You are on page 1of 10

Plant Disease  2021  105:3101-3110  https://doi.org/10.

1094/PDIS-01-21-0006-RE

Diversity, Pathogenicity, and Fungicide Sensitivity of Fungal Species Associated


with Late-Season Rots of Wine Grape in the Mid-Atlantic United States
Scott D. Cosseboom and Mengjun Hu†
Department of Plant Science and Landscape Architecture, University of Maryland, College Park, MD 20742

Abstract
Late-season bunch rots cause major losses in grape production every respective controls in the first season, but only A. uvarum and Botrytis
year in the Mid-Atlantic United States, but the causal agents are not cinerea caused this in the second season. Also, wounding was found to
well characterized. In this study, 265 fungal isolates were collected have a detrimental effect on cluster weight, which was significantly
from rotten grapes from 2014 to 2020 and identified to the genus level influenced by inoculation timing and cultivar. Lastly, A. uvarum and
according to internal transcribed spacer sequences. The most prevalent N. rosae were tested for sensitivity to azoxystrobin, boscalid, and dife-
of the 15 genera were Botrytis, Colletotrichum, Aspergillus, Alternaria, noconazole. The A. uvarum isolates were found to be more sensitive to
Pestalotiopsis, and Neopestalotiopsis. Of these, isolates within three boscalid and difenoconazole in general, with varying sensitivity to azox-
prevalent, yet understudied, genera were identified to be Aspergillus ystrobin. N. rosae isolates were resistant to boscalid and azoxystrobin
uvarum, Alternaria alternata, and Neopestalotiopsis rosae. The patho- but displayed much higher sensitivity to difenoconazole. Evidence
genicity of these three fungal species was evaluated in two field trials by from the isolate collection and field trials demonstrates that A. uvarum
artificially inoculating wounded and nonwounded grapes (Vitis vinifera) could be a significant pathogen of wine grapes in the Mid-Atlantic
of four cultivars at the phenological stages of bloom, veraison, and pre- United States. Results from this study will be useful for the identifica-
harvest. Upon ripening, fruit were weighed and assessed for severity of tion and management of the understudied Alternaria, Aspergillus, and
multiple diseases. On nonwounded fruit, A. uvarum caused significantly Neopestalotiopsis fruit rots of wine grapes.
higher disease severity than the control in both seasons. On wounded
fruit, each inocula caused significantly higher disease than the Keywords: fungi, fruit rots, pathogenicity, wine grape

In grape production, fruit rotting diseases that occur late in the sea- was more prevalent than Alternaria tenuissima (Mikusova et al.
son cause serious losses every year, especially in the Mid-Atlantic 2014). Within Aspergillus, 36 species have been isolated from grapes,
region of the United States. These diseases are known to be caused with most belonging to Aspergillus section Nigri, the black aspergilli
by Botrytis and Colletotrichum spp. (Adamo 2016; Oliver 2016), (Rousseaux et al. 2014). In 2014, the genus Pestalotiopsis was split
and multiple other fungi have been associated with rotten fruit in the into the genera Pestalotiopsis, Neopestalotiopsis, and Pseudopestalo-
Mid-Atlantic United States, including Alternaria, Aspergillus, and tiopsis (Maharachchikumbura et al. 2014), and multiple species of Pes-
Pestalotiopsis (Kepner and Swett 2018). In fact, >70 fungi associated talotiopsis and Neopestalotiopsis have been isolated from rotten grapes
with fruit rots have been reported on grapes, but their roles in causing (Deng et al. 2013; Jayawardena et al. 2015; Kepner and Swett 2018;
fruit rots in different cultivars largely remain unknown (Wilcox et al. Xu et al. 1999).
2015). Additionally, the genera of many previously reported fungi The second challenge is that fruit rot disease pressure is typically
have been expanded, and multiple species per genera have been iso- high in the Mid-Atlantic United States. First, warm, humid, and wet
lated from diseased grapes (Jayawardena et al. 2015; Lorenzini and ripening seasons provide an ideal environment for fungal infection.
Zapparoli 2014; Rousseaux et al. 2014; Walker et al. 2011). Of the Second, a large proportion of Mid-Atlantic U.S. grape acreage contains
many fungal species associated with grape late-season rots, it is unclear “French” cultivars (Vitis vinifera) that are desirable for their wine qual-
which are most prevalent and contribute significantly to fruit rots in the ity but tend to be susceptible to some foliar and fruit rotting diseases.
emerging Mid-Atlantic U.S. wine grape industry. Other “French-American” hybrid cultivars (V. vinifera × Vitis spp.) are
The relatively new wine grape industry in this region lacks a history also planted in this region and tend to be more disease tolerant (Wilcox
of foundational pathology research. The first obstacle in controlling et al. 2015). The susceptibility of grape cultivars to fruit rots besides
late-season rots of grape is identification of the causal agent. Symptoms black rot, Botrytis bunch rot, and ripe rot has not been thoroughly
on fruit can appear similar between diseases, and multiple fungi can be investigated.
isolated from the same fruit (Samuelian et al. 2011). In Botrytis, two One further obstacle is that species within the same genus isolated
primary species have been identified causing grape late-season rots from rotten fruit may have similar morphologies but differ in fungicide
with Botrytis cinerea being the most prevalent followed by Botrytis sensitivity (Dowling et al. 2020). Research has been conducted on the
pseudocinerea (Walker et al. 2011). Similarly, Alternaria alternata fungicide sensitivity of the frequently isolated fungi A. alternata
(Avenot and Michailides 2020; Iacomi-Vasilescu et al. 2004), B. cin-

erea (Adamo 2016), and Colletotrichum spp. (Oliver 2016); however,
Corresponding author: M. Hu; mjhu@umd.edu little is known about the sensitivity of Aspergillus and Pestalotiopsis
spp. associated with late-season rots. Aspergillus niger and Aspergil-
Funding: This work was supported by the Foundation for Food and Agri-
culture Research under award no. FF-NIA20-0000000062. lus flavus were tested for sensitivity to thiophanate-methyl, tebucona-
zole, iprodione (Thomidis et al. 2009), boscalid, captan (Serey et al.
*The e-Xtra logo stands for “electronic extra” and indicates there are sup- 2007), and pyraclostrobin (Latorre et al. 2002) using single fungicide
plementary materials published online. concentrations in vitro or on detached fruit. Other studies suggested
that Pestalotiopsis spp. were sensitive to carbendazim and iprodione
The author(s) declare no conflict of interest. (Saju et al. 2011), but carbendazim-resistant isolates have been found
(Yong et al. 2014).
Accepted for publication 26 February 2021. The latest study, as of this writing, on the various late-season rots in
the Mid-Atlantic United States isolated multiple fungi from grapes and
© 2021 The American Phytopathological Society assessed the pathogenicity of Aspergillus japonicus, A. alternata, and

Plant Disease / October 2021 3101


Pestalotiopsis telopeae on wounded and nonwounded detached fruit through detached fruit assays (Jayawardena et al. 2015). The b-tubulin
(Kepner and Swett 2018). All three fungi were able to initiate fruit and calmodulin genes of Aspergillus isolates were amplified with pri-
rot consistently on wounded fruit, and fungi in these genera have mers Bt2a/Bt2b and Cmd5/Cmd6, respectively. The ITS region and
also been isolated from asymptomatic (Barbetti 1980; Da Rocha RNA polymerase subunit II (rpb2) gene of Alternaria isolates was
Rosa et al. 2002; Kakalıkova et al. 2009; Serra et al. 2006; Steel et al. amplified with primers ITS1/4 and Rpb2-5f2/Rpb2-7cr, respectively.
2007) and symptomatic berries in other studies (Barkai-Golan 1980; The b-tubulin and translation elongation factor 1-a (tef1) genes of
Jayawardena et al. 2015; Latorre et al. 2002; Lorenzini and Zapparoli Neopestalotiopsis isolates were amplified with primers BT2A/BT2B
2014; Nair 1985). These fungi may play a significant role in fruit rots of and 728F/EF2, respectively. Each isolate was identified to the species
grape, but the ability of these fungi to cause disease in a field setting has level by comparing the sequences with previously published sequences
not yet been assessed. It is also not known if cultivars differ in suscep- of species in each genus (Jayawardena et al. 2015; Maharachchikumbura
tibility or if grapes are ontogenically resistant as with black rot (Molitor et al. 2014; Samson et al. 2014; Woudenberg et al. 2013).
and Berkelmann-Loehnertz 2011), Botrytis bunch rot (McClellan and In-field pathogenicity evaluation. Three fungi, A. alternata,
Hewitt 1973), and powdery mildew (Gadoury et al. 2003), where the Aspergillus uvarum, and Neopestalotiopsis rosae, were tested for their
bloom period tends to be more susceptible to infection. To address ability to cause fruit rot diseases in a vineyard after being collected and
these knowledge gaps, a multiyear study was undertaken in Mid- identified to the species level as described above. Three isolates of each
Atlantic U.S. vineyards to understand the diversity of species associated fungus plus three isolates of the well-established pathogen B. cinerea
with late season rots; the in-field pathogenicity of the most prevalent yet were selected to be prepared as inoculum. To induce sporulation, A.
understudied species among different cultivars and inoculation condi- uvarum and B. cinerea were cultured on PDA at 22 C in the dark,
tions; and the fungicide sensitivity of associated fungi. A. alternata was cultured on quarter-strength PDA at 22 C in the
dark, and N. rosae was cultured on PDA at 25 C under constant fluores-
cent light. Spore suspensions of a mixture of the isolates of each species
Materials and Methods were created by flooding the plates with sterile water and liberating the
Isolate collection and identification. In the late summer and early conidia by rubbing with a sterile cell spreader. The spore suspensions
fall seasons from 2014 to 2020, ripe grapes with fruit rot symptoms were filtered through sterile, double-layered cheesecloth and diluted to
from 18 Mid-Atlantic U.S. (Maryland, Pennsylvania, and Virginia) 1 × 105 conidia/ml and were kept in a chilled atomizer for no longer
vineyards were collected, and fungi were isolated from them. If the than 4 h before inoculation in the field (Amiri et al. 2018).
fungi on the rotten fruit were sporulating, then spores were aseptically The trials took place at an experimental vineyard at the University of
transferred to a 100-mm Petri dish containing potato dextrose agar Maryland Wye Research and Education Center in 2019 and 2020. At
(PDA). Fruit without sporulation were surface sterilized, cut into small three phenological timings, bloom (BBCH 65), veraison (BBCH 83),
pieces, and plated onto PDA. The fruit were sterilized by submerging and preharvest (BBCH 85), healthy-appearing grape clusters of four
in a solution of 1% sodium hypochlorite for 1 min followed by rinsing cultivars, Chardonnay, Chambourcin, Cabernet Franc, and Merlot,
with sterile deionized water. After fungi had grown on the medium, a were inoculated with spore suspensions or water as a control by mist-
hyphal tip was transferred to a new plate of PDA containing ten ing with an atomizer until runoff (Lorenz et al. 1995). For the first and
25-mm2 pieces of sterile filter paper and incubated for 3 to 4 days at second years, 30 and 16 clusters per pathogen–timing–cultivar combi-
22 C. Once the mycelium had grown to the edge of the plate, these fil- nation were inoculated, respectively. Half of the clusters per
ter paper pieces were dried in a desiccator for 2 weeks, then placed in a pathogen–timing–cultivar combination were wounded by piercing
1.5-ml centrifuge tube containing silica gel beads and stored at –20 C. the cuticle of 10 berries per cluster with a sterile toothpick. Clusters
The isolates were then categorized into groups according to their inoculated at the bloom timing were wounded at preharvest. For the
morphology, and up to 15 isolates of each morphological group other two inoculation timing treatments of veraison and preharvest,
were saved per vineyard. DNA was extracted from mycelia of a ran- the clusters were wounded immediately before inoculation. Because
dom selection of isolates from each morphology group (Chi et al. the cultivars in the trial matured at different rates, the inoculation,
2009). The internal transcribed spacer (ITS) region of these isolates wounding, and harvest dates differed by cultivar (Supplementary Table
was amplified through PCR with previously described primers (ITS1 S1). Any fungicides applied during the two trials were not effective
and ITS4) and thermocycling protocol. All primer sequences and refer- against the inocula and were primarily targeting foliar downy mildew
ences of thermocycling protocols used in this study are listed in Table 1. with limited applications for powdery mildew.
All PCRs in this study contained a total volume of 20 ll with 1× Taq Immediately after inoculation, the clusters were covered with white,
Master Mix (Apex Bioresearch Products, San Diego, CA), 1 ll of wax-paper bags to promote infection by prolonging wetness duration
DNA template, 0.5 lM of each primer, and purified water. Purification and to prevent unwanted infection or injury from other fungi, birds,
and sanger sequencing of amplicons was conducted by the Arizona or insects (Karajeh 2018). The bags were left on the clusters until har-
State University Genomic Facility (Tempe, AZ), and the sequences vest. For the bloom-inoculated clusters, the bags were temporarily
were identified using the tool BLAST (https://blast.ncbi.nlm.nih.gov/ removed at the preharvest stage, wounded, and immediately replaced.
Blast.cgi). The wounded-bloom inoculation treatment was not included for the B.
Five isolates each of Aspergillus spp., Alternaria spp., and Neopes- cinerea and water inoculum in 2019. At harvest, the clusters were cut
talotiopsis spp. identified by the ITS sequencing were selected for fur- from the vines, weighed, and evaluated for disease severity. Disease
ther characterization because of their prevalence in Mid-Atlantic U.S. severity was measured by visually estimating the percentage of berries
vineyards and understudied nature. Neopestalotiopsis was chosen over with disease per cluster. If more than one disease was observed on a
Pestalotiopsis because it was shown to be a more severe pathogen cluster, the severity of each disease was noted. Fungi causing the

Table 1. Primers used for PCR amplification and sequencing


Primer name Gene Primer sequence (59-39) Reference
BT2A b-tubulin GGTAACCAAATCGGTGCTGCTTTC Glass and Donaldson (1995)
BT2B b-tubulin ACCCTCAGTGTAGTGACCCTTGGC Glass and Donaldson (1995)
Cmd5 Calmodulin CCGAGTACAAGGAGGCCTTC Hong et al. (2005)
Cmd6 Calmodulin CCGATAGAGGTCATAACGTGG Salah et al. (2019)
ITS1 ITS TCCGTAGGTGAACCTGCGG White et al. (1990)
ITS4 ITS TCCTCCGCTTATTGATATGC White et al. (1990)
Rpb2-5f2 Rpb2 GGGGWGAYCAGAAGAAGGC Sung et al. (2007)
Frpb2-7cr Rpb2 CCCATRGCTTGYTTRCCCAT Liu et al. (1999)
728f Tef1 CATCGAGAAGTTCGAGAAGG Carbone and Kohn (1999)
EF2 Tef1 GGARGTACCAGTSATCATGTT O’Donnell and Cigelnik (1997)

3102 Plant Disease / Vol. 105 No. 10


different diseases were isolated from the clusters and were identified with five different inocula (A. alternata, A. uvarum, B. cinerea, N.
according to their morphology. rosae, and water) was compared with the disease severity on clusters
Sensitivity to azoxystrobin, boscalid, and difenoconazole. The inoculated with water. This was conducted separately for each year,
five isolates of A. uvarum and N. rosae that were identified above and disease, inoculation timing, and wound treatment with a Steel multiple
used as inoculum were tested for their sensitivity to azoxystrobin comparisons test with the water inoculum as the control. The water
(Abound Flowable; Syngenta Crop Protection, Basel, Switzerland), control treatment was not included for nonwounded clusters at the
boscalid (Endura; BASF Crop Protection, Research Triangle Park, bloom timing in 2019 and, therefore, no statistical analysis was con-
NC) and difenoconazole (Inspire; Syngenta Crop Protection). Serial ducted on this disease severity data. The cluster weight data were
dilutions of the fungicides mixed with PDA were used to determine square root transformed to meet analysis-of-variance assumptions of
the effective concentration that inhibits mycelial growth by 50% homogeneity of variance and normal distribution of residuals. The
(EC50). The media were autoclaved and cooled to 55 C before the fun- effect of the inoculum, wounding, cultivar, and inoculation timing
gicide was mixed into the media in a laminar flow hood. Azoxystrobin and interactions of these factors on the square root of the cluster
was amended at the concentrations of 0.01, 0.1, 0.3, 1, 3, 10, and 100 weights was evaluated by analysis of variance. Post hoc comparisons
lg/ml; boscalid at the concentrations of 0.1, 1, 3, 10, 30, and 100 lg/ of the variables on cluster weight were conducted with a Tukey’s Hon-
ml; and difenoconazole at the concentrations of 0.01, 0.03, 0.3, 1, 3, 10, estly Significant Difference test. The square root-transformed cluster
and 100 lg/ml. A 25-mm2 agar plug of each isolate was transferred to weight data were then back transformed to present results. The fungi-
the center of three plates of each concentration for all fungicides and cide sensitivity data were analyzed by determining the percent inhibi-
three plates of nonamended PDA. The plates were then incubated at tion at each concentration, then regressing the log of the concentration
22 C until mycelial growth on nonamended plates reached at least against the average percent inhibition at each concentration. This
80% of the diameter of the 100-mm Petri dishes. At this point, the regression was then used to estimate the concentration that inhibits
diameter of mycelial growth on each concentration was measured in growth by 50% compared with the control for each isolate.
two perpendicular directions using digital calipers. This experiment
was conducted twice.
Statistical analysis. All statistical analyses in this project were con- Results
ducted using the software JMP Pro 14.1 (SAS Institute, Cary, NC). The Isolate collection and identification. From 2014 to 2020, clusters
severity of four different diseases (Alternaria fruit rot, Aspergillus fruit with various late-season bunch rot symptoms were collected. In total,
rot, Botrytis bunch rot, and Neopestalotiopsis fruit rot) was recorded at 265 fungal isolates were obtained from 37 cultivars grown in 18 vine-
harvest. The severity of each disease occurring on clusters inoculated yards located in Maryland (238 isolates), Pennsylvania (18 isolates),
and Virginia (nine isolates). Sequences of the ITS region revealed
that the isolates belonged to 15 genera, and the most common genera
were Botrytis, Colletotrichum, Aspergillus, Alternaria, Pestalotiopsis,
and Neopestalotiopsis (Figs. 1 and 2). Isolates collected from V. vinif-
era cultivars (as opposed to hybrid cultivars) constituted 87, 94, and
100% of total isolates collected from Maryland, Pennsylvania, and Vir-
ginia, respectively. The five isolates each of Aspergillus, Alternaria,
and Neopestalotiopsis chosen for multilocus sequencing were identi-
fied as A. uvarum, A. alternata, and N. rosae.

Fig. 1. Genera of fungal isolates (n = 265) collected from ripe, rotten grape berries
during 2014 to 2020 from 18 Mid-Atlantic U.S. vineyards that had been identified Fig. 2. Genera of isolates collected from wine grapes with late-season fruit rot
through sequencing of the internal transcribed spacer region. symptoms from 2014 to 2020 in Pennsylvania, Maryland, and Virginia.

Plant Disease / October 2021 3103


A variety of symptoms was observed on the clusters sampled in this appeared as a pure white web of mycelium, at times surrounded by
study, and it was not always possible to predict which fungus would be small black bumps (Fig. 4C). Three other diseases were immediately
isolated from each cluster. However, some symptoms and fungi were identified as Botrytis bunch rot, ripe rot, and powdery mildew. The
consistently associated with each other. Botrytis spp. were associated characterization of each disease except powdery mildew fruit rot was
with soft fruit, water-soaked lesions, and profuse whitish gray sporula- confirmed by isolation and morphological identification of the causal
tion (Botrytis bunch rot; Fig. 3C and D). Colletotrichum spp. were iso- agents. Ripe rot severity was low in both years, and powdery mildew
lated from fruit that were shriveled, had rough skin, or orange was not present in the first year but was very severe in the second year
sporulation (ripe rot; Fig. 3B, E, and F). A. uvarum was mostly isolated on the foliage and the fruit (data not shown).
from clusters with shriveled berries and light to dark brown powdery In general, all four fruit rotting diseases of interest occurred in the
sporulation (Aspergillus fruit rot; Fig. 3A and F). A. alternata was iso- experiment in both years and disease tended to be more severe in
lated from berries with dense, dark brown to dark green mycelium 2020 than in 2019 except for Botrytis bunch rot. On nonwounded
(Alternaria fruit rot; Fig. 3C). Pestalotiopsis and Neopestalotiopsis fruit, the primary disease occurring on the clusters was usually asso-
were consistently isolated from fruit with white mycelial growth ciated with the inoculum. However, it was common for more than
(Pestalotiopsis/Neopestalotiopsis fruit rot; Fig. 3E and F). A. uvarum, one disease to occur on fruit inoculated with a single pathogen,
A. alternata, Pestalotiopsis, and Neopestalotiopsis were also frequently and fungi of different genera could be isolated from them. For
isolated from clusters displaying ambiguous symptoms of shriveling, example, four different diseases were observed on fruit inoculated
bruising, or rough skin, and fungi of more than one genus could often with B. cinerea in 2020 (Fig. 5). The water inoculation treatment
be isolated from the same cluster (Fig. 3C, E, and F). for nonwounded fruit resulted in relatively little disease severity
Pathogenicity of A. alternata, A. uvarum, B. cinerea, and N. of either of the four diseases in both years, except for Alternaria
rosae. The mature fruit of the four cultivars in the trials were evaluated and Neopestalotiopsis fruit rot in 2020 (Fig. 5A and D). After
for disease and six different disease symptoms were most common. the treatments for nonwounded fruit, clusters inoculated with A.
The causal agents of Alternaria, Aspergillus, and Neopestalotiopsis uvarum at the preharvest timing in 2019 and at the veraison timing
fruit rots were isolated and morphologically identified as A. alternata, in 2020 had significantly greater Aspergillus fruit rot severity than
A. uvarum, and N. rosae, respectively. Alternaria fruit rot appeared as the control (Fig. 5B). No other inocula caused its associated disease
prostrate growth of dense, dark green to dark brown mycelium on the to be significantly greater than the control in either year on non-
fruit surface (Fig. 4A). Clusters with Aspergillus fruit rot had tan to wounded fruit (Fig. 5). Over both seasons, nonwounded Chardon-
dark brown tufts of conidiophores that could be stained pink or red nay clusters tended to have more severe fruit rots than the other
by the berry skin (Fig. 4B). Berries with Aspergillus fruit rot infections cultivars, whereas Chambourcin tended to have the least severe fruit
also tended to easily detach from the rachis. Neopestalotiopsis fruit rot rot on nonwounded clusters (Supplementary Fig. S1).

Fig. 3. Clusters of wine grapes collected from Mid-Atlantic U.S. vineyards with various symptoms of late-season bunch rot from which A, Aspergillus; B, Colletotrichum; C,
Alternaria and Botrytis; D, Botrytis; E, Colletotrichum and Neopestalotiopsis; and F, Aspergillus, Colletotrichum, and Pestalotiopsis spp. were isolated.

3104 Plant Disease / Vol. 105 No. 10


Wounded fruit tended to have greater disease severity than non- Discussion
wounded fruit in both years, and it was also common for clusters inoc-
ulated with one fungus to exhibit diseases associated with other fungi Each of the 15 fungal genera identified from the Mid-Atlantic U.S.
(Fig. 6). For example, compared with the control, significant levels of isolates collected in this study have been previously associated with
Aspergillus fruit rot occurred on clusters inoculated with A. alternata, late-season rots of grapes. Each genus has also been demonstrated to
A. uvarum, and N. rosae in 2019, but only on clusters inoculated with be pathogenic on detached fruit except for Sporobolomyces, a yeast
A. uvarum in 2020 (Fig. 6B). Significantly higher Alternaria fruit rot associated with sour rot of grapes, and Curvularia (Barbetti 1980; Jaya-
severity was observed only on fruit inoculated with A. alternata in wardena et al. 2015; Lederer et al. 2013; Lorenzini et al. 2015; Rajput
2019 at the preharvest timing (Fig. 6A). Also, only clusters inoculated et al. 2020; Rousseaux et al. 2014; Steel et al. 2007). For the fungal spe-
with B. cinerea had significant severities of Botrytis bunch rot in both cies included in the fungicide sensitivity and pathogenicity experiments,
years (Fig. 6C), while clusters inoculated with N. rosae had signifi- A. alternata and A. uvarum have been isolated from diseased grapes
cantly higher Neopestalotiopsis fruit rot in 2019 only (Fig. 6D). The (Mikusova et al. 2010; Steel et al. 2007), but N. rosae has not. N. rosae
severity of disease on different wounded cultivars was not consistent has been reported as a pathogen of eucalyptus (Santos et al. 2020), blue-
between the 2 years (Supplementary Fig. S2). berry (Rodrıguez-Galvez et al. 2020), and strawberry (Rebollar-Alviter
In 2019 and 2020, the factors of cultivar, inoculation timing, and et al. 2020; Baggio et al. 2021). Because of the subdivision of the genus
wounding had a significant impact on cluster weight, while the inocu- Pestalotiopsis into Pestalotiopsis, Neopestalotiopsis, and Pseudopesta-
lum did not. In 2019, there was no interaction between wounding and lotiopsis (Maharachchikumbura et al. 2014), it is possible that N. rosae
any other variable on cluster weight, and wounded fruit were found to has been endemic to Mid-Atlantic U.S. grapes for years but never char-
weigh less than nonwounded fruit. In 2019, there was a significant cul- acterized as this species. It is important to note that the collection and
tivar × inoculation timing interaction, and Chambourcin inoculated and identification of isolates within Alternaria, Aspergillus, and Neopes-
bagged at veraison and preharvest had the highest weight according to talotiopsis was limited, and other species within these genera are
a post hoc Tukey’s Honestly Significant Difference test (Fig. 7). In likely present in Mid-Atlantic U.S. vineyards (Kepner and Swett
2020 there was a significant cultivar × inoculation timing interaction 2018). A more focused study on one of these genera may reveal other
and a significant wounding × inoculation timing interaction, but no species that differ in pathogenicity and fungicide sensitivity. Further-
interaction between all three factors. Similar to the first year, treatments more, the limited number of isolates collected from Virginia and
for wounded fruit tended to result in lower cluster weights than the Pennsylvania likely did not accurately represent the distribution of
treatments for nonwounded fruit at all three inoculation timings in fungal species associated with late-season rots in these states.
2020 (Supplementary Fig. S3). Further, Cabernet Franc and Merlot This was the first study to conduct baseline sensitivity testing of N.
clusters inoculated at bloom weighed significantly more than those rosae and A. uvarum to azoxystrobin, boscalid, and difenoconazole.
inoculated at veraison and preharvest (Fig. 7). These fungicides were chosen because they represent three important
Azoxystrobin, boscalid, and difenoconazole sensitivity. Five chemical classes of fungicides (quinone outside inhibitors, succinate
isolates of A. uvarum and N. rosae were tested for their sensitivity to dehydrogenase inhibitors, and demethylation inhibitors) that are
the fungicides azoxystrobin, boscalid, and difenoconazole. All N. rosae already used for control of other grape bunch rots. The N. rosae isolates
isolates were insensitive to boscalid and azoxystrobin (EC50 > 100 lg/ in this study were found to be insensitive to boscalid and azoxystrobin;
ml), but sensitivity to difenoconazole ranged from 0.30 to 5.06 lg/ml however, more isolates may need to be tested along with molecular
(Table 2). Sensitivity of A. uvarum isolates ranged 0.63 to >100 lg/ml analyses to determine if this insensitivity is intrinsic or acquired.
for azoxystrobin, 0.59 to 4.73 µg/ml for boscalid, and 0.18 to 1.12 lg/ Only one other study has begun to assess the fungicide sensitivity of
ml for difenoconazole. this species. Rebollar-Alviter (2020) found N. rosae isolates were

Fig. 4. Disease symptoms of A, Alternaria fruit rot, B, Aspergillus fruit rot, and C, Neopestalotiopsis fruit rot from clusters inoculated with A. alternata, A. uvarum, and N.
rosae, respectively. Clusters were sourced from a replicated field trial in Maryland.

Plant Disease / October 2021 3105


sensitive to field rates of captan, fludioxonil + cyprodinil, difenocona- A. uvarum was demonstrated to be the most pathogenic of these fungi,
zole, prochloraz, and iprodione in vitro. Prior research on A. uvarum is at times causing higher disease severity than the well-established path-
also limited, but the five A. uvarum isolates in this study were found to ogen B. cinerea. The ability of A. uvarum to cause significant disease
have similar sensitivities to closely related fungi within Aspergillus severity on nonwounded fruit confirms this fungus as a primary path-
section Nigri tested for sensitivity to the clinically used demethylation ogen (Fig. 5B). A. alternata, B. cinerea, and N. rosae did not cause sig-
inhibitors itraconazole and voriconazole (Badali et al. 2016). Also, nificant levels of disease severity on nonwounded fruit. Because B.
three of five isolates of A. uvarum were insensitive to azoxystrobin, cinerea is a known primary pathogen of grape, these results hint that
suggesting that resistance is widespread and quinone outside inhibitors environmental conditions may not have been ideal for infection
would not be an effective option for preventing Aspergillus fruit rot. by this fungus. Conditions may have not been ideal for the other
The pathogenicity of the inocula A. alternata, A. uvarum, and N. inocula as well. During the bloom, veraison, and preharvest inoc-
rosae in a vineyard setting was tested for the first time in this study. ulations, the weather was often very warm and sunny, and the

Fig. 5. Average severity of A, Alternaria, B, Aspergillus, C, Botrytis, and D, Neopestalotiopsis fruit rot across nonwounded clusters of four wine grape cultivars inoculated with
four fungi and water at bloom, veraison, and preharvest in field trials conducted in 2019 and 2020. One (*P < 0.05), two (**P < 0.005), or three (***P < 0.0005) asterisks
indicate that the disease severity was significantly greater than the water-inoculated treatment that was set as the control for the Steel multiple comparisons tests. Each
disease/year/inoculation timing combination was evaluated with the Steel multiple comparisons test separately, and error bars represent standard error.

3106 Plant Disease / Vol. 105 No. 10


water-based spore suspensions may have quickly dried, not allow- the inoculum, increasing the chance for infection (Broome et al. 1995).
ing the fungi to infect the fruit. The bags may have also altered the ambient environment around the
Bagging was a critical component of the field experiments conducted clusters throughout the season, providing shade but also perhaps trap-
in this study because the grape ripening season in Maryland often entails ping heat. Drain holes on the downward facing side of the bags may
high disease pressure and no fungicides were applied to the clusters dur- have aided in equalizing the temperature inside and outside of the
ing the trials. Bags have been shown to effectively reduce the severity of bags. Although the bagging resulted in less than natural development
fruit rotting diseases (Karajeh 2018), and clusters that were protected by conditions for the clusters, nontarget disease issues were avoided and
bags throughout the season appeared healthier than nonbagged fruit conclusions of the pathogenicity of each inocula were not in jeopardy
(data not shown). As the bags were placed on the fruit immediately after because disease severity was compared relative to control clusters
inoculation, the bags may have also prolonged the wetness duration of that were bagged in the same manner.

Fig. 6. Average severity of A, Alternaria, B, Aspergillus, C, Botrytis, and D, Neopestalotiopsis fruit rot across wounded clusters of four wine grape cultivars inoculated with
four fungi and water at bloom, veraison, and preharvest in a field trial conducted in 2019 and 2020. One (*P < 0.05), two (**P < 0.005), or three (***P < 0.0005) asterisks
indicate that the disease severity was significantly greater than the water-inoculated treatment that was set as the control for the Steel multiple comparisons tests. Each
disease/year/inoculation timing combination was evaluated with the Steel multiple comparisons test separately, and error bars represent standard error. The wounded-bloom
inoculation treatment was not included for the B. cinerea and water inoculum in 2019.

Plant Disease / October 2021 3107


The wounding treatments for fruit facilitated infection by the inoc- cultivars except Chambourcin. Previous studies have suggested that
ula, and higher disease severity was observed on these clusters. On powdery mildew infection occurring early in the season can cause
the wounded fruit, A. uvarum and B. cinerea were found to cause sig- wounds in the berry cuticle, providing an entry point for secondary fun-
nificantly more Aspergillus fruit rot (Fig. 6B) and Botrytis bunch rot gal pathogens (Gadoury et al. 2003). Although clusters appearing
(Fig. 6C) than the water-inoculated control in both years, respectively. healthy were selected for inoculations, latent or diffuse powdery mil-
Interestingly, both A. alternata and N. rosae were only found to cause dew infections are not visible to the unaided eye (Gadoury et al.
significantly more Alternaria (Fig. 6A) or Neopestalotiopsis fruit rot 2007). It is therefore possible that the high levels of Alternaria and
(Fig. 6D) than the control in 2019 at the preharvest timing, respec- Neopestalotiopsis fruit rots on water-inoculated clusters in 2020 were
tively. Also, during the preharvest timing in 2019, clusters inoculated secondary infections resulting from the severe powdery mildew epi-
with the A. uvarum and B. cinerea inoculum displayed significantly demic. Interestingly, little to no Botrytis bunch rot was observed on
less Alternaria and Neopestalotiopsis fruit rot than the control (Fig. clusters that were not inoculated with the pathogen. The conditions
6A and D). This is likely because these inocula caused very high in the field may not have been suitable for Botrytis bunch rot to natu-
severities of their respective diseases (Aspergillus fruit rot and Botrytis rally occur during the two seasons of this study, and little Botrytis
bunch rot), thereby outcompeting other less aggressive pathogens like bunch rot was observed in other Maryland vineyards during this season
A. alternata or N. rosae. The ability of these four fungi to cause signif- (data not shown). The prevention of wounding appears to be critical to
icant levels of disease on wounded fruit means that these fungi were controlling the majority of the late-season bunch rots.
able to act as secondary pathogens. Cultivar was not a factor in the statistical analysis of disease severity,
Highly severe diseases emerged that were not inoculated with an however there may be differences in susceptibility between the culti-
associated pathogen. For example, there were significant levels of vars tested. French-American hybrid grapevines (e.g., Chambourcin)
Aspergillus fruit rot on wounded clusters not inoculated with the asso- tend to be more tolerant to some foliar and fruit rotting diseases than
ciated pathogen A. uvarum (Fig. 6B). There were also high levels of French cultivars (e.g., Cabernet Franc, Chardonnay, and Merlot)
Alternaria fruit rot on wounded clusters inoculated with water (Fig. (Wilcox et al. 2015). This trend could be seen on the nonwounded fruit,
6A). This was likely the result of native inoculum in the vineyard, where the average disease severity on Chambourcin was consistently
with infection aided by the fruit wounding. Other naturally occurring low compared with the other cultivars (Supplementary Fig. S1). This
diseases appeared on the fruit such as ripe rot and powdery mildew. was not true on the wounded fruit, for wounding may have allowed
Powdery mildew was widespread and very severe in 2020 in all the pathogens to overcome varietal resistances (Supplementary Fig.

Fig. 7. Average cluster weight of four wine grape cultivars from replicated field trials in 2019 and 2020 that were inoculated with five inocula at the phenological stages of
bloom, veraison, and preharvest. Error bars represent standard error and bars labeled with different letters are significantly different according to a post hoc Tukey’s Honestly
Significant Difference test conducted independently for each year.

Table 2. Range and mean of EC50 values (µg/ml) for five isolates of Aspergillus uvarum and five isolates of Neopestalotiopsis rosae tested for sensitivity to
azoxystrobin, boscalid, and difenoconazole in two separate trials
A. uvarum N. rosae
Fungicide Range Mean ± SE Range Mean ± SE
Azoxystrobin 0.63 to >100 NCa >100 NCa
Boscalid 0.59 to 4.73 2.51 ± 0.53 >100 NCa
Difenoconazole 0.18 to 1.12 0.62 ± 0.10 0.30 to 5.06 1.81 ± 0.61
a
Some isolates had higher EC50 values than could be measured (above 100 lg/ml); therefore, standard error (SE) was not calculated. NC = not calculated.

3108 Plant Disease / Vol. 105 No. 10


S2). It is also noteworthy that most isolates (88%) were collected from Barbetti, M. J. 1980. Bunch rot of Rhine Riesling grapes in the lower south-west
V. vinifera cultivars, as opposed to hybrid cultivars. This could be of Western Australia. Aust. J. Exp. Ag. 20:247-251.
Barkai-Golan, R. 1980. Species of Aspergillus causing post-harvest fruit decay in
because of varietal resistance resulting in fewer hybrid clusters being
Israel. Mycopathologia 71:13-16.
sampled or because a greater proportion of Mid-Atlantic U.S. acreage Broome, J. C., English, J. T., Marois, J. J., Latorre, B. A., and Aviles, J. C.
consists of V. vinifera cultivars. 1995. Development of an infection model for Botrytis bunch rot of grapes
The phenological stage of the grape cluster is an important factor for based on wetness duration and temperature. Phytopathology 85:97-102.
disease infection for Botrytis bunch rot, black rot, and powdery mildew Carbone, I., and Kohn, L. M. 1999. A method for designing primer sets for spe-
(Gadoury et al. 2003; McClellan and Hewitt 1973; Molitor and ciation studies in filamentous ascomycetes. Mycologia 91:553-556.
Chi, M.-H., Park, S.-Y., and Lee, Y.-H. 2009. A quick and safe method for fungal
Berkelmann-Loehnertz 2011), and this may be an important factor DNA extraction. Plant Pathol. J. 25:108-111.
for A. uvarum infection as well. In both years, the inoculation of A. uva- Da Rocha Rosa, C. A., Palacios, V., Combina, M., Fraga, M. E., Rekson, A. D.
rum at veraison and preharvest resulted in significantly more Aspergil- O., Magnoli, C. E., and Dalcero, A. M. 2002. Potential ochratoxin A pro-
lus fruit rot than the control (Figs. 5 and 6). The inoculation timing also ducers from wine grapes in Argentina and Brazil. Food Addit. Contam. 19:
had a significant effect on the weight of the clusters; however, this was 408-414.
likely because of the protective bagging rather than the inoculum Deng, J. X., Sang, H. K., Hwang, Y. S., Lim, B. S., and Yu, S. H. 2013. Post-
harvest fruit rot caused by Pestalotiopsis sp. on grape in Korea. Australas.
because the inoculum did not have a significant effect on the weight. Plant Dis. Notes 8:111-114.
The protective bags were placed on the clusters at the point of inocu- Dowling, M., Peres, N., Villani, S., and Schnabel, G. 2020. Managing Colletotri-
lation and may have protected the fruit from subsequent detrimental chum on fruit crops: a “complex” challenge. Plant Dis. 104:2301-2316.
effects from the environment and other diseases and pests. For exam- Gadoury, D. M., Seem, R. C., Ficke, A., and Wilcox, W. F. 2003. Ontogenic
ple, clusters bagged and inoculated at bloom in 2020 tended to weigh resistance to powdery mildew in grape berries. Phytopathology 93:
547-555.
more than those bagged and inoculated at veraison and preharvest
Gadoury, D. M., Seem, R. C., Wilcox, W. F., Henick-Kling, T., Conterno, L.,
(Supplementary Fig. S3). Bloom-inoculated clusters were protected Day, A., and Ficke, A. 2007. Ecology and epidemiology effects of diffuse col-
from external pressures like the severe powdery mildew outbreak in onization of grape berries by Uncinula necator on bunch rots, berry microflora,
2020 from bloom until harvest (approximately 4 months), while the and juice and wine quality. Phytopathology 97:1356-1365.
other clusters were only protected from veraison until harvest (approx- Glass, N. L., and Donaldson, G. C. 1995. Development of primer sets designed
imately 2 months) or from preharvest until harvest (about 1 month). for use with the PCR to amplify conserved genes from filamentous ascomy-
cetes. Appl. Environ. Microbiol. 61:1323-1330.
However, bagging at the fragile bloom stage also posed risks and dam- Hong, S.-B., Go, S.-J., Shin, H.-D., Frisvad, J. C., and Samson, R. A. 2005. Poly-
aged clusters in this trial, at times breaking the inflorescence phasic taxonomy of Aspergillus fumigatus and related species. Mycologia 97:
completely off the grapevine. This may explain the lower cluster 1316-1329.
weights observed on bloom bagged clusters in 2019. If all clusters Iacomi-Vasilescu, B., Avenot, H., Bataille-Simoneau, N., Laurent, E., Guenard,
to be inoculated in these trials were bagged at bloom, only to be M., and Simoneau, P. 2004. In vitro fungicide sensitivity of Alternaria species
removed for the inoculation and wounding treatments at veraison pathogenic to crucifers and identification of Alternaria brassicicola field iso-
lates highly resistant to both dicarboximides and phenylpyrroles. Crop Prot.
and preharvest, a clearer difference between inoculation timings may 23:481-488.
have occurred. Jayawardena, R. S., Zhang, W., Liu, M., Maharachchikumbura, S. S. N., Zhou,
The diversity, pathogenicity, and fungicide sensitivity of under- Y., Huang, J., Nilthong, S., Wang, Z., Li, X., Yan, J., and Hyde, K. D. 2015.
studied fungal species involved in late-season rots of wine grapes in Identification and characterization of Pestalotiopsis-like fungi related to grape-
the Mid-Atlantic U.S. region was assessed in this study. The primary vine diseases in China. Fungal Biol. 119:348-361.
 arova, A. 2009. First report of Alterna-
Kakalıkova, L'., Jankura, E., and Srob
fruit rotting pathogens were found to be A. uvarum, Botrytis spp.,
ria bunch rot of grapevines in Slovakia. Australas. Plant Dis. Notes 4:
and Colletotrichum spp., while A. alternata and N. rosae appear to 68-69.
act as secondary pathogens in the Mid-Atlantic United States. This is Karajeh, M. R. 2018. Pre-harvest bagging of grape clusters as a non-chemical
the first study confirming the in-field pathogenicity of A. uvarum, physical control measure against certain pests and diseases of grapevines.
which caused the most disease when infection occurred during the Org. Agric. 8:259-264.
veraison or preharvest stages. Chemical management efforts should Kepner, C., and Swett, C. L. 2018. Previously unrecognized diversity within fun-
gal fruit rot pathosystems on Vitis vinifera and hybrid white wine grapes in
be focused on primary pathogens, and boscalid and difenoconazole Mid-Atlantic vineyards. Australas. Plant Pathol. 47:181-188.
were found to be effective against A. uvarum. An integrated approach Latorre, B. A., Viertel, S. C., and Spadaro, I. 2002. Severe outbreaks of bunch
can be taken to reduce other fruit rotting diseases by growing disease- rots caused by Rhizopus stolonifer and Aspergillus niger on table grapes in
resistant hybrid cultivars and by preventing fruit wounding from pri- Chile. Plant Dis. 86:815.
mary pathogens, insects, weather events, and birds. Lederer, M. A., Nielsen, D. S., Toldam-Andersen, T. B., Herrmann, J. V., and
Arneborg, N. 2013. Yeast species associated with different wine grape varie-
ties in Denmark. Acta Agr. Scand. 63:89-96.
Acknowledgments Liu, Y. J., Whelen, S., and Hall, B. D. 1999. Phylogenetic relationships among
ascomycetes: evidence from an RNA polymerase II subunit. Mol. Biol. Evol.
The authors thank Joe Fiola at the University of Maryland and Cain Hickey and
16:1799-1808.
Bryan Hed at Penn State for assisting in the collection of samples. They also thank
Lorenz, D. H., Eichhorn, K. W., Bleiholder, H., Klose, R., Meier, U., and Weber,
Anita Schoeneberg and Michael Newell at the University of Maryland for their
E. 1995. Growth stages of the grapevine: phenological growth stages of the
technical support. Some isolates used in this study were kindly provided by Cassan-
grapevine (Vitis vinifera L. ssp. vinifera) − codes and descriptions according
dra Swett at the University of California, Davis. The content of this manuscript is
to the extended BBCH scale. Aust. J. Grape Wine Res. 1:100-103.
solely the responsibility of the authors and does not necessarily represent the offi-
Lorenzini, M., Cappello, M. S., and Zapparoli, G. 2015. Isolation of Neofusicoc-
cial views of the Foundation for Food and Agriculture Research.
cum parvum from withered grapes: strain characterization, pathogenicity and
its detrimental effects on passito wine aroma. J. Appl. Microbiol. 119:1335-
1344.
Literature Cited Lorenzini, M., and Zapparoli, G. 2014. Characterization and pathogenicity of
Adamo, N. R. 2016. Fungicide resistance of Botrytis cinerea from Virginia wine Alternaria spp. strains associated with grape bunch rot during post-harvest
grapes, strawberry, and ornamental crops. Master’s thesis. Virginia Polytech- withering. Int. J. Food Microbiol. 186:1-5.
nic Institute and State University, Blacksburg, VA. Maharachchikumbura, S. S. N., Hyde, K. D., Groenewald, J. Z., Xu, J., and
Amiri, A., Zuniga, A. I., and Peres, N. A. 2018. Potential impact of populations Crous, P. W. 2014. Pestalotiopsis revisited. Stud. Mycol. 79:121-186.
drift on Botrytis occurrence and resistance to multi- and single-site fungicides McClellan, W. D., and Hewitt, W. B. 1973. Early Botrytis rot of grapes: time of
in Florida Southern highbush blueberry fields. Plant Dis. 102:2142-2148. infection and latency of Botrytis cinerea Pers. in Vitis vinifera L. Phytopathol-
Avenot, H. F., and Michailides, T. J. 2020. Occurrence and extent of boscalid ogy 63:1151-1157.
resistance in populations of Alternaria alternata from California pistachio  arova, A. 2010. Con-
Mikusova, P., Ritieni, A., Santini, A., Juhasova, G., and Srob
orchards. Plant Dis. 104:306-314. tamination by moulds of grape berries in Slovakia. Food Addit. Contam. 27:
Badali, H., Fakhim, H., Zarei, F., Nabili, M., Vaezi, A., Poorzad, N., Dolatabadi, S., 738-747.
and Mirhendi, H. 2016. In vitro activities of five antifungal drugs against oppor-  arova, A. 2014. Alternaria mycotoxins asso-
Mikusova, P., Sulyok, M., and Srob
tunistic agents of Aspergillus nigri complex. Mycopathologia 181:235-240. ciated with grape berries in vitro and in situ. Biologia 69:173-177.
Baggio, J. S., Forcelini, B. B., Wang, N. Y., Ruschel, R. G., Mertely, J. C., and Molitor, D., and Berkelmann-Loehnertz, B. 2011. Simulating the susceptibility of
Peres, N. A. 2021. Outbreak of leaf spot and fruit rot in Florida strawberry clusters to grape black rot infections depending on their phenological develop-
caused by Neopestalotiopsis spp. Plant Dis. 105:305-315. ment. Crop Prot. 30:1649-1654.

Plant Disease / October 2021 3109


Nair, N. G. 1985. Fungi associated with bunch rot of grapes in the Hunter Valley. Samuelian, S. K., Greer, L. A., Savocchia, S., and Steel, C. C. 2011. Detection
Aust. J. Ag. R. 36:435-442. and monitoring of Greeneria uvicola and Colletotrichum acutatum develop-
O’Donnell, K., and Cigelnik, E. 1997. Two divergent intragenomic rDNA ITS2 ment on grapevines by real-time PCR. Plant Dis. 95:298-303.
types within a monophyletic lineage of the fungus Fusarium are nonortholo- Santos, G. S., Mafia, R. G., Aguiar, A. M., Zarpelon, T. G., Damacena, M. B.,
gous. Mol. Phylogenet. Evol. 7:103-116. Barros, A. F., and Ferreira, M. A. 2020. Stem rot of eucalyptus cuttings caused
Oliver, C. 2016. Investigation of wine grape cultivar and cluster developmental by Neopestalotiopsis spp. in Brazil. J. Phytopathol. 168:311-321.
stage susceptibility to grape ripe rot caused by two fungal species complexes, Serey, R. A., Torres, R., and Latorre, B. A. 2007. Pre-and post-infection activity
Colletotrichum gloeosporioides, and C. acutatum, and the evaluation of poten- of new fungicides against Botrytis cinerea and other fungi causing decay of
tial controls. Master’s thesis. Virginia Polytechnic Institute and State Univer- table grapes. Cienc. Investig. Agrar. 34:215-224.
sity, Blacksburg, VA. Serra, R., Mendonça, C., and Ven^ancio, A. 2006. Fungi and ochratoxin A detected in
Rajput, N. A., Hussainullah, Huo, C., Cao, J., Atiq, M., Atif, R. M., Lodhi, A. healthy grapes for wine production. Lett. Appl. Microbiol. 42:42-47.
M., Syed, R. N., Sarfraz, S., Hameed, A., and Zhao, Z. 2020. First report of Steel, C. C., Greer, L. A., and Savocchia, S. 2007. Studies on Colletotrichum
Curvularia verruculosa causing leaf spot disease of grape (Vitis vinifera) in acutatum and Greeneria uvicola: two fungi associated with bunch rot of grapes
Afghanistan. J. Plant Pathol. 102:1337. in sub-tropical Australia. Aust. J. Grape Wine Res. 13:23-29.
Rebollar-Alviter, A., Silva-Rojas, H. V., Fuentes-Aragon, D., Acosta-Gonzalez, Sung, G.-H., Sung, J.-M., Hywel-Jones, N. L., and Spatafora, J. W. 2007. A
U., Martınez-Ruiz, M., and Parra-Robles, B. E. 2020. An emerging strawberry multi-gene phylogeny of Clavicipitaceae (Ascomycota, Fungi): identification
fungal disease associated with root rot, crown rot and leaf spot caused by Neo- of localized incongruence using a combinational bootstrap approach. Mol.
pestalotiopsis rosae in Mexico. Plant Dis. 104:2054-2059. Phylogenet. Evol. 44:1204-1223.
Rodrıguez-Galvez, E., Hilario, S., Lopes, A., and Alves, A. 2020. Diversity and Thomidis, T., Michailides, T., and Exadaktylou, E. 2009. Contribution of patho-
pathogenicity of Lasiodiplodia and Neopestalotiopsis species associated with gens to peach fruit rot in Northern Greece and their sensitivity to iprodione,
stem blight and dieback of blueberry plants in Peru. Eur. J. Plant Pathol. carbendazim, thiophanate-methyl and tebuconazole fungicides. J. Phytopathol.
157:89-102. 157:194-200.
Rousseaux, S., Diguta, C. F., Radoï-Matei, F., Alexandre, H., and Guilloux- Walker, A. S., Gautier, A., Confais, J., Martinho, D., Viaud, M., Le P^echeur, P.,
Benatier, M. 2014. Non-Botrytis grape-rotting fungi responsible for earthy Dupont, J., and Fournier, E. 2011. Botrytis pseudocinerea, a new cryptic spe-
and moldy off-flavors and mycotoxins. Food Microbiol. 38:104-121. cies causing gray mold in French vineyards in sympatry with Botrytis cinerea.
Saju, K. A., Mech, S., Deka, T. N., and Biswas, A. K. 2011. In vitro evaluation Phytopathology 101:1433-1445.
of biocontrol agents, botanicals and fungicides against Pestalotiopsis sp. White, T. J., Bruns, T., Lee, S., and Taylor, J. 1990. PCR Protocols: A Guide to
infecting large cardamom (Amomum subulatum Roxb.). J. Spices Aromat. Methods and Applications. Academic Press, New York.
Crops 20:89-92. Wilcox, W. F., Gubler, W. D., and Uyemoto, J. K. 2015. Compendium of Grape
Salah, H., Lackner, M., Houbraken, J., Theelen, B., Lass-Fl€orl, C., Boekhout, T., Diseases, Disorders, and Pests, 2nd ed. APS Press, St. Paul, MN.
Almaslamani, M., and Taj-Aldeen, S. J. 2019. The emergence of rare clinical Woudenberg, J. H. C., Groenewald, J. Z., Binder, M., and Crous, P. W. 2013.
Aspergillus species in Qatar: molecular characterization and antifungal suscep- Alternaria redefined. Stud. Mycol. 75:171-212.
tibility profiles. Front. Microbiol. 10:1677. Xu, L., Kusakari, S.-i., Hosomi, A., Toyoda, H., and Ouchi, S. 1999. Postharvest dis-
Samson, R. A., Visagie, C. M., Houbraken, J., Hong, S. B., Hubka, V., Klaassen, ease of grape caused by Pestalotiopsis species. Jpn. J. Phytopathol. 65:305-311.
C. H. W., Perrone, G., Seifert, K. A., Susca, A., and Tanney, J. B. 2014. Yong, Y. C., Chen, Y. J., Fang, B. Y., and Chung, W. H. 2014. Sensitivity of
Phylogeny, identification and nomenclature of the genus Aspergillus. Stud. Pestalotiopsis spp. from guava to benzimidazoles in Taiwan. Plant Pathol.
Mycol. 78:141-173. Bull. 23:271-275.

3110 Plant Disease / Vol. 105 No. 10

You might also like