Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 31
Filter
1.
Zhonghua Wai Ke Za Zhi ; 56(9): 665-669, 2018 Sep 01.
Article in Chinese | MEDLINE | ID: mdl-30157572

ABSTRACT

Many factors contribute to a successful total knee arthroplasty, and postoperative coronal lower limb alignment has always been a focus of joint surgeons. Previous researches have suggested that neutral alignment can bring higher prosthesis survival rate and better knee function. However, the theory has been challenged in recent years.In this article, the author introduces the axis, alignment and osteotomy of total knee arthroplasty briefly and reviews the studies on the neutral alignment and kinematic alignment of recent years in order to provide some advice for the clinical operation.


Subject(s)
Arthroplasty, Replacement, Knee , Knee Joint , Knee Prosthesis , Humans , Knee Joint/surgery , Leg/anatomy & histology , Leg/surgery , Osteoarthritis, Knee/surgery
2.
Cell Death Discov ; 2: 16013, 2016.
Article in English | MEDLINE | ID: mdl-27551507

ABSTRACT

To better understand the molecular mechanisms responsible for light-induced damage in retinal pigmented epithelial (RPE) cells, we developed an automated device to recapitulate intense light exposure. When compared with human fibroblasts, ARPE-19 cells that had been exposed to blue-rich light-emitting diode-light of 10 000 Lux at 37 °C for 9 h displayed dramatic cellular apoptosis. Collectively, gene expression profiling and qPCR demonstrated that growth arrest and DNA damage-45α (GADD45α) expression was markedly upregulated. Transient knockdown of GADD45α partially attenuated light-damage-induced apoptosis in ARPE-19 cells, whereas GADD45α overexpression dramatically increased it. These results demonstrate the critical function of GADD45α in light-induced RPE cellular apoptosis. Quantitative reverse transcription-PCR and western blotting revealed that the upregulation of GADD45α was under direct control of p53. Moreover, treatment with Ly294002, an inhibitor of AKT phosphorylation, further promoted GADD45α gene transcription in both non-light and light-damaged ARPE-19 cells. Treatment also exacerbated RPE cellular apoptosis after light exposure, confirming that inhibition of Akt phosphorylation increases GADD45α expression. Collectively, our findings reveal that light irrigation induces human RPE cellular apoptosis through upregulation of GADD45α expression mediated through both the p53 and phosphatidylinositol 3-kinase-AKT signaling pathways. These results provide new insights into human retinal diseases elicited by light damage and open a new avenue for disease prevention and treatment.

3.
Int J Syst Evol Microbiol ; 66(11): 4766-4771, 2016 Nov.
Article in English | MEDLINE | ID: mdl-27530392

ABSTRACT

A Gram-stain-negative, nutritionally fastidious bacterium (PLS229T) causing pear leaf scorch was identified in Taiwan and previously grouped into Xylella fastidiosa. Yet, significant variations between PLS229T and Xylellafastidiosa were noted. In this study, PLS229T was evaluated phenotypically and genotypically against representative strains of Xylellafastidiosa, including strains of the currently known subspecies of Xylellafastidiosa, Xylella fastidiosa subsp. multiplex and 'Xylella fastidiosasubsp.pauca'. Because of the difficulty of in vitro culture characterization, emphases were made to utilize the available whole-genome sequence information. The average nucleotide identity (ANI) values, an alternative for DNA-DNA hybridization relatedness, between PLS229T and Xylellafastidiosa were 83.4-83.9 %, significantly lower than the bacterial species threshold of 95 %. In contrast, sequence similarity of 16S rRNA genes was greater than 98 %, higher than the 97 % threshold to justify if two bacterial strains belong to different species. The uniqueness of PLS229T was also evident by observing only about 87 % similarity in the sequence of the 16S-23S internal transcribed spacer (ITS) between PLS229T and strains of Xylellafastidiosa, discovering significant single nucleotide polymorphisms at 18 randomly selected housekeeping gene loci, observing a distinct fatty acid profile for PLS229T compared with Xylellafastidiosa, and PLS229T having different observable phenotypes, such as different susceptibility to antibiotics. A phylogenetic tree derived from 16S rRNA gene sequences showed a distinct PLS229T phyletic lineage positioning it between Xylellafastidiosa and members of the genus Xanthomonas. On the basis of these data, a novel species, Xylella taiwanensis sp. nov. is proposed. The type strain is PLS229T (=BCRC 80915T=JCM 31187T).


Subject(s)
Phylogeny , Plant Diseases/microbiology , Pyrus/microbiology , Xylella/classification , Bacterial Typing Techniques , Base Composition , DNA, Bacterial/genetics , Fatty Acids/chemistry , Nucleic Acid Hybridization , Plant Leaves/microbiology , RNA, Ribosomal, 16S/genetics , Sequence Analysis, DNA , Taiwan , Xylella/genetics , Xylella/isolation & purification
4.
Genome Announc ; 2(2)2014 Mar 20.
Article in English | MEDLINE | ID: mdl-24652975

ABSTRACT

The draft genome sequence of Xylella fastidiosa pear leaf scorch strain PLS229, isolated from the pear cultivar Hengshan (Pyrus pyrifolia) in Taiwan, is reported here. The bacterium has a genome size of 2,733,013 bp, with a G+C content of 53.1%. The PLS229 genome was annotated and has 3,259 open reading frames and 50 RNA genes.

5.
Plant Dis ; 98(7): 990, 2014 Jul.
Article in English | MEDLINE | ID: mdl-30708865

ABSTRACT

Sesame (Sesamum indicum L.), an annual plant, is grown as an oilseed crop and the seeds are used in bakery products in Taiwan. In June 2013, plants exhibiting symptoms including phyllody and abnormal stem curling were observed in sesame fields in Pitou Township, Changhua County, Taiwan. Incidence of infected plants was estimated to be greater than 90% within a single field. Phytoplasmas associated with sesame exhibiting phyllody, witches'-broom, or virescence have been classified as strains of 16SrI-B in Myanmar (GenBank Accession No. AB558132), 16SrII-A in Thailand (JN006075), 16SrII-D in Oman (EU072505) and India (KF429486), 16SrIV-C in Iran (JF508515), and 16SrVI-A (KF156894) and 16SrIX (KC139791) in Turkey (1). Three symptomatic and four asymptomatic plants were uprooted and transplanted in a greenhouse for further study. Transmission electron microscopy (TEM) revealed clusters of phytoplasma cells ranging from 300 to 800 nm in diameter only in phloem sieve elements of stems of three symptomatic and two asymptomatic plants. Comparable tissues from two other symptomless plants were devoid of phytoplasma cells. Total DNA was extracted with a modified CTAB method (2) from plant tissues (100 mg each) including stem, leaf, petiole, and root from the same plants used for TEM work. Analyses by a nested PCR using universal primer pairs P1/P7 (5'-AAGAGTTTGATCCTGGCTCAGGATT/5'-CGTCCTTCATCGGCTCTT) followed by R16F2n/R16R2 (5'-GAAACGACTGCTAAGACTGG/5'-TGACGGGCGGTGTGTACAAACCCCG) were performed to detect putative phytoplasma DNA (3). Each primer pair amplified a single PCR product of either 1.8 or 1.2 kb, respectively, only from the three symptomatic and two asymptomatic plant tissues that had phytoplasma cells in their sieve elements. It is likely that these two asymptomatic plants were in the early stage of infection before symptoms became noticeable. The nested PCR products (1.2 kb) amplified from the symptomatic plants were cloned separately and sequenced (GenBank Accession Nos. KF923391, KF923392, and KF923393). BLAST analysis of the sequences revealed that they shared 99.2% sequence identity with strains reported from India and Thailand (KF429486 and JN006075), which were classified to the 16SrII-D and 16SrII-A subgroups, respectively. Moreover, iPhyClassifier software (4) was used to perform sequence comparison and generate a virtual restriction fragment length polymorphism (RFLP) profile. The 16S rDNA sequences shared 99.4% identity with that of the 'Candidatus Phytoplasma australasiae' (Y10097) and the RFLP patterns were identical to that of the 16SrII-A subgroup, indicating the Taiwanese strain is a 'Ca. P. australasiae'-related strain. To our knowledge, this is the first report of a 16SrII-A subgroup phytoplasma causing phyllody and abnormal stem curling on sesame in Taiwan. The occurrence of phytoplasma on sesame could have direct implications for the cultivation of this economically important oilseed plant and the bakery industry in Taiwan. References: (1) M. Catal et al. Plant Dis. 97:835, 2013. (2) T. M. Fulton et al. Plant Mol. Biol. Rep. 13:207, 1995. (3) D. E. Gundersen and I. M. Lee. Phytopathol. Mediterr. 35:144, 1996. (4) Y. Zhao et al. Int. J. Syst. Evol. Microbiol. 59:2582, 2009.

6.
Plant Dis ; 98(7): 991, 2014 Jul.
Article in English | MEDLINE | ID: mdl-30708870

ABSTRACT

Roselle (Hibiscus sabdariffa L.), an annual plant with acidic taste, has been used for making juice, jelly, and other baking additives in Taiwan. In September 2013, symptoms including phyllody and wrinkled leaves were observed on roselle plants in a field in Tantsu Township, Taichung County, Taiwan. Incidence of the infected plants was estimated to be greater than 80% within a single field. A phytoplasma was recently reported as the causal agent of roselle phyllody and reddening of leaves in India and classified as a group 16SrV-D strain (1). Samples including stems, flowers, and leaves were collected from four symptomatic and one asymptomatic roselle plants from the field. Transmission electron microscopy revealed clusters of phytoplasma cells ranging from 400 to 750 nm in diameter only in phloem sieve elements of petioles and stems of symptomatic plants. These cells were not observed in asymptomatic plants. Total DNA was extracted from plant tissues (100 mg each) including stems, petioles, and mid veins of leaves by a modified CTAB method (2). Analyses by a nested PCR assay using universal primer pairs P1/P7 followed by R16F2n/R16R2 were performed to detect putative phytoplasma (1). Each primer pair amplified a single PCR product 1.8 kb and 1.2 kb long, respectively, only from tissues of the four symptomatic plants. The nested PCR products (1.2 kb) amplified from three independent symptomatic plants were cloned separately and sequenced by automatic DNA sequencing method with ABI3730 DNA Analyzer (Applied Biosystems) at the Biotechnology Center, National Chung Hsing University, Taichung, Taiwan (GenBank Accession Nos. KF923397, KF923398, and KF923399). BLAST analysis of the sequences revealed that they shared 99.8% sequence identity with those of 16SrI group phytoplasma strains, e.g., garlic yellows phytoplasma, torenia yellows phytoplasma, and periwinkle leaf yellowing phytoplasma (AB750363, FJ437568, and GU361754). Moreover, i PhyClassifier software (3) was used to perform sequence comparison and generate a virtual restriction fragment length polymorphism (RFLP) profile for the sequences derived from the symptomatic roselle samples. The 16S rDNA sequences shared 99.6% identity with those of the 'Candidatus Phytoplasma asteris' reference strain (M30790) and the RFLP patterns were identical to that of the 16SrI group. However, this strain may represent a new subgroup because the shared similarity coefficient was only 0.94, which is within the values set for a new subgroup (3). Taken together, these results indicate the phytoplasma infecting roselle in Taiwan is a 'Ca. P. asteris'-related strain belonging to the 16SrI group. To our knowledge, this is the first report of a 16SrI group phytoplasma causing wrinkled leaves and phyllody on roselle in Taiwan. The occurrence of phytoplasma on roselle could have direct implication for the bakery and juice industries in Taiwan. References: (1) C. Biswas et al. Phytoparasitica 41:539, 2013. (2) I. Echevarría-Machado et al. Mol. Biotechnol. 31:129, 2005. (3) W. Wei et al. Int. J. Syst. Evol. Microbiol. 57:1855, 2007.

7.
Plant Dis ; 96(4): 582, 2012 Apr.
Article in English | MEDLINE | ID: mdl-30727409

ABSTRACT

Purple coneflower (Echinacea purpurea), widely grown as an ornamental and medicinal plant, is a perennial flowering plant that is native to eastern North America. In July 2011, symptoms indicative of phytoplasma disease, including floral virescence, phyllody, and witches'-broom (WB), were observed to be affecting plants in coneflower fields in Wufeng, Taichung City, Taiwan. Incidence of infected plants was estimated to be greater than 90% within a single field. Phytoplasmas previously associated with purple coneflower WB disease have all been classified as aster yellows group (16SrI) strains (GenBank Accession Nos. EU333395, AY394856, EU416172, and EF546778) except for pale purple coneflower (Echinacea pallida) WB in Australia, which was identified as a subgroup 16SrII-D member (2). Three diseased plants were uprooted and transplanted in a greenhouse for further study. Transmission electron microscopy revealed clusters of phytoplasma cells ranging from 170 to 490 nm in diameter in phloem sieve elements of virescent and phylloid flowers and stems from diseased plants. Comparable tissues from symptomless plants were devoid of phytoplasma. Total DNA was extracted from plant tissue samples (50 to 100 mg each) including stems, leaves, and flowers by a modified CTAB method (1) from three symptomatic plants as well as from three asymptomatic coneflower plants seedlings. Analyses by a nested PCR using universal primer pairs P1/P7 followed by R16F2n/R16R2 were performed to detect putative phytoplasma (2). Each primer pair amplified a single PCR product of either 1.8 or 1.2 kb, respectively, from diseased plant tissues only. The nested PCR products (1.2 kb) amplified from phylloid flowers of the three diseased plants were cloned separately and sequenced (GenBank Accession Nos. JN885460, JN885461, and JN885462). Blast analysis of the sequences revealed a 99.7 to 99.8% sequence identity with those of Echinacea WB phytoplasma strain EWB5 and EWB6 (GenBank Accession Nos. JF340076 and JF340080), which reportedly belonged to the 16SrII-D subgroup (2). Moreover, iPhyClassifier software (3) was used to perform sequence comparison and generate the virtual restriction fragment length polymorphism (RFLP) profile. The 16S rDNA sequences share a 99.4 to 99.5% similarity with that of the 'Candidatus Phytoplasma australasiae' reference strain (Y10097) and the RFLP patterns are identical to that of the 16SrII-A subgroup. Taken together, these results indicated that the phytoplasma infecting purple coneflower in Taiwan is a 'Ca. Phytoplasma australasiae'-related strain and belongs to the 16SrII-A subgroup. To our knowledge, this is the first report of a 16SrII-A subgroup phytoplasma causing WB disease on purple coneflower in Taiwan. The occurrence of phytoplasma on purple coneflower could have direct implication for the economically important ornamental, medicinal plant, and floral industry in Taiwan, especially to the growers and breeders that eagerly promote the purple coneflower industry. References: (1) T. M. Fulton et al. Plant Mol. Biol. Rep. 13:207, 1995. (2) T. L. Pearce et al. Plant Dis. 95:773, 2011. (3) Y. Zhao et al. Int. J. Syst. Evol. Microbiol. 59:2582, 2009.

8.
Plant Dis ; 94(8): 1065, 2010 Aug.
Article in English | MEDLINE | ID: mdl-30743454

ABSTRACT

In November 2008, betelvines (Piper betle L., Piperaceae) exhibiting leaf blight symptoms were observed in central Taiwan. Infections resulted in a 30 to 70% loss of leaf yield in the investigated betel leaf-producing facilities. Symptoms began with small, necrotic, water-soaked spots that progressed to circular to irregularly shaped brown lesions, 5 to 10 mm in diameter, with chlorotic halos on leaves; some lesions started from the edge of leaves and later fused to form dried, necrotic margins. Bacteria-like streaming fluid was visible from the edges of freshly cut lesions at the junctions of chlorotic and necrotic leaf tissues when observed with a light microscope at ×100. When the streaming fluid was streaked onto King's medium B (3), a slow-growing, gram-negative, nonfluorescent bacterium was identified from the whitish colonies that consistently developed on the medium. Five bacterial isolates from three lesions were characterized with fatty acid methyl ester analysis (Agilent Technologies, Santa Clara, CA) and Sherlock Microbial Identification System (Microbial IDentification Inc., Newark, DE), and for each isolate, the bacterium was confirmed as Acidovorax avenae subsp. citrulli with a similarity index >0.70. In addition, the Biolog system (Biolog, Hayward, CA) and 16S ribosomal RNA sequence identity comparison were performed to confirm that the five betelvine-isolated bacteria were A. avenae subsp. citrulli based on a similarity of 0.54 with Biolog and 99% sequence identity for 16S rRNA gene. Koch's postulates were fulfilled by infiltrating a bacterial suspension of 3 × 105 CFU/ml into 40 leaves of four greenhouse-grown, disease-free, mature betelvine plants. After inoculation, plants were kept in a humidified greenhouse at 28°C to favor symptom development and symptoms similar to those observed in the greenhouse were evident at 7 days post inoculation (dpi) on all bacterium-infiltrated leaves. Control leaves infiltrated with distilled water remained symptomless. Bacteria showing morphological and biochemical similarities (2) to the ones used for inoculation were isolated from all of the inoculated betelvine leaves. In addition, a bacterial suspension at 3 × 108 CFU/ml was sprayed at the amount of 5 ml per plant onto 6 to 10 plants each of 4-week-old disease-free seedlings of watermelon (Citrullus lanatus (Thunb.) Matsum & Nakai, cv. Empire No. 2), oriental sweet melon (Cucumis melo L. var. saccharinus Naudin, cv. Silver Beam), and waxgourd (Benincasa hispida (Thunb.) Cogn., cv. Cheerer) for bioassays, and the inoculated seedlings were enclosed in plastic bags for 36 h at 28°C. Water-soaked lesions were observed on leaves of watermelon and waxgourd at 2 dpi and on sweet melon at 4 dpi on all inoculated plants but not on distilled water-sprayed control plants, indicating that A. avenae subsp. citrulli strains from betelvine could also infect melon plants. A. avenae subsp. citrulli was previously identified as the causal agent of bacterial fruit blotch on melon and bitter gourd in Taiwan (1). To our knowledge, this is the first report that A. avenae subsp. citrulli can naturally infect betelvine, a noncucurbit crop, to elicit bacterial leaf blight disease. References: (1) A.-H. Cheng and T.-C. Huang. Plant Pathol. Bull. 7:216, 1998. (2) J. B. Jones et al. Page 121 in: Laboratory Guide for Identification of Plant Pathogenic Bacteria. 3rd ed. The American Phytopathological Society, St. Paul, MN, 2001. (3) E. O. King et al. J. Lab. Clin. Med. 44:301, 1954.

9.
Mol Plant Microbe Interact ; 15(10): 1014-24, 2002 Oct.
Article in English | MEDLINE | ID: mdl-12437299

ABSTRACT

The ability of Pseudomonas syringae to grow in planta is thought to be dependent upon the Hrp (type III secretion) system and multiple effector proteins that this system injects into plant cells. ORF5 in the conserved effector locus of the P. syringae pv. tomato DC3000 Hrp pathogenicity island was shown to encode a Hrp-secreted protein and to have a similarly secreted homolog encoded in an effector-rich pathogenicity island located elsewhere in the genome. These putative effector genes were designated hopPtoA1 and hopPtoA2, respectively. DNA gel blot analysis revealed that sequences hybridizing with hopPtoA1 were widespread among P. syringae pathovars, and some strains, like DC3000, appear to have two copies of the gene. uidA transcriptional fusions revealed that expression of hopPtoA1 and hopPtoA2 can be activated by the HrpL alternative sigma factor. hopPtoA1 and hopPtoA1/hopPtoA2 double mutants were not obviously different from wild-type P. syringae pv. tomato DC3000 in their ability to produce symptoms or to increase their total population size in host tomato and Arabidopsis leaves. However, confocal laser-scanning microscopy of GFP (green fluorescent protein)-labeled bacteria in Arabidopsis leaves 2 days after inoculation revealed that the frequency of undeveloped individual colonies was higher in the hopPtoA1 mutant and even higher in the hopPtoA1/hopPtoA2 double mutant. These results suggest that hopPtoA1 and hopPtoA2 contribute redundantly to the formation of P. syringae pv. tomato DC3000 colonies in Arabidopsis leaves.


Subject(s)
Bacterial Proteins/genetics , DNA-Binding Proteins , Plants/microbiology , Pseudomonas/genetics , Sigma Factor , Amino Acid Sequence , Arabidopsis/microbiology , Bacterial Outer Membrane Proteins/genetics , Bacterial Outer Membrane Proteins/metabolism , Bacterial Proteins/metabolism , Chromosome Mapping , Green Fluorescent Proteins , Luminescent Proteins/genetics , Luminescent Proteins/metabolism , Solanum lycopersicum/microbiology , Microscopy, Confocal , Molecular Sequence Data , Mutation , Open Reading Frames/genetics , Open Reading Frames/physiology , Phenotype , Polysaccharide-Lyases/genetics , Polysaccharide-Lyases/metabolism , Pseudomonas/growth & development , Pseudomonas/pathogenicity , Sequence Homology, Amino Acid
10.
Proc Natl Acad Sci U S A ; 98(23): 13306-11, 2001 Nov 06.
Article in English | MEDLINE | ID: mdl-11687633

ABSTRACT

Early pathological manifestations of Crohn's disease (CD) include vascular disruption, T cell infiltration of nerve plexi, neuronal degeneration, and induction of T helper 1 cytokine responses. This study demonstrates that disruption of the enteric glial cell network in CD patients represents another early pathological feature that may be modeled after CD8(+) T cell-mediated autoimmune targeting of enteric glia in double transgenic mice. Mice expressing a viral neoself antigen in astrocytes and enteric glia were crossed with specific T cell receptor transgenic mice, resulting in apoptotic depletion of enteric glia to levels comparable in CD patients. Intestinal and mesenteric T cell infiltration, vasculitis, T helper 1 cytokine production, and fulminant bowel inflammation were characteristic hallmarks of disease progression. Immune-mediated damage to enteric glia therefore may participate in the initiation and/or the progression of human inflammatory bowel disease.


Subject(s)
Autoimmune Diseases/immunology , Crohn Disease/immunology , Enterocolitis/immunology , Neuroglia/immunology , Adolescent , Adult , Animals , Base Sequence , CD8-Positive T-Lymphocytes/immunology , Case-Control Studies , Central Nervous System/immunology , Central Nervous System/metabolism , Crohn Disease/etiology , DNA Primers , Enteric Nervous System/immunology , Enteric Nervous System/metabolism , Enterocolitis/etiology , Female , Glial Fibrillary Acidic Protein/genetics , Glial Fibrillary Acidic Protein/metabolism , Glial Fibrillary Acidic Protein/physiology , Hemagglutinin Glycoproteins, Influenza Virus/genetics , Hemagglutinin Glycoproteins, Influenza Virus/metabolism , Humans , Immunohistochemistry , Male , Mice , Mice, Transgenic , Middle Aged , Receptors, Antigen, T-Cell/genetics , Receptors, Antigen, T-Cell/immunology , Receptors, Antigen, T-Cell/physiology
11.
Proc Natl Acad Sci U S A ; 97(16): 8770-7, 2000 Aug 01.
Article in English | MEDLINE | ID: mdl-10922033

ABSTRACT

Pseudomonas syringae is a member of an important group of Gram-negative bacterial pathogens of plants and animals that depend on a type III secretion system to inject virulence effector proteins into host cells. In P. syringae, hrp/hrc genes encode the Hrp (type III secretion) system, and avirulence (avr) and Hrp-dependent outer protein (hop) genes encode effector proteins. The hrp/hrc genes of P. syringae pv syringae 61, P. syringae pv syringae B728a, and P. syringae pv tomato DC3000 are flanked by an exchangeable effector locus and a conserved effector locus in a tripartite mosaic Hrp pathogenicity island (Pai) that is linked to a tRNA(Leu) gene found also in Pseudomonas aeruginosa but without linkage to Hrp system genes. Cosmid pHIR11 carries a portion of the strain 61 Hrp pathogenicity island that is sufficient to direct Escherichia coli and Pseudomonas fluorescens to inject HopPsyA into tobacco cells, thereby eliciting a hypersensitive response normally triggered only by plant pathogens. Large deletions in strain DC3000 revealed that the conserved effector locus is essential for pathogenicity but the exchangeable effector locus has only a minor role in growth in tomato. P. syringae secretes HopPsyA and AvrPto in culture in a Hrp-dependent manner at pH and temperature conditions associated with pathogenesis. AvrPto is also secreted by Yersinia enterocolitica. The secretion of AvrPto depends on the first 15 codons, which are also sufficient to direct the secretion of an Npt reporter from Y. enterocolitica, indicating that a universal targeting signal is recognized by the type III secretion systems of both plant and animal pathogens.


Subject(s)
Bacterial Proteins/metabolism , Pseudomonas/metabolism , Bacterial Proteins/genetics , Mosaicism , Pseudomonas/genetics
12.
Proc Natl Acad Sci U S A ; 97(9): 4856-61, 2000 Apr 25.
Article in English | MEDLINE | ID: mdl-10781092

ABSTRACT

The plant pathogenic bacterium Pseudomonas syringae is divided into pathovars differing in host specificity, with P. syringae pv. syringae (Psy) and P. syringae pv. tomato (Pto) representing particularly divergent pathovars. P. syringae hrp/hrc genes encode a type III protein secretion system that appears to translocate Avr and Hop effector proteins into plant cells. DNA sequence analysis of the hrp/hrc regions in Psy 61, Psy B728a, and Pto DC3000 has revealed a Hrp pathogenicity island (Pai) with a tripartite mosaic structure. The hrp/hrc gene cluster is conserved in all three strains and is flanked by a unique exchangeable effector locus (EEL) and a conserved effector locus (CEL). The EELs begin 3 nt downstream of the stop codon of hrpK and end, after 2.5-7.3 kb of dissimilar intervening DNA with tRNA(Leu)-queA-tgt sequences that are also found in Pseudomonas aeruginosa but without linkage to any Hrp Pai sequences. The EELs encode diverse putative effectors, including HopPsyA (HrmA) in Psy 61 and proteins similar to AvrPphE and the AvrB/AvrC/AvrPphC and AvrBsT/AvrRxv/YopJ protein families in Psy B728a. The EELs also contain mobile genetic element sequences and have a G + C content significantly lower than the rest of the Hrp Pai or the P. syringae genome. The CEL carries at least seven ORFs that are conserved between Psy B728a and Pto DC3000. Deletion of the Pto DC3000 EEL slightly reduces bacterial growth in tomato, whereas deletion of a large portion of the CEL strongly reduces growth and abolishes pathogenicity in tomato.


Subject(s)
Bacterial Proteins/genetics , Multigene Family , Plants/microbiology , Pseudomonas/genetics , Pseudomonas/pathogenicity , Base Sequence , Chromosome Mapping , Conserved Sequence , Genes, Bacterial , Solanum lycopersicum/microbiology , Molecular Sequence Data , Open Reading Frames , Plasmids , RNA, Transfer, Leu/genetics , Virulence/genetics
13.
Proc Natl Acad Sci U S A ; 97(5): 2247-52, 2000 Feb 29.
Article in English | MEDLINE | ID: mdl-10681465

ABSTRACT

Bacterial surface appendages called pili often are associated with DNA and/or protein transfer between cells. The exact function of pili in the transfer process is not understood and is a matter of considerable debate. The Hrp pilus is assembled by the Hrp type III protein secretion system of Pseudomonas syringae pv. tomato (Pst) strain DC3000. In this study, we show that the hrpA gene, which encodes the major subunit of the Hrp pilus, is required for secretion of putative virulence proteins, such as HrpW and AvrPto. In addition, the hrpA gene is required for full expression of genes that encode regulatory, secretion, and effector proteins of the type III secretion system. hrpA-mediated gene regulation apparently is through effect on the mRNA level of two previously characterized regulatory genes, hrpR and hrpS. Ectopic expression of the hrpRS gene operon restored gene expression, but not protein secretion, in the hrpA mutant. Three single amino acid mutations at the HrpA carboxyl terminus were identified that affect the secretion or regulatory function of the HrpA protein. These results define an essential role of the Hrp pilus structural gene in protein secretion and coordinate regulation of the type III secretion system in Pst DC3000.


Subject(s)
Bacterial Proteins/genetics , DNA-Binding Proteins/genetics , Escherichia coli Proteins , Gene Expression Regulation, Bacterial , Pili, Sex , Polysaccharide-Lyases/genetics , Pseudomonas/genetics , RNA Helicases , Transcription Factors , Amino Acids , Bacterial Proteins/metabolism , Bacterial Proteins/physiology , DEAD-box RNA Helicases , Genes, Bacterial , Mutagenesis , Operon , Polysaccharide-Lyases/metabolism , RNA, Bacterial , Transcription, Genetic
14.
J Bacteriol ; 181(7): 2298-301, 1999 Apr.
Article in English | MEDLINE | ID: mdl-10094714

ABSTRACT

The complete hrp-hrc-hrmA cluster of Pseudomonas syringae pv. syringae 61 encodes 28 polypeptides. A saprophytic bacterium carrying this cluster is capable of secreting HrpZ-a harpin encoded by hrpZ-in an hrp-dependent manner, which suggests that this cluster contains sufficient components to assemble functional type III secretion machinery. Sequence data show that HrcJ and HrcC are putative outer membrane proteins, and nonpolar mutagenesis demonstrates they are all required for HrpZ secretion. In this study, we investigated the cellular localization of the HrcC and HrcJ proteins by Triton solubilization, sucrose-gradient isopycnic centrifugation, and immunogold labeling of the bacterial cell surface. Our results indicate that HrcC is indeed an outer membrane protein and that HrcJ is located between both membranes. Their membrane localization suggests that they might be involved in the formation of a supramolecular structure for protein secretion.


Subject(s)
Bacterial Outer Membrane Proteins/metabolism , Pseudomonas/metabolism , Pseudomonas/ultrastructure
15.
J Bacteriol ; 180(17): 4523-31, 1998 Sep.
Article in English | MEDLINE | ID: mdl-9721291

ABSTRACT

The species Pseudomonas syringae encompasses plant pathogens with differing host specificities and corresponding pathovar designations. P. syringae requires the Hrp (type III protein secretion) system, encoded by a 25-kb cluster of hrp and hrc genes, in order to elicit the hypersensitive response (HR) in nonhosts or to be pathogenic in hosts. DNA sequence analysis of the hrpC and hrpRS operons of P. syringae pv. syringae 61 (brown spot of beans), P. syringae pv. glycinea U1 (bacterial blight of soybeans), and P. syringae pv. tomato DC3000 (bacterial speck of tomatos) revealed that the 13 genes comprising the right half of the hrp cluster (including those in the previously sequenced hrpZ operon) are conserved and identically arranged. The hrpC operon is comprised of hrpF, hrpG, hrcC, hrpT, and hrpV. hrcC encodes a putative outer membrane protein that is conserved in all type III secretion systems. The other four genes appear to be characteristic of group I Hrp systems, such as those possessed by P. syringae and Erwinia amylovora. The predicted products of these four genes in P. syringae pv. syringae 61 are HrpF (8 kDa), HrpG (15.4 kDa), HrpT (7.5 kDa), and HrpV (13.4 kDa). HrpT is a putative outer membrane lipoprotein. HrpF, HrpG, and HrpV are all hydrophilic proteins lacking N-terminal signal peptides. The HrpG, HrcC, HrpT, and HrpV proteins of P. syringae pathovars syringae and tomato (the two most divergent pathovars) had at least 76% amino acid identity with each other, whereas the HrpF proteins of these two pathovars had only 36% amino acid identity. The HrpF proteins of P. syringae pathovars syringae and glycinea also showed significant similarity to the HrpA pilin protein of P. syringae pathovar tomato. Functionally nonpolar mutations were introduced into each of the genes in the hrpC operon of P. syringae pv. syringae 61 by insertion of an nptII cartridge lacking a transcription terminator. The mutants were assayed for their ability to elicit the HR in nonhost tobacco leaves or to multiply and cause disease in host bean leaves. Mutations in hrpF, hrcC, and hrpT abolished or greatly reduced the ability of P. syringae pv. syringae 61 to elicit the HR in tobacco. The hrpG mutant had only weakly reduced HR activity, and the activity of the hrpV mutant was indistinguishable from that of the wild type. Each of the mutations could be complemented, but surprisingly, the hrpV subclone caused a reduction in the HR elicitation ability of the DeltahrpV::nptII mutant. The hrpF and hrcC mutants caused no disease in beans, whereas the hrpG, hrpT, and hrpV mutants had reduced virulence. Similarly, the hrcC mutant grew little in beans, whereas the other mutants grew to intermediate levels in comparison with the wild type. These results indicate that HrpC and HrpF have essential functions in the Hrp system, that HrpG and HrpT contribute quantitatively but are not essential, and that HrpV is a candidate negative regulator of the Hrp system.


Subject(s)
Genes, Bacterial , Glycine max/microbiology , Mutation , Operon , Plant Proteins/genetics , Pseudomonas/genetics , Solanum lycopersicum/microbiology , Amino Acid Sequence , Base Sequence , Genetic Complementation Test , Molecular Sequence Data , Phenotype , Sequence Homology, Amino Acid
16.
J Bacteriol ; 180(17): 4532-7, 1998 Sep.
Article in English | MEDLINE | ID: mdl-9721292

ABSTRACT

Mutations in the five hrp and hrc genes in the hrpC operon of the phytopathogen Pseudomonas syringae pv. syringae 61 have different effects on bacterial interactions with host and nonhost plants. The hrcC gene within the hrpC operon encodes an outer membrane component of the Hrp secretion system that is conserved in all type III protein secretion systems and is required for most pathogenic phenotypes and for secretion of the HrpZ harpin to the bacterial milieu. The other four genes (in order), hrpF, hrpG, (hrcC), hrpT, and hrpV, appear to be unique to the group I hrp clusters found in certain phytopathogens (e.g., P. syringae and Erwinia amylovora) and are less well understood. We initiated an examination of their role in Hrp regulation and secretion by determining the effects of functionally nonpolar nptII cartridge insertions in each gene on the production and secretion of HrpZ, as determined by immunoblot analysis of cell fractions. P. syringae pv. syringae 61 hrpF, hrpG, and hrpT mutants were unable to secrete HrpZ, whereas the hrpV mutant overproduced and secreted the protein. This suggested that HrpV is a negative regulator of HrpZ production. Further immunoblot assays showed that the hrpV mutant produced higher levels of proteins encoded by all three of the major hrp operons tested-HrcJ (hrpZ operon), HrcC (hrpC operon), and HrcQB (hrpU operon)-and that constitutive expression of hrpV in trans abolished the production of each of these proteins. To determine the hierarchy of HrpV regulation in the P. syringae pv. syringae 61 positive regulatory cascade, which is composed of HrpRS (proteins homologous with sigma54-dependent promoter-enhancer-binding proteins) and HrpL (alternate sigma factor), we tested the ability of constitutively expressed hrpV to repress the activation of HrcJ production that normally accompanies constitutive expression of hrpL or hrpRS. No repression was observed, indicating that HrpV acts upstream of HrpRS in the cascade. The effect of HrpV levels on transcription of the hrpZ operon was determined by monitoring the levels of beta-glucuronidase produced by a hrpA'::uidA transcriptional fusion plasmid in different P. syringae pv. syringae 61 strains. The hrpV mutant produced higher levels of beta-glucuronidase than the wild type, a hrcU (type III secretion) mutant produced the same level as the wild type, and the strain constitutively expressing hrpV in trans produced low levels equivalent to that of a hrpS mutant. These results suggest that HrpF, HrpG, and HrpT are all components of the type III protein secretion system whereas HrpV is a negative regulator of transcription of the Hrp regulon.


Subject(s)
Gene Expression Regulation, Bacterial , Genes, Bacterial , Pseudomonas/genetics , Mutation , Operon , Phenotype , Transcription, Genetic
17.
J Bacteriol ; 179(5): 1497-504, 1997 Mar.
Article in English | MEDLINE | ID: mdl-9045805

ABSTRACT

By using transposon insertional mutagenesis and deletion analyses, a recombinant clone containing the region upstream of the acoABCD operon of Klebsiella pneumoniae was found to be required for acetoin-inducible expression of the operon in Escherichia coli. The nucleotide sequence of the region was determined, and it displayed an open reading frame of 2,763 bp that is transcribed divergently to the acoABCD operon. This gene, designated acoK, is capable of encoding a protein with an overall 58.4% amino acid identity with MalT, the transcriptional activator of the E. coli maltose regulon. A conserved sequence for nucleotide binding at the N-terminal region, as well as a helix-turn-helix motif belonging to the LuxR family of transcriptional regulators at the C terminus, was also identified. Primer extension analysis identified two transcription initiation sites, S1 and S2, located 319 and 267 bp, respectively, upstream of the putative start codon of acoK. Several copies of NtrC recognition sequence [CAC-(N11 to N18)-GTG] were found in the promoter regions of both the acoK gene and the acoABCD operon. Acetoin-dependent expression of the acoABCD operon could be restored in the E. coli acoK mutants by supplying a plasmid carrying an intact acoK, suggesting a transactivating function of the gene product. The AcoK protein overproduced in E. coli was approximately 100 kDa, which is in good agreement with the molecular mass deduced from the nucleotide sequence. A specific DNA binding property and an ATPase activity of the purified AcoK were also demonstrated.


Subject(s)
Acetoin Dehydrogenase/genetics , Bacterial Proteins , DNA-Binding Proteins , Genes, Regulator , Klebsiella pneumoniae/genetics , Operon , Trans-Activators/genetics , Acetoin/metabolism , Acetoin/pharmacology , Acetoin Dehydrogenase/biosynthesis , Adenosine Triphosphatases/metabolism , Adenosine Triphosphate/metabolism , Amino Acid Sequence , Base Sequence , DNA, Bacterial/metabolism , Escherichia coli/genetics , Gene Expression Regulation, Bacterial , Genes, Bacterial , Helix-Turn-Helix Motifs , Klebsiella pneumoniae/enzymology , Molecular Sequence Data , Mutagenesis, Insertional , RNA, Bacterial/genetics , RNA, Messenger/genetics , Trans-Activators/chemistry , Trans-Activators/metabolism
18.
J Biochem ; 119(6): 1118-23, 1996 Jun.
Article in English | MEDLINE | ID: mdl-8827447

ABSTRACT

The acoD gene, which encodes a dihydrolipoamide dehydrogenase component of the acetoin dehydrogenase enzyme system of Klebsiella pneumoniae was isolated and the nucleotide sequence determined. The gene is capable of encoding a protein of 465 amino acid residues with conserved binding domains for NAD and FAD, and two redox-active cysteine residues. The acoD gene product exhibited a Michaelis constant of 170 microM for NAD, while NADP can not be used as a substrate. The purified enzyme appeared to be a dimer of the acoD gene product. It did not associate tightly with the E1 and E2 components of either acetoin dehydrogenase or 2-oxoglutarate dehydrogenase to form an active multi-enzyme complex.


Subject(s)
Dihydrolipoamide Dehydrogenase/genetics , Genes, Bacterial/genetics , Klebsiella pneumoniae/genetics , Acetoin Dehydrogenase/metabolism , Amino Acid Sequence , Base Sequence , Chromatography, Gel , Dihydrolipoamide Dehydrogenase/isolation & purification , Dihydrolipoamide Dehydrogenase/metabolism , Escherichia coli/genetics , Escherichia coli/metabolism , Kinetics , Klebsiella pneumoniae/enzymology , Molecular Sequence Data , Multienzyme Complexes/metabolism , NAD/metabolism
19.
Microb Pathog ; 20(5): 255-61, 1996 May.
Article in English | MEDLINE | ID: mdl-8861391

ABSTRACT

Non-mucoid mutants of a Klebsiella pneumoniae wild type strain CG43 were generated by transposon mutagenesis. One of the mutants was incapable of fermenting galactose and was designated CG43-17. Alterations of the bacterial surface including capsule, lipopolysaccharides, and several species of outer membrane proteins were noted. The mutant was avirulent to mice and became highly sensitive to human serum. The defects could not be complemented by the gaIETK operon. Diminished activity of UDP-glucose pyrophosphorylase in CG43-17 suggested that it is a gaIU mutant. This possibility was confirmed because the parental phenotypes could be fully restored in the mutant by transforming it with a human liver cDNA encoding UDP-glucose pyrophosphorylase.


Subject(s)
Bacterial Outer Membrane Proteins/genetics , Klebsiella Infections/microbiology , Klebsiella pneumoniae/genetics , Animals , Humans , Klebsiella pneumoniae/pathogenicity , Mice , Mutagenesis , Mutation , Virulence/genetics
20.
Mol Plant Microbe Interact ; 8(5): 733-46, 1995.
Article in English | MEDLINE | ID: mdl-7579617

ABSTRACT

Pseudomonas syringae pv. syringae 61 contains a 25-kb hrp cluster that is sufficient to elicit the hypersensitive response (HR) in nonhost plants. Previous studies have shown that mutations in complementation groups VIII, IX, and XI in the hrp cluster abolished the ability of the bacterium to cause the HR. The sequence of a 3.7-kb SmaI-SstI fragment covering groups VIII and IX now reveals five open reading frames (ORFs) in the same transcript, designated as hrpU, hrpW, hrpO, hrpX, and hrpY, and predicted to encode proteins of 14,795, 23,211, 9,381, 28,489, and 39,957 Da, respectively. The hrpU, hrpW, hrpO, hrpX, and hrpY genes are homologous and arranged colinearly with the yscQ/spa33/spaO, yscR/spa24/spaP, yscS/spa9/spaQ, yscT/spa29/spaR, and yscU/spa40/spaS genes of Yersinia spp., Shigella flexneri, and Salmonella typhimurium, respectively. These proteins also show similarity to Fli/Flh proteins of Bacillus and enteric bacteria. The Ysc and Spa proteins are involved in the secretion of virulence factors, like the Yop and Ipa proteins. Fli/Flh proteins are involved in flagellar biogenesis. The sequence of a 2.9-kb EcoRV-EcoRI DNA fragment containing mainly group XI revealed five ORFs, designated hrpC, hrpD, hrpE, hrpF, and hrpG, predicted to encode proteins of 29,096, 15,184, 21,525, 7,959, and 13,919 Da, respectively. The first three genes belong to an operon containing hrpZ, which encodes an extracellular protein that elicits the HR. hrpF and hrpG are two potential ORFs upstream of hrpH in the hrpH operon. HrpC is homologous to Yersinia YscJ, Pseudomonas solanacearum HrpI, Xanthomonas compestris pv. vesicatoria HrpB3, and Rhizobium fredii NolT. HrpE is similar to YscL of Yersinia spp. P. s. pv. syringae 61 Hrp proteins are most similar to Ysc proteins among those homologs. TnphoA insertions in hrpC, hrpE, hrpW, hrpX, and hrpY abolished the ability of P. s.pv. syringae 61 to secrete HrpZ (harpinPss), as determined by immunoblot analysis of cell-bound and culture supernatant fractions. Thus, many of the proteins required for flagellar biogenesis and virulence protein secretion in plant and animal pathogens may have a common ancestry.


Subject(s)
Bacterial Outer Membrane Proteins/genetics , Genes, Bacterial , Multigene Family , Pseudomonas/genetics , Yersinia/genetics , Amino Acid Sequence , Bacterial Outer Membrane Proteins/metabolism , Base Sequence , Cloning, Molecular , DNA, Bacterial , DNA-Directed RNA Polymerases/metabolism , Escherichia coli/genetics , Molecular Sequence Data , Sequence Homology, Amino Acid , Viral Proteins
SELECTION OF CITATIONS
SEARCH DETAIL
...