scholarly journals First report of Colletotrichum tropicale causing anthracnose on Passiflora edulis in Brazil

Plant Disease ◽  
2021 ◽  
Author(s):  
Jackeline Laurentino da Silva ◽  
Walisson Ferreira da Silva da Silva ◽  
Luiz Eduardo Monteiro Lopes ◽  
Maria Jussara dos Santos Silva ◽  
Janaíne Rossane Araújo Silva-Cabral ◽  
...  

Brazil is the world's largest producer and consumer of yellow passion fruit (Passiflora edulis f. flavicarpa), mainly for the manufacture of concentrate and frozen juice as well as for fresh consumption (Faleiro et al. 2005). Between June and July 2018, passion fruit plants with symptoms of anthracnose were observed in commercial planting in the municipality of Coruripe (20 ha), northeastern state of Alagoas, Brazil. Approximately 70% of the plants showed leaves with relatively large, watery, circular spots that affected 30% of the leaf surface. Small fragments taken from the transition region of symptomatic tissue were superficially disinfected in 70% ethanol for 30 s and in 1% NaClO for 1 min, rinsed in sterile distilled water (SDW), dried on filter paper, plated on potato dextrose agar (PDA-Kasvi) incubated at 25°C under white light and 12 h photoperiod, for 3 days. Two isolates were obtained and deposited in the Collection of Phytopathogens at the Universidade Federal de Alagoas (COUFAL0281 and COUFAL0282). To identify the isolates, partial sequences of the glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and β-tubulin (TUB2) genes and of the rDNA-ITS (ITS) region were amplified by PCR. The partial sequences were deposited at GenBank (MT299339, MT334694, MT310553, MT299340, MT334695 and MT310554). Based on the BLASTn analysis, sequences of the PCR products showed high nucleotide similarity with sequences of the species C. tropicale (CBS 124949/ex-type and ICMP 18672), for GAPDH (98.94% and 100%), TUB2 (99 and 100%) genes and ITS (100%). This result was also confirmed in the phylogenetic tree of Bayesian Inference assembled with concatenated data (GAPDH, TUB2 and ITS). The colonies of the isolates were white with a white reverse, with dense mycelium, and mean growth rate of 7.54 mm/day, after 7 days on PDA medium at 25° C. Conidia were subcylindrical with rounded ends, hyaline, smooth walls and measured 13.63-20.59μm (= 17.54μm; n= 50) in length and 4.40-7.93 μm (= 5.88 μm; n= 50) in width. Appressoria were melanized, subglobose, irregular and measured 7.44 - 18.57 μm (= 10.04 μm; n= 50) in length and 5.49-10.16 μm (= 7.66 μm; n= 50) in width. These morphological characteristics were consistent with those described for Colletotrichum tropicale E.I. Rojas, S.A. Rehner & Samuels (Rojas et al. 2010). To confirm pathogenicity, 30 µL of a 106 conidia/mL sterile distilled water (SDW) conidia suspension, together with a drop of 20% Tween were deposited on the adaxial surface of passion fruit leaves wounded with a sterile needle, with four repetitions. The control consisted of leaves inoculated only with SDW. The leaves were placed in a plastic Gerbox box with sterilized filter paper moistened with SDW and maintained in a Biochemistry Oxygen Demand (BOD) incubator stove at 25 ºC and photoperiod of 12 h. After 7 days, typical anthracnose symptoms were observed on inoculated leaves. The pathogen was re-isolated and confirmed by morphological characterization, according to Koch's postulates. No symptoms were observed in the negative control. The occurrence of this species has been frequently reported in several other crops grown in northeastern Brazil (Silva et al. 2017; Veloso et al. 2018; Vieira et al. 2018; Costa et al. 2019). Additionally, many of these crops are grown in close proximity to the passion fruit orchards, thus favoring pathogen movement between hosts, probably, due to the anthropic influence, circulation of animals and insects, as well as wind driven rain splashes. However, this is first report of C. tropicale in Passiflora edulis in the world.

Plant Disease ◽  
2020 ◽  
Author(s):  
Wu Zhang ◽  
Xue Li Niu ◽  
Jin Yu Yang

As an economically important tropical and subtropical fruit crop, passion fruit (Passiflora edulis Sims) is widely planted in many provinces of southern China. In April 2017, postharvest fruit rot was observed on 15% to 25% of passion fruit in several fruit markets of Zhanjiang City in Guangdong Province. Initial disease symptoms on infected fruit were irregular, brown, water-soaked lesions, which enlarged into large black and sunken patches. Lesions were usually covered with an abundance of little black dots (pycnidia) and black-gray hyphae. For the pathogen isolation, fifteen symptomatic fruit were randomly collected from three local markets. Fourteen single-spore fungal isolates with similar morphology ware isolated from the infected tissues. Two isolates (ZW 49-1 and ZW 50-1) were randomly selected to further study. The colonies on PDA were initially greyish-white and became dark-gray with age. Abundant globular and irregular pycnidia were observed after incubation at 25 °C for 3 weeks. The conidia of the fungus were initially hyaline, unicellular, apex rounded, thick-walled, and ellipsoid, becoming dark brown, bicellular with longitudinal striations at maturity, 26.4 ± 2.5 × 13.4 ± 1.2 μm (n = 50). The morphology of the fungus resembled Lasiodiplodia theobromae (Pat.) Griff. & Maubl. (Phillips et al. 2013). To confirm species identification, the partial internal transcribed spacer (ITS) region of rDNA, translation elongation factor-alpha (EF1-α) and β-tubulin (TUB) gene were amplified from genomic DNA of the two isolates with the ITS1/ITS4, EF1-688F/EF1-986R, and Bt2a/Bt2b primers, respectively (Glass and Donaldson 1995; Alves et al. 2008; White et al. 1990). Base on the BLASTn analysis, the ITS (MT644473, MT644474), EF1-α (MT649210, MT649211) and TUB (MT649212, MT649213) sequences of both isolates were 100%, 99% and 100% similarity to the L. theobromae CBS 164.96 ex-type sequences in the NCBI database (AY640255, AY640258, and KU887532, respectively) (Phillips et al. 2013). For pathogenicity test, asymptomatic passion fruit were previously disinfested in 0.5 % sodium hypochlorite and superficially wounded with a sterile needle. Five-mm-diameter plugs with mycelial taken from 5-day-old PDA colonies were placed on the wounds. Sterile PDA plugs were used as negative controls. Each treatment had five replicates and the test was repeated twice. Fruit were maintained in plastic boxes to keep at 25°C for one week. One week after inoculation, gray mycelia had covered a majority of the fruit surface and caused a black, sunken rot. The inoculated fungus was reisolated and confirmed as L. theobromae by morphological characteristics. The mock inoculated fruit remained asymptomatic. The occurrence of fruit rot on passion fruit caused by L. theobromae was reported in Taiwan, China recently (Huang et al., 2019). To our knowledge, this is the first report of L. theobromae causing postharvest fruit rot on passion fruit in the Chinese mainland.


Plant Disease ◽  
2013 ◽  
Vol 97 (9) ◽  
pp. 1248-1248 ◽  
Author(s):  
N. B. Lima ◽  
M. W. Marques ◽  
S. J. Michereff ◽  
M. A. Morais ◽  
M. A. G. Barbosa ◽  
...  

From April to June 2010, mango fruits (Mangifera indica L.) (cv. Tommy Atkins) showing post-harvest anthracnose symptoms were collected during a survey conducted in São Francisco Valley, northeastern Brazil. Fruits affected by anthracnose showed sunken, prominent, dark brown to black decay spots. Small pieces (4 to 5 mm) of necrotic tissues were surface sterilized for 1 min in 1.5% NaOCl, washed twice with sterile distilled water, and plated onto potato dextrose agar (PDA) amended with 0.5 g liter–1 streptomycin sulfate. Plates were incubated at 25°C in the dark for 5 to 7 days and colonies that were morphologically similar to species of Colletotrichum were transferred to PDA (1). Identification was made using morphological characteristics and phylogenetic analysis. Two isolates (CMM 4101 and CMM 4102) presented colonies that had white aerial mycelia and orange conidial mass, varying between colorless and pale orange in reverse. Conidia were hyaline, cylindrical, and aseptate 14.52 (10.40 to 20.20) μm long and 4.90 (3.80 to 6.50) μm wide, length/width ratio = 3.0. Mycelial growth rate was 5.20 mm per day at 25°C. Morphological and cultural characterizations were consistent with the description of Colletotrichum karstii (3). PCR amplification by universal primers (ITS1/ITS4) and DNA sequencing of the internal transcribed spacer (ITS1-5.8S-ITS2 rRNA gene cluster) were conducted to confirm the identifications. Analysis of representative sequences (GenBank Accession Nos. HM585409 and HM585406) suggested that the isolated pathogen was C. karstii. Using published ITS data for C. karstii (3), a phylogenetic analysis was made via Bayesian inference, which shows that the isolated fungi belong to the C. karstii clade. Sequences of the isolates obtained in this study were deposited in GenBank (KC295235 and KC295236), and cultures were deposited in the Culture Collection of Phytopathogenic Fungi of the Universidade Federal Rural de Pernambuco (CMM, Recife, Brazil). Pathogenicity tests were conducted with the C. karstii strains on mango fruits cv. Tommy Atkins. Mycelial plugs taken from the margin of actively growing colonies (PDA) of each isolate were applied in shallow wounds (0.4 cm in diameter) at the medium region of the each fruit. PDA discs without fungal growing were used as controls. Inoculated fruits were placed in plastic containers lined with paper towels wetted in distilled water. The containers were partially sealed with plastic bags to maintain high humidity and incubated at 25°C in the dark. The plastic bags and paper towels were removed after 24 h, and fruits were kept at the same temperature. The experiment was arranged in a completely randomized design with four replicates per treatment (isolate) and four fruits per replicate. Typical anthracnose symptoms were observed after 10 days in mango fruits. C. karstii was successfully reisolated from symptomatic mango fruits to fulfill Koch's postulates. C. karstii was previously described from Orchidaceae in southwest China and the United States (2,3). To our knowledge, this is the first report of C. karstii causing mango anthracnose in Brazil and worldwide. References: (1) U. Damm et al. Stud. Mycol. 73:1, 2012. (2) I. Jadrane et al. Plant Dis. 96:1227, 2012. (3) Yang et al. Cryptogamie Mycol. 32:229, 2011.


Plant Disease ◽  
2021 ◽  
Author(s):  
G. Y. Shi ◽  
Quan Zeng ◽  
Y. W. Wei ◽  
Chun Jin Hu ◽  
X. L. Ye ◽  
...  

Violet passion fruit (Passiflora edulis Sims) is an important tropical and subtropical perennial evergreen vine with large-scale cultivation in Guangxi, China. Between May and September 2020, anthracnose symptoms occurred on passion fruit (cultivar Tainong No. 1) in Xingye county (22°77′13″N, 110°07′80″E) in Guangxi province, China. The disease incidence varied from 25 to 60% in different orchards. Initial symptoms on young fruits appeared as multiple tiny water-soaked, oval to irregular pale greenish spots. As the disease progressed, the lesions became medium brown, with sunken cavities. Under humid conditions, acervuli containing masses of conidia and dark setae were found on the lesions. The affected fruits became shriveled. Tissue pieces (5 × 5 mm) were cut out from infected fruits, surface sterilized in 75% ethanol for 15 s and 0.1% HgCl2 for 2 min, washed three times with sterile water, placed onto potato dextrose agar (PDA), and incubated at 28 °C for three days. Of the 29 Colletotrichum isolates obtained , the isolate B13 was selected for morphological characterization. B13 was purified by single spore isolation and incubated on PDA at 25°C under continuous fluorescent light irradiation, producing white to pale yellow colonies with dense aerial mycelia. The reverse side of the colony was pale yellowish to olive. Conidia were hyaline, unicellular, straight, cylindrical, with both ends slightly round or one end round and the other slightly pointed, measuring 10.5 to 18.8 (average 16.4) × 5.4 to 7.2 (average 6.3) µm (n = 50). Appressoria were light brown to dark black, smooth-walled, lobed, often with a roundish outline, sometimes also triangular, 7.2 to 10.9 (average 9.1) × 6.8 to 9.2 (average 8.2) µm (n = 50). Morphological characteristics of the isolate matched those of Colletotrichum brasiliense (Damm et al. 2012). The internal transcribed spacers (ITS), actin (ACT), glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and beta-tubulin (TUB2) genes of strain B13 were sequenced using the method and primers of Damm et al. (2012). Sequences of the amplified DNA regions were submitted to GenBank (ITS: MW198820; ACT: MW266083; GAPDH: MW266084; and TUB2: MW266085). A concatenated maximum likelihood phylogenetic tree was built using MEGA 7.0.21 in which B13 clustered with C. brasiliense and clearly separated from other Colletotrichum spp. Pathogenicity of B13 was assayed using one-year-old plants of violet passion fruit cultivar ‘Tainong No. 1’. Conidial suspensions were prepared from 7-day-old cultures grown on PDA at 28°C Sterile distilled water was used to dislodge conidia from the culture dish and the conidial concentration was adjusted to 1 × 106 spores mL-1 using a haemocytometer. Fruits were rinsed with sterilized water and wounded with a sterile needle at three locations. Three fruits were inoculated by spraying with 20 mL of the conidial suspension. Control fruits were sprayed with distilled water. Fruits were then covered with plastic bags to maintain high relative humidity . After 9 days, all inoculated fruits developed brown spots with sunken cavities, resembling symptoms observed in the field, and controls remained symptomless. Fungal cultures with phenotypic features similar to C. brasiliense were re-isolated from the symptomatic fruits, verifying C. brasiliense as the causal agent of the disease based on Koch’s postulates. C. boninense, C. gloeosporioides, C.queenslandicum, C. brevisporum, and C. karstii were reported as causal agents of anthracnose on passion fruit (Júnior et al.2010; Power et al. 2010; James et al.2014; Du et al.2017; Ran et al.2020). To the best of our knowledge, this is the first report of C. brasiliense causing anthracnose on passion fruit in China.


Plant Disease ◽  
2011 ◽  
Vol 95 (5) ◽  
pp. 616-616 ◽  
Author(s):  
J. Kim ◽  
O. Choi ◽  
J.-H. Kwon

Sweet persimmon (Diospyros kaki L.), a fruit tree in the Ebenaceae, is cultivated widely in Korea and Japan, the leading producers worldwide (2). Sweet persimmon fruit with flyspeck symptoms were collected from orchards in the Jinju area of Korea in November 2010. The fruit had fungal clusters of black, round to ovoid, sclerotium-like fungal bodies with no visible evidence of a mycelial mat. Orchard inspections revealed that disease incidence ranged from 10 to 20% in the surveyed area (approximately 10 ha) in 2010. Flyspeck symptoms were observed on immature and mature fruit. Sweet persimmon fruit peels with flyspeck symptoms were removed, dried, and individual speck lesions transferred to potato dextrose agar (PDA) and cultured at 22°C in the dark. Fungal isolates were obtained from flyspeck colonies on 10 sweet persimmon fruit harvested from each of three orchards. Fungal isolates that grew from the lesions were identified based on a previous description (1). To confirm identity of the causal fungus, the complete internal transcribed spacer (ITS) rDNA sequence of a representative isolate was amplified and sequenced using primers ITS1 and ITS4 (4). The resulting 552-bp sequence was deposited in GenBank (Accession No. HQ698923). Comparison with ITS rDNA sequences showed 100% similarity with a sequence of Zygophiala wisconsinensis Batzer & Crous (GenBank Accession No. AY598855), which infects apple. To fulfill Koch's postulates, mature, intact sweet persimmon fruit were surface sterilized with 70% ethanol and dried. Three fungal isolates from this study were grown on PDA for 1 month. A colonized agar disc (5 mm in diameter) of each isolate was cut from the advancing margin of a colony with a sterilized cork borer, transferred to a 1.5-ml Eppendorf tube, and ground into a suspension of mycelial fragments and conidia in a blender with 1 ml of sterile, distilled water. The inoculum of each isolate was applied by swabbing a sweet persimmon fruit with the suspension. Three sweet persimmon fruit were inoculated per isolate. Three fruit were inoculated similarly with sterile, distilled water as the control treatment. After 1 month of incubation in a moist chamber at 22°C, the same fungal fruiting symptoms were reproduced as observed in the orchards, and the fungus was reisolated from these symptoms, but not from the control fruit, which were asymptomatic. On the basis of morphological characteristics of the fungal colonies, ITS sequence, and pathogenicity to persimmon fruit, the fungus was identified as Z. wisconsinensis (1). Flyspeck is readily isolated from sweet persimmon fruit in Korea and other sweet persimmon growing regions (3). The exposure of fruit to unusual weather conditions in Korea in recent years, including drought, and low-temperature and low-light situations in late spring, which are favorable for flyspeck, might be associated with an increase in occurrence of flyspeck on sweet persimmon fruit in Korea. To our knowledge, this is the first report of Z. wisconsinensis causing flyspeck on sweet persimmon in Korea. References: (1) J. C. Batzer et al. Mycologia 100:246, 2008. (2) FAOSTAT Database. Retrieved from http://faostat.fao.org/ , 2008. (3) H. Nasu and H. Kunoh. Plant Dis. 71:361, 1987. (4) T. J. White et al. Page 315 in: PCR Protocols: A Guide to Methods and Applications. M. A. Innis et al., eds. Academic Press, Inc., New York, 1990.


Plant Disease ◽  
2021 ◽  
Author(s):  
Jiahao Lai ◽  
Guihong Xiong ◽  
Bing Liu ◽  
Weigang Kuang ◽  
Shuilin Song

Blueberry (Vaccinium virgatum), an economically important small fruit crop, is characterized by its highly nutritive compounds and high content and wide diversity of bioactive compounds (Miller et al. 2019). In September 2020, an unknown leaf blight disease was observed on Rabbiteye blueberry at the Agricultural Science and Technology Park of Jiangxi Agricultural University in Nanchang, China (28°45'51"N, 115°50'52"E). Disease surveys were conducted at that time, the results showed that disease incidence was 90% from a sampled population of 100 plants in the field, and this disease had not been found at other cultivation fields in Nanchang. Leaf blight disease on blueberry caused the leaves to shrivel and curl, or even fall off, which hindered floral bud development and subsequent yield potential. Symptoms of the disease initially appeared as irregular brown spots (1 to 7 mm in diameter) on the leaves, subsequently coalescing to form large irregular taupe lesions (4 to 15 mm in diameter) which became curly. As the disease progressed, irregular grey-brown and blighted lesion ran throughout the leaf lamina from leaf tip to entire leaf sheath and finally caused dieback and even shoot blight. To identify the causal agent, 15 small pieces (5 mm2) of symptomatic leaves were excised from the junction of diseased and healthy tissue, surface-sterilized in 75% ethanol solution for 30 sec and 0.1% mercuric chloride solution for 2 min, rinsed three times with sterile distilled water, and then incubated on potato dextrose agar (PDA) at 28°C for 5-7 days in darkness. Five fungal isolates showing similar morphological characteristics were obtained as pure cultures by single-spore isolation. All fungal colonies on PDA were white with sparse creeping hyphae. Pycnidia were spherical, light brown, and produced numerous conidia. Conidia were 10.60 to 20.12 × 1.98 to 3.11 µm (average 15.27 × 2.52 µm, n = 100), fusiform, sickle-shaped, light brown, without septa. Based on morphological characteristics, the fungal isolates were suspected to be Coniella castaneicola (Cui 2015). To further confirm the identity of this putative pathogen, two representative isolates LGZ2 and LGZ3 were selected for molecular identification. The internal transcribed spacer region (ITS) and large subunit (LSU) were amplified and sequenced using primers ITS1/ITS4 (Peever et al. 2004) and LROR/LR7 (Castlebury and Rossman 2002). The sequences of ITS region (GenBank accession nos. MW672530 and MW856809) showed 100% identity with accessions numbers KF564280 (576/576 bp), MW208111 (544/544 bp), MW208112 (544/544 bp) of C. castaneicola. LSU gene sequences (GenBank accession nos. MW856810 to 11) was 99.85% (1324/1326 bp, 1329/1331 bp) identical to the sequences of C. castaneicola (KY473971, KR232683 to 84). Pathogenicity was tested on three blueberry varieties (‘Rabbiteye’, ‘Double Peak’ and ‘Pink Lemonade’), and four healthy young leaves of a potted blueberry of each variety with and without injury were inoculated with 20 μl suspension of prepared spores (106 conidia/mL) derived from 7-day-old cultures of LGZ2, respectively. In addition, four leaves of each variety with and without injury were sprayed with sterile distilled water as a control, respectively. The experiment was repeated three times, and all plants were incubated in a growth chamber (a 12h light and 12h dark period, 25°C, RH greater than 80%). After 4 days, all the inoculated leaves started showing disease symptoms (large irregular grey-brown lesions) as those observed in the field and there was no difference in severity recorded between the blueberry varieties, whereas the control leaves showed no symptoms. The fungus was reisolated from the inoculated leaves and confirmed as C. castaneicola by morphological and molecular identification, fulfilling Koch’s postulates. To our knowledge, this is the first report of C. castaneicola causing leaf blight on blueberries in China. The discovery of this new disease and the identification of the pathogen will provide useful information for developing effective control strategies, reducing economic losses in blueberry production, and promoting the development of the blueberry industry.


Plant Disease ◽  
2012 ◽  
Vol 96 (10) ◽  
pp. 1580-1580
Author(s):  
J. H. Park ◽  
K. S. Han ◽  
J. Y. Kim ◽  
H. D. Shin

Sweet basil, Ocimum basilicum L., is a fragrant herb belonging to the family Lamiaceae. Originated in India 5,000 years ago, sweet basil plays a significant role in diverse cuisines across the world, especially in Asian and Italian cooking. In October 2008, hundreds of plants showing symptoms of leaf spot with nearly 100% incidence were found in polyethylene tunnels at an organic farm in Icheon, Korea. Leaf spots were circular to subcircular, water-soaked, dark brown with grayish center, and reached 10 mm or more in diameter. Diseased leaves defoliated prematurely. The damage purportedly due to this disease has reappeared every year with confirmation of the causal agent made again in 2011. A cercosporoid fungus was consistently associated with disease symptoms. Stromata were brown, consisting of brown cells, and 10 to 40 μm in width. Conidiophores were fasciculate (n = 2 to 10), olivaceous brown, paler upwards, straight to mildly curved, not geniculate in shorter ones or one to two times geniculate in longer ones, 40 to 200 μm long, occasionally reaching up to 350 μm long, 3.5 to 6 μm wide, and two- to six-septate. Conidia were hyaline, acicular to cylindric, straight in shorter ones, flexuous to curved in longer ones, truncate to obconically truncate at the base, three- to 16-septate, and 50 to 300 × 3.5 to 4.5 μm. Morphological characteristics of the fungus were consistent with the previous reports of Cercospora guatemalensis A.S. Mull. & Chupp (1,3). Voucher specimens were housed at Korea University herbarium (KUS). An isolate from KUS-F23757 was deposited in the Korean Agricultural Culture Collection (Accession No. KACC43980). Fungal DNA was extracted with DNeasy Plant Mini DNA Extraction Kits (Qiagen Inc., Valencia, CA). The complete internal transcribed spacer (ITS) region of rDNA was amplified with the primers ITS1/ITS4 and sequenced. The resulting sequence of 548 bp was deposited in GenBank (Accession No. JQ995781). This showed >99% similarity with sequences of many Cercospora species, indicating their close phylogenetic relationship. Isolate of KACC43980 was used in the pathogenicity tests. Hyphal suspensions were prepared by grinding 3-week-old colonies grown on PDA with distilled water using a mortar and pestle. Five plants were inoculated with hyphal suspensions and five plants were sprayed with sterile distilled water. The plants were covered with plastic bags to maintain a relative humidity of 100% for 24 h and then transferred to a 25 ± 2°C greenhouse with a 12-h photoperiod. Typical symptoms of necrotic spots appeared on the inoculated leaves 6 days after inoculation, and were identical to the ones observed in the field. C. guatemalensis was reisolated from symptomatic leaf tissues, confirming Koch's postulates. No symptoms were observed on control plants. Previously, the disease was reported in Malawi, India, China, and Japan (2,3), but not in Korea. To our knowledge, this is the first report of C. guatemalensis on sweet basil in Korea. Since farming of sweet basil has recently started on a commercial scale in Korea, the disease poses a serious threat to safe production of this herb, especially in organic farming. References: (1) C. Chupp. A Monograph of the Fungus Genus Cercospora. Ithaca, NY, 1953. (2) D. F. Farr and A. Y. Rossman. Fungal Databases. Systematic Mycology & Microbiology Laboratory, ARS, USDA. Retrieved from http://nt.ars-grin.gov/fungaldatabases/ , May 5, 2012. (3) J. Nishikawa et al. J. Gen. Plant Pathol. 68:46, 2002.


Plant Disease ◽  
2011 ◽  
Vol 95 (7) ◽  
pp. 874-874 ◽  
Author(s):  
Y. M. Shen ◽  
C. H. Chao ◽  
H. L. Liu

Gynura bicolor (Roxb. ex Willd.) DC., known as Okinawa spinach or hong-feng-cai, is a commonly consumed vegetable in Asian countries. In May 2010, plants with blight and wilt symptoms were observed in commercial vegetable farms in Changhua, Taiwan. Light brown-to-black blight lesions developed from the top of the stems to the petioles and extended to the base of the leaves. Severely infected plants declined and eventually died. Disease incidence was approximately 20%. Samples of symptomatic tissues were surface sterilized in 0.6% NaOCl and plated on water agar. A Phytophthora sp. was consistently isolated and further plated on 10% unclarified V8 juice agar, with daily radial growths of 7.6, 8.6, 5.7, and 2.4 mm at 25, 30, 35, and 37°C, respectively. Four replicates were measured for each temperature. No hyphal growth was observed at 39°C. Intercalary hyphal swellings and proliferating sporangia were produced in culture plates flooded with sterile distilled water. Sporangia were nonpapillate, obpyriform to ellipsoid, base tapered or rounded, and 43.3 (27.5 to 59.3) × 27.6 (18.5 to 36.3) μm. Clamydospores and oospores were not observed. Oospores were present in dual cultures with an isolate of P. nicotianae (p731) (1) A2 mating type, indicating that the isolate was heterothallic. A portion of the internal transcribed spacer sequence was deposited in GenBank (Accession No. HQ717146). The sequence was 99% identical to that of P. drechsleri SCRP232 (ATCC46724) (3), a type isolate of the species. The pathogen was identified as P. drechsleri Tucker based on temperature growth, morphological characteristics, and ITS sequence homology (3). To evaluate pathogenicity, the isolated P. drechsleri was inoculated on greenhouse-potted G. bicolor plants. Inoculum was obtained by grinding two dishes of the pathogen cultured on potato dextrose agar (PDA) with sterile distilled water in a blender. After filtering through a gauze layer, the filtrate was aliquoted to 240 ml. The inoculum (approximately 180 sporangia/ml) was sprayed on 24 plants of G. bicolor. An equal number of plants treated with sterile PDA processed in the same way served as controls. After 1 week, incubation at an average temperature of 29°C, blight and wilt symptoms similar to those observed in the fields appeared on 12 inoculated plants. The pathogen was reisolated from the lesions of diseased stems and leaves, fulfilling Koch's postulates. The controls remained symptomless. The pathogenicity test was repeated once with similar results. G. bicolor in Taiwan has been recorded to be infected by P. cryptogea (1,2), a species that resembles P. drechsleri. The recorded isolates of P. cryptogea did not have a maximal growth temperature at or above 35°C (1,2), a distinctive characteristic to discriminate between the two species (3). To our knowledge, this is the first report of P. drechsleri being associated with stem and foliar blight of G. bicolor. References: (1) P. J. Ann. Plant Pathol. Bull. 5:146, 1996. (2) H. H. Ho et al. The Genus Phytophthora in Taiwan. Institute of Botany, Academia Sinica, Taipei, 1995. (3) R. Mostowfizadeh-Ghalamfarsa et al. Fungal Biol. 114:325, 2010.


Plant Disease ◽  
2021 ◽  
Author(s):  
Qing Qu ◽  
Liu Shiwei ◽  
Ning Liu ◽  
Yunxia Liu ◽  
Jia Hui ◽  
...  

Abelmoschus manihot (Linn. ) Medicus (A. manihot) is an annual to perennial herb of the Malvaceae okra, mainly distributed in Guangdong, Guangxi, Fujian, Hunan, Hubei provinces. It can not only be used as an ornamental flower, but also has important economic and medicinal value. Last year, 10% A. manihot in 1,000 acres were observed with stalk rot in the Zhongshang Agricultural Industrial Park, 50 meters east of Provincial Highway 235 in Gaoyang County of Hebei province. Internal discoloration of the stem began brown to black discoloration of the vascular system and became hollow, with the mycelium growing on the surface. Stems from symptomatic plants (approximately 5 mm2) were dissected, washed free of soil, then soaked in 75% ethanol for 16 s to surface-sterilize, and 40 s in HgCl2, then rinsed three times in sterile water. After being dried with blotting paper, five pieces were placed on potato dextrose agar (PDA). After cultured 2 or 3 days, five isolates were purified and re-cultured on PDA in the dark at 25°C. The color of the colony was white. The hyphae were radial in PDA, and the aerial hyphae were flocculent, well-developed with luxuriant branches. The colonies were white and floccus, and the aerial hyphae were well developed, branched and without septum on corn meal agar (CMA). The sporangia were large or petal shaped, composed of irregular hyphae, terminal or intermediate , with the size of (31.6-88.4) μm ×(12.7- 14.6) μm. Vesicles were spherical, terminal or intermediate, ranging from 14.6 to 18.5μm. Oogonia were globose, terminal and smooth which stipe was straight. Antheridia were clavate or baggy and mostly intercalary, sometimes terminal. Oospores were aplerotic, 21.5 to 30.0 μm in diameter, 1.6 to 3.1 μm in wall thickness. The isolates morphological characteristics were consistent with P. aphanidermatum (van der Plaats-Niterink 1981, Wu et al. 2021 ). To identify the isolates, universal primers ITS1/ITS4 (White et al. 1993) were used for polymerase chain reaction–based molecular identification. The amplification region was sequenced by Sangon Biotech (Shanghai, China) and submitted to GenBank (MW819983). BLAST analysis showed that the sequence was 100% identical to Pythium aphanidermatum. Pathogenicity tests were conducted 3 times, with 4 treatments and 2 controls each time. The plants treated were 6 months old. Then the hyphae growing on PDA for 7 days were cut into four pieces. Next, they were inoculated into the soil of the A. manihot. Negative control was inoculated only with PDA for 7 days ( Zhang et al. 2000). The plants were then placed in a greenhouse under 28°C, 90% relative humidity. After inoculated 20 to 30 days, the infected plants showed stalk rot, the same symptoms as observed on the original plants. The control plants didn’t display symptoms. Pythium aphanidermatum was re-isolated from infected stems and showed the same characteristics as described above and was identical in appearance to the isolates used to inoculate the plants. To our knowledge, this is the first report of Pythium aphanidermatum infecting A. manihot stem and causing stalk rot in China. It may become a significant problem for A. manihot. Preliminary management practices are needed for reducing the cost and losses of production.


Plant Disease ◽  
2021 ◽  
Author(s):  
Jiahao Lai ◽  
Tongke Liu ◽  
Bing Liu ◽  
Weigang Kuang ◽  
Shuilin Song

Sweet potato [Ipomoea batatas (L.) Lam], is an extremely versatile vegetable that possesses high nutritional values. It is also a valuable medicinal plant having anti-cancer, antidiabetic, and anti-inflammatory activities. In July 2020, leaf spot was observed on leaves of sweet potato in Nanchang, China (28°45'51"N, 115°50'52"E), which affected the growth and development of the crop and caused tuberous roots yield losses of 25%. The disease incidence (total number of diseased plants / total number of surveyed plants × 100%) was 57% from a sampled population of 100 plants in the field. Symptomatic plants initially exhibited small, light brown, irregular-shaped spots on the leaves, subsequently coalescing to form large irregular brown lesions and some lesions finally fell off. Fifteen small pieces (each 5 mm2) of symptomatic leaves were excised from the junction of diseased and healthy tissue, surface sterilized in 75% ethanol solution for 30 sec and 0.1% mercuric chloride solution for 2 min, rinsed three times with sterile distilled water and incubated on potato dextrose agar (PDA) plates at 28°C in darkness. A total of seven fungal isolates with similar morphological characteristics were obtained as pure cultures by single-spore isolation. After 5 days of cultivation at 28°C, dark brown or blackish green colonies were observed, which developed brown, thick-walled, simple, or branched, and septate conidiophores. Conidia were 18.28 to 24.91 × 7.46 to 11.69 µm (average 21.27 × 9.48 µm, n = 100) in size, straight or slightly curved, middle cell unequally enlarged, brown to dark brown, apical, and basal cells slightly paler than the middle cells, with three septa. Based on morphological characteristics, the fungal isolates were suspected to be Curvularia plantarum (Raza et al. 2019). To further confirm the identification, three isolates (LGZ1, LGZ4 and LGZ5) were selected for molecular identification. The internal transcribed spacer region (ITS), glyceraldehyde-3-phosphate-dehydrogenase (GAPDH), and translation elongation factor 1-alpha (EF1-α) genes were amplified and sequenced using primers ITS1/ITS4 (Peever et al. 2004), gpd1/gpd2 (Berbee et al. 1999), EF-983F/EF-2218R (Rehner and Buckley 2005), respectively. The sequences of ITS region of the three isolates (accession nos. MW581905, MZ209268, and MZ227555) shared 100% identity with those of C. plantarum (accession nos. MT410571-72, MN044754-55). Their GAPDH gene sequences were identical (accession nos. MZ224017-19) and shared 100% identity with C. plantarum (accession nos. MN264120, MT432926, and MN053037-38). Similarly, EF1-α gene sequences were identical (accession nos. MZ224020-22) and had 100% identity with C. plantarum (accession nos. MT628901, MN263982-83). A maximum likelihood phylogenetic tree was built based on concatenated data from the sequences of ITS, GAPDH, and EF-1α by using MEGA 5. The three isolates LGZ1, LGZ4, and LGZ5 clustered with C. plantarum. The fungus was identified as C. plantarum by combining morphological and molecular characteristics. Pathogenicity tests were conducted by inoculating a conidial suspension (106 conidia/ml) on three healthy potted I. batatas plants (five leaves wounded with sterile needle of each potted plant were inoculated). In addition, fifteen wounded leaves of three potted plants were sprayed with sterile distilled water as a control. All plants were maintained in a climate box (12 h light/dark) at 25°C with 80% relative humidity. All the inoculated leaves started showing light brown flecks after 7 days, whereas the control leaves showed no symptoms. The pathogenicity test was conducted three times. The fungus was reisolated from all infected leaves of potted plants and confirmed as C. plantarum by morphological and molecular identification, fulfilling Koch’s postulates. To our knowledge, this is the first report of C. plantarum causing leaf spot on sweet potato in China. The discovery of this new disease and the identification of the pathogen will contribute to the disease management, provide useful information for reducing economic losses caused by C. plantarum, and lay a foundation for the further research of resistance breeding.


Plant Disease ◽  
2020 ◽  
Author(s):  
Xue Li ◽  
Jie Li ◽  
Hua Yong Bai ◽  
Kecheng Xu ◽  
Ruiqi Zhang ◽  
...  

Rubber tree (Hevea brasiliensis (Willd. ex Adr. Juss) Müll. Arg.) is used for the extraction of natural rubber and is an economically and socially important estate crop commodity in many Asian countries such as Indonesia, Malaysia, Thailand, India, Sri Lanka, China and several countries in Africa (Pu et al, 2007). Xishuangbanna City and Wenshan City are the main rubber cultivation areas in Yunnan Province, China. In November 2012, rubber tree showing typical wilt symptoms (Fig. 1 A) and vascular stains (Fig. 1 B) were found in Mengla County, Xishuangbanna City. This disease was destructive in these trees and plant wilt death rate reached 5%. The diseased wood pieces (0.5cm long) from trunk of rubber was surface disinfected with 75% ethanol for 30s and 0.1% mercuric chloride (HgCl2) for 2min, rinsed three times with sterile distilled water, plated onto malt extract agar medium (MEA), and incubated at 28℃. After 7 days, fungal-like filaments were growing from the diseased trunk. Six cultures from 6 rubber trunk were obtained and incubated on MEA at 28℃, after 7 days to observe the cultural features. The mycelium of each culture was white initially on MEA, and then became dark green. Cylindrical endoconidia apices rounded, non-septate, smooth, single or borne in chains (8.9 to 23.6 × 3.81 to 6.3μm) (Fig. 1 C). Chlamydospores (Fig. 1 D) were abundant, thick walled, smooth, forming singly or in chains (11.1 to 19.2 × 9.4 to 12.0μm). The mould fungus was identifed as Chalaropsis based on morphology (Paulin-Mahady et al. 2002). PCR amplification was carried out for 3 isolates, using rDNA internal transcribed spacer (ITS) primer pairs ITS1F and ITS4 (Thorpe et al. 2005). The nucleotide sequences were deposited in the GenBank data base and used in a Blast search of GenBank. Blast analysis of sequenced isolates XJm8-2-6, XJm8-2 and XJm10-2-6 (accessions KJ511486, KJ511487, KJ511489 respectively) had 99% identity to Ch. thielavioides strains hy (KF356186) and C1630 (AF275491). Thus the pathogen was identified as Ch. thielavioides based on morphological characteristics and rDNA-ITS sequence analysis. Pathogenicity test of the isolate (XJm8-2) was conducted on five 1-year-old rubber seedlings. The soil of 5 rubber seedlings was inoculated by drenching with 40 ml spore suspension (106 spores / ml). Five control seedlings were inoculated with 40 ml of sterile distilled water. All the seedlings were maintained in a controlled greenhouse at 25°C and watered weekly. After inoculated 6 weeks, all the seedlings with spore suspension produced wilt symptoms, as disease progressed, inoculated leaves withered (Fig. 1 E) and vascular stains (Fig. 1 F) by 4 months. While control seedlings inoculated with sterile distilled water remained healthy. The pathogen re-isolated from all inoculated symptomatic trunk was identical to the isolates by morphology and ITS analysis. But no pathogen was isolated from the control seedlings. The pathogenicity assay showed that Ch. thielavioides was pathogenic to rubber trees. Blight caused on rubber tree by Ceratocystis fimbriata previously in Brazil (Valdetaro et al. 2015), and wilt by Ch. thielavioides was not reported. The asexual states of most species in Ceratocystis are “chalara” or “thielaviopsis” (de Beer et al. 2014). To our knowledge, this is the first report of this fungus causing wilt of rubber in China. The spread of this disease may pose a threat to rubber production in China.


Sign in / Sign up

Export Citation Format

Share Document