, this work identifies Alph as a general attenuator of MAPK signaling in Drosophila.  相似文献   

17.
18.
In this study, we attempted to understand the role of an orphan gene amyR in Erwinia amylovora, a functionally conserved ortholog of ybjN in Escherichia coli, which has recently been characterized. Amylovoran, a high molecular weight acidic heteropolymer exopolysaccharide, is a virulent factor of E. amylovora. As reported earlier, amylovoran production in an amyR knockout mutant was about eight-fold higher than that in the wild type (WT) strain of E. amylovora. When a multicopy plasmid containing the amyR gene was introduced into the amyR mutant or WT strains, amylovoran production was strongly inhibited. Furthermore, amylovoran production was also suppressed in various amylovoran-over-producing mutants, such as grrSA containing multicopies of the amyR gene. Consistent with amylovoran production, an inverse correlation was observed between in vitro expression of amyR and that of amylovoran biosynthetic genes. However, both the amyR knockout mutant and over-expression strains showed reduced levan production, another exopolysaccharide produced by E. amylovora. Virulence assays demonstrated that while the amyR mutant was capable of inducing slightly greater disease severity than that of the WT strain, strains over-expressing the amyR gene did not incite disease on apple shoots or leaves, and only caused reduced disease on immature pear fruits. Microarray studies revealed that amylovoran biosynthesis and related membrane protein-encoding genes were highly expressed in the amyR mutant, but down-regulated in the amyR over-expression strains in vitro. Down-regulation of amylovoran biosynthesis genes in the amyR over-expression strain partially explained why over-expression of amyR led to non-pathogenic or reduced virulence in vivo. These results suggest that AmyR plays an important role in regulating exopolysaccharide production, and thus virulence in E. amylovora.  相似文献   

19.
20.
Clostridium difficile is a Gram-positive spore-forming pathogen and a leading cause of nosocomial diarrhea. C. difficile infections are transmitted when ingested spores germinate in the gastrointestinal tract and transform into vegetative cells. Germination begins when the germinant receptor CspC detects bile salts in the gut. CspC is a subtilisin-like serine pseudoprotease that activates the related CspB serine protease through an unknown mechanism. Activated CspB cleaves the pro-SleC zymogen, which allows the activated SleC cortex hydrolase to degrade the protective cortex layer. While these regulators are essential for C. difficile spores to outgrow and form toxin-secreting vegetative cells, the mechanisms controlling their function have only been partially characterized. In this study, we identify the lipoprotein GerS as a novel regulator of C. difficile spore germination using targeted mutagenesis. A gerS mutant has a severe germination defect and fails to degrade cortex even though it processes SleC at wildtype levels. Using complementation analyses, we demonstrate that GerS secretion, but not lipidation, is necessary for GerS to activate SleC. Importantly, loss of GerS attenuates the virulence of C. difficile in a hamster model of infection. Since GerS appears to be conserved exclusively in related Peptostreptococcaeace family members, our results contribute to a growing body of work indicating that C. difficile has evolved distinct mechanisms for controlling the exit from dormancy relative to B. subtilis and other spore-forming organisms.  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