Next Article in Journal
Study on Livelihood Resilience of Rural Residents under the Rural Revitalization Strategy in Ethnic Areas of Western Sichuan, China
Previous Article in Journal
The Responses of C, N, P and Stoichiometric Ratios to Biochar and Vermicompost Additions Differ from Alfalfa and a Mine Soil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effects of the Fungicide Myclobutanil on Soil Enzyme Activity

by
Diana Larisa Roman
1,2,†,
Mariana Adina Matica
3,†,
Alecu Ciorsac
4,
Bianca Vanesa Boros
1,2 and
Adriana Isvoran
1,2,*
1
Department of Biology-Chemistry, Faculty of Chemistry, Biology, Geography, West University of Timisoara, 16 Pestalozzi, 300115 Timisoara, Romania
2
Advanced Environmental Research Laboratories (AERL), 4 Oituz, 300086 Timisoara, Romania
3
Department of Chemistry-Biology, Institute for Advanced Environmental Research (ICAM), West University of Timisoara, Oituz 4C, 300086 Timisoara, Romania
4
Department of Physical Education and Sport, University Politehnica Timișoara, 300006 Timișoara, Romania
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Agriculture 2023, 13(10), 1956; https://doi.org/10.3390/agriculture13101956
Submission received: 10 September 2023 / Revised: 3 October 2023 / Accepted: 5 October 2023 / Published: 7 October 2023
(This article belongs to the Section Agricultural Soils)

Abstract

:
(1) Background: The use of pesticides, although needed to protect crops and increase production, represents an environmental and human health issue. Therefore, measures must be taken in order to develop a better understanding of the risks involved in the overuse of these compounds. Myclobutanil is a chiral triazole fungicide widely used for the protection of crops against fungal diseases. Published data have shown that, although effective in preventing fungal infections, high doses of myclobutanil can affect the soil environment. The aim of this study was to evaluate the effect of different doses of myclobutanil on soil enzyme activity, as well as the possible specificity of the interactions of the two stereoisomers of myclobutanil with these enzymes. (2) Methods: A combination of experimental and computational approaches was considered. An experimental method was applied in order to assess the effect of different doses of myclobutanil on the activity of dehydrogenase, phosphatase, catalase, urease and protease. The computational approach was based on the molecular docking of the two enantiomers of myclobutanil with the above-mentioned enzymes to assess the possible enantioselectivity of the interactions. (3) Results: High doses of myclobutanil significantly affected the enzymatic activity of dehydrogenase and led to a slight increase in the activity of catalase. Molecular docking data showed that both enantiomers of myclobutanil were able to bind to the active sites of dehydrogenase, phsosphatase and protease, with higher interacting energies observed for (S)-myclobutanil, the enantiomer known to be less active against target organisms but have a higher toxicity against non-target organisms. (4) Conclusions: The results of our study confirm the need to implement better management practices regarding the use of myclobutanil (and of pesticides in general) by using the enantiomer that is most effective on target organisms and less toxic to non-target organisms.

Graphical Abstract

1. Introduction

Pesticides are chemicals that include a wide range of compounds, the main categories being represented by herbicides, fungicides and insecticides. These chemicals are widely developed and used to protect crops, but are also found as contaminants in air, water and soil [1]. Pesticides affect both their target organisms and numerous non-target species and are recognized for their harmful impact on the environment and human health [2]. Thus, new information on their effects is needed to ensure the effectiveness of protective measures.
Among pesticides, the class of triazole fungicides is essential to agriculture and food safety, being used to protect crops against yeast and mildew, their target organisms [3]. The global application amount of triazole fungicides was around 20,000 metric tons in 2020 [4]. These fungicides are proven to offer long-lasting control of numerous target plant pathogens and are considered to have moderate risk for developing fungicide resistance [5]. The biological action of these chemicals is based on their interaction with the target enzyme 14α-demethylase (a member of the cytochrome 450 family) that results in the inhibition of the synthesis of the fungal-specific sterol, ergosterol [6]. Ergosterol is the main sterol component of fungal membranes. It is involved in maintaining the structural integrity, thickness, permeability, and fluidity of the membrane and modulating the activity of several membrane-bound enzymes. The inhibition of ergosterol synthesis has a fungistatic effect because an increase in membrane permeability occurs and adversely affects the activity of several membrane-associated enzymes [7].
The presence of triazoles in the environment may influence not only the target species (fungi) but all kingdoms as well due to their nonspecific interactions with enzymes belonging to the cytochrome P450 family, resulting in the non-target effects of these fungicides [8].
Published data reveal that triazole fungicides usually affect soil enzyme activity, as discussed in a recent review [4]. DA was considerably reduced by high doses of difenoconazole, hexaconazole, mycobutanil, paclo-butrazol, tebuconazole, and triadimefon. Urease (UA) and phosphatase (PhA) activities caused strong decreases in the presence of difenoconazole, propiconazole and tebuconazole. Tebuconazole use led to decreased catalase (CA), aryl sulfatase, nitrate reductase, β-glucosidase, and invertase activities. Another study reveals that triticonazole decreased DA, did not affect PhA, but increased UA [9].
Literature data show that both the antifungal activity and non-target toxicity of several triazole fungicides are enantioselective [10,11,12,13,14,15,16,17,18]. There was no enantioselectivity of triticonazole stereoisomers in their interaction with soil enzymes, but such selectivity was recorded for the two steroisomers in the interactions with plasma proteins and human cytochromes [9]. Enantioselectivity in interactions with plasma proteins and human cytochromes was also recorded for difenoconazole enantiomers [10]. Triazole fungicides have also disrupted the structure of soil microbial communities (usually resulting in a decrease in the soil microbial population) [4] and have shown toxicological effects on organisms living in aquatic environments [19].
In the class of triazole fungicides, myclobutanil, with the IUPAC name (RS)-2-(4-chlorophenyl)-2-(1H-1,2,4-triazol-1-ylmethyl) hexanenitrile, is a broad-spectrum systemic fungicide. It is approved for use in many countries worldwide, including Iceland, Norway, Australia, the USA, and countries from Asia and Africa [20]. The fungicide may be applied at multiple plant growth stages (from seeds to grown season) but is also used as food additive and wood preservative. The target organisms of myclobutanil include ascomycetes, basidiomycetes and imperfect fungi. Myclobutanil controls the populations of these microorganisms on various crops including almonds, apricots, cereals, cotton, cucumbers, grapes, peaches, plums, strawberries, etc. [21].
Residues of myclobutanil were found in crops, in the soils in which they were grown [22], and in water streams [23,24]. The persistence of myclobutanil in various soil types was determined under both laboratory and field conditions, and the reported dissipation half-life (DT50) values ranged from 11.0–19.2 days [25,26,27] up to 574 days in anaerobic soil [28]. Others studies have discovered that myclobutanil residues are present in air [29] and food [30,31], the accumulation of myclobutanil in the food chain also being due to its ability to persist in the environment. Due to the presence and persistence of the fungicide in the environment, moderate adverse effects have been recorded from myclobutanil on non-target species: aquatic organisms, birds, bees, earthworms and rodents [32].
To the best of our knowledge, there are only several studies published in the literature on the effects of myclobutanil on soil enzyme activity. A small dose of myclobutanil applied to tea orchard soil produced an increase in DA, but higher doses led to decreased DA [33]. When using sandy tea garden soil, DA was slightly inhibited in the presence of myclobutanil during the first 30 days, but this effect disappeared over time [34].
Myclobutanil, as a commercial product, is an isomeric equimolar mixture of the (R)- and (S)-stereoisomers [35]. Literature data show that there are significant differences between the effects on target organisms produced by the racemic mixture and the effects produced by the enantiomers of myclobutanil. (R)-myclobutanil showed a higher antifungal activity than (S)-myclobutanil and also than the racemic mixture [36]. The enantioselective effects of myclobutanil isomers were also observed for non-target organisms. In the case of aquatic organisms, racemic myclobutanil showed greater toxicity than its enantiomers, which suggests a toxic synergistic effect between enantiomers. Also, the (S)-myclobutanil isomer revealed slightly greater toxicity than (R)-myclobutanil on these organisms [37]. Furthermore, the degradation of the two enantiomers of myclobutanil was distinct in aerobic soils, (R)-myclobutanil being preferentially degraded compared to (S)-myclobutanil [37]. Another study pointed to a preferential degradation of (R)-myclobutanil in cucumber crops, which led to an enrichment of (S)-myclobutanil in plants and soil [38]. Despite the proven differences between the biological activities of the myclobutanil enantiomers, this fungicide is still sold as a racemic mixture.
This study aims to evaluate, under laboratory conditions for 28 days, the effects of different doses of myclobutanil on the soil enzyme activity of the following: catalase, dehydrogenase, phosphatase, protease and urease. Numerous other enzymes, such as amidase, β-glucosidase, cellulose and chitinase, are detected in soil; their activities are correlated with soil organic matter and total nitrogen and may be also affected by pesticides [39]. The reason for considering dehydrogenase, phosphatase, protease and urease in the present study is that they are the most commonly used enzymes in the assessment of the biological status of soils and are also considered as potential indicators of soil quality [40]. Catalase is considered because its activity in soil has been shown to correlate with DA [41]. In addition, catalase, dehydrogenase, phosphatase, protease and urease activities are affected by other triazole fungicides [9,42]. Last but not least, since published data reveal the distinct biological activities of the two enantiomers of myclobutanil, molecular docking is considered in order to assess the specificity of the interactions of the enantiomers with these enzymes.
Considering the DT50 values for myclobutanil in soil, 11.0–19.2 days [11,12,13], the present study was implemented over a period of 28 days to cover the dissipation half-life. Moreover, numerous studies in the specialized literature mention this period as appropriate for experiments seeking to evaluate the effects of pesticides on the activity of soil enzymes under laboratory conditions.
To the best of our knowledge, this is the first study assessing the effects of myclobutanil on the activities of soil catalases, phosphatases, proteases, and ureases. Furthermore, the molecular docking approach is considered in order to better describe the possible enantioselective effects of the two stereoisomers of myclobutanil on the soil enzyme activities.

2. Materials and Methods

2.1. Fungicide

The experiments were performed using the commercial product namely “SysthaneTM Forte” (DAW AgroSciences, Indianapolis, IN, USA) containing 240 g/L myclobutanil. This fungicide has one chiral center and, consequently, two stereoisomers: (R)-myclobutanil and (S)-myclobutanil (Figure 1).
For the molecular docking study, the 3D structures of the two enantiomers of myclobutanil were extracted from PubChem database [43].

2.2. Soil Sampling

The soil used in experiments was of chernozem type and collected from a field located close to Timisoara city, Romania, (45°45′14.54″ N, 21°18′16.66″ E) where pesticides were never used. Soil samples were collected from the top layer of soil (about 20 cm) from 5 distinct spots in quantities of 20 kg each. The soil was ground, sieved (2 mm) and scooped via random sampling to obtain sub-samples of 10 kg of soil, which were immediately processed [44].
SysthaneTM Forte was sprayed onto soil samples at different doses (Table 1) and other soil samples without myclobutanil were used as controls.
Samples were maintained in laboratory conditions at 22 °C ± 2 °C and the soil moisture was 19% ± 3%. Soil moisture was adjusted every 3–4 days.

2.3. Enzymatic Activity Analyses

Dehydrogenase (EC 1.1.), urease (EC 3.5.1.5), phosphatase (EC 3.1.3.2), catalase (EC 1.11.1.6) and protease (EC 3.4) activities were evaluated in this study. For this purpose, the spectrophotometric method and a UV/Vis T90 spectrophotometer (PG Instruments, England) were used. The determination of DA followed the method described by Schinner and his coworkers [45]. For the determination of UA, the method proposed by Alef and Nannipieri [46] was taken into account. PhA was obtained according to the method described by Dick [47]. CA was determined using the permanganometric method described by Drăgan-Bularda [48]. PA was estimated on the basis of the reaction of ninhydrin with the amino acids that resulted from the hydrolysis of the gelatin used as a substrate, which was the method proposed by Drăgan-Bularda [48]. These methods have previously been applied in order to evaluate the effects on the activity of these enzymes of the herbicides aclonifen [49], S-metolachlor [50] and oxyfluorfen [51], and on the triazole fungicides triticonazole [9] and difenoconazole [52]. The soil treated with myclobutanil at different doses and the control soil were incubated for a period of 28 days under laboratory conditions, and the activity of each enzyme was determined every 7 days. Each experiment was implemented in triplicate and the same researcher performed the measurements on the same day.

2.4. Molecular Docking Study

The two stereoisomers of myclobutanil were docked with three-dimensional structures of enzymes found in soil: catalase, dehydrogenase, phosphatase, protease and urease. The three-dimensional structures of these enzymes, which belong to microorganisms found in soil (fungi, bacteria and Rizhobium sp.) in complex with their specific substrates or inhibitors (when available), were extracted from the Protein Data Bank (PDB) [53] and are presented in Table 2.
In the case of catalase, AlphaFold [61] structural models, namely, AF-P42234-F1 for Bacillus subtilis and AF-Q1M498-F1 for Rhizobium leguminosarum, were taken into account because these proteins had no three-dimensional structures. To locate the active site regions of the two catalases for which only structural models existed, the AlphaFold models were superposed onto the structure of the catalase belonging to the fungus Komagataella pastoris (PDB ID 6RJN). Very good matches of the superimposed structures were obtained: (i) RMSD = 0.851 Å for 410 alpha carbon pairs out of the 480 total pairs when superposing the structures of catalases belonging to Komagataella pastoris and Bacillus subtillis; and (ii) RMSD = 0.764 Å for 395 alpha carbon pairs out of 480 total pairs when superposing the structures of catalases belonging to Komagataella pastoris and Rhizobium meliloti.
A similar situation was registered for ureases. In the case of urease, AlphaFold models were identified for urease belonging to the fungus Aspergillus fumigatus (AF-P28296-F1) and to bacteria Rhizobium leguminosarum (AF-A0A0Q7ALB8-F1). Very good matches of the superimposed structures were also obtained in these cases: (i) RMSD = 0.585 Å for 544 alpha carbon pairs out of the 565 total pairs when superposing the structures of ureases belonging to Bacillus pasteurii and Aspergillus fumigatus; and (ii) RMSD = 0.550 Å for 550 alpha carbon pairs out of 567 total pairs when superposing the structures of ureases belonging to Bacillus pasterurii and Rhizobium leguminosarum.
The AlphaFold model was available for the protease belonging to Rhizobium legumi-nosarum, but the structure of this enzyme did not have very good matches with the structures of proteases belonging to Streptomyces griseus and Aspergillus clavatus, respectively.
The three-dimensional structures of the myclobutanil enantiomers and those of the investigated enzymes were prepared for molecular docking using of the DockPrep facility under Chimera 1.16 software [62]. SwissDock server [63] was used to implement the molecular docking study. Blind and precise docking were selected. Chimera software was also used to analyze the docking results.
In order to identify the residues that are involved in the non-covalent interactions of myclobutanil enantiomers with the investigated enzymes, the Protein-Ligand Interaction Profiler (PLIP) software was used [64]. Starting from the three-dimensional structure of the investigated complex, PLIP allows the detection and visualization of protein–ligand interactions at the atomic level: hydrogen bonds, hydrophobic contacts, pi-stacking and pi–cation interactions, salt bridges, water bridges and halogen bonds [64].

2.5. Statistical Analysis

The statistical analysis of data was performed using PAST 4.03 software [65]. Data normality was assessed through Shapiro–Wilk W test and ANOVA analysis following a distribution test. Parametric tests were used for normally distributed data, these being Levene’s test for homogeneity of variance among treatments and Tukey’s post hoc test for analysis of variance. Non-parametric tests were used for non-normally distributed data. These tests were Kruskal–Wallis test and Dunn’s post hoc test for further analysis of variance.

3. Results

3.1. Effects of Myclobutanil on the Activity of Enzymes Found in Soil

Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6 present the activities of soil dehydrogenases, urease, phosphatase, catalase and protease over a period of 28 days for the soil samples treated with different doses of myclobutanil, as observed every 7 days after treatment. The error bars in these figures reveal some uncertainty in the measured activities, which can be explained by the variations in soil moisture during the experiment. Statistically significant differences were observed over time during the test period and, in some cases, between effects caused by different doses of myclobutanil (Table 3, Table 4 and Table 5).
The statistical analysis only reveals statistically significant differences for DA over time in the case of the double dose, with the DA in the soil in the 28th day being significantly different from that found on the other test days. Also, throughout the test period, the double dose showed statistically significant differences from the control and the other two doses (Table 3, above the diagonal).
For soil phosphatases, statistically significant differences were observed over time for both the control and the three doses tested, but no statistically significant differences were observed between the control and the three doses over the entire test period. For both the control and the three doses, significant differences were observed between the day-21 and day-28 activities (Table 3, below the diagonal). A sharp increase in PhA was found to occur on 28 compared to day 21. A possible explanation for this is that the soil moisture on day 21 was lower than that on day 28. After the increase in moisture, the activity of enzyme increased as it is known that phosphatase needs water to cleave the phosphoric acid monoester into phosphate ion and alcohol.
Statistically significant differences were observed over time in the case of UA, both for the control and for the entire dose, with significant differences being highlighted between the activity on day 7 and 28. Between the control and applied doses, no statistically significant differences were observed throughout the entire test period (Table 4, above the diagonal).
In terms of CA, some statistically significant differences were observed over time for both control and tested doses, with activity on day 7 being significantly different from activity on day 14. A comparison of the effects produced by the different doses of myclobutanil revealed statistically significant differences between CA for the double dose, CA for the control and, respectively, the other doses on test days 21 and 28 (Table 4, below the diagonal).
Regarding PA, statistically significant differences were observed over time. At a half dose and double dose, significant differences were found between PA on day 7 and day 21. For the full dose, PA on day 7 showed significant differences from the other three test days. No statistically significant differences were observed between PA for the control and the three doses, respectively, over the entire test period (Table 5, above the diagonal).

3.2. Molecular Docking Study Regarding the Interactions of the Enantiomers of Myclobutanil with Soil Enzymes

The molecular docking results regarding the interactions of the myclobutanil enantiomers with soil enzymes are presented in Figure 7, Figure 8, Figure 9 and Figure 10, in Supplementary Figures S1–S10 and in Table 6, respectively. Figure 7, Figure 8 and Figure 9 reveal the binding modes corresponding to the highest interaction energies for the binding of myclobutanil enantiomers to the active sites of Clostridium beijerinckii dehydrogenase (Figure 7), Bacillus subtilis phosphatase (Figure 8) and Streptomyces griseus protease (Figure 9), respectively. For the dehydrogenases belonging to Aspergillus fumigatus and to Rhizobium leguminosarum, the binding modes of the myclobutanil enantiomers to the enzyme active sites are revealed in Supplementary Figures S1 and S2, respectively. In the cases of phosphatases belonging to Aspergillus niger and Rhizobium leguminosarum, the binding modes corresponding to the active sites of enzymes are shown in Supplementary Figures S3 and S4, respectively. Supplementary Figures S5 and S6, respectively, show the binding modes corresponding to the active sites of Aspergillus clavatus and Rhizobium leguminosarum proteases.
Table 6 contains the values of the interaction energies for the binding modes corresponding to the binding of myclobutanil enantiomers to the active sites of the enzymes resulting from molecular docking.
Figure 7 illustrates that the orientations of (R)- and (S)-myclobutanil are distinct when binding to the active site of Clostridium beijerinckii dehydrogenase. The orientations of the myclobutanil enantiomers are similar when bound to the catalytic site of Aspergillus fumigatus dehydrogenase (Supplementary Figure S1), but there are distinct orientations for the two enantiomers bound to the active site of Rhizobium leguminosarum dehydrogenase (Supplementary Figure S2). There are also small differences in the interaction energies for each enantiomer with investigated dehydrogenases (Table 6). The interactions of myclobutanil enantiomers with Clostridium beijerinckii dehydrogenase are stronger than those with Aspergillus fumigatus dehydrogenase, which are also stronger than those with Rhizobium leguminosarum dehydrogenase (Supplementary Table S1). The binding of myclobutanil to the active site of all investigated dehydrogenases is in good correlation with the experimental results, revealing a significant decrease in DA in the presence of the high dose of fungicide.
In the case of the interaction of myclobutanil enantiomers with Bacillus subtilis phosphatase, there is quite a similar positioning for the two enantiomers (Figure 8), but there are quite distinct orientations for the myclobutanil enantiomers bound to Aspergillus niger (Supplementary Figure S3) and, respectively, to Rhizobium leguminosarum (Supplementary Figure S4). The interaction energies corresponding to these binding modes are quite distinct for the two enantiomers, the higher interacting energies being registered for the Bacillus subtilis phosphatase.
Distinct orientations and different interaction energies (Table 6) of myclobutanil enantiomers are observed for their interactions with proteases belonging to Streptomyces griseus (Figure 9), Aspergillus clavatus (Supplementary Figure S5) and Rhizobium leguminosarum (Supplementary Figure S6).
Even though the two enantiomers of myclobutanil are able to bind to the active sites of phosphatase and protease, the activities of these enzymes were not significantly affected by the presence of the fungicide. The interaction energies between myclobutanil and phosphatase and protease, respectively, were lower than those corresponding to interactions with dehydrogenase.
Enantiomers of myclobutanil with distinct orientations bound to the active site of Komagataella pastoris catalase (Supplementary Figure S7) but did not bind to the active sites of catalases from Bacillus subtilis (Figure 10a) and Rhizobium meliloti (Supplementary Figure S8). They also did not bind to any of the ureases belonging to Bacillus pasteurii (Figure 10b), Aspergillus fumigatus (Supplementary Figure S9) and Rhizobium leguminosarum (Supplementary Figure S10). This is in good correlation with experimental results indicating that UA was not significantly affected by myclobutanil and that CA registered a slight increase.
In addition to the distinct orientations, there are also small differences in the interaction energies for the binding modes of the two enantiomers of myclobutanil to soil enzymes (Table 6).
The interactions between the two myclobutanil enantiomers and the soil enzymes were analyzed using PLIP software and the results are shown in Table 7. This table also contains the enzymes residues involved in the interactions with the ligands that are present in the crystallographic structures, when available: NADPH—dihydro-nicotinamide-adenine-dinucleotide phosphate, signaling peptide GLN-ARG-GLY-MET-ILE, inhibitory tetrapetide ACE-PRO-ALA-PRO-PHE, and NAG—2-acetamido-2-deoxy-beta-D-glucopyranose.
The information presented in Table 7 reveals that the two enantiomers interact with amino acids that are involved in the binding of known substrates/inhibitors to the investigated enzymes and confirms their binding to the active sites of these enzymes. Furthermore, the two enantiomers usually do not interact with similar amino acids of the same enzyme. There are several distinct amino acids involved in these interactions that underlie the enantioselectivity of binding.

4. Discussion

4.1. Experimental Approach

The results presented in this study complete our knowledge of the spectra of the effects of triazole fungicides on soil enzyme activities with information on the effects of myclobutanil. Additionally, they emphasize that high doses of myclobutanil produce decreases in the DA for a period of 28 days. Regarding the PhA, myclobutanil does not significantly affect it regardless of the dose and this is also true for UA and PA. On the contrary, an increase in CA is registered for the double dose of myclobutanil. It should be noticed that, during the period of investigations, there are changes in all the enzymes activities, both for the control and for the soil samples treated with myclobutanil. This can be due to the variations in soil moisture correlated with changes in the physicochemical parameters and bacterial community of soil and underlying the high complexity of soil systems.
Some of these results are in quite good agreement with other published data concerning the triazole fungicides, as is presented in the following. Regarding the effects of triazole fungicides on DA, in a study performed by Zhang et al., the application of a small dose of myclobutanil to tea orchard soil induced, in the first 10 days, an increase in the DA, but higher doses induced decreased DA [33]. Another study revealed that the DA in sandy loam tea orchard soil was inhibited to some extent in the presence of myclobutanil for the 30 days [34]. DA also registered decreased values in the presence of other triazole fungicides: triticonazole [9], difenoconazole, hexaconazole, paclobutrazole, tebuconazole and triadimefon [42]. In the case of UA, a slight decrease was registered in the presence of the triazole fungicides difenoconazole, propiconazole and tebuconazole [42]. Among the triazole fungicides, difenoconazole produced a slight increase in the PA [42]. PhA seemed to be unaffected by the presence myclobutanil and a similar result wasobserved for triadimefon; however, several other triazole fungicides (difenoconazole, propiconazole and tebuconazole) produced a decrease in PhA [42].
Little is known about the effects of triazole fungicides on CA. The present study emphasized that the presence of myclobutanil slightly increased the CA, but other studies revealed that tebuconazole produced a decrease in CA [66,67]. These dissimilar finding may be due to the different experimental approaches. This study was performed in laboratory conditions on soil without crop, while that produced by Saha and coworkers [66] was implemented in the field in peanut soil [66,67]. Additionally, the study made performed Baćmaga and coworkers [67] considered the sandy clay soil cultured with spring barley [66,67].
Specific literature data reveal that triazole fungicides may also affect the activity of other soil enzymes such as chitinases [68], glucosidases, arylsulfatases and invertases [67]. All these outcomes confirm the high sensitivity of soil enzymes to the presence of triazole fungicides.
One of the limitations of this study is that the physicochemical properties of soil and the bacterial community of soil were not assessed. The activities of enzymes found in soil, the soil physicochemical properties and the soil microbial communities are closely related, with soil physicochemical properties determining the composition of communities of soil microorganisms [44]. These microorganisms directly influence soil enzyme activities, with some of enzymes operating intracellularly in living microbial cells and others being secreted extracellularly by the living microorganisms [69]. Furthermore, soil dehydrogenase is an enzyme occurring in the viable microbial cells and its decrease in the presence of myclobutanil suggests a diminution of soil microbial population. Consequently, the effects of myclobutanil on the investigated enzymes indirectly emphasizes that there are interactions between myclobutanil, the bacterial community, and the physical and chemical characteristics of the soil and that myclobutanil may lead to modified soil physicochemical properties and affect the populations of the microorganisms living in soil, the population of fungi as target organisms and the population of yeasts and bacteria as non-target organisms. The decrease in DA suggests the reduction in microbial biomass in the presence of myclobutanil, but the slight increase in CA suggests that some bacteria may use the fungicide as a source of nutrients. This underlines that myclobutanil may also alter the structure of microorganism communities in soil.

4.2. Computational Approach

The outcomes of the molecular docking complete the information derived from the experimental approach. The highest binding energies were recorded for the two enantiomers of myclobutanil to dehydrogenases, which is in good correlation with the significant registered decrease in DA recorded for the soil treated with the double dose of myclobutanil. Even though both enantiomers of myclobutanil were able to bind to the active sites of phosphatases and proteases, experimental data showed that the activities of these enzymes were not significantly affected by the presence of myclobutanil. Phosphatases and proteases were enzymes secreted into the soil by some microorganisms, while dehydrogenases are intracellular enzymes. Moreover, the energies were higher for the interactions of myclobutanil with dehydrogenases compared to the energies corresponding to the interactions with phosphatases and proteases.
Usually, the orientations of the enantiomers in the interactions with the active sites of the enzymes are different. There are also differences between the binding energies for the two enantiomers and this emphasizes their enantioselectivity. (S)-myclobutanil, the enantiomer that revealed lower antifungal activity and higher toxicity against non-target organisms, presents relatively higher interaction energies with these enzymes in many cases, suggesting its higher inhibitory potential.
The enantiomers of myclobutanil could only bind to the active site of catalase from Komagataella pastoris. However, this was not the case for the other investigated catalases. The interaction energies for these bonds were lower compared to those recorded for dehydrogenases. The two enantiomers did not bind to the active sites of the other investigated catalases. This was potentially due to the structures of these enzymes, which revealed long and narrow channels before reaching the active site to prevent the binding of numerous chemicals [70]. This was in good agreement with the experimental result that revealed myclobutanil did not lead to CA decrease and with published data revealing catalase activity to be considered relatively stable in soil [71].
Myclobutanil enantiomers did not bind to urease. This was is agreement with the experimental results, which showed that myclobutanil produced no changes in UA. A similar result was observed for the triticonazole enantiomers [9] and for aclonifen and its metabolites [49] that could not bind to the active site of this enzyme. This was explained by the obtuseness, rigidity and hydrophilicity of the catalytic site of urease [60].
Recent data in the literature reveal the enantioselective effects of pesticides and/or of their metabolites on both target and non-target organisms, with enantioselectivity differing between toxicological effects or species [72]. Regarding triazole fungicides, enantioselective environmental and human health effects were observed for difenoconazole [10], hexaconazole [11], mefentrifluconazole [12], prothioconazole [13], tebuconazole [14], and triticonazole [9].
The limitation of this computational study is that the dose of myclobutanil cannot be taken into account. The advantage of using a molecular docking study is reflected by the fact that it allows researchers to take into account the racemic structure of the commercial product.

5. Conclusions

The outcomes of the present study reveal that the triazole fungicide myclobutanil negatively affected the activity of soil dehydrogenase when used in high doses. Furthermore, both enantiomers of myclobutanil were able to bind to the active sites of dehydrogenase, phosphatase and protease, the interacting energies being to some extent higher for the (S)-myclobutanil, the enantiomer that is known to be less active against target organisms and to manifest higher toxicity against non-target organisms. This underlines the enantioselective effects of the triazole fungicide myclobutanil on the activities of enzymes found in soil. It also emphasizes that the racemic mixtures that are used in crop and food management practices should be replaced by the enantiomer that reveals the highest activity against target organisms and the lowest toxicity against non-target organisms.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agriculture13101956/s1, Figure S1: Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Aspergillus fumigatus dehydrogenase (PDB ID 7RK5) (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The substrate 1,4-dihydronicotinamide adenine dinucleotide (yellow sticks) illustrates the position of the active site; Figure S2: Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Rhizobium leguminosarum dehydrogenase (PDB ID 8C54) (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The substrate 1,4-dihydronicotinamide adenine dinucleotide (yellow sticks) illustrates the position of the active site; Figure S3: Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Aspergillus niger phosphatase (PDB ID 1QFX) (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The substrate 2-acetamido-2-deoxy-beta-D-glucopyranose (yellow sticks) illustrates the position of the active site; Figure S4: Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Rhizobium leguminosarum phosphatase (PDB ID 2VQR) (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The active site was found taking into consideration the position of the Ca2+ and Mn2+ ions (not shown); Figure S5: Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Aspergillus clavatus protease (PDB ID 7Z6T) (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The active site was found taking into consideration the position of the Ca2+ and Zn2+ ions (not shown); Figure S6: Best binding modes of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Rhizobium leguminosarum protease (AF-A0A0Q7ALB8-F1) (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic); Figure S7: Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Komagataella pastoris catalase (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The substrate dihy-dro-nicotinamide-adenine-dinucleotide phosphate (yellow sticks) illustrates the position of the active site; Figure S8: Binding of (R)-myclobutanil (red sticks) (a) and of (S)-myclobutanil (green sticks) (b) to site of Rhizobium meliloti catalase (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The position substrate dihy-dro-nicotinamide-adenine-dinucleotide phosphate (yellow sticks) illustrates the active site by superposing the structures of Rhizobium meliloti and Komagataella pastoris catalases; Figure S9: Binding of (R)-myclobutanil (red sticks) (a) and of (S)-myclobutanil (green sticks) (b) to Aspergillus fumigatus (AF-Q6A3P9-F1) urease (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The myclobutanil enantiomers do not bind to the active site, those position is illustrated by the citrate ion (yellow sticks); Figure S10: Binding of (R)-myclobutanil (red sticks) (a) and of (S)-myclobutanil (green sticks) (b) to Rhizobium leguminosarum (AF-Q1MCV9-F1, solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The myclobutanil enantiomers do not bind to the active site, those position is illustrated by the citrate ion (yellow sticks).

Author Contributions

Conceptualization, A.I.; methodology, A.I. and D.L.R.; formal analysis, D.L.R., M.A.M. and A.C.; investigation, D.L.R., M.A.M. and A.C.; data curation, D.L.R. and M.A.M.; statistical analysis—B.V.B.; writing—original draft preparation, A.I.; writing—review and editing, A.I., M.A.M. and B.V.B.; supervision, A.I.; project administration, D.L.R.; funding acquisition, D.L.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant from the Ministry of Research, Innovation, and Digitization, CNCS/CCCDI-UEFISCDI, project number PN-III-P1-1.1-PD-2019-0255, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aktar, M.W.; Sengupta, D.; Chowdhury, A. Impact of Pesticides Use in Agriculture: Their Benefits and Hazards. Interdiscip. Toxicol. 2009, 2, 1–12. [Google Scholar] [CrossRef] [PubMed]
  2. Zúñiga-Venegas, L.A.; Hyland, C.; Muñoz-Quezada, M.T.; Quirós-Alcalá, L.; Butinof, M.; Buralli, R.; Cardenas, A.; Fernandez, R.A.; Foerster, C.; Gouveia, N.; et al. Health Effects of Pesticide Exposure in Latin American and the Caribbean Populations: A Scoping Review. Environ. Health Perspect. 2022, 130, 096002. [Google Scholar] [CrossRef] [PubMed]
  3. Zubrod, J.P.; Bundschuh, M.; Arts, G.; Brühl, C.A.; Imfeld, G.; Knäbel, A.; Payraudeau, S.; Rasmussen, J.J.; Rohr, J.; Scharmüller, A.; et al. Fungicides: An Overlooked Pesticide Class? Environ. Sci. Technol. 2019, 53, 3347–3365. [Google Scholar] [CrossRef]
  4. Souders, C.L., 2nd; Perez-Rodriguez, V.; El Ahmadie, N.; Zhang, X.; Tischuk, C.; Martyniuk, C.J. Investigation Into the Sub-lethal Effects of the Triazole Fungicide Triticonazole in Zebrafish (Danio rerio) Embryos/Larvae. Environ. Toxicol. 2020, 35, 254–267. [Google Scholar] [CrossRef]
  5. Jørgensen, L.N.; Heick, T.M. Azole Use in Agriculture, Horticulture, and Wood Preservation—Is It Indispensable? Front. Cell. Infect. Microbiol. 2021, 11, 730297. [Google Scholar] [CrossRef]
  6. Börjesson, E.; Stenström, J.; Johnsson, L.; Torstensson, L. Comparison of Triticonazole Dissipation after Seed or Soil Treatment. J. Environ. Qual. 2003, 32, 1258–1261. [Google Scholar] [CrossRef]
  7. Jackson, C.J.; Lamb, D.C.; Manning, N.J.; Kelly, D.E.; Kelly, S.L. Mutations in Saccharomyces cerevisiae sterol C5-desaturase Conferring Resistance to the CYP51 Inhibitor Fluconazole. Biochem. Biophys. Res. Commun. 2003, 309, 999–1004. [Google Scholar] [CrossRef]
  8. Jaklová Dytrtová, J.; Bělonožníková, K.; Jakl, M.; Ryšlavá, H. Triazoles and Aromatase: The impact of Copper Cocktails. Environ. Pollut. 2020, 266, 115201. [Google Scholar] [CrossRef]
  9. Roman, D.L.; Voiculescu, D.I.; Matica, M.A.; Baerle, V.; Filimon, M.N.; Ostafe, V.; Isvoran, A. Assessment of the Effects of Triticonazole on Soil and Human Health. Molecules 2022, 27, 6554. [Google Scholar] [CrossRef]
  10. Voiculescu, D.I.; Roman, D.L.; Ostafe, V.; Isvoran, A. A Cheminformatics Study Regarding the Human Health Risks Assessment of the Stereoisomers of Difenoconazole. Molecules 2022, 27, 4682. [Google Scholar] [CrossRef]
  11. Liu, T.; Fang, K.; Liu, Y.; Zhang, X.; Han, L.; Wang, X. Enantioselective Residues and Toxicity Effects of the Chiral Triazole Fungicide Hexaconazole in Earthworms (Eisenia fetida). Environ. Pollut. 2021, 270, 116269. [Google Scholar] [CrossRef]
  12. Li, L.; Sun, X.; Zhao, X.; Xiong, Y.; Gao, B.; Zhang, J.; Shi, H.; Wang, M. Absolute Configuration, Enantioselective Bioactivity, and Degradation of the Novel Chiral Triazole Fungicide Mefentrifluconazole. J. Agric. Food Chem. 2021, 69, 4960–4967. [Google Scholar] [CrossRef] [PubMed]
  13. Xiang, Q.; Zhou, Y.; Tan, C. Enantioselective Toxic Effects of Prothioconazole Toward Scenedesmus obliquus. Molecules 2023, 28, 4774. [Google Scholar] [CrossRef] [PubMed]
  14. Cui, N.; Xu, H.; Yao, S.; He, Y.; Zhang, H.; Yu, Y. Chiral Triazole Fungicide Tebuconazole: Enantioselective Bioaccumulation, Bioactivity, Acute Toxicity, and Dissipation in Soils. Environ. Sci. Pollut. Res. 2018, 25, 25468–25475. [Google Scholar] [CrossRef] [PubMed]
  15. Smith, S.W. Chiral Toxicology: It’s the Same Thing…Only Different. Toxicol. Sci. 2009, 110, 4–30. [Google Scholar] [CrossRef]
  16. Garrison, A.W.; Avants, J.K.; Jones, W.J. Microbial Transformation of Triadimefon to Triadimenol in Soils: Selective Production Rates of Triadimenol Stereoisomers Affect Exposure and Risk. Environ. Sci. Technol. 2011, 45, 2186–2193. [Google Scholar] [CrossRef] [PubMed]
  17. Li, L.; Gao, B.; Wen, Y.; Zhang, Z.; Chen, R.; He, Z.; Kaziem, A.E.; Shi, H.; Wang, M. Stereoselective Bioactivity, Toxicity and Degradation of the Chiral Triazole Fungicide Bitertanol. Pest Manag. Sci. 2020, 76, 343–349. [Google Scholar] [CrossRef]
  18. Yang, X.; Gong, R.; Chu, Y.; Liu, S.; Xiang, D.; Li, C. Mechanistic Insights into Stereospecific Antifungal Activity of Chiral Fungicide Prothioconazole against Fusarium oxysporum F. sp. cubense. Int. J. Mol. Sci. 2022, 23, 2352. [Google Scholar] [CrossRef]
  19. Roman, D.L.; Voiculescu, D.I.; Ostafe, V.; Ciorsac, A.; Isvoran, A. A Review of the Toxicity of Triazole Fungicides Approved to be Used in European Union to the Soil and Aqueous Environment. Ovidius Univ. Ann. Chem. 2022, 33, 113–120. [Google Scholar] [CrossRef]
  20. Myclobutanil Approval. Available online: http://sitem.herts.ac.uk/aeru/ppdb/en/Reports/478.htm (accessed on 19 June 2023).
  21. Soler-Rodríguez, F.; Oropesa Jiménez, A.L. Myclobutanil. In Encyclopedia of Toxicology, 3rd ed.; Wexler, P., Ed.; Academic Press: Oxford, UK, 2014; pp. 420–423. [Google Scholar]
  22. Dong, Z.; Cui, K.; Liang, J.; Guan, S.; Fang, L.; Ding, R.; Wang, J.; Li, T.; Zhao, S.; Wang, Z. The Widespread Presence of Triazole Fungicides in Greenhouse Soils in Shandong Province, China: A systematic Study on Human Health and Ecological Risk Assessments. Environ. Pollut. 2023, 328, 121637. [Google Scholar] [CrossRef]
  23. Smiley, P.C.; King, K.W.; Fausey, N.R. Annual and Seasonal Differences in Pesticide Mixtures Within Channelized Agricultural Headwater Streams in Central Ohio. Agric. Ecosyst. Environ. 2014, 193, 83–95. [Google Scholar] [CrossRef]
  24. Zhao, P.; Lei, S.; Xing, M.; Xiong, S.; Guo, X. Simultaneous Enantioselective Determination of Six Pesticides in Aqueous Environmental Samples by Chiral Liquid Chromatography with Tandem Mass Spectrometry. J. Sep. Sci. 2018, 41, 1287–1297. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, X.; Dong, F.; Wang, X.; Zheng, Y. The Dissipation Rates of Myclobutanil and Residue Analysis in Wheat and Soil Using Gas Chromatography-ion Trap Mass Spectrometry. Int. J. Environ. Anal. Chem. 2009, 89, 957–967. [Google Scholar] [CrossRef]
  26. Han, B.; Chen, L.; Peng, L. Degradation of Four Fungicides in Tropical Soils from Hainan, China. J. Food Agric. Environ. 2009, 7, 197–200. [Google Scholar]
  27. Wang, X.; Li, Y.; Xu, G.; Sun, H.; Xu, J.; Zheng, X.; Wang, F. Dissipation and Residues of Myclobutanil in Tobacco and Soil Under Field Conditions. Bull. Environ. Contam. Toxicol. 2012, 88, 759–763. [Google Scholar] [CrossRef]
  28. FAO. Pesticide Residues in Food, Joint FAO/WHO Meeting on Pesticide Residues; FAO and WHO: Rome, Italy, 2014. Available online: https://www.fao.org/fileadmin/templates/agphome/documents/Pests_Pesticides/JMPR/Report2014/JMPR_2014_Full_Report.pdf (accessed on 10 July 2023).
  29. Di Filippo, P.; Pomata, D.; Riccardi, C.; Buiarelli, F.; De Gennaro, M.; Console, C.; Laurendi, V.; Puri, D. Determination of Pesticides in the Respirable Fraction of Airborne Particulate Matter by High-performance Liquid Chromatography–Tandem Mass Spectrometry. Anal. Lett. 2018, 51, 600–612. [Google Scholar] [CrossRef]
  30. Kovacova, J.; Kocourek, V.; Kohoutkova, J.; Lansky, M.; Hajslova, J. Production of Apple-based Baby Food: Changes in Pesticide Residues. Food Addit. Contam. Part A 2014, 31, 1089–1099. [Google Scholar] [CrossRef]
  31. Freeman, S.; Kaufman-Shriqui, V.; Berman, T.; Varsano, R.; Shahar, D.R.; Manor, O. Children’s Diets, Pesticide Uptake, and Implications for Risk Assessment: An Israeli Case Study. Food Chem. Toxicol. 2016, 87, 88–96. [Google Scholar] [CrossRef]
  32. Lewis, K.A.; Tzilivakis, J.; Warner, D.J.; Green, A. An International Database for Pesticide Risk Assessments and Management. Hum. Ecol. Risk Assess. Int. J. 2016, 22, 1050–1064. [Google Scholar] [CrossRef]
  33. Zhang, D.; Wu, Y.; Zhang, X.; Zhu, Y. Persistence of Myclobutanil and Its Impact on Soil Microbial Biomass C and Dehydrogenase Enzyme Activity in Tea Orchard Soils. Eurasian J. Soil Sci. 2017, 6, 106–113. [Google Scholar] [CrossRef]
  34. Marín-Benito, J.M.; Herrero-Hernández, E.; Andrades, M.S.; Sánchez-Martín, M.J.; Rodríguez-Cruz, M.S. Effect of Different Organic Amendments on the Dissipation of Linuron, Diazinon and Myclobutanil in an Agricultural Soil Incubated for Different Time Periods. Sci. Total Environ. 2014, 476–477, 611–621. [Google Scholar] [CrossRef]
  35. Myclobutanil Isomers. Available online: http://sitem.herts.ac.uk/aeru/iupac/Reports/478.htm (accessed on 2 June 2023).
  36. Sun, M.; Liu, D.; Qiu, X.; Zhou, Q.; Shen, Z.; Wang, P.; Zhou, Z. Acute Toxicity, Bioactivity, and Enantioselective Behavior with Tissue Distribution in Rabbits of Myclobutanil Enantiomers. Chirality 2014, 26, 784–789. [Google Scholar] [CrossRef]
  37. Li, Y.; Dong, F.; Liu, X.; Xu, J.; Han, Y.; Zheng, Y. Enantioselectivity in Tebuconazole and Myclobutanil Non-target Toxicity and Degradation in Soils. Chemosphere 2015, 122, 145–153. [Google Scholar] [CrossRef]
  38. Dong, F.; Cheng, L.; Liu, X.; Xu, J.; Li, J.; Li, Y.; Kong, Z.; Jian, Q.; Zheng, Y. Enantioselective Analysis of Triazole Fungicide Myclobutanil in Cucumber and Soil under Different Application Modes by Chiral Liquid Chromatography/Tandem Mass Spectrometry. J. Agric. Food Chem. 2012, 60, 1929–1936. [Google Scholar] [CrossRef]
  39. Arunrat, N.; Sereenonchai, S.; Sansupa, C.; Kongsurakan, P.; Hatano, R. Effect of Rice Straw and Stubble Burning on Soil Physicochemical Properties and Bacterial Communities in Central Thailand. Biology 2023, 12, 501. [Google Scholar] [CrossRef] [PubMed]
  40. Furtak, K.; Gałązka, A. Enzymatic Activity as a Popular Parameter Used to Determine the Quality of the Soil Environment. Pol. J. Agron. 2019, 37, 22–30. [Google Scholar] [CrossRef]
  41. Brzezińska, M.; Włodarczyk, T. Enzymes of Intracellular Redox Trasformations (Oxidoreductases). Acta Agroph. 2005, 3, 11–26. [Google Scholar]
  42. Roman, D.L.; Voiculescu, D.I.; Filip, M.; Ostafe, V.; Isvoran, A. Effects of Triazole Fungicides on Soil Microbiota and on the Activities of Enzymes Found in Soil: A Review. Agriculture 2021, 11, 893. [Google Scholar] [CrossRef]
  43. Kim, S.; Chen, J.; Cheng, T.; Gindulyte, A.; He, J.; He, S.; Li, Q.; Shoemaker, B.A.; Thiessen, P.A.; Yu, B.; et al. PubChem 2019 Update: Improved Access to Chemical Data. Nucleic Acids Res. 2018, 47, D1102–D1109. [Google Scholar] [CrossRef]
  44. Imakumbili, M.L.E.; Semu, E.; Semoka, J.M.R.; Abass, A.; Mkamilo, G. Soil Nutrient Adequacy for Optimal Cassava Growth, Implications on Cyanogenic Glucoside Production: A Case of Konzo-Affected Mtwara Region, Tanzania. PLoS ONE 2019, 14, e0216708. [Google Scholar] [CrossRef]
  45. Schinner, F.; Ohlinger, R.; Kandeler, E.; Margesin, R. Enzymes Involved in Intracellular Metabolism. In Methods in Soil Biology, 1st ed.; Schinner, F., Ed.; Springer: Berlin/Heidelberg, Germany, 1996; p. 426. [Google Scholar]
  46. Alef, K.; Nannipieri, P. Enzyme activities. In Methods in Applied Soil Microbiology and Biochemistry; Alef, K., Nannipieri, P., Eds.; Academic Press: London, UK, 1995; pp. 311–373. [Google Scholar]
  47. Dick, R.P.; Breakwell, D.P.; Turco, R.F. Soil Enzyme Activities and Biodiversity Measurements as Integrative Microbiological Indicators. In Methods for Assessing Soil Quality; Soil Science Society of America: Madison, WI, USA, 1997; pp. 247–271. [Google Scholar]
  48. Drăgan-Bularda, M. Lucrări Practice de microbiologie General; Universitatea Babes Bolyai: Cluj-Napoca, Romania, 2000; pp. 175–191. [Google Scholar]
  49. Caraba, M.N.; Roman, D.L.; Caraba, I.V.; Isvoran, A. Assessment of the Effects of the Herbicide Aclonifen and Its Soil Metabolites on Soil and Aquatic Environments. Agriculture 2023, 13, 1226. [Google Scholar] [CrossRef]
  50. Filimon, M.N.; Roman, D.L.; Caraba, I.V.; Isvoran, A. Assessment of the Effect of Application of the Herbicide S-Metolachlor on the Activity of Some Enzymes Found in Soil. Agriculture 2021, 11, 469. [Google Scholar] [CrossRef]
  51. Filimon, M.N.; Roman, D.L.; Bordean, D.M.; Isvoran, A. Impact of the Herbicide Oxyfluorfen on the Activities of Some Enzymes Found in Soil and on the Populations of Soil Microorganisms. Agronomy 2021, 11, 1702. [Google Scholar] [CrossRef]
  52. Filimon, M.N.; Voia, S.O.; Vladoiu, D.L.; Isvoran, A.; Ostafe, V. Temperature Dependent Effect of Difenoconazole on Enzymatic Activity From Soil. J. Serb. Chem. Soc. 2015, 80, 1127–1137. [Google Scholar] [CrossRef]
  53. Berman, H.M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.N.; Weissig, H.; Shindyalov, I.N.; Bourne, P.E. The Protein Data Bank. Nucleic Acids Res. 2000, 28, 235–242. [Google Scholar] [CrossRef]
  54. Gómez, S.; Navas-Yuste, S.; Payne, A.M.; Rivera, W.; López-Estepa, M.; Brangbour, C.; Fullà, D.; Juanhuix, J.; Fernández, F.J.; Vega, M.C. Peroxisomal Catalases From the Yeasts Pichia pastoris and Kluyveromyces lactis as Models for Oxidative Damage in Higher Eukaryotes. Free Radic. Biol. Med. 2019, 141, 279–290. [Google Scholar] [CrossRef]
  55. Korkhin, Y.; Frolow, F.; Bogin, O.; Peretz, M.; Kalb, A.J.; Burstein, Y. Crystalline Alcohol Dehydrogenases From the Mesophilic Bacterium Clostridium beijerinckii And the Ahermophilic Bacterium Thermoanaerobium brockii: Preparation, Characterization and Molecular Symmetry. Acta Crystallogr. D 1996, 52, 882–886. [Google Scholar] [CrossRef]
  56. Serrière, J.; Robert, X.; Perez, M.; Gouet, P.; Guillon, C. Biophysical Characterization and Crystal Structure of the Feline Immunodeficiency Virus p15 matrix protein. Retrovirology 2013, 10, 64. [Google Scholar] [CrossRef]
  57. Kostrewa, D.; Wyss, M.; D’Arcy, A.; van Loon, A.P. Crystal Structure of Aspergillus niger pH 2.5 Acid Phosphatase at 2. 4 A Resolution. J. Mol. Biol. 1999, 288, 965–974. [Google Scholar] [CrossRef]
  58. Jonas, S.; van Loo, B.; Hyvönen, M.; Hollfelder, F. A New Member of the Alkaline Phosphatase Superfamily with a Formylglycine Nucleophile: Structural and Kinetic Characterisation of a Phosphonate Monoester Hydrolase/Phosphodiesterase from Rhizobium leguminosarum. J. Mol. Biol. 2008, 384, 120–136. [Google Scholar] [CrossRef]
  59. James, M.N.G.; Sielecki, A.R.; Brayer, G.D.; Delbaere, L.T.J.; Bauer, C.A. Structures of Product and Inhibitor Complexes of Streptomyces griseus Protease A at 1.8 Å Resolution: A Model for Serine Protease Catalysis. J. Mol. Biol. 1980, 144, 43–88. [Google Scholar] [CrossRef] [PubMed]
  60. Benini, S.; Kosikowska, P.; Cianci, M.; Mazzei, L.; Vara, A.G.; Berlicki, Ł.; Ciurli, S. The Crystal Structure of Sporosarcina pasteurii Urease in A Complex with Citrate Provides New Hints for Inhibitor Design. J. Biol. Inorg. Chem. 2013, 18, 391–399. [Google Scholar] [CrossRef] [PubMed]
  61. Varadi, M.; Anyango, S.; Deshpande, M.; Nair, S.; Natassia, C.; Yordanova, G.; Yuan, D.; Stroe, O.; Wood, G.; Laydon, A. AlphaFold Protein Structure Database: Massively Expanding the Structural Coverage of Protein-sequence Space with High-accuracy Models. Nucleic Acids Res. 2022, 50, D439–D444. [Google Scholar] [CrossRef] [PubMed]
  62. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A Visualization System for Exploratory Research and Analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef]
  63. Grosdidier, A.; Zoete, V.; Michielin, O. SwissDock, A Protein-small Molecule Docking Web Service Based on EADock DSS. Nucleic Acids Res. 2011, 39, W270–W277. [Google Scholar] [CrossRef]
  64. Salentin, S.; Schreiber, S.; Haupt, V.J.; Adasme, M.F.; Schroeder, M. PLIP: Fully Automated Protein–ligand Interaction Profiler. Nucleic Acids Res. 2015, 43, W443–W447. [Google Scholar] [CrossRef]
  65. Hammer, O.; Harper, D.; Ryan, P. PAST: Paleontological Statistics Software Package for Education and Data Analysis. Palaeontol. Electron. 2001, 4, 1–9. [Google Scholar]
  66. Saha, A.; Pipariya, A.; Bhaduri, D. Enzymatic Activities and Microbial Biomass in Peanut Field Soil as Affected by the Foliar Application of Tebuconazole. Environ. Earth Sci. 2016, 75, 558. [Google Scholar] [CrossRef]
  67. Baćmaga, M.; Wyszkowska, J.; Kucharski, J. Response of Soil Microorganisms and Enzymes to the Foliar Application of Helicur 250 EW Fungicide on Horderum vulgare L. Chemosphere 2020, 242, 125163. [Google Scholar] [CrossRef]
  68. Vlădoiu, D.L.; Filimon, M.N.; Ostafe, V.; Isvoran, A. Effects Of Herbicides And Fungicides On The Soil Chitinolytic Activity. A Molecular Docking Approach. Ecol. Chem. Eng. S 2015, 22, 439–450. [Google Scholar] [CrossRef]
  69. Fan, L.; Tarin, M.W.K.; Zhang, Y.; Han, Y.; Rong, J.; Cai, X.; Chen, L.; Shi, C.; Zheng, Y. Patterns of Soil Microorganisms and Enzymatic Activities of Various Forest Types in Coastal Sandy Land. Glob. Ecol. Conserv. 2021, 28, e01625. [Google Scholar] [CrossRef]
  70. Hara, I.; Ichise, N.; Kojima, K.; Kondo, H.; Ohgiya, S.; Matsuyama, H.; Yumoto, I. Relationship Between the Size of the Bottleneck 15 Å from Iron in the Main Channel and the Reactivity of Catalase Corresponding to the Molecular Size of Substrates. Biochemistry 2007, 46, 11–22. [Google Scholar] [CrossRef] [PubMed]
  71. Brzezińska, M.; Włodarczyk, T.; Stępniewski, W.; Przywara, G. Soil Aeration Status and Catalase Activity. Acta Agroph. 2005, 5, 555–565. [Google Scholar]
  72. Ji, C.; Song, Z.; Tian, Z.; Feng, Z.; Fan, L.; Shou, C.; Zhao, M. Enantioselectivity in the Toxicological Effects of Chiral Pesticides: A Review. Sci. Total Environ. 2023, 857, 159656. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Three-dimensional structures and IUPAC names of the stereoisomers of myclobutanil: nitrogen—blue, carbon—brown, chloride—green, hydrogen—white.
Figure 1. Three-dimensional structures and IUPAC names of the stereoisomers of myclobutanil: nitrogen—blue, carbon—brown, chloride—green, hydrogen—white.
Agriculture 13 01956 g001
Figure 2. Soil dehydrogenases activity (DA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Figure 2. Soil dehydrogenases activity (DA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Agriculture 13 01956 g002
Figure 3. Soil phosphatase activity (PhA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Figure 3. Soil phosphatase activity (PhA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Agriculture 13 01956 g003
Figure 4. Soil urease activity (UA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Figure 4. Soil urease activity (UA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Agriculture 13 01956 g004
Figure 5. Soil catalase activity (CA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Figure 5. Soil catalase activity (CA) for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Agriculture 13 01956 g005
Figure 6. Soil protease (PA) activity for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Figure 6. Soil protease (PA) activity for every experimental variant and determined every 7 days for a period of 28 days: C—control soil, D—normal dose of myclobutanil (0.1 L/ha).
Agriculture 13 01956 g006
Figure 7. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Clostridium beijerinckii dehydrogenase (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The substrate dihydro-nicotinamide-adenine-dinucleotide phosphate (yellow sticks) illustrates the position of the active site.
Figure 7. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Clostridium beijerinckii dehydrogenase (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The substrate dihydro-nicotinamide-adenine-dinucleotide phosphate (yellow sticks) illustrates the position of the active site.
Agriculture 13 01956 g007
Figure 8. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Bacillus subtilis phosphatase (mesh hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The position of the active site is illustrated by the signaling inhibitory peptide (yellow sticks).
Figure 8. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Bacillus subtilis phosphatase (mesh hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The position of the active site is illustrated by the signaling inhibitory peptide (yellow sticks).
Agriculture 13 01956 g008
Figure 9. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Streptomyces griseus protease (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The position of the active site is illustrated by the inhibitor tetrapeptide ACE-PRO-ALA-PRO-PHE (yellow sticks).
Figure 9. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the active site of Streptomyces griseus protease (solid hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic). The position of the active site is illustrated by the inhibitor tetrapeptide ACE-PRO-ALA-PRO-PHE (yellow sticks).
Agriculture 13 01956 g009
Figure 10. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the: (a) Bacillus subtilis catalase (mesh hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic), the position of the active site being illustrated by the cofactor hem (yellow sticks); (b) Bacillus pasteurii urease (mesh hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic), the position of the active site being illustrated by the ligand citrate ion (yellow sticks).
Figure 10. Binding of (R)-myclobutanil (red sticks) and of (S)-myclobutanil (green sticks) to the: (a) Bacillus subtilis catalase (mesh hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic), the position of the active site being illustrated by the cofactor hem (yellow sticks); (b) Bacillus pasteurii urease (mesh hydrophobicity surface, blue regions are hydrophilic and orange regions are hydrophobic), the position of the active site being illustrated by the ligand citrate ion (yellow sticks).
Agriculture 13 01956 g010
Table 1. Experimental variants considered in this study: D—normal dose of myclobutanil.
Table 1. Experimental variants considered in this study: D—normal dose of myclobutanil.
Experimental VariantC (Control)½ DD2D
Myclobutanil dose0—untreated dose0.1 L/ha0.2 L/ha0.4 L/ha
Table 2. Crystallographic structures of enzyme found in soil and considered in the present study: PDB ID—identifier of the structural file in Protein Data Bank, AF—AlphaFold structural model.
Table 2. Crystallographic structures of enzyme found in soil and considered in the present study: PDB ID—identifier of the structural file in Protein Data Bank, AF—AlphaFold structural model.
EnzymeOrganismPDB ID/AF
catalase 2Bacillus subtilisAF-P42234-F1
catalaseKomagataella pastoris6RJN [54]
catalaseRhizobium melilotiAF-Q9X576-F1
dehydrogenaseClostridium beijerinckii1KEV chain A [55]
dehydrogenaseAspergillus fumigatus7RK5 chain A
dehydrogenaseRhizobium leguminosarum8C54 chain A
phosphataseBacillus subtilis4I9C chain A [56]
phosphataseAspergillus niger1QFX [57]
phosphataseRhizobium leguminosarum2VQR [58]
protease AStreptomyces griseus4SGA chain E [59]
metalloproteaseAspergillus clavatus7Z6T
proteaseRhizobium leguminosarumAF-A0A0Q7ALB8-F1
ureaseBacillus pasteurii4AC7 chain C [60]
ureaseAspergillus fumigatusAF-Q6A3P9-F1
ureaseRhizobium leguminosarumAF-Q1MCV9-F1
Table 3. Results of Dunn’s post hoc test for enzymatic activities of soil dehydrogenases (above diagonal) and of enzymatic activities of soil phosphatases (below diagonal). Statistically significant data (p < 0.05) are written in bold.
Table 3. Results of Dunn’s post hoc test for enzymatic activities of soil dehydrogenases (above diagonal) and of enzymatic activities of soil phosphatases (below diagonal). Statistically significant data (p < 0.05) are written in bold.
Dose/
Time
(Days)
p Values
Control1/2DD2D
7142128714212871421287142128
Control7 0.740.670.720.340.460.640.570.600.890.650.250.020.030.04<0.01
140.04 0.930.980.200.280.890.370.840.850.900.140.010.010.02<0.01
210.630.11 0.950.170.240.960.320.910.780.980.120.010.010.02<0.01
28<0.010.350.01 0.190.270.910.350.860.830.930.130.010.010.02<0.01
1/2D70.420.200.750.03 0.830.150.700.140.270.160.850.160.250.30<0.01
140.110.640.260.160.41 0.230.860.200.380.230.690.110.170.21<0.01
210.870.050.75<0.010.520.14 0.300.950.740.990.11<0.010.010.01<0.01
28<0.010.470.020.840.040.230.01 0.270.470.300.570.070.120.15<0.01
1D70.250.350.500.060.730.640.320.10 0.700.940.09<0.010.010.01<0.01
140.010.680.040.600.090.380.020.750.18 0.760.200.010.020.03<0.01
210.620.110.990.010.760.260.740.020.51<0.05 0.11<0.010.010.01<0.01
28<0.010.14<0.010.580.010.05<0.010.450.020.28<0.01 0.230.340.39<0.01
2D70.090.680.230.180.380.950.130.260.600.410.240.06 0.800.72<0.01
140.090.700.220.190.370.930.120.270.580.430.230.060.98 0.91<0.01
210.780.070.840.010.600.180.910.010.380.030.83<0.010.160.15 <0.01
28<0.010.22<0.010.770.010.09<0.010.620.030.41<0.010.790.100.11<0.01
Table 4. Results of Tukey’s post hoc test for enzymatic activities of soil urease (above diagonal) and of enzymatic activities of soil catalases (below diagonal). Statistically significant data (p < 0.05) are written in bold.
Table 4. Results of Tukey’s post hoc test for enzymatic activities of soil urease (above diagonal) and of enzymatic activities of soil catalases (below diagonal). Statistically significant data (p < 0.05) are written in bold.
Dose/
Time
(Days)
p Values
Control1/2DD2D
7142128714212871421287142128
Control7 1.00<0.01<0.010.321.000.010.011.001.000.010.040.690.52<0.01<0.01
140.03 0.01<0.010.871.000.080.060.981.000.050.260.990.97<0.01<0.01
210.250.34 1.000.520.011.001.00<0.010.021.000.980.200.320.991.00
280.160.470.80 0.32<0.011.001.00<0.010.011.000.920.100.171.001.00
1/2D70.530.010.080.04 0.790.950.910.110.930.901.001.001.000.040.20
140.040.940.370.520.01 0.050.040.991.000.040.190.980.93<0.01<0.01
210.970.040.270.170.50<0.05 1.00<0.010.111.001.000.670.820.690.98
280.630.100.500.360.270.120.66 <0.010.091.001.000.590.750.760.99
1D70.680.010.120.070.830.010.650.37 0.95<0.010.010.350.22<0.01<0.01
140.030.910.280.41<0.010.850.030.080.01 0.080.351.000.99<0.01<0.01
210.430.180.720.540.160.210.460.760.230.15 1.000.570.730.780.99
280.740.080.410.290.340.090.770.880.460.060.65 0.950.990.320.81
2D70.26<0.010.020.010.61<0.010.240.110.47<0.010.050.14 1.000.010.06
140.080.700.560.740.020.760.090.210.030.620.340.16<0.01 0.020.10
210.23<0.010.020.010.57<0.010.220.090.43<0.01<0.050.130.95<0.01 1.00
280.21<0.010.020.010.53<0.010.190.080.40<0.010.040.110.91<0.010.95
Table 5. Results of Tukey’s post hoc test for enzymatic activities of soil protease (above diagonal). Statistically significant data (p < 0.05) are written in bold.
Table 5. Results of Tukey’s post hoc test for enzymatic activities of soil protease (above diagonal). Statistically significant data (p < 0.05) are written in bold.
Dose/
Time
(Days)
p Values
Control1/2DD2D
7142128714212871421287142128
Control7 <0.01<0.010.621.000.02<0.010.041.000.01<0.01<0.010.40<0.01<0.01<0.01
14 1.000.490.011.001.001.000.011.000.981.000.711.000.951.00
21 0.29<0.011.001.000.99<0.011.001.001.000.481.000.991.00
28 0.890.910.180.970.880.660.030.231.000.230.020.44
1/2D7 0.07<0.010.121.000.02<0.01<0.010.71<0.01<0.010.01
14 0.991.000.061.000.671.000.981.000.581.00
21 0.95<0.011.001.001.000.321.001.001.00
28 0.111.000.510.981.000.980.411.00
1D7 0.02<0.01<0.010.71<0.01<0.010.01
14 0.921.000.851.000.861.00
21 1.000.061.001.000.99
28 0.391.001.001.00
2D7 0.400.040.65
14 1.001.00
21 0.97
28
Table 6. Interaction energies for the best binding modes of the myclobutanil enantiomers to soil enzymes.
Table 6. Interaction energies for the best binding modes of the myclobutanil enantiomers to soil enzymes.
Enzyme/OrganismΔG (kcal/mol)
(R)-myclobutanil(S)-myclobutanil
Catalase Bacillus subtilis--
Komagataella pastoris−7.06−7.19
Rhizobium meliloti--
DehydrogenaseClostridium beijerinckii−9.25−9.68
Aspergillus fumigatus−8.77−8.49
Rhizobium leguminosarum−7.04−6.76
PhosphataseBacillus subtilis−8.04−8.25
Aspergillus niger−5.92−5.57
Rhizobium leguminosarum−7.33−7.25
ProteaseStreptomyces griseus−7.15−7.32
Aspergillus clavatus−5.83−4.75
Rhizobium leguminosarum−6.80−7.62
Table 7. Enzyme residues involved in the interactions with the enantiomers of myclobutanil obtained using Protein-Ligand Interaction Profiler (PLIP) software and compared with the enzymes residues involved in the interactions with the ligands that are present in the crystallographic structures: NADPH—dihydro-nicotinamide-adenine-dinucleotide phosphate, NAG—2-acetamido-2-deoxy-beta-D-glucopyranose.
Table 7. Enzyme residues involved in the interactions with the enantiomers of myclobutanil obtained using Protein-Ligand Interaction Profiler (PLIP) software and compared with the enzymes residues involved in the interactions with the ligands that are present in the crystallographic structures: NADPH—dihydro-nicotinamide-adenine-dinucleotide phosphate, NAG—2-acetamido-2-deoxy-beta-D-glucopyranose.
EnzymeLigand/SubstrateInteracting Residues
Clostridium beijerinckii
dehydrogenase
(R)-myclobutanilARG200, LYS340
(S)-myclobutanilILE175, ARG200, LYS340
NADPHTHR38, SER39, ILE175, ALA177, VAL178, GLY179, SER199, ARG200, GLU247, ASN266, TYR267, LYS340
Aspergillus fumigatus dehydrogenase(R)-myclobutanilPHE45, ASN200, ASN312, ARG385
(S)-myclobutanilPHE45, ASN200, ASP239, LYS307, ARG385, LYS393
NADPHGLY44, PHE45, GLN74, ASP77, ASP199, ASN200, VAL238, THR242
Rhizobium leguminosarum dehydrogenase(R)-myclobutanilILE159, PRO161, SER189, GLN190, PHE237, VAL243, PHE394
(S)-myclobutanilILE159, LYS186, ALA219, VAL243
NADPHTHR160, PRO161, LYS186, PRO187, SER240, GLU261
Bacillus subtilis phosphatase(R)-myclobutanilTYR66, TYR153, SER186, LEU230, ARG223, ASN227
(S)-myclobutanilTYR66, TYR153, GLN183, ARG223, ASN227, LEU230
GLN-ARG-GLY-MET-ILETYR66, TYR152, TYR153, LYS155, GLN183, LEU187, ASP194, ARG223, TYR226, ASN227, SER260, GLN263, PHE266, TYR300, GLU303, ALA334, ASP335, ASP 338
Aspergillus niger phosphatase(R)-myclobutanilVAL147, LEU312
(S)-myclobutanilVAL266, PHE345, GLY346
NAGTHR300, LEU432, ASN439
Rhizobium leguminosarum phosphatase(R)-myclobutanilTYR105, PHE138, ASN141, TYR215
(S)-myclobutanilTYR105, LEU125, PHE138, ASN141, TYR215
Streptomyces griseus protease(R)-myclobutanilPHE1, PRO2, ALA3, TYR171, GLY193, THR226
(S)-myclobutanilPHE1, PRO2, ALA3, TYR171, ALA192, GLY193
ACE-PRO-ALA-PRO-PHEHIS57, VAL169, ASN170, TYR171, ALA192, GLN192, PRO192, GLY193, ASP194, SER195, SER214, GLY215, GLY216
Aspergillus clavatus protease(R)-myclobutanilLEU204, VAL274
(S)-myclobutanilLEU204, VAL274
Rhizobium leguminosarum protease(R)-myclobutanilTRP61, PHE35, ARG145
(S)-myclobutanilARG91, PHE135, ARG145
(R)-myclobutanilTYR189, LEU445, VAL448, PHE449, ARG194,
Komagataella pastoris catalase(S)-myclobutanilTYR189, ARG194, ASN204, ALA444, LEU445
NADPHPRO142, HIS185, SER192, ASN204, VAL293, TRP294, HIS296, GLN455
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Roman, D.L.; Matica, M.A.; Ciorsac, A.; Boros, B.V.; Isvoran, A. The Effects of the Fungicide Myclobutanil on Soil Enzyme Activity. Agriculture 2023, 13, 1956. https://doi.org/10.3390/agriculture13101956

AMA Style

Roman DL, Matica MA, Ciorsac A, Boros BV, Isvoran A. The Effects of the Fungicide Myclobutanil on Soil Enzyme Activity. Agriculture. 2023; 13(10):1956. https://doi.org/10.3390/agriculture13101956

Chicago/Turabian Style

Roman, Diana Larisa, Mariana Adina Matica, Alecu Ciorsac, Bianca Vanesa Boros, and Adriana Isvoran. 2023. "The Effects of the Fungicide Myclobutanil on Soil Enzyme Activity" Agriculture 13, no. 10: 1956. https://doi.org/10.3390/agriculture13101956

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

Article Metrics

Back to TopTop