scholarly journals First Report of Red-Fleshed Apple Anthracnose Caused by Colletotrichum siamense

Plant Disease ◽  
2021 ◽  
Author(s):  
Fengying Han ◽  
Yu-tong Zhang ◽  
Zaize Liu ◽  
Lei Ge ◽  
Lian-Dong Wang ◽  
...  

The red-fleshed apple (Malus niedzwetzkyana) produces a colored fruit and rich anthocyanins and it has become popular among consumers in Shandong (Yang et al 2020). In recent years, anthracnose diseases have been reported in red-fleshed apple orchards and nurseries in Shandong province, China. The incidence of anthracnose in the red-fleshed apple plantings ranges from 50-90%. Initially, anthracnose lesions on fruit begin as sub-circular shaped, sunken, pale brown. Over time black lesions enlarged and coalesced into large necrotic areas. The sunken centers of mature lesion became filled with slimy pink sporulation. In September 2015, fifteen fruit with anthracnose symptoms and sporulation were collected, and 11 single-spore isolates were obtained. Three representative isolates (JNTW11, JNTW2, JNTW33) were used for morphological and molecular characterization. On PDA, the colonies were initially white and turned into pale brown in three days. Orange-brown pigmentation was produced near the center on the reverse. Aerial mycelium was cottony, dense, pale white to pale gray. Acervuli developed visible orange-pink conidial masses. Conidiophores were colorless, septate, not branched or branched at the base. Conidia were 1-celled, hyaline, subcylindrical, oblong, attenuated with blunt ends, and the average size was 16.7 ± 1.5 × 6.1 ± 0.9 μm (n = 50). Appressoria were brown, obovoid or irregular, 9.2 ± 1.6 × 8.0 ± 1.8 μm (n = 20). The morphological characters matched the descriptions of Colletotrichum gloeosporioides sensu lato (Cannon et al. 2008). Isolates JNTW11, JNTW2, and JNTW33 were subject to bioinformatic characterization by partial sequencing of 6 genetic loci including the ribosomal internal transcribed spacer (ITS), actin (ACT), beta-tub2 (TUB2), calmodulin (CAL), chitin synthase (CHS-1), and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) (Weir et al, 2012). The ITS (MT577037, MT577040, MT577042), ACT (MT767712, MT767715, MT767717), TUB2 (MT767723, MT767726, MT767728), CAL (MT767689, MT767692, MT767694), CHS-1(MT767700, MT767703, MT767705), and GAPDH (MT767734, MT767737, MT767739) sequences were deposited in GenBank. The six sets of sequence data were concatenated “ITS-GAPDH-ACT-CHS-1-TUB2-CAL”, and the aligned sequences (2,007 bp) had 99.0% similarity to ex-type C. siamense ICMP18578. In a maximum likelihood phylogenetic tree, the highest log likelihood was -9148.55, and the isolates tested were in the C. siamense cluster with 96 % bootstrap support. Thus, the isolates were identified as C. siamense on the basis of multilocus phylogenetic analyses and morphological characters. To complete Koch’s postulates, several healthy red-fleshed apple fruit (‘Jiuhong’, 1 month prior to harvest) were inoculated using colonized and uncolonized hyphal plugs and a blank agar as a control. All inoculated fruit were placed in sterile tissue culture bottles containing 2 layers of wet paper towels at 28 °C under a 12 h light/dark cycle. All fruit developed anthracnose symptoms in 7 days while the controls did not develop any symptoms. The symptoms were similar to those collected from fruit in the field, and same fungus was re-isolated from the lesions. Presently it was known that C. acutatum, C. asianum, C. chrysophilum, C. cuscutae, C. fioriniae, C. fragariae, C. fructicola, C. gloeosporioides, C. godetiae, C. kahawae, C. karstii, C. limetticola, C. melonis, C. noveboracense, C. nymphaeae, C. paranaense, C. rhombiforme, C. salicis, and C. theobromicola could infect M. coronaria, M. domestica, M. prunifolia, M. pumila, and M. sylvestris worldwide. To our knowledge, this is the first report of C. siamense as a pathogen of M. niedzwetzkyana. This finding provides crucial information for the management of anthracnose disease in China.

Plant Disease ◽  
2021 ◽  
Author(s):  
Shengbo Han ◽  
Yanyong Cao ◽  
Jie Zhang ◽  
Jie Wang ◽  
Lili Zhang ◽  
...  

In a field survey from 2017 to 2019, Fusarium stalk rot symptoms including discolored, disintegrated stalk pith tissues and lodged plants were observed in maize hybrid lines Fuyu1611, Jidan66, and Danyu8439 grown in fields in Anshan (40o49′39′′N, 122 o34′6′′E), Liaoning province. Its incidence ranged from 15% to 20% and caused a yield loss of up to 30%. Infected pieces of stem tissues were dissected and then sterilized with 1% NaOCl for 1 min, 70% ethanol for 1 min, rinsed 3 times with sterilized ddH2O, and dried with filter paper in hood. Three pieces were placed onto Potato dextrose agar (PDA) and incubated at 25 °C for 5 days. The colonies were single-spore subcultured on PDA at 25 °C for 2 weeks (Leslie and Summerell 2006). Morphological features were observed on PDA and carnation leaf agar (CLA). The average mycelial growth rate was 4.5 to 10.3 mm/day at 25 °C on PDA. The colonies produced aerial mycelia, varying from dense white to grayish-rose, and secreted red pigments in the agar (Fig. 1A; 1B). Macroconidia produced on CLA were long and relatively slender, commonly 4- to 7-septate, averaging 85.6 × 5.2 μm, with thick walls and pronounced dorsiventral curvature with a distinctly foot-shaped and elongated basal cell and an apical cell that was whip-like (Fig. 1C). Microconidia were rarely observed on PDA or CLA. Morphological characteristics of the isolates were similar to the features of Fusarium longipes as previously described (Leslie and Summerell 2006). The portions of three phylogenic loci (EF1-α, RPB1, RPB2) were PCR amplified using the primer pairs EF1/EF2 (O'Donnell et al, 1998), lonR1F/lonR1R (5-TTTTCCTCACCAAGGAGCAGATCATG-3 and 5-CCAATGGACTGGGCAGCCAAAACGCC-3) and lonR2F/lonR2R (5-TATACATTTGCCTCCACTCTTTCCCAT-3 and 5-CGGAGCTTGCGTCCGGTGTGGCCGTTG-3) and sequenced. The consensus sequences were submitted to GenBank (MT513215 and MT997083 for TEF, MT513213 and MT997088 for RPB1; MT513214 and MW020572 for RPB2). BLASTn searches indicated that the nucleotide sequences of the three loci of the two isolates shared 94.52% to 99.69% identity with sequences of F. longipes strains deposited in the GenBank, Fusarium-ID and Fusarium MLST databases (Supplementary Table 1, 3, 4). A phylogram inferred via maximum likelihood analysis of the combined EF-1α, RPB1, RPB2 partial sequence data of Fusarium species (Supplementary Table 2) was inferred using the CIPRIES website (https://www.phylo.org). Isolates LNAS-05-A and LNAS-09-A clustered with F. longipes, with 98% bootstrap support (Fig. 2). Pathogenicity tests were conducted on three-leaf-stage seedlings and flowering-stage c.v. Zhengdan958 and B104 plants according to previously described methods (Ye et al., 2013; Zhang et al. 2016) with minor modifications. Three days after the roots of the seedlings were inoculated with 1 × 106 macroconidia solution, the leaves and stems exhibited typical wilt symptoms (Fig. 1D). Twenty flowering-stage maize plants were drilled individually at the second or third node above the soil using an electric drill (Bosch TSR1080-2-Li) to create a hole (8 mm in diameter). An approximately 0.5 mL mycelia plug (125 mL homogenized hyphal mats + 75 mL sterilized ddH2O) was injected into the hole and covered with Vaseline. Sterilized PDA plugs were used as a control. The stalk tissue of the split internodes turned dark brown and the brown area expanded above and below the injection site by 14 dpi. All of the inoculated plants developed characteristic stalk rot symptoms, whereas no symptoms were observed in the controls (Fig. 1E). The pathogen was re-isolated, and its identity was confirmed by sequencing the above mentioned loci. F. longipes was generally regarded as a tropical Fusarium species (Leslie and Summerell 2006). This is the first report that F. cf. longipes can cause stalk rot of maize under filed condition in a temperate, typical corn belt region of China.


Plant Disease ◽  
2012 ◽  
Vol 96 (2) ◽  
pp. 293-293
Author(s):  
A. J. Palmateer ◽  
T. L. B. Tarnowski ◽  
P. Lopez

Sansevieria Thunberg, a member of the Agavaceae, contains around 60 species indigenous to Africa, Arabia, and India. Several species and their cultivars are commercially produced for use as interior and landscape foliage plants. During August 2010, several local nurseries submitted Sansevieria trifasciata samples to the Florida Extension Plant Diagnostic Clinic in Homestead. Leaves had round, water-soaked lesions and as the disease progressed, lesions rapidly enlarged and coalesced, resulting in severe leaf blight. Both young and mature leaves were affected. Closer examination of mature lesions revealed numerous brownish black acervuli that were produced in concentric rings, which is characteristic of anthracnose. The fungus was identified as Colletotrichum sansevieriae Nakamura based on typical cultural characteristics, conidial and appressoria morphology (1). Conidia were straight, cylindrical, obtuse at the apex, slightly acute at the base with a truncate attachment point, and 12.5 to 33 (18.4) × 4 to 8.9 (6.5) μm (n = 50). Hyphopodia were ovate, dark brown, single celled, and 6.2 to 8.7 (7.7) × 6.3 to 7.5 (7.3) μm (n = 25). Colonies on potato dextrose agar (PDA) were grayish white, felted with aerial mycelium, reverse gray to dark olivaceous gray, and partly cream in color. Sequences of the rDNA internal transcribed spacer (ITS) regions of two isolates (GenBank Accession Nos. JF911349 and JF911350) exhibited 99% nucleotide identity to an isolate of C. sansevieriae (GenBank Accession No. HQ433226) collected from diseased sansevieria in Australia. In addition, a maximum parsimony analysis (MEGA v.5.0) indicates that the two C. sansevieriae isolates from Florida are monophyletic (86% bootstrap support) with the type species from Japan (SA-1-2 AB212991; SA-1-1 AB212990) and the Australian isolate. Pathogenicity of our sequenced isolates was evaluated in greenhouse experiments. Twelve- to fourteen-week-old sansevieria plants were inoculated with conidial suspensions (1 × 106 conidia/ml) of C. sansevieriae. Inoculum or autoclaved water was sprayed over the foliage until runoff. Four plants of each of two economically important cultivars, Laurentii and Moonshine, were sprayed per treatment and the experiment was repeated twice. Inoculated plants were placed in a greenhouse at 29°C with 70 to 85% relative humidity. Plants were observed for disease development, which occurred within 10 days of inoculation for both cultivars. No symptoms developed on the control plants. Foliar lesions closely resembled those observed in the affected nurseries. C. sansevieriae was consistently reisolated from symptomatic tissue collected from greenhouse experiments. On the basis of molecular phylogenetics and distinguishing morphological characters, Nakamura et al. erected C. sansevieriae as a novel species that appears to be restricted to the host sansevieria (1). To our knowledge, this is the first report of C. sansevieriae causing anthracnose of sansevieria in Florida. Reference: (1) M. Nakamura et al. J. Gen. Plant Pathol. 72:253, 2006.


Plant Disease ◽  
2021 ◽  
Author(s):  
Jie Tang ◽  
YiLin Du ◽  
LiXiang Lai ◽  
Qin Yang

Camellia oleifera, an evergreen small tree or shrub with high medicinal and ecological values, is mainly distributed in subtropical regions of China. Camellia oil obtained from Camellia oleifera seeds is rich in unsaturated fatty acids and unique flavors, and has become a rising high-quality edible vegetable oil in south of China (Zhuang 2008). The tea-oil tree Camellia oleifera plays important economic and ecological roles in Hunan province. During collecting trips, seeds of C. oleifera with disease symptoms have been observed in almost all oil-tea forests. In lab, the seeds can be infected by wounds and directly, however, wound infection is more rapid. In oil-tea forests, the wound of seed is often caused by external factors such as mechanical and insects. Symptomatic seeds exhibited brown rot symptoms with irregular, black spots, brown necrosis of the kernels, and accounted for 65% of the surveyed seeds (Fig. 1). Rotted seeds were surface-sterilized for 1 min in 75% ethanol, 3 min in 1% sodium hypochlorite, then rinsed for 2 min in sterile water and blotted on dry sterile filter paper. Discolored seed tissues were cut into pieces of 3 mm × 3 mm using a sterile scalpel, placed on potato dextrose agar (PDA) medium, and then incubated for 7 days at 25°C with a 12-h photoperiod. After 7 days of incubation, circular fungal colonies with dense aerial mycelium, produced black, wet spore masses. Four-septate conidia were ellipsoidal to obovoid, measuring 24 (22 to 26) × 6.5 (6 to 7) µm (n = 30). Conidia had three median cells, which were dark brown, with a single basal hyaline appendage, 4 (3.5 to 4.5) µm long, and two to four (usually three) apical hyaline appendages, 32 (27 to 35) µm long, similar to these recorded by Crous et al. (2011). Two single-spore isolates cultured on PDA medium were selected for DNA extraction. The ITS region was amplified using primers ITS5 and ITS4 (White et al. 1990). The partial translation elongation factor 1-alpha (tef1-α) gene region was amplified using primers EF1-728F (O'Donnell et al. 1998) and EF-2 (Carbone & Kohn 1999). The partial β-tubulin (tub2) was amplified using primers T1 and Bt2b (Glass & Donaldson 1995). The sequences of ITS (MW391815), tef1-α (MW398222), and tub2 (MW398223) were submitted to GenBank. BLAST analysis demonstrated that these sequences were 99%~100% similar to the sequences of ITS (MH553959), tef1-α (MH554377), and tub2 (MH554618) published for Neopestalotiopsis protearum. Phylogenetic analysis revealed that all the representative isolates recovered from symptomatic Camellia oleifera seeds showed 91% bootstrap support with Neopestalotiopsis protearum isolate in references (Fig. 2). Pathogenicity tests were conducted on 20 healthy seeds. We wounded the seeds by a sterilized needle on the middle position, and put the 5-mm-diameter agar plugs with actively grown mycelia (strain HNWC04) or pure PDA on the wound. We then covered the wounds with clean masking tape to prevent contamination and desiccation. After inoculation, the seeds were kept at 90 to 100% relative humidity at 25°C in a greenhouse for 3 weeks and monitored daily for lesion development. Twenty days after inoculation, all the seeds inoculated presented similar typical symptoms observed under natural conditions, whereas the control seeds showed no symptoms. Koch’s postulates were fulfilled by reisolating the same fungus and verifying its colony and morphological characters as Neopestalotiopsis protearum. To our knowledge, this is the first report of Neopestalotiopsis protearum causing oil-tea seed rot in China.


Plant Disease ◽  
2022 ◽  
Author(s):  
Kecheng Xu ◽  
Ruiqi Zhang ◽  
Jie Li ◽  
Xue Li ◽  
Jing Yang ◽  
...  

The rubber tree (Hevea brasiliensis) is an important economic resource for the rubber and latex industry. During November 2013 and June 2016, rubber trees showing typical wilt symptoms were found in Mengla, Xishuangbannan, Yunnan, China (N 21° 28', E 101° 33'). Symptomatic trees initially exhibited wilting of foliage on individual branches, then spread to the whole canopy, finally followed by death of the whole tree. Dark-blue to black discoloration was observed in the inner bark and affected xylem, a grayish layer of fungal growth and sporulation occasionally. The disease was detected on 20% of trees surveyed. The diseased tissues of three rubber trees were surface disinfected with 75% ethanol for 30 s and 0.1% mercuric chloride (HgCl2) for 2 min, rinsed three times with sterile distilled water, plated onto potato dextrose agar (PDA), and incubated at 25°C. After 7 days, a fungus was consistently observed growing from the tissue. Three single-spore isolates were obtained. In culture, colonies reaching 69 mm diam within 10 days, mycelium was initially white, then becoming celadon. After 5 days of perithecium formation, observed perithecia were black, globose (173.1 - 237.9 × 175.6 - 217.2 μm) and showed a long black neck (507.3 - 794.1 μm). Ascospore with outer cell wall forming a brim, hat-shaped at the tips of ostiolar hyphae (3.43 × 5.63 μm). Cylindrical endoconidia (10.5 - 39.7 × 3.5 - 6.6 μm) were hyaline. Chain of barrel-shaped conidia (7.2 - 9.5 × 4.1 - 6.2 μm) was found. Aleuroconidia were ovoid or obpyriform, and smooth (10.2 - 14.1 × 8.4 - 10.6 μm). Morphological characteristics of the fungus were consistent with the description of Ceratocystis fimbriata (Engelbrecht and Harrington 2005). The genomic DNA was extracted from isolates (XJm10-2-5, XJm8-2-5, XJm4) using the Chelex-100 method (Xu et al. 2020). The ITS region of rDNA was sequenced using the procedures of Thorpe et al. (2005). Analysis of ITS sequence data (GenBank accessions KJ511488, KJ511485, KT963149) showed that the isolates were 100% homologous to those of the isolates on Punica granatum and Colocasia esculenta from China (GenBank accessions KT963152, MH793673) by BLAST analysis. Neighbor-joining phylogenetic analyse were performed using MEGA 6.06 based on ITS sequences (Fig. 1). Analyses showed that all isolates located on the same clade with all C. fimbriata with a high bootstrap support. Therefore, the fungus was identified as C. fimbriata based on morphology and molecular evidences. Pathogenicity of C. fimbriata isolated from this study was tested by inoculation of three one-year-old pot-grown (3L) seedlings of rubber tree. The soil of three seedlings was inoculated by drenching with 30 ml spore suspension (2.0 × 106 spores / ml). Three control plants were inoculated with 30 ml of sterile distilled water. The experiment was repeated three times. The plants were kept in a controlled greenhouse at 25°C and watered weekly. After the inoculation for one month, all the isolates produced typical wilt symptoms, while control plants showed no symptoms. The original fungus was successfully re-isolated from inoculated trees and identified as C. fimbriata according to the methods described above. The pathogenicity assay showed that C. fimbriata was pathogenic to rubber trees. C. fimbriata was first reported on rubber tree in Brazil (Albuquerque et al. 1972; Silveira et al. 1985). To the best of our knowledge, this is the first report of C. fimbriata causing wilt of rubber tree in China. This finding contributes to understanding the diversity of this pathogen, and it appears to be a significant threat to rubber trees in its ecosystem.


MycoKeys ◽  
2018 ◽  
Vol 36 ◽  
pp. 83-105 ◽  
Author(s):  
Jing Yang ◽  
Jian-Kui Liu ◽  
Kevin D. Hyde ◽  
E.B. Gareth Jones ◽  
Zuo-Yi Liu

A survey of freshwater fungi on submerged wood in China and Thailand resulted in the collection of three species in Dictyocheirospora and four species in Dictyosporium including two new species in the latter genus. Morphological characters and phylogenetic analyses based on ITS, LSU and TEF1α sequence data support their placement in Dictyocheirospora and Dictyosporium (Dothideomycetes). An updated backbone tree is provided for the family Dothideomycetes. Descriptions and illustrations of the new taxa and re-collections are provided. Four new combinations are proposed for Dictyocheirospora.


Phytotaxa ◽  
2019 ◽  
Vol 419 (1) ◽  
pp. 28-38 ◽  
Author(s):  
KE-KE ZHANG ◽  
SINANG HONGSANAN ◽  
DANUSHKA S. TENNAKOON ◽  
SHENG-LI TIAN ◽  
NING XIE

Phaeosphaeria chinensis sp. nov. was found on dead leaves, collected from Guangdong Province, China. Morphology of the new species was compared with other Phaeosphaeria species and related genera of Phaeosphaeriaceae. Phylogenetic analyses of combined ITS, LSU, SSU and TEF-1 sequence data based on maximum parsimony (MP), maximum likelihood (ML) and Bayesian inference (BI) revealed that P. chinensis as a distinct species within the Phaeosphaeria with high bootstrap support. The comparison of the new species with other Phaeosphaeria species and a comprehensive description and micrographs are provided. The linkage of sexual and asexual morphs of the new species is also showed.


The Auk ◽  
2000 ◽  
Vol 117 (2) ◽  
pp. 321-336 ◽  
Author(s):  
John Klicka ◽  
Kevin P. Johnson ◽  
Scott M. Lanyon

AbstractHistorically, a paucity of comparative morphological characters has led to much debate regarding relationships within and among the major lineages of New World nine-primaried oscines. More recently, DNA-DNA hybridization studies have provided novel and testable hypotheses of relationships, although no consensus has been reached. For 40 songbird taxa, we obtained 1,929 base pairs (bp) of DNA sequence from the mitochondrial cytochrome-b (894 bp) and NADH dehydrogenase subunit 2 (1,035 bp) genes. Phylogenetic analyses confirm the monophyly of this assemblage as traditionally defined. The lineages delineated historically on morphological grounds are retained; finches (Fringillinae) are sister to a well-supported clade (Emberizinae) containing blackbirds (Icterini), sparrows (Emberizini), wood-warblers (Parulini), tanagers (Thraupini), and cardinal-grosbeaks (Cardinalini). However, each tribe individually is either paraphyletic or polyphyletic with respect to most recent songbird classifications. Our results suggest that Euphonia is not a tanager but perhaps represents a derived form of cardueline finch. Piranga, traditionally considered a typical tanager, is a cardinaline in all of our analyses. Calcarius falls outside the sparrow lineage in all of our analyses, but its true affinities remain unclear. Elements of four different AOU families are represented in our clade Thraupini. The inclusion of several “tanager-finches” (Haplospiza, Diglossa, Tiaris, Volatinia, Sporophila) and a nectarivore (Coereba) in this clade is consistent with findings from other molecular phylogenies in suggesting that convergence in feeding specializations among some lineages has confounded traditional morphological classifications. We obtained a novel arrangement of relationships among tribes in our “best” topology; Cardinalini is sister to the rest of the Emberizinae assemblage (as defined by Sibley and Ahlquist [1990]), and Thraupini is sister to a clade containing Icterini, Emberizini, and Parulini. Despite nearly 2,000 bp of sequence for each taxon, and a high degree of stability across most weighting schemes and analytical methods, most nodes lack strong bootstrap support. The ND2 gene provided higher resolution than did cytochrome b, but combining genes provided the most highly supported and resolved topology. We consider the phylogeny a working hypothesis to be used as a guide for further studies within the nine-primaried oscine assemblage.


Plant Disease ◽  
2014 ◽  
Vol 98 (5) ◽  
pp. 690-690 ◽  
Author(s):  
L. P. Kou ◽  
V. L. Gaskins ◽  
Y. G. Luo ◽  
W. M. Jurick

Apples are grown and stored for 9 to 12 months under controlled atmosphere conditions in the United States. During storage, apples are susceptible to various fungal pathogens, including several Alternaria species (2). Alternaria tenuissima (Nees) Wiltshire causes dry core rot (DCR) on apples during storage and has recently occurred in South Africa (1). Losses range widely, but typically occur at 6 to 8% annually due to this disease (2). In February 2013, ‘Nittany’ apples with round, dark-colored, dry, spongy lesions were obtained from wooden bins in a commercial cold storage facility located in Pennsylvania. Symptomatic fruits were transported to the lab, rinsed with sterile water, and the lesions were sprayed with 70% ethanol until runoff and wiped dry. The skin was aseptically removed with a scalpel, and asymptomatic tissue was placed onto potato dextrose agar (PDA) and incubated at 25°C. Two single-spore isolates were propagated on PDA and permanent cultures were maintained as slants and stored at 4°C. The fungus produced a cottony white mycelium that turned olive-green to brown with abundant aerial hyphae and had a dark brown to black reverse on PDA. Isolates were identified as Alternaria based on conidial morphology as the spores were slightly melanized and obclavate to obpyriform catentulate with longitudinal and transverse septa attached in unbranched chains on simple short conidiophores. Conidia ranged from 10 to 70 μm long (mean 27.7 μm) and 5 to 15 μm wide (mean 5.25 μm) (n = 50) with 1 to 6 transverse and 0 to 2 longitudinal septa. Conidial beaks, when present, were short (5 μm or less) and tapered. Mycelial genomic DNA was extracted, and a portion of the histone gene (357 bp) was amplified via gene specific primers (Alt-His3-F/R) using conventional PCR (Jurick II, unpublished). The forward and reverse sequences were assembled into a consensus representing 2× coverage and MegaBLAST analysis showed that both isolates were 100% identical to Alternaria tenuissima isolates including CR27 (GenBank Accession No. AF404622.1) that caused DCR on apple fruit during storage in South Africa. Koch's postulates were conducted using 10 organic ‘Gala’ apple fruit that were surface sterilized with soap and water, sprayed with 70% ethanol, and wiped dry. The fruit were aseptically wounded with a nail to a 3 mm depth, inoculated with 50 μl of a conidial suspension (1 × 104 conidia/ml), and stored at 25°C in 80 count boxes on paper trays for 21 days. Mean lesion diameters on inoculated ‘Gala’ apple fruit were 19.1 mm (±7.4), water only controls (n = 10 fruit) were symptomless, and the experiment was repeated. Symptoms observed on artificially inoculated ‘Gala’ apple fruit were similar to the decay observed on ‘Nittany’ apples from cold storage. Based on our findings, it is possible that A. tenuissima can cause decay that originates from wounded tissue in addition to dry core rot, which has been reported (1). Since A. tenuissima produces potent mycotoxins, even low levels of the pathogen could pose a health problem for contaminated fruit destined for processing and may impact export to other countries. To the best of our knowledge, this is the first report of alternaria rot caused by A. tenuissima on apple fruit from cold storage in the United States. References: (1) J. C. Combrink et al. Decid. Fruit Grow. 34:88, 1984. (2) M. Serdani et al. Mycol. Res. 106:562, 2002. (3) E. E. Stinson et al. J. Agric. Food Chem. 28:960, 1980.


Phytotaxa ◽  
2014 ◽  
Vol 174 (4) ◽  
pp. 187 ◽  
Author(s):  
Lakshmi Attigala ◽  
Jimmy K. Triplett ◽  
Hashendra-Suvini Kathriarachchi ◽  
Lynn G Clark

Kuruna, a new temperate woody bamboo (Poaceae, Bambusoideae, Arundinarieae) genus from Sri Lanka, is recognized based on chloroplast sequence data from five markers (coding: ndhF 3’ end; non-coding: rps16-trnQ, trnC-rpoB, trnD-trnT, trnT-trnL). This genus represents the twelfth major lineage of temperate woody bamboos and is characterized by pachymorph culm bases with short necks, unicaespitose clumps, culm leaf girdles ca. 1 mm wide, usually abaxially hispid culm leaves with non-irritating hairs, persistent foliage leaf sheaths, complete branch sheathing and acute to biapiculate palea apices. Maximum Parsimony, Bayesian Inference and Maximum Likelihood analyses of a combined data set consistently strongly supported the monophyly of this Sri Lankan temperate woody bamboo clade. Although the Kishino-Hasegawa test is unable to reject the alternative hypothesis of monophyly of the Sri Lankan clade plus Bergbambos tessellata from South Africa, Kuruna and Bergbambos are distinguishable by a combination of morphological characters. A few additional cpDNA markers not previously used in phylogenetic analyses of Arundinarieae were tested to evaluate their utility in this taxonomically difficult tribe.


Parasitology ◽  
2005 ◽  
Vol 130 (6) ◽  
pp. 669-677 ◽  
Author(s):  
X. Y. WU ◽  
N. B. CHILTON ◽  
X. Q. ZHU ◽  
M. Q. XIE ◽  
A. X. LI

Sequences of the first internal transcribed spacer (ITS-1) and the D1-D3 domains of the large subunit (LSU) of the ribosomal DNA (rDNA) were determined for multiple specimens of 4 operational taxonomic units (OTUs) of the monogenean, Pseudorhabdosynochus lantauensis. OTUs were defined based on their collecting localities, host and/or morphological characteristics. All P. lantauensis specimens of one group (OTUs 1 and 3) differed in their sequences of the ITS-1 and partial LSU rDNA when compared with specimens of a second group (OTUs 2 and 4) by 12% and 2%, respectively. Results of the phylogenetic analyses of the LSU rDNA sequence data showed total (100%) bootstrap support for the separation of P. lantauensis into 2 distinct clades. At least 11 of the 18 nucleotide differences in the LSU sequence between the two P. lantauensis clades were derived (i.e. autapomorphic) characters when the morphologically distinct species, P. epinepheli and P. coioidesis, were used as outgroups. Furthermore, there were several autapomorphic character states for each P. lantauensis clade. This provides sufficient evidence to reject the null hypothesis that P. lantauensis represents a single species. Morphological and morphometric differences between these two clades provided additional strong support for the separation of P. lantauensis into two species. These two parasite species were found to co-exist on one of the two species of serranid fish (i.e. Epinephelus coioides) examined in the South China Sea (Guangdong Province, China).


Sign in / Sign up

Export Citation Format

Share Document