Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/352467625

Antifungal activity of zinc oxide nanoparticles in Fusarium oxysporum‐


Solanum lycopersicum pathosystem under controlled conditions

Article in Journal of Phytopathology · June 2021


DOI: 10.1111/jph.13023

CITATIONS READS

39 743

6 authors, including:

Agustín Hernández Juárez Rebeca Betancourt


Universidad Autónoma Agraria Antonio Narro (UAAAN) Centro de Investigación en Química Aplicada
67 PUBLICATIONS 227 CITATIONS 38 PUBLICATIONS 467 CITATIONS

SEE PROFILE SEE PROFILE

Yisa María Ochoa-Fuentes


Universidad Autónoma Agraria Antonio Narro (UAAAN)
153 PUBLICATIONS 623 CITATIONS

SEE PROFILE

All content following this page was uploaded by Yisa María Ochoa-Fuentes on 24 January 2022.

The user has requested enhancement of the downloaded file.


Received: 10 December 2020 | Revised: 20 May 2021 | Accepted: 1 June 2021

DOI: 10.1111/jph.13023

ORIGINAL ARTICLE

Antifungal activity of zinc oxide nanoparticles in Fusarium


oxysporum-­Solanum lycopersicum pathosystem under controlled
conditions

Ana María González-­Merino1 | Agustín Hernández-­Juárez1 |


Rebeca Betancourt-­Galindo2 | Yisa María Ochoa-­Fuentes1 |
Luis Alonso Valdez-­Aguilar3 | Mónica Lorena Limón-­Corona1

1
Departamento de Parasitología,
Universidad Autónoma Agraria Antonio Abstract
Narro, Saltillo, México Antifungal activity of zinc oxide (ZnO) and ZnO nanoparticles (ZnO NPs) was evalu-
2
Centro de Investigación en Química
ated on the control of Fusarium oxysporum Schltdl. (Nectriaceae) under laboratory
Aplicada, Saltillo, México
3
Departamento de Horticultura, Universidad
and greenhouse conditions. In vitro evaluation, poisoned culture media was pre-
Autónoma Agraria Antonio Narro, Saltillo, pared, and an explant was placed in the centre of solid medium. The experimental
México
design was completely randomized with 18 treatments. Mycelial growth and co-
Correspondence nidia concentration were evaluated. Subsequently, three treatments (3,000, 1,500,
Agustín Hernández-­Juárez, Departamento
de Parasitología, Universidad Autónoma
100 ppm) of ZnO NPs and ZnO were chosen for their evaluation in the greenhouse in
Agraria Antonio Narro, Calzada Antonio tomato plants of Floradade variety under a randomized block design. Inoculation was
Narro # 1923, C.P. 25315. Buenavista,
Saltillo, Coahuila, México.
carried out with a 1x107 conidia per mL suspension of F. oxysporum, when the plants
Email: chinoahj14@hotmail.com presented the third pair of true leaves. Later, the application of different concentra-
tions of ZnO and ZnO NPs was carried out; in this investigation, the incidence and
severity and plant height were evaluated into account to determine the treatment ef-
fect on F. oxysporum. In vitro, the best treatments in mycelial growth inhibition were
the high concentrations of ZnO NPs from 1,600 to 3,000 ppm with 81%–­83%, and
in the sporulation of the fungus, they were also those that inhibited from 82.57% to
83.85%. In greenhouse, the treatments that reached the highest plant height were
ZnO NPs from 1,500 to 3,000 ppm, with a range of 166.0–­175.40 cm, with a severity
on the scale of 0.40–­0.80 and an incidence of 20%–­40%. ZnO NPs have a potential
application as an antifungal agent and can be used to control the spread of F. oxyspo-
rum in tomato plants, in addition to improving the promoter effect related to the Zinc
activity as a precursor in auxins synthesis, cytokinins and gibberellins biosynthesis,
as well as the induction of higher activity of antioxidant enzymes useful in response
to the pathogens attack.

KEYWORDS

disease, nanoparticles, tomato, toxicity

Journal of Phytopathology. 2021;00:1–12. wileyonlinelibrary.com/journal/jph © 2021 Wiley-VCH GmbH | 1


2 | GONZÁLEZ-­MERINO et al.

1 | I NTRO D U C TI O N 2019; Rajiv et al., 2013) resulted in a reduction in the presence of


diseases.
The tomato Solanum lycopersicum L. (Solanaceae) is the most There is little information on applications of ZnO NPs as fungi-
important vegetable worldwide. China is the most important cides for F. oxysporum in tomato, which is why this research has the
producer and tomato consumer, while the United States is the objective of evaluating the antifungal activity of zinc oxide nanopar-
main importer and Mexico is the main exporter of this vegetable ticles for the control of F. oxysporum under controlled laboratory
(FIRA, 2019). Mexico stands out with a production of 3,441,639.37 conditions and greenhouse in tomato S. lycopersicum.
t and a production value of $ 29 874 007.50 (MXN) (SIAP, 2020),
in addition, whose cultivation, harvest and marketing generate
72 thousand direct jobs and approximately 10.7 million indirect 2 | M ATE R I A L S A N D M E TH O DS
jobs. The production of this crop is affected by a large number
of phytopathogenic microorganisms that affect its performance, 2.1 | Experiment location
such as bacteria, fungi and viruses. One of the main fungi that
most affects the crop are of the genus Fusarium Link (Nectriaceae) Research was conducted at the Entomologia Molecular Laboratory
that cause losses between 21% and 47% in protected and open-­ and greenhouse of Universidad Autónoma Agraria Antonio Narro, in
field crops (Cardona-­Piedrahíta & Castaño-­Z apata, 2019; Mejía & Saltillo, Coahuila, Mexico.
Hernández, 2001).
Fusarium oxysporum Schltdl. (Nectriaceae) is the causal agent
of tomato wilt disease. The main symptoms of the disease include 2.2 | Obtaining zinc oxide
chlorosis, progressive wilting of leaves, discoloration of vascular
tissue, growth retardation and up to causing the death of the plant Zinc oxide powder was obtained commercially at Sanely Marketer.
(Abdallah et al., 2016, 2019).
In conventional agriculture, chemical fungicides are the main tool
for the control of phytopathogenic fungi. The continuous and exces- 2.3 | Synthesis and characterization of
sive use of these products has resulted in the development of resis- nanoparticles
tance to the fungicides used (Coromoto & Reyes, 2018). Chemical
products enter the human body and remain in it, causing diseases ZnO NPs were synthesized at Centro de Investigación en Química
such as cancer, permanent damage in children born with genetic Aplicada, in Saltillo, Coahuila. They were synthesized through
damage with deterioration, environmental damage such as soil ster- controlled precipitation according to Betancourt et al. (2010)
ilization and the desertification of large regions (Rivero et al., 2014). technique, by the chemical hydrolysis method as follow: 13.7 g
Currently, the search for control alternatives has increased, such of Zn(O2CCH 3)2 and 600 ml of ethanol were placed in a ball flask
as the use of nanotechnology in agriculture; whose applications in- with three necks. This solution was constantly stirred at a tem-
clude fertilizers to increase plant growth and yield, pesticides for perature of 75℃ under reflux for 2 hr. Then, an aqueous solu-
pest and disease management, and sensors to monitor soil quality tion of 0.22 M NaOH and an additional 100 ml of distilled H2O
and plant health (Servin et al., 2015). were added to complete the reaction mixture. Constant stir-
Nanoparticles (NPs) are materials between 1 and 100 nanome- ring was continued for 24 hr. Subsequently, the ZnO NPs ob-
ters (nm) in size and exist as metalloids, metal oxides, non-­metals tained immersed in ethanol were recovered by centrifugation
and carbon nanomaterials, etc., whose small size, large surface area at 15,000 rpm for 5 min for their recovery. The precipitate was
and high reactivity have allowed their use in disease control and as washed two times with ethanol and dried in an oven at 60℃ for
nanofertilizers (Elmer & White, 2018). Nanoparticles have gained im- 24 hr. The dried ZnO NPs were crushed in an agate mortar to ob-
portance as novel antimicrobial agents due to their strong germicidal tain a fine powder and stored at room temperature until use. The
properties, against a wide range of multi-­resistant microorganisms size and morphology of nanoparticles were measured by means
(Singh et al., 2010). of a high-­r esolution transmission electronic microscope (HRTEM)
Zinc oxide nanoparticles (ZnO NPs) in different crops of eco- Titan 80–­3 00 kV (FEI Company).
nomic interest have been reported to have properties such as
growth promoters, bactericides and fungicides (Esparza, 2016;
Méndez-­Argüello et al., 2016); in the latter, it has been pointed 2.4 | Isolation and purification of F. oxysporum
out that their antifungal activity against Penicillium expansum
Link. (Eurotiales: Trichocomaceae), Botrytis cinerea Pers.: Fr. To obtain the strains of the fungus, plant tissue infected by F. ox-
(Helotiales: Sclerotinicaeae), Aspergillus flavus Link., Aspergillus ysporum was collected from the tomato crop; for its isolation, the
niger, P.E.L. van Tieghem, Aspergillus fumigatus Fresenius tissue was sectioned with a scalpel, then it was disinfected with
(Eurotiales: Trichocomaceae), Fusarium culmorum (Wm.G. Sm.) Sacc. a triple wash with 3% hypochlorite, distilled water sterile (dH2O),
(Hypocreales: Nectriaceae) y F. oxysporum (Jamdagni et al., 2018, 70% alcohol and dH20 and later it was seeded in culture medium
GONZÁLEZ-­MERINO et al. | 3

Potato Dextrose Agar (PDA) (Bioxon®) and incubated at 27 ± 2℃ in a 100 ppm) and an absolute control (PDA = 0 ppm) with 4 replicates,
®
growth chamber (Yamato ), and from their growth, the strains were and it considered a Petri dish as one repetition.
purified by hypha point.

2.6.3 | Mycelial growth evaluation and conidia


2.5 | Morphological and molecular identification concentration

Morphological identification was carried out by mycelium struc- The evaluation consisted of measuring the mycelial growth, in two
tures and their respective conidia on slide with a lactophenol and axes (horizontal and vertical) of the phytopathogen, every 24 hr,
cotton blue solution. Its observation was performed at 40 and until the filling of the control (8 days), using as support a calliper
100X in a compound microscope, and for identification at the genus (vernier) with digital compound (Titan®). To count the conidia, the
level, it was supported by the taxonomic keys for imperfect fun- mycelium from Petri dishes was collected in dH2O, the suspen-
gal genera of Barnett and Hunter (1998) and at the species level sion was shaken to homogenize the concentration of conidia, and
with the keys of Leslie and Summerell (2006). Molecular identifi- then, 10 µl was taken and placed in the hemocytometer (Marienfeld
cation was performed by extracting DNA from the fungus, using Germany), following the steps and formulas of the guide provided
the CTAB protocol (Lee & Taylor, 1990). DNA was resuspended in by Celeromics (Total Cells Counted / Number of Frames × 10,000).
30 µl of nuclease-­free water and stored at −20℃. The polymerase With the data obtained in the evaluation, the per cent of inhibition
chain reaction (PCR) was carried out in 25 µl containing 12.5 µl of of mycelial growth was calculated using the formula described by
Gotaq® DNA Polymerase (Promega corporation), 1 µl of each oligo- Orberá et al. (2009), taking as 100% the mycelial growth of the con-
nucleotide, FOF1: (5'-­ACATACCACTTGTTGCCTCG-­3') and FOR1: trol (0 ppm).
(5'-­CGCCAATCAATTTGAGGAACG-­3 ') at 0.35 µM and 20 ng of
genomic DNA, and the final volume was adjusted with nuclease-­free
( )
R1 − R2
PIRG = × 100
water. The amplification conditions were as follows: an initial dena- R1
turation of 1 min at 95℃, followed by 25 cycles of 94℃ for 1 min,
52℃ for 30 s, 72℃ for 1 min and a final extension of 7 min at 72℃. Where:
The length of the amplified fragments was 340 bp, which were visu- PIRG: is the per cent of inhibition of radial growth.
alized on a 1% agarose gel (Mishra et al., 2003). R1: represents the average value of the radius of the fungus
growth (control).
R2: is the average value of the radius of the inhibited colony
2.6 | Evaluation in vitro (treatment).

2.6.1 | Culture medium poisoned with ZnO NPs


2.7 | Evaluation in vivo
PDA culture medium added with 5g/L of malt extract was prepared.
It was placed in a pressure cooker (Presto Model 79,291, capacity 2.7.1 | Vegetative material
21 L) for 15 min at 121℃ and 15 pounds for sterilization.
ZnO NPs and ZnO were prepared in glass tubes with 10 ml of Tomato seedlings of S. lycopersicum Floradade variety type ball
deionized water, and the corresponding concentrations of nanopar- were used (Fax de Occidente S.A. de C.V.). Seeds were planted in
ticles and ZnO were added. The ZnO NPs were dispersed by soni- expanded polystyrene germination trays of 200 cavities in peat
cation in three cycles of 15 min. After the sterilization, the culture moss-­perlite substrate in a 3:1 ratio, and the transplant was car-
medium was allowed to cool, which was completed with the previ- ried out 4 weeks after planting in polyethylene containers with a
ously sonicated NPs solution, stirred for a period of 15–­20 min to capacity of 8 L, in conditions of 25ºC ± 2ºC and relative humidity
achieve homogenization of the solution with the medium and finally of 60 ± 10% and automated ventilation to reduce heat and renewal
each treatment was emptied into Petri dishes (90 × 15 mm). 5 mm of carbon dioxide.
explant of F. oxysporum was placed in the centre. The Petri dishes
were incubated at a temperature of 27 ± 2ºC for 8 days in a growth
chamber (Yamato®). 2.7.2 | Inoculation and application of treatments

When the seedlings presented the third pair of true leaves, the first
2.6.2 | Experimental design inoculation of F. oxysporum was performed with a suspension con-
taining 1 × 107 conidia. The pathogen was inoculated again 15 and
The in vitro effect of F. oxysporum by ZnO NPs and ZnO was evalu- 30 days after the first inoculation to clearly observe the manifesta-
ated in 8 concentrations (3,000, 2000, 1,600, 1,200, 800, 400, 200, tion of the disease. The different concentrations of ZnO NPs and
4 | GONZÁLEZ-­MERINO et al.

TA B L E 1 Escala para evaluar la


Class Intensity of disease
severidad por Fusarium oxysporum f. sp.
0 No visible symptoms of the disease radicis-­lycopersici en tomate
1 Necrotic spots in hypocotyl
2 Wilted leaves, darkening at the base of the hypocotyl, or decreased plant growth
3 Wilting, necrotic lesions 1–­5 cm and decreased plant growth
4 6–­10 cm necrotic lesion, defoliation and decreased growth
5 Death of the plant

ZnO were sprayed onto the foliage with a manual sprayer (RLFlo phytopathogen. Diseased plant tissue was disinfected and placed on
Master) 1900 ml capacity, the next day after each F. oxysporum PDA as previously described.
inoculation.

2.8 | Analysis of results


2.7.3 | Experimental design
With the percentage of mycelial inhibition and conidia concentra-
In vivo effect over F. oxysporum by ZnO NPs and ZnO was evalu- tion, probit analysis was performed and the values of the inhibitris
ated in three concentrations chosen based on the in vitro evaluation concentration (IC50 and IC90) and their fiducial limits at 95% reliabil-
(3,000, 1,500, 100 ppm) and an absolute control (without inocula- ity were determined. To determine the effect of the treatments on
tion with F. oxysporum = TA) and a control inoculated with spore mycelial inhibition, sporulation inhibition, incidence, severity and
suspension (=T0) with five repiclates each, considered a plant as a plant height in the different treatments, an analysis of variance and
repetition. comparison of means of the treatments with a Tukey multiple range
test (p < .05) were used. The data of mycelial inhibition, sporulation
inhibition and incidence expressed in percentage were transformed
2.7.4 | Evaluation by arcsine square root for analysis. SAS/STAT software was used for
all analyses (SAS, 2002).
The plant height (cm) was determined with the support of a Gripper
tape measure (Truper) at the same time as the damage evaluation by
F. oxysporum. The evaluation of the damage severity by F. oxysporum 3 | R E S U LT S
was carried out 75 days after the first pathogen inoculation, when
the control plants (T0) presented characteristic symptoms, caused High-­resolution transmission electronic microscope micrograph
by the pathogen, using the symptomatological scale of Apodaca-­ of ZnO nanoparticles is shown in Figure 1, which displays semi-­
Sánchez et al. (2004) and Clavijo (2014) (Table 1) and the incidence spherical. The ZnO exhibited a morphology semi-­spherical with a
was calculated with the following formula: I = (IA∕N) × 100where I = size in the range of 10–­4 0 nm with average diameter of (23.44 nm).
Incidence, is the measure of the per cent of individuals affected (IA) In the inhibition of mycelial growth, significant differences were
and showing symptoms of the disease (%) in a population (N). found, whose treatment with the greatest effect on F. oxysporum
was the ZnO NPs at 1,600 and 3,000 ppm with 83.3% and 81.05%
inhibition. The ZnO concentrations of 1,600, 2000 and 3,000 ppm
2.7.5 | Agronomic crop management had the greatest effect with 76.6%, 76% and 78.63% inhibition. In
the low dose treatments, ZnO NPs at 400, 200 and 100 ppm to-
The crop nutritional part was covered with foliar fertiliizer Ferti plus gether with ZnO at 400 ppm inhibited from 30.78% to 47.28% of
+2 L/ha (AGROformuladora DELTA® S.A. DE C.V.) applied to the sub- mycelial growth. However, at low concentrations of ZnO (200 and
strate four times during the crop development. An application of the 100 ppm) they did not show any significant difference compared
insecticide Imidacloprid Confidel 350SC, 1 L/ha (AGROformuladora with the control (Table 2, Figures 2 and 3).
®
DELTA S.A. DE C.V) was made for the management of the whitefly In the conidia count, the treatments with the greatest inhibition
Trialeurodes vaporariorum Westwood (Hemiptera: Aleyrodidae). effect with a percentage greater than 80% are the ZnO NPs of 1,600
and 3,000 ppm, and in the ZnO treatments with the greatest effect
are 1,200–­3,000 ppm with an inhibition of 65.59%–­69.87%, while the
2.7.6 | Phytopathogen recovery low concentrations of ZnO NPs (200, 400 and 800 ppm) together with
ZnO at 800 ppm presented a percentage of inhibition between 55.73%
At the end of the data collection of results, healthy tissue with symp- and 59.73% of sporulation. The lowest concentration of ZnO NPs and
toms of the disease was sectioned to determine the presence of the ZnO at 100 ppm inhibited 41.71% and 28.08%, respectively (Table 2).
GONZÁLEZ-­MERINO et al. | 5

61%, the absolute control (without inoculation) 115% and the treat-
ments with ZnO NPs up to 118%, even higher than the control with-
out inoculation (Figures 4 and 5).

4 | D I S CU S S I O N

In recent years, nanoparticles have been developed for use in


agriculture, with action in the improvement of physiological
processes, promoted as mineral stimulants, particularly essen-
tial micronutrients necessary for host defence, as resistance in-
ducers, reducing the presence of pathogenic organisms (Lamsal
et al., 2011; Servin et al., 2015). These characteristics highlight
the potential application of nanoparticles as antifungal agents to
control the spread and infection of several pathogens (Raskar &
Laware, 2014).
Among the nanoparticles used for crop protection, ZnO NPs are
most recurrent because of their excellent chemical and thermal sta-
bility, low production cost and are not harmful to the environment
F I G U R E 1 Transmission electron microscopy micrograph (Rajiv et al., 2013; Sabri et al. 2013, taken from Kriti et al., 2020).
showing the spherical morphology of ZnO NPs The role of ZnO release from NPs in antimicrobial activity was
tested using microparticles and nanoparticles, demonstrating that
the nano-­specific effects of ZnO play a contributing role in the
For the 50% inhibition of mycelial growth, a lower concentration level of antifungal effects. Bioactivity was more effective with
of ZnO NPs is required compared with ZnO, which requires almost NPs, confirming that nanometer size enhances the effectiveness of
twice as much to inhibit the same percentage of growth. Similarly, ZnO, increasing its toxicity relative to larger microparticles, inhib-
for the 50% inhibition of spores a lower concentration of ZnO NPs iting mycelial growth of the fungus Fusarium graminearum (Dimkpa
is required, while with zinc oxide almost three times as much is re- et al., 2013).
quired to inhibit the same percentage of growth (Table 3). The ability of metal oxide NPs through foliar application to pro-
From the greenhouse evaluation, at the end of the tests, the phy- vide disease resistance is a relatively new and underexplored con-
topathogen was recovered from healthy and diseased tissue and was cept. Initially, it was attributed that the increase in crop growth and
re-­isolated, again coinciding with the F. oxysporum species. yield was the result of inhibition or control of some phytopathogen
It was observed that ZnO NPs at 1,500 ppm presented signifi- by ZnO NPs (Servin et al., 2015). On the other hand, it was consid-
cantly the highest plant growth with 175.4 cm, followed by ZnO NPs ered that the growth-­promoting effect of ZnO NPs can be attributed
at 3,000 ppm with a plant height of 166 cm, the latter without signif- to the activity of Zinc, essential micronutrients for overall plant
icant differences with the absolute control that presented 157.4 cm growth and development and precursor of auxin production that
in plant height. The ZnO compared with the ZnO NPs showed signifi- promote division, cell elongation, have influence on the reactivity
cantly less growth from 106 to 140 cm and with much less growth of indoleacetic acid, which functions as a hormonal phytostimulant,
the inoculated control without protection against F. oxysporum with associated with the biosynthesis of cytokinins and gibberellins; as
a height of 73 cm in plants that ended up dead (Table 4). well as the induction of increased activity of antioxidant enzymes
In terms of disease damage, there were significant differences, useful in the response to pathogen attack (Kriti et al., 2020; Méndez-­
with a higher incidence in the plants corresponding to the inoculated Argüello et al., 2016).
control and ZnO, whereas the plants with application of NPs had ZnO NPs (23.44 nm) demonstrated an effective control over
significantly lower incidence (Table 4). Regarding the severity, there F. oxysporum growth and can be an alternative for the management
were significant differences between the treatments, without sever- of the phytopathogen, as described by Esparza (2016) in their eval-
ity in the absolute control without inoculation, followed by the ZnO uation of the antifungal activity of ZnO NPs on F. oxysporum, who
NPs with less severity with 0.4–­1.2 in the symptomatological scale observed a 91.13% inhibition of growth at 1,000 mg/L concentra-
of damage by F. oxysporum, whereas the ZnO presented the highest tion. However, this observation may vary depending on the phyto-
severity (2.0–­4.8 on the scale) together with the inoculated control pathogen to be controlled; in this sense, Esparza-­Rivera et al. (2014)
without protection against the pathogen with the highest severity observed that ZnO NPs with a size of 200 nm at 2000 ppm against
(Figures 4 and 5). Phytophthora capsici Leonian (Peronosporaceae) and Rhizoctonia
In the present investigation, with respect to the inoculated plants solani J. G. Kühn (Ceratobasidiaceae) inhibited mycelial growth be-
(T0), the plants treated with ZnO showed an increase in height by tween 14% and 16%.
6 | GONZÁLEZ-­MERINO et al.

TA B L E 2 Percent inhibition means,


Concentration Mycelial growth inhibition Inhibition of
sporulation inhibition per cent, ± standard
Treatment (ppm) (%)a,b sporulation (%)a,b
deviation of Fusarium oxysporum in PDA
Testigo 0 0 ± 0j j 0 ± 0.00 h medium at different concentrations of
ZnO 3,000 78.63 ± 1.72 bcd 69.87 ± 1.92 c ZnO and ZnO NPs at 8 days of evaluation

ZnO 2000 76.00 ± 0.91 de 66.81 ± 1.21 c


ZnO 1,600 76.60 ± 0.51 cd 68.06 ± 1.17 c
ZnO 1,200 65.85 ± 2.73 f 65.59 ± 0.49 c
ZnO 800 62.63 ± 0.57 g 56.83 ± 2.79 d
ZnO 400 46.32 ± 1.36 h 48.38 ± 1.88 e
ZnO 200 0 ± 0.00 j 28.81 ± 3.17 g
ZnO 100 0 ± 0.00 j 28.08 ± 2.32 g
ZnO NPs 3,000 81.05 ± 1.84 ab 83.85 ± 0.39 a
ZnO NPs 2000 79.35 ± 0.41 bc 76.38 ± 1.60 b
ZnO NPs 1,600 83.30 ± 0.91 a 82.57 ± 1.49 a
ZnO NPs 1,200 73.25 ± 1.12 e 75.56 ± 0.97 b
ZnO NPs 800 66.40 ± 1.98 f 59.72 ± 2.43 d
ZnO NPs 400 47.28 ± 0.74 h 57.42 ± 1.20 d
ZnO NPs 200 33.35 ± 0.83 i 55.73 ± 1.47 d
ZnO NPs 100 30.78 ± 1.09 i 41.71 ± 5.48 f
CV 2.32 3.79
R2 0.99 0.99
df 16,67 16,67
F 2,386.22c 417.49c
Pr>F <0.0001 <0.0001

Abbreviation: ppm, parts per million.


a
Means with the same letter in same column are not significantly different (Tukey; p < .05).
b
Data for analysis are transformed by arcsine square root.
c
Indicate significance contrast value F to p < .001.

Rajiv et al. (2013) synthesized ZnO nanoparticles from Parthenium Regarding the sporulation of the fungus, with the treatment with
hysterophorus L. (Asteraceae) in different sizes and explored their an- ZnO NPs, its effect was correlated with the mycelial growth, inhib-
tifungal activity against A. flavus and A. niger, showing that ZnO NPs iting the conidia to a greater degree, compared with the control and
are good antifungal agents, inhibiting these phytopathogens with a the eight treatments with ZnO, an effect that can be attributed to
size of 27 ± 5 nm, in addition to being respectful to the environment. the fact that the nanoparticles induce damage to the hypha of the
Dimkpa et al. (2013) ratified the antifungal effect of ZnO NPs on fungus and at the same time to the conidia, a characteristic observed
Fusarium graminearum Schwabe (Nectriaceae), reducing the growth by Jo et al. (2009) and Lamsal et al. (2011) with Ag NPs (silver), who
of the phytopathogen, in addition to being supplemented with indicated that NPs influence the formation of spore colonies, damag-
the biological control agent Pseudomonas chlororaphis O6 Bergey ing and penetrating the cell membrane, and the disease progression
(Pseudomonadaceae), by not inhibiting the metabolites of the bac- of plant pathogenic fungi.
teria, though it inhibit the development of F. graminearum, opening On the one hand, He et al. (2011) describe that ZnO NPs can
the possibility of using them as formulations to complement existing exhibit different antifungal activities. On the other hand, they ob-
strategies to improve crop health. served that P. expansum in PDA with ZnO NPs not only produced
Recently, Pariona et al. (2020) studied the in vitro antifungal ac- conidia but also inhibited their development and distorted conidio-
tivity of ZnO in three forms of particles, including NPs with a size phores, whereas the biomass of B. cinerea developed mainly hyphae,
of 18 ± 2 nm, for three species of fungi: F. oxysporum f.sp. lycop- with apparently more resistance to ZnO NPs, conserving the fine
ersici, Fusarium solani (Mart.) Sacc. (Nectriaceae) and Colletotrichum structure of the mycelia, although the surface of the hyphae of the
gloeosporioides (Penz.) Penz & Sacc. (Glomerellaceae), finding very fungi was deformed.
similar control results with 1 mg/ml against F. oxysporum with 54% Wani and Shah (2012) reported a high rate of inhibition in
and 53% inhibition by ZnO and ZnO NPs, respectively. At the same the germination of spores of the fungi Alternaria alternata (Fr.)
concentration in F. solani, the inhibition was 65% and 55% and for Keissl. (Pleosporales: Pleosporaceae), F. oxysporum, Rhizopus sto-
C. gloeosporioides 60% and 59% by ZnO and ZnO NPs, respectively. lonifer (Ehrenb.) Vuill. and Mucor plumbeus Bonord. (Mucorales:
GONZÁLEZ-­MERINO et al. | 7

F I G U R E 2 Antifungal activity of
(a) (b) (c)
ZnO NPs on Fusarium oxysporum in
vitro test: (a) 0 ppm, (b) 3,000 ppm, (c)
2000 ppm, (d) 1,600 ppm, (e) 1,200 ppm,
(f) 800 ppm, (g) 400 ppm, (h) 200 ppm and
(i) 100 ppm

(d) (e) (f)

(g) (h) (i)

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

F I G U R E 3 Antifungal activity of ZnO


on Fusarium oxysporum in vitro test: (a)
0 ppm, (b) 3,000 ppm, (c) 2000 ppm, (d)
1,600 ppm, (e) 1,200 ppm, (f) 800 ppm,
(g) 400 ppm, (h) 200 ppm and (i) 100
to ppm
8 | GONZÁLEZ-­MERINO et al.

TA B L E 3 Concentration inhibitory of
ppm
mycelial growth and sporulation inhibition
Fiducial limits 95% Fiducial limits 95% and fiducial limits of ZnO and ZnO NPs
applied to Fusarium oxysporum in PDA at
Treatments IC50 Lower Upper IC90 Lower Upper 8 days of evaluation
Mycelial growth inhibition (%)
ZnO NPs 364.40 286.54 446.37 5,017 3,537 8,081
ZnO 776.04 465.94 1,191 3,329 1943 11,401
Sporulation inhibition (%)
ZnO NPs 186.46 111.94 264.71 8,327 4,744 20,327
ZnO 545.68 416.40 694.56 17,072 9,245 43,671

Abbrevaitions: IC50, 50% growth inhibitory concentration; IC90, 90% growth inhibitory
concentration; ppm, parts per million.

TA B L E 4 Plant height, incidence and severity of damage by Fusarium oxysporum ± standard deviation at different concentrations of ZnO
NPs and ZnO at 75 days after the first inoculation

Concentration
Treatments (ppm) Plant height (cm)a Incidence (%)a,b Severitya,c

Absolute control = TA 0 157.4 ± 13.2 ab 0.0 ± 0.0 b 0 ± 0.0 c


Inoculated control = T0 0 73.0 ± 6.02 d 100.0 ± 0.0 a 5.0 ± 0.0 a
ZnO 3,000 140.0 ± 24.3 b 100.0 ± 0.0 a 2.0 ± 0.0 b
ZnO 1,500 108.4 ± 15.2 c 100.0 ± 0.0 a 2.0 ± 0.0 b
ZnO 100 105.6 ± 13.4 c 100.0 ± 0.0 a 4.8 ± 0.45 a
ZnO NPs 3,000 166.0 ± 18.1 ab 40.0 ± 54.7 ab 0.8 ± 1.10 bc
ZnO NPs 1,500 175.4 ± 12.8 a 20.0 ± 44.7 b 0.4 ± 0.89 c
ZnO NPs 100 136.6 ± 13.9 bc 60.0 ± 54.7 ab 1.2 ± 1.10 bc
CV 11.62 48.65 32.19
R2 0.85 0.65 0.90
df 7,39 7,39 7,39
F 25.50 d 8.43d 42.82d
Pr > F <0.0001 <0.0001 <0.0001

Abbreviation: ppm, parts per million.


a
Means with the same letter in same column are not significantly different (Tukey; p < .05).
b
Data for analysis are transformed by arcsine square root.
c
Symptomatic scale of Apodaca-­Sánchez et al. (2004) andClavijo (2014)
d
Indicate significance contrast value F to p < .001.

(a) (b)

F I G U R E 4 Tomato plants S.
lycopersicum (a) inoculated with Fusarium
oxysporum, (b) absolute control without
inoculation of the phytopathogen
GONZÁLEZ-­MERINO et al. | 9

F I G U R E 5 Control of Fusarium (a) (d)


oxysporum with ZnO NPs and ZnO in
Solanum lycopersicum tomato plants. (a)
ZnO NPs at 3,000 ppm, (b) ZnO NPs at
1,500 ppm, (c) ZnO NPs at 100 ppm, (d)
ZnO at 3,000 ppm, (e) ZnO at 1,500 ppm,
(f) ZnO at 100 ppm

(b) (e)

(c) (f)

Rhizpodaceae) to the exposure of ZnO NPs and NPs of magnesium The effect of ZnO NPs on plant growth is not fully distinguished.
oxide (MgO) at concentrations of 100 mg/L. In this regard, Navarro et al. (2008) suggest that nano-­sized materials
Among the antifungal properties, attributed to the nanoscale with larger surface areas could more efficiently absorb, translocate
surface-­to-­volume ratio of ZnO NPs (Kriti et al., 2020), is the in- and retain nutrients in plants.
duction of deformations in the structure of fungal hyphae, nota- In other crops such as Capsicum annuum L. (Solanaceae), Méndez-­
ble thinning of fibres in the hyphae and tendency to agglutinate Argüello et al. (2015) evaluated ZnO NPs doped with Ag at 2.5% in
(Arciniegas-­Grijalba et al., 2017; Yehia & Ahmed, 2013); causing liq- the growth and production of biomass in seedlings, finding only an
uefaction of the cytoplasmic content, making it less electron-­dense, increase in height of 16.8% compared with control plants.
with a significant detachment of the cell wall, which causes cell Most of the studies on NPs in plant pathology have examined
death of the pathogen, causing growth to stop (Arciniegas-­Grijalba the direct antifungal and antibacterial activity against the pathogen
et al., 2017; Sawai & Yoshikawa, 2004; Yehia & Ahmed, 2013). in question (Ocsoy et al., 2013; Saharan et al., 2015; Strayer-­Scherer
Additionally, ZnO NPs generate various other reactive oxygen spe- et al., 2018; Wani & Shah, 2012; Zabrieski et al., 2015).
cies, such as hydroxyl radicals and singlet oxygen, which in turn stim- In the present study, it was demonstrated that foliar application
ulate cell death (Jamdagni et al., 2018). of ZnO NPs inhibits the development of the disease and positively
10 | GONZÁLEZ-­MERINO et al.

affects plant growth, presumably through better mineral nutrition samplings and experiments, revised the original draft and prepared
and host defence. The effect of ZnO at the nanoscale (25 nm) on the final manuscript. Rebeca Betancourt-­Galindo designed the re-
plant growth and development was verified by Prasad et al. (2012) search and methodology and conducted samplings and experiments,
in the peanut crop Arachis hypogaea L. (Fabaceae), observing that and involved in writing—­review and editing the final version of the
at 1,000 ppm it promoted the germination of seeds, plant vigour, manuscript. Yisa María Ochoa-­Fuentes analyzed data, revised the
as well as increased stem and root growth, early flowering and in- original draft and prepared the final manuscript. Luis Alonso Valdez-­
creased chlorophyll content and pod yield, although the increase in Aguilar revised the original draft and prepared the final manuscript
concentration (2000 ppm) presented adverse effects such as toxic- and involved in writing—­review and editing the final version of the
ity in the growth and performance. manuscript; Mónica Lorena Limón-­Corona conducted samplings and
Stampoulis et al. (2009) investigated the effects of five nano- experiments, revised the original draft and prepared the final manu-
materials including ZnO NPs in Cucurbita pepo L. (Cucurbitaceae) script. All authors have read and agreed to the published version of
in a hydroponic medium at 1,000 mg / L, finding that the NPs did the manuscript.
not affect germination, but did present toxicity on root length and
biomass. PEER REVIEW
In watermelon plants, Citrullus lanatus (Thunb.) Matsum. and The peer review history for this article is available at https://publo​
Nakai, Elmer et al. (2018) evaluated the efficiency of the ZnO NPs ns.com/publo​n/10.1111/jph.13023.
and CuO on F. oxysporum and observed that the plants treated with
the nanoparticles had a range of disease classification values signifi- DATA AVA I L A B I L I T Y S TAT E M E N T
cantly lower than the control plants that were not treated. Data are available on request from the authors. All the data are in-
Recently, different antifungal activity of ZnO NPs effective in in- corporated in the manuscript.
terfering with the metabolism of phytopathogenic fungi was deter-
mined; it was observed that ZnO NPs at a concentration of 20 ppm ORCID
were effective in inhibiting spore germination on Bipolaris sorokin- Agustín Hernández-­Juárez https://orcid.
iana (Sorokin) Shoemaker (Pleosporaceae) and at 10 ppm effective org/0000-0001-7059-4471
for Alternaria brassicicola (Schwein.) Wiltshire (Pleosporaceae) and
at 100 ppm significantly inhibit mycelial growth of both pathogens REFERENCES
(Kriti et al., 2020). Abdallah, R. A. B., Jabnoun-­Khiareddine, H., Nefzi, A., Ayed, F., & Daami-­
The lower mycelial growth in vitro and the reduction of the Remadi, M. (2019). Field suppression of Fusarium wilt and microbial
population shifts in tomato rhizosphere following soil treatment with
disease in the crop suggest an effective action of ZnO NPs, on the
two selected endophytic bacteria. Eurasian Journal of Soil Science, 8,
growth of the fungus F. oxysporum; in addition, enhancing the pro- 208–­220.
moter effect by the action of zinc as a precursor of various phys- Apodaca-­Sánchez, M. A., Zavaleta-­Mejía, E., Osada-­Kawasoe, S., & García-­
iological mechanisms in plants highlights the feasibility of these Espinoza, R. (2004). Pudrición de la corona de chile (Capsicum annum
L.) en Sinaloa, México. Revista Mexicana de Fitopatología, 22, 22–­29.
nanoparticles as an alternative management of F. oxysporum in
Arciniegas-­Grijalba, P. A., Patiño-­Portela, M. C., Mosquera-­Sánchez, L. P.,
tomato. Guerrero-­Vargas, J. A., & Rodríguez-­Páez, J. E. (2017). ZnO nanopar-
ticles (ZnO NPs) and their antifungal activity against coffee fungus
Erythricium salmonicolor. Applied Nanoscience, 7, 225–­241. https://
doi.org/10.1007/s1320​4-­017-­0561-­3
5 | CO N C LU S I O N
Aydi Ben Abdallah, R., Mokni-­Tlili, S., Nefzi, A., Jabnoun-­Khiareddine, H.,
& Daami-­Remadi, M. (2016). Biocontrol of Fusarium wilt and growth
ZnO NPs showed antifungal activity inhibiting in vitro mycelial promotion of tomato plants using endophytic bacteria isolated from
growth and sporulation of Fusarium oxysporum. Foliar application Nicotiana glauca organs. Biological Control, 97, 80–­88. https://doi.
of ZnO NPs decreased the incidence and severity of the disease org/10.1016/j.bioco​ntrol.2016.03.005
Barnett, H. L., & Hunter, B. B. (1998). Illustrated genera of imperfect fungi.
caused by F. oxysporum, allowing the growth of tomato plants. ZnO
APS Press.
NPs have the potential as a biostimulant to promote plant growth Betancourt, G. R., Berlanga, D. M. L., Puente, U. B., Rodríguez, F. O., &
and to be used in the prevention and control of plant deterioration Sanchez-­Valdes, S. (2010). Surface modification of ZnO nanoparti-
by phytopathogenic microorganisms. cles. Materials Science Forum, 644, 61–­6 4.
Cardona-­Piedrahíta, L. F., & Castaño-­Zapata, J. (2019). Comparación
de métodos de inoculación de Fusarium oxysporum f. sp. lycopersici
C O N FL I C T O F I N T E R E S T (Sacc.) Snyder & Hansen, causante del marchitamiento vascular del
The authors declare no conflict of interest. tomate. Revista De La Academia Colombiana De Ciencias Exactas,
Físicas Y Naturales, 43, 227–­233.
Clavijo, C. S. D. (2014) Búsqueda de resistencia a la pudrición causada por
AU T H O R C O N T R I B U T I O N S
Fusarium spp. en Capsicum. Universidad Nacional de Colombia. Thesis.
Ana María González-­Merino designed the research and methodol- Coromoto, A. Y., & Reyes, I. (2018). Microorganismos promotores de cre-
ogy, and conducted samplings and experiments. Agustín Hernández-­ cimiento en el biocontrol de Alternaria alternata en tomate (Solanum
Juárez designed the research and methodology, conducted lycopersicum L.). Bioagro, 30, 59–­66.
GONZÁLEZ-­MERINO et al. | 11

Dimkpa, C. O., McLean, J. E., Britt, D. W., & Anderson, A. J. (2013). Méndez-­Argüello, B., Vera-­Reyes, I., Mendoza-­Mendoza, E., García-­
Antifungal activity of ZnO nanoparticles and their interactive ef- Cerda, L. A., Puente Urbina, B. A., & Lira-­Saldívar, R. H. (2016).
fect with a biocontrol bacterium on growth antagonism of the plant Promoción del crecimiento en plantas de Capsicum annuum por
pathogen Fusarium graminearum. BioMetals, 26, 913–­924. https://doi. nanopartículas de óxido de zinc. Nova Scientia, 8, 140–­156. https://
org/10.1007/s1053​4-­013-­9667-­6 doi.org/10.21640/​ns.v8i17.544
Elmer, W., De La Torre-­Roche, R., Pagano, L., Majumdar, S., Zuverza-­ Mishra, P. K., Fox, R. T., & Culham, A. (2003). Development of a PCR-­
Mena, N., Dimkpa, C., Gardea-­Torresdey, J., & White, J. C. (2018). based assay for rapid and reliable identification of pathogenic
Effect of metalloid and metal oxide nanoparticles on Fusarium Fusaria. FEMS Microbiology Letters, 218, 329–­332. https://doi.
wilt of watermelon. Plant Disease, 102, 1394–­1401. https://doi. org/10.1111/j.1574-­6968.2003.tb115​37.x
org/10.1094/PDIS-­10-­17-­1621-­RE Navarro, E., Piccapietra, F., Wagner, B., Marconi, F., Kaegi, R., Odzak,
Elmer, W., & White, J. C. (2018). The future of nanotechnology in plant N., Sigg, L., & Behra, R. (2008). Toxicity of silver Nanoparticles to
pathology. Annual Review of Phytopathology, 56, 111–­133. https://doi. Chlamydomonas reinhardtii. Environmental Science and Technology, 42,
org/10.1146/annur​ev-­phyto​- ­0 8041​7-­050108 8959–­8964.
Esparza, A. I. J. E. (2016) Potencial antimicrobial de nanopartículas metáli- Ocsoy, I., Paret, M. L., Arslan, O. M., Kunwar, S., Chen, T., You, M., &
cas en microorganismos fitopatógenos y su potencial como promotores Tan, W. (2013). Nanotechnology in plant disease management: DNA-­
de crecimiento en plantas. Universidad Autónoma Agraria Antonio directed silver nanoparticles on graphene oxide as an antibacterial
Narro. Thesis. against Xanthomonas perforans. ACS Nano, 7, 8972–­8980.
Esparza-­Rivera, E., Lira-­Saldivar, R. H., Hernández-­Suárez, M., Orberá, R. T. M., Serrat, D. M. J., & González, G. Z. (2009). Potencialidades
Betancourt-­Galindo, R., García-­Cerda, L. A., & Puente-­Urbina, B. de bacterias aerobias formadoras de endosporas para el biocontrol
(2014). Actividad antimicrobial de nanopartículas de cobre y óxido de en plantas ornamentales. Fitosanidad, 13, 95–­100.
zinc contra bacterias y hongos fitopatógenos. Conference 36 Congreso Pariona, N., Paraguay-­Delgado, F., Basurto-­Cereceda, S., Morales-­
Internacional de Metalurgia y Materiales. Mendoza, J. E., Hermida-­Montero, L. A., & Martinez-­Enriquez, A. I.
FIRA (Fideicomisos Instituidos en Relación con la Agricultura). (2019) (2020). Shape-­dependent antifungal activity of ZnO particles against
Panorama Agroalimentario. Tomate Rojo. Internet Resource: https:// phytopathogenic fungi. Applied Nanoscience, 10, 435–­4 43.
www.infor​u ral.com.mx/wp-­conte​nt/uploa​d s/2019/06/Panor​a ma-­ Prasad, T. N. V. K. V., Sudhakar, P., Sreenivasulu, Y., Latha, P., Munaswamy,
Agroa​limen​t ario​-­Tomat​e-­rojo-­2019.pdf (verified Dec 08, 2020) V., Reddy, K. R., Sreeprasad, T. S., Sajanlal, P. R., & Pradeep, T. (2012).
He, L., Liu, Y., Mustapha, A., & Lin, M. (2011). Antifungal activity of zinc Effect of nanoscale zinc oxide particles on the germination, growth
oxide nanoparticles against Botrytis cinerea and Penicillium expansum. and yield of peanut. Journal of Plant Nutrition, 35, 905–­927. https://
Microbiological Research, 166, 207–­215. https://doi.org/10.1016/j. doi.org/10.1080/01904​167.2012.663443
micres.2010.03.003 Rajiv, P., Rajeshwari, S., & Venckatesh, R. (2013). Bio-­Fabrication of zinc
Jamdagni, P., Khatri, P., & Rana, J. S. (2018). Green synthesis of zinc oxide oxide nanoparticles using leaf extract of Parthenium hysterophorus
nanoparticles using flower extract of Nyctanthes arbor-­tristis and L. and its size-­dependent antifungal activity against plant fungal
their antifungal activity. Journal of King Saud University-­Science, 30, pathogens. Spectrochimica Acta Part A: Molecular and Biomolecular
168–­175. Spectroscopy, 112, 384–­387.
Jamdagni, P., Rana, J. S., & Khatri, P. (2019). Antioxidant activity and Raskar, S. V., & Laware, S. L. (2014). Effect of zinc oxide nanoparticles on
antifungal fractional inhibitory concentration indices of zinc oxide cytology and seed germination in onion. The International Journal of
nanoparticles in combination with carbendazim, mancozeb, and thi- Current Microbiology and Applied Science, 3, 467–­473.
ram. Micro & Nano Letters, 14, 1037–­1040. https://doi.org/10.1049/ Rivero, J. C. A., Díaz, G. J. A., & López, N. J. I. (2014). Agricultura
mnl.2019.0104 orgánica vs agricultura moderna como factores en la salud pública.
Jo, Y. K., Kim, B. H., & Jung, G. (2009). Antifungal activity of silver ions ¿Sustentabilidad? Horizonte Sanitario, 4, 28–­4 0. https://doi.
and nanoparticles on phytopathogenic fungi. Plant Disease, 93, 1037–­ org/10.19136/​hs.a4n1.304
1043. https://doi.org/10.1094/PDIS-­93-­10-­1037 S.A.S. Institute. 2002. (2002). The SAS system for windows, release 9.0.
Kriti, A., Ghatak, A., & Mandal, N. (2020). Antimycotic efficacy of zinc SAS, Institute.
nanoparticle on dark-­spore forming phytopathogenic fungi. Journal Saharan, V., Sharma, G., Yadav, M., Choudhary, M. K., Sharma, S. S., Pal,
of Pharmacognosy and Phytochemistry, 9, 750–­754. A., Raliya, R., & Biswas, P. (2015). Synthesis and in vitro antifungal
Lamsal, K., Kim, S. W., Jung, J. H., Kim, Y. S., Kim, K. S., & Lee, Y. S. (2011). efficacy of Cu-­chitosan nanoparticles against pathogenic fungi of to-
Application of silver nanoparticles for the control of Colletotrichum mato. International Journal of Biological Macromolecules, 75, 346–­353.
species in vitro and pepper anthracnose disease in field. Mycobiology, Sawai, J., & Yoshikawa, T. (2004). Quantitative evaluation of antifungal
39, 194–­199. activity of metallic oxide powders (MgO, CaO and ZnO) by an indirect
Lee, S. B., & Taylor, J. W. (1990). Isolation of DNA from fungal mycelia and conductimetric assay. Journal of Applied Microbiology, 96, 803–­8 09.
single spores. In M. A. Innis, D. H. Gelfand, J. J. Sninsky, & T. J. White Servin, A., Elmer, W., Mukherjee, A., De la Torre-­Roche, R., Hamdi, H.,
(Eds.), PCR protocols a guide to methods and applications (pp. 282–­314). White, J. C., Bindraban, P., & Dimkpa, C. (2015). A review of the use
Academic Press. of engineered nanomaterials to suppress plant disease and enhance
Leslie, J. F., & Summerell, B. A. (2006). The Fusarium laboratory manual. crop yield. Journal of Nanoparticle Research, 17, 17–­92. https://doi.
Blackwell Publishing. org/10.1007/s1105​1-­015-­2907-­7
Mejía, A. J., & Hernández, M. (2001). Evaluación de azoxystrobin en el SIAP (Servicio de Información Agroalimentaria y Pesquera). (2020)
control de la candelilla temprana (Alternaria solani) en el cultivo de Anuario estadístico de la producción agrícola. Cierre de la producción
tomate. Revista De La Facultad De Agronomía, Universidad Del Zulia, agrícola. Internet Resource: http://www.siap.go b.mx. (verified Dec
18, 106–­116. 01, 2020).
Méndez-­Argüello, B., Lira-­Saldívar, R. H., Ruíz-­Torres, N. A., Cárdenas-­ Singh, B. K., Rakesh, E. S., Yadav, V. P. S., & Singh, D. K. (2010). Adoption
Flores, A., Ponce-­Zambrano, R., Vera-­Reyes, I., Mendoza-­Mendoza, of commercial cut flower production technology in Meerut. Indian
E., García-­Cerda, L. A., & De los Santos, G., (2015). Influencia de Research Journal of Extension Education, 10, 50–­53.
nanopartículas de óxido de zinc puras y dopadas con plata en el crec- Stampoulis, D., Sinha, S. K., & White, J. C. (2009). Assay-­dependent
imiento y producción de biomasa en plántulas de chile. XVI Congreso phytotoxicity of nanoparticles to plants. Environmental Science and
Nacional de Biotecnología y Bioingeniería. Technology, 43, 9473–­9479. https://doi.org/10.1021/es901​695c
12 | GONZÁLEZ-­MERINO et al.

Strayer-­S cherer, A., Liao, Y. Y., Young, M., Ritchie, L., Vallad, G. E., on plant pathogenic isolates of Pythium. Ecotoxicology, 24, 1305–­
Santra, S., Freeman, J. H., Clark, D., Jones, J. B., & Paret, M. L. 1314. https://doi.org/10.1007/s1064​6-­015-­1505-­x
(2018). Advanced copper composites against copper-­tolerant
Xanthomonas perforans and tomato bacterial spot. Phytopathology,
108, 196–­205.
How to cite this article: González-­Merino, A. M., Hernández-­
Wani, A. H., & Shah, M. A. (2012). A unique and profound effect of MgO
Juárez, A., Betancourt-­Galindo, R., Ochoa-­Fuentes, Y. M.,
and ZnO nanoparticles on some plant pathogenic fungi. Journal of
Applied Pharmaceutical Science, 2, 40–­4 4. Valdez-­A guilar, L. A., & Limón-­Corona, M. L. (2021).
Yehia, R. S., & Ahmed, O. F. (2013). In vitro study of the antifungal ef- Antifungal activity of zinc oxide nanoparticles in Fusarium
ficacy of zinc oxide nanoparticles against Fusarium oxysporum and oxysporum-­Solanum lycopersicum pathosystem under
Penicilium expansum. African Journal of Microbiology Research, 7,
controlled conditions. Journal of Phytopathology, 00, 1–­12.
1917–­1923.
Zabrieski, Z., Morrell, E., Hortin, J., Dimkpa, C., McLean, J., Britt, D., & https://doi.org/10.1111/jph.13023
Anderson, A. (2015). Pesticidal activity of metal oxide nanoparticles

View publication stats

You might also like