scholarly journals First Report of Brown Root Rot on Alfalfa Caused by Phoma sclerotioides in the Continental United States

Plant Disease ◽  
1999 ◽  
Vol 83 (11) ◽  
pp. 1071-1071 ◽  
Author(s):  
C. R. Hollingsworth ◽  
F. A. Gray

Phoma sclerotioides G. Preuss ex Sacc. (previously named Plenodomus meliloti Dearn. & G.B. Sanford) is associated with root rot and extensive winterkill of leguminous forage crops, such as clover (Trifolium and Melilotus spp.), sainfoin (Onobrychis viciifolia), and alfalfa (Medicago sativa). Winterkill and root rot of irrigated alfalfa were observed for the first time in a field of cv. Multiplier in western Wyoming during the spring of 1996. Dark brown to black, sunken, rotting lesions were noted on upper secondary roots and taproots of dead and living diseased plants. Superficial and embedded beaked pycnidia and pycnosclerotia were observed near root lesions. A Phoma sp. isolated from a diseased plant in Farson, WY, was maintained on potato dextrose and half-strength V8-juice agars. Beaked pycnidia, typical of P. sclerotioides, were observed in culture when grown at 10°C for 2 months. A pathogenicity test was performed on cv. Multiplier. Two barley seeds colonized by a Phoma sp. derived from a Wyoming isolate were positioned on taproots of healthy, greenhouse-grown, 5-month-old plants ≈2.5 cm below the crown and were covered with a small piece of sterile cotton. Three replicate samples (24 plants inoculated and 24 plants uninoculated per replicate) were winter-hardened for 4 weeks (15.6°C/10°C, day/night, for 2 weeks, followed by 10°C/7.2°C, day/night, for 2 weeks) and placed outside during January 1998 in Laramie, WY, for a 4-month winter exposure period. Plants were rated for disease during June 1998. A disease severity rating of 1 to 5 was assigned to each experimental unit, where 1 = no disease and 5 = dead plant. The percentage of diseased plants at each severity rating for all inoculated plants was 1 = 19%, 2 = 33%, 3 = 31%, 4 = 13%, and 5 = 4%. Mycelium typical of P. sclerotioides was found on 99% of inoculated plant roots whether or not they had pycnidia. Pycnidia were found on the lower stems and petioles of some inoculated plants. Three percent of control plants also developed brown root rot (BRR) symptoms (taproot lesions or discoloration) by June 1998. The percentage of diseased plants at each severity rating for all uninoculated plants was 1 = 96%, 2 = 4%, and 3 through 5 = 0%. Aboveground propagule placement likely contributed to the spread of BRR by raindrop splash and wind-driven plant debris to adjacent alfalfa. Most inoculated plants had immature pycnidia or protopycnidia (94%), whereas 6.9% of the plants also had fully mature, beaked pycnidia. Pure fungal cultures were obtained from several diseased roots and compared with the original Wyoming Phoma sp. culture and a Canada isolate of P. sclerotioides (ATCC no. 56515) (2): colony, pycnidial, and conidial morphologies were identical, completing Koch's postulates. This is the first report of BRR on alfalfa in the continental United States. References: (1) J. G. N. Davidson. 1990. Brown root rot. Pages 29–31 in: Compendium of Alfalfa Disease. 2nd ed. The American Phytopathological Society, St. Paul, MN. (2) C. R. Hollingsworth et al. Phytopathology 88(suppl.):S39, 1999.

Plant Disease ◽  
2004 ◽  
Vol 88 (7) ◽  
pp. 769-769 ◽  
Author(s):  
R. C. Larsen ◽  
C. R. Grau ◽  
G. J. Vandemark ◽  
T. J. Hughes ◽  
B. D. Hudelson

Brown root rot (BRR) has been associated with winterkill of alfalfa (Medicago sativa L.) in the temperate regions of North America where winters are severe (1). Although suspected, BRR has not been associated with winterkill of alfalfa in the upper Midwestern United States. Alfalfa plants exhibiting symptoms resembling those induced by the causal agent Phoma sclerotioides G. Preuss ex Sacc. were collected from fields in Marinette, Pierce, and Marathon counties in Wisconsin during the spring and early summer of 2003. Symptoms included stunting and decline in 1- to 3-year-old plants that were slow to break dormancy in the early spring. Roots frequently exhibited dark brown lesions or were entirely decayed. Advanced lesions often formed dark bands around the circumference of tap and secondary roots. Beaked pycnidial structures typical of P. sclerotioides were also observed on many samples with advanced lesions. Plants with symptoms of BRR were also observed in Clark, Langlade, Lincoln, Oconto, Shawno, Taylor, and Wood counties. Several lesion areas of tissue on the tap and lateral roots of each sample were excised with a sterile scalpel. Total DNA was extracted using the Fast DNA kit (Bio 101, Carlsbad, CA). In addition, soil samples were collected in the root rhizosphere of symptomatic plants from four fields in two counties. Soil DNA was extracted with the Ultra-Clean DNA soil extraction kit (Mo Bio, Solana Beach, CA). DNA extractions were diluted 1:10 or 1:50, and samples were evaluated for the presence of P. sclerotioides using polymerase chain reaction (PCR)-based sequence-characterized amplified region (SCAR) markers according to the method described previously (4). Amplicons of the expected size (499 bp) were detected from alfalfa roots sampled from Marathon (4 of 4), Marinette (4 of 5), and Pierce (4 of 4) counties but not in roots from healthy controls produced in the greenhouse at Prosser, WA. PCR amplicons were also produced from all field soil samples in Marathon and Marinette counties. Proof of pathogenicity via Koch's postulates for this host-pathogen system was not attempted because of the extensive time period required (1). However, characteristic beaked pycnidia were present, and the pathogen was identified using PCR on DNA from roots symptomatic of BRR. Detection of BRR has been limited in the United States to Wyoming (2), but has been thought to occur in other states with severe winters (3). To our knowledge, this is the first report of P. sclerotioides in Wisconsin. References: (1) J. G. N. Davidson. Brown root rot. Pages 29–31 in: Compendium of Alfalfa Diseases. 2nd ed. D. L. Stuteville and D. C. Erwin, eds. The American Phytopathological Society, St. Paul, MN, 1990. (2) F. A. Gray et al. Pages 27–28 in: Proc. 10th Western Alfalfa Improv. Conf., 1997. (3) C. R. Hollingsworth et al. Can. J Plant Pathol. 25:215, 2003. (4) R. C. Larsen et al. Plant Dis. 86:928, 2002.


Plant Disease ◽  
2009 ◽  
Vol 93 (3) ◽  
pp. 317-317 ◽  
Author(s):  
M. J. Wunsch ◽  
R. Kersbergen ◽  
A. U. Tenuta ◽  
M. H. Hall ◽  
G. C. Bergstrom

Brown root rot (BRR), caused by the fungal pathogen Phoma sclerotioides G. Preuss ex Sacc. (synonym Plenodomus meliloti Dearn. & G.B. Sanford), is associated with yield loss of alfalfa (Medicago sativa L.) in regions with severe winters (1). In the spring of 2007, 9 to 69 alfalfa plants were collected from each of five production fields in Maine, 10 fields in Ontario, and nine fields in Pennsylvania. All alfalfa stands existed at least two winters. P. sclerotioides was isolated from alfalfa root and crown lesions from five fields in Maine (Penobscot, Somerset and Waldo counties), seven fields from southwestern (Woodstock and Niagara), south-central (Lindsay and Belleville), and southeastern Ontario (near Ottawa), and four fields in Pennsylvania (Columbia, Crawford, and Jefferson counties; 41.1 to 41.6°N). BRR incidence was 9 to 29% in Maine, 5 to 29% in Ontario, and 8 to 22% in Pennsylvania. In Ontario, some lesions girdled the crown; in three fields in Maine, large pycnidia characteristic of P. sclerotioides were present on alfalfa crowns and overwintered stems. On potato dextrose agar, conidia (5 to 8 × 2 to 3 μm, unicellular, hyaline, and ovoid) and pycnidia (0.33 to 1.15 mm in diameter with multiple beaks) of single-conidium isolates were characteristic of P. sclerotioides (2). Diagnostic PCR (3) of isolates resulted in a single amplicon of expected size (500 bp). The internal transcribed spacer (ITS) 1, 5.8S, and ITS2 of the rDNA were sequenced for 12 representative isolates, and sequences (GenBank Accession Nos. FJ179151 to FJ179162) were 95.5 to 100% identical to P. sclerotioides ATCC isolate 56515 over a 488-bp alignment. Eight months after seeding, potted ‘Vernal’ alfalfa was inoculated (4), kept at 4°C for 8 weeks, 0 to –2°C for 12 weeks, 4°C for 8 weeks, and 10 to 15°C for 7 weeks. Of 108 plants inoculated with the Maine isolates, 35 developed severe cortical lesions and 16 died. Of 18 plants inoculated with the Ontario isolates, 16 developed severe cortical lesions and eight died. Of 18 plants inoculated with a Pennsylvania isolate, 11 developed severe cortical lesions and five died. Lesions were typical of BRR: light to very dark brown, sometimes with a darker border, and often containing abundant pycnidia. Plant mortality was associated with lesions that girdled the root and crown. Of 18 plants in the control treatment, three developed severe cortical lesions and none died. BRR is common in Alberta, Saskatchewan, and Manitoba, but in eastern Canada it has been reported only in Nova Scotia. To our knowledge, this is the first report of BRR in Maine, Ontario, and Pennsylvania and the southernmost report of BRR in eastern North America. References: (1) B. Berkenkamp et al. Can. J. Plant Sci. 71:211, 1991. (2) G. H. Boerema et al. Persoonia 15:431, 1994. (3) R. C. Larsen et al. Plant Dis. 86:928, 2002. (4) M. J. Wunsch et al. Plant Dis. 91:1293, 2007.


Plant Disease ◽  
2008 ◽  
Vol 92 (4) ◽  
pp. 653-653 ◽  
Author(s):  
M. J. Wunsch ◽  
M. A. Dillon ◽  
R. Torres ◽  
H. F. Schwartz ◽  
G. C. Bergstrom

Brown root rot (BRR), caused by the fungal pathogen Phoma sclerotioides G. Preuss ex Sacc. (synonym Plenodomus meliloti Dearn. & G.B. Sanford), is associated with winterkill, slow emergence from winter dormancy, and yield loss of alfalfa (Medicago sativa L.) (1,2). BRR is a problem in regions with severe winters and is common in Alaska and Alberta, Saskatchewan and Manitoba, Canada. It was first observed in the continental United States in Wyoming during 1996 (2) and has subsequently been found in Idaho, Minnesota, Montana, New Hampshire, New York, Vermont, and Wisconsin. In the intermountain valleys of northern New Mexico and western Colorado, winters can be severe; alfalfa winterkill events occur periodically, but it is unknown if BRR is present. In May 2006, alfalfa plants were collected from production fields in Huerfano, Otero, and Rio Grande counties in Colorado and Rio Arriba and Taos counties in New Mexico and assessed for BRR. Two to three fields were sampled per county and 20 or 40 plants were collected per field. All fields existed for at least two winters. Fields sampled in Rio Grande County exhibited severe winterkill, with most plants completely girdled by crown lesions. Plants from other fields exhibited a range of root and crown rots. Isolation of P. sclerotioides was attempted from all plants with a previously described protocol (4). The pathogen was isolated from crown lesions of one alfalfa plant each from Rio Grande and Taos counties. Both lesions extended into the cortex. On potato dextrose agar and water agar with barley (4), single-conidium cultures of each isolate produced large pycnidia (0.35 to 0.80 mm in diameter) with multiple beaks, white cirri darkening to yellow with age, and unicellular, hyaline, ovoid conidia 5 to 7 μm long by 2 μm wide. Diagnostic PCR of the cultures using P. sclerotioides-specific primers (3) resulted in a single amplicon of expected size (500 bp). The internal transcribed spacer (ITS) 1, 5.8S, and ITS2 of the rDNA were amplified and sequenced using primers ITS1 and ITS4. The ITS sequences (GenBank Accession Nos. EU265669 and EU265670) were >98% identical to P. sclerotioides ATCC isolate 56515 over 503 bp of aligned sequence. Potted ‘Vernal’ alfalfa was inoculated 4 months after seeding, kept at 4°C for 5.5 weeks, 0 to –2°C for 12 weeks, and 4°C for 3 weeks. Of the 14 plants inoculated with the Colorado isolate, 11 developed cortical lesions and 8 winterkilled. Of the 23 plants inoculated with the New Mexico isolate, 22 developed cortical lesions and 16 winterkilled. Lesions were light to very dark brown, sometimes with a darker border and often containing abundant pycnidia. Winterkill was associated with lesions girdling the crown. P. sclerotioides was isolated from the lesions. To our knowledge, this is the southernmost report of BRR in North America and the first report of BRR in New Mexico and Colorado. The incidence and severity of BRR in the region surveyed appear to be considerably lower than in the more northern regions. References: (1) B. Berkenkamp et al. Can. J. Plant Sci. 71:211, 1991. (2) C. R. Hollingsworth et al. Can. J. Plant Pathol. 25:215, 2003. (3) R. C. Larsen et al. Plant Dis. 86:928, 2002. (4) M. J. Wunsch et al. Plant Dis. 91:1293, 2007.


Plant Disease ◽  
2017 ◽  
Vol 101 (6) ◽  
pp. 1038 ◽  
Author(s):  
J. Beckerman ◽  
H. Nisonson ◽  
N. Albright ◽  
T. Creswell

Plant Disease ◽  
2021 ◽  
Author(s):  
Charles Krasnow ◽  
Nancy Rechcigl ◽  
Jennifer Olson ◽  
Linus Schmitz ◽  
Steven N. Jeffers

Chrysanthemum (Chrysanthemum × morifolium) plants exhibiting stem and foliage blight were observed in a commercial nursery in eastern Oklahoma in June 2019. Disease symptoms were observed on ~10% of plants during a period of frequent rain and high temperatures (26-36°C). Dark brown lesions girdled the stems of symptomatic plants and leaves were wilted and necrotic. The crown and roots were asymptomatic and not discolored. A species of Phytophthora was consistently isolated from the stems of diseased plants on selective V8 agar (Lamour and Hausbeck 2000). The Phytophthora sp. produced ellipsoid to obpyriform sporangia that were non-papillate and persistent on V8 agar plugs submerged in distilled water for 8 h. Sporangia formed on long sporangiophores and measured 50.5 (45-60) × 29.8 (25-35) µm. Oospores and chlamydospores were not formed by individual isolates. Mycelium growth was present at 35°C. Isolates were tentatively identified as P. drechsleri using morphological characteristics and growth at 35°C (Erwin and Ribeiro 1996). DNA was extracted from mycelium of four isolates, and the internal transcribed spacer (ITS) region was amplified using universal primers ITS 4 and ITS 6. The PCR product was sequenced and a BLASTn search showed 100% sequence similarity to P. drechsleri (GenBank Accession Nos. KJ755118 and GU111625), a common species of Phytophthora that has been observed on ornamental and vegetable crops in the U.S. (Erwin and Ribeiro 1996). The gene sequences for each isolate were deposited in GenBank (accession Nos. MW315961, MW315962, MW315963, and MW315964). These four isolates were paired with known A1 and A2 isolates on super clarified V8 agar (Jeffers 2015), and all four were mating type A1. They also were sensitive to the fungicide mefenoxam at 100 ppm (Olson et al. 2013). To confirm pathogenicity, 4-week-old ‘Brandi Burgundy’ chrysanthemum plants were grown in 10-cm pots containing a peat potting medium. Plants (n = 7) were atomized with 1 ml of zoospore suspension containing 5 × 103 zoospores of each isolate. Control plants received sterile water. Plants were maintained at 100% RH for 24 h and then placed in a protected shade-structure where temperatures ranged from 19-32°C. All plants displayed symptoms of stem and foliage blight in 2-3 days. Symptoms that developed on infected plants were similar to those observed in the nursery. Several inoculated plants died, but stem blight, dieback, and foliar wilt were primarily observed. Disease severity averaged 50-60% on inoculated plants 15 days after inoculation. Control plants did not develop symptoms. The pathogen was consistently isolated from stems of symptomatic plants and verified as P. drechsleri based on morphology. The pathogenicity test was repeated with similar results. P. drechsleri has a broad host range (Erwin and Ribeiro 1996; Farr et al. 2021), including green beans (Phaseolus vulgaris), which are susceptible to seedling blight and pod rot in eastern Oklahoma. Previously, P. drechsleri has been reported on chrysanthemums in Argentina (Frezzi 1950), Pennsylvania (Molnar et al. 2020), and South Carolina (Camacho 2009). Chrysanthemums are widely grown in nurseries in the Midwest and other regions of the USA for local and national markets. This is the first report of P. drechsleri causing stem and foliage blight on chrysanthemum species in the United States. Identifying sources of primary inoculum may be necessary to limit economic loss from P. drechsleri.


Plant Disease ◽  
2011 ◽  
Vol 95 (10) ◽  
pp. 1316-1316 ◽  
Author(s):  
M. M. Díaz Arias ◽  
G. P. Munkvold ◽  
L. F. Leandro

Fusarium spp. are widespread soilborne pathogens that cause important soybean diseases such as damping-off, root rot, Fusarium wilt, and sudden death syndrome. At least 12 species of Fusarium, including F. proliferatum, have been associated with soybean roots, but their relative aggressiveness as root rot pathogens is not known and pathogenicity has not been established for all reported species (2). In collaboration with 12 Iowa State University extension specialists, soybean roots were arbitrarily sampled from three fields in each of 98 Iowa counties from 2007 to 2009. Ten plants were collected from each field at V2-V3 and R3-R4 growth stages (2). Typical symptoms of Fusarium root rot (2) were observed. Symptomatic and asymptomatic root pieces were superficially sterilized in 0.5% NaOCl for 2 min, rinsed three times in sterile distilled water, and placed onto a Fusarium selective medium. Fusarium colonies were transferred to carnation leaf agar (CLA) and potato dextrose agar and later identified to species based on cultural and morphological characteristics. Of 1,230 Fusarium isolates identified, 50 were recognized as F. proliferatum based on morphological characteristics (3). F. proliferatum isolates produced abundant, aerial, white mycelium and a violet-to-dark purple pigmentation characteristic of Fusarium section Liseola. On CLA, microconidia were abundant, single celled, oval, and in chains on monophialides and polyphialides (3). Species identity was confirmed for two isolates by sequencing of the elongation factor (EF1-α) gene using the ef1 and ef2 primers (1). Identities of the resulting sequences (~680 bp) were confirmed by BLAST analysis and the FUSARIUM-ID database. Analysis resulted in a 99% match for five accessions of F. proliferatum (e.g., FD01389 and FD01858). To complete Koch's postulates, four F. proliferatum isolates were tested for pathogenicity on soybean in a greenhouse. Soybean seeds of cv. AG2306 were planted in cones (150 ml) in autoclaved soil infested with each isolate; Fusarium inoculum was applied by mixing an infested cornmeal/sand mix with soil prior to planting (4). Noninoculated control plants were grown in autoclaved soil amended with a sterile cornmeal/sand mix. Soil temperature was maintained at 18 ± 1°C by placing cones in water baths. The experiment was a completely randomized design with five replicates (single plant in a cone) per isolate and was repeated three times. Root rot severity (visually scored on a percentage scale), shoot dry weight, and root dry weight were assessed at the V3 soybean growth stage. All F. proliferatum isolates tested were pathogenic. Plants inoculated with these isolates were significantly different from the control plants in root rot severity (P = 0.001) and shoot (P = 0.023) and root (P = 0.013) dry weight. Infected plants showed dark brown lesions in the root system as well as decay of the entire taproot. F. proliferatum was reisolated from symptomatic root tissue of infected plants but not from similar tissues of control plants. To our knowledge, this is the first report of F. proliferatum causing root rot on soybean in the United States. References: (1) D. M. Geiser et al. Eur. J. Plant Pathol. 110:473, 2004. (2) G. L. Hartman et al. Compendium of Soybean Diseases. 4th ed. The American Phytopathologic Society, St. Paul, MN, 1999. (3) J. F. Leslie and B. A. Summerell. The Fusarium Laboratory Manual. Blackwell Publishing, Oxford, UK, 2006. (4) G. P. Munkvold and J. K. O'Mara. Plant Dis. 86:143, 2002.


Plant Disease ◽  
2002 ◽  
Vol 86 (9) ◽  
pp. 928-932 ◽  
Author(s):  
R. C. Larsen ◽  
C. R. Hollingsworth ◽  
G. J. Vandemark ◽  
M. A. Gritsenko ◽  
F. A. Gray

A rapid technique for identification and detection of Phoma sclerotioides, the causal agent of brown root rot of alfalfa, has been developed using polymerase chain reaction (PCR). Amplification products obtained from random amplified polymorphic DNA (RAPD) reactions were cloned and sequenced, and two extended primer sets were designed from the resulting data that were used to detect sequence-characterized DNA markers. A single 499-bp DNA amplification product was consistently obtained from primers PSB12499 that was specific for 19 isolates of P.sclerotioides but was not produced from Phoma medicaginis or Phoma betae, or from other soilborne pathogens including Aphanomyces euteiches, Rhizoctonia solani, Fusarium oxysporum, Pythium ultimum, or Phytophthora infestans. A 499-bp amplification product was also produced from root tissue known to be infected with the fungus as verified by microscopic examination. A similar PCR product was obtained from soil samples collected from fields with an established infection of P. sclerotioides on alfalfa. This PCR-based assay enables detection of P. sclerotioides from alfalfa root tissue and in soil samples in a single day, including extraction of DNA, compared with standard methods that require up to 100 days for identification using agar media.


Plant Disease ◽  
2011 ◽  
Vol 95 (9) ◽  
pp. 1187-1187
Author(s):  
J. J. Sadowsky ◽  
T. D. Miles ◽  
A. M. C. Schilder

Necrotic stems and leaves were observed on 2- to 4-month-old, rooted microshoot plants (Vaccinium corymbosum L. ‘Liberty’ and ‘Bluecrop’, V. angustifolium Aiton ‘Putte’, and V. corymbosum × V. angustifolium ‘Polaris’) in a Michigan greenhouse in 2008 and 2009. As the disease progressed, leaves fell off and 80 to 100% of the plants died in some cases. Root rot symptoms were also observed. A fungus was isolated from stem lesions. On potato dextrose agar (PDA), cultures first appeared light tan to orange, then rusty brown and zonate with irregular margins. Chains of orange-brown chlamydospores were abundant in the medium. Macroconidiophores were penicillately branched and had a stipe extension of 220 to 275 × 2.5 μm with a narrowly clavate vesicle, 3 to 4 μm wide at the tip. Conidia were hyaline and cylindrical with rounded ends, (1-)3-septate, 48 to 73 × 5 to 7 (average 60 × 5.5) μm and were held together in parallel clusters. Perithecia were globose to subglobose, yellow, 290 to 320 μm high, and 255 to 295 μm in diameter. Ascospores were hyaline, 2- to 3-septate, guttulate, fusoid with rounded ends, slightly curved, and 30 to 88 × 5 to 7.5 (average 57 × 5.3) μm. On the basis of morphology, the fungus was identified as Calonectria colhounii Peerally (anamorph Cylindrocladium colhounii Peerally) (1,2). The internal transcribed spacer region (ITS1 and ITS2) of the ribosomal DNA and the β-tubulin gene were sequenced (GenBank Accession Nos. HQ909028 and JF826867, respectively) and compared with existing sequences using BLASTn. The ITS sequence shared 99% maximum identity with that of Ca. colhounii CBS 293.79 (GQ280565) from Java, Indonesia, and the β-tubulin sequence shared 97% maximum identity with that of Ca. colhounii CBS 114036 (DQ190560) isolated from leaf spots on Rhododendron sp. in North Carolina. The isolate was submitted to the Centraalbureau voor Schimmelcultures in the Netherlands (CBS 129628). To confirm pathogenicity, 5 ml of a conidial suspension (1 × 105/ml) were applied as a foliar spray or soil drench to four healthy ‘Bluecrop’ plants each in 10-cm plastic pots. Two water-sprayed and two water-drenched plants served as controls. Plants were misted intermittently for 2 days after inoculation. After 7 days at 25 ± 3°C, drench-inoculated plants developed necrotic, sporulating stem lesions at the soil line, while spray-inoculated plants showed reddish brown leaf and stem lesions. At 28 days, three drench-inoculated and one spray-inoculated plant had died, while others showed stem necrosis and wilting. No symptoms were observed on control plants. Fungal colonies reisolated from surface-disinfested symptomatic stem, leaf, and root segments appeared identical to the original isolate. Cy. colhounii was reported to cause a leaf spot on blueberry plants in nurseries in China (3), while Ca. crotalariae (Loos) D.K. Bell & Sobers (= Ca. ilicicola Boedijn & Reitsma) causes stem and root rot of blueberries in North Carolina (4). To our knowledge, this is the first report of Ca. colhounii causing a disease of blueberry in Michigan or the United States. Because of its destructive potential, this pathogen may pose a significant threat in blueberry nurseries. References: (1) P. W. Crous. Taxonomy and Pathology of Cylindrocladium (Calonectria) and Allied Genera. The American Phytopathological Society, St. Paul, MN, 2002. (2) L. Lombard et al. Stud. Mycol. 66:31, 2010. (3) Y. S. Luan et al. Plant Dis. 90:1553, 2006. (4) R. D. Milholland. Phytopathology 64:831, 1974.


Plant Disease ◽  
2009 ◽  
Vol 93 (4) ◽  
pp. 433-433 ◽  
Author(s):  
A. Garibaldi ◽  
G. Gilardi ◽  
D. Bertetti ◽  
M. L. Gullino

Aquilegia flabellata (Ranunculaceae), fan columbine, is a perennial herbaceous plant with brilliant blue-purple flowers with white petal tips. It can also be grown for cut flower production. In April of 2008, in several nurseries located near Biella (northern Italy), a leaf blight was observed on 10 to 15% of potted 30-day-old plants grown on a sphagnum peat substrate at 15 to 20°C and relative humidity of 80 to 90%. Semicircular, water-soaked lesions developed on leaves just above the soil line at the leaf-petiole junction and later along the leaf margins. Lesions expanded over several days along the midvein until the entire leaf was destroyed. Blighted leaves turned brown, withered, and abscised. Severely infected plants died. Diseased tissue was disinfested for 10 s in 1% NaOCl, rinsed with sterile water, and plated on potato dextrose agar (PDA) amended with 25 mg/liter streptomycin sulfate. A fungus with the morphological characteristics of Rhizoctonia solani was consistently recovered, then transferred and maintained in pure culture. Ten-day-old mycelium grown on PDA at 22 ± 1°C appeared light brown, rather compact, and had radial growth. Sclerotia were not present. Isolates obtained from affected plants successfully anastomosed with tester isolate AG 4 (AG 4 RT 31, obtained from tobacco plants). Results were consistent with other reports on anastomosis reactions (2). Pairings were also made with tester isolates of AG 1, 2.1, 2.2, 3, 6, 7, 11, and BI with no anastomoses observed between the recovered and tester isolates. The internal transcribed spacer (ITS) region of rDNA was amplified using primers ITS4/ITS6 and sequenced. BLASTn analysis (1) of the 648-bp fragment showed a 100% homology with the sequence of R. solani AG-4 AB000018. The nucleotide sequence has been assigned GenBank Accession No. FJ 534555. For pathogenicity tests, the inoculum of R. solani was prepared by growing the pathogen on PDA for 10 days. Five plants of 30-day-old A. flabellata were grown in 3-liter pots. Inoculum consisting of an aqueous suspension of PDA and mycelium disks (5 g of mycelium + agar per plant) was placed at the collar of plants. Five plants inoculated with water and PDA fragments alone served as control treatments. Plants were maintained in a greenhouse at temperatures between 20 and 24°C. The first symptoms, similar to those observed in the nursery, developed 7 days after the artificial inoculation. R. solani was consistently reisolated from infected leaves and stems. Control plants remained healthy. The pathogenicity test was carried out twice with similar results. The presence of R. solani AG1-IB on A. flabellata has been reported in Japan (4), while in the United States, Rhizoctonia sp. is described on Aquilegia sp. (3). This is, to our knowledge, the first report of leaf blight of A. flabellata caused by R. solani in Italy as well as in Europe. References: (1) S. F. Altschul et al. Nucleic Acids Res. 25:3389, 1997. (2) D. E. Carling. Grouping in Rhizoctonia solani by hyphal anastomosis reactions. In: Rhizoctonia Species: Taxonomy, Molecular Biology, Ecology, Pathology and Disease Control. Kluwer Academic Publishers, The Netherlands, 1996. (3) D. F. Farr et al. Fungi on Plants and Products in the United States. The American Phytopathological Society, St Paul, MN, 1989. (4) E. Imaizumi et al. J. Gen. Plant Pathol. 66:210, 2000.


Plant Disease ◽  
2010 ◽  
Vol 94 (6) ◽  
pp. 788-788 ◽  
Author(s):  
A. Garibaldi ◽  
D. Bertetti ◽  
M. T. Amatulli ◽  
M. L. Gullino

Persimmon (Diospyros kaki L.) is widely grown in Italy, the leading producer in Europe. In the fall of 2009, a previously unknown rot was observed on 3% of fruit stored at temperatures between 5 and 15°C in Torino Province (northern Italy). The decayed area was elliptical, firm, and appeared light brown to dark olive-green. It was surrounded by a soft margin. The internal decayed area appeared rotten, brown, and surrounded by bleached tissue. On the decayed tissue, black pycnidia that were partially immersed and up to 0.5 mm in diameter were observed. Light gray conidia produced in the pycnidia were unicellular, ovoid or lacriform, and measured 3.9 to 6.7 × 2.3 to 3.5 (average 5.0 × 2.9) μm. Fragments (approximately 2 mm) were taken from the margin of the internal diseased tissues, cultured on potato dextrose agar (PDA), and incubated at temperatures between 23 and 26°C under alternating light and darkness. Colonies of the fungus initially appeared ash colored and then turned to dark greenish gray. After 14 days of growth, pycnidia and conidia similar to those described on fruit were produced. The internal transcribed spacer (ITS) region of rDNA was amplified using the primers ITS4/ITS6 and sequenced. BLAST analysis (1) of the 502-bp segment showed a 100% similarity with the sequence of Phacidiopycnis washingtonensis Xiao & J.D. Rogers (GenBank Accession No. AY608648). The nucleotide sequence has been assigned the GenBank Accession No. GU949537. Pathogenicity tests were performed by inoculating three persimmon fruits after surface disinfesting in 1% sodium hypochlorite and wounding. Mycelial disks (10 mm in diameter), obtained from PDA cultures of one strain were placed on wounds. Three control fruits were inoculated with plain PDA. Fruits were incubated at 10 ± 1°C. The first symptoms developed 6 days after the artificial inoculation. After 15 days, the rot was very evident and P. washingtonensis was consistently reisolated. Noninoculated fruit remained healthy. The pathogenicity test was performed twice. Since P. washingtonensis was first identified in the United States on decayed apples (2), ‘Fuji’, ‘Gala’, ‘Golden Delicious’, ‘Granny Smith’, ‘Red Chief’, and ‘Stark Delicious’, apple fruits also were artificially inoculated with a conidial suspension (1 × 106 CFU/ml) of the pathogen obtained from PDA cultures. For each cultivar, three surface-disinfested fruit were wounded and inoculated, while three others served as mock-inoculated (sterile water) controls. Fruits were stored at temperatures ranging from 10 to 15°C. First symptoms appeared after 7 days on all the inoculated apples. After 14 days, rot was evident on all fruit inoculated with the fungus, and P. washingtonensis was consistently reisolated. Controls remained symptomless. To our knowledge, this is the first report of the presence of P. washingtonensis on persimmon in Italy, as well as worldwide. The occurrence of postharvest fruit rot on apple caused by P. washingtonensis was recently described in the United States (3). In Italy, the economic importance of the disease on persimmon fruit is currently limited, although the pathogen could represent a risk for apple. References: (1) S. F. Altschul et al. Nucleic Acids Res. 25:3389, 1997. (2) Y. K. Kim and C. L. Xiao. Plant Dis. 90:1376, 2006. (3) C. L. Xiao et al. Mycologia 97:473, 2005.


Sign in / Sign up

Export Citation Format

Share Document