(Below N is a link to NCBI taxonomic web page and E link to ESTHER at designed phylum.) > cellular organisms: NE > Bacteria: NE > Proteobacteria: NE > Gammaproteobacteria: NE > Pseudomonadales: NE > Pseudomonadaceae: NE > Pseudomonas: NE > Pseudomonas aeruginosa group: NE > Pseudomonas aeruginosa: NE
Warning: This entry is a compilation of different species or line or strain with more than 90% amino acide identity. You can retrieve all strain data
(Below N is a link to NCBI taxonomic web page and E link to ESTHER at designed phylum.) Pseudomonas aeruginosa C3719: N, E.
Pseudomonas aeruginosa 2192: N, E.
Pseudomonas aeruginosa VRFPA04: N, E.
Pseudomonas aeruginosa VRFPA01: N, E.
Pseudomonas aeruginosa LESB58: N, E.
Pseudomonas aeruginosa UCBPP-PA14: N, E.
Pseudomonas aeruginosa PA7: N, E.
Pseudomonas aeruginosa 39016: N, E.
Pseudomonas aeruginosa BL12: N, E.
Pseudomonas aeruginosa PAO581: N, E.
Pseudomonas aeruginosa MPAO1/P2: N, E.
Pseudomonas aeruginosa ATCC 14886: N, E.
Pseudomonas aeruginosa BL01: N, E.
Pseudomonas aeruginosa ATCC 700888: N, E.
Pseudomonas aeruginosa ATCC 25324: N, E.
Pseudomonas aeruginosa BL16: N, E.
Pseudomonas aeruginosa BWH050: N, E.
Pseudomonas aeruginosa DHS01: N, E.
Pseudomonas aeruginosa BWHPSA005: N, E.
Pseudomonas aeruginosa BWHPSA003: N, E.
Pseudomonas aeruginosa PADK2_CF510: N, E.
Pseudomonas aeruginosa VRFPA05: N, E.
Pseudomonas aeruginosa BWHPSA028: N, E.
Pseudomonas aeruginosa M9A.1: N, E.
Pseudomonas aeruginosa PA21_ST175: N, E.
Pseudomonas aeruginosa BWHPSA046: N, E.
Pseudomonas aeruginosa UDL: N, E.
Pseudomonas aeruginosa BWHPSA027: N, E.
Pseudomonas aeruginosa 3574: N, E.
Pseudomonas aeruginosa CF127: N, E.
Pseudomonas aeruginosa PAO1-VE13: N, E.
Pseudomonas aeruginosa BWHPSA024: N, E.
Pseudomonas aeruginosa PA1: N, E.
Pseudomonas aeruginosa BWHPSA043: N, E.
Pseudomonas aeruginosa DK2: N, E.
Pseudomonas aeruginosa BL05: N, E.
Pseudomonas aeruginosa BWH035: N, E.
Pseudomonas aeruginosa VRFPA02: N, E.
Pseudomonas aeruginosa C51: N, E.
Pseudomonas aeruginosa S54485: N, E.
Pseudomonas aeruginosa BWHPSA001: N, E.
Pseudomonas aeruginosa BWHPSA019: N, E.
Pseudomonas aeruginosa BWHPSA002: N, E.
Pseudomonas aeruginosa BWHPSA038: N, E.
Pseudomonas aeruginosa BWH051: N, E.
Pseudomonas aeruginosa BWHPSA047: N, E.
Pseudomonas aeruginosa BWHPSA022: N, E.
Pseudomonas aeruginosa BL03: N, E.
Pseudomonas aeruginosa 3577: N, E.
Pseudomonas aeruginosa str. Stone 130: N, E.
Pseudomonas aeruginosa c7447m: N, E.
Pseudomonas aeruginosa BWHPSA044: N, E.
Pseudomonas aeruginosa C52: N, E.
Pseudomonas aeruginosa BL25: N, E.
Pseudomonas aeruginosa 18A: N, E.
Pseudomonas aeruginosa LESlike5: N, E.
Pseudomonas aeruginosa PAO1-VE2: N, E.
Pseudomonas aeruginosa 3576: N, E.
Pseudomonas aeruginosa 3573: N, E.
Pseudomonas aeruginosa BL17: N, E.
Pseudomonas aeruginosa U2504: N, E.
Pseudomonas aeruginosa BL24: N, E.
Pseudomonas aeruginosa CF5: N, E.
Pseudomonas aeruginosa BL22: N, E.
Pseudomonas aeruginosa BWHPSA007: N, E.
Pseudomonas aeruginosa PAO1-GFP: N, E.
Pseudomonas aeruginosa VRFPA08: N, E.
Pseudomonas aeruginosa M8A.4: N, E.
Pseudomonas aeruginosa BWHPSA015: N, E.
Pseudomonas aeruginosa S35004: N, E.
Pseudomonas aeruginosa PAO579: N, E.
Pseudomonas aeruginosa M8A.1: N, E.
Pseudomonas aeruginosa VRFPA06: N, E.
Pseudomonas aeruginosa CIG1: N, E.
Pseudomonas aeruginosa BL06: N, E.
Pseudomonas aeruginosa LESlike1: N, E.
Pseudomonas aeruginosa MSH-10: N, E.
Pseudomonas aeruginosa BL07: N, E.
Pseudomonas aeruginosa PA1R: N, E.
Pseudomonas aeruginosa M10: N, E.
Pseudomonas aeruginosa IGB83: N, E.
Pseudomonas aeruginosa BL02: N, E.
Pseudomonas aeruginosa PAO1: N, E.
Pseudomonas aeruginosa BWHPSA020: N, E.
Pseudomonas aeruginosa ID4365: N, E.
Pseudomonas aeruginosa C40: N, E.
Pseudomonas aeruginosa 19660: N, E.
Pseudomonas aeruginosa BWHPSA004: N, E.
Pseudomonas aeruginosa MTB-1: N, E.
Pseudomonas aeruginosa BWH036: N, E.
Pseudomonas aeruginosa BWHPSA010: N, E.
Pseudomonas aeruginosa LESlike7: N, E.
Pseudomonas aeruginosa 3581: N, E.
Pseudomonas aeruginosa BWH030: N, E.
Pseudomonas aeruginosa C41: N, E.
Pseudomonas aeruginosa VRFPA07: N, E.
Pseudomonas aeruginosa BWH057: N, E.
Pseudomonas aeruginosa HB13: N, E.
Pseudomonas aeruginosa BWH059: N, E.
Pseudomonas aeruginosa CF27: N, E.
Pseudomonas aeruginosa PA99: N, E.
Pseudomonas aeruginosa HB15: N, E.
Pseudomonas aeruginosa BWH049: N, E.
Pseudomonas aeruginosa LESlike4: N, E.
Pseudomonas aeruginosa CF77: N, E.
Pseudomonas aeruginosa MH27: N, E.
Pseudomonas aeruginosa SCV20265: N, E.
Pseudomonas aeruginosa BWH032: N, E.
Pseudomonas aeruginosa BWHPSA008: N, E.
Pseudomonas aeruginosa BL11: N, E.
Pseudomonas aeruginosa PA14: N, E.
Pseudomonas aeruginosa BL04: N, E.
Pseudomonas aeruginosa BWHPSA014: N, E.
Pseudomonas aeruginosa BWH053: N, E.
Pseudomonas aeruginosa BWH054: N, E.
Pseudomonas aeruginosa CF614: N, E.
Pseudomonas aeruginosa NCGM2.S1: N, E.
Pseudomonas aeruginosa MSH10: N, E.
Pseudomonas aeruginosa 3579: N, E.
Pseudomonas aeruginosa MSH3: N, E.
Pseudomonas aeruginosa C20: N, E.
Pseudomonas aeruginosa BWH058: N, E.
Pseudomonas aeruginosa Z61: N, E.
Pseudomonas aeruginosa YL84: N, E.
Pseudomonas aeruginosa BWHPSA018: N, E.
Pseudomonas aeruginosa BWH055: N, E.
Pseudomonas aeruginosa BL15: N, E.
Pseudomonas aeruginosa BWHPSA042: N, E.
Pseudomonas aeruginosa BWHPSA021: N, E.
Pseudomonas aeruginosa CF18: N, E.
Pseudomonas aeruginosa BWH033: N, E.
Pseudomonas aeruginosa 148: N, E.
Pseudomonas aeruginosa PA103: N, E.
Pseudomonas aeruginosa BL23: N, E.
Pseudomonas aeruginosa MPAO1/P1: N, E.
Pseudomonas aeruginosa 3580: N, E.
Pseudomonas aeruginosa 62: N, E.
Pseudomonas aeruginosa 6077: N, E.
Pseudomonas aeruginosa C48: N, E.
Pseudomonas aeruginosa BWHPSA041: N, E.
Pseudomonas aeruginosa BWHPSA023: N, E.
Pseudomonas aeruginosa BWHPSA045: N, E.
Pseudomonas aeruginosa BL14: N, E.
Pseudomonas aeruginosa BL08: N, E.
Pseudomonas aeruginosa BWHPSA017: N, E.
Pseudomonas aeruginosa BWHPSA039: N, E.
Pseudomonas aeruginosa DHS29: N, E.
Pseudomonas aeruginosa CI27: N, E.
Pseudomonas aeruginosa X24509: N, E.
Pseudomonas aeruginosa BWHPSA040: N, E.
Pseudomonas aeruginosa 3578: N, E.
Pseudomonas aeruginosa PS75: N, E.
Pseudomonas aeruginosa RP73: N, E.
Pseudomonas aeruginosa BWHPSA013: N, E.
Pseudomonas aeruginosa PA96: N, E.
Pseudomonas aeruginosa E2: N, E.
Pseudomonas aeruginosa BWH031: N, E.
Pseudomonas aeruginosa JJ692: N, E.
Pseudomonas aeruginosa MH38: N, E.
Pseudomonas aeruginosa B136-33: N, E.
Pseudomonas aeruginosa BWHPSA009: N, E.
Pseudomonas aeruginosa PA38182: N, E.
Pseudomonas aeruginosa WC55: N, E.
Pseudomonas aeruginosa BWHPSA026: N, E.
Pseudomonas aeruginosa X13273: N, E.
Pseudomonas aeruginosa C23: N, E.
Pseudomonas aeruginosa BL18: N, E.
Pseudomonas aeruginosa SG17M: N, E.
Pseudomonas aeruginosa PA45: N, E.
Pseudomonas aeruginosa BWH060: N, E.
Pseudomonas aeruginosa BWHPSA016: N, E.
Pseudomonas aeruginosa NCMG1179: N, E.
Pseudomonas aeruginosa BWHPSA025: N, E.
Pseudomonas aeruginosa BWH029: N, E.
Pseudomonas aeruginosa BL21: N, E.
Pseudomonas aeruginosa BL20: N, E.
Pseudomonas aeruginosa BWHPSA006: N, E.
Pseudomonas aeruginosa M18: N, E.
Pseudomonas aeruginosa BWHPSA037: N, E.
Pseudomonas aeruginosa PAK: N, E.
Pseudomonas aeruginosa BWH056: N, E.
Pseudomonas aeruginosa BL09: N, E.
Pseudomonas aeruginosa PS50: N, E.
Pseudomonas aeruginosa BWHPSA011: N, E.
Pseudomonas aeruginosa BL19: N, E.
Pseudomonas aeruginosa M8A.2: N, E.
Pseudomonas aeruginosa BWH052: N, E.
Pseudomonas aeruginosa BWHPSA012: N, E.
Pseudomonas aeruginosa LES400: N, E.
Pseudomonas aeruginosa BL10: N, E.
Pseudomonas aeruginosa VRFPA03: N, E.
Pseudomonas aeruginosa LES431: N, E.
Pseudomonas aeruginosa PS42: N, E.
Pseudomonas aeruginosa BL13: N, E.
Pseudomonas aeruginosa M8A.3: N, E.
Pseudomonas aeruginosa LESB65: N, E.
Pseudomonas aeruginosa 3575: N, E.
Pseudomonas aeruginosa DK1: N, E.
Pseudomonas sp. 42A2: N, E.
Molecular evidence
Database
No mutation 1 structure: 1EX9: Pseudomonas aeruginosa lipase complexed with RC-(RP,SP)-1,2-dioctylcarbamoyl-glycero-3-O-octylphosphonate No kinetic
LegendThis sequence has been compared to family alignement (MSA) red => minority aminoacid blue => majority aminoacid color intensity => conservation rate title => sequence position(MSA position)aminoacid rate Catalytic site Catalytic site in the MSA MKKKSLLPLGLAIGLASLAASPLIQASTYTQTKYPIVLAHGMLGFDNILG VDYWFGIPSALRRDGAQVYVTEVSQLDTSEVRGEQLLQQVEEIVALSGQP KVNLIGHSHGGPTIRYVAAVRPDLIASATSVGAPHKGSDTADFLRQIPPG SAGEAVLSGLVNSLGALISFLSSGSTGTQNSLGSLESLNSEGAARFNAKY PQGIPTSACGEGAYKVNGVSYYSWSGSSPLTNFLDPSDAFLGASSLTFKN GTANDGLVGTCSSHLGMVIRDNYRMNHLDEVNQVFGLTSLFETSPVSVYR QHANRLKNASL
Lipases are essential and widely used biocatalysts. Hence, the production of lipases requires a detailed understanding of the molecular mechanism of its folding and secretion. Lipase A from Pseudomonas aeruginosa, PaLipA, constitutes a prominent example that has additional relevance because of its role as a virulence factor in many diseases. PaLipA requires the assistance of a membrane-integrated steric chaperone, the lipase-specific foldase Lif, to achieve its enzymatically active state. However, the molecular mechanism of how Lif activates its cognate lipase has remained elusive. Here, we show by molecular dynamics simulations at the atomistic level and potential of mean force computations that Lif catalyzes the activation process of PaLipA by structurally stabilizing an intermediate PaLipA conformation, particularly a beta-sheet in the region of residues 17-30, such that the opening of PaLipA's lid domain is facilitated. This opening allows substrate access to PaLipA's catalytic site. A surprising and so far not fully understood aspect of our study is that the open state of PaLipA is unstable compared to the closed one according to our computational and in vitro biochemical results. We thus speculate that further interactions of PaLipA with the Xcp secretion machinery and/or components of the extracellular matrix contribute to the remaining activity of secreted PaLipA. (c) 2019 Wiley Periodicals, Inc.
        
Title: Understanding domain movements and interactions of Pseudomonas aeruginosa lipase with lipid molecule tristearoyl glycerol: A molecular dynamics approach Thiruvengadam K, Baskaran SK, Pennathur G Ref: J Mol Graph Model, 85:190, 2018 : PubMed
Lipases are biocatalysts which exhibit optimal activity at the aqueous-lipid interface. Molecular Dynamics (MD) Simulation studies on lipases have revealed the structural changes occurring in the enzyme, at the loop-helix-loop, often designated as the "lid", which is responsible for its interfacial activation. In recent years, MD simulation of lipases at molecular level have been studied in detail, whereas very few studies are carried over on its interaction with lipid molecules. Hence, in the current study we have investigated molecular interaction of bacterial lipase (Pseudomonas aeruginosa lipase, PAL) with a lipid molecule (tristearoyl glycerol, TGL). This provides an insight into the interfacial activation of the enzyme. The lipid molecule was placed near the lids of the enzyme and MD simulations were performed for 100 ns to understand the nature and site of the interaction. The results clearly indicate that, the presence of a lipid molecule near the lids affects the motion of the enzyme through changes in conformation. Lipid molecule near the lids reduces the movements of both lids, and the TGL molecule was observed moving towards the active site. The movement of the lids, surface accessibility and the domain movements of PAL are discussed and the results provide valuable insight in to the role played by the two lids in the interfacial activation of PAL with TGL.
A special class of proteins adopts an inactive conformation in aqueous solution and activates at an interface (such as the surface of lipid droplet) by switching their conformations. Lipase, an essential enzyme for breaking down lipids, serves as a model system for studying such interfacial proteins. The underlying conformational switch of lipase induced by solvent condition is achieved through changing the status of the gated substrate-access channel. Interestingly, a lipase was also reported to exhibit pressure activation, which indicates it is drastically active at high hydrostatic pressure. To unravel the molecular mechanism of this unusual phenomenon, we examined the structural changes induced by high hydrostatic pressures (up to 1500 MPa) using molecular dynamics simulations. By monitoring the width of the access channel, we found that the protein undergoes a conformational transition and opens the access channel at high pressures (>100 MPa). Particularly, a disordered amphiphilic alpha5 region of the protein becomes ordered at high pressure. This positive correlation between the channel opening and alpha5 ordering is consistent with the early findings of the gating motion in the presence of a water-oil interface. Statistical analysis of the ensemble of conformations also reveals the essential collective motions of the protein and how these motions contribute to gating. Arguments are presented as to why heightened sensitivity to high-pressure perturbation can be a general feature of switchable interfacial proteins. Further mutations are also suggested to validate our observations. Proteins 2016; 84:820-827. (c) 2016 Wiley Periodicals, Inc.
Lipases are essential and widely used biocatalysts. Hence, the production of lipases requires a detailed understanding of the molecular mechanism of its folding and secretion. Lipase A from Pseudomonas aeruginosa, PaLipA, constitutes a prominent example that has additional relevance because of its role as a virulence factor in many diseases. PaLipA requires the assistance of a membrane-integrated steric chaperone, the lipase-specific foldase Lif, to achieve its enzymatically active state. However, the molecular mechanism of how Lif activates its cognate lipase has remained elusive. Here, we show by molecular dynamics simulations at the atomistic level and potential of mean force computations that Lif catalyzes the activation process of PaLipA by structurally stabilizing an intermediate PaLipA conformation, particularly a beta-sheet in the region of residues 17-30, such that the opening of PaLipA's lid domain is facilitated. This opening allows substrate access to PaLipA's catalytic site. A surprising and so far not fully understood aspect of our study is that the open state of PaLipA is unstable compared to the closed one according to our computational and in vitro biochemical results. We thus speculate that further interactions of PaLipA with the Xcp secretion machinery and/or components of the extracellular matrix contribute to the remaining activity of secreted PaLipA. (c) 2019 Wiley Periodicals, Inc.
        
Title: Understanding domain movements and interactions of Pseudomonas aeruginosa lipase with lipid molecule tristearoyl glycerol: A molecular dynamics approach Thiruvengadam K, Baskaran SK, Pennathur G Ref: J Mol Graph Model, 85:190, 2018 : PubMed
Lipases are biocatalysts which exhibit optimal activity at the aqueous-lipid interface. Molecular Dynamics (MD) Simulation studies on lipases have revealed the structural changes occurring in the enzyme, at the loop-helix-loop, often designated as the "lid", which is responsible for its interfacial activation. In recent years, MD simulation of lipases at molecular level have been studied in detail, whereas very few studies are carried over on its interaction with lipid molecules. Hence, in the current study we have investigated molecular interaction of bacterial lipase (Pseudomonas aeruginosa lipase, PAL) with a lipid molecule (tristearoyl glycerol, TGL). This provides an insight into the interfacial activation of the enzyme. The lipid molecule was placed near the lids of the enzyme and MD simulations were performed for 100 ns to understand the nature and site of the interaction. The results clearly indicate that, the presence of a lipid molecule near the lids affects the motion of the enzyme through changes in conformation. Lipid molecule near the lids reduces the movements of both lids, and the TGL molecule was observed moving towards the active site. The movement of the lids, surface accessibility and the domain movements of PAL are discussed and the results provide valuable insight in to the role played by the two lids in the interfacial activation of PAL with TGL.
A special class of proteins adopts an inactive conformation in aqueous solution and activates at an interface (such as the surface of lipid droplet) by switching their conformations. Lipase, an essential enzyme for breaking down lipids, serves as a model system for studying such interfacial proteins. The underlying conformational switch of lipase induced by solvent condition is achieved through changing the status of the gated substrate-access channel. Interestingly, a lipase was also reported to exhibit pressure activation, which indicates it is drastically active at high hydrostatic pressure. To unravel the molecular mechanism of this unusual phenomenon, we examined the structural changes induced by high hydrostatic pressures (up to 1500 MPa) using molecular dynamics simulations. By monitoring the width of the access channel, we found that the protein undergoes a conformational transition and opens the access channel at high pressures (>100 MPa). Particularly, a disordered amphiphilic alpha5 region of the protein becomes ordered at high pressure. This positive correlation between the channel opening and alpha5 ordering is consistent with the early findings of the gating motion in the presence of a water-oil interface. Statistical analysis of the ensemble of conformations also reveals the essential collective motions of the protein and how these motions contribute to gating. Arguments are presented as to why heightened sensitivity to high-pressure perturbation can be a general feature of switchable interfacial proteins. Further mutations are also suggested to validate our observations. Proteins 2016; 84:820-827. (c) 2016 Wiley Periodicals, Inc.
        
Title: Solvent-Dependent Gating Motions of an Extremophilic Lipase from Pseudomonas aeruginosa Johnson QR, Nellas RB, Shen T Ref: Biochemistry, 51:6238, 2012 : PubMed
Understanding how organic solvent-stable proteins can function in anhydrous and often complex solutions is essential for the study of the interaction of protein and molecular immiscible interfaces and the design of efficient industrial enzymes in nonaqueous solvents. Using an extremophilic lipase from Pseudomonas aeruginosa as an example, we investigated the conformational dynamics of an organic solvent-tolerant enzyme in complex solvent milieux. Four 100-ns molecular dynamics simulations of the lipase were performed in solvent systems: water, hexane, and two mixtures of hexane and water, 5% and 95% (w/w) hexane. Our results show a solvent-dependent structural change of the protein, especially in the region that regulates the admission of the substrate. We observed that the lipase is much less flexible in hexane than in aqueous solution or at the immiscible interface. Quantified by the size of the accessible channel, the lipase in water has a closed-gate conformation and no access to the active site, while in the hexane-containing systems, the lipase is at various degrees of open-gate state, with the immiscible interface setup being in the widely open conformation ensembles. The composition of explicit solvents in the access channel showed a significant influence on the conformational dynamics of the protein. Interestingly, the slowest step (bottleneck) of the hexane-induced conformational switch seems to be correlated with the slow dehydration dynamics of the channel.
Pseudomonas aeruginosa is a common opportunistic bacterial pathogen that causes a variety of infections in humans. Populations of P. aeruginosa are dominated by common clones that can be isolated from diverse clinical and environmental sources. To determine whether specific clones are associated with corneal infection, we used a portable genotyping microarray system to analyze a set of 63 P. aeruginosa isolates from patients with corneal ulcers (keratitis). We then used population analysis to compare the keratitis isolates to a wider collection of P. aeruginosa from various nonocular sources. We identified various markers in a subpopulation of P. aeruginosa associated with keratitis that were in strong disequilibrium with the wider P. aeruginosa population, including oriC, exoU, katN, unmodified flagellin, and the carriage of common genomic islands. The genome sequencing of a keratitis isolate (39016; representing the dominant serotype O11), which was associated with a prolonged clinical healing time, revealed several genomic islands and prophages within the accessory genome. The PCR amplification screening of all 63 keratitis isolates, however, provided little evidence for the shared carriage of specific prophages or genomic islands between serotypes. P. aeruginosa twitching motility, due to type IV pili, is implicated in corneal virulence. We demonstrated that 46% of the O11 keratitis isolates, including 39016, carry a distinctive pilA, encoding the pilin of type IV pili. Thus, the keratitis isolates were associated with specific characteristics, indicating that a subpopulation of P. aeruginosa is adapted to cause corneal infection.
        
Title: Differential behaviour of Pseudomonas sp. 42A2 LipC, a lipase showing greater versatility than its counterpart LipA Bofill C, Prim N, Mormeneo M, Manresa A, Pastor FI, Diaz P Ref: Biochimie, 92:307, 2010 : PubMed
Growth of Pseudomonas sp. 42A2 on oleic acid releases polymerized hydroxy-fatty acids as a result of several enzymatic conversions that could involve one or more lipases. To test this hypothesis, the lipolytic system of strain Pseudomonas sp. 42A2 was analyzed, revealing the presence of at least an intracellular carboxylesterase and a secreted lipase. Consensus primers derived from a conserved region of bacterial lipase subfamilies I.1 and I.2 allowed isolation of two secreted lipase genes, lipA and lipC, highly homologous to those of Pseudomonas aeruginosa PAO1. Homologous cloning of the isolated lipA and lipC genes was performed in Pseudomonas sp. 42A2 for LipA and LipC over-expression. The overproduced lipases were further purified and characterized, both showing preference for medium fatty acid chain-length substrates. However, significant differences could be detected between LipA and LipC in terms of enzyme kinetics and behaviour pattern. Accordingly, LipA showed maximum activity at moderate temperatures, and displayed a typical Michaelis-Menten kinetics. On the contrary, LipC was more active at low temperatures and displayed partial interfacial activation, showing a shift in substrate specificity when assayed at different temperatures, and displaying increased activity in the presence of certain heavy metal ions. The versatile properties shown by LipC suggest that this lipase could be expressed in response to variable environmental conditions.
        
Title: Co-expression of the lipase and foldase of Pseudomonas aeruginosa to a functional lipase in Escherichia coli Madan B, Mishra P Ref: Applied Microbiology & Biotechnology, 85:597, 2010 : PubMed
The lipA gene, a structural gene encoding for protein of molecular mass 48 kDa, and lipB gene, encoding for a lipase-specific chaperone with molecular mass of 35 kDa, of Pseudomonas aeruginosa B2264 were co-expressed in heterologous host Escherichia coli BL21 (DE3) to obtain in vivo expression of functional lipase. The recombinant lipase was expressed with histidine tag at its N terminus and was purified to homogeneity using nickel affinity chromatography. The amino acid sequence of LipA and LipB of P. aeruginosa B2264 was 99-100% identical with the corresponding sequence of LipA and LipB of P. aeruginosa LST-03 and P. aeruginosa PA01, but it has less identity with Pseudomonas cepacia (Burkholderia cepacia) as it showed only 37.6% and 23.3% identity with the B. cepacia LipA and LipB sequence, respectively. The molecular mass of the recombinant lipase was found to be 48 kDa. The recombinant lipase exhibited optimal activity at pH 8.0 and 37 degrees C, though it was active between pH 5.0 and pH 9.0 and up to 45 degrees C. K (m) and V (max) values for recombinant P. aeruginosa lipase were found to be 151.5 +/- 29 microM and 217 +/- 22.5 micromol min(-1) mg(-1) protein, respectively.
Pseudomonas aeruginosa isolates have a highly conserved core genome representing up to 90% of the total genomic sequence with additional variable accessory genes, many of which are found in genomic islands or islets. The identification of the Liverpool Epidemic Strain (LES) in a children's cystic fibrosis (CF) unit in 1996 and its subsequent observation in several centers in the United Kingdom challenged the previous widespread assumption that CF patients acquire only unique strains of P. aeruginosa from the environment. To learn about the forces that shaped the development of this important epidemic strain, the genome of the earliest archived LES isolate, LESB58, was sequenced. The sequence revealed the presence of many large genomic islands, including five prophage clusters, one defective (pyocin) prophage cluster, and five non-phage islands. To determine the role of these clusters, an unbiased signature tagged mutagenesis study was performed, followed by selection in the chronic rat lung infection model. Forty-seven mutants were identified by sequencing, including mutants in several genes known to be involved in Pseudomonas infection. Furthermore, genes from four prophage clusters and one genomic island were identified and in direct competition studies with the parent isolate; four were demonstrated to strongly impact on competitiveness in the chronic rat lung infection model. This strongly indicates that enhanced in vivo competitiveness is a major driver for maintenance and diversifying selection of these genomic prophage genes.
One of the hallmarks of the Gram-negative bacterium Pseudomonas aeruginosa is its ability to thrive in diverse environments that includes humans with a variety of debilitating diseases or immune deficiencies. Here we report the complete sequence and comparative analysis of the genomes of two representative P. aeruginosa strains isolated from cystic fibrosis (CF) patients whose genetic disorder predisposes them to infections by this pathogen. The comparison of the genomes of the two CF strains with those of other P. aeruginosa presents a picture of a mosaic genome, consisting of a conserved core component, interrupted in each strain by combinations of specific blocks of genes. These strain-specific segments of the genome are found in limited chromosomal locations, referred to as regions of genomic plasticity. The ability of P. aeruginosa to shape its genomic composition to favor survival in the widest range of environmental reservoirs, with corresponding enhancement of its metabolic capacity is supported by the identification of a genomic island in one of the sequenced CF isolates, encoding enzymes capable of degrading terpenoids produced by trees. This work suggests that niche adaptation is a major evolutionary force influencing the composition of bacterial genomes. Unlike genome reduction seen in host-adapted bacterial pathogens, the genetic capacity of P. aeruginosa is determined by the ability of individual strains to acquire or discard genomic segments, giving rise to strains with customized genomic repertoires. Consequently, this organism can survive in a wide range of environmental reservoirs that can serve as sources of the infecting organisms.
        
Title: Activation Mechanisms for Pseudomonas species Lipase by Cardiovascular Drugs in Vitro Lin MC, Lai GW, Lin CS, Lin G Ref: Journal of the Chinese Chemical Society, 54:1601, 2007 : PubMed
The goal of this work was to determine the reaction mechanism for Pseudomonas species lipase activation by cardiovascular drugs such as lovastatin, simvastatin, amlodipine besylate, nifedipine, and hydralazine hydrochloride based on the results from enzyme kinetics. These drugs are the essential activators of Pseudomonas species lipase in the presence of triton-X 100 or taurochloate in vitro. Moreover, QSAR studies show that the pKA values are correlated with the molecular weights of these drugs.
Organic solvent-tolerant Pseudomonas aeruginosa LST-03 secretes an organic solvent-stable lipase, LST-03 lipase. The gene of the LST-03 lipase (Lip9) and the gene of the lipase-specific foldase (Lif9) were cloned and expressed in Escherichia coli. In the cloned 2.6 kbps DNA fragment, two open reading frames, Lip9 consisting of 933 nucleotides which encoded 311 amino acids and Lif9 consisting of 1,020 nucleotides which encoded 340 amino acids, were found. The overexpression of the lipase gene (lip9) was achieved when T7 promoter was used and the signal peptide of the lipase was deleted. The expressed amount of the lipase was greatly increased and overexpressed lipase formed inclusion body in E. coli cell. The collected inclusion body of the lipase from the cell was easily solubilized by urea and activated by using lipase-specific foldase of which 52 or 58 amino acids of N-terminal were deleted. Especially, the N-terminal methionine of the lipase of which the signal peptide was deleted was released in E. coli and the amino acid sequence was in agreement with that of the originally-produced lipase by P. aeruginosa LST-03. Furthermore, the overexpressed and solubilized lipase of which the signal peptide was deleted was more effectively activated by lipase-specific foldase.
In a previous paper, the combinatorial active-site saturation test (CAST) was introduced as an effective strategy for the directed evolution of enzymes toward broader substrate acceptance. CASTing comprises the systematic design and screening of focused libraries around the complete binding pocket, but it is only the first step of an evolutionary process because only the initial libraries of mutants are considered. In the present study, a simple method is presented for further optimization of initial hits by combining the mutational changes obtained from two different libraries. Combined lipase mutants were screened for hydrolytic activity against six notoriously difficult substrates (bulky carboxylic acid esters) and improved mutants showing significantly higher activity were identified. The enantioselectivity of the mutants in the hydrolytic kinetic resolution of two substrates was also studied, with the best mutant-substrate combination resulting in a selectivity factor of E=49. Finally, the catalytic profile of the evolved mutants in the hydrolysis of simple nonbranched carboxylic acid esters, ranging from acetate to palmitate, was studied for theoretical reasons.
        
Title: Directed evolution of Pseudomonas aeruginosa lipase for improved amide-hydrolyzing activity Fujii R, Nakagawa Y, Hiratake J, Sogabe A, Sakata K Ref: Protein Engineering Des Sel, 18:93, 2005 : PubMed
A lipase from Pseudomonas aeruginosa was subjected to directed molecular evolution for increased amide-hydrolyzing (amidase) activity. A single round of random mutagenesis followed by screening for hydrolytic activity for oleoyl 2-naphthylamide as compared with that for oleoyl 2-naphthyl ester identified five mutants with 1.7-2.0-fold increased relative amidase activities. Three mutational sites (F207S, A213D and F265L) were found to affect the amidase/esterase activity ratios. The combination of these mutations further improved the amidase activity. Active-site titration using a fluorescent phosphonic acid ester allowed the molecular activities for the amide and the ester to be determined for each mutant without purification of the lipase. A double mutant F207S/A213D gave the highest molecular activity of 1.1 min(-1) for the amide, corresponding to a 2-fold increase compared with that of the wild-type lipase. A structural model of the lipase indicated that the mutations occurred at the sites near the surface and remote from the catalytic triad, but close to the calcium binding site. This study is a first step towards understanding why lipases do not hydrolyze amides despite the similarities to serine proteases in the active site structure and the reaction mechanism and towards the preparation of a general acyl transfer catalyst for the biotransformation of amides.
        
Title: Expanding the range of substrate acceptance of enzymes: combinatorial active-site saturation test Reetz MT, Bocola M, Carballeira JD, Zha D, Vogel A Ref: Angew Chem Int Ed Engl, 44:4192, 2005 : PubMed
Herein, we describe a method for expanding the scope of substrate acceptance of a given enzyme with the aim of including a wide range of structurally different compounds. Two straightforward steps are required: the design and the generation of relatively small focused libraries of enzyme mutants produced by randomization at several sets of two spatially close amino acid positions around the active site. The choice of two amino acids which are spatially close to one another allows for potential synergistic conformational effects arising from side-chain orientations, an unpredictable phenomenon which cannot be brought about by single-site saturation mutagenesis. The optimal choice of the respective pairs of amino acids is guided by the 3D structure of the wild-type (WT) enzyme with a bound substrate. Geometric inspection allows the definition of sites at which the side chains of the individual amino acids in each pair point toward the binding site of the WT enzyme
        
Title: Learning from directed evolution: theoretical investigations into cooperative mutations in lipase enantioselectivity Bocola M, Otte N, Jaeger KE, Reetz MT, Thiel W Ref: Chembiochem, 5:214, 2004 : PubMed
Molecular modeling with classical force-fields has been used to study the reactant complex and the tetrahedral intermediate in lipase-catalyzed ester hydrolysis in 20 enzyme/substrate combinations. The R and S enantiomers of alpha-methyldecanoic acid ester served as substrates for the wild-type lipase from Pseudomonas aeruginosa and nine selected mutants. After suitable preparation of initial structures from an available wild-type crystal structure, each system was subjected to 1 ns CHARMM force-field molecular dynamics simulations. The resulting geometric and energetic changes allow interpretation of some experimentally observed effects of mutations, particularly with regard to the "hot spots" at residues 155 and 162. The replacement S155F enhances S enantiopreference through a steric relay involving Leu162. The double mutation S53P + L162G improves S enantioselectivity by creating a new binding pocket for the S enantiomer with an additional stabilizing hydrogen bond to His83. The simulations provide insight into remote and cooperative effects of mutations.
        
Title: Controlling the enantioselectivity of enzymes by directed evolution: practical and theoretical ramifications Reetz MT Ref: Proc Natl Acad Sci U S A, 101:5716, 2004 : PubMed
A fundamentally new approach to asymmetric catalysis in organic chemistry is described based on the in vitro evolution of enantioselective enzymes. It comprises the appropriate combination of gene mutagenesis and expression coupled with an efficient high-throughput screening system for evaluating enantioselectivity (enantiomeric excess assay). Several such cycles lead to a "Darwinistic" process, which is independent of any knowledge concerning the structure or the mechanism of the enzyme being evolved. The challenge is to choose the optimal mutagenesis methods to navigate efficiently in protein sequence space. As a first example, the combination of error-prone mutagenesis, saturation mutagenesis, and DNA-shuffling led to a dramatic enhancement of enantioselectivity of a lipase acting as a catalyst in the kinetic resolution of a chiral ester. Mutations at positions remote from the catalytically active center were identified, a surprising finding, which was explained on the basis of a novel relay mechanism. The scope and limitations of the method are discussed, including the prospect of directed evolution of stereoselective hybrid catalysts composed of robust protein hosts in which transition metal centers have been implanted.
        
Title: Characterization of the lipA gene encoding the major lipase from Pseudomonas aeruginosa strain IGB83 Martinez A, Soberon-Chavez G Ref: Applied Microbiology & Biotechnology, 56:731, 2001 : PubMed
The lipases produced by Pseudomonas have a wide range of potential biotechnological applications. Pseudomonas aeruginosa IGB83 was isolated as a highly lipolytic strain which produced a thermotolerant and alkaline lipase. In the present work, we have characterized the P. aeruginosa IGB83 gene (lipA) encoding this enzyme. We describe the construction of a lipA mutant and report on the effect of two carbon sources on lipase expression.
        
Title: Crystal structure of pseudomonas aeruginosa lipase in the open conformation. The prototype for family I.1 of bacterial lipases Nardini M, Lang DA, Liebeton K, Jaeger KE, Dijkstra BW Ref: Journal of Biological Chemistry, 275:31219, 2000 : PubMed
The x-ray structure of the lipase from Pseudomonas aeruginosa PAO1 has been determined at 2.54 A resolution. It is the first structure of a member of homology family I.1 of bacterial lipases. The structure shows a variant of the alpha/beta hydrolase fold, with Ser(82), Asp(229), and His(251) as the catalytic triad residues. Compared with the "canonical" alpha/beta hydrolase fold, the first two beta-strands and one alpha-helix (alphaE) are not present. The absence of helix alphaE allows the formation of a stabilizing intramolecular disulfide bridge. The loop containing His(251) is stabilized by an octahedrally coordinated calcium ion. On top of the active site a lid subdomain is in an open conformation, making the catalytic cleft accessible from the solvent region. A triacylglycerol analogue is covalently bound to Ser(82) in the active site, demonstrating the position of the oxyanion hole and of the three pockets that accommodate the sn-1, sn-2, and sn-3 fatty acid chains. The inhibited enzyme can be thought to mimic the structure of the tetrahedral intermediate that occurs during the acylation step of the reaction. Analysis of the binding mode of the inhibitor suggests that the size of the acyl pocket and the size and interactions of the sn-2 binding pocket are the predominant determinants of the regio- and enantio-preference of the enzyme.
        
Title: Substitutions of Ser for Asn-163 and Pro for Leu-264 are important for stabilization of lipase from Pseudomonas aeruginosa Shinkai A, Hirano A, Aisaka K Ref: J Biochem, 120:915, 1996 : PubMed
The lipase gene from Pseudomonas aeruginosa was randomly mutated by error-prone PCR to obtain thermostable mutants, followed by screening for thermostable mutant lipases. Out of about 2,600 transformants, four thermostable clones were obtained. Their nucleotide sequences showed that they had two or three amino acid substitutions. Analysis of the thermal stabilization of these mutant lipases indicated that Asn-163 to Ser and Leu-264 to Pro mutations were essential for the increased stability of the lipase. We expressed a mutant lipase (StLipA-5) having only the Asn-163 to Ser mutation and another (StLipA-6) having only the Leu-264 to Pro mutation in P. aeruginosa PAO1161, purified them, and then confirmed that the temperature which causes a 50% decrease in the activity of the non-treated enzyme on treatment for 30 min was increased by 1.5 and 3 degrees C, respectively, compared to the wild-type enzyme. However, the thermal stability of the mutant lipase (StLipA-7) having both mutations was increased only by 2.5 degrees C. These mutant lipases were stabilized through a decrease in activation entropy. Kinetic studies showed that the Kcat/K(m) values of StLipA-5, StLipA-6, and StLipA-7 were decreased by 14.4, 52.9, and 26.0%, respectively. Interestingly, the pH-stabilities of StLipA-6 and StLipA-7 were also increased, especially at alkaline pH. Based on these results, the tertiary structure and mechanism of stabilization of the lipase were discussed.
An extracellular lipase secreted by Pseudomonas aeruginosa TE3285 was purified. A genomic library of this strain was constructed in lambda EMBL3, and a DNA fragment 2.7 kb long containing the lipase gene, lipA, was isolated with an oligonucleotide probe synthesized on the basis of the partial amino acid sequence of a purified preparation of the enzyme. Nucleotide sequence analysis showed an open reading frame of 933 bases, and the deduced amino acid sequence agreed well with the molecular mass and partial amino acid sequences of mature lipase. The results of alignment of the amino acid sequences of five lipases from Pseudomonas species considered together with the published crystal structure studied with human pancreatic lipase showed that Ser82, His251, and Asp209 were catalytic residues and that a surface loop from residues 172 to 204 was responsible for the substrate specificity. About 50 bases downstream of lipA, there was another gene, lipB. The sequence of lipB was highly homologous to that of putative modulators of the production of active lipases in other Pseudomonas species. Expression plasmids encoding lipA followed by the complete or incomplete lipB gene downstream of the lac promoter of pUC18 were constructed. lipA was expressed in Escherichia coli 1100 only in the presence of the complete lipB gene.
The structural gene (lipA) coding for the extracellular lipase of Pseudomonas aeruginosa PAO1 has been cloned on plasmid pSW118. Nucleotide sequence analysis revealed a gene of 936 bp. lipA codes for a proenzyme of 311 amino acids including a leader sequence of 26 amino acids. The mature protein was predicted to have a M(r) of 30134, an isoelectric point of 5.6, and a consensus sequence (IGHSHGG) typical of lipases. Furthermore it is highly homologous (greater than 60%) to other lipases from various pseudomonads. The lipA gene failed to hybridize detectably with genomic DNA from other Pseudomonas species except P. alcaligenes, even under relaxed stringency. Located 220 bp downstream of the lipA gene, is an open reading frame (ORF2, lipH) which encodes a hydrophilic protein (283 amino acids; M(r) 33587) that shows some homology to the limA gene product of P. cepacia. In complementation tests of lipase-defective mutants, lipH was shown to be necessary for expression of active extracellular lipase in P. aeruginosa PAO1.
        
Title: Extracellular lipase of Pseudomonas aeruginosa: biochemical characterization and effect on human neutrophil and monocyte function in vitro Jaeger KE, Kharazmi A, Hoiby N Ref: Microb Pathog, 10:173, 1991 : PubMed
Lipase was isolated from P. aeruginosa by ultrafiltration of sterile-filtered culture supernatant. Gel filtration on Sepharose 4B yielded a broad peak corresponding to a molecular mass range of 100 to 1000 kDa. Electron microscopy of a negatively stained lipase preparation after Sepharose 4B chromatography revealed spherical particles with diameters ranging from 5 to 20 nm. Biochemical characterization and SDS polyacrylamide gel electrophoresis suggested that these particles consisted of protein and carbohydrate including lipopolysaccharide with the major enzyme activity being lipase. Various concentrations of this lipase preparation were preincubated with human peripheral blood neutrophils and monocytes. The chemotaxis and chemiluminescence of these cells were then determined. It was shown that lipase inhibited the monocyte chemotaxis and chemiluminescence, whereas it had no or very little effect on neutrophils. The inhibitory effect was concentration dependent and was abolished by heat treatment of the enzyme at 100 degrees C. Since monocytes are one of the important cells of the host defence system the inhibition of the function of these cells may contribute to the pathogenesis of infections caused by P. aeruginosa.