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 78 毫秒
1.
2.
3.
4.
5.
Specific therapy is not available for hantavirus cardiopulmonary syndrome caused by Andes virus (ANDV). Peptides capable of blocking ANDV infection in vitro were identified using antibodies against ANDV surface glycoproteins Gn and Gc to competitively elute a cyclic nonapeptide-bearing phage display library from purified ANDV particles. Phage was examined for ANDV infection inhibition in vitro, and nonapeptides were synthesized based on the most-potent phage sequences. Three peptides showed levels of viral inhibition which were significantly increased by combination treatment with anti-Gn- and anti-Gc-targeting peptides. These peptides will be valuable tools for further development of both peptide and nonpeptide therapeutic agents.Andes virus (ANDV), an NIAID category A agent linked to hantavirus cardiopulmonary syndrome (HCPS), belongs to the family Bunyaviridae and the genus Hantavirus and is carried by Oligoryzomys longicaudatus rodents (11). HCPS is characterized by pulmonary edema caused by capillary leak, with death often resulting from cardiogenic shock (9, 16). ANDV HCPS has a case fatality rate approaching 40%, and ANDV is the only hantavirus demonstrated to be capable of direct person-to-person transmission (15, 21). There is currently no specific therapy available for treatment of ANDV infection and HCPS.Peptide ligands that target a specific protein surface can have broad applications as therapeutics by blocking specific protein-protein interactions, such as preventing viral engagement of host cell receptors and thus preventing infection. Phage display libraries provide a powerful and inexpensive tool to identify such peptides. Here, we used selection of a cyclic nonapeptide-bearing phage library to identify peptides capable of binding the transmembrane surface glycoproteins of ANDV, Gn and Gc, and blocking infection in vitro.To identify peptide sequences capable of recognizing ANDV, we panned a cysteine-constrained cyclic nonapeptide-bearing phage display library (New England Biolabs) against density gradient-purified, UV-treated ANDV strain CHI-7913 (a gift from Hector Galeno, Santiago, Chile) (17, 18). To increase the specificity of the peptides identified, we eluted phage by using monoclonal antibodies (Austral Biologicals) prepared against recombinant fragments of ANDV Gn (residues 1 to 353) or Gc (residues 182 to 491) glycoproteins (antibodies 6B9/F5 and 6C5/D12, respectively). Peptide sequences were determined for phage from iterative rounds of panning, and the ability of phage to inhibit ANDV infection of Vero E6 cells was determined by immunofluorescent assay (IFA) (7). Primary IFA detection antibodies were rabbit polyclonal anti-Sin Nombre hantavirus (SNV) nucleoprotein (N) antibodies which exhibit potent cross-reactivity against other hantavirus N antigens (3). ReoPro, a commercially available Fab fragment which partially blocks infection of hantaviruses in vitro by binding the entry receptor integrin β3 (5), was used as a positive control (80 μg/ml) along with the original antibody used for phage elution (5 μg/ml). As the maximum effectiveness of ReoPro in inhibiting hantavirus entry approaches 80%, we set this as a threshold for maximal expected efficacy for normalization. The most-potent phage identified by elution with the anti-Gn antibody 6B9/F5 bore the peptide CPSNVNNIC and inhibited hantavirus entry by greater than 60% (61%) (Table (Table1).1). From phage eluted with the anti-Gc antibody 6C5/D12, those bearing peptides CPMSQNPTC and CPKLHPGGC also inhibited entry by greater than 60% (66% and 72%, respectively).

TABLE 1.

Peptide-bearing phage eluted from ANDV
Phage% Inhibition (SD)aP valueb
Phage bearing the following peptides eluted with anti-Gn antibody 6B9/F5
    Group 1 (<30% inhibition)
        CDQRTTRLC8.45 (15.34)0.0002
        CPHDPNHPC9.94 (7.72)0.333
        CQSQTRNHC11.76 (13.25)0.0001
        CLQDMRQFC13.26 (9.92)0.0014
        CLPTDPIQC15.70 (14.05)0.0005
        CPDHPFLRC16.65 (15.22)0.8523
        CSTRAENQC17.56 (16.50)0.0004
        CPSHLDAFC18.98 (20.06)0.0017
        CKTGHMRIC20.84 (7.47)0.0563
        CVRTPTHHC20.89 (27.07)0.1483
        CSGVINTTC21.57 (19.61)0.0643
        CPLASTRTC21.65 (5.98)0.004
        CSQFPPRLC22.19 (8.26)0.0004
        CLLNKQNAC22.34 (7.78)0.001
        CKFPLNAAC22.89 (6.15)0.0001
        CSLTPHRSC23.63 (16.74)0.0563
        CKPWPMYSC23.71 (6.68)0.0643
        CLQHDALNC24.01 (7.60)1
        CNANKPKMC24.67 (11.67)0.0004
        CPKHVLKVC25.30 (28.36)0.0003
        CTPDKKSFC26.91 (11.15)0.399
        CHGKAALAC27.22 (32.53)0.005
        CNLMGNPHC28.08 (21.35)0.0011
        CLKNWFQPC28.64 (18.49)0.0016
        CKEYGRQMC28.76 (29.33)0.0362
        CQPSDPHLC29.44 (31.22)0.0183
        CSHLPPNRC29.70 (17.37)0.0061
    Group 2 (30-59% inhibition)
        CSPLLRTVC33.05 (20.26)0.0023
        CHKGHTWNC34.17 (12.50)0.0795
        CINASHAHC35.62 (13.03)0.3193
        CWPPSSRTC36.75 (26.95)0.0006
        CPSSPFNHC37.78 (7.11)0.0001
        CEHLSHAAC38.47 (7.60)0.0115
        CQDRKTSQC38.74 (9.12)0.1802
        CTDVYRPTC38.90 (25.03)0.006
        CGEKSAQLC39.11 (27.52)0.0013
        CSAAERLNC40.13 (6.33)0.0033
        CFRTLEHLC42.07 (5.01)0.0608
        CEKLHTASC43.60 (27.92)0.1684
        CSLHSHKGC45.11 (49.81)0.0864
        CNSHSPVHC45.40 (28.80)0.0115
        CMQSAAAHC48.88 (44.40)0.5794
        CPAASHPRC51.84 (17.09)0.1935
        CKSLGSSQC53.90 (13.34)0.0145
    Group 3 (60-79% inhibition)
        CPSNVNNIC61.11 (25.41)0.1245
Negative control0 (6.15)
6B9/F5 (5 μg/ml)26.77 (5.33)
ReoPro (80 μg/ml)79.86 (4.88)
Phage bearing the following peptides eluted with anti-Gc antibody 6C5/D12
    Group 1 (<30% inhibition)
        CHPGSSSRC1.01 (7.03)0.0557
        CSLSPLGRC10.56 (13.62)0.7895
        CTARYTQHC12.86 (3.83)0.3193
        CHGVYALHC12.91 (7.32)0.0003
        CLQHNEREC16.79 (13.72)0.0958
        CHPSTHRYC17.23 (14.53)0.0011
        CPGNWWSTC19.34(9.91)0.1483
        CGMLNWNRC19.48 (19.42)0.0777
        CPHTQFWQC20.44 (13.65)0.0008
        CTPTMHNHC20.92 (11.68)0.0001
        CDQVAGYSC21.79 (23.60)0.0063
        CIPMMTEFC24.33 (9.28)0.2999
        CERPYSRLC24.38 (9.09)0.0041
        CPSLHTREC25.06 (22.78)0.1202
        CSPLQIPYC26.30 (34.29)0.4673
        CTTMTRMTC (×2)29.27 (8.65)0.0001
    Group 2 (30-59% inhibition)
        CNKPFSLPC30.09 (5.59)0.4384
        CHNLESGTC31.63 (26.67)0.751
        CNSVPPYQC31.96 (6.51)0.0903
        CSDSWLPRC32.95 (28.54)0.259
        CSAPFTKSC33.40 (10.64)0.0052
        CEGLPNIDC35.63 (19.90)0.0853
        CTSTHTKTC36.28 (13.42)0.132
        CLSIHSSVC36.40 (16.44)0.8981
        CPWSTQYAC36.81 (32.81)0.5725
        CTGSNLPIC36.83 (31.64)0.0307
        CSLAPANTC39.73 (4.03)0.1664
        CGLKTNPAC39.75 (16.98)0.2084
        CRDTTPWWC40.08 (18.52)0.0004
        CHTNASPHC40.26 (4.77)0.5904
        CTSMAYHHC41.89 (8.61)0.259
        CSLSSPRIC42.13 (29.75)0.2463
        CVSLEHQNC45.54 (6.55)0.5065
        CRVTQTHTC46.55 (8.45)0.3676
        CPTTKSNVC49.28 (14.00)0.3898
        CSPGPHRVC49.50 (42.60)0.0115
        CKSTSNVYC51.20 (4.60)0.0611
        CTVGPTRSC57.30 (11.31)0.0176
    Group 3 (60-79% inhibition)
        CPMSQNPTC65.60 (13.49)0.014
        CPKLHPGGC71.88 (27.11)0.0059
Negative control0.26 (4.53)
6C5/D12 (5 μg/ml)22.62 (8.40)
ReoPro (80 μg/ml)80.02 (76.64)
Open in a separate windowaStandard deviations of four experiments are shown in parentheses. Peptide-bearing phage were added at 109 phage/μl.bP values for the pairwise amino acid alignment score of each peptide versus that of integrin β3 were determined using an unpaired Student''s t test. P values considered statistically significant are shown in bold.To determine whether the peptide sequences of any of the identified inhibitory phage showed homology to integrin β3, a known entry receptor for pathogenic hantaviruses (6, 7), we used the Gap program to perform a pairwise amino acid alignment of each peptide versus the extracellular portion of integrin β3 and determined P values for the alignments. Of 45 phage eluted with the anti-Gn antibody, 6B9/F5, 27 of the peptide sequences showed homology to integrin β3 (P < 0.05), and 9 were highly significant (P ≤ 0.0005) (Fig. (Fig.1A).1A). Of the latter, CKFPLNAAC and CSQFPPRLC map to the hybrid domain (Fig. (Fig.1B),1B), which is proximal to the plexin-semaphorin-integrin domain (PSI) containing residue D39, shown to be critical for viral entry in vitro (19). Five sequences (CPSSPFNH, CPKHVLKVC, CNANKPKMC, CQSQTRNHC, and CDQRTTRLC) map to the I-like (or βA) domain near the binding site of ReoPro (2). Finally, CLPTDPIQC maps to the epidermal growth factor 4 (EGF-4) domain, and CSTRAENQC aligns to a portion of β3 untraceable in the crystal structure, specifically the linker region between the hybrid domain and EGF-1. Although this represents a disordered portion of the protein (22), the location of this loop proximal to the PSI domain is worth noting, due to the role of the PSI domain in facilitating viral entry (19). Therefore, 60% of phage eluted with the anti-Gn antibody showed some homology to integrin β3, and those with highly significant P values predominantly mapped to or proximal to regions of known interest in viral entry.Open in a separate windowFIG. 1.Inhibitory peptides identified through phage panning against ANDV show homology to integrin β3. (A) Alignment of phage peptide sequences with P values for integrin β3 pairwise alignment of less than 0.05. Residues comprising the signal peptide, transmembrane, and cytoplasmic domains, which were not included during pairwise alignment, are underlined. Residues 461 to 548, which are missing in the crystal structure, are italicized. Residues involved in the ReoPro binding site are highlighted in green (2). Residue D39 of the PSI domain is highlighted in yellow (19). Peptides are shown above the sequence of integrin β3, with antibody 6C5/D12-eluted sequences shown in blue text and sequences eluted with antibody 6B9/F5 shown in red. Peptide sequences with alignment P values of ≤0.0005 are highlighted in yellow. Percent inhibition of the peptide-bearing phage is shown in parentheses. (B) View of integrin αvβ3 (PDB ID 1U8C [23]). αv is shown in blue ribbon diagram, and β3 is shown in salmon-colored surface representation, with specific domains circled. Residues corresponding to the ReoPro binding site are shown in green, as in panel A, and D39 is shown in yellow. Regions corresponding to 6C5/D12-eluted peptides with P values of ≤0.0005 for alignment with integrin β3 (highlighted in panel A) are shown in blue, and those corresponding to 6B9/F5-eluted peptides with P values of ≤0.0005 for alignment with integrin β3 are shown in red. Alignment of peptide PLASTRT (P value of 0.0040) adjacent to D39 of the PSI domain is shown in magenta. Graphics were prepared using Pymol (DeLano Scientific LLC, San Carlos, CA).Of the 41 peptide-bearing phage eluted with the anti-Gc antibody 6C5/D12, 14 showed sequence homology to integrin β3 (P < 0.05), 4 of which had P values of ≤0.0005 (Fig. (Fig.1A).1A). Of the latter, sequence CTTMTRMTC mapped to the base of the I-like domain (Fig. (Fig.1B),1B), while CHGVYALHC and CRDTTPWWC mapped to the EGF-3 domain. Finally, sequence CTPTMHNHC mapped to the linker region untraceable in the crystal structure. Therefore, in contrast to peptide sequences identified by competition with the anti-Gn antibody, sequences identified by competition with the anti-Gc antibody 6C5/D12 appear to be mostly unrelated to integrin β3.As a low level of pathogenic hantavirus infection can be seen in cells lacking integrin β3, such as CHO cells (19), we asked if any of the identified peptide sequences could represent a previously unidentified receptor. We used the Basic Local Alignment Search Tool to search a current database of human protein sequences for potential alternate receptors represented by these peptides. However, none of the alignments identified proteins that are expressed at the cell surface, eliminating them as potential candidates for alternate viral entry receptors. This suggests that the majority of the peptides identified here likely represent novel sequences for binding ANDV surface glycoproteins.To determine whether synthetic peptides would also block infection, we synthesized cyclic peptides based on the 10 most-potent peptide-bearing phage. These peptides, in the context of phage presentation, showed levels of inhibition ranging from 44 to 72% (Table (Table2).2). When tested by IFA at 1 mM, four of the synthetic peptides showed inhibition levels significantly lower than those of the same peptide presented in the context of phage. This is not surprising, as steric factors due to the size of the phage and the multivalent presentation of peptide in the context of phage may both contribute to infection inhibition (8). However, there was no significant difference in inhibition by synthetic peptide versus peptide-bearing phage for six of the sequences, implying that inhibition in the context of phage was due solely to the nature of the peptide itself and not to steric factors or valency considerations contributed by the phage, which contrasts with our previous results, determined by using phage directed against αvβ3 integrin (10).

TABLE 2.

Synthetic cyclic peptides inhibit ANDV infection
TargetSample% Inhibition bya:
Peptide-bearing phageSynthetic peptide
GnCMQSAAAHC48.88 (44.40)59.66 (11.17)
GcCTVGPTRSC57.30 (11.31)46.47 (7.61)
GnCPSNVNNIC61.11 (25.41)44.14 (10.74)
GnCEKLHTASC43.60 (27.92)34.87 (9.26)
GcCPKLHPGGC71.88 (27.11)30.95 (7.73)b
GnCSLHSHKGC45.11 (49.81)29.79 (9.34)
GcCPMSQNPTC65.60 (13.49)18.19 (8.55)b
GnCKSLGSSQC53.90 (13.34)18.10 (7.55)b
GnCNSHSPVHC45.40 (28.80)15.52 (10.48)
GnCPAASHPRC51.84 (17.09)0 (10.72)b
Integrin β3ReoPro80.10 (7.72)
Gn6B9/F5 antibody42.72 (6.75)
Gc6C5/D12 antibody31.04 (7.81)
Open in a separate windowaStandard deviations of the results of at least four experiments are shown in parentheses.bMean percent inhibition between phage and synthetic peptide differs significantly (P < 0.05).The three most-potent synthetic peptides were examined for their ability to inhibit ANDV entry in a dose-dependent manner. The concentration of each peptide that produces 50% of its maximum potential inhibitory effect was determined. As shown in Fig. Fig.2A,2A, the 50% inhibitory concentration for each of the peptides was in the range of 10 μM, which from our experience is a reasonable potency for a lead compound to take forward for optimization.Open in a separate windowFIG. 2.Activities of synthetic peptides in inhibition of ANDV infection in vitro. (A) Peptides were examined for their ability to block ANDV infection of Vero E6 cells in a dose-dependent manner by IFA. (B) Peptides were tested in parallel for the ability to block infection of Vero E6 cells by ANDV, SNV, HTNV, and PHV. (C) Peptides were tested, singly or in combination, for the ability to block ANDV infection of Vero E6 cells. For all experiments, controls included media, ReoPro at 80 μg/ml, and monoclonal antibodies 6C5/D12 and 6B9/F5 at 5 μg/ml. All peptides were used at 1 mM. Data points represent n = 2 to 6, with error bars showing the standard errors of the means. Statistical analyses were performed on replicate samples using an unpaired Student''s t test.In order to determine the specificity of the three most-potent synthetic cyclic peptides in blocking ANDV, we examined them for inhibition of ANDV infection versus two other pathogenic hantaviruses, SNV and Hantaan virus (HTNV), or the nonpathogenic hantavirus Prospect Hill virus (PHV). As shown in Fig. Fig.2B,2B, ReoPro, which binds integrin β3, showed inhibition of infection by each of the pathogenic hantavirus strains, known to enter cells via β3, but not the nonpathogenic PHV, which enters via integrin β1 (6, 7). In contrast, peptides selected for the ability to bind ANDV were highly specific inhibitors of ANDV versus SNV, HTNV, or PHV. The specificities of peptides eluted by the anti-Gn monoclonal antibody are not surprising, as they are likely due to global differences in the Gn amino acid sequence. Specifically, sequence homologies between ANDV and SNV, HTNV, and PHV are 61%, 36%, and 51%, respectively, for the region corresponding to the immunogen for antibody 6B9/F5. Although homology between the immunogen for antibody 6C5/D12 and the corresponding Gc region of these viruses is somewhat higher (82% with SNV, 63% with HTNV, and 71% with PHV), the possibility that the monoclonal antibody used here recognizes a three-dimensional epitope lends itself to the high specificity of the peptides.The current model for cellular infection by hantaviruses (14) is as follows. Viral binding of the host cell surface target integrin is followed by receptor-mediated endocytosis and endosome acidification. Lowered pH induces conformational changes in Gn and/or Gc, which facilitate membrane fusion and viral release into the cytosol. As there is currently little information available about whether one glycoprotein is dominant in mediating infection, and as neutralizing epitopes have been found on both Gn and Gc glycoproteins (1, 4, 12, 13, 20), we examined whether combining anti-Gn- and anti-Gc-targeted synthetic peptides would lead to an increased infection blockade compared to those for single treatments. As shown in Fig. Fig.2C,2C, the combination of anti-Gn and anti-Gc peptides CMQSAAAHC and CTVGPTRSC resulted in a significant increase in infection inhibition (P = 0.0207 for CMQSAAAHC, and P = 0.0308 for CTVGPTRSC) compared to that resulting from single treatments. Although the high specificity of the peptides for ANDV makes it unlikely that this combination treatment will lead to more cross-reactivity with other pathogenic hantaviruses, this can be determined only by additional testing. Regardless, these data suggest a unique role for each of these viral proteins in the infection process as well as the benefits of targeting multiple viral epitopes for preventing infection.To our knowledge, the peptides reported here are the first identified that directly target ANDV, and this work further illustrates the power of coupling phage display and selective elution techniques in the identification of novel peptide sequences capable of specific protein-protein interactions from a large, random pool of peptide sequences. These novel peptide inhibitors (R. S. Larson, P. R. Hall, H. Njus, and B. Hjelle, U.S. patent application 61/205,211) provide leads for the development of more-potent peptide or nonpeptide organics for therapeutic use against HCPS.  相似文献   

6.
7.
GTP cyclohydrolase I (GCYH-I) is an essential Zn2+-dependent enzyme that catalyzes the first step of the de novo folate biosynthetic pathway in bacteria and plants, the 7-deazapurine biosynthetic pathway in Bacteria and Archaea, and the biopterin pathway in mammals. We recently reported the discovery of a new prokaryotic-specific GCYH-I (GCYH-IB) that displays no sequence identity to the canonical enzyme and is present in ∼25% of bacteria, the majority of which lack the canonical GCYH-I (renamed GCYH-IA). Genomic and genetic analyses indicate that in those organisms possessing both enzymes, e.g., Bacillus subtilis, GCYH-IA and -IB are functionally redundant, but differentially expressed. Whereas GCYH-IA is constitutively expressed, GCYH-IB is expressed only under Zn2+-limiting conditions. These observations are consistent with the hypothesis that GCYH-IB functions to allow folate biosynthesis during Zn2+ starvation. Here, we present biochemical and structural data showing that bacterial GCYH-IB, like GCYH-IA, belongs to the tunneling-fold (T-fold) superfamily. However, the GCYH-IA and -IB enzymes exhibit significant differences in global structure and active-site architecture. While GCYH-IA is a unimodular, homodecameric, Zn2+-dependent enzyme, GCYH-IB is a bimodular, homotetrameric enzyme activated by a variety of divalent cations. The structure of GCYH-IB and the broad metal dependence exhibited by this enzyme further underscore the mechanistic plasticity that is emerging for the T-fold superfamily. Notably, while humans possess the canonical GCYH-IA enzyme, many clinically important human pathogens possess only the GCYH-IB enzyme, suggesting that this enzyme is a potential new molecular target for antibacterial development.The Zn2+-dependent enzyme GTP cyclohydrolase I (GCYH-I; EC 3.5.4.16) is the first enzyme of the de novo tetrahydrofolate (THF) biosynthesis pathway (Fig. (Fig.1)1) (38). THF is an essential cofactor in one-carbon transfer reactions in the synthesis of purines, thymidylate, pantothenate, glycine, serine, and methionine in all kingdoms of life (38), and formylmethionyl-tRNA in bacteria (7). Recently, it has also been shown that GCYH-I is required for the biosynthesis of the 7-deazaguanosine-modified tRNA nucleosides queuosine and archaeosine produced in Bacteria and Archaea (44), respectively, as well as the 7-deazaadenosine metabolites produced in some Streptomyces species (33). GCYH-I is encoded in Escherichia coli by the folE gene (28) and catalyzes the conversion of GTP to 7,8-dihydroneopterin triphosphate (55), a complex reaction that begins with hydrolytic opening of the purine ring at C-8 of GTP to generate an N-formyl intermediate, followed by deformylation and subsequent rearrangement and cyclization of the ribosyl moiety to generate the pterin ring in THF (Fig. (Fig.1).1). Notably, the enzyme is dependent on an essential active-site Zn2+ that serves to activate a water molecule for nucleophilic attack at C-8 in the first step of the reaction (2).Open in a separate windowFIG. 1.Reaction catalyzed by GCYH-I, and metabolic fate of 7,8-dihydroneopterin triphosphate.A homologous GCYH-I is found in mammals and other higher eukaryotes, where it catalyzes the first step of the biopterin (BH4) pathway (Fig. (Fig.1),1), an essential cofactor in the biosynthesis of tyrosine and neurotransmitters, such as serotonin and l-3,4-dihydroxyphenylalanine (3, 52). Recently, a distinct class of GCYH-I enzymes, GCYH-IB (encoded by the folE2 gene), was discovered in microbes (26% of sequenced Bacteria and most Archaea) (12), including several clinically important human pathogens, e.g., Neisseria and Staphylococcus species. Notably, GCYH-IB is absent in eukaryotes.The distribution of folE (gene product renamed GCYH-IA) and folE2 (GCYH-IB) in bacteria is diverse (12). The majority of organisms possess either a folE (65%; e.g., Escherichia coli) or a folE2 (14%; e.g., Neisseria gonorrhoeae) gene. A significant number (12%; e.g., B. subtilis) possess both genes (a subset of 50 bacterial species is shown in Table Table1),1), and 9% lack both genes, although members of the latter group are mainly intracellular or symbiotic bacteria that rely on external sources of folate. The majority of Archaea possess only a folE2 gene, and the encoded GCYH-IB appears to be necessary only for the biosynthesis of the modified tRNA nucleoside archaeosine (44) except in the few halophilic Archaea that are known to synthesize folates, such as Haloferax volcanii, where GCYH-IB is involved in both archaeosine and folate formation (13, 44).

TABLE 1.

Distribution and candidate Zur-dependent regulation of alternative GCYH-I genes in bacteriaa
OrganismcPresence of:
folEfolE2
Enterobacteria
    Escherichia coli+
    Salmonella typhimurium+
    Yersinia pestis+
    Klebsiella pneumoniaeb++a
    Serratia marcescens++a
    Erwinia carotovora+
    Photorhabdus luminescens+
    Proteus mirabilis+
Gammaproteobacteria
    Vibrio cholerae+
    Acinetobacter sp. strain ADP1++a
    Pseudomonas aeruginosa++a
    Pseudomonas entomophila L48++a
    Pseudomonas fluorescens Pf-5++a
    Pseudomonas syringae++a
    Pseudomonas putida++a
    Hahella chejuensis KCTC 2396++a
    Chromohalobacter salexigens DSM 3043++a
    Methylococcus capsulatus++a
    Xanthomonas axonopodis++a
    Xanthomonas campestris++a
    Xylella fastidiosa++a
    Idiomarina loihiensis+
    Colwellia psychrerythraea++
    Pseudoalteromonas atlantica T6c++a
    Pseudoalteromonas haloplanktis TAC125++
    Alteromonas macleodi+
    Nitrosococcus oceani++
    Legionella pneumophila+
    Francisella tularensis+
Betaproteobacteria
    Chromobacterium violaceum+
    Neisseria gonorrhoeae+
    Burkholderia cepacia R18194++
    Burkholderia cenocepacia AU 1054++
    Burkholderia xenovorans+
    Burkholderia mallei+
    Bordetella pertussis+
    Ralstonia eutropha JMP134+
    Ralstonia metallidurans++
    Ralstonia solanacearum+
    Methylobacillus flagellatus+
    Nitrosomonas europaea+
    Azoarcus sp.++
Bacilli/Clostridia
    Bacillus subtilisd++
    Bacillus licheniformis++
    Bacillus cereus+
    Bacillus halodurans++
    Bacillus clausii+
    Geobacillus kaustophilus+
    Oceanobacillus iheyensis+
    Staphylococcus aureus+
Open in a separate windowaGenes that are preceded by candidate Zur binding sites.bZur-regulated cluster is on the virulence plasmid pLVPK.cExamples of organisms with no folE genes are in boldface type.dZn-dependent regulation of B. subtilis folE2 by Zur was experimentally verified (17).Expression of the Bacillus subtilis folE2 gene, yciA, is controlled by the Zn2+-dependent Zur repressor and is upregulated under Zn2+-limiting conditions (17). This led us to propose that the GCYH-IB family utilizes a metal other than Zn2+ to allow growth in Zn2+-limiting environments, a hypothesis strengthened by the observation that an archaeal ortholog from Methanocaldococcus jannaschii has recently been shown to be Fe2+ dependent (22). To test this hypothesis, we investigated the physiological role of GCYH-IB in B. subtilis, an organism that contains both isozymes, as well as the metal dependence of B. subtilis GCYH-IB in vitro. To gain a structural understanding of the metal dependence of GCYH-IB, we determined high-resolution crystal structures of Zn2+- and Mn2+-bound forms of the N. gonorrhoeae ortholog. Notably, although the GCYH-IA and -IB enzymes belong to the tunneling-fold (T-fold) superfamily, there are significant differences in their global and active-site architecture. These studies shed light on the physiological significance of the alternative folate biosynthesis isozymes in bacteria exposed to various metal environments, and offer a structural understanding of the differential metal dependence of GCYH-IA and -IB.  相似文献   

8.
9.
Twelve cluster groups of Escherichia coli O26 isolates found in three cattle farms were monitored in space and time. Cluster analysis suggests that only some O26:H11 strains had the potential for long-term persistence in hosts and farms. As judged by their virulence markers, bovine enterohemorrhagic O26:H11 isolates may represent a considerable risk for human infection.Shiga toxin (Stx)-producing Escherichia coli (STEC) strains comprise a group of zoonotic enteric pathogens (42). In humans, infections with some STEC serotypes result in hemorrhagic or nonhemorrhagic diarrhea, which can be complicated by hemolytic-uremic syndrome (HUS) (49). These STEC strains are also designated “enterohemorrhagic E. coli” (EHEC). Consequently, EHEC strains represent a subgroup of STEC with a high pathogenic potential for humans. Strains of the E. coli serogroup O26 were originally classified as enteropathogenic E. coli due to their association with outbreaks of infantile diarrhea in the 1940s. In 1977, Konowalchuk et al. (37) recognized that these bacteria produced Stx, and 10 years later, the Stx-producing E. coli O26:H11/H− strains were classified as EHEC. EHEC O26 strains constitute the most common non-O157 EHEC group associated with diarrhea and HUS in Europe (12, 21, 23, 24, 26, 27, 55, 60). Reports on an association between EHEC O26 and HUS or diarrhea from North America including the United States (15, 30, 33), South America (51, 57), Australia (22), and Asia (31, 32) provide further evidence for the worldwide spread of these organisms. Studies in Germany and Austria (26, 27) on sporadic HUS cases between 1996 and 2003 found that EHEC O26 accounted for 14% of all EHEC strains and for ∼40% of non-O157 EHEC strains obtained from these patients. A proportion of 11% EHEC O26 strains was detected in a case-control study in Germany (59) between 2001 and 2003. In the age group <3 years, the number of EHEC O26 cases was nearly equal to that of EHEC O157 cases, although the incidence of EHEC O26-associated disease is probably underestimated because of diagnostic limitations in comparison to the diagnosis of O157:H7/H− (18, 34). Moreover, EHEC O26 has spread globally (35). Beutin (6) described EHEC O26:H11/H−, among O103:H2, O111:H, O145:H28/H−, and O157:H7/H−, as the well-known pathogenic “gang of five,” and Bettelheim (5) warned that we ignore the non-O157 STEC strains at our peril.EHEC O26 strains produce Stx1, Stx2, or both (15, 63). Moreover, these strains contain the intimin-encoding eae gene (11, 63), a characteristic feature of EHEC (44). In addition, EHEC strains possess other markers associated with virulence, such as a large plasmid that carries further potential virulence genes, e.g., genes coding for EHEC hemolysin (EHEC-hlyA), a catalase-peroxidase (katP), and an extracellular serine protease (espP) (17, 52). The efa1 (E. coli factor for adherence 1) gene was identified as an intestinal colonization factor in EHEC (43). EHEC O26 represents a highly dynamic group of organisms that rapidly generate new pathogenic clones (7, 8, 63).Ruminants, especially cattle, are considered the primary reservoir for human infections with EHEC. Therefore, the aim of this study was the molecular characterization of bovine E. coli field isolates of serogroup O26 using a panel of typical virulence markers. The epidemiological situation in the beef herds from which the isolates were obtained and the spatial and temporal behavior of the clonal distribution of E. coli serogroup O26 were analyzed during the observation period. The potential risk of the isolates inducing disease in humans was assessed.In our study, 56 bovine E. coli O26:H11 isolates and one bovine O26:H32 isolate were analyzed for EHEC virulence-associated factors. The isolates had been obtained from three different beef farms during a long-term study. They were detected in eight different cattle in farm A over a period of 15 months (detected on 10 sampling days), in 3 different animals in farm C over a period of 8 months (detected on 3 sampling days), and in one cow on one sampling day in farm D (Table (Table1)1) (28).

TABLE 1.

Typing of E. coli O26 isolates
Sampling day, source, and isolateSerotypeVirulence profile by:
fliC PCR-RFLPstx1 genestx2 geneStx1 (toxin)Stx2 (toxin)Subtype(s)
efa1 genebEHEC-hlyA genekatP geneespP genePlasmid size(s) in kbCluster
stx1/stx2eaetirespAespB
Day 15
    Animal 6 (farm A)
        WH-01/06/002-1O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/06/002-2O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/06/002-3O26:H11H11++stx1ββββ+/++++110, 127
    Animal 8 (farm A)
        WH-01/08/002-2O26:H11H11++stx1ββββ+/++++110, 127
    Animal 26 (farm A)
        WH-01/26/001-2O26:H11H11++stx1ββββ+/++++130, 127
        WH-01/26/001-5O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/26/001-6O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/26/001-7O26:H11H11++stx1ββββ+/−+++110, 127
Day 29
    Animal 2 (farm A)
        WH-01/02/003-1O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-2O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-5O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-6O26:H11H11++stx1ββββ+/+++110, 126
        WH-01/02/003-7O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-8O26:H11H11++stx1ββββ−/++++110, 126
        WH-01/02/003-9O26:H11H11++stx1ββββ+/++++1106
        WH-01/02/003-10O26:H11H11++stx1ββββ+/++++1106
    Animal 26 (farm A)
        WH-01/26/002-2O26:H11H11++stx1ββββ+/++++130, 125
        WH-01/26/002-5O26:H11H11++stx1ββββ+/++++130, 125
        WH-01/26/002-8O26:H11H11++stx1ββββ+/++++130, 125
        WH-01/26/002-9O26:H11H11++stx1ββββ+/++110, 125
        WH-01/26/002-10O26:H11H11++stx1ββββ+/++++130, 125
Day 64
    Animal 20 (farm A)
        WH-01/20/005-3O26:H11H11++stx1ββββ+/+130, 2.52
Day 78
    Animal 29 (farm A)
        WH-01/29/002-1O26:H11H11++stx1ββββ+/−+130, 12, 2.54
        WH-01/29/002-2O26:H11H11++stx1ββββ+/++++130, 12, 2.54
        WH-01/29/002-3O26:H11H11++stx1ββββ+/++++130, 12, 2.54
        WH-01/29/002-4O26:H11H11++stx1ββββ+/++++130, 12, 2.54
        WH-01/29/002-5O26:H11H11++stx1ββββ+/++130, 12, 2.54
Day 106
    Animal 27 (farm A)
        WH-01/27/005-2O26:H11H11++stx1ββββ+/−+++145, 110, 123
        WH-01/27/005-5O26:H11H11++stx1ββββ+/++++130, 12, 2.55
        WH-01/27/005-6O26:H11H11++stx1ββββ+/+130, 12, 2.55
Day 113
    Animal 7 (farm C)
        WH-04/07/001-2O26:H11H11++++stx1/stx2ββββ+/+++55, 35, 2.511
        WH-04/07/001-4O26:H11H11++++stx1/stx2ββββ+/++++5512
        WH-04/07/001-6O26:H11H11++++stx1/stx2ββββ+/++++5512
Day 170
    Animal 22 (farm C)
        WH-04/22/001-1O26:H11H11++stx1ββββ+/++++110, 12, 6.312
        WH-04/22/001-4O26:H11H11++stx1ββββ+/++++110, 12, 6.312
        WH-04/22/001-5O26:H11H11++stx1ββββ+/++++110, 12, 6.312
Day 176
    Animal 14 (farm D)
        WH-03/14/004-8O26:H11H11++stx1ββββ+/+++11010
Day 218
    Animal 27 (farm A)
        WH-01/27/009-1O26:H11H11++++stx1/stx2ββββ+/++++110, 129
        WH-01/27/009-2O26:H11H11++++stx1/stx2ββββ+/++++110, 129
        WH-01/27/009-3O26:H11H11++++stx1/stx2ββββ+/++++110, 128
        WH-01/27/009-8O26:H11H11++++stx1/stx2ββββ+/++110, 128
        WH-01/27/009-9O26:H11H11++++stx1/stx2ββββ+/++++110, 129
Day 309
    Animal 29 (farm A)
        WH-01/29/010-1O26:H11H11++stx1ββββ+/++++110, 35, 124
        WH-01/29/010-2O26:H11H11++stx1ββββ+/++130, 55, 358
        WH-01/29/010-3O26:H11H11++stx1ββββ+/++++130, 35, 128
Day 365
    Animal 8 (farm C)
        WH-04/08/008-6O26:H11H11++stx1ββββ+/++++110, 5512
Day 379
    Animal 9 (farm A)
        WH-01/09/016-2O26:H32H32++stx1/stx2−/−145, 130, 1.81
    Animal 27 (farm A)
        WH-01/27/014-3O26:H11H11++stx1ββββ+/++++110, 129
        WH-01/27/014-4O26:H11H11++stx1ββββ+/++++110, 129
        WH-01/27/014-5O26:H11H11++stx1ββββ+/++++110, 128
Day 407
    Animal 29 (farm A)
        WH-01/29/013-4O26:H11H11++stx1ββββ+/++++110, 12, 2.58
        WH-01/29/013-7O26:H11H11++stx1ββββ+/++++110, 12, 2.58
Day 478
    Animal 27 (farm A)
        WH-01/27/017-1O26:H11H11++++stx1/stx2ββββ+/++++110, 128
        WH-01/27/017-5O26:H11H11++++stx1/stx2ββββ+/++++110, 128
        WH-01/27/017-6O26:H11H11++++stx1/stx2ββββ+/++++1108
        WH-01/27/017-7O26:H11H11++++stx1/stx2ββββ+/++++1108
        WH-01/27/017-10O26:H11H11+++stx1ββββ+/++++130, 12, 2.58
Open in a separate windowastx1/stx2, gene stx1 or stx2.befa1 was detected by two hybridizations (with lifA1-lifA2 and lifA3-lifA4 probes). +/+, complete gene; +/− or −/+, incomplete gene; −/−, efa1 negative.The serotyping of the O26 isolates was confirmed by the results of the fliC PCR-restriction fragment length polymorphism (RFLP) analysis performed according to Fields et al. (25), with slight modifications described by Zhang et al. (62). All O26:H11 isolates showed the H11 pattern described by Zhang et al. (62). In contrast, the O26:H32 isolate demonstrated a different fliC RFLP pattern that was identical to the H32 pattern described by the same authors. It has been demonstrated that EHEC O26:H11 strains belong to at least four different sequence types (STs) in the common clone complex 29 (39). In the multilocus sequence typing analysis for E. coli (61), the tested five EHEC O26:H11 isolates (WH-01/02/003-1, WH-01/20/005-3, WH-01/27/009-9, WH-03/14/004-8, and WH-04/22/001-1) of different farms and clusters were characterized as two sequence types (ST 21 and ST 396). The isolates from farms A and C belong to ST 21, the most frequent ST of EHEC O26:H11 isolates found in humans and animals (39), but the single isolate from farm D was characterized as ST 396.Typing and subtyping of genes (stx1 and/or stx2, eae, tir, espA, espB, EHEC-hlyA, katP, and espP) associated with EHEC were performed with LightCycler fluorescence PCR (48) and different block-cycler PCRs. To identify the subtypes of the stx2 genes and of the locus of enterocyte effacement-encoding genes eae, tir, espA, and espB, the PCR products were digested by different restriction endonucleases (19, 26, 46). The complete pattern of virulence markers was detected in most bovine isolates examined in our study. An stx1 gene was present in all O26 isolates. In addition, an stx2 gene was found in nine O26:H11 isolates in farm A and in three isolates of the same type in farm C, as well as in the O26:H32 isolate. Both Stx1 and Stx2 were closely related to families of Stx1 and Stx2 variants or alleles. EHEC isolates with stx2 genes are significantly more often associated with HUS and other severe disease manifestations than isolates with an stx1 gene, which are more frequently associated with uncomplicated diarrhea and healthy individuals (13). In contrast to STEC strains harboring stx2 gene variants, however, STEC strains of the stx2 genotype were statistically significantly associated with HUS (26). The stx2 genotype was found in all O26 isolates with an stx2 gene, while the GK3/GK4 amplification products after digestion with HaeIII and FokI restriction enzymes showed the typical pattern for this genotype described by Friedrich et al. (26). The nucleotide sequences of the A and B subunits of the stx2 gene of the selected bovine O26:H11 isolate WH-01/27/017-1 (GenBank accession no. EU700491) were identical to the stx2 genes of different sorbitol-fermenting EHEC O157:H− strains associated with human HUS cases and other EHEC infections in Germany (10) and 99.3% identical in their DNA sequences to the stx2 gene of the EHEC type strain EDL933, a typical O157:H7 isolate from an HUS patient. A characteristic stx1 genotype was present in all O26 isolates. The nucleotide sequences of the A and B subunits of the stx1 gene of the tested bovine O26:H11 isolate WH-01/27/017-1 (GenBank accession no. EU700490) were nearly identical to those of the stx1 genes of the EHEC O26:H11 reference type strains H19 and DEC10B, which had been associated with human disease outbreaks in Canada and Australia. Nucleotide exchanges typical for stx1c and stx1d subtypes as described by Kuczius et al. (38) were not found. All bovine O26:H11 strains produced an Stx1 with high cytotoxicity for Vero cells tested by Stx enzyme-linked immunosorbent assay and Vero cell neutralization assay (53). The Stx2 cytotoxicity for Vero cells was also very high in the O26:H11 isolates.Not only factors influencing the basic and inducible Stx production are important in STEC pathogenesis. It has been suggested that the eae and EHEC-hlyA genes are likely contributors to STEC pathogenicity (2, 3, 13, 50). Ritchie et al. (50) found both genes in all analyzed HUS-associated STEC isolates. In all O26:H11 isolates we obtained, stx genes were present in combination with eae genes. Only the O26:H32 isolate lacked an eae gene. To date, 10 distinct variants of eae have been described (1, 19, 36, 45, 47). Some serotypes were closely associated with a particular intimin variant: the O157 serogroup was linked to γ-eae, the O26 serogroup to β-eae, and the O103 serogroup to ɛ-eae (4, 19, 20, 58). Our study confirms these associations. All bovine O26:H11 isolates were also typed as members of the β-eae subgroup. A translocated intimin receptor gene (tir gene) and the type III secreted proteins encoded by the espA and espB genes were found in all 56 O26:H11 isolates but not in the O26:H32 isolate. These other tested locus of enterocyte effacement-associated genes belonged to the β-subgroups. These results are in accord with the results of China et al. (19), who detected the pathotypes β-eae, β-tir, β-espA, and β-espB in all investigated human O26 strains. Like the eae gene, the EHEC-hlyA gene was found in association with severe clinical disease in humans (52). Aldick et al. (2) showed that EHEC hemolysin is toxic (cytolytic) to human microvascular endothelial cells and may thus contribute to the pathogenesis of HUS. In our study, the EHEC-hlyA gene was detected in 50 of the 56 bovine E. coli O26:H11 isolates which harbored virulence-associated plasmids of different sizes (Table (Table1).1). The presence of virulence-associated plasmids corresponded to the occurrence of additional virulence markers such as the espP and katP genes (17). The katP gene and the espP gene were detected in 49 and 50 of the 56 O26:H11 isolates, respectively. The espP gene was missing in six of the seven bovine O26:H11 isolates in which the katP genes were also absent. Both genes were not found in the O26:H32 isolate (Table (Table1).1). Although we found large plasmids of the same size in O26:H11 isolates, they lacked one or more of the plasmid-associated virulence factors (Table (Table1).1). Two DNA probes were used to detect the efa1 genes by colony hybridization. (DNA probes were labeled with digoxigenin [DIG] with lifA1-lifA2 and lifA3-lifA4 primers [14] using the PCR DIG probe synthesis kit [Roche Diagnostics, Mannheim, Germany]; DIG Easy Hyb solution [Roche] was used for prehybridization and hybridization.) Positive results with both DNA probes were obtained for 52 of 56 E. coli O26:H11 isolates. A positive signal was only found in three isolates with the lifA1-lifA2 DNA probe and in one isolate with the lifA3-lifA4 probe. An efa1 gene was not detected in the O26:H32 isolate (Table (Table11).We also analyzed the spatial and temporal behavior of the O26:H11/H32 isolates in the beef herds by cluster analysis (conducted in PAUP* for Windows version 4.0, 2008 [http://paup.csit.fsu.edu/about.html]). This was performed with distance matrices using the neighbor-joining algorithm, an agglomerative cluster method which generates a phylogenetic tree. The distance matrices were calculated by pairwise comparisons of the fragmentation patterns produced by genomic typing through pulsed-field gel electrophoresis analysis with four restriction endonucleases (XbaI, NotI, BlnI, and SpeI) and the presence or absence of potential virulence markers (Fig. (Fig.11 and Table Table1).1). To this end, the total character difference was used, which counts the pairwise differences between two given patterns. During a monitoring program of 3 years in four cattle farms (29), different O26:H11 cluster groups and one O26:H32 isolate were detected in three different farms. The genetic distance of the O26:H32 isolate was very high relative to the O26:H11 isolates. Therefore, the O26:H32 isolate was outgrouped. The O26:H11 isolates of each farm represented independent cluster groups. The single isolate from farm D fitted better to the isolates from farm C than to those from farm A. This finding is in accord with the geographical distance between the farms. The fact that the farms were located in neighboring villages may suggest that direct or indirect connections between the farms were possible (e.g., by person contacts or animal trade). However, the isolates from farm C and farm D belonged to different sequence types (ST 21 and ST 396), which may argue against a direct connection. Interestingly, O26:H11 isolates with and without stx2 genes were detected in the same clusters. This phenomenon was observed in both farm A and farm C. In farm A, the isolates with additional stx2 genes were found in animal 27 and were grouped in clusters 8 and 9 (day 218). An stx2 gene was repeatedly found (four isolates) in the same animal (animal 27). The isolates grouped in cluster 8 on a later day of sampling (day 478). All other O26:H11 isolates grouped in the same clusters and obtained from the same animals (27 and 29) on different sampling days lacked an stx2 gene. Also, the isolates obtained from animal 27 on previous sampling days, which grouped in clusters 3 and 5, exhibited no stx2 genes. In farm C, the three isolates with additional stx2 genes obtained from animal 7 grouped in clusters 11 and 12. An stx2 gene was absent from all other O26:H11 isolates grouped in the same cluster 12 on later sampling days, and no other isolates of cluster 11 were found later on. However, we detected members of many clusters over relatively long periods (clusters 5, 8, and 9 in farm A and cluster 12 in farm C), but members of other clusters were only found on single occasions. This patchy temporal pattern is apparently not a unique property of O26:H11, as we found similar results for cluster groups of other EHEC serotypes of bovine origin (28). The isolates grouped in the dominant cluster 8 were found on 5 of 9 sampling days over a period of 10 months. In contrast, we found the members of clusters 4, 5, 9, and 12 only on two nonconsecutive sampling days. The period during which isolates of these groups were not detected was particularly long for cluster 4 (231 days). We also observed the coexistence of different clusters over long periods in the same farm and in the same cattle (clusters 8 and 9), while one of the clusters dominated. Transmission of clusters between cattle was also observed. These results suggest that some of the EHEC O26:H11 strains had the potential for a longer persistence in the host population, while others had not. The reasons for this difference are not yet clear. Perhaps the incomplete efa1 gene found in isolates of clusters which were only detected once might explain why some strains disappeared rapidly. Efa1 has been discussed as a potential E. coli colonization factor for the bovine intestine used by non-O157 STEC, including O26 (54, 56). The O165:H25 cluster detected during a longer period in farm B may have disappeared after it had lost its efa1 gene (28). The precise biological activity of Efa1 in EHEC O26 is not yet known, but it has been demonstrated that the molecule is a non-Stx virulence determinant which can increase the virulence of EHEC O26 in humans (8).Open in a separate windowFIG. 1.Neighbor-joining tree of bovine E. coli O26:H11/H32 strains based on the restriction pattern obtained after digestion with XbaI, NotI, BlnI, and SpeI.We distinguished 12 different clusters, but complete genetic identity was only found in two isolates. The variations in the O26:H11 clusters may be due to increasing competition between the bacterial populations of the various subtypes in the bovine intestine or to potential interactions between EHEC O26:H11 and the host.The ephemeral occurrence of additional stx2 genes in different clusters and farms may be the result of recombination events due to horizontal gene transfer (16). The loss of stx genes may occur rapidly in the course of an infection, but the reincorporation by induction of an stx-carrying bacteriophage into the O26:H11 strains is possible at any time (9, 40). Nevertheless, an additional stx2 gene may increase the dangerousness of the respective EHEC O26:H11 strains. While all patients involved in an outbreak caused by an EHEC O26:H11 strain harboring the gene encoding Stx2 developed HUS (41), the persons affected by another outbreak caused by an EHEC O26:H11 strain that produced exclusively Stx1 had only uncomplicated diarrhea (60).In conclusion, our results showed that bovine O26:H11 isolates can carry virulence factors of EHEC that are strongly associated with EHEC-related disease in humans, particularly with severe clinical manifestations such as hemorrhagic colitis and HUS. Therefore, strains of bovine origin may represent a considerable risk for human infection. Moreover, some clusters of EHEC O26:H11 persisted in cattle and farms over longer periods, which may increase the risk of transmission to other animals and humans even further.  相似文献   

10.
11.
12.
13.
Melioidosis has been considered an emerging disease in Brazil since the first cases were reported to occur in the northeast region. This study investigated two municipalities in Ceará state where melioidosis cases have been confirmed to occur. Burkholderia pseudomallei was isolated in 26 (4.3%) of 600 samples in the dry and rainy seasons.Melioidosis is an endemic disease in Southeast Asia and northern Australia (2, 4) and also occurs sporadically in other parts of the world (3, 7). Human melioidosis was reported to occur in Brazil only in 2003, when a family outbreak afflicted four sisters in the rural part of the municipality of Tejuçuoca, Ceará state (14). After this episode, there was one reported case of melioidosis in 2004 in the rural area of Banabuiú, Ceará (14). And in 2005, a case of melioidosis associated with near drowning after a car accident was confirmed to occur in Aracoiaba, Ceará (11).The goal of this study was to investigate the Tejuçuoca and Banabuiú municipalities, where human cases of melioidosis have been confirmed to occur, and to gain a better understanding of the ecology of Burkholderia pseudomallei in this region.We chose as central points of the study the residences and surrounding areas of the melioidosis patients in the rural areas of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50′W) (Fig. (Fig.1).1). There are two well-defined seasons in each of these locations: one rainy (running from January to May) and one dry (from June to December). A total of 600 samples were collected at five sites in Tejuçuoca (T1, T2, T3, T4, and T5) and five in Banabuiú (B1, B2, B3, B4, and B5), distributed as follows (Fig. (Fig.2):2): backyards (B1 and T1), places shaded by trees (B2 and T2), water courses (B3 and T3), wet places (B4 and T4), and stock breeding areas (B5 and T5).Open in a separate windowFIG. 1.Municipalities of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50″W).Open in a separate windowFIG. 2.Soil sampling sites in Banabuiú and Tejuçuoca.Once a month for 12 months (a complete dry/rainy cycle), five samples were gathered at five different depths: at the surface and at 10, 20, 30 and 40 cm (Table (Table1).1). The samples were gathered according to the method used by Inglis et al. (9). Additionally, the sample processing and B. pseudomallei identification were carried out as previously reported (1, 8, 9).

TABLE 1.

Distribution of samples with isolates by site and soil depth
Sitesa and depth (cm)No. of B. pseudomallei isolates in samples from:
Banabuiú (n = 300)Tejuçuoca (n = 300)Total (n = 600)
B1/T13
    Surface2
    10
    201
    30
    40
B2/T21
    Surface1
    10
    20
    30
    40
B3/T315
    Surface2
    102
    204
    303
    404
B4/T45
    Surface
    101
    201
    3011
    401
B5/T52
    Surface
    10
    20
    302
    40
Total62026
Open in a separate windowaSites designated with B are in Banabuiú, and sites designated with T are in Tejuçuoca. See the text for details.The data on weather and soil composition were obtained from specialized government institutions, such as FUNCEME, IPECE, and EMBRAPA. The average annual temperature in both municipalities is between 26 and 28°C. In 2007, the annual rainfall in Tejuçuoca was 496.8 mm, and that in Banabuiú was 766.8 mm. There are a range of soil types in both Tejuçuoca and Banabuiú: noncalcic brown, sodic planossolic, red-yellow podzolic, and litholic. In Banabuiú, there are also alluvial and cambisol soils. The characteristic vegetation in both municipalities is caatinga (scrublands).There were isolates of B. pseudomallei in 26 (4.3%) of the 600 samples collected. The bacterium was isolated at a rate (3%) similar to that previously reported (9). The bacterium isolation occurred in both the dry (53.8%) and the rainy (46.2%) seasons. Tejuçuoca represented 76.9% (20/26) of the strains isolated. Four sites in Tejuçuoca (T1, T3, T4, and T5) and three in Banabuiú (B1, B2, and B4) presented isolates of the bacterium (Table (Table1).1). The isolation of the B. pseudomallei strains varied from the surface down to 40 cm. However, 17 of the 26 positive samples (65.3%) were found at depths between 20 and 40 cm (Table (Table1).1). Only two isolates were found at the surface during the dry season.A study in Vietnam (13) and one in Australia (9) reported the presence of B. pseudomallei near the houses of melioidosis patients. In our study, the same thing happened. Site T3 (15/26; 57.6%) was located 290 m from the patient''s house, as reported by the Rolim group (14).B. pseudomallei was isolated from a sheep paddock in Australia, where animals sought shelter below mango and fig trees (17). In our study, the bacterium was isolated at site T5, a goat corral alongside the house where the outbreak occurred in Tejuçuoca. Four sites in places shaded by trees yielded positive samples (30.7%) in both Tejuçuoca (palm trees) and Banabuiú (mango trees). Additionally, B. pseudomallei was isolated on three occasions from a cornfield (site 4B) located alongside the house of the melioidosis patient in Banabuiú.In the main areas of endemicity, the disease is more prevalent in the rainy season (4, 5, 16). The outbreak in Tejuçuoca was related to rainfall (14). Besides the association of cases of the disease with rainfall itself, the isolation of B. pseudomallei in soil and water was also demonstrated during the dry season (12, 15). An Australian study isolated strains from soil and water during the dry and rainy seasons (17). A Thai study also reported B. pseudomallei in the dry season (18). In our study, the isolation of B. pseudomallei took place either at the end of the wet season or in the dry months. Fourteen of the positive samples (53.8%) were collected during the dry season, albeit near a river or reservoir (sites T3 and B4).Physical, biological, and chemical soil features appear to influence the survival of B. pseudomallei (6, 10). In the present study, the soil was classified as litholic with sandy or clayey textures. It is susceptible to erosion, and when there is a lack of water, it is subject to salinization. During the dry season, the clay layer becomes dried, cracked, and very hard. During the rainy season, it becomes soggy and sticky. The isolation of B. pseudomallei in the dry season is possibly related to the capacity for adaptation of this soil, since the extreme conditions of lithosols do not prevent the bacterial growth and survival.It has been shown that B. pseudomallei is more often isolated at depths between 25 and 45 cm (17). In our study, 65.3% of the positive samples were taken at depths between 20 and 40 cm. Moreover, of these 17 samples, 10 (58.8%) were collected during the dry months. Also, unlike in other regions, two positive samples were taken from the surface in the period without rainfall.The rainfall in Tejuçuoca and Banabuiú is generally low, and temperatures do not vary significantly during the year. Therefore, the isolation of B. pseudomallei in these places occurs outside the rainfall, temperature, and moisture conditions observed in other regions of endemicity. Our data thus suggest that peculiar environmental features, such as soil composition, might favor the multiplication of B. pseudomallei in northeast Brazil.  相似文献   

14.
Short-chain alcohol dehydrogenase, encoded by the gene Tsib_0319 from the hyperthermophilic archaeon Thermococcus sibiricus, was expressed in Escherichia coli, purified and characterized as an NADPH-dependent enantioselective oxidoreductase with broad substrate specificity. The enzyme exhibits extremely high thermophilicity, thermostability, and tolerance to organic solvents and salts.Alcohol dehydrogenases (ADHs; EC 1.1.1.1.) catalyze the interconversion of alcohols to their corresponding aldehydes or ketones by using different redox-mediating cofactors. NAD(P)-dependent ADHs, due to their broad substrate specificity and enantioselectivity, have attracted particular attention as catalysts in industrial processes (5). However, mesophilic ADHs are unstable at high temperatures, sensitive to organic solvents, and often lose activity during immobilization. In this relation, there is a considerable interest in ADHs from extremophilic microorganisms; among them, Archaea are of great interest. The representatives of all groups of NAD(P)-dependent ADHs have been detected in genomes of Archaea (11, 12); however, only a few enzymes have been characterized, and the great majority of them belong to medium-chain (3, 4, 14, 16, 19) or long-chain iron-activated ADHs (1, 8, 9). Up to now, a single short-chain archaeal ADH from Pyrococcus furiosus (10, 18) and only one archaeal aldo-keto reductase also from P. furiosus (11) have been characterized.Thermococcus sibiricus is a hyperthermophilic anaerobic archaeon isolated from a high-temperature oil reservoir capable of growth on complex organic substrates (15). The complete genome sequence of T. sibiricus has been recently determined and annotated (13). Several ADHs are encoded by the T. sibiricus genome, including three short-chain ADHs (Tsib_0319, Tsib_0703, and Tsib_1998) (13). In this report, we describe the cloning and expression of the Tsib_0319 gene from T. sibiricus and the purification and the biochemical characterization of its product, the thermostable short-chain ADH (TsAdh319).The Tsib_0319 gene encodes a protein with a size of 234 amino acids and the calculated molecular mass of 26.2 kDa. TsAdh319 has an 85% degree of sequence identity with short-chain ADH from P. furiosus (AdhA; PF_0074) (18). Besides AdhA, close homologs of TsAdh319 were found among different bacterial ADHs, but not archaeal ADHs. The gene flanked by the XhoI and BamHI sites was PCR amplified using two primers (sense primer, 5′-GTTCTCGAGATGAAGGTTGCTGTGATAACAGGG-3′, and antisense primer, 5′-GCTGGATCCTCAGTATTCTGGTCTCTGGTAGACGG-3′) and cloned into the pET-15b vector. TsAdh319 was overexpressed, with an N-terminal His6 tag in Escherichia coli Rosetta-gami (DE3) and purified to homogeneity by metallochelating chromatography (Hi-Trap chelating HP column; GE Healthcare) followed by gel filtration on Superdex 200 10/300 GL column (GE Healthcare) equilibrated in 50 mM Tris-HCl (pH 7.5) with 200 mM NaCl. The homogeneity and the correspondence to the calculated molecular mass of 28.7 kDa were verified by SDS-PAGE (7). The molecular mass of native TsAdh319 was 56 to 60 kDa, which confirmed the dimeric structure in solution.The standard ADH activity measurement was made spectrophotometrically at the optimal pH by following either the reduction of NADP (in 50 mM Gly-NaOH buffer; pH 10.5) or the oxidation of NADPH (in 0.1 M sodium phosphate buffer; pH 7.5) at 340 nm at 60°C. The enzyme exhibited a strong preference for NADP(H) and broad substrate specificity (Table (Table1).1). The highest oxidation rates were found with pentoses d-arabinose (2.0 U mg−1) and d-xylose (2.46 U mg−1), and the highest reduction rates were found with dimethylglyoxal (5.9 U mg−1) and pyruvaldehyde (2.2 U mg−1). The enzyme did not reduce sugars which were good substrates for the oxidation reaction. The kinetic parameters of TsAdh319 determined for the preferred substrates are shown in Table Table2.2. The enantioselectivity of the enzyme was estimated by measuring the conversion rates of 2-butanol enantiomers. TsAdh319 showed an evident preference, >2-fold, for (S)-2-butanol over (RS)-2-butanol. The enzyme stereoselectivity is confirmed by the preferred oxidation of d-arabinose over l-arabinose (Table (Table1).1). The fact that TsAdh319 is metal independent was supported by the absence of a significant effect of TsAdh319 preincubation with 10 mM Me2+ for 30 min before measuring the activity in the presence of 1 mM Me2+ or EDTA (Table (Table3).3). TsAdh319 also exhibited a halophilic property, so the enzyme activity increased in the presence of NaCl and KCl and the activation was maintained even at concentration of 4 M and 3 M, respectively (Table (Table33).

TABLE 1.

Substrate specificity of TsAdh319
SubstrateaRelative activity (%)
Oxidation reactionb
    Methanol0
    2-Methoxyethanol0
    Ethanol36
    1-Butanol80
    2-Propanol100
    (RS)-(±)-2-Butanol86
    (S)-(+)-2-Butanol196
    2-Pentanol67
    1-Phenylmethanol180
    1.3-Butanediol91
    Ethyleneglycol0
    Glycerol16
    d-Arabinose*200
    l-Arabinose*17
    d-Xylose*246
    d-Ribose*35
    d-Glucose*146
    d-Mannose*48
    d-Galactose*0
    Cellobiose*71
Reduction reactionc
    Pyruvaldehyde100
    Dimethylglyoxal270
    Glyoxylic acid36
    Acetone0
    Cyclopentanone0
    Cyclohexanone4
    3-Methyl-2-pentanone*13
    d-Arabinose*0
    d-Xylose*0
    d-Glucose*0
    Cellobiose*0
Open in a separate windowaSubstrates were present in 250 mM or 50 mM (*) concentrations.bRelative rates, measured under standard conditions, were calculated by defining the activity for 2-propanol as 100%, which corresponds to 1.0 U mg−1. Data are averages from triplicate experiments.cRelative rates, measured under standard conditions, were calculated by defining the activity for pyruvaldehyde as 100%, which corresponds to 2.2 U mg−1. Data are averages from triplicate experiments.

TABLE 2.

Apparent Km and Vmax values for TsAdh319
Coenzyme or substrateApparent Km (mM)Vmax (U mg−1)kcat (s−1)
NADPa0.022 ± 0.0020.94 ± 0.020.45 ± 0.01
NADPHb0.020 ± 0.0033.16 ± 0.111.51 ± 0.05
2-Propanol168 ± 291.10 ± 0.090.53 ± 0.04
d-Xylose54.4 ± 7.41.47 ± 0.090.70 ± 0.04
Pyruvaldehyde17.75 ± 3.384.26 ± 0.402.04 ± 0.19
Open in a separate windowaActivity was measured under standard conditions with 2-propanol. Data are averages from triplicate experiments.bActivity was measured under standard conditions with pyruvaldehyde. Data are averages from triplicate experiments.

TABLE 3.

Effect of various ions and EDTA on TsAdh319a
CompoundConcn (mM)Relative activity (%)
None0100
NaCl400206
600227
4,000230
KCl600147
2,000200
3,000194
MgCl21078
CoCl210105
NiSO410100
ZnSO41079
FeSO41074
EDTA1100
580
Open in a separate windowaThe activity was measured under standard conditions with 2-propanol; relative rates were calculated by defining the activity without salts as 100%, which corresponds to 0.9 U mg−1. Data are averages from duplicate experiments.The most essential distinctions of TsAdh319 are the thermophilicity and high thermostability of the enzyme. The optimum temperature for the 2-propanol oxidation catalyzed by TsAdh319 was not achieved. The initial reaction rate of oxidation increased up to 100°C (Fig. (Fig.1).1). The Arrhenius plot is a straight line, typical of a single rate-limited thermally activated process, but there is no obvious transition point due to the temperature-dependent conformational changes of the protein molecule. The activation energy for the oxidation of 2-propanol was estimated at 84.0 ± 5.8 kJ·mol−1. The thermostability of TsAdh319 was calculated from residual TsAdh319 activity after preincubation of 0.4 mg/ml enzyme solution in 50 mM Tris-HCl buffer (pH 7.5) containing 200 mM NaCl at 70, 80, 90, or 100°C. The preincubation at 70°C or 80°C for 1.5 h did not cause a decrease in the TsAdh319 activity, but provoked slight activation. The residual TsAdh319 activities began to decrease after 2 h of preincubation at 70°C or 80°C and were 10% and 15% down from the control, respectively. The determined half-life values of TsAdh319 were 2 h at 90°C and 1 h at 100°C.Open in a separate windowFIG. 1.Temperature dependence of the initial rate of the 2-propanol reduction by TsAdh319. The reaction was initiated by enzyme addition to a prewarmed 2-propanol-NADP mixture. The inset shows the Arrhenius plot of the same data.Protein thermostability often correlates with such important biotechnological properties as increased solvent tolerance (2). We tested the influence of organic solvents at a high concentration (50% [vol/vol]) on TsAdh319 by using either preincubation of the enzyme at a concentration of 0.2 mg/ml with solvents for 4 h at 55°C or solvent addition into the reaction mixture to distinguish the effect of solvent on the protein stability and on the enzyme activity. TsAdh319 showed significant solvent tolerance in both cases (Table (Table4),4), and the effects of solvents could be modulated by salts, acting apparently as molecular lyoprotectants (17). Furthermore, TsAdh319 maintained 57% of its activity in 25% (vol/vol) 2-propanol, which could be used as the cosubstrate in cofactor regeneration (6).

TABLE 4.

Influence of various solvents on TsAdh319 activitya
SolventRelative activity (%)bRelative activity (%)c
Buffer without NaClBuffer with 600 mM NaCl
None100100100
DMSOd98040
DMFAe1011341
Methanol98259
Acetonitrile9500
Ethyl acetate470*33*
Chloroform10579*81*
n-Hexane10560*118*
n-Decane3691*107*
Open in a separate windowaThe activity measured at the standard condition with 2-propanol as a substrate. Data are averages from triplicate experiments.bPreincubation for 4 h at 55°C in the presence of 50% (vol/vol) of solvent prior the activity assay.cWithout preincubation, solvent addition to the reaction mixture up to 50% (vol/vol) or using the buffer saturated by a solvent (*).dDMSO, dimethyl sulfoxide.eDMFA, dimethylformamide.From all the aforesaid we may suppose TsAdh319 or its improved variant to be interesting both for the investigation of structural features of protein tolerance and for biotechnological applications.  相似文献   

15.
Escherichia coli isolates (72 commensal and 10 O157:H7 isolates) were compared with regard to physiological and growth parameters related to their ability to survive and persist in the gastrointestinal tract and found to be similar. We propose that nonhuman hosts in E. coli O157:H7 strains function similarly to other E. coli strains in regard to attributes relevant to gastrointestinal colonization.Escherichia coli is well known for its ecological versatility (15). A life cycle which includes both gastrointestinal and environmental stages has been stressed by both Savageau (15) and Adamowicz et al. (1). The gastrointestinal stage would be subjected to acid and detergent stress. The environmental stage is implicit in E. coli having transport systems for fungal siderophores (4) as well as pyrroloquinoline quinone-dependent periplasmic glucose utilization (1) because their presence indicates evolution in a location containing fungal siderophores and pyrroloquinoline quinone (1).Since its recognition as a food-borne pathogen, there have been numerous outbreaks of food-borne infection due to E. coli O157:H7, in both ground beef and vegetable crops (6, 13). Cattle are widely considered to be the primary reservoir of E. coli O157:H7 (14), but E. coli O157:H7 does not appear to cause disease in cattle. To what extent is E. coli O157:H7 physiologically unique compared to the other naturally occurring E. coli strains? We feel that the uniqueness of E. coli O157:H7 should be evaluated against a backdrop of other wild-type E. coli strains, and in this regard, we chose the 72-strain ECOR reference collection originally described by Ochman and Selander (10). These strains were chosen from a collection of 2,600 E. coli isolates to provide diversity with regard to host species, geographical distribution, and electromorph profiles at 11 enzyme loci (10).In our study we compared the 72 strains of the ECOR collection against 10 strains of E. coli O157:H7 and six strains of E. coli which had been in laboratory use for many years (Table (Table1).1). The in vitro comparisons were made with regard to factors potentially relevant to the bacteria''s ability to colonize animal guts, i.e., acid tolerance, detergent tolerance, and the presence of the Entner-Doudoroff (ED) pathway (Table (Table2).2). Our longstanding interest in the ED pathway (11) derives in part from work by Paul Cohen''s group (16, 17) showing that the ED pathway is important for E. coli colonization of the mouse large intestine. Growth was assessed by replica plating 88 strains of E. coli under 40 conditions (Table (Table2).2). These included two LB controls (aerobic and anaerobic), 14 for detergent stress (sodium dodecyl sulfate [SDS], hexadecyltrimethylammonium bromide [CTAB], and benzalkonium chloride, both aerobic and anaerobic), 16 for acid stress (pH 6.5, 6.0, 5.0, 4.6, 4.3, 4.2, 4.1, and 4.0), four for the ability to grow in a defined minimal medium (M63 glucose salts with and without thiamine), and four for the presence or absence of a functional ED pathway (M63 with gluconate or glucuronate). All tests were done with duplicate plates in two or three separate trials. The data are available in Tables S1 to S14 in the supplemental material, and they are summarized in Table Table22.

TABLE 1.

E. coli strains used in this study
E. coli strain (n)Source
ECOR strains (72)Thomas Whittman
Laboratory adapted (6)
    K-12 DavisPaul Blum
    CG5C 4401Paul Blum
    K-12 StanfordPaul Blum
    W3110Paul Blum
    BTyler Kokjohn
    AB 1157Tyler Kokjohn
O157:H7 (10)
    FRIK 528Andrew Benson
    ATCC 43895Andrew Benson
    MC 1061Andrew Benson
    C536Tim Cebula
    C503Tim Cebula
    C535Tim Cebula
    ATCC 43889William Cray, Jr.
    ATCC 43890William Cray, Jr.
    ATCC 43888Willaim Cray, Jr.
    ATCC 43894William Cray, Jr.
Open in a separate window

TABLE 2.

Physiological comparison of 88 strains of Escherichia coli
Growth medium or conditionOxygencNo. of strains with type of growthb
ECOR strains (n = 72)
Laboratory strains (n = 6)
O157:H7 strains (n = 10)
GoodPoorNoneVariableGoodPoorNoneVariableGoodPoorNoneVariable
LB controlaBoth72000600010000
1% SDSAerobic6930060008002
5% SDSAerobic6840060008200
1% SDSAnaerobic53154023101702
5% SDSAnaerobic0684004200704
CTABd (all)Both00720006000100
0.05% BACAerobic31158202220091
0.2% BACAerobic01710105000100
0.05% BACAnaerobic2367001500091
0.2% BACAnaerobic00720006000100
pH 6.5Both72000600010000
pH 6Both72000600010000
pH 5Both7020060009001
pH 4.6Both70200600010000
pH 4.3Aerobic14015731203205
pH 4.3Anaerobic6930031201100
pH 4.1 or 4.2Aerobic00720NDgND
pH 4.0Both0072000600091
M63 with supplemente
    GlucoseAerobicf6912050109010
    GlucoseAnaerobicf7002050109010
    GluconateBoth6912050109010
    GlucuronateAerobic6822050109010
    GlucuronateAnaerobic6912050109010
Open in a separate windowaEight LB controls were run, two for each set of LB experiments: SDS, CTAB, benzalkonium chloride (BAC), and pH stress.bGrowth was measured as either +++, +, or 0 (good, poor, and none, respectively), with +++ being the growth achieved on the LB control plates. “Variable” means that two or three replicates did not agree. All experiments were done at 37°C.c“Anaerobic” refers to use of an Oxoid anaerobic chamber. Aerobic and anaerobic growth data are presented together when the results were identical and separately when the results were not the same or the anaerobic set had not been done. LB plates were measured after 1 (aerobic) or 2 (anaerobic) days, and the M63 plates were measured after 2 or 3 days.dCTAB used at 0.05, 0.2%, and 0.4%.eM63 defined medium (3) was supplemented with glucose, gluconate, or glucuronate, all at 0.2%.fIdentical results were obtained with and without 0.0001% thiamine.gND, not determined.  相似文献   

16.
The effect of eliminating d-lactate synthesis in poly(3-hydroxybutyrate) (PHB)-accumulating recombinant Escherichia coli (K24K) was analyzed using glycerol as a substrate. K24KL, an ldhA derivative, produced more biomass and had altered carbon partitioning among the metabolic products, probably due to the increased availability of carbon precursors and reducing power. This resulted in a significant increase of PHB and ethanol synthesis and a decrease in acetate production. Cofactor measurements revealed that cultures of K24K and K24KL had a high intracellular NADPH content and that the NADPH/NADP+ ratio was higher than the NADH/NAD+ ratio. The ldhA mutation affected cofactor distribution, resulting in a more reduced intracellular state, mainly due to a further increase in NADPH/NADP+. In 60-h fed-batch cultures, K24KL reached 41.9 g·liter−1 biomass and accumulated PHB up to 63% ± 1% (wt/wt), with a PHB yield on glycerol of 0.41 ± 0.03 g·g−1, the highest reported using this substrate.Poly(3-hydroxybutyrate) (PHB) is the best-known and most common polyhydroxyalkanoate (PHA). PHAs are polymers with thermoplastic properties that are totally biodegradable by microorganisms present in most environments and that can be produced from different renewable carbon sources (38). Accumulated as intracellular granules by many bacteria under unfavorable conditions (1, 21), PHAs are carbon and energy reserves and also act as electron sinks, enhancing the fitness and stress resistance of bacteria and contributing to redox balance (12, 30). Escherichia coli offers a well-defined physiological environment for the construction and manipulation of various metabolic pathways to produce different bioproducts, such as PHB, from cost-effective carbon sources.In recent years, a significant increase in the production of biodiesel has caused a sharp fall in the cost of glycerol, the main by-product of biodiesel synthesis. As a result, glycerol has become a very attractive substrate for bacterial fermentations (10), specially for reduced products, such as PHB (36). The E. coli strain used in this work, K24K, carries phaBAC, the structural genes responsible for PHB synthesis, from Azotobacter sp. strain FA8 (23) (Table (Table1).1). The pha genes in K24K are expressed from a chimeric promoter and consequently are not subject to the genetic regulatory systems present in natural PHA producers. Because of this, it can be assumed that regulation of PHA synthesis in the recombinants is restricted by enzyme activity levels, modulated principally by substrate availability. In most natural producers, and also in PHB-producing E. coli recombinants, PHB is synthesized through the condensation of two molecules of acetyl-coenzyme A (acetyl-CoA), catalyzed by an acetoacetyl-CoA transferase or 3-ketothiolase, resulting in acetoacetyl-CoA. This compound is subsequently reduced by an NAD(P)H-dependent acetoacetyl-CoA reductase to R-(−)-3-hydroxybutyryl-CoA, which is then polymerized by a specific PHA synthase (34).

TABLE 1.

E. coli strains, plasmids, and oligonucleotides used in this study
Strain, plasmid, or oligonucleotideRelevant characteristicsbReference or source
E. coli strains
    K1060aFfadE62 lacI60 tyrT58(AS) fabB5 mel-129
    K24Same as K1060, carrying pJP24; Apr23
    K24KSame as K1060, carrying pJP24K; Apr Kmr23
    ALS786aF λrph-1 ΔldhA::kan; Kmr14
    K24LTSame as K1060 but ΔldhA::kan by K1060 × P1(ALS786), carrying pJP24; Apr KmrThis work
    K24KLSame as K1060 but ΔldhA by allelic replacement, carrying pJP24K; KmrThis work
    TA3522aF λ Δ(his-gnd)861 hisJo-7012
    TA3514aSame as TA3522 but pta-20019
    TA3522LSame as TA3522 but ΔldhA::kan by TA3522 × P1(ALS786); KmrThis work
    TA3514LSame as TA3514 but ΔldhA::kan by TA3514 × P1(ALS786); KmrThis work
Plasmids
    pQE32Expression vector, ColE1 ori; AprQiagen GmbH, Hilden, Germany
    pJP24pQE32 derivative expressing a 4.3-kb BamHI-HindIII insert containing the phaBAC genes from Azotobacter sp. strain FA8 under the control of a T5 promoter/lac operator element; Apr23
    pJP24KpJP24 derivative; Apr Kmr23
    pCP20Helper plasmid used for kan excision; Saccharomyces cerevisiae FLP λ cI857 λ PRrepA(Ts); Apr Cmr7
Oligonucleotides
    ΔldhA-F5′-TAT TTT TAG TAG CTT AAA TGT GAT TCA ACA TCA CTG GAG AAA GTC TTA TGG TGT AGG CTG GAG CTG CTT C-3′This work
    ΔldhA-R5′-CTC CCC TGG AAT GCA GGG GAG CGG CAA GAT TAA ACC AGT TCG TTC GGG CAC ATA TGA ATA TCC TCC TTA G-3′This work
Open in a separate windowaStrain obtained through the E. coli Genetic Stock Center, Yale University, New Haven, CT.bFor oligonucleotides, the ATG codon of ldhA is underlined and the sequences with homology to FRT-kan-FRT in the template plasmid pKD4 (11) are shown in boldface.Cells growing on glycerol are in a more reduced intracellular state than cells grown on glucose under similar conditions of oxygen availability. This has a significant effect on the intracellular redox state, which causes the cells to direct carbon flow toward the synthesis of more-reduced products when glycerol is used than when glucose is used in order to achieve redox balance (31). When metabolic product distribution was analyzed in bioreactor cultures of K24K using glucose or glycerol as the substrate, product distributions with the two substrates were found to be different, as glycerol-grown cultures produced smaller amounts of acetate, lactate, and formate and more ethanol than those grown on glucose. However, PHB production from glycerol was lower than that from glucose, except under conditions of low oxygen availability (13).Manipulations to enhance the synthesis of a metabolic product include several approaches to increase the availability of the substrates needed for its formation or to inhibit competing pathways. The effect of eliminating competing pathways on PHB production from glucose has been investigated through the inactivation of different genes, such as those encoding enzymes participating in the synthesis of acetate (ackA, pta, and poxB) or d-lactate (ldhA). A pta mutant, which produces very little acetate (6), and an frdA ldhA double mutant (40) had increased PHB accumulation from glucose. A recent report using an ackA pta poxB ldhA adhE mutant under microaerobic conditions attained similar results (17). The inactivation of ldhA has also been shown to have an important effect on the metabolic product distribution in recombinant E. coli with glycerol as the carbon source, promoting ethanol synthesis (28). In the present work we analyzed the effect of ldhA inactivation in strain K24K using glycerol as the carbon source, with special emphasis on changes in carbon distribution and in the intracellular redox state, determined through cofactor levels.  相似文献   

17.
18.
Vertebrate genomic assemblies were analyzed for endogenous sequences related to any known viruses with single-stranded DNA genomes. Numerous high-confidence examples related to the Circoviridae and two genera in the family Parvoviridae, the parvoviruses and dependoviruses, were found and were broadly distributed among 31 of the 49 vertebrate species tested. Our analyses indicate that the ages of both virus families may exceed 40 to 50 million years. Shared features of the replication strategies of these viruses may explain the high incidence of the integrations.It has long been appreciated that retroviruses can contribute significantly to the genetic makeup of host organisms. Genes related to certain other viruses with single-stranded RNA genomes, formerly considered to be most unlikely candidates for such contribution, have recently been detected throughout the vertebrate phylogenetic tree (1, 6, 13). Here, we report that viruses with single-stranded DNA (ssDNA) genomes have also contributed to the genetic makeup of many organisms, stretching back as far as the Paleocene period and possibly the late Cretaceous period of evolution.Determining the evolutionary ages of viruses can be problematic, as their mutation rates may be high and their replication may be rapid but also sporadic. To establish a lower age limit for currently circulating ssDNA viruses, we analyzed 49 published vertebrate genomic assemblies for the presence of sequences derived from the NCBI RefSeq database of 2,382 proteins from known viruses in this category, representing a total of 23 classified genera from 7 virus families. Our survey uncovered numerous high-confidence examples of endogenous sequences related to the Circoviridae and to two genera in the family Parvoviridae: the parvoviruses and dependoviruses (Fig. (Fig.11).Open in a separate windowFIG. 1.Phylogenetic tree of vertebrate organisms and history of ssDNA virus integrations. Times of integration of ancestral dependoviruses (yellow icosahedrons), parvoviruses (blue icosahedrons), and circoviruses (triangles) are approximate.The Dependovirus and Parvovirus genomes are typically 4 to 6 kb in length, include 2 major open reading frames (encoding replicase proteins [Rep and NS1, respectively] and capsid proteins [Cap and VP1, respectively]), and have characteristic hairpin structures at both ends (Fig. (Fig.2).2). For replication, these viruses depend on host enzymes that are recruited by the viral replicase proteins to the hairpin regions, where self-primed viral DNA synthesis is initiated (2). Circovirus genomes are typically ∼2-kb circles. DNA of the type species, porcine circovirus 1 (PCV-1), contains a stem-loop structure within the origin of replication (Fig. (Fig.2),2), and the largest open reading frame includes sequences that are homologous to the Parvovirus replicase open reading frame (9, 11). The circoviruses also depend on host enzymes for replication, and DNA synthesis is self-primed from a 3′-OH end formed by endonucleolytic cleavage of the stem-loop structure (4). The frequency of Dependovirus infection is estimated to be as high as 90% within an individual''s lifetime. None of the dependoviruses have been associated with human disease, but related viruses in the family Parvoviridae (e.g., erythrovirus B19 and possibly human bocavirus) are pathogenic for humans, and members of both the Parvoviridae and the Circoviridae can cause a variety of animal diseases (2, 4).Open in a separate windowFIG. 2.Schematics illustrating the structure and organization of Parvoviridae and Circoviridae genomes and origins of several of the longest-integrated ancestral viral sequences found in vertebrates. Integrations were aligned to the Dependovirus adeno-associated virus 2 (AAV2), the Parvovirus minute virus of mice (MVM), and the Circovirus porcine circovirus 1 (PCV-1). The inverted terminal repeat (ITR) sequences in the Dependovirus and Parvovirus genomes are depicted on an expanded scale. A linear representation of the circular genome of PCV-1 is shown with the 10-bp stem-loop structure on an expanded scale. Horizontal lines beneath the maps indicate the lengths of similar sequences that could be identified by BLAST. The numbers indicate the locations of amino acids in the viral proteins where the sequence similarities in the endogenous insertions start and end. The actual ancestral virus-derived integrated sequences may extend beyond the indicated regions.With some ancestral endogenous sequences that we identified, phylogenetic comparisons can be used to estimate age. For example, as a Dependovirus-like sequence is present at the same location in the genomes of mice and rats, the ancestral virus must have existed before their divergence, more than 20 million years ago. Some Circovirus- and Dependovirus-related integrations also predate the split between dog and panda, about 42 million years ago. However, in most other cases, we rely on an indirect method for estimating age (1). As genomic sequences evolve, they accumulate new stop codons and insertion/deletion-induced frameshifts. The rates of these events can be tied directly to the rates of neutral sequence drift and, therefore, the time of evolution. To apply this method, we first performed a BLAST search of vertebrate genomes for all known ssDNA virus proteins (BLAST options, -p tblastn -M BLOSUM62 -e 1e−4). Candidate sequences were then recorded, along with 5 kb of flanking regions, and then again aligned against the database of ssDNA viruses to find the most complete alignment (BLAST options, -t blastx -F F -w 15 -t 1500 -Z 150 -G 13 -E 1 -e 1e−2). Detected alignments were then compared with a neutral model of genome evolution, as described in the supplemental material, and the numbers of stop codons and frameshifts were converted into the expected genomic drift undergone by the sequences. The age of integration was then estimated from the known phylogeny of vertebrates (7, 10). Using these methods, we discovered that as many as 110 ssDNA virus-related sequences have been integrated into the 49 vertebrate genomes considered, during a time period ranging from the present to over 40 to 60 million years ago (Table (Table1;1; see also Tables S1 to S3 in the supplemental material).

TABLE 1.

Selected endogenous sequences in vertebrate genomes related to single-stranded DNA viruses
Virus group and vertebrate speciesInitial genomic search using TBLASTN
Best sequence homology identified using BLASTX
Predicted nucleotide drift (%)Integration labelAge (million yr) or timing of integration based on sequence aging
Chromosomal or scaffold locationProteinBLAST E value/% sequence identityMost similar virusaProteinCoordinatesNo. of stop codons/frameshifts
Circoviruses
    CatScaffold_62068Rep6E−05/37Canary circovirusRep4-2833/7 in 268 aab14.2fcECLG-182
Scaffold_24038Rep6E−06/51Columbid circovirusRep44-3174/5 in 231 aac15.2fcECLG-287
    DogChr5dRep7E−16/46Raven circovirusRep16-2636/5 in 250 aa17.6cfECLG-198
Chr22Rep1E−14/43Beak and feather disease virusRep7-2642/1 in 261 aac4.5cfECLG-254
    OpossumChr3Rep4E−46/44Finch circovirusRep2-2910/2 in 282 aa2.3mdECLG12
Cap6-360/0 in 30 aa
Dependoviruses
    DogChrXRep6E−05/55AAV5Rep239-4453/4 in 200 aa14.0cfEDLG-178
    DolphinGeneScaffold1475Rep8E−39/39Avian AAV DA1Rep79-4863/4 in 379 aac6.6ttEDLG-255
Cap4E−61/47Cap1-7384/7 in 678 aac
    ElephantScaffold_4Rep0/55AAV5Rep3-5890/0 in 579 aa0.0laEDLGRecent
    HyraxGeneScaffold5020Cap3E−34/53AAV3Cap485-7350/5 in 256 aa7.0pcEDLG-129
Scaffold_19252Rep9E−72/47Bovine AAVRep2-3488/4 in 348 aa14.3pcEDLG-260
    MegabatScaffold_5601Rep2E−13/31AAV2Rep315-4791/5 in 175 aa13.1pvEDLG-376
    MicrobatGeneScaffold2026Rep1E−117/50AAV2Rep1-6172/5 in 612 aa5.8mlEDLG-127
Cap9E−33/51Cap1-7312/9 in 509 aac
Scaffold_146492Cap6E−32/42AAV2Cap479-7320/3 in 252 aa4.2mlEDLG-219
    MouseChr1Rep2E−06/34AAV2Rep4-2063/5 in 191 aa17.1mmEDLG-139
Chr3Rep2E−24/31AAV5Rep71-47812/7 in 389 aa16.5mmEDLG-237
Cap2E−22/45Cap22-72412/10 in 649aac
Chr8Rep1E−08/46AAV2Rep314-4733/3 in 147 aa13.8mmEDLG-331
Cap1-1371/2 in 114 aa
    PandaScaffold2359Rep2E−06/37Bovine AAVRep238-4262/3 in 186 aa10.4amEDLG-159
    PikaScaffold_9941Rep4E−14/28AAV5Rep126-4152/2 in 282 aa5.4opEDLG14
    PlatypusChr2Rep9E−10/35Bovine AAVRep297-4374/3 in 138 aa17.1oaEDLG-179
Cap272-4191/2 in 150 aac
Contig12430Rep2E−09/47Bovine AAVRep353-4503/1 in 123 aa12.0oaEDLG-255
Cap2E−05/32Cap253-3672/1 in 116 aa
    RabbitChr10Rep3E−97/39AAV2Rep1-6193/9 in 613 aa9.3ocEDLG43
Cap5E−50/45Cap1-72310/9 in 675 aa
    RatChr13Rep2E−09/33AAV2Rep4-1752/4 in 177 aa13.3rnEDLG-128
Chr2Rep4E−18/40AAV5Rep1-46112/12 in 454 aa22.7rnEDLG-251
Chr19Rep2E−07/33AAV5Rep329-4642/4 in 136 aa16.1rnEDLG-335
Cap31-1332/1 in 93 aa
    TarsierScaffold_178326Rep4E−14/23AAV5Rep96-4652/3 in 356 aa5.3tsEDLG23
Parvoviruses
    Guinea pigScaffold_188Rep3E−24/46Porcine parvovirusRep313-5675/3 in 250 aa12.3cpEPLG-140
Cap1E−16/36Cap10-68911/12 in 672 aa
Scaffold_27Rep1E−50/39Canine parvovirusRep11-6401/4 in 616 aa5.3cpEPLG-217
Cap1E−38/39Porcine parvovirusCap3-7192/14 in 700 aa
    TenrecScaffold_260946Rep2E−20/38LuIII virusRep406-5984/4 in 190 aa19.0etEPLG-260
Cap11-63916/15 in 595 aa
    RatChr5Rep6E−10/56Canine parvovirusRep1-2820/0 in 312 aa0.6rnEPLGRecent
Cap0/62Cap637-6670/2 in 760 aa
Rep0/631-751
    OpossumChr3Rep2E−39/33LuIII virusRep7-57011/3 in 502 aa10.9mdEPLG-256
Cap7E−8/33Cap11-72914/7 in 704 aa
Chr6Rep6E−58/44Porcine parvovirusRep16-5633/7 in 534 aac4.6mdEPLG-324
Cap6E−60/38Cap10-7152/5 in 707 aac
    WallabyScaffold_108040Rep4E−74/62Canine parvovirusRep341-6450/0 in 287 aa1.3meEPLG-37
Cap8E−37/32Cap35-7380/4 in 687 aa
Scaffold_72496Rep2E−61/42Porcine parvovirusRep23-5674/3 in 531 aa5.7meEPLG-630
Cap2E−31/38Cap10-5326/4 in 514 aa
Scaffold_88340Rep7E−37/55Mouse parvovirus 1Rep344-5660/3 in 223 aa6.7meEPLG-1636
Cap7E−22/33Cap11-7136/9 in 700 aa
Open in a separate windowaSome ambiguity in choosing the most similar virus is possible. We generally used the alignment with the lowest E value in the BLAST results. However, one or two points in the exponent of an E value were sometimes sacrificed to achieve a longer sequence alignment.baa, amino acids.cThese sequences have long insertions compared to the present-day viruses. In all cases tested, these insertions originated from short interspersed elements (SINEs). These insertions were excluded from the counts of stop codons and frameshifts and the estimation of integration age.dChr, chromosome.It is important to recognize that there is an intrinsic limit on how far back in time we can reach to identify ancient endogenous viral sequences. First, the sequences must be identified with confidence by BLAST or similar programs. This requirement places a lower limit on sequence identity at about 20 to 30% of amino acids, or about 75% of nucleotides (nucleotides evolve nearly 2.5 times slower than the amino acid sequence they encode). Second, the related, present-day virus must have evolved at a rate that is not much higher than that of the endogenous sequences. The viruses for which ancestral endogenous sequences were identified in this study exhibit sequence drift similar to that associated with mammalian genomes. Setting this rate at 0.14% per million years of evolution (8), we arrive at 90 million years as the theoretical limit for the oldest sequences that can be identified using our methods. This limit drops to less than 35 million years for endogenous viral sequences in rodents and even lower for sequences related to viruses that evolve faster than mammalian genomes.The most widespread integrations found in our survey are derived from the dependoviruses. These include nearly complete genomes related to adeno-associated virus (AAV) in microbat, wallaby, dolphin, rabbit, mouse, and baboon (Fig. (Fig.2).2). We did not detect inverted terminal repeats in several integrations tested, even though repeats are common in the present-day dependoviruses. This result could be explained by sequence decay or the absence of such structures in the ancestral viruses. However, we do see sequences that resemble degraded hairpin structures to which Dependovirus Rep proteins bind, with an example from microbat integration mlEDLG-1 shown in Fig. Fig.3.3. The second most widespread endogenous sequences are related to the parvoviruses. They are found in 6 of 49 vertebrate species considered, with nearly complete genomes in rat, opossum, wallaby, and guinea pig (Fig. (Fig.22).Open in a separate windowFIG. 3.Hairpin structure of the inverted terminal repeat of adeno-associated virus 2 (left) and a candidate degraded hairpin structure located close to the 5′ end of the mlEDLG-1 integration in microbats (right). Structures and mountain plots were generated using default parameters of the RNAfold program (5), with nucleotide coloring representing base-pairing probabilities: blue is below average, green is average, and red is above average. Mountain plots represent hairpin structures based on minimum free energy (mfe) calculations and partition function (pf) calculations, as well as the centroid structure (5). Height is expressed in numbers of nucleotides; position represents nucleotide.The Dependovirus AAV2 has strong bias for integration into human chromosome 19 during infection, driven by a host sequence that is recognized by the viral Rep protein(s). Rep mediates the formation of a synapse between viral and cellular sequences, and the cellular sequences are nicked to serve as an origin of viral replication (14). The related integrations in mice and rats, located in the same chromosomal locations, might be explained by such a mechanism. However, the extent of endogenous sequence decay and the frequency of stop codons indicate that these integrations occurred some 30 to 35 million years ago, implying that they are derived from a single event in a rodent ancestor rather than two independent integration events at the same location. Similarly, integrations EDLG-1 in dog and panda lie in chromosomal regions that can be readily aligned (based on University of California—Santa Cruz [UCSC] genome assemblies) and show sequence decay consistent with the age of the common ancestor, about 42 million years. Endogenous sequences related to the family Parvoviridae can thus be traced to over 40 million years back in time, and viral proteins related to this family have remained over 40% conserved.Sequences related to circoviruses were detected in five vertebrate species (Table (Table11 and Table S1 in the supplemental material). At least one of these sequences, the endogenous sequence in opossum, likely represents a recent integration. Several integrations in dog, cat, and panda, on the other hand, appear to date from at least 42 million years ago, which is the last time when pandas and dogs shared a common ancestor. We see evidence for this age in data from sequence degradation (Table (Table1),1), phylogenetic analyses of endogenous Circovirus-like genomes (see Fig. S2 in the supplemental material), and genomic synteny where integration ECLG-3 is surrounded by genes MTA3 and ARID5A in both dog and panda and integration ECLG-2 lies 35 to 43 kb downstream of gene UPF3A. In fact, Circovirus integrations may even precede the split between dogs and cats, about 55 million years ago, although the preliminary assembly and short genomic contigs for cats make synteny analysis impossible.The most common Circovirus-related sequences detected in vertebrate genomes are derived from the rep gene. We speculate that, like those of the Parvoviridae, the ancestral Circoviridae sequences might have been copied using a primer sequence in the host DNA that resembled the viral origin and was therefore recognized by the virus Rep protein. Higher incidence of rep gene identifications may represent higher conservation of this gene with time, or alternatively, possession of these sequences may impart some selective advantage to the host species. The largest Circovirus-related integration detected, in the opossum, comprises a short fragment of what may have been the cap gene immediately adjacent to and in the opposite orientation from the rep gene. This organization is similar to that of the present day Circovirus genome in which these genes share a promoter in the hairpin regions but are translated in opposite directions (Fig. (Fig.22).In summary, our results indicate that sequences derived from ancestral members of the families Parvoviridae and Circoviridae were integrated into their host''s genomes over the past 50 million years of evolution. Features of their replication strategies suggest mechanisms by which such integrations may have occurred. It is possible that some of the endogenous viral sequences could offer a selective advantage to the virus or the host. We note that rep open reading frame-derived proteins from some members of these families kill tumor cells selectively (3, 12). The genomic “fossils” we have discovered provide a unique glimpse into virus evolution but can give us only a lower estimate of the actual ages of these families. However, numerous recent integrations suggest that their germ line transfer has been continuing into present times.   相似文献   

19.
Two amino acids, Leu149 and Val223, were identified as proteolytically sensitive when Pseudozyma antarctica lipase (PalB) was heterologously expressed in Escherichia coli. The functional expression was enhanced using the double mutant for cultivation. However, the recombinant protein production was still limited by PalB misfolding, which was resolved by DsbA coexpression.Though Escherichia coli retains popularity as a host for recombinant protein production, it has technical limitations for expressing eukaryotic proteins. Host/protein incompatibility is critical, since eukaryotic proteins heterologously expressed in E. coli are recognized as a foreign object that can induce heat shock responses and drive protease-mediated degradation. E. coli contains several proteases situated in various intracellular compartments, including the cytoplasm, periplasm, inner membrane, and outer membrane (7, 9). While these proteases play a pivotal role in maintaining cell physiology by clearing up abnormal proteins (4), they can potentially degrade recombinant proteins (6). Nevertheless, the mechanisms associated with protein degradation and protein substrate specificity are incompletely understood.Since the proteolysis of recombinant proteins is a common issue, approaches to overcome this limitation have been extensively explored. These include using protease-deficient mutants as an expression host (10, 18), lowering the cultivation temperature (3), coexpressing folding factors (19, 20), applying protein fusion technology (2, 8), eliminating protease cleavage sites (5, 15), and modifying target protein hydrophobicity (11). While genetic elimination of proteolytically sensitive sequences without impacting the target protein''s bioactivity appears to be technically attractive, there is no effective prediction or identification of potential protease cleavage sites. Proteolytically sensitive sequences might not be specific to certain proteases. Various factors affecting the folding, solubility, steric hindrance, and binding of the substrate protein, e.g., neighboring domains or amino acids, can contribute to the sensitivity of proteolytic sites, making sequence manipulation for proteolysis alleviation more difficult.Herein, the recombinant protein''s sensitivity toward intracellular proteolysis was genetically alleviated to enhance the functional expression of PalB, an industrial enzyme with justified applications (1), in E. coli and to understand the relationship between proteolytic specificity and substrate protein sequence. Since there are three intramolecular disulfide bonds whose formation can potentially affect PalB structure and bioactivity, PalB expression in the oxidative periplasm of E. coli was explored (20, 22). However, the expressed PalB protein was extremely sensitive to intracellular proteolysis, resulting in complete degradation. To overcome this problem, PalB mutant variants that were stable against proteolysis were derived (Table (Table11).

TABLE 1.

Strains, plasmids, and oligonucleotides
Strain, plasmid, or oligonucleotideRelevant genotype or DNA sequencea
Strain
    BL21(DE3)FompT dcm lon hsdSB (rB mB) gal λ(DE3[lacIind1sam7 nin5 lacUV5-T7 gene 1])
Plasmids
    pARDsbA (20)ParaB::dsbA, Ori (pACYC184), Cmr
    pARDsbC (20)ParaB::dsbA, Ori (pACYC184), Cmr
    pETG (20)PT7-pelB::palB, Ori (pBR322), Apr
    pETGM-2pETG derivative containing two mutations of Leu149Val and Val223Ile in palB (by ep-PCR with pETG and LB-P10/LB-P11)
    pETGM-3pETG derivative containing mutation of Leu149Gly in palB (by SDMb with pETG and P104/P114)
    pETGM-4pETG derivative containing mutation of Leu149Val in palB (by SDM with pETG and P103/P113)
    pETGM-5pETG derivative containing mutation of Leu149Ala in palB (by SDM with pETG and P105/P115)
    pETGM-6pETG derivative containing mutation of Leu149Met in palB (by SDM with pETG and P106/P116)
    pETGM-7pETG derivative containing mutation of Leu149Ile in palB (by SDM with pETG and P107/P117)
    pETGM-8pETG derivative containing mutation of Val223Phe in palB (by SDM with pETG and P112/P122)
    pETGM-9pETG derivative containing mutation of Val223Met in palB (by SDM with pETG and P111/P121)
    pETGM-10pETG derivative containing mutation of Val223Ala in palB (by SDM with pETG and P110/P120)
    pETGM-11pETG derivative containing mutation of Val223Gly in palB (by SDM with pETG and P109/P119)
    pETGM-12pETG derivative containing mutation of Val223Ile in palB (by SDM with pETGM-2 and P136/P137, i.e., by reverting the mutation of Val149 to Leu in pETGM-2)
Oligonucleotides
    LB-P105′-GGCCATGGGTCTACCTTCCGGTTCGG-3′
    LB-P115′-CTGAATTCTCAGGGGGTGACGATGCCGGAGCAGG-3′
    P1035′-CCTCTCGATGCAGTCGCGGTTAGTGCACCC-3′
    P1135′-GGGTGCACTAACCGCGACTGCATCGAGAGG-3′
    P1045′-CCTCTCGATGCAGGCGCCGTTAGTGCACCC-3′
    P1145′-GGGTGCACTAACGGCGCCTGCATCGAGAGG-3′
    P1055′-CCTCTCGATGCAGCCGCGGTTAGTGCACCC-3′
    P1155′-GGGTGCACTAACCGCGGCTGCATCGAGAGG-3′
    P1065′-CCTCTCGATGCCATGGCGGTTAGTGCACCC-3′
    P1165′-GGGTGCACTAACCGCCATGGCATCGAGAGG-3′
    P1075′-CCTCTCGATGCGATCGCGGTTAGTGCACCC-3′
    P1175′-GGGTGCACTAACCGCGATCGCATCGAGAGG-3′
    P1095′-GGGCCGCTGTTCGGGATCGACCATGCAGGC-3′
    P1195′-GCCTGCATGGTCGATCCCGAACAGCGGCCC-3′
    P1105′-GGGCCGCTGTTCGCGATCGACCATGCAGGC-3′
    P1205′-GCCTGCATGGTCGATCGCGAACAGCGGCCC-3′
    P1115′-GGGCCGCTGTTCATGATCGACCATGCAGGC-3′
    P1215′-GCCTGCATGGTCGATCATGAACAGCGGCCC-3′
    P1125′-GGGCCGCTGTTCTTTATCGACCATGCAGGC-3′
    P1225′-GCCTGCATGGTCGATAAAGAACAGCGGCCC-3′
    P1365′-CCTCTCGATGCACTGGCGGTTAGTGCACCC-3′
    P1375′-GGGTGCACTAACCGCCAGTGCATCGAGAGG-3′
Open in a separate windowaFor oligonucleotides, italic indicates mutated nucleotides, underlining indicates restriction site (silent mutation), and bold indicates a mutation codon.bSDM, site-directed mutagenesis.The results for functional expression of various PalB variants are summarized in Table Table22 and Fig. Fig.1.1. Neither PalB activity nor PalB-related polypeptide was detectable for the BL21(DE3)(pETG) culture sample, suggesting that the expressed gene product was degraded. Error-prone PCR (ep-PCR) was conducted using pETG as the template and LB-P10/LB-P11 as the primers. Potential mutants were screened on tributyrin agar plates. A colony developed a large halo, and the harboring plasmid (i.e., pETGM-2) was purified for sequencing. Two mutations, Leu149Val and Val223Ile, were identified in the expressed PalB mutant (i.e., M-2). High PalB activity was obtained for the BL21(DE3) (pETGM-2) culture sample. However, a PalB-related polypeptide was detected in both the soluble and insoluble fractions, implying potential protein misfolding. It was previously reported that the functional expression of PalB could be limited by disulfide bond formation associated with folding (21). It was intriguing to observe the structural effect of these two simple mutations on improving PalB stabilization and functional expression.Open in a separate windowFIG. 1.Western blotting analysis of the culture samples for the expression of various PalB mutants. Both soluble and insoluble fractions are shown. The number (n = 2 to 12) represents the PalB mutant (M-n) with the use of BL21(DE3) harboring the expression plasmid pETGM-n, summarized in Tables Tables11 and and2.2. “G” represents the control experiment using BL21(DE3) harboring pETG. “C” and “I” represent the cultures without and with IPTG induction, respectively.

TABLE 2.

Cultivation performance for production of PalB mutant variantsa
PlasmidCell density (OD600)b
Activity (U/liter/OD600)cMutation in PalB
w/o IPTGw/IPTG
pETG3.1 ± 0.03.3 ± 0.1NDNone
pETGM-23.4 ± 0.13.3 ± 0.0155Leu149Val, Val223Ile
pETGM-33.2 ± 0.13.3 ± 0.0NDLeu149Gly
pETGM-43.4 ± 0.13.6 ± 0.160Leu149Val
pETGM-53.3 ± 0.13.5 ± 0.1NDLeu149Ala
pETGM-62.7 ± 0.13.0 ± 0.1NDLeu149Met
pETGM-73.5 ± 0.03.2 ± 0.038Leu149Ile
pETGM-83.0 ± 0.03.6 ± 0.050Val223Phe
pETGM-92.9 ± 0.12.8 ± 0.0NDVal223Met
pETGM-103.4 ± 0.13.3 ± 0.1NDVal223Ala
pETGM-113.5 ± 0.03.5 ± 0.130Val223Gly
pETGM-123.6 ± 0.13.5 ± 0.157Val223Ile
Open in a separate windowaBL21(DE3) was used as an expression host to harbor various plasmids. PalB enzyme assay was conducted at 37°C and pH 8.0 using a pH stat with olive oil as a substrate. One unit of enzyme activity is defined as the amount of enzyme required to liberate one μmole of fatty acid per min.bOD600, optical density at 600 nm; w/o and w/IPTG, without and with isopropyl-β-d-thiogalactopyranoside, respectively.cThese activities represent IPTG-induced cultures. No activity was detected for all control cultures without IPTG induction. ND, not detected.Site-directed mutagenesis was conducted to introduce various single mutations of Leu149 and Val223 for a further understanding of the structural effect. Note that only hydrophobic amino acid residues were selected for the replacement to avoid any major structural disturbances. Among the five single mutations on Leu149, only Leu149Val and Leu149Ile were associated with a slight improvement in functional expression, and Leu149Val was more effective than Leu149Ile, whereas the others behaved similarly to the wild type. Among the five single mutations at Val223, only Val223Phe, Val223Gly, and Val223Ile showed a slight improvement in functional expression, with Val223Ile as the most effective one, whereas the others showed no improvement. The results suggest that Leu149 and Val223 are two amino acids critically affecting the proteolytic susceptibility of PalB in E. coli, with a synergistic effect on PalB stabilization from Leu149Val and Val223Ile mutations.Several amino acids were identified as being involved in the catalytic mechanism (17), such as Ser107-Asp189-His226, forming the catalytic triad, and Thr42 and Gly108, acting as the hydrogen bond donors in the oxyanion hole of the active site. However, Leu149 is located at the entrance of the partially formed lid of the protein structure, whereas Val223 is situated near one of the catalytic triad residues, His226, in the active site. The substitutions at Leu149 and/or Val223 with a PalB stabilization effect potentially convert the overall protein conformation into a less proteolytically sensitive form without changing the enzyme kinetics or activity. It was interesting to observe a similarly positive effect upon replacing Val223 with either a bulky Phe or a small Gly, implying that the molecular size of this amino acid residue was not critical for alleviation of the proteolytic sensitivity. Note that PalB-related polypeptide was present in the insoluble fraction for all the PalB mutants with improved functional expression, suggesting another expression hurdle of protein misfolding.To overcome the protein misfolding, coexpression of two periplasmic folding factors, DsbA and DsbC, was explored, and the cultivation results are summarized in Fig. Fig.2.2. DsbA coexpression significantly boosted functional expression, with specific PalB activity of more than twofold that of the control culture and a slight reduction of insoluble PalB. However, the improving effect associated with DsbC coexpression was minimal. The results suggest that the initiation but not isomerization for disulfide bond formation in the periplasm became limiting. The chaperone activity associated with DsbA could assist the folding of the expressed PalB double mutant.Open in a separate windowFIG. 2.Effect of DsbA or DsbC coexpression on the functional expression of the PalB double mutant. Cultivations were performed in a bioreactor containing 1 liter LB medium and operated at pH 7.0, 28°C, and 650 rpm. The cultures were induced with 0.1 mM isopropyl-β-d-thiogalactopyranoside and/or 0.2 g/liter arabinose at a ∼0.5 optical density at 600 nm (OD600) cell density. (A) Time profiles of cell density; (B) time profiles of the specific PalB activity [note that there was no detectable activity for BL21(DE3) (pETG) culture samples]; (C) Western blotting analysis of the final samples of the four cultures. Both soluble and insoluble fractions are shown. Lanes 1 and 5, BL21(DE3) (pETG); lanes 2 and 6, BL21(DE3) (pETGM-2); lane 3 and 7, BL21(DE3) (pETGM-2, pARDsbC); lanes 4 and 8, BL21(DE3) (pETGM-2, pARDsbA).While many studies based on rational mutagenesis, directed evolution, and gene shuffling are conducted to derive PalB mutants with improved enzyme properties, such as thermostability, bioactivity, or enantioselectivity (12-14, 16), this is an approach to improve protein manufacturing by deriving the variants with less proteolytic susceptibility. Though it has been perceived that the proteolytic specificity can be intrinsically determined by the protein substrate sequence, experimental demonstration of this is rarely reported. This study complements the lack of such experimental demonstration by showing the proteolytic specificity and sensitivity of PalB heterologously expressed in E. coli can be drastically altered by simple amino acid substitutions. It also demonstrates the application of molecular manipulation to enhance recombinant protein production in E. coli.  相似文献   

20.
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