Next Article in Journal
Biodiversity of Fungi in Freshwater Ecosystems of Italy
Next Article in Special Issue
Detection Method of Fungal Spores Based on Fingerprint Characteristics of Diffraction–Polarization Images
Previous Article in Journal
Multivariate Statistical Techniques: A New Approach to Identify the Commercial Properties of Mixtures of Flours of Lentinula edodes and Cocoa Pod Husk
Previous Article in Special Issue
Host Range and Loop-Mediated Isothermal Amplification Detection of Globisporangium sylvaticum from Guizhou, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Characterization and Pathogenicity of Alternaria spp. Associated with Black Rot of Sweet Cherries in Italy

1
Department of Agricultural, Forest and Food Sciences (DISAFA), University of Torino, Largo Braccini 2, 10095 Grugliasco, TO, Italy
2
AGROINNOVA—Interdepartmental Centre for Innovation in the Agro-environmental Sector, University of Torino, Largo Braccini 2, 10095 Grugliasco, TO, Italy
*
Author to whom correspondence should be addressed.
J. Fungi 2023, 9(10), 992; https://doi.org/10.3390/jof9100992
Submission received: 4 September 2023 / Revised: 30 September 2023 / Accepted: 5 October 2023 / Published: 7 October 2023
(This article belongs to the Special Issue Monitoring, Detection and Surveillance of Fungal Plant Pathogens)

Abstract

:
Black rot is limiting the production of sweet cherries in Italy. Dark brown to black patches and sunken lesions on fruits are the most common symptoms of Alternaria black rot on sweet cherry fruits. We isolated 180 Alternaria spp. from symptomatic cherry fruits ‘Kordia’, ‘Ferrovia’, and ‘Regina’ harvested in Northern Italy, over three years, from 2020 to 2022. The aim was to identify and characterize a selection of forty isolates of Alternaria spp. based on morphology, pathogenicity, and combined analysis of rpb2, Alt-a1, endoPG and OPA10-2. The colonies were dark greyish in the center with white margins. Ellipsoidal or ovoid shaped conidia ranging from 19.8 to 21.7 μm in length were observed under a microscope. Based on the concatenated session of four gene regions, thirty-three out of forty isolates were identified as A. arborescens species complex (AASC), and seven as A. alternata. Pathogenicity was evaluated on healthy ‘Regina’ sweet cherry fruits. All the tested strains were pathogenic on their host. This study represents the first characterization of Alternaria spp. associated with black rot of cherries in Italy and, to the best of our knowledge, it is also the first report of AASC as an agent of black rot of sweet cherries in Italy.

1. Introduction

Sweet cherry (Prunus avium L.) is an economically important stone fruit which belongs to the genus Prunus within the Rosaceae family. Italy is the seventh sweet cherry producer in the world after Turkey, United States of America, Chile, Iran, Uzbekistan, and Spain and the second in Europe [1]. In Italy, sweet cherry is cultivated on an area of 28,609 ha, with an annual production of approximately 107,905 t [2].
Sweet cherries are highly perishable fruit, with a shelf life of 7–14 days [3] and they are susceptible, during cold storage, to decay caused by postharvest pathogens, including Monilinia spp., agent of brown rot; Botrytis cinerea, agent of gray mold; Penicillium expansum, agent of blue mold; and Rhizopus spp., agent of soft rot [4,5,6,7]. Cherry production could be threatened also by black rot caused by Alternaria species, which is emerging as a major fungal disease [8,9,10]. Symptoms of black rot of Alternaria appear as dark brown to black patches on the outer surface of fruits. These patches gradually increase in size and surface and, under high-humidity conditions, may develop white to light brown, fluffy and moldy growth, and ultimately can cause complete fruit rot [8,11,12].
The genus Alternaria was originally described by Nees von Esenbeck in 1816; it is part of the Pleosporaceae family [13] and includes endophytic, pathogenic, and saprobic species [14,15,16,17]. Alternaria spp. are distributed all over the world and may infect over 4000 plant species [13,15,17], such as vegetables, cereals and fruit trees in field and during storage [18,19,20,21,22,23]. Black rot caused by A. alternata (Fr.) Keissl. was recently reported on cherries in Chile [9] and in China [8,10]. A. alternata is also associated with brown to black spots on cherry leaves [24,25]. Alternaria rots and spots are becoming more and more frequent on cherry and on other fruit crops due to climate change, characterized by the increase of average temperature, high-humidity conditions and stressed plants, weakened by abiotic stresses. Biological and environmental factors are causing a shift in the fruit microbiological ecology, which is influencing the pre- and postharvest development of Alternaria spp. [26].
In the past, Alternaria spp. were identified based on macro- and micro-morphological characteristics [16,27]. Nowadays, DNA-based molecular techniques are used to identify Alternaria at species level to avoid morphological variations depending on the environmental conditions [13,15,28,29,30,31,32]. Different molecular approaches have been used for the identification of Alternaria section, resulting in several taxonomic revisions [13,16,28,33,34,35,36,37]. Most small-spored Alternaria species with concatenated conidia belong to the section Alternaria [16], and A. alternata (Fr.) Keissl. and A. arborescens Simmons are important representative plant pathogens of this section [16]. To the best of our knowledge, Alternaria spp. were not previously reported and characterized as agents of black rot of sweet cherries in Italy.
The aims of this study are to isolate and identify the Alternaria species associated with black rot of sweet cherries, based on morphological and phylogenetic analysis, and to confirm their pathogenicity and virulence on sweet cherries.

2. Materials and Methods

2.1. Sampling and Isolation

A monitoring of postharvest diseases was conducted in sweet cherry harvested from orchards located in Piedmont, Northern Italy, from 2020 to 2022. A total of 180 isolates were isolated from symptomatic cherry fruits belonging to the varieties ‘Regina’, ‘Kordia’, and ‘Ferrovia’ in the packinghouses of Piedmont (Figure 1A,B). Sweet cherry fruits were surface disinfected with 1% sodium hypochlorite for 1 min, rinsed in sterile water for 1 min, and dried on sterile filter paper. Then five pieces of black rotten fruits were cut at the margin between healthy and infected tissues and plated on potato dextrose agar (PDA, VWR international, Leuven, Belgium) [29] containing streptomycin (0.025 g/L). The PDA plates were incubated at 22 ± 1 °C for 3 days. Pure cultures were obtained by transferring the mycelium plug from the edge of the colonies and placed in fresh PDA plates. Isolates used in this study were maintained and kept at −80 °C in the culture collection of the University of Turin, Torino, Italy (Table 1).

2.2. Micro and Macro-Morphological Characteristics

Alternaria isolates were plated on PDA medium (in triplicate) and incubated at 25 ± 1 °C for the macro-morphological characteristics (color, margin, diameter, and texture), according to Simmons et al. [27]. Mean radial growth was measured after 7 days of incubation (Table 1). For the microscopic features (conidia and conidiophore) the isolates were grown onto Potato Carrot Agar (PCA, HiMedia Laboratories, Mumbai, India) under 12 h light and 12 h dark cycle for 15 days [27]. Thirty conidia per isolate were examined using an Eclipse 55i microscope (Nikon, Tokyo, Japan) at 40× magnification.

2.3. DNA Extraction and PCR Amplification

Genomic DNA of 40 Alternaria isolates was extracted with an E.Z.N.A. Fungal DNA mini kit (Omega Bio-tek, Darmstadt, Germany) from 0.1 g mycelium of 7-day-old culture grown on PDA (VWR international) according to the manufacturer’s instructions. The quality and concentration of extracted DNA was determined using NanoDrop 2000 spectrophotometer (Thermo scientific, Wilmington, DE, USA). The primers Alt-for and Alt-rev [38] were used to amplify part of Alternaria major allergen gene (Alt-a1). The partial endopolygalacturonase gene (endoPG) was amplified using PG3 and PG2b primers [39,40]. The primer sets RPB2-5f2 and fRPB2-7cr [41,42] were used to amplify the part of RNA polymerase second largest subunit (rpb2). The primers OPA 10-2R and OPA 10-2L [28] were used to amplify part of an anonymous gene region (OPA10-2). The amplification of all four loci were performed according to PCR amplified conditions described by Prencipe et al. [43]. The PCR cycling conditions adopted for rpb2 and Alt-a1 were described by Woudenberg et al. [44], and for endoPG and OPA 10-2, by Andrew et al. [28]. The amplification products were analyzed on 1% agarose (VWR International, Milan, Italy) after staining with GelRedTM. PCR products were purified with the PCR Purification Kit (QIAquick®, Hilden, Germany) following manufacturer instructions, before sequencing by Macrogen Europe B. V. (Amsterdam, The Netherlands).

2.4. Phylogenetic Analysis

Phylogenetic analysis was performed using sequences generated in this study and reference sequences of Alternaria spp. [32] (Table S1). After cutting the trimmed regions in Geneious v. 11.1.5 program (Auckland, New Zealand) and manual correction in MEGA v. 7, a dataset 2229 bp of 281 bp for Alt-a1, 479 bp for endoPG, 634 bp for OPA10-2, and 835 bp for rpb2 was obtained. The sequences were aligned using CLUSTALW in MEGA v. 7 [45]. Phylogenetic analysis was performed using the concatenated dataset (rpb2, OPA 10-2, Alt-a1 and endoPG) for the identification of Alternaria isolates at species level. The phylogeny was based on maximum parsimony (MP) and Bayesian inference (BI) used in a concatenated analysis. For BI, the best fit evolutionary model for each partitioned locus was estimated using MrModeltest v. 2.3 [46] and incorporated into the analysis. MrBayes v. 3.2.5 [47] was used to generate phylogenetic trees under optimal criteria per partition. The Markov chain Monte Carlo (MCMC) analysis used four chains and started from a random tree topology. The heating parameter was set at 0.2 and trees were sampled every 1000 generations. The analysis stopped when the average standard deviation of split frequencies was below 0.01. The MP analysis was performed using Phylogenetic Analysis using Parsimony (PAUP) v. 4.0b10 [48]. Phylogenetic relationships were estimated by heuristic searches with 100 random addition sequences. Tree bisection-reconnection was used, with the branch swapping option set on ‘best trees’, with all characters equally weighted and alignment gaps treated as fifth state. Tree length (TL), consistency index (CI), retention index (RI) and rescaled consistence index (RC) were calculated for parsimony, and the bootstrap analysis [49] was based on 1000 replications. Sequences generated in this study were deposited in GenBank (Table S1).

2.5. Pathogenicity Assay

The same isolates used for molecular analysis were used for the pathogenicity test. Conidia of tested Alternaria isolates, produced on 21-day-old PDA cultures incubated at 22 ± 1 °C under a 12 h photoperiod, were used to obtain conidial suspensions, which were prepared by scrapping off the conidia from the surface of PDA plates with sterile water and 10 μL of Tween 20, which were then filtered through four layers of sterile gauze, as described by Prencipe et al. [43]. The concentration of the collected spore suspension was adjusted to 105 conidia/mL. Pathogenicity tests were conducted on healthy fruits of cherries ‘Regina’ using the method described by Ahmad et al. [8]. Fruit surfaces were disinfected with 1% sodium hypochlorite for 1 min, washed with distilled water and air dried for 5 min. The experiment was conducted using ten fruits per isolate. Each fruit was inoculated with 1 µL of conidial suspension by creating one wound (1 mm2) with a sterile needle. Control fruits were treated with sterilized distilled water. Inoculated fruits were placed in plastic trays and covered with a plastic film and incubated at 20 ± 2 °C until symptoms appeared. After 10 days of inoculation, rot diameters were measured. Re-isolations were conducted from the inoculated fruits as described above. Each isolate was tested twice.

2.6. Statistical Analysis

The statistical analysis was performed using SPSS software (IBM SPSS Statistics v. 28.0.1.0). Rot diameters obtained in the pathogenicity test were subjected to the analysis of variance (ANOVA) and the mean values were separated by Tukey test (p ≤ 0.05).

3. Results

3.1. Fungal Isolation, Identification, and Morphological Characterization

Typical symptoms of dark brown to black patches and sunken lesions were observed on sweet cherry fruits. Out of 180 isolates obtained from symptomatic sweet cherry fruits, 40 isolates were identified as Alternaria spp. based on micro and macro-morphological observations [27] (Table 1).
Isolates of Alternaria spp. isolated from infected sweet cherry fruits were identified based on their morphological characteristics [27]. All the Alternaria isolates formed aerial mycelium on PDA, and most of the colonies were dark greyish in the center with white margins after 7 days of incubation at 25 ± 1 °C under 12 h light and dark cycle. The mycelium becomes dark brown after 10 to 14 days. The mean radial growth of colony was 4.96 ± 0.52 cm (Table 1, Figure 1C,D). The conidiophores were light brown and the conidia were ellipsoidal or ovoid with 1–4 transverse septa, a mean of 21.01 ± 0.39 μm in length and 11.11 ± 0.26 μm in width (Table 1, Figure 1E,F).

3.2. Phylogenetic Analysis

The sequences obtained in this study were subjected to a BLAST search in NCBI’s GenBank (https://blast.ncbi.nlm.nih.gov/Blast.cgi; accessed on 10 January 2023) nucleotide database for preliminary identification. A multilocus phylogenetic analysis was conducted using forty isolates of Alternaria spp. based on the sequences from four genes (rpb2, OPA 10-2, Alt-a1 and endoPG) and sixty-two reference sequences [32], including the outgroup Alternaria nobilis. A total of 184 nucleotides were parsimony-informative, 276 were variable and parsimony-uninformative, and 1769 were constant. A maximum of 1000 equally maximum parsimony (MP) trees were saved (Tree length = 771, CI = 0.642, RI = 0.864 and RC = 0.555). Bootstrap support values from the MP analysis are incorporated on the Bayesian tree in Figure 2. For the Bayesian analyses (BI), MrModeltest suggested that all partitions should be analyzed with Dirichlet state frequency distributions. The following models were recommended by MrModeltest and used: K80 for Alt-a1, SYM + G for endoPG and rpb2, and K80 + I + G for OPA 10-2. In the BI, the Alt-a1 partition had 80 unique site patterns, the endoPG partition had 92 unique site patterns, the OPA 10-2 partition had 136 unique site patterns, the rpb2 partition had 155 unique site patterns and the analysis ran for 6,595,000 generations, resulting in 6596 trees, of which 4947 trees were samples used to calculate the posterior probabilities. Based on multilocus phylogenetic analysis, thirty-three out of forty strains clustered with AASC, whereas the remaining seven strains belonged to A. alternata (Figure 2 and Table 1).

3.3. Pathogenicity Assay

The tested Alternaria strains were pathogenic on artificially inoculated sweet cherries. Symptoms appeared as the development of black rotted spots and sunken lesions all over the inoculation point after 7 days post inoculation (Figure 3). Lesions increased in diameter with disease progress and morphologically appeared similar to those observed in naturally infected fruits (Figure 3). In our pathogenicity, no significant differences were observed among the Alternaria strain/species tested combinations. Only two strains (Ch4 and Ch17) from AASC showed significant differences: the isolate Ch-17 showed the highest rot diameter (17.0 ± 2.13 mm), whereas the isolate Ch-4 showed the lowest rot diameter (8.43 ± 2.03 mm) (Table 1). To fulfill Koch’s postulates, re-isolation was carried out on all the symptomatic inoculated fruits, and isolates were identified as Alternaria spp. using the rpb2 gene [41]. Healthy control fruits did not develop any symptoms.

4. Discussion

In our study, isolates of Alternaria spp. associated with black rot of sweet cherries in Italy were identified as A. alternata and A. arborescens species complex (AASC) based on morphological and phylogenetic analysis at species level. During three years of investigation, all orchards showed the presence of A. alternata and AASC. In the present study, AASC was the predominant fungus isolated from rotted cherries. Along with Alternaria spp., we found the co-occurrence of other secondary fungal pathogens associated with rotten fruits, including Cladosporium spp., Monilinia spp., Penicillium spp. and Botrytis cinerea. In several studies, AASC has been reported as a causal agent of black rot of mandarin [50], blueberry [51], pomegranate fruit [52], and Japanese plum [53]; core rot of apple [54], leaf blotch and fruit spot diseases of apple [55]; black spots on fruit and leaves of pear [40]; and brown spots on sweet orange and lemon in Italy [56]. No information is reported about the association of AASC with sweet cherries. In contrast, A. alternata has been reported as being associated with postharvest rot of sweet cherry [15], black spot of cherry fruits in China [10], and leaf spot disease of sweet cherry in Greece [25] and Turkey [24]. Furthermore, A. alternata has been reported as the main causal agent of black rot of cherry tomatoes [57], and fruit rot on strawberry [51,58], mandarin [50], pomegranate fruit [52,59], Japanese plum [53], Solanum muricatum [60], and opium poppy [61]. Most studies did not report Alternaria spp. as a main pathogen, but in our study, we isolated Alternaria spp. as a major pathogen from sweet cherry fruits. In our study, all isolates of Alternaria spp. were identified based on cultural and morphological characteristics, as they all formed conidia, similar to the observations of Prencipe et al. [43] and Şimşek et al. [24]. Our Alternaria isolates were dark grayish with white margins and formed aerial mycelium on PDA, with an average size of conidia (19.7–21.7 × 10.8–11.6) on PCA. These morphological and cultural characteristics were similar to previous studies [40,62], but different from those described by Şimşek et al. [24]. This phenomenon could be due to the morphological plasticity of Alternaria spp. [17] and the fact that the morphology of conidia is dependent on conidial age and culture conditions [37].
The taxonomy of small-spored Alternaria spp. suffered from controversies because Alternaria spp. shared similar morphological characteristics and the size of conidia [33,50,63]. Molecular-based assays could be used for the correct identification of Alternaria spp. along with morphological characteristics [32,33,64]. Molecular analysis also had some challenges to overcome, as Alternaria section Alternaria cannot be recognized using standard genetic loci due to the little or no variation in molecular markers [13,28,36,43]. Previous studies suggested that the identification criteria for low resolution of species delimitation in small-spored Alternaria spp. are only significant when employing the combination of different genes together [14,17,27]. To overcome the issues, the phylogeny of Alternaria sections was solved by using nine gene regions (Alt a 1, endoPG, gapdh, ITS, LSU, OPA10-2, rpb2, SSU and tef1) by Woudenberg et al. [32]. The concatenated session of six gene regions of Alt a 1, endoPG, ITS, OPA10-2, rpb2, and tef1 was able to separate AASC from A. alternata [14]. In our study, we excluded the slowly evolving gene tef 1 and used the genes proposed by Prencipe et al. [43]. Moreover, previous studies confirmed that alt a1 [65] and OPA10-2 [43] were sufficient to separate the A. alternata from AASC. The four loci (Alt-a1, endoPG, opa10-2 and rpb2) used for the phylogenetic analysis performed in this research allowed us to identify thirty-three isolates as members of AASC, and seven isolates as A. alternata. The combined analysis of phylogenetic trees showed similarity with previous studies [14,43]. However, the phylogeny of our study obtained from the concatenated session of four genes has low bootstrap value in agreement with previous studies [35,53]. Considering our phylogenetic analysis results, the inclusion of more genes, such as gaphd, LSU and SSU, could increase the discrimination power, as previously proposed by Zhang et al. [17].
In our pathogenicity test, sweet cherry fruits were wounded and inoculated with conidial suspension of Alternaria strains. Pathogenicity results showed that all the Alternaria strains were pathogenic, had high virulence, and produced irregular lesions when fruits were inoculated by conidial suspension. According to previous studies, A. alternata species were pathogenic on sweet cherry fruits inoculated with conidial suspension and lesions were observed [8]. Additionally, Prencipe et al. [43] reported that isolates of A. alternata and AASC were pathogenic when inoculated with conidial suspension on wounded European pear. In previous studies, Alternaria spp. was confirmed to be an opportunistic, saprophytic and weak pathogen that enters the plant tissues through natural openings and wounds [24,66,67], when the plant becomes more susceptible to diseases [68]. AASC strains caused lesions ranging from 8.43 to 17.0 mm in size while the lesion size of A. alternata strains ranged from 9.06 to 13.81 mm. According to previous studies [55,69], pathogenicity may be isolate-dependent instead of species-dependent. Our results showed that there was little difference among the tested Alternaria spp.
In conclusion, the present work describes for the first time the presence of AASC as an agent of black rot on sweet cherry fruits in Italy. Investigation should verify if other species of Alternaria could be involved in black rot of cherry fruit. Moreover, more isolates from different geographical areas should be included to explore the genetic diversity of the causal agents of black rot of cherry. Future studies will focus on developing and testing effective disease management strategies both in field and during postharvest. Furthermore, future studies will focus on characterizing the mycotoxin production potential of Alternaria species on sweet cherries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jof9100992/s1, Table S1: Collection details and GenBank accession numbers of strains isolated from sweet cherries in Italy and reference strains included in this study for the phylogenetic analysis.

Author Contributions

Conceptualization, M.W. and D.S.; methodology, M.W., D.S. and S.P.; software, M.W. and V.G.; validation, D.S., V.G. and M.W.; formal analysis, M.W.; investigation, M.W.; resources, D.S.; data curation, M.W.; writing—original draft preparation, M.W.; writing—review and editing, V.G., S.P. and D.S; supervision, D.S.; project administration, D.S.; funding acquisition, D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was conducted within the framework of the PRIMA StopMedWaste project, which is funded by PRIMA, a program supported by the European Union.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. FAOSTAT. Available online: https://www.fao.org/faostat/en/#data/QCL/visualize (accessed on 27 February 2023).
  2. ISTAT. Available online: http://dati.istat.it/Index.aspx?QueryId=33705&lang=en (accessed on 14 March 2023).
  3. Padilla-Zakour, O.I.; Ryona, I.; Cooley, H.J.; Robinson, T.L.; Osborne, J.; Freer, J. Shelf-Life Extension of Sweet Cherries by Field Management, Post-Harvest Treatments and Modified Atmosphere Packaging. N. Y. Fruit. Q. 2007, 15, 3–6. [Google Scholar]
  4. Chiabrando, V.; Garavaglia, L.; Giacalone, G. The Postharvest Quality of Fresh Sweet Cherries and Strawberries with an Active Packaging System. Foods 2019, 8, 335. [Google Scholar] [CrossRef]
  5. Sehirli, S.; Karabulut, O.A.; Ilhan, K.; Sehirli, A. Use and Efficiency of Disinfectants within a Hydrocooler System for Postharvest Disease Control in Sweet Cherry. Int. J. Fruit. Sci. 2020, 20, S1590–S1606. [Google Scholar] [CrossRef]
  6. Maghenzani, M.; Chiabrando, V.; Santoro, K.; Spadaro, D.; Giacalone, G. Essential oil vapour treatment (Thymus vulgaris and Satureja montana) on postharvest quality of sweet cherry (cv Ferrovia). J. Food Nutr. Res. 2018, 57, 161–169. [Google Scholar]
  7. Mari, M.; Spadaro, D.; Casals, C.; Collina, M.; De Cal, A.; Usall, J. Postharvest Diseases of Stone Fruits. In Postharvest Pathology of Fresh Horticultural Produce; Palou, L., Smilanick, J.L., Eds.; CRC Press: Boca Raton, FL, USA, 2019; pp. 111–140. ISBN 9781138630833. [Google Scholar]
  8. Ahmad, T.; Liu, Y.; Shujian, H.; Moosa, A. First Record of Alternaria alternata Causing Postharvest Fruit Rot of Sweet Cherry (Prunus avium) in China. Plant Dis. 2020, 104, 2030. [Google Scholar] [CrossRef]
  9. Cancino, S.; Lolas, M.A.; Galdós, L.; Hernández, Y.; Ferrada, E.; Riveros, P.; Blanco-Ulate, B.; Díaz, G.A. Occurrence of Alternaria alternata and A. tenuissima Causing Black Rot in Cherry Fruits (Prunus avium) in Central Chile. Plant Dis. 2023. [Google Scholar] [CrossRef] [PubMed]
  10. Zhao, Y.Z.; Liu, Z.H. First Report of Black Spot Disease Caused by Alternaria alternata on Cherry Fruits in China. Plant Dis. 2012, 96, 1580. [Google Scholar] [CrossRef] [PubMed]
  11. Latinović, N.; Radišek, S.; Latinović, J. First Report of Alternaria alternata Causing Fruit Rot on Fig (Ficus carica) in Montenegro. Plant Dis. 2014, 98, 424. [Google Scholar] [CrossRef]
  12. Xiang, M.L.; Li, S.C.; Wu, F.; Zhao, X.Y.; Wang, Y.B.; An, X.X.; Zhang, Y.N.; Chen, M. First Report of Alternaria alternata Causing Fruit Rot on Tetradium ruticarpum in China. Plant Dis. 2021, 105, 1194. [Google Scholar] [CrossRef]
  13. Lawrence, D.P.; Gannibal, P.B.; Peever, T.L.; Pryor, B.M. The Sections of Alternaria: Formalizing Species-Group Concepts. Mycologia 2013, 105, 530–546. [Google Scholar] [CrossRef]
  14. Matić, S.; Tabone, G.; Garibaldi, A.; Gullino, M.L. Alternaria Leaf Spot Caused by Alternaria Species: An Emerging Problem on Ornamental Plants in Italy. Plant Dis. 2020, 104, 2275–2287. [Google Scholar] [CrossRef] [PubMed]
  15. Thomma, B.P.H.J. AlternariasSpp.: From General Saprophyte to Specific Parasite. Mol. Plant Pathol. 2003, 4, 225–236. [Google Scholar] [CrossRef]
  16. Woudenberg, J.H.C.; Groenewald, J.Z.; Binder, M.; Crous, P.W. Alternaria Redefined. Stud. Mycol. 2013, 75, 171–212. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, M.-J.; Zheng, X.-R.; Li, H.; Chen, F.-M. Alternaria alternata, the Causal Agent of a New Needle Blight Disease on Pinus Bungeana. J. Fungi 2023, 9, 71. [Google Scholar] [CrossRef] [PubMed]
  18. Abata, L.K.; Paz, I.A.; Viera, W.; Flores, F.J. First Report of Alternaria Rot Caused by Alternaria alternata on Peach in Ecuador. Plant Dis. 2016, 100, 2323. [Google Scholar] [CrossRef]
  19. Dogan, A.; Cat, A.; Catal, M.; Erkan, M. First Report of Alternaria alternata Causing Postharvest Decay in Fig (Ficus carica L. Cv. Bursa Siyahi) Fruit in Turkey. J. Biotechnol. 2018, 280, S84. [Google Scholar] [CrossRef]
  20. Gur, L.; Reuveni, M.; Cohen, Y. Occurrence and Etiology of Alternaria Leaf Blotch and Fruit Spot of Apple Caused by Alternaria alternata f. sp. mali on cv. Pink Lady in Israel. Eur. J. Plant Pathol. 2017, 147, 695–708. [Google Scholar] [CrossRef]
  21. Moslemi, A.; Ades, P.K.; Groom, T.; Nicolas, M.E.; Taylor, P.W.J. Alternaria infectoria and Stemphylium herbarum, Two New Pathogens of Pyrethrum (Tanacetum cinerariifolium) in Australia. Australas. Plant Pathol. 2017, 46, 91–101. [Google Scholar] [CrossRef]
  22. Munhuweyi, K.; Lennox, C.L.; Meitz-Hopkins, J.C.; Caleb, O.J.; Opara, U.L. Major Diseases of Pomegranate (Punica granatum L.), Their Causes and Management—A Review. Sci. Hortic. 2016, 211, 126–139. [Google Scholar] [CrossRef]
  23. Ustun, R.; Cat, A.; Uzun, B.; Catal, M. First Report of Alternaria Alternata Causing Leaf Spot Disease on Soybean (Glycine max) in Antalya Province of Turkey. Plant Dis. 2019, 103, 3284. [Google Scholar] [CrossRef]
  24. Şimşek, A.; Dinler, H.; Uysal Morca, A. Identification and Pathogenicity of Alternaria Alternata Causing Leaf Spot Disease on Sweet Cherry in Province of Turkey. J. Plant Dis. Prot. 2022, 129, 1355–1366. [Google Scholar] [CrossRef]
  25. Thomidis, T.; Tsipouridis, C. First Report of Alternaria Leaf Spot on Cherry Trees in Greece. Plant Dis. 2006, 90, 680. [Google Scholar] [CrossRef]
  26. Van de Perre, E.; Jacxsens, L.; Liu, C.; Devlieghere, F.; De Meulenaer, B. Climate impact on Alternaria moulds and their mycotoxins in fresh produce: The case of the tomato chain. Food Res. Int. 2015, 68, 41–46. [Google Scholar] [CrossRef]
  27. Simmons, E.G. Alternaria: An Identification Manual; CBS Fungal Biodiversity Centre: Utrecht, The Netherlands, 2007; ISBN 9070351684. [Google Scholar]
  28. Andrew, M.; Peever, T.L.; Pryor, B.M. An Expanded Multilocus Phylogeny Does Not Resolve Morphological Species within the Small-Spored Alternaria Species Complex. Mycologia 2009, 101, 95–109. [Google Scholar] [CrossRef]
  29. Ma, G.; Bao, S.; Zhao, J.; Sui, Y.; Wu, X. Morphological and Molecular Characterization of Alternaria Species Causing Leaf Blight on Watermelon in China. Plant Dis. 2021, 105, 60–70. [Google Scholar] [CrossRef]
  30. Somma, S.; Pose, G.; Pardo, A.; Mulè, G.; Pinto, V.F.; Moretti, A.; Logrieco, A.F. AFLP Variability, Toxin Production, and Pathogenicity of Alternaria Species from Argentinean Tomato Fruits and Puree. Int. J. Food Microbiol. 2011, 145, 414–419. [Google Scholar] [CrossRef]
  31. Stewart, J.E.; Andrew, M.; Bao, X.; Chilvers, M.I.; Carris, L.M.; Peever, T.L. Development of Sequence Characterized Amplified Genomic Regions (SCAR) for Fungal Systematics: Proof of Principle Using Alternaria, Ascochyta and Tilletia. Mycologia 2013, 105, 1077–1086. [Google Scholar] [CrossRef]
  32. Woudenberg, J.H.C.; Seidl, M.F.; Groenewald, J.Z.; de Vries, M.; Stielow, J.B.; Thomma, B.P.H.J.; Crous, P.W. Alternaria Section Alternaria: Species, Formae Speciales or Pathotypes? Stud. Mycol. 2015, 82, 1–21. [Google Scholar] [CrossRef]
  33. Lawrence, D.P.; Rotondo, F.; Gannibal, P.B. Biodiversity and Taxonomy of the Pleomorphic Genus Alternaria. Mycol. Prog. 2016, 15, 3. [Google Scholar] [CrossRef]
  34. Ozkilinc, H.; Rotondo, F.; Pryor, B.M.; Peever, T.L. Contrasting Species Boundaries between Sections Alternaria and Porri of the Genus Alternaria. Plant Pathol. 2018, 67, 303–314. [Google Scholar] [CrossRef]
  35. Peever, T.L.; Ibañez, A.; Akimitsu, K.; Timmer, L.W. Worldwide Phylogeography of the Citrus Brown Spot Pathogen, Alternaria alternata. Phytopathology 2002, 92, 794–802. [Google Scholar] [CrossRef]
  36. Peever, T.L.; Su, G.; Carpenter-Boggs, L.; Timmer, L.W. Molecular Systematics of Citrus-Associated Alternaria Species. Mycologia 2004, 96, 119. [Google Scholar] [CrossRef]
  37. Pryor, B.M.; Michailides, T.J. Morphological, Pathogenic, and Molecular Characterization of Alternaria Isolates Associated with Alternaria Late Blight of Pistachio. Phytopathology 2002, 92, 406–416. [Google Scholar] [CrossRef]
  38. Hong, S.G.; Cramer, R.A.; Lawrence, C.B.; Pryor, B.M. Alt a 1 Allergen Homologs from Alternaria and Related Taxa: Analysis of Phylogenetic Content and Secondary Structure. Fungal Genet. Biol. 2005, 42, 119–129. [Google Scholar] [CrossRef] [PubMed]
  39. Isshiki, A.; Akimitsu, K.; Yamamoto, M.; Yamamoto, H. Endopolygalacturonase Is Essential for Citrus Black Rot Caused by Alternaria Citri but Not Brown Spot Caused by Alternaria alternata. Mol. Plant Microbe Interact. 2001, 14, 749–757. [Google Scholar] [CrossRef] [PubMed]
  40. Isshiki, A.; Akimitsu, K.; Nishio, K.; Tsukamoto, M.; Yamamoto, H. Purification and Characterization of an Endopolygalacturonase from the Rough Lemon Pathotype of Alternaria alternata, the Cause of Citrus Brown Spot Disease. Physiol. Mol. Plant Pathol. 1997, 51, 155–167. [Google Scholar] [CrossRef]
  41. Sung, G.H.; Sung, J.M.; Hywel-Jones, N.L.; Spatafora, J.W. A Multi-Gene Phylogeny of Clavicipitaceae (Ascomycota, Fungi): Identification of Localized Incongruence Using a Combinational Bootstrap Approach. Mol. Phylogenet Evol. 2007, 44, 1204–1223. [Google Scholar] [CrossRef]
  42. Liu, Y.J.; Whelen, S.; Hall, B.D. Phylogenetic Relationships among Ascomycetes: Evidence from an RNA Polymerse II Subunit. Mol. Biol. Evol. 1999, 16, 1799–1808. [Google Scholar] [CrossRef]
  43. Prencipe, S.; Meloni, G.R.; Nari, L.; Schiavon, G.; Spadaro, D. Pathogenicity, Molecular Characterization and Mycotoxigenic Potential of Alternaria Spp. Agents of Black Spots on Fruit and Leaves of Pyrus Communis in Italy. Phytopathology 2022, 113, 309–320. [Google Scholar] [CrossRef]
  44. Woudenberg, J.H.C.; Truter, M.; Groenewald, J.Z.; Crous, P.W. Large-Spored Alternaria Pathogens in Section Porri Disentangled. Stud. Mycol. 2014, 79, 1–47. [Google Scholar] [CrossRef]
  45. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed]
  46. Nylander, J.A.A. MrModeltest Version 2. Program Distributed by the Author. Evolutionary Biology Centre, Uppsala University, Uppsala.-References-Scientific Research Publishing. 2004. Available online: https://www.scirp.org/(S(i43dyn45teexjx455qlt3d2q))/reference/ReferencesPapers.aspx?ReferenceID=1276217 (accessed on 29 March 2023).
  47. Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian Phylogenetic Inference and Model Choice Across a Large Model Space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [PubMed]
  48. Swofford, D.L. PAUP. Phylogenetic Analysis Using Parsimony (and Other Methods). Version 4. Sinauer Associates, Sunderland.-References-Scientific Research Publishing. 2003. Available online: https://www.scirp.org/(S(czeh2tfqyw2orz553k1w0r45))/reference/ReferencesPapers.aspx?ReferenceID=1085917 (accessed on 29 March 2023).
  49. Hillis, D.M.; Bull, J.J. An Empirical Test of Bootstrapping as a Method for Assessing Confidence in Phylogenetic Analysis. Syst. Biol. 1993, 42, 182–192. [Google Scholar] [CrossRef]
  50. Wang, F.; Saito, S.; Michailides, T.J.; Xiao, C.L. Phylogenetic, Morphological, and Pathogenic Characterization of Alternaria Species Associated with Fruit Rot of Mandarin in California. Plant Dis. 2021, 105, 1555–1567. [Google Scholar] [CrossRef]
  51. Zhu, X.Q.; Xiao, C.L. Phylogenetic, Morphological, and Pathogenic Characterization of Alternaria Species Associated with Fruit Rot of Blueberry in California. Phytopathology 2015, 105, 1555–1567. [Google Scholar] [CrossRef]
  52. Kanetis, L.; Testempasis, S.; Goulas, V.; Samuel, S.; Myresiotis, C.; Karaoglanidis, G.S. Identification and Mycotoxigenic Capacity of Fungi Associated with Pre- and Postharvest Fruit Rots of Pomegranates in Greece and Cyprus. Int. J. Food Microbiol. 2015, 208, 84–92. [Google Scholar] [CrossRef]
  53. Riquelme, D.; Zuniga, C.; Tapia, E. First Report of Fruit Rot of Sweet Cultivars of Japanese Plum Caused by Alternaria alternata, A. arborescens, and A. tenuissima in Chile. Plant Dis. 2021, 105, 4167. [Google Scholar] [CrossRef]
  54. Ntasiou, P.; Myresiotis, C.; Konstantinou, S.; Papadopoulou-Mourkidou, E.; Karaoglanidis, G.S. Identification, Characterization and Mycotoxigenic Ability of Alternaria spp. Causing Core Rot of Apple Fruit in Greece. Int. J. Food Microbiol. 2015, 197, 22–29. [Google Scholar] [CrossRef]
  55. Harteveld, D.O.C.; Akinsanmi, O.A.; Drenth, A. Multiple Alternaria Species Groups Are Associated with Leaf Blotch and Fruit Spot Diseases of Apple in Australia. Plant Pathol. 2013, 62, 289–297. [Google Scholar] [CrossRef]
  56. Aiello, D.; Guarnaccia, V.; Azzaro, A.; Polizzi, G. Alternaria Brown Spot on New Clones of Sweet Orange and Lemon in Italy. Phytopathol. Mediterr. 2020, 59, 131–145. [Google Scholar] [CrossRef]
  57. Yang, J.; Chen, Y.Z.; Yu-Xuan, W.; Tao, L.; Zhang, Y.D.; Wang, S.R.; Zhang, G.C.; Zhang, J. Inhibitory Effects and Mechanisms of Vanillin on Gray Mold and Black Rot of Cherry Tomatoes. Pestic. Biochem. Physiol. 2021, 175, 104859. [Google Scholar] [CrossRef] [PubMed]
  58. Al-Rahbi, B.A.A.; Al-Sadi, A.M.; Al-Mahmooli, I.H.; Al-Maawali, S.S.; Al-Mahruqi, N.M.T.; Velazhahan, R. Meyerozyma guilliermondii SQUCC-33Y Suppresses Postharvest Fruit Rot of Strawberry Caused by Alternaria Alternata. Australas. Plant Pathol. 2021, 50, 349–352. [Google Scholar] [CrossRef]
  59. Ezra, D.; Shulhani, R.; Bar Ya’Akov, I.; Harel-Beja, R.; Holland, D.; Shtienberg, D. Factors Affecting the Response of Pomegranate Fruit to Alternaria alternata, the Causal Agent of Heart Rot. Plant Dis. 2019, 103, 315–323. [Google Scholar] [CrossRef] [PubMed]
  60. Chen, M.; Jia, M.S.; Li, S.C.; Xiao, L.H.; Wang, Y.B.; Peng, W.W.; Chen, J.Y.; Xiang, M.L. First Report of Postharvest Fruit Rot in Solanum muricatum Caused by Alternaria alternata in Southwest China. Plant Dis. 2022, 106, 2520. [Google Scholar] [CrossRef]
  61. Guo, L.W.; He, S.H.; Gao, Z.R.; Wu, Z.T.; Duan, R.Q.; Yang, K.Z.; Wei, Y.J.; He, X.H. First Report of Fruit Rot in Opium Poppy (Papaver somniferum) Caused by Alternaria alternata in China. Plant Dis. 2020, 104, 3264. [Google Scholar] [CrossRef]
  62. Basım, E.; Basım, H.; Abdulai, M.; Baki, D.; Öztürk, N. Identification and Characterization of Alternaria alternata Causing Leaf Spot of Olive Tree (Olea europaea) in Turkey. Crop Prot. 2017, 92, 79–88. [Google Scholar] [CrossRef]
  63. Serdani, M.; Kang, J.C.; Andersen, B.; Crous, P.W. Characterisation of Alternaria Species-Groups Associated with Core Rot of Apples in South Africa. Mycol. Res. 2002, 106, 561–569. [Google Scholar] [CrossRef]
  64. Pryor, B.M.; Gilbertson, R.L. Molecular Phylogenetic Relationships amongst Alternaria Species and Related Fungi Based upon Analysis of Nuclear ITS and Mt SSU RDNA Sequences. Mycol. Res. 2000, 104, 1312–1321. [Google Scholar] [CrossRef]
  65. Elfar, K.; Zoffoli, J.P.; Latorre, B.A. Identification and Characterization of Alternaria Species Associated with Moldy Core of Apple in Chile. Plant Dis. 2018, 102, 2158–2169. [Google Scholar] [CrossRef]
  66. Prusky, D. Pathogen Quiescence in Postharvest Diseases. Annu. Rev. Phytopathol. 1996, 34, 413–434. [Google Scholar] [CrossRef]
  67. Rotem, J. The Genus Alternaria: Biology, Epidemiology, and Pathogenicity; The American Phytopathological Society: Saint Paul, MN, USA, 1994; ISBN 978-0-89054-152-4. [Google Scholar]
  68. Mmbaga, M.T.; Shi, A.; Kim, M.-S. Identification of Alternaria alternata as a Causal Agent for Leaf Blight in Syringa Species. Plant Pathol J 2011, 27, 120–127. [Google Scholar] [CrossRef]
  69. Fontaine, K.; Fourrier-Jeandel, C.; Armitage, A.D.; Boutigny, A.L.; Crépet, M.; Caffier, V.; Gnide, D.C.; Shiller, J.; Le Cam, B.; Giraud, M.; et al. Identification and Pathogenicity of Alternaria Species Associated with Leaf Blotch Disease and Premature Defoliation in French Apple Orchards. PeerJ 2021, 9, e12496. [Google Scholar] [CrossRef]
Figure 1. (A,B) Symptoms of Alternaria black rot on naturally infected sweet cherry cv. Regina; (C) colony growth of AASC (strain T8) after 7 days on PDA; (D) colony growth of A. alternata (strain GR13) after 7 days on PDA; (E) conidia of AASC (strain T8) obtained after 15 days on PCA; (F) conidia of A. alternata (strain GR13) obtained after 15 days on PCA. Scale bar: (E,F) = 10 µm.
Figure 1. (A,B) Symptoms of Alternaria black rot on naturally infected sweet cherry cv. Regina; (C) colony growth of AASC (strain T8) after 7 days on PDA; (D) colony growth of A. alternata (strain GR13) after 7 days on PDA; (E) conidia of AASC (strain T8) obtained after 15 days on PCA; (F) conidia of A. alternata (strain GR13) obtained after 15 days on PCA. Scale bar: (E,F) = 10 µm.
Jof 09 00992 g001
Figure 2. Consensus phylogram of 4947 trees resulting from a Bayesian analysis of the combined rpb2, Alta-1, endoPG and OPA 10-2 sequence alignments of the Alternaria species. Bootstrap support values and Bayesian posterior probability values are indicated at the nodes. The isolates obtained in this study are in red. The tree was rooted with Alternaria nobilis AC1.
Figure 2. Consensus phylogram of 4947 trees resulting from a Bayesian analysis of the combined rpb2, Alta-1, endoPG and OPA 10-2 sequence alignments of the Alternaria species. Bootstrap support values and Bayesian posterior probability values are indicated at the nodes. The isolates obtained in this study are in red. The tree was rooted with Alternaria nobilis AC1.
Jof 09 00992 g002
Figure 3. Symptoms of Alternaria fruit rot on sweet cherries after 10 days post inoculation (A) AASC (strain X7); (B) A. alternata (strain Ch23); (C) healthy control.
Figure 3. Symptoms of Alternaria fruit rot on sweet cherries after 10 days post inoculation (A) AASC (strain X7); (B) A. alternata (strain Ch23); (C) healthy control.
Jof 09 00992 g003
Table 1. Strain name, species identification, year of isolation, colony growth (cm), size of conidia (length-width) and mean rot diameter obtained with the pathogenicity test for the strains isolated from sweet cherry fruits.
Table 1. Strain name, species identification, year of isolation, colony growth (cm), size of conidia (length-width) and mean rot diameter obtained with the pathogenicity test for the strains isolated from sweet cherry fruits.
Isolate NameFungal SpeciesYear of IsolationCherry CultivarColony Growth (cm ± SD)Size of Conidia (Length × Width; μm)Rot Diameter (mm ± SD) 1
D4AASC2020Kordia4.94 ± 0.5921.73 × 11.3112.06 b–d ± 1.52
T2AASC2020Ferrovia5.14 ± 0.2820.13 × 11.199.56 bc ± 1.22
T3AASC2020Ferrovia4.98 ± 0.5521.68 × 11.2312.21 bc ± 1.66
T6A. alternata2020Ferrovia4.90 ± 0.5920.64 × 11.1813.81 b–d ± 1.82
T8AASC2020Ferrovia4.96 ± 0.5520.57 × 11.1010.65 b–d ± 1.86
T9AASC2020Ferrovia5.08 ± 0.3521.53 × 11.2710.75 bc ± 2.31
GR1AASC2020Regina4.90 ± 0.5821.73 × 11.2411.39 b–d ± 1.88
GR3A. alternata2020Regina4.70 ± 0.5120.83 × 11.129.06 bc ± 1.81
GR6AASC2020Regina4.8 8± 0.5320.89 × 10.9914.94 b–d ± 2.77
GR8AASC2020Regina4.94 ± 0.5920.94 × 11.2011.94 b–d ± 1.68
GR13A. alternata2020Regina4.82 ± 0.5120.93 × 11.219.06 bc ± 1.45
W2AASC2020Kordia5.02 ± 0.3321.07 × 11.1611.06 b–d ± 1.53
W6AASC2020Kordia5.30 ± 0.4120.43 × 11.1211.40 b–d ± 2.29
Q1AASC2020Kordia5.00 ± 0.7819.76 × 11.6414.43 b–d ± 2.64
X7AASC2020Kordia4.96 ± 0.6220.84 × 11.3115.25 cd ± 2.94
Ch1AASC2021Regina4.70 ± 0.7021.25 × 11.119.38 bc ± 1.98
Ch2AASC2021Regina5.16 ± 0.1520.59 × 10.8610.94 bc ± 2.59
Ch4AASC2021Regina4.94 ± 0.5521.30 × 11.218.43 ab ± 2.03
Ch5AASC2021Regina4.76 ± 0.9221.32 × 11.1513.93 b–d ± 1.94
Ch11AASC2021Regina5.12 ± 0.1320.75 × 11.189.11 bc ± 1.61
Ch12AASC2021Regina4.96 ± 0.5520.88 × 11.0210.71 bc ± 3.54
Ch13AASC2021Regina4.94 ± 0.5921.16 × 11.179.12 bc ± 1.66
Ch14AASC2021Regina5.26 ± 0.4920.20 × 10.8410.06 b–d ± 1.59
Ch16AASC2021Regina5.12 ± 0.7721.40 × 11.3111.75 b–d ± 1.88
Ch17AASC2021Regina5.08 ± 0.7421.32 × 11.1417.00 d ± 2.13
Ch18AASC2021Regina4.94 ± 0.4721.71 × 11.2314.25 b–d ± 1.89
Ch19AASC2021Regina4.74 ± 0.8121.10 × 11.2212.64 b–d ± 1.61
Ch21A. alternata2021Regina5.06 ± 0.3421.02 × 10.8811.00 bc ± 1.98
Ch22AASC2021Regina5.00 ± 0.2120.87 × 11.0710.06 bc ± 1.66
Ch23A. alternata2022Regina5.12 ± 0.5320.90 × 11.0311.50 b–d ± 1.90
Ch26AASC2022Regina5.18 ± 0.4821.04 × 11.1211.94 b–d ± 1.87
Ch27AASC2022Regina4.86 ± 0.7721.12 ×11.099.25 bc ± 1.80
Ch35AASC2022Regina4.98 ± 0.4121.03 × 11.1610.78 b–d ± 1.93
Ch37A. alternata2022Regina4.52 ± 0.7821.12 × 11.199.17 bc ± 1.52
Ch39AASC2022Regina4.74 ± 0.5721.07 × 10.9811.50 bc ± 1.93
Ch40AASC2022Regina5.18 ± 0.2621.12 × 11.1110.50 bc ± 1.74
Ch42AASC2022Regina5.04 ± 0.2720.94 × 11.0312.06 b–d ± 1.52
Ch43A. alternata2022Regina4.68 ± 0.6821.05 × 11.0110.88 bc ± 1.23
Ch45AASC2022Regina4.92 ± 0.5921.20 × 11.2910.81 bc ± 1.94
Ch48AASC2022Regina4.88 ± 0.5021.11 × 11.059.33 bc ± 1.97
Healthy control--Regina--0 a
1 Values are mean of the rot diameter on nine cherry fruits. Values with the same letter are not different according to Tukey’s test (p ≤ 0.05).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Waqas, M.; Prencipe, S.; Guarnaccia, V.; Spadaro, D. Molecular Characterization and Pathogenicity of Alternaria spp. Associated with Black Rot of Sweet Cherries in Italy. J. Fungi 2023, 9, 992. https://doi.org/10.3390/jof9100992

AMA Style

Waqas M, Prencipe S, Guarnaccia V, Spadaro D. Molecular Characterization and Pathogenicity of Alternaria spp. Associated with Black Rot of Sweet Cherries in Italy. Journal of Fungi. 2023; 9(10):992. https://doi.org/10.3390/jof9100992

Chicago/Turabian Style

Waqas, Muhammad, Simona Prencipe, Vladimiro Guarnaccia, and Davide Spadaro. 2023. "Molecular Characterization and Pathogenicity of Alternaria spp. Associated with Black Rot of Sweet Cherries in Italy" Journal of Fungi 9, no. 10: 992. https://doi.org/10.3390/jof9100992

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop