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 96 毫秒
1.
3‐deoxy‐d ‐manno‐octulosonic acid‐lipid A (Kdo2‐lipid A) is the essential component of lipopolysaccharide in most Gram‐negative bacteria and the minimal structural component to sustain bacterial viability. It serves as the active component of lipopolysaccharide to stimulate potent host immune responses through the complex of Toll‐like‐receptor 4 (TLR4) and myeloid differentiation protein 2. The entire biosynthetic pathway of Escherichia coli Kdo2‐lipid A has been elucidated and the nine enzymes of the pathway are shared by most Gram‐negative bacteria, indicating conserved Kdo2‐lipid A structure across different species. Yet many bacteria can modify the structure of their Kdo2‐lipid A which serves as a strategy to modulate bacterial virulence and adapt to different growth environments as well as to avoid recognition by the mammalian innate immune systems. Key enzymes and receptors involved in Kdo2‐lipid A biosynthesis, structural modification and its interaction with the TLR4 pathway represent a clear opportunity for immunopharmacological exploitation. These include the development of novel antibiotics targeting key biosynthetic enzymes and utilization of structurally modified Kdo2‐lipid A or correspondingly engineered live bacteria as vaccines and adjuvants. Kdo2‐lipid A/TLR4 antagonists can also be applied in anti‐inflammatory interventions. This review summarizes recent knowledge on both the fundamental processes of Kdo2‐lipid A biosynthesis, structural modification and immune stimulation, and applied research on pharmacological exploitations of these processes for therapeutic development.  相似文献   

2.
The lipopolysaccharide (LPS) isolated from certain important Gram-negative pathogens including a human pathogen Yersinia pestis and opportunistic pathogens Burkholderia mallei and Burkholderia pseudomallei contains d-glycero-d-talo-oct-2-ulosonic acid (Ko), an isosteric analog of 3-deoxy-d-manno-oct-2-ulosonic acid (Kdo). Kdo 3-hydroxylase (KdoO), a Fe2+/α-KG/O2 dependent dioxygenase from Burkholderia ambifaria and Yersinia pestis is responsible for Ko formation with Kdo2-lipid A as a substrate, but in which stage KdoO functions during the LPS biosynthesis has not been established. Here we purify KdoO from B. ambifaria (BaKdoO) to homogeneity for the first time and characterize its substrates. BaKdoO utilizes Kdo2-lipid IVA or Kdo2-lipid A as a substrate, but not Kdo-lipid IVAin vivo as well as in vitro and Kdo-(Hep)kdo-lipid A in vitro. These data suggest that KdoO is an inner core assembly enzyme that functions after the Kdo-transferase KdtA but before the heptosyl-transferase WaaC enzyme during the Ko-containing LPS biosynthesis.  相似文献   

3.
Lipid A coats the outer surface of the outer membrane of Gram‐negative bacteria. In Francisella tularensis subspecies novicida lipid A is present either as the covalently attached anchor of lipopolysaccharide (LPS) or as free lipid A. The lipid A moiety of Francisella LPS is linked to the core domain by a single 2‐keto‐3‐deoxy‐D‐manno‐octulosonic acid (Kdo) residue. F. novicida KdtA is bi‐functional, but F. novicida contains a membrane‐bound Kdo hydrolase that removes the outer Kdo unit. The hydrolase consists of two proteins (KdoH1 and KdoH2), which are expressed from adjacent, co‐transcribed genes. KdoH1 (related to sialidases) has a single predicted N‐terminal transmembrane segment. KdoH2 contains 7 putative transmembrane sequences. Neither protein alone catalyses Kdo cleavage when expressed in E. coli. Activity requires simultaneous expression of both proteins or mixing of membranes from strains expressing the individual proteins under in vitro assay conditions in the presence of non‐ionic detergent. In E. coli expressing KdoH1 and KdoH2, hydrolase activity is localized in the inner membrane. WBB06, a heptose‐deficient E. coli mutant that makes Kdo2‐lipid A as its sole LPS, accumulates Kdo‐lipid A when expressing the both hydrolase components, and 1‐dephospho‐Kdo‐lipid A when expressing both the hydrolase and the Francisella lipid A 1‐phosphatase (LpxE).  相似文献   

4.
Monophosphoryl lipid A (MPLA) species, including MPL (a trade name of GlaxoSmithKline) and GLA (a trade name of Immune Design, a subsidiary of Merck), are widely used as an adjuvant in vaccines, allergy drugs, and immunotherapy to boost the immune response. Even though MPLA is a derivative of lipopolysaccharide (LPS), a component of the outer membrane of Gram-negative bacteria, bacterial strains producing MPLA have not been found in nature nor engineered. In fact, MPLA generation involves expensive and laborious procedures based on synthetic routes or chemical transformation of precursors isolated from Gram-negative bacteria. Here, we report the engineering of an Escherichia coli strain for in situ production and accumulation of MPLA. Furthermore, we establish a succinct method for purifying MPLA from the engineered E. coli strain. We show that the purified MPLA (named EcML) stimulates the mouse immune system to generate antigen-specific IgG antibodies similarly to commercially available MPLA, but with a dramatically reduced manufacturing time and cost. Our system, employing the first engineered E. coli strain that directly produces the adjuvant EcML, could transform the current standard of industrial MPLA production.  相似文献   

5.
The effects of cell wall mutation on the oxygenation of linoleic acid (M.W. 280) by recombinant Escherichia coli expressing the CYP102A2 gene encoding self-sufficient P450 monooxygenase of Bacillus subtilis was investigated. After the CYP102A2 gene was heterologously expressed in E. coli W3110 and its isogenic lipopolysaccharide (LPS) structural mutant strains, their whole-cell biotransformation activities were compared. The mutants used in this study had previously been designated as MLK53, MLK1067, and MLK986. These strains carry one or two defined mutations in the secondary acyl fatty acids of the LPS lipid A constituent. The CYP102A2 gene was overexpressed in both wild type E. coli W3110 and its mutant strains, with the specific activity ranging from 1.7 to 2.1 U/mg protein. Interestingly, the whole-cell biotransformation activity of those recombinant biocatalysts differed significantly. Indeed, MLK986 possessing the tetraacylated LPS showed a higher oxygenation activity of linoleic acid than those in wild type or other mutant strains having hexa- or penta-acylated LPSs. These results suggest that the biotransformation efficiency of E. coli-based biocatalysts, especially for medium- to large-sized lipophilic organic substrates, can be enhanced via engineering their LPS, which is known to function as a formidable barrier for hydrophobic molecules.  相似文献   

6.
The lipopolysaccharide of Vibrio cholerae has been reported to contain a single 3-deoxy-d-manno-octulosonic acid (Kdo) residue that is phosphorylated. The phosphorylated Kdo sugar further links the hexa-acylated V. cholerae lipid A domain to the core oliogosaccharide and O-antigen. In this report, we confirm that V. cholerae possesses the enzymatic machinery to synthesize a phosphorylated Kdo residue. Further, we have determined that the presence of the phosphate group on the Kdo residue is necessary for secondary acylation in V. cholerae. The requirement for a secondary substituent on the Kdo residue (either an additional Kdo sugar or a phosphate group) was also found to be critical for secondary acylation catalyzed by LpxL proteins from Bordetella pertussis, Escherichia coli, and Haemophilus influenzae. Although three putative late acyltransferase orthologs have been identified in the V. cholerae genome (Vc0212, Vc0213, and Vc1577), only Vc0213 appears to be functional. Vc0213 functions as a myristoyl transferase acylating lipid A at the 2′-position of the glucosamine disaccharide. Generally acyl-ACPs serve as fatty acyl donors for the acyltransferases required for lipopolysaccharide biosynthesis; however, in vitro assays indicate that Vc0213 preferentially utilizes myristoyl-CoA as an acyl donor. This is the first report to biochemically characterize enzymes involved in the biosynthesis of the V. cholerae Kdo-lipid A domain.Lipopolysaccharide (LPS),2 the major surface molecule in the outer membrane of Gram-negative bacteria, is composed of three domains: lipid A, core oligosaccharide, and O-antigen (1). The core oligosaccharide is further divided into two distinct regions: inner and outer core. The inner core consists of the Kdo sugars, which are responsible for linking the core region to the lipid A moiety of LPS. Lipid A is the hydrophobic anchor of LPS and is the only portion of LPS required for activating the host innate immune response by interacting with Toll-like receptor 4 and the accessory molecule, MD2.Kdo-lipid A biosynthesis is a well conserved and ordered process among Gram-negative bacteria; however, not all Gram-negative bacteria produce similar lipid A structures (2). In Escherichia coli, the biosynthesis of the Kdo-lipid A domain occurs via a nine-step process, resulting in the production of a hexa-acylated lipid A structure known as Kdo2-lipid A. Kdo2-lipid A has long been thought to be essential for the viability of E. coli; however, viable suppressor strains have been isolated that lack the Kdo sugar (3). The late steps of the biosynthetic pathway involve the addition of the Kdo sugars and the secondary or “late” acyl chains. The enzyme responsible for the addition of the Kdo sugars is the Kdo transferase (WaaA). In E. coli, this enzyme is bifunctional, transferring two Kdo sugars to the lipid A precursor, lipid IVA (4); however, other Gram-negative bacteria have been shown to possess a monofunctional or trifunctional WaaA, as is the case in Haemophilus influenzae (5) or Chlamydia trachomatis (6), respectively.Previous reports have shown that in E. coli, the addition of the Kdo sugars is critical for the functionality of the secondary acyltransferases (LpxL, LpxM, and LpxP). The E. coli late acyltransferase LpxL catalyzes the transfer of laurate (C12:0) to the acyl chain linked at the 2′-position of Kdo2-lipid IVA (7). LpxM then catalyzes the addition of a myristate (C14:0) to the 3′-linked acyl chain of the penta-acylated lipid A precursor (8). When E. coli experience cold shock conditions (temperatures below 20 °C), the late acyltransferase LpxP transfers a palmitoleate (C16:1) to the 2′-position of Kdo2-lipid IVA, replacing the C12:0 acyl chain transferred by LpxL (9). Lipid A secondary acyltransferases have been shown to primarily utilize acyl-acyl carrier proteins (acyl-ACPs) as their acyl chain donor; however, a recent report by Six et al. (10) has shown that purified E. coli LpxL is capable of utilizing acyl-coenzyme A (acyl-CoA) as an alternative acyl donor at a lesser rate.The Gram-negative bacteria Vibrio cholerae is the causative agent of the severe diarrheal disease cholera. Cholera is transmitted via the fecal-oral route by ingestion of contaminated drinking water or food. The World Health Organization reported ∼130,000 cases of cholera in 2005 with the majority occurring in Africa. There are two serogroups of V. cholerae capable of epidemic and pandemic disease: O1 and O139 (11). Previous structural analyses have revealed that these serogroups possess very different lipid A structures. The V. cholerae O1 lipid A structure was reported as hexa-acylated, bearing secondary acyl chains at the 2- and 2′-positions of phosphorylated Kdo-lipid A (1113); however, V. cholerae O139 was reported as having an octa-acylated lipid A (see Fig. 1) (11, 14).Open in a separate windowFIGURE 1.Comparison of E. coli K12 lipid A species to V. cholerae O1 and V. cholerae O139 lipid A species. The covalent modifications of lipid A are indicated with dashed bonds, and the lengths of the acyl chains are indicated below each structure. The lipid A of E. coli K12 is a hexa-acylated structure, bearing two secondary acyl chains at the 2′- and 3′-positions. The E. coli lipid A structure is glycosylated at the 6′-position with two Kdo moieties and is phosphorylated at the 1- and 4′-positions of the disaccharide backbone. Similar to E. coli, the lipid A species of V. cholerae serogroup O1 is hexa-acylated, but with a symmetrical acyl chain distribution. The proposed lipid A structure of V. cholerae O139 is the octa-acylated structure. Both V. cholerae serogroups O1 and O139 reported lipid A species have a single Kdo sugar that is phosphorylated (red) and a phosphoethanolamine (magenta) attached to the 1-phosphate.Our report focuses on V. cholerae O1 El Tor, which is the predominant disease-causing strain worldwide. Because little attention has been given to the Kdo-lipid A domain of V. cholerae, we investigated the assembly of the inner core structure of V. cholerae O1 LPS and the late acylation steps. This report demonstrates the importance of a secondary negative charge on the primary Kdo sugar of lipid A for late acyltransferase activity in V. cholerae and in other Gram-negative bacteria. Also, we have identified the putative V. cholerae late acyltransferase, Vc0213 as the LpxL homolog, transferring a myristate (C14:0) to the 2′-position of V. cholerae lipid A. These initial findings provide us with the groundwork needed to investigate the modifications of the V. cholerae Kdo-lipid A structure, which may serve as attractive vaccine targets in future research.  相似文献   

7.

Background  

Lipid A is the highly immunoreactive endotoxic center of lipopolysaccharide (LPS). It anchors the LPS into the outer membrane of most Gram-negative bacteria. Lipid A can be recognized by animal cells, triggers defense-related responses, and causes Gram-negative sepsis. The biosynthesis of Kdo2-lipid A, the LPS substructure, involves with nine enzymatic steps.  相似文献   

8.
The innate immune systems of humans and other animals are activated by lipopolysaccharides (LPS), which are glucosamine-based phospholipids that form the outer leaflet of the outer membranes of Gram-negative bacteria. Activation involves interactions of LPS with the innate immunity-receptor comprised of toll-like receptor 4 in complex with so-called MD-2 protein and accessory proteins, such as CD14 and LPS binding protein. The Lipid Metabolites and Pathways Strategy (LIPID MAPS) Consortium has isolated in large amounts a nearly homogeneous LPS, Kdo2-Lipid A, and demonstrated that it activates macrophages via toll-like receptor 4. The active form of LPS, monomer or aggregate, is controversial. We have therefore examined the aggregation behavior and other physical properties of Kdo2-Lipid A. Differential scanning calorimetry of Kdo2-Lipid A suspensions revealed a gel-to-liquid crystalline phase transition at 36.4°C (Tm). The nominal critical aggregation concentration, determined by dynamic light scattering, was found to be 41.2 ± 1.6 nM below the Tm (25°C), but only 8.1 ± 0.3 nM above the Tm (37°C). The specific molecular volume of Kdo2-Lipid A, obtained by densitometry measurements was found to be 3159 ± 71 Å3 at 25°C, from which the number of molecules in each aggregate was estimated to be 5.8 × 105. The aggregation behavior of Kdo2-Lipid A in the presence of lipoprotein-deficient serum suggests that Re LPS monomers and multimers are the active units for the immune system in the CD14-dependent and -independent pathways, respectively.  相似文献   

9.
This paper constitutes the first report on the Alr1105 of Anabaena sp. PCC7120 which functions as arsenate reductase and phosphatase and offers tolerance against oxidative and other abiotic stresses in the alr1105 transformed Escherichia coli. The bonafide of 40.8 kDa recombinant GST+Alr1105 fusion protein was confirmed by immunoblotting. The purified Alr1105 protein (mw 14.8 kDa) possessed strong arsenate reductase (Km 16.0 ± 1.2 mM and Vmax 5.6 ± 0.31 μmol min?1 mg protein?1) and phosphatase activity (Km 27.38 ± 3.1 mM and Vmax 0.077 ± 0.005 μmol min?1 mg protein?1) at an optimum temperature 37 °C and 6.5 pH. Native Alr1105 was found as a monomeric protein in contrast to its homologous Synechocystis ArsC protein. Expression of Alr1105 enhanced the arsenic tolerance in the arsenate reductase mutant E. coli WC3110 (?arsC) and rendered better growth than the wild type W3110 up to 40 mM As (V). Notwithstanding above, the recombinant E. coli strain when exposed to CdCl2, ZnSO4, NiCl2, CoCl2, CuCl2, heat, UV-B and carbofuron showed increase in growth over the wild type and mutant E. coli transformed with the empty vector. Furthermore, an enhanced growth of the recombinant E. coli in the presence of oxidative stress producing chemicals (MV, PMS and H2O2), suggested its protective role against these stresses. Appreciable expression of alr1105 gene as measured by qRT-PCR at different time points under selected stresses reconfirmed its role in stress tolerance. Thus the Alr1105 of Anabaena sp. PCC7120 functions as an arsenate reductase and possess novel properties different from the arsenate reductases known so far.  相似文献   

10.
Gram-negative bacteria, including Escherichia coli, release outer membrane vesicles (OMVs) that are derived from the bacterial outer membrane. OMVs contribute to bacterial cell–cell communications and host–microbe interactions by delivering components to locations outside the bacterial cell. In order to explore the molecular machinery involved in OMV biogenesis, the role of a major OMV protein was examined in the production of OMVs from E. coli W3110, which is a widely used standard E. coli K-12 strain. In addition to OmpC and OmpA, which are used as marker proteins for OMVs, an analysis of E. coli W3110 OMVs revealed that they also contain abundant levels of FliC, which is also known as flagellin. A membrane-impermeable biotin-labeling reagent did not label FliC in intact OMVs, but labeled FliC in sonically disrupted OMVs, suggesting that FliC is localized in the lumen of OMV. Compared to the parental strain expressing wild-type fliC, an E. coli strain with a fliC-null mutation produced reduced amounts of OMVs based on both protein and phosphate levels. In addition, an E. coli W3110-derived strain with a null-mutation in flgK, which encodes flagellar hook-associated protein that is essential along with FliC for flagella synthesis, also produced fewer OMVs than the parental strain. Taken together, these results indicate that the ability to form flagella, including the synthesis of flagella proteins, affects the production of E. coli W3110 OMVs.  相似文献   

11.
Ribosomal ribonucleic acid (RNA) and lipopolysaccharide (LPS) from P. aeruginosa were compared with respect to their protective activities in mice against an infection with P. aeruginosa. This study is concentrated on the protective activity of RNA. RNA isolated from purified ribosomes did not contain LPS as determined with the Limulus test. Injection of RNA with the adjuvant dimethyldioctadecylammonium bromide (DDA) protected mice against P. aeruginosa without inducing LPS-specific antibodies. C3H/HeJ mice which are relatively insensitive to the protective activity of LPS could be protected with RNA. The protective activities of RNA and LPS from a mutant strain of P. aeruginosa, PAC 605, containing defective lipopolysaccharide, were compared with the protective activities of RNA and LPS from the parent strain, PAC IR. The protective activity of LPS from PAC 605 was 1000 fold lower than the protective activity of LPS from PAC IR. RNA preparations of both strains induced similar percentages of survival. The protective activity of ribosomal RNA from P. aeruginosa was nonspecific since mice were also protected against a heterologous serotype of P. aeruginosa and against Escherichia coli. RNA from ribosomes of P. aeruginosa, E. coli and the non-lipopolysaccharide containing Saccharomyces cerevisiae had similar protective activities. No protection was obtained with the ribonucleic acid from the E. coli phage MS 2. It is concluded that ribosomal RNA has protective activities distinct from those of LPS.  相似文献   

12.
The Ter-15 mutant derived from E. coli K12 W2252-11U? RCstr (wild type I) is found to be sensitive to φx174 phage infection. Lipopolysaccharide extracted from this mutant inactivates the phage, and has core oligosaccharides identical in amounts to those in the lipopolysaccharide from wild type cells.In contrast, the Ter-21 mutant derived from E. coli K12 W2252-11U? RCrel (wild type II) is not sensitive to this phage infection, and its lipopolysaccharide does not inactivate the phage. Its lipopolysaccharide sugars are found to be D-glucose and D-ribose, thus differing from the lipopolysaccharide sugars of the wild type cells.  相似文献   

13.
Escherichia coli K-12 utilizes 3-(3-hydroxyphenyl)propionate (3HPP) as a sole carbon and energy source. Among the genes in its catabolic cluster in the genome, mhpT was proposed to encode a hypothetical transporter. Since no transporter for 3HPP uptake has been identified, we investigated whether MhpT is responsible for 3HPP uptake. MhpT fused with green fluorescent protein was found to be located at the periphery of cells by confocal microscopy, consistent with localization to the cytoplasmic membrane. Gene knockout and complementation studies clearly indicated that mhpT is essential for 3HPP catabolism in E. coli K-12 W3110 at pH 8.2. Uptake assays with 14C-labeled substrates demonstrated that strain W3110 and strain W3110ΔmhpT containing recombinant MhpT specifically transported 3HPP but not benzoate, 3-hydroxybenzoate, or gentisate into cells. Energy dependence assays suggested that MhpT-mediated 3HPP transport was driven by the proton motive force. The change of Ala-272 of MhpT to a histidine, surprisingly, resulted in enhanced transport activity, and strain W3110ΔmhpT containing the MhpT A272H mutation had a slightly higher growth rate than the wild-type strain at pH 8.2. Hence, we demonstrated that MhpT is a specific 3HPP transporter and vital for E. coli K-12 W3110 growth on this substrate under basic conditions.  相似文献   

14.
During lipopolysaccharide biosynthesis in several pathogens, including Burkholderia and Yersinia, 3-deoxy-d-manno-oct-2-ulosonic acid (Kdo) 3-hydroxylase, otherwise referred to as KdoO, converts Kdo to d-glycero-d-talo-oct-2-ulosonic acid (Ko) in an Fe(II)/α-ketoglutarate (α-KG)/O2-dependent manner. This conversion renders the bacterial outer membrane more stable and resistant to stresses such as an acidic environment. KdoO is a membrane-associated, deoxy-sugar hydroxylase that does not show significant sequence identity with any known enzymes, and its structural information has not been previously reported. Here, we report the biochemical and structural characterization of KdoO, Minf_1012 (KdoMI), from Methylacidiphilum infernorum V4. The de novo structure of KdoMI apoprotein indicates that KdoOMI consists of 13 α helices and 11 β strands, and has the jelly roll fold containing a metal binding motif, HXDX111H. Structures of KdoMI bound to Co(II), KdoMI bound to α-KG and Fe(III), and KdoMI bound to succinate and Fe(III), in addition to mutagenesis analysis, indicate that His146, His260, and Asp148 play critical roles in Fe(II) binding, while Arg127, Arg162, Arg174, and Trp176 stabilize α-KG. It was also observed that His225 is adjacent to the active site and plays an important role in the catalysis of KdoOMI without affecting substrate binding, possibly being involved in oxygen activation. The crystal structure of KdoOMI is the first completed structure of a deoxy-sugar hydroxylase, and the data presented here have provided mechanistic insights into deoxy-sugar hydroxylase, KdoO, and lipopolysaccharide biosynthesis.  相似文献   

15.
16.

Background

The receptor for advanced glycation end-products (RAGE) has been suggested to modulate lung injury in models of acute pulmonary inflammation. To study this further, model systems utilizing wild type and RAGE knockout (KO) mice were used to determine the role of RAGE signaling in lipopolysaccharide (LPS) and E. coli induced acute pulmonary inflammation. The effect of intraperitoneal (i.p.) and intratracheal (i.t.) administration of mouse soluble RAGE on E. coli injury was also investigated.

Methodology/Principal Findings

C57BL/6 wild type and RAGE KO mice received an i.t. instillation of LPS, E. coli, or vehicle control. Some groups also received i.p. or i.t. administration of mouse soluble RAGE. After 24 hours, the role of RAGE expression on inflammation was assessed by comparing responses in wild type and RAGE KO. RAGE protein levels decreased in wild type lung homogenates after treatment with either LPS or bacteria. In addition, soluble RAGE and HMGB1 increased in the BALF after E. coli instillation. RAGE KO mice challenged with LPS had the same degree of inflammation as wild type mice. However, when challenged with E. coli, RAGE KO mice had significantly less inflammation when compared to wild type mice. Most cytokine levels were lower in the BALF of RAGE KO mice compared to wild type mice after E. coli injury, while only monocyte chemotactic protein-1, MCP-1, was lower after LPS challenge. Neither i.p. nor i.t. administration of mouse soluble RAGE attenuated the severity of E. coli injury in wild type mice.

Conclusions/Significance

Lack of RAGE in the lung does not protect against LPS induced acute pulmonary inflammation, but attenuates injury following live E. coli challenge. These findings suggest that RAGE mediates responses to E. coli-associated pathogen-associated molecular pattern molecules other than LPS or other bacterial specific signaling responses. Soluble RAGE treatment had no effect on inflammation.  相似文献   

17.
A previously isolated T-even-type PP01 bacteriophage was used to detect its host cell, Escherichia coli O157:H7. The phage small outer capsid (SOC) protein was used as a platform to present a marker protein, green fluorescent protein (GFP), on the phage capsid. The DNA fragment around soc was amplified by PCR and sequenced. The gene alignment of soc and its upstream region was g56-soc.2-soc.1-soc, which is the same as that for T2 phage. GFP was introduced into the C- and N-terminal regions of SOC to produce recombinant phages PP01-GFP/SOC and PP01-SOC/GFP, respectively. Fusion of GFP to SOC did not change the host range of PP01. On the contrary, the binding affinity of the recombinant phages to the host cell increased. However, the stability of the recombinant phages in alkaline solution decreased. Adsorption of the GFP-labeled PP01 phages to the E. coli cell surface enabled visualization of cells under a fluorescence microscope. GFP-labeled PP01 phage was not only adsorbed on culturable E. coli cells but also on viable but nonculturable or pasteurized cells. The coexistence of insensitive E. coli K-12 (W3110) cells did not influence the specificity and affinity of GFP-labeled PP01 adsorption on E. coli O157:H7. After a 10-min incubation with GFP-labeled PP01 phage at a multiplicity of infection of 1,000 at 4°C, E. coli O157:H7 cells could be visualized by fluorescence microscopy. The GFP-labeled PP01 phage could be a rapid and sensitive tool for E. coli O157:H7 detection.  相似文献   

18.
To elucidate the minimal lipopolysaccharide (LPS) structure needed for the viability of Escherichia coli, suppressor-free strains lacking either the 3-deoxy-d-manno-oct-2-ulosonic acid transferase waaA gene or derivatives of the heptosyltransferase I waaC deletion with lack of one or all late acyltransferases (lpxL/M/P) and/or various outer membrane biogenesis factors were constructed. Δ(waaC lpxL lpxM lpxP) and waaA mutants exhibited highly attenuated growth, whereas simultaneous deletion of waaC and surA was lethal. Analyses of LPS of suppressor-free waaA mutants grown at 21 °C, besides showing accumulation of free lipid IVA precursor, also revealed the presence of its pentaacylated and hexaacylated derivatives, indicating in vivo late acylation can occur without Kdo. In contrast, LPS of Δ(waaC lpxL lpxM lpxP) strains showed primarily Kdo2-lipid IVA, indicating that these minimal LPS structures are sufficient to support growth of E. coli under slow-growth conditions at 21/23 °C. These lipid IVA derivatives could be modified biosynthetically by phosphoethanolamine, but not by 4-amino-4-deoxy-l-arabinose, indicating export defects of such minimal LPS. ΔwaaA and Δ(waaC lpxL lpxM lpxP) exhibited cell-division defects with a decrease in the levels of FtsZ and OMP-folding factor PpiD. These mutations led to strong constitutive additive induction of envelope responsive CpxR/A and σE signal transduction pathways. Δ(lpxL lpxM lpxP) mutant, with intact waaC, synthesized tetraacylated lipid A and constitutively incorporated a third Kdo in growth medium inducing synthesis of P-EtN and l-Ara4N. Overexpression of msbA restored growth of Δ(lpxL lpxM lpxP) under fast-growing conditions, but only partially that of the Δ(waaC lpxL lpxM lpxP) mutant. This suppression could be alleviated by overexpression of certain mutant msbA alleles or the single-copy chromosomal MsbA-498V variant in the vicinity of Walker-box II.Lipopolysacharides (LPS)4 are the major amphiphilic constituents of the outer leaflet of the outer membrane (OM) of Gram-negative bacteria, including Escherichia coli. LPS share a common architecture composed of a membrane-anchored phosphorylated and acylated β(1→6)-linked GlcN disaccharide, termed lipid A, to which a carbohydrate moiety of varying size is attached (1, 2). The latter may be divided into a lipid A proximal core oligosaccharide and, in smooth-type bacteria, a distal O-antigen. LPS always contain 3-deoxy-α-d-manno-oct-2-ulosonic acid (Kdo) linked to the lipid A.The physiological importance of the Kdo/lipid A region is reflected by its specific position within the pathway of LPS biosynthesis. In E. coli K-12, a bisphosphorylated lipid A precursor molecule with two amide and two ester-bound (R)-3-hydroxymyristate residues (lipid IVA) is synthesized from UDP-GlcNAc, following 6 distinct enzyme reactions (1). This intermediate serves as an acceptor for the Kdo transferase (WaaA), which transfers two Kdo residues from CMP-Kdo to yield an α(2→4)-linked Kdo disaccharide-attached α(2→6) to the non-reducing GlcN residue of lipid IVA (3). The latter reaction product, termed Kdo2-lipid IVA, comprises a key intermediate of LPS biosynthesis that acts 2-fold as a specific substrate: (i) for glycosyltransferases catalyzing further steps of the core oligosaccharide biosynthesis (4) and (ii) for acyltransferases that complete the lipid A moiety by the transfer of 2 additional fatty acids to the (R)-3-hydroxyl groups of both acyl chains, which are directly bound to position 2′ and 3′ of the non-reducing GlcN residue (1). Three acyltransferases, encoded by paralogous genes, have been described in E. coli K-12, which catalyze the latter enzyme reactions using acyl carrier protein-activated fatty acids as co-substrates (510). At ambient temperatures, a lauroyl residue is first transferred by LpxL (6) to the OH group of the amide-bound (R)-3-hydroxymyristate residue at position 2′. This catalytic step is partially replaced at low temperature (12 °C) by LpxP, which transfers palmitoleate to the same position in ∼80% of the LPS molecules (7). The free OH group of the ester-bound (R)-3-hydroxymyristate residue at position 3′ within both pentaacylated intermediates is then myristoylated by LpxM to give a hexaacylated lipid A moiety (Fig. 3) (5).Open in a separate windowFIGURE 3.Chemical structure of tetraacylated lipid IVA precursor (A) and Kdo2-lipid IVA (B). R1 represents C12:0 or C16:1; R2, C14:0; R3 and R4 are under LPS-modifying conditions P-EtN and l-Ara4N, respectively, and R5, C16:0.Consistent with the essentiality of LPS in E. coli, all the genes, whose products are required for committed steps of biosynthesis of lipid IVA and subsequent transfer of Kdo to it, are essential (1, 2). However, individually neither the subsequent steps of addition of the secondary lauroyl and myristoyl residues to the distal glucosoamine unit by LpxL and LpxM to synthesize hexaacylated lipid A nor the later glycosylation of hexaacylated Kdo2-lipid A is essential for viability of bacteria like E. coli K-12 under defined growth conditions (8). Although Re mutants that possess LPS with only hexaacylated Kdo2-lipid A or mutants that synthesize complete LPS core with only lipid IVA are viable, they are impaired in several growth properties, including constitutive induction of RpoE signal transduction in Re mutants (8, 1113). A triple null mutant, which lacks all 3 late acyltransferases, is viable but only in slow-growth conditions in accordance with lipid IVA being a poor substrate of the lipid A transporter MsbA (8). Mutants impaired in the synthesis of Kdo, which synthesize only lipid IVA lacking any glycosylation, can be constructed, but they require additional suppressor mutations either in msbA, or the yhjD gene (14, 15). Strains that potentially can only synthesize Kdo2-lipid IVA have not been reported up to now. Thus, suppressor-free minimal LPS structures that can support growth of E. coli K-12 bacteria known up to now have genetic compositions of Δ(lpxL lpxM lpxP) or Re mutants.We describe the construction and characterization of suppressor-free ΔwaaA and Δ(waaC lpxL lpxM lpxP) mutants, synthesizing either free lipid IVA derivatives or Kdo2-lipid IVA LPS, respectively. Analyses of lipid A of ΔwaaA also revealed the presence of free penta- and hexaacylated lipid A derivatives, arising due to incorporation of secondary acyl chains. Such suppressor-free strains could be constructed only in slow-growth conditions at lower temperatures. Growth of Δ(waaC lpxL lpxM lpxP) could be restored by extragenic chromosomal MsbA-D498V suppressor mutation or by the overexpression of the msbA wild-type gene product. The LPS of Δ(waaC lpxL lpxM lpxP) and lipid IVA precursor of ΔwaaA was found to be substituted by P-EtN, but not l-Ara4N, under LPS-modifying growth conditions. Deletion of late acyltransferases in ΔwaaC or deletion of the waaA gene resulted in constitutively elevated levels of periplasmic protease HtrA, due to additive induction of the envelope stress responsive CpxR/A two-component system and σE pathway.  相似文献   

19.
Xanthomonas axonopodis pv. citri (Xac) causes citrus canker, provoking defoliation and premature fruit drop with concomitant economical damage. In plant pathogenic bacteria, lipopolysaccharides are important virulence factors, and they are being increasingly recognized as major pathogen-associated molecular patterns for plants. In general, three domains are recognized in a lipopolysaccharide: the hydrophobic lipid A, the hydrophilic O-antigen polysaccharide, and the core oligosaccharide, connecting lipid A and O-antigen. In this work, we have determined the structure of purified lipopolysaccharides obtained from Xanthomonas axonopodis pv. citri wild type and a mutant of the O-antigen ABC transporter encoded by the wzt gene. High pH anion exchange chromatography and matrix-assisted laser desorption/ionization mass spectrum analysis were performed, enabling determination of the structure not only of the released oligosaccharides and lipid A moieties but also the intact lipopolysaccharides. The results demonstrate that Xac wild type and Xacwzt LPSs are composed mainly of a penta- or tetra-acylated diglucosamine backbone attached to either two pyrophosphorylethanolamine groups or to one pyrophosphorylethanolamine group and one phosphorylethanolamine group. The core region consists of a branched oligosaccharide formed by Kdo2Hex6GalA3Fuc3NAcRha4 and two phosphate groups. As expected, the presence of a rhamnose homo-oligosaccharide as O-antigen was determined only in the Xac wild type lipopolysaccharide. In addition, we have examined how lipopolysaccharides from Xac function in the pathogenesis process. We analyzed the response of the different lipopolysaccharides during the stomata aperture closure cycle, the callose deposition, the expression of defense-related genes, and reactive oxygen species production in citrus leaves, suggesting a functional role of the O-antigen from Xac lipopolysaccharides in the basal response.  相似文献   

20.

Escherichia coli strains W3110 and BL21 were engineered for the production of plasmid DNA (pDNA) under aerobic and transitions to microaerobic conditions. The gene coding for recombinase A (recA) was deleted in both strains. In addition, the Vitreoscilla hemoglobin (VHb) gene (vgb) was chromosomally inserted and constitutively expressed in each E. coli recA mutant and wild type. The recA inactivation increased the supercoiled pDNA fraction (SCF) in both strains, while VHb expression improved the pDNA production in W3110, but not in BL21. Therefore, a codon-optimized version of vgb was inserted in strain BL21recA, which, together with W3110recAvgb+, was tested in cultures with shifts from aerobic to oxygen-limited regimes. VHb expression lowered the accumulation of fermentative by-products in both strains. VHb-expressing cells displayed higher oxidative activity as indicated by the Redox Sensor Green fluorescence, which was more intense in BL21 than in W3110. Furthermore, VHb expression did not change pDNA production in W3110, but decreased it in BL21. These results are useful for understanding the physiological effects of VHb expression in two industrially relevant E. coli strains, and for the selection of a host for pDNA production.

  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