scholarly journals First Report of Fusarium verticillioides Causing Fusariosis on Triticale Grain in Serbia

Plant Disease ◽  
2021 ◽  
Author(s):  
Iva Savić ◽  
Milica Nikolic ◽  
Ana Nikolic ◽  
Vesna Kandić ◽  
Ivana Vico ◽  
...  

Triticale (Triticosecale Wittmack) is a cereal crop species developed to combine positive traits of wheat and rye into a single plant (Arendt and Zannini, 2013). Due to its high protein content, favourable amino acid composition and nutritional values higher than maize, triticale is an excellent component for preparing feed for domestic animals (Đekić et al. 2019). In May 2017, discoloured spikes of triticale were observed in a field in Zemun Polje, Serbia, with a disease incidence between 15 and 20%. A total of 400 kernels obtained from discoloured spikes were surface sterilised with 1% sodium hypochlorite solution for 3 minutes and then rinsed 3 times with sterilized water and plated on potato dextrose agar (PDA) at 25°C incubated in the dark for 7 days. A total of 15 isolates of genus Fusarium on PDA formed cottony white to greyish purple colony with dark yellow to purple grey reverse and average colony. Obtained isolates were transferred to carnation leaf agar (CLA) and synthetic nutrition agar (SNA), for 7-day incubation in the dark at 25°C. Ten isolates formed “rabbit ear” monophialides which is the main characteristic of Fusarium verticillioides (Sacc.) Nirenberg. Microconidia are formed from monophialids in long chains. Their shape is oval with a flattened base, without septa (4.0-18.0 x 1.5-4.5 µm). Shape of macroconidia is slightly falcate to almost straight, (3.5-4.5 x 31-58 µm). Chlamydospores are not produced (Leslie and Summerell, 2006). The genomic DNA from one F. verticillioides isolate was extracted using the DNeasy Plant Mini kit (Qiagen, Hilden) according to the manufacturer’s protocol. Molecular identification was confirmed by sequencing the internal transcribed spacer (ITS), translation elongation factor-1 alpha (EF-1α) and RPB2 (O′Donell, 2000). Sequence was deposited in GenBank as MZ664391, MZ666958 and MZ666957. This sequence was compared with the sequences of F. verticillioides strains registered in the GenBank database based on nucleotide similarity. It showed 100% identity to the sequences MT180471.1 (ITS), MN861767.1 (EF-1 α) and MT264836.1 (RPB2). Pathogenicity of all isolates was tested on 20 randomly selected triticale spikes in four replicates (Mesterházy et al. 1999). Inoculation was performed when half of the plants reached the flowering stage, by spraying with 20 ml of spore suspension (1 × 106 spores/ml from 7-day-old cultures on PDA). F. verticillioides RBG 1603 Q27 was used as a positive control. Inoculated spikes were covered with PVC bags for 48h. Discoloured spikes were observed after 3 week incubation in a greenhouse at 20°C with a 16h light/8h dark photoperiod. All inoculated spikes showed symptoms, similar to those from field infections. Control spikes were symptomless. The fungus was reisolated and was morphologically identical to the original isolates, thus completing Koch’s postulates. Based on morphological, molecular and pathogenic features the isolated fungus was identified as F. verticillioides. Fusarium species infect crop plants worldwide resulting in yield loss and reduced grain quality due to mycotoxin contamination (Asam et al, 2017). In Serbia F. verticillioides was isolated from wheat and barley grains (Stanković et al., 2012) and to the best of our knowledge, this is the first report of F. verticillioides on triticale grain in Serbia.

Plant Disease ◽  
2020 ◽  
Author(s):  
Fangmin Hao ◽  
Quanyu Zang ◽  
Weihong Ding ◽  
Erlei Ma ◽  
Yunping Huang ◽  
...  

Melon (Cucumis melo L.) is a member of the Cucurbitaceae family, an important economical and horticultural crop, which is widely grown in China. In May 2020, fruit rot disease with water-soaked lesions and pink molds on cantaloupe melons was observed in several greenhouses with 50% disease incidence in Ningbo, Zhejiang Province in China. In order to know the causal agent, diseased fruits were cut into pieces, surface sterilized for 1 min with 1% sodium hypochlorite (NaClO), 2 min with 75% ethyl alcohol, rinsed in sterile distilled water three times (Zhou et al. 2018), and then placed on potato dextrose agar (PDA) medium amended with streptomycin sulfate (100 μg/ml) plates at 25°C for 4 days. The growing hyphae were transferred to new PDA plates using the hyphal tip method, putative Fusarium colonies were purified by single-sporing. Twenty-five fungal isolates were obtained and formed red colonies with white aerial mycelia at 25°C for 7 days, which were identified as Fusarium isolates based on the morphological characteristics and microscopic examination. The average radial mycelial growth rate of Fusarium isolate Fa-25 was 11.44 mm/day at 25°C in the dark on PDA. Macroconidia were stout with curved apical and basal cells, usually with 4 to 6 septa, and 29.5 to 44.2 × 3.7 to 5.2 μm on Spezieller Nährstoffarmer agar (SNA) medium at 25°C for 10 days (Leslie and Summerell 2006). To identify the species, the internal transcribed spacer (ITS) region and translational elongation factor 1-alpha (TEF1-α) gene of the isolates were amplified and cloned. ITS and TEF1-α was amplified using primers ITS1/ITS4 and EF1/EF2 (O’Donnell et al. 1998), respectively. Sequences of ITS (545 bp, GenBank Accession No. MT811812) and TEF1-α (707 bp, GenBank Acc. No. MT856659) for isolate Fa-25 were 100% and 99.72% identical to those of F. asiaticum strains MSBL-4 (ITS, GenBank Acc. MT322117.1) and Daya350-3 (TEF1-α, GenBank Acc. KT380124.1) in GenBank, respectively. A phylogenetic tree was established based on the TEF1-α sequences of Fa-25 and other Fusarium spp., and Fa-25 was clustered with F. asiaticum. Thus, both morphological and molecular characterizations supported the isolate as F. asiaticum. To confirm the pathogenicity, mycelium agar plugs (6 mm in diameter) removed from the colony margin of a 2-day-old culture of strain Fa-25 were used to inoculate melon fruits. Before inoculation, healthy melon fruits were selected, soaked in 2% NaClO solution for 2 min, and washed in sterile water. After wounding the melon fruits with a sterile needle, the fruits were inoculated by placing mycelium agar plugs on the wounds, and mock inoculation with mycelium-free PDA plugs was used as control. Five fruits were used in each treatment. The inoculated and mock-inoculated fruits were incubated at 25°C with high relative humidity. Symptoms were observed on all inoculated melon fruits 10 days post inoculation, which were similar to those naturally infected fruits, whereas the mock-inoculated fruits remained symptomless. The fungus re-isolated from the diseased fruits resembled colony morphology of the original isolate. The experiment was conducted three times and produced the same results. To our knowledge, this is the first report of fruit rot of melon caused by F. asiaticum in China.


Plant Disease ◽  
2021 ◽  
Author(s):  
Oliul Hassan ◽  
Taehyun Chang

In South Korea, ovate-leaf atractylodes (OLA) (Atractylodes ovata) is cultivated for herbal medicine. During May to June 2019, a disease with damping off symptoms on OLA seedlings were observed at three farmer fields in Mungyeong, South Korea. Disease incidence was estimated as approximately 20% based on calculating the proportion of symptomatic seedlings in three randomly selected fields. Six randomly selected seedlings (two from each field) showing damping off symptoms were collected. Small pieces (1 cm2) were cut from infected roots, surface-sterilized (1 minute in 0.5% sodium hypochlorite), rinsed twice with sterile water, air-dried and then plated on potato dextrose agar (PDA, Difco, and Becton Dickinson). Hyphal tips were excised and transferred to fresh PDA. Six morphologically similar isolates were obtained from six samples. Seven-day-old colonies, incubated at 25 °C in the dark on PDA, were whitish with light purple mycelia on the upper side and white with light purple at the center on the reverse side. Macroconidia were 3–5 septate, curved, both ends were pointed, and were 19.8–36.62 × 3.3–4.7 µm (n= 30). Microconidia were cylindrical or ellipsoid and 5.5–11.6 × 2.5–3.8 µm (n=30). Chlamydospores were globose and 9.6 –16.3 × 9.4 – 15.0 µm (n=30). The morphological characteristics of present isolates were comparable with that of Fusarium species (Maryani et al. 2019). Genomic DNA was extracted from 4 days old cultures of each isolate of SRRM 4.2, SRRH3, and SRRH5, EF-1α and rpb2 region were amplified using EF792 + EF829, and RPB2-5f2 + RPB2-7cr primer sets, respectively (Carbone and Kohn, 1999; O'Donnell et al. 2010) and sequenced (GenBank accession number: LC569791- LC569793 and LC600806- LC600808). BLAST query against Fusarium loci sampled and multilocus sequence typing database revealed that 99–100% identity to corresponding sequences of the F. oxysporum species complex (strain NRRL 28395 and 26379). Maximum likelihood phylogenetic analysis with MEGA v. 6.0 using the concatenated sequencing data for EF-1α and rpb2 showed that the isolates belonged to F. oxysporum species complex. Each three healthy seedlings with similar sized (big flower sabju) were grown for 20 days in a plastic pot containing autoclaved peat soil was used for pathogenicity tests. Conidial suspensions (106 conidia mL−1) of 20 days old colonies per isolate (two isolates) were prepared in sterile water. Three pots per strain were inoculated either by pouring 50 ml of the conidial suspension or by the same quantity of sterile distilled water as control. After inoculation, all pots were incubated at 25 °C with a 16-hour light/8-hour dark cycle in a growth chamber. This experiment repeated twice. Inoculated seedlings were watered twice a week. Approximately 60% of the inoculated seedlings per strain wilted after 15 days of inoculation and control seedlings remained asymptomatic. Fusarium oxysporum was successfully isolated from infected seedling and identified based on morphology and EF-1α sequences data to confirm Koch’s postulates. Fusarium oxysporum is responsible for damping-off of many plant species, including larch, tomato, melon, bean, banana, cotton, chickpea, and Arabidopsis thaliana (Fourie et al. 2011; Hassan et al.2019). To the best of our knowledge, this is the first report on damping-off of ovate-leaf atractylodes caused by F. oxysporum in South Korea. This finding provides a basis for studying the epidemic and management of the disease.


Plant Disease ◽  
2021 ◽  
Author(s):  
Tingting Zhu ◽  
Linxuan Li ◽  
Antonios Petridis ◽  
George Xydis ◽  
Maozhi Ren

Ligusticum chuanxiong (known as Chuanxiong in China) is a traditional edible-medicinal herb, which has been playing important roles in fighting against COVID-19 (Ma et al. 2020). In March 2021, we investigated stem rot of Chuanxiong in six adjacent fields (~100 ha) in Chengdu, Sichuan Province, China. The disease incidence was above 5% in each field. Symptomatic plants showed stem rot, watersoaked lesions, and blackening with white hyphae present on the stems. Twelve symptomatic Chuanxiong plants (2 plants/field) were sampled. Diseased tissues from the margins of necrotic lesions were surface sterilized in 75% ethanol for 45 s, and 2% NaClO for 5 min. Samples were then rinsed three times in sterile distilled water and cultured on potato dextrose agar (PDA) at 25ºC for 72 h. Fourteen fungal cultures were isolated from 18 diseased tissues, of which eight monosporic isolates showed uniform characteristics. The eight fungal isolates showed fluffy white aerial mycelia and produced yellow pigments with age. Mung bean broth was used to induce sporulation. Macroconidia were sickle-shaped, slender, 3- to 5-septate, and averaged 50 to 70 μm in length. Based on morphological features of colonies and conidia, the isolates were tentatively identified as Fusarium spp. (Leslie and Summerell 2006). To identify the species, the partial translation elongation factor 1 alpha (TEF1-α) gene was amplified and sequenced (O’Donnell et al. 1998). TEF1-α sequences of LCSR01, LCSR02 and LCSR05 isolates (GenBank nos. MZ169386, MZ169388 and MZ169387) were 100%, 99.72% and 99.86% identical to that of F. asiaticum strain NRRL 26156, respectively. The phylogenetic tree based on TEF1-α sequences showed these isolates clustered with F. asiaticum using Neighbor-Joining algorithm. Furthermore, these isolates were identified using the specific primer pair Fg16 F/R (Nicholson et al. 1998). The results showed these isolates (GenBank nos. MZ164938, MZ164939 and MZ164940) were 100% identical to F. asiaticum NRRL 26156. Pathogenicity test of the isolate LCSR01 was conducted on Chuanxiong. After wounding Chuanxiong stalks and rhizomes with a sterile needle, the wounds were inoculated with mycelia PDA plugs. A total of 30 Chuanxiong rhizomes and stalks were inoculated with mycelia PDA plugs, and five mock-inoculated Chuanxiong rhizomes and stalks served as controls. After inoculation, the stalks and rhizomes were kept in a moist chamber at 25°C in the dark. At 8 days post inoculation (dpi), all inoculated stalks and rhizomes exhibited water-soaked and blackened lesions. At 10 dpi, the stalks turned soft and decayed, and abundant hyphae grew on the exterior of infected plants, similar to those observed in the field. No disease symptoms were observed on the control plants. The pathogen was re-isolated from the inoculated tissues and the identity was confirmed as described above. Ten fungal cultures were re-isolated from the 10 inoculated tissues, of which nine fungal cultures were F. asiaticum, fulfilling Koch’s postulates. To our knowledge, this is the first report of F. asiaticum causing stem rot of Chuanxiong in China. Chuanxiong has been cultivated in rotation with rice over multiple years. This rotation may have played a role in the increase in inoculum density in soil and stem rot epidemics in Chuanxiong. Diseased Chuanxiong may be contaminated with the mycotoxins produced by F. asciaticum, 3-acetyldeoxynivalenol or nivalenol, which may deleteriously affect human health. Therefore, crop rotations should be considered carefully to reduce disease impacts.


Plant Disease ◽  
2021 ◽  
Author(s):  
Francisco Beluzán ◽  
Diego Olmo ◽  
Maela León ◽  
Paloma Abad-Campos ◽  
Josep Armengol

Nectarine (Prunus persica (L.) Batsch var. nucipersica (Suckow) C. K. Schneid.) is a fruit crop widely cultivated throughout the Mediterranean basin. In Spain, it is mainly grown in eastern regions of the country. In March 2018, 5-year-old nectarine trees showing twig canker symptoms were observed after a rainy spring period in a 0.5 ha orchard located at Alaior, Menorca island (Spain). Cankers were frequent on affected trees (approximately, 80% of the total trees), thus leading to shoot blight. Ten twig segments of one-year old wood with cankers were cut, washed under running tap water, surface disinfected for 1 min in a 1.5% sodium hypochlorite solution and rinsed twice in sterile distilled water. Small pieces (2 mm) of affected tissues were taken from the margin of the cankers and plated on potato dextrose agar (PDA) supplemented with 0.5 g/L of streptomycin sulphate (PDAS). The plates were then incubated at 25 ºC in the dark for 7 to 10 d. Actively growing colonies were first hyphal-tipped and then transferred to PDA and 2% water agar supplemented with sterile pine needles and incubated at 21-22ºC under a 12h/12h near UV / darkness cycle during 21 d (León et al. 2020). Colonies were white at first, becoming light cream, with visible solitary and aggregate pycnidia at maturity. Alpha conidia were aseptate, fusiform, hyaline, multi-guttulated (mean ± SD = 7.4 ± 0.7 × 2.8 ± 0.4 µm, n = 100). Beta and gamma conidia were not observed. The morphological and cultural characteristics of the isolates were congruent with those of Diaporthe spp. (Gomes et al. 2013). The ITS1-5.8S-ITS2 (ITS) region and fragments of β-tubulin (tub2), the translation elongation factor 1-alpha (tef1-α) gene regions, histone H3 (his3) and calmodulin (cal) genes of representative isolate DAL-59 were amplified and sequenced (Santos et al. 2017). The BLASTn analysis revealed 100% similarity with sequences of D. mediterranea (Synonym D. amygdali) (Hilário et al. 2021) isolate DAL-34 from almond (ITS: MT007489, tub2: MT006686, tef1-α: MT006989, his3: MT007095 and cal: MT006761). Sequences of isolate DAL-59 were deposited in GenBank Database (ITS: MT007491, tub2: MT006688, tef1-α: MT006991, his3: MT007097 and cal: MT006763). Pathogenicity tests were conducted using one-year-old potted plants of nectarine cv. Boreal, which were inoculated with isolate DAL-59. In each plant, a 3 mm wound was made in the center of the main branch (about 30 cm length) with a scalpel. Colonized agar plugs with 3 mm diameter, which were obtained from active 10-day-old colonies growing on PDA, were inserted underneath the epidermis and the wounds sealed with Parafilm. Inoculated plants were incubated in a growth chamber at 23 ºC with 12 h of light per day. Controls were inoculated with uncolonized PDA plugs. There were twelve plants per treatment, which were arranged in a completely randomized design. Five days after inoculation necrosis development was observed in the area of inoculation. Wilting and twig blight symptoms over the lesion occurred 3-wk after inoculation and pycnidia were detected, while the controls remained asymptomatic. Diaporthe amygdali was re-isolated from symptomatic tissues and identified as described above to satisfy Koch’s postulates. To our knowledge, this is the first report of D. amygdali causing twig canker and shoot blight disease on nectarine in Spain.


Plant Disease ◽  
2021 ◽  
Author(s):  
Yue Sun ◽  
Rui Wang ◽  
Kaibin Qiao ◽  
Hongyu Pan ◽  
Fengting Wang ◽  
...  

Bush lily (Clivia miniata) is an important indoor flower. It is the city flower of Changchun City and has important ornamental and medicinal value in China where it is culitvated on an area of 125 hectare. During the summer of 2018, symptoms of a leaf sheath rot disease were observed on bush lily in 103 greenhouses in Changchun city, Jilin Province. The disease incidence ranged from 25 to 60% in 11 surveyed greenhouses. At the early stage, the diseased plants displayed symptoms as initial leaf sheath lesions. Progressively, the whole leaves wilted, and even the plant ultimately died. Once a leaf exhibits leaf sheath lesions, the whole plant’s ornamental value significantly drops. To identify the pathogen, symptomatic leaves were cut into pieces, surface sterilized, placed on potato dextrose agar (PDA) and incubated for 7 days at 25°C in the dark (Cao et al. 2013; the e-Xtra description for details). Fusarium single-spore isolates were obtained from characteristic colonies (Leslie et al. 2006). Two single-spore isolates were selected for further study. The isolates were identified as Fusarium spp. based on microscopic morphology on PDA. Fusarium-like colonies were white to slightly yellow with abundant cottony mycelia. Single or two-celled (single septum) microconidia were reniform or oval, 8.0 to 9.6×4.0 to 6.0m in size. The elongated conidiophores bearing microconidia in monophialides were observed (Summerbell et al. 2002). Macroconidia were abundant, sickle shaped, 18.8 to 34.8×6.4 to 6.8m, with one to three septa (Taylor et al. 2019). For molecular identification, five regions of ITS, EF1-α, RPB1, RPB2 and β-tubulin genes were amplified and sequenced. Sequences of five different regions exhibited at least 97.98% similiarity with the corresponding DNA sequences in F. solani species complex (FSSC) (the e-Xtra description for details). The phylogenetic analysis based on the EF1-α, RPB1, RPB2 and β-tubulin region sequences revealed that the isolated strain in this study was clustered with only F. solani species in the phylogenetic tree for each region. Based on morphological and molecular analysis, the isolated fungal strains were identified as F. solani. Pathogenicity was confirmed by injecting a conidial suspension (106 spores/mL) of the isolated strains in to surface surface-disinfested leaf sheath of 2-year-old potted healthy plants. As a negative control, four plants were injected with sterilized water. All plants were kept in a greenhouse with controlled conditions: 26°C, 50% to 75% relative humidity. The similar rot symptoms were observed on the leaf sheathes in the inoculated plants 30 days after inoculation whereas the control plants remained asymptomatic. The fungi reisolated from the experimental plants were confirmed to be F. solani by morphology and sequences analysis, thus completing Koch’s postulates. To the best of our knowledge, this is the first report of F. solani causing leaf sheath rot of bush lily in China, where this pathogen has been reported to cause rot diseases of other economically important ornamental plants such as Phalaenopsis, Dendrobium according to the U.S. National Fungus Collections (Farr et al. 2020). In recent years, other Fusarium species have been reported to cause rot diseases on bush lily, including F. proliferatum and F. oxysporum (Farr et al. 2020). This study will also provide critical information on the causal agent for growers to implement disease management strategies.


Plant Disease ◽  
2014 ◽  
Vol 98 (7) ◽  
pp. 1002-1002 ◽  
Author(s):  
H.-W. Choi ◽  
S. K. Hong ◽  
Y. K. Lee ◽  
W. G. Kim

In July 2010, flower rot of thread-leaf coreopsis (Coreopsis verticillata) was found in a garden in the Icheon City, Korea. The disease affected about 20 to 50% of a 100 m2 area. The disease was characterized by the appearance of pinkish mycelia on the stigmata and inflorescences of flowers. In some cases, flowers failed to bloom or turned brown before opening fully. Fragments (each 5 × 5 mm) of the symptomatic tissue were surface-sterilized with 1% NaOCl for 1 min, and then rinsed twice in sterilized distilled water. The tissue pieces were placed on water agar (WA) and incubated at 25°C for 4 to 6 days. Twenty-two isolates of Fusarium species were obtained from the diseased flowers. All isolates were identified as Fusarium succisae based on their morphological characteristics on carnation leaf agar (CLA) medium and DNA sequences of the translation elongation factor 1-alpha gene (1). Macroconidia and sporodochia were sparsely produced on CLA medium. Microconidia were abundant, borne in false heads, oval or allantoid and sometimes pyriform, and measured 4.2 to 13 × 2.2 to 5.4 μm. Chlamydospores were absent. The EF-1α gene was amplified from three isolates by PCR assay and the amplification products were sequenced (2). The nucleotide sequences obtained were deposited in GenBank with accession numbers KF514658, KF514659, and KF514660. BLASTn analysis showed 99% homology with the EF-1α sequence of F. succisae NRRL13613 (GenBank Accession No. AF160291). Pathogenicity tests were conducted with inoculation of flowers on Coreopsis verticillata. Spore suspension was prepared by flooding 7-day-old cultures on potato dextrose agar with sterilized 2% (w/v) sugar solution. When the plants started to have buds, the isolates were inoculated by placing one drop (20 μl) of spore suspension (1 × 106 spores ml−1) into the buds. Fifteen buds of the plants were arranged into three replications. The control was treated with sterilized 2% sugar solution. Inoculated plants were kept in a greenhouse at 25/20°C (12 h/12 h). Three weeks after inoculation, the symptoms were observed on buds with mycelial production. Control plants had no mycelia on buds. F. succisae was re-isolated from the inoculated flowers. To our knowledge, this is the first report of flower rot of thread-leaf coreopsis caused by F. succisae. References: (1) J. F. Leslie and B. A. Summerell. The Fusarium Laboratory Manual. Blackwell Publishing, Ames, IA, 2006. (2) K. O'Donnell et al. Proc. Nat. Acad. Sci. 95:2044, 1998.


Plant Disease ◽  
2014 ◽  
Vol 98 (8) ◽  
pp. 1156-1156 ◽  
Author(s):  
H. Zhang ◽  
W. Luo ◽  
Y. Pan ◽  
J. Xu ◽  
J. S. Xu ◽  
...  

Fusarium is an important genus of fungal pathogens that are responsible for devastating diseases, such as Fusarium ear rot on maize, which may result in yield losses and/or mycotoxin contamination. In September 2013, a survey to determine population composition of Fusarium species on maize was conducted at 22 fields in 18 counties in Gansu Province. Maize ears with clear symptoms (with a white to pink- or salmon-colored mold at the ear tip) were collected. Symptomatic seeds were surface-sterilized with 70% ethanol and 10% sodium hypochlorite and rinsed three times with sterile water to eliminate hypochlorite residues. After drying on sterile filter paper, the seeds were placed on potato dextrose agar (PDA) and incubated at 25°C in the dark for 3 days. Mycelium that was characteristic of Fusarium spp. (2) was purified by transferring single spores to fresh PDA. Fusarium species were identified by morphological characteristics (2), multilocus genotyping assay (MLGT) (3), and sequence analysis of the translation elongation factor-1α (TEF) gene. Several Fusarium species were identified and Fusarium verticillioides and F. proliferatum were the predominant species. Based on MLGT, two strains from Chenghong County were identified as F. meridionale with NIV chemotype, a species in F. graminearum species complex (FGSC). Morphological characteristics were also identical to FGSC. Colonies grew rapidly on PDA and produce relatively large amounts of dense mycelia and red pigments. Slender, thick-walled, and moderately curved or straight macroconidia were observed with 5- to 6-septate. Furthermore, conidia on SNA also showed typical characteristics of F. meridionale, as the dorsal and ventral lines were often parallel and gradually curved. Sequences comparison of the partial translation elongation factor (TEF-1α, 644 bp) gene (1) was used to validate these observations. BLASTn analysis with the FUSARIUM-ID database revealed 100% sequence identity to F. meridionale (GenBank Accession No. KJ137017). Thus, both morphological and molecular criteria supported identification of the strains as F. meridionale. A pathogenicity test was performed on Zhengdan958, the maize variety with the largest planted acreage in China. Four days after silk emergence, 2 ml conidial suspension (105 macroconidia/ml) of each isolate were injected into each of 10 maize ears through silk channel. Control plants were inoculated with sterile distilled water. Typical FER symptoms (reddish-white mold) was observed on inoculated ears and no symptoms were observed on water controls. Koch's postulates were fulfilled by re-isolating the same fungus from the infected seeds. F. meridionale was one of the pathogens causing Fusarium head blight on wheat and barley in China and produced nivalenol (4,5) and it also has been isolated from maize in Korea and Nepal. To our knowledge, this is the first report of F. meridionale causing Fusarium ear rot on maize in China. Further studies on biological characteristics such as temperature sensibility and fungicide resistance are needed to gain a better understanding of this new pathogen. References: (1) D. M. Geiser et al. Eur. J. Plant Pathol. 110:473, 2004. (2) J. F. Leslie and B. A. Summerell. The Fusarium Laboratory Manual. Blackwell Publishing, Ames, IA, 2006. (3) T. J. Ward et al. Fungal Genet. Biol. 45:473, 2008. (4) L. Yang et al. Phytopathology 98:719, 2008. (5) H. Zhang et al. Plos one 7:e31722, 2012.


Plant Disease ◽  
2008 ◽  
Vol 92 (7) ◽  
pp. 1132-1132 ◽  
Author(s):  
L. Riccioni ◽  
A. Haegi ◽  
M. Valvassori

Lentil (Lens culinaris Medik.) is a traditional crop in Sicily, Italy. Near Villalba (Caltanissetta), a local lentil landrace, “Lenticchia di Villalba”, is commonly grown. From 2002 to 2004, wilt was observed in five lentil fields (≈1 ha each) at rates from 5 to 20%. Affected plants were yellow and stunted with discoloration in the vascular tissue of stems and crowns. Pieces of brown vascular tissue from stems were disinfested in 2% sodium hypochlorite for 2 min, rinsed with sterile distilled water, placed on potato dextrose agar, and incubated at 23°C. Isolates with morphological characteristics of Fusarium oxysporum Schlecht.:Fr. (2) were consistently recovered from affected plants. For molecular identification of five isolates, the rDNA internal transcribed spacer (ITS) region and a portion of the elongation factor EF-1α were sequenced using ITS5/4 and EF1/2 primers, respectively (1). Two sequences of the ITS region were obtained: a 468-bp sequence from isolates ER1259, ER1260, and ER1275 (submitted as GenBank Accession No. EU159118) and a 483-bp sequence from isolates ER1274 and ER1276 (submitted as GenBank Accession No. EU281661). The two sequences shared 93% similarity. A sequence homology search using the NCBI BLAST program revealed that the first sequence had 100% homology with the ITS sequences of more than 50 F. oxysporum isolates of various formae speciales in GenBank and the second shared 100% homology with the ITS sequences of five isolates of F. redolens Wollenw. (e.g., GenBank Accession No. X94169 of the strain CBS 360.87). Amplification of the EF-1α produced a sequence from isolates ER1274 and ER1276 (submitted as GenBank Accession No. EU281660) with 99 to 100% homology to sequences of F. redolens and a sequence from strains ER1259, ER1275, and ER1260 (submitted as GenBank Accession No. EU281659) with 100% homology to that of more than 50 F. oxysporum strains in GenBank. Although F. redolens and F. oxysporum are morphologically similar, recent molecular studies have shown that they are distinct and phylogenetically distant species (3). On the basis of genetic sequences, isolates ER1274 and ER1276 were identified as F. redolens. These isolates were evaluated for pathogenicity on lentil. For each isolate, 10 2-week-old seedlings of “Lenticchia di Villalba” were inoculated by submerging roots in a suspension of 2.5 × 106 conidia/ml for 10 min. Plants were put into separate tubes containing 70 ml of a nutritional liquid medium (7 ml of HydroPlus Olikani per liter; Yara, Nanterre, France) and incubated in a growth chamber at 20°C with 12 h of light per day. Seedlings dipped in sterile water served as the control treatment. The pathogenicity test was repeated twice. Inoculated seedlings started to wilt 1 week after inoculation and developed root rot and vascular discoloration. After 2 weeks, 70% of the inoculated plants were affected by both isolates and 40 and 10% died when inoculated with ER1274 and ER1276 isolates, respectively. F. redolens was consistently reisolated from the stems of wilted plants. Noninoculated plants remained healthy. Currently, only F. oxysporum f. sp. lentis Vasud. and Sriniv. has been reported as the cause of Fusarium wilt of lentil. To our knowledge, this is the first report of F. redolens as a pathogen on lentil. References: (1) R. P. Baayen et al. Phytopathology 91:1037, 2001. (2) P. E. Nelson et al. Fusarium Species: An Illustrated Manual for Identification. The Pennsylvania State University Press, University Park, 1983. (3) K. O'Donnell et al. Mycologia 90:465, 1998.


Plant Disease ◽  
2012 ◽  
Vol 96 (1) ◽  
pp. 145-145 ◽  
Author(s):  
S. T. Koike ◽  
T. R. Gordon ◽  
S. C. Kirkpatrick

In 2010 in Yuma, AZ, field-grown fennel (Foeniculum vulgare, Apiaceae) exhibited previously undescribed disease symptoms. The lower stems in contact with soil developed a brown decay and leaves on these stems became chlorotic. White mycelium and orange sporodochia were observed on affected tissues near the soil line. Diseased stems later wilted, died, and resulted in reduced quality of the fennel; these plants were not harvested. Disease distribution was patchy and prevalence was approximately 5%. Symptomatic tissues were surface sterilized in a dilute (1%) bleach solution for 3 min and tissues from the margins of the decay were placed into petri plates containing acidified corn meal agar (2 ml of 25% lactic acid/liter). Isolations consistently resulted in the recovery of a presumptive Fusarium species. Isolates were transferred to carnation leaf agar and incubated at 22°C under fluorescent lights for 10 days. Morphologies of all isolates were identical, with macroconidia being long and slender, slightly curved, with elongated, bent apical cells and notched basal cells. Conidia were borne on monophialides. Microconidia were sparse and chlamydospores were not observed. For two isolates, a portion of the translation elongation factor 1-alpha gene (TEF) was amplified with primers ef1 and ef2 (3). Based on a comparison of 668 base pairs, both isolates had the same sequence, which differed by one base pair from an accession (GQ915502.1) of Fusarium avenaceum in GenBank. The same single base pair also separated the two fennel isolates from an isolate of F. avenaceum (GL 13) previously recovered from Eustoma grandiflorum (=Lisianthus russellianus) (2). Thus, both morphological and molecular criteria support identification of the recovered fungus as F. avenaceum (Fries) Saccardo. Partial TEF sequences were deposited in GenBank (Accession Nos. JN254784, JN254785, and JN254786 for the two fennel isolates and GL 13, respectively). All isolates are archived in the Department of Plant Pathology at University of California, Davis. Pathogenicity was tested by cutting shallow slits into fennel stems, inserting one colonized agar plug into each cut, and wrapping the stems with Parafilm. Five isolates from fennel were tested on 10 stems each. Control plants were inoculated with uncolonized agar plugs. Plants were maintained at 24 to 26°C in a greenhouse. After 6 to 8 days, a brown decay developed on 70 to 90% of Fusarium-inoculated stems at the points of inoculation. Foliage later became chlorotic and F. avenaceum was recovered from all symptomatic stems. Control plants were symptomless. The experiment was completed two times and results were the same. In addition, F. avenaceum isolate GL13 from E. grandiflorum (2) was inoculated onto fennel plants with the same method. However, these inoculated plants remained symptomless. To our knowledge, this is the first report of a stem and crown rot disease of fennel caused by F. avenaceum. Apparently, the only other published account of a Fusarium disease of fennel is root rot caused by F. solani (1). The inability of the Eustoma isolate of F. avenaceum to cause disease in fennel suggests that these two crown rot pathogens may have restricted host ranges. References: (1) J. H. Gupta and V. P. Srivastava. Indian J. Mycol. Plant Pathol. 8:206, 1979. (2) S. T. Koike et al. Plant Dis. 80:1429, 1996. (3) K. O'Donnell et al. Proc. Nat. Acad. Sci. U.S.A. 95:2044, 1998.


Plant Disease ◽  
2021 ◽  
Author(s):  
Andjelka Prokić ◽  
Tamara Marković ◽  
Jelena Menković ◽  
Milan Ivanovic ◽  
Aleksa Obradoviċ

Arugula (Eruca vesicaria subsp. sativa (Miller) Thell., syn. Brassica eruca L.), is an annual cruciferous crop that is increasingly grown for fresh consumption in Serbia. In November 2018, a few detached leaves of cultivated arugula originating from a local producer, showing necrotic lesions, were observed in a fresh vegetable market in Belgrade, Serbia. Information about the disease incidence and severity was not available. Intensity of the observed symptoms was low, but it could be a consequence of the produce quality selection for the market. The leaves developed irregular chlorotic lesions starting from the leaf edge, and tissue within some of them turned dark brown and necrotic (Fig. 1a). From the lesions on different leaves, smooth, bright yellow pigmented, round and opalescent bacterial colonies were isolated on nutrient agar (NA) medium after 72 h of incubation at 26°C. Six bacterial isolates, obtained from three leaf subsamples which induced hypersensitive reaction in tobacco leaves (Nicotiana tabacum L. cv. Samsun), were selected for further studies. On yeast - dextrose – CaCO3 medium, the strains formed characteristic creamy yellow, mucoid, opaque and convex colonies. All isolates were Gram-negative, strictly aerobic, non-fluorescent and catalase positive, did not produce oxidase nor arginine dehydrolase, and did not show pectynolitic activity on potato tuber slices. They hydrolyzed starch, gelatine and esculin, used glucose and sucrose, but not arabinose as a carbon source, and did not reduce nitrates. They grew at 36°C, and tolerated 5% NaCl and 0.02% triphenyl-tetrazolium chloride (Lelliott and Stead, 1987). These growth characteristics were similar as for the reference Xanthomonas campestris pv. campestris (Xcc) strain KFB 105, used in all tests as a positive control (Obradović et al., 2000). The isolates were further characterized by polymerase chain reaction (PCR) using primers DLH120/DLH125, specific for the hrpF gene region of X. campestris according to Berg et al. (2005). Specific DNA fragment of 619 bp was amplified for all tested isolates. Amplification and partial sequencing of the gyrB gene of four isolates was performed using set of primers described by Parkinson et al. (2007). All obtained partial gyrB sequences were identical to each other. According to BLAST analysis (GenBank Acc. Nos. MW508894 - MW508897) they shared 100% of sequence identity with different Xcc strains and 99.5 % with the X.c. pv. raphani pathotype strain, deposited in the NCBI GenBank database. Pathogenicity of the isolates was tested by spraying leaves of 3-week old E. sativa seedlings grown in a commercial potting mix in a greenhouse, with a 24 h-old bacterial culture suspended in sterile distilled water (107 CFU/ml). Xcc strain KFB 105 was used as positive and sterile distilled water as negative control. Inoculated plants were incubated under plastic bags for 48 h and further maintained in a greenhouse at approx. 28°C. On inoculated plants, chlorotic lesions, spreading from the leaf margins, further coalescing into irregular, V-shaped tissue necrosis associated with blackening of veins, developed up to two weeks after inoculation (Fig. 1b, c). The colonies reisolated from symptomatic leaves were identified using PCR, as described above. Based on studied characteristics, all six isolates associated with arugula leaf lesions in Serbia belong to a clonal population. They were identified as X. campestris pv. campestris, the causal agent of black rot, a major disease affecting crucifers, including arugula worldwide (Romero et al., 2008; Rosenthal, et al., 2018). So far, it has been described on Brassica oleracea and B. napus in Serbia (Obradović et al., 2001; Popović et al., 2019). This is the first report of Xcc infecting arugula in this country. The severity of the symptoms developed on artificially inoculated plants indicated significant potential of the pathogen to affect arugula crop in conditions favoring infection. Being a minor crop, accurate information about severity of arugula diseases in Serbia is not available. Lack of crop rotation and close proximity of other Xcc host species on a farm could contribute to further spreading of this problem. Follow up of this arugula disease should reveal the distribution, population structure and genetic diversity of Xcc strains affecting this crop in Serbia.


Sign in / Sign up

Export Citation Format

Share Document