SciELO - Scientific Electronic Library Online

 
vol.52 número2Efectos del silicato de calcio en la protección de las semillas de Brachiaria contra Sclerotium rolfsiiEficiencia de hongos entomopatógenos contra Rhyzopertha dominica (Fabricius) (Coleoptera: Bostrichidae) bajo condiciones de laboratorio índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

  • No hay articulos citadosCitado por SciELO

Links relacionados

  • No hay articulos similaresSimilares en SciELO

Compartir


Revista de la Facultad de Ciencias Agrarias. Universidad Nacional de Cuyo

versión impresa ISSN 1853-8665versión On-line ISSN 1853-8665

Rev. Fac. Cienc. Agrar., Univ. Nac. Cuyo vol.52 no.2 Mendoza dic. 2020

 

ORIGINAL ARTICLE

Characterization of Fusarium spp., a Phytopathogen of avocado (Persea americana Miller var. drymifolia (Schltdl. and Cham.)) in Michoacán, México

Caracterización de Fusarium spp., fitopatógeno de aguacate (Persea americana Miller var. drymifolia (Schltdl. y Cham.)) en Michoacán, México

Guillermo Gregorio Olalde-Lira 1, Yurixhi Atenea Raya Montaño 1, Patricio Apáez Barrios 2, Margarita Vargas-Sandoval 1, Martha Elena Pedraza Santos 1, Tania Raymundo 3, Ricardo Valenzuela 3, Ma. Blanca Nieves Lara-Chávez 1*

1 Universidad Michoacana de San Nicolás de Hidalgo. Facultad de Agrobiología "Presidente Juárez". Laboratorio de Fitopatología. Paseo Lázaro Cárdenas 2290. Emiliano Zapata. Melchor Ocampo. Código Postal 60170. Uruapan. Michoacán. México. * chavez12001@yahoo.com.mx

2 Universidad Michoacana de San Nicolás de Hidalgo. Facultad de Ciencias Agropecuarias. Prolongación de la calle Mariano Jiménez. Col. El Varillero S/N. C. P. 60660. Apatzingán. Michoacán. México.

3 Instituto Politécnico Nacional. Escuela Nacional de Ciencias Biológicas. Departamento de Botánica. Laboratorio de Micología. Prol. Carpio y Plan de Ayala s/n.Col. Santo Tomás, Alcaldía Miguel Hidalgo. C. P. 11340. Ciudad de México. CDMX. México.

Originales: Recepción: 21/12/2018 - Aceptación: 04/05/2020


ABSTRACT

Avocado has great socioeconomic importance in Mexico because of the benefits it generates for the production chain participants and the significant foreign exchange earnings engendered by the export of its fruit. However, this crop has phytosanitary problems, caused mainly by fungi, among which the genus Fusarium stands out. The objective of this study was to characterize Fusarium species that caused root rot in avocado trees in Michoacan, Mexico. In 19 isolates of Fusarium spp., polymerase chain reactions (PCR) with primers coding for elongation factor and calmodulin genes were performed. These sequences were compared in homology using BLAST analysis and aligned in MEGA 6.0. Cladograms were created based on maximum verisimilitude. The pathogenicity of the isolates was evaluated based on their virulence and severity in the avocado plants. Morphological and molecular analyses showed that 15 isolates belonged to F. oxysporum Schl and four to F. solani Mart. All isolates were pathogenic, with virulence ranging from 16 to 56 days. All isolates produced root rot and yellowing of leaves, with 63% producing wilting and 16% producing apical necrosis, the latter being the most severe.

Keywords: Plant pathogen; Root rot; Crop disease; Ascomycetes; Fusarium oxysporum Schl; Fusarium solani Mart

RESUMEN

El aguacate en México tiene gran importancia socioeconómica por los beneficios que genera para los participantes de la cadena productiva y la generación de divisas por la exportación de su fruta. Sin embargo, este cultivo presenta problemas fitosanitarios ocasionados por hongos entre los que destacan el género Fusarium. El objetivo de esta investigación fue caracterizar especies de Fusarium que producen pudrición de raíz en árboles de aguacate en Michoacán, México. En 19 aislados de Fusarium spp. la reacción en cadena de la polimerasa (PCR) se hizo con iniciadores que codifican para genes de factor de elongación y calmodulina. Estas secuencias se compararon en homología por medio de análisis BLAST y se alinearon en MEGA 6.0. Los cladogramas se hicieron con base a máxima verisimilitud. La patogenicidad de los aislados se evaluó de acuerdo con su virulencia y severidad en plantas de aguacate. Los análisis morfológicos y moleculares concluyeron que 15 aislados pertenecen a Fusarium oxysporum Schl. y cuatro a Fusarium solani Mart. Todos los aislados fueron patogénicos, presentaron con su virulencia que osciló entre 16 a 56 días. El total de los aislados produjo pudrición radical y amarillamiento de las hojas, marchitez en 63% y necrosis apical en 16%, este último síntoma fue el más severo.

Palabras clave: Patógeno vegetal; Pudrición de raíz; Enfermedades de cultivos; Ascomycetes; Fusarium oxysporum Schl; Fusarium solani Mart


INTRODUCTION

In 2016, 583,978 ha globally were sown with avocado, yielding 5,689,985 t (13). During that same year, in Mexico, 180,546.43 ha with a production of 1,889.353.50 t was sown with avocado. The state of Michoacan accounted for 85% of this total (38). This strong performance in Michoacan was because the crop was established in an area that contained benign agroecological conditions (soil and climate) that favored its reproductive development (19).

However, this species has undergone genetic and physiological changes resulting from the massive establishment and intensive management of its production systems, thus limiting production and affecting its quality (39, 46). Plant diseases caused by Fusarium spp. in the tropics have increased in importance with the introduction of high-yield production systems and genetically uniform cultivars (36, 44). Symptoms caused by Fusarium spp. in avocado trees include leaf yellowing, interrupted growth of vegetative flushes, small leaves, and the premature abscission of leaves. If rotting is severe, the trees can die weeks after the first foliar symptoms appear or they barely survive for long periods. At the root level, the observed symptoms include the rotting of small roots with a dark brown-black hue that, over time, spreads to the thickest roots. The symptoms described differ from those caused by Armillaria mellea or Rosellinia sp. owing to the absence of rhizomorphs or abundant white mycelia, which are characteristics of these species and, for Phytophthora cinnamomi, due to the absence of secondary roots (23, 32, 33).

Fungal species have traditionally been identified by morphological techniques based on the comparison of characters. These techniques are often complicated and slow, given that different variables must be evaluated. This is in addition to the observation and recognition of the characteristics of the isolates (45).

Alternative diagnostic techniques, such as molecular tools, have developed in recent decades, Polymerase chain reaction (PCR)-based protocols and specific primers provide better-founded diagnoses, as well as taxonomic identifications at the species level (10). The limitations associated with biochemical, morphological, and cytological variability have been overcome by the development of DNA markers (2).

Although the taxonomy and identification of species of the genus Fusarium have been studied based on several genetic markers, it has not yet been possible to find a suitable marker for the identification of all its species (47). The high degree of variability within F. oxysporum has shown the complexity of the species (3).

A wide variety of molecular markers have been used to analyze the diversity of plant pathogenic Fusarium species at the genomic level: AFLPs in F. graminearum and F. asiaticum (34) and RAPD and ISSR in F. graminearum and F. culmorum (28, 40, 43). Attempts have been made to characterize F. oxysporum lines using different techniques, such as IGS and ITS regions; however, these have not provided favorable results (40, 41).

Ribosomal RNA genes have been used in phylogenetic studies of fungi. Some of the criticism against using rDNA markers include if they are used instead of biased base composition then this might influence the branching order of the tree and the unequal rates of evolution among lineages, nonindependence of sites and invariable sites, and among-site rate variation may contribute to artifactual topologies in the trees. To compensate for this, several protein-coding genes have been explored as phylogenetic markers (20).

Studies have shown that phylogenies obtained from the elongation factor gene are congruent with other molecular phylogenies for the recovery of monophyly from groups such as Metazoa, Fungi, Magnoliophyta, and Euglenozoa. This marker is important for providing a highly informative section between closely related species (17, 20). However, coding sequences of genes such as calmodulin (31) have been employed for molecular phylogenetic analysis of the Gibberella fujikuroi complex and other Fusarium species.

In the avocado belt of Michoacan, Mexico, there are no studies of species of the genus Fusarium that cause root rot in avocado trees. Therefore, there is no information on the phenotypic, genetic, and pathogenic diversity of Fusarium species. Thus, the objectives of the present study were to isolate and morphologically and molecularly identify Fusarium species, obtained from avocado roots, and determine their pathogenicity in avocado plants. The hypothesis was that in the Michoacan avocado belt, Fusarium spp. is associated with avocado trees that have root rot symptoms.

MATERIALS AND METHODS

Targeted samplings were conducted in 10 localities of nine municipalities (Ario de Rosales, Los Reyes, Nuevo Parangaricutiro, Periban, Tacambaro, Tancitaro, Tingambato, Uruapan (2), and Ziracuaretiro) located in the so-called avocado belt of the state of Michoacán, México, with different elevations and agro-ecological zones (19). In each locality, three root samples were collected from five trees that showed wilting symptoms. Washing and disinfection of the root fragments were performed in the laboratory according to the protocol described by other researchers (1, 42). A total of 10 tissue sections per sample were seeded in a Petri dish with Bioxon® potato dextrose agar (PDA) culture medium. When the colony had grown, hyphal tips were transferred to Petri dishes containing Spezieller Nährstoffarmer Agar (SNA) with two pieces of 0.5 cm sterile filter paper to induce sporulation (24).

Monosporic cultures

Monosporic cultures were obtained for all isolates grown in SNA. From each isolate, two 5-mm disks of medium and mycelium were placed in 5-mL glass tubes containing 3 mL of sterile distilled water. The tubes were briefly shaken to separate the conidia from the mycelium and 25 μL was taken from the conidial suspension and dispersed in dishes containing 2% agar water. The dishes were incubated at 25°C in the darkness for 24 h. The germinated conidia were observed under a stereoscopic microscope Leica S9E and then individually transferred to dishes containing SNA and sterile paper, and incubated at 25°C. Once the cultures covered most of the medium in the dish, 3 mL of 25% sterile glycerol was added to remove the conidia from the mycelium. The conidial suspension (1 mL) was transferred to 2 mL cryogenic tubes and stored at -80°C. Aliquots of the stored conidial suspension of each isolate were used for subsequent characterization.

Cultural and morphological characterization

To stimulate the development of the cultures and conidia, fungal were incubated at 25°C in PDA medium and CLA (carnation leaves in 2% agar water) (24). After 15, 30, and 60 days of incubation, the following morphological characters were analyzed: type and thickness of the mycelium, and the colony color and agar pigmentation based on the Mathuen Handbook of Color (21) and the Color Picker tool for Mac OS X. The morphological characters of 30 macroconidia, microconidia, and chlamydospores were measured by calculating the ranges and mean values (24). In addition, information on the shape, size, and number of septa of macroconidia and microconidia, presence of mono and/ or polyphialides, chlamydospores, and sterile coiled hyphae was recorded.

For the growth rate, mycelial disks 1 cm in diameter were taken and placed in the center of five Petri dishes containing PDA medium, after which they were incubated at 21°C. The growth was measured after 72 h. The daily growth rate (mm) was obtained from the resulting difference between the final diameter minus the initial diameter divided by the total number of days evaluated (24).

Nucleic acid extraction

DNA extraction was performed using the protocol of Cenis (7). The pellet was resuspended in TE buffer and stored at -20°C until subsequent use. Quality was assessed by gel electrophoresis at 120 volts for 20 min and staining was performed with ethidium bromide, was after which it was observed in a MiniBis Pro® ultraviolet photo-documenter. The concentration and purity of the extracted DNA were measured in a Thermo Scientific NanoDrop spectrophotometer; 1 μL-1 of the sample was placed in a BioRad® SmartSpec ultraviolet spectrophotometer, fitted with a Hellma-Analytics® TrayCell cell, and evaluated by readings at 260 and 280 nm.

Polymerase chain reaction and DNA sequencing

DNAs from fungal isolates were amplified by PCR based on sequences coding for elongation factor EF-1α (ATGGGTAAGGARGACAAGAC) and EF-2 (GGARGTACCAGTSATCATGGT) and calmodulin Cal-228F (GAGTTCAAGGAGGCCTTCTCCC) and Cal-737F (CATCTTTCTGGCCATCATGG). For amplification, a PCR reaction mixture composed of 2.5 μL 1M Tris pH 9.0, 2 μL dNTPs, 0.5 μL forward primer, 0.5 μL reverse primer at a final concentration of 200 nM, 11.3 μL dH2 O, 3 μL of 50mM MgCl2, 0.2 μL Taq polymerase, and 2.5 μL of isolated DNA was used. PCR was performed in a Life Technologies® Veriti® thermal cycler under the following conditions: an initial denaturation cycle of 95°C for 4 min, followed by 35 denaturation cycles at 95°C for 1 min, annealing at 55°C for 1 min, and extension at 72°C for 2 min; final extension was performed at 72°C for 10 min (20, 29).

Amplification samples were electrophoretically analyzed in agarose gels at 2% Tris-borate-EDTA X (TBEX) with a molecular weight marker of 100 base pairs at 120 volts for 5 min and then at 80 volts for 60 min; the gels were then visualized in a MiniBis Pro® ultraviolet photo-documenter. The PCR product was purified and sequenced.

Phylogenetic analysis

The obtained sequences were compared in homology with sequences stored in GenBank using BLAST analysis, from where those that presented the best similarity index as well as those whose morphological characteristics and antecedents were relevant for the study were obtained. These sequences were aligned and edited using the MEGA 6.0 software program. Later, the best evolutionary model was sought for the construction of cladograms based on the analysis of maximum parsimony and maximum likelihood. The cladograms obtained were edited with the FigTree 1.4.2 application for Mac OSX.

Pathogenicity tests

Isolates were grown in CLA at 25°C for 15 days. Conidia were collected in sterile water and diluted to 6,072 conidia/ mL-1. The conidial suspension was used to inoculate 4-month-old avocado plants grown in sterile substrate and inoculations of the conidial suspensions were performed (42), for which 100 mL-1 of the solution of each isolate was added near the root of the plant, without causing damage to them, and five plants were inoculated per treatment including a control without inoculum. Subsequently, the severity and virulence of each isolate were evaluated.

A completely randomized experimental design was used. The virulence data were subjected to an analysis of variance and the means of the treatments were compared with Tukey’s test (p≤0.05). Analyses were performed using SAS (35) version 9.1.

RESULTS AND DISCUSSION

Cultural and morphological analysis A total of 19 isolates of Fusarium spp. were obtained from nine municipalities within the avocado belt of the state of Michoacan, Mexico. Monosporic cultures produced abundant, dense, aerial, and floccose mycelia in the PDA medium.

The colonies pigmented the culture medium, with 30% presenting gray coloration, 25% violet, 25% blue, and 20% yellow. Variation in colony coloration is a character of the genus Fusarium (5); however, pigments are not a representative taxonomic character of species and may vary within the same isolate, except for the yellow pigment that is characteristic of F. thapsinum Klittich (24, 47) (figure 1).

Figure 1. Comparison of 19 Fusarium isolates in PDA culture medium, obtained from 10 localities in the avocado belt of Michoacán, México.

Figura 1. Comparación de 19 aislados de Fusarium en medio de cultivo PDA, obtenidos de 10 localidades en la franja aguacatera de Michoacán, México.

The growth rates ranged between 1.65 and 5.13 mm/day, which were similar to those obtained for F. oxysporum Schl (16), indicating that it can grow up to 1 mm per day at a temperature of 25°C; however, for the isolates of yellow coloration, its growth rate was 2.98–3.17 mm/day.

In the CLA medium, the isolates produced abundant microconidia, which were oval, without septa, and arranged in fake heads. These characteristics match those reported previously for F. oxysporum and F. solani (27). The macroconidia, scarce in all isolates, were elongated, slightly curved, and septate. In the isolates F4, F6, F14, and F19, the macroconidia had five septa, with the apical cell rounded and the basal cell "foot"-shaped, being 8.98 to 9.42 μm wide and 51.12 to 56.32 μm long. For the remainder of the isolates, the macroconidia presented three septa, the apical cell was curved, and the basal cell "foot"-shaped, being 6.17 to 13.63 μm wide and 35.65 to 49.83 μm long. The macroconidia are considered the most important character for the identification of Fusarium species and in the CLA medium they are usually uniform (6, 24, 27, 35, 47). All isolates produced abundant chlamydospores of 12.5-17.6 μm long, globose shaped and smooth walled, terminals and intercalary in the hyphae, and were lonely and/or in small chains.

In the isolates F4, F6, F14, and F19, the chlamydospores were larger (19.3-20.4 μm diameter) than those of the other isolates, which matched with the size of the chlamydospores reported for F. solani (24, 27). Moreover, these four isolates presented rolled hyphae, both aerially and immersed in the culture medium. Rolled hyphae are also produced by F. sterilihyphosum, F. mexicanum, F. circinatum, and F. pseudocircinatum (24, 27, 35). Nevertheless, F. sterilihyphosum presents longer and thinner conidia, and absent chlamydospores (27). F. circinatum and F. pseudocircinatum present violet pigmentation in PDA medium and an absence of chlamydospores (24, 27, 30). In contrast, the isolates F4, F6, F14, and F19 produced abundant chlamydospores and yellow pigmentation in the culture medium (24) (figure 2, page 308).

Figure 2. Reproductive structures of Fusarium spp. A) Macroconidia, B) Rolled hyphae, C) Monophialides, D) Chlamydospores.

Figura 2. Estructuras reproductivas de Fusarium spp. A) Macroconidios, B) Hifas enrolladas, C) Monofialidess, D) Clamidosporas.

Molecular identification

The DNA extracted from the 19 isolates had a concentration between 102 and 398 ng μL-1, whereas the 260/280 nm absorbance ratio was 1.79-1.92, which coincides with the findings published previously (47), where a DNA sample with a value of -1.8 was accepted. The PCR based on the elongation factor and calmodulin genes provided fragments of a size greater than 600 base pairs; this size is among those reported for species of the genus Fusarium (2, 20, 29, 47). The selected GenBank sequences for phylogenetic analysis with elongation factor and calmodulin genes, by their similarity, are represented in table 1 (page 308) and table 2 (page 309).

Table 1. Coding sequences for the elongation factor gene obtained from GenBank.

Tabla 1. Secuencias codificantes para el gen factor de elongación obtenidas del GenBank.

Table 2. Coding sequences for calmodulin gene obtained from GenBank.

Tabla 2. Secuencias codificantes para el gen calmodulina obtenidas del GenBank.

Calmodulin

Phylogenetic analysis of the calmodulin gene coding region was performed with 640 characters in a GTR substitution model. The cladogram for maximum plausibility showed a monophyletic group that grouped the isolates of F. solani Mart, two monophyletic groups of F. oxysporum Schl, and a sister polyphyletic group of these. Results coincided with those from other research (31, 48); therefore, calmodulin-coding sequences are effective in the phylogenetic analysis of highly related Fusarium species (figure 3, page 309).

Figure 3. Cladogram product of the analysis of Fusarium sequences coding for calmodulin gene based on maximum likelihood with 500 repetitions in bootstrap.

Figura 3. Cladograma producto del análisis de secuencias de Fusarium codificantes del gen calmodulina con base en máxima verosimilitud con 500 repeticiones en bootstrap.

Elongation factor

Phylogenetic analysis of the elongation factor gene coding region was based on 632 characters and a GTR+G+I substitution model. The cladogram for maximum likelihood showed a different topology, with seven monophyletic groups, five of them for F. oxysporum, one for F. solani, and a group with those species that shared the character of a coiled hyphae. This resolution coincides with other researchers (18, 26, 47, 48) who stated that phylogenetic analyses performed based on protein-coding markers, and especially the region for the elongation factor, had a higher resolution than other primers such as calmodulin and ITS (figure 4).

Figure 4. Cladogram representative of the analysis of sequences of Fusarium spp. coding for elongation factor gene realized based on maximum likelihood with 500 repetitions in bootstrap.

Figura 4. Cladograma representativo del análisis de secuencias de Fusarium spp. codificantes para el gen factor de elongación realizado con base en Máxima verosimilitud con 500 repeticiones en bootstrap.

Pathogenicity tests

The 19 isolates inoculated in avocado plants were pathogenic and they varied in time, virulence, and severity of the disease. For virulence, the analysis of variance performed on the obtained data detected highly significant differences (p≤0.05). The time elapsed for the first symptoms to appear varied between species. For F. solani, it fluctuated between 51 and 56 days, whereas for F. oxysporum it ranged from 16 to 46 days (figure 5, page 311).

Figure 5. Virulence of the 19 Fusarium spp. isolates in avocado plants. Means with a different letter on a bar are statistically different (Tukey’ s test p≤0.05).

Figura 5. Virulencia de 19 aislados de Fusarium spp. en plantas de aguacate. Medias con una letra diferente sobre cada barra son estadísticamente diferentes (Prueba de Tukey p≤0,05).

The disease severity determined in avocado seedlings inoculated with the 19 isolates of the two species of the genus Fusarium showed variability, which was reflected by reduced plant growth.

At the root level, the 19 isolates showed a reduction in volume with respect to the control, as well as a reddish-brown coloration in their tissue indicating necrosis (figure 6, page 312).

Figure 6. Symptoms observed in avocado root Persea americana Mill. drymifolia variety caused by the isolates of F. oxysporum and F. solani. (A) Healthy roots, (B) Reddish-brown coloration roots, (C) Apical necrosis, (D) Healthy plant, (E) Wilting of tips, (F) Death of seedling.

Figura 6. Síntomas observados en raíz de aguacate Persea americana Mill. variedad drymifolia ocasionados por las cepas de F. oxysporum y F. solani. (A) Raíces sanas, (B) Raíces de coloración marrón-rojiza, (C) Necrosis apical, (D), Planta sana, (E) Marchitamiento de puntas, (F) Muerte de plántula.

These symptoms coincided with those reported in vanilla (22) and asparagus (11, 12), where F. oxysporum caused root rot in the former case, and root and crown rot in mature plants, seedlings, and young transplants in the latter case (8, 25). Similar symptoms were caused by F. paranaense belonging to the F. solani species complex in the root rot of soybeans (Glycine max (L.) Merrill). It has been reported that F. solani is secondary in citrus root rot because it does not cause plant death, despite colonizing a large part of the root, nevertheless it reduces the vigor of the plant (8), as noted in the present study.

In the aerial part of the plant, Fusarium isolates produced variability in the symptoms. Thus, in four isolates, yellowing was observed, in 12 apical wilting, and in three apical death compared with the control plant that remained healthy. These symptoms have previously been reported in other studies (4, 15, 22), which observed that F. oxysporum caused general decay, chlorosis, and necrosis in the stems. It has also been reported that F. solani in lisianthus (Eustoma grandiflorum) produced wilting, defoliation, and subsequent death of the plant (42). It has been reported that in a group of Fusarium species isolated from maize, F. subglutinans and F. verticillioides were the most pathogenic and produced similar symptoms to those observed initially in the present study (14).

The symptoms observed were yellowing, wilting, and apical necrosis. The most virulent isolate was F7, which was obtained in the municipality of Los Reyes, and corresponds to F. oxysporum Schl. This coincides with the results reported where F. oxysporum f. sp. radicis-vanillie caused yellowing and defoliation at 15 days after inoculation (22). F. solani has been shown to cause root rot at 35 days (8), which differed from that reported in other studies, where F. solani produced root and crown rot 2 weeks after inoculation (15).

CONCLUSIONS

Cultural characterization and phylogenetic analysis derived from molecular characterization confirmed that 15 isolates belonged to F. oxysporum (F1, F2, F3, F5, F7, F8, F9, F10, F11, F12, F13, F15, F16, F17, and F18) and four to F. solani (F4, F6, F14, and F19).

The EF-1α obtained the best resolution for the identification of the Fusarium species. F. oxysporum and F. solani produced root rot in avocado trees in the avocado belt of the state of Michoacan, Mexico. F. oxysporum and F. solani were pathogenic in avocado plants (P. americana Miller drymifolia variety).

The species of F. oxysporum were more virulent than the ones from F. solani by producing root rot quickly, and thus yellowing, wilting, and apical necrosis in the aerial parts and finally plant death.

The F7 isolate, collected in the municipality of Los Reyes, was the most virulent and caused the death of avocado plants in 16 days. The isolates of F. solani produced less damage, with F4 and F14 only causing apical wilting.

The pathogenic variability exhibited by the Fusarium spp. isolates in avocado seedlings could help to establish strategies that enable the control and management of this phytopathogen more efficient for avocado cultivation in Michoacán, México, from the nursery stage to field plantations.

REFERENCES

1. Agrios, N. G. 2005. Plant pathology. Department of Plant Pathology University of Florida. Fifth edition. Elsevier Academic Press. USA. 922 p.         [ Links ]

2. Arif, M.; Zaidi, N. W.; Haq, Q. M. R.; Singh, Y. P.; Taj, G.; Kar, C. S.; Singh, U. S. 2015. Morphological and comparative genomic analyses of pathogenic and non-pathogenic Fusarium solani isolated from Dalbergia sisso. Molecular Biology Reports. 42(6):1107-22. https:// doi:10.1007/s11033-014-3849-3.         [ Links ]

3. Baayen, R. P.; O’Donnell, K.; Bonants, P. J. M.; Cigelnik, E. K.L.; Roebroeck, E.; Waalwijk, C. 2000. Gene genealogies and AFLP analyses in the Fusarium oxysporum complex identify monophyletic and nonmonophyletic formae speciales causing wilt and rot disease. Phytopathology 90: 891-900. https://doi.org/10.1094/PHYTO.2000.90.8.891.

4. Bonacci, M.; Barros, G. 2019. Genetic diversity and pathogenicity on root seedlings from three soybean cultivars of Fusarium graminearum isolated from maize crop residues. Revista de la Facultad de Ciencias Agrarias. Universidad Nacional de Cuyo. Mendoza. Argentina. 51(1): 147-160.         [ Links ]

5. Booth, C. 1986. The genus Fusarium. Commonwealth Mycological Institute. Kew. Reino Unido. 237.         [ Links ]

6. Britz, H.; Steenkamp, E. T.; Coutinho, T. A.; Wingfield, B. D.; Marasas, W. F.; Winfield, M. J. 2002. Two new species of Fusarium section Liseola associated with mango malformation. Mycology 94: 22-730.         [ Links ]

7. Cenis, J. L. 1992. Rapid extraction of fungal DNA for PCR amplification. Nucleic Acids Research. 20: 9.         [ Links ]

8. Costa, S. S.; Matos, K. S.; Tessman, D. J.; Seixas, C.; Pfenning, L. H. 2016. Fusarium paranaense sp. nov., a member of the Fusarium solani species complex causes root rot on soybean in Brazil. Fungal Biology. 120: 51-60. https://doi.org/10.1016/j.funbio.2015.09.005        [ Links ]

9. Dandurand, L. M.; Menge, J. A. 1992. Influence of Fusarium solani on citrus root rot caused by Phytophthora parasitica and Phytophthora citrophthora. Plant and Soil. 144: 13-21. https://doi.org/10.1007/BF00018840        [ Links ]

10. Edwards, S.; O’Callaghan, J.; Dobson, A. D. W. 2002. PCR-Based detection and quantification of mycotoxigenic fungi. Mycological research 106: 1005-1025. Doi: https://doi. org/10.1017/S0953756202006354

11. Elmer, W. H. 2015. Management of Fusarium crown and root rot of asparagus. Crop Protection. 73: 2-6. https://doi.org/10.1016/j.cropro.2014.12.005        [ Links ]

12. Elmer, W. H.; Johnson, D. A.; Mink, G. J. 1996. Epidemiology and management of diseases causal to asparagus decline. Plant Disease. 80: 117-125.         [ Links ]

13. FAOSTAT 2016. Organización de las Naciones Unidas para la Alimentación y la Agricultura. http://www.fao.org/faostat/es/#data/QC (Accesada en Marzo 2016).         [ Links ]

14. Figueroa-Rivera, M. G.; Rodriguez-Guerra, R.; Guerrero-Aguilar, B. Z.; González-Chavira, M. M.; Pons-Hernández, J. L. 2010. Caracterización de especies de Fusarium asociadas a la pudrición de raíz de maíz en Guanajuato, Mexico. Revista Mexicana de Fitopatología. 28: 124-134.         [ Links ]

15. Gaetán, S. A.; Madia, M. S.; Perez, A. 2007. Recent outbreak of Fusarium crown and root rot caused by Fusarium solani in marjoram in Argentina. Australian Plant Disease Notes. 2: 15-16. doi 10.1071/DN07006.         [ Links ]

16. Garces de Granada, E.; Orozco de Amezquita, M.; Bautista, G. R.; Valencia, H. 2001. Fusarium oxysporum, el hongo que nos falta conocer. Acta Biol. Colombiana. 6: 1-20.         [ Links ]

17. Geiser, D.; Aoki, T.; Bacon, C.; Baker, S.; Bhattacharyya, M.; Brandt, M.; Brown, D.; Burgess, L.; Chulze, S.; Coleman, J.; Correll, J.; Covert, S.; Crous, P.; Cuomo, C.; Sybren De Hoog, G.; Di Pietro; A.; Elmer, W.; Epstein, L.; Frandsen, R.; Freeman, S.; Gagkaeva, T.; Glenn, A.; Gordon, T.; Gregory, N.; Gregory, N.; Hammond-Kosack, K.; Hanson, L.; Jimenez-Gasco, Ma. del M.; Kang, S.; Corby Kistler, H.; Kuldau, G.; Leslie, J.; Logrieco, A.; Lu, G.; Lysøe, E.; Ma, Li-Jun; McCormick, S.; Migheli, Q.; Moretti, A.; Munaut, F.; O`Donnell, K.; Pfening, L.; Ploetz, R.; Proctor, R.; Rehner, S.; Robert, V.; Rooney, A.; Salleh, B.; Scandiani, Ma. M.; Scauflaire, J.; Short, D.; Steenkamp, E.; Suga, H.; Summerell, B.; Sutton, D.; Thrane, U.; Trail, F.; Van D., A.; Van E., H.; Viljoen, A.; Waalwijk, C.; Ward, T.; Wingfield, M.; Xu, J.; Yang, X.; Yli-Mattila, T; Zhang, N. 2013. One fungus, one name; defining the genus Fusarium in a scientifically robust way that preserves longstanding use. Phytopathology. 103: 400-408. https://doi.org/10.1094/PHYTO-07-12-0150-LE        [ Links ]

18. Gupta, V. K.; Jain, P. K.; Misra, A. K.; Gaur, R.; Gaur, R. K. 2010. Comparative molecular analysis of Fusarium solani isolates by RFLP and RAPD. Microbiology. 79: 772-776.         [ Links ]

19. Gutiérrez, C. M.; Lara, C. B. N.; Guillén, A. H.; Chávez, B. A. T. 2010. Agroecología de la Franja Aguacatera del Estado de Michoacán. México. Interciencia. 35: 647-653.         [ Links ]

20. Knutsen, A. K.; Torp, M.; Holst-Jensen, A. 2004. Phylogenetic analyses of the Fusarium pose, Fusarium sporotrichioides and Fusarium langsethiae species complex based on partial sequences of the translation elongation factor-1 alpha gene. International Journal of Food Microbiology, 95: 287-295. DOI:10.1016/j.ijfoodmicro.2003.12.007        [ Links ]

21. Kornerup, A.; Wanscher, J. W. 1978. Methuen Handbook of Colour. 3rd. ed. Eyre Methuen London. UK. 22.         [ Links ]

22. Koyyappurath, S.; Atuahiva, T.; Guen, R. Le; Batina, H.; Squin, S. Le; Gautheron, N.; Edel Hermann V.; Peribe, J.; Jahiel, M.; Steinberg, C.; Liew, E. C.; Alabouvette, C.; Beese, P.; Dron, M.; Sache, I.; Laval, V.; Grisoni, M. 2015. Fusarium oxysporum f. sp. radicis-vanillae is the causal agent of root and stem rot of vanilla. Plant Pathology. 1-14. DOI:10.1111/ ppa.12445        [ Links ]

23. Lara, C. B. N. 2008. Tesis doctoral Variabilidad fenotípica y patogénica de Phytophthora cinnamomi Rands en la franja aguacatera de Michoacán. México. Universidad Autónoma de Nayarit. Xalisco. Nayarit.         [ Links ]

24. Leslie, J. F.; Summerell, B. A. 2006. The Fusarium laboratory manual. Blackwell Publishing. Iowa. USA. 387 p.         [ Links ]

25. López-Lima, D.; Carrión, G.; Sánchez-Nava, P.; Desgarennes, D.; Villain, L. 2020. Fungal diversity and Fusarium oxysporum pathogenicity associated with coffee corky-root disease in México. Revista de la Facultad de Ciencias Agrarias. Universidad Nacional de Cuyo. Mendoza. Argentina. 52(1): 276-292.         [ Links ]

26. Madania, A.; ltawil, M.; Naffaa, W.; Volker, P.; Hawat, M. 2013. Morphological and molecular characterization of Fusarium isolated from maize in Syria. Journal of Phytoathology, 161: 452-458. https://doi.org/10.1111/jph.12085        [ Links ]

27. Martinez, F. E.; Martinez, J. P.; Guillén, S. D.; Peña, C. G.; Hernández, V. M. 2015. Diversidad de Fusarium en las raíces de caña de azúcar (Saccharum o cinarum) en el estado de Morelos. México. Revista Mexicana de Micología. 42: 33-43.         [ Links ]

28. Mishra, P. K.; Tewari, J. P.; Clear, R. M.; Turkington, T. K. 2004. Molecular genetic variation and geographical structuring in Fusarium graminearum. Annals of Applied Biology. 145: 299-307. DOI:10.1111/j.1744-7348.2004.tb00387.x        [ Links ]

29. Mulé G.; Susca, A.; Stea, G.; Moretti, A. 2004. Specific detection of the toxigenic species Fusarium proliferatum from asparagus plants using primers based on calmodulin gene sequences. Microbiology Letters. 230: 235-240. https://doi.org/10.1016/S0378- 1097(03)00926-1        [ Links ]

30. Ochoa Fuentes, Y. M.; Cerna Chávez, E.; Gallegos Morales, G.; Landeros Flores, J.; Delgado Ortiz, J. C.; Hernández Camacho, S.; Rodriguez Guerra, R.; Olalde Portugal, V. 2012. Identificación de especies de Fusarium en semillas de ajo en Aguascalientes. México. Revista Mexicana de Micología. 36: 27-31.         [ Links ]

31. O’Donnell, K.; Nirenberg, H. I.; Aoiki, T.; Cigelnik, E. 2000. A multigene phylogeny of the Gibberella fujikuroi species complex: Detection of additionally phylogenetically distinct species. Mycoscience, 41: 61-78. https://doi.org/10.1007/BF02464387.

32. Ohr, H. D.; Zentmyer, G. A.; Menge, J. A. 1991. Avocado root rot. University of California Cooperative Extension Publication. 110: 3-20.         [ Links ]

33. Pegg, K. G.; Coates, L. L.; Korsten, L.; Harding, R. M. 2007. Enfermedades foliares, del fruto y el suelo. En: Whiley, A. W.; Schaffer, B. y Wolstenholme, B. N. 2007. El palto: botánica, producción y usos. Traducción. Ediciones Universitarias de Valparaiso. Litogarín. Valparaiso. p. 275-310.         [ Links ]

34. Qu, B.; Li, H. P.; Zhang, J. B.; Xu, Y. B.; Huang, T.; Wu, A. B. 2008. Geographic distribution and genetic diversity of Fusarium graminearum and Fusarium asiaticum on wheat spikes throughout China. Plant Pathology. 57: 15-24. https://doi.org/10.1111/j.1365- 3059.2007.01711.x        [ Links ]

35. Rodríguez, A. G.; Betancourt, R. R.; Rodríguez, F. R.; Velázquez, J. J.; Fernández, S. P.; Gómez, D. 2012. Vegetative compatibility groups characterization of Fusarium mexicanum causing mango malformation in Jalisco, México. Revista Mexicana de Fitopatología. 30: 128-140.         [ Links ]

36. Salleh, B. 1994. Current status and control of plant diseases caused by Fusarium in Malaysia. En Biology and Control of Crop Pathogens. Rifai, M. A.; Scott, E. S.; Quebral, F. C.; Dharmaputra, O. S. Biotrop Special Publication Nº 54. Biotrop. Bogor. Indonesia.         [ Links ]

37. SAS Institute. 2017 SAS 9.4 Companion for Windows. 5th ed. SAS Institute Inc. Cary. NC. USA. 700 p.         [ Links ]

38. SIAP. 2016. El Servicio de Información Agroalimentaria y Pesquera. Avances de siembra y cosecha por cultivo. SAGARPA. http://www.siap.gob.mx/cierre-de-la-produccionagricola-porestado/ Consultado 24 de mayo de 2017.         [ Links ]

39. Téliz, O. D.; A. Mora, A. 2007. Enfermedades del aguacate. El aguacate y su manejo integrado. 2ª ed. Mundi Prensa. Ixtapaluca. Edo. de México. p. 171-205.         [ Links ]

40. Thangavelu, R.; Kumar, M. K.; Devi, G. P.; Mustaffa, M. M. 2012. Genetic diversity of Fusarium oxysporum f. sp. cubense Isolates (Foc) of India by inter simple sequence repeats (ISSR) analysis. Molecular Biotechnology. 51: 203-211. DOI 10.1007/s12033-011-9457-8        [ Links ]

41. Tomoika, K.; Hirooka, Y.; Takezaki, A.; Aoiki, T.; Sato, T. 2011. Fusarium root rot of praire gentian caused by a species belonging to the Fusarium solani species complex. Journal of General Plant Pathology. 77: 132-135. DOI10.1007/s10327-011-0295-0        [ Links ]

42. Trigiano, N. R.; Windham, T. M.; Windham, S. A. 2004. Plant pathology, concepts and laboratory exercises. CRC PRESS. 413 p.         [ Links ]

43. Ullstrup, A. 1970. Method for inoculating corn ears with Gibberella zeae and Diplodia maidis. Plant Disease report. 54: 658-662        [ Links ]

44. Vogelgsang, S.; Sulyok, M.; Bänziger, I.; Krska, R.; Schuhmacher, R.; Forrer, H. R. 2008. Effect of fungal strain and cereal substrate on in vitro mycotoxin production by Fusarium poae and Fusarium avenaceum. Food Additives and Contaminants. 25: 745-757. https://doi. org/10.1080/02652030701768461        [ Links ]

45.Watanabe, M.; Yonezawa, T.; Lee, K.; Kumagai, S.; Sugita-Konishi, Y.; Goto, K.; Hara-Kudo, Y. 2011. Evaluation of genetic markers for identifying isolates of the species of the genus Fusarium. Journal of the Science of Food and Agriculture. 91: 2500-2504. Doi org/10.1002/jsfa.4507        [ Links ]

46. Whiley, A. W.; Schaffer, B.; Wolstenholme, B. N. 2007. El palto. Botánica, producción y usos. Ediciones Universitarias de Valparaiso. p. 9-11.         [ Links ]

47. Wilfinger, W. W.; Mackey, K.; Chomczynski, P. 1997. Effect of pH ionic strength on the spectrophotometric assessment of nucleic acid purity. BioThechniques. 22: 474-481. DOI: 10.2144/97223st01        [ Links ]

48. Zaccardelli, L.; Vitale, S.; Luongo, L.; Meright, M.; Corazza, L. 2008. Morphological and molecular characterization of Fusarium solani isolates. Journal of Phytopathology. 156: 534-541. https://doi.org/10.1111/j.1439-0434.2008.01403.x        [ Links ]

ACKNOWLEDGEMENTS

The authors would like to thank P/PFCE-2017-16MSU0014T-04 for the financial support provided to conduct this research, as well as for its publication.

Creative Commons License Todo el contenido de esta revista, excepto dónde está identificado, está bajo una Licencia Creative Commons