Next Article in Journal
Copper Speciation in Wine Growing-Drain Waters: Mobilization, Transport, and Environmental Diffusion
Previous Article in Journal
Potential Use of Biochar as a Mitigation Strategy for Salinity-Related Issues in Tomato Plants (Solanum lycopersicum L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dissipation Kinetics, Leaching, and Ecological Risk Assessment of S-Metolachlor and Benfluralin Residues in Soil

by
Paraskevas Parlakidis
1,*,
George S. Adamidis
1,
Georgios D. Gikas
2,
Sofia Vasiliou
1,
Melpomeni Kissa
1,
Konstantinos Doitsinis
1,
Christos Alexoudis
1 and
Zisis Vryzas
1
1
Laboratory of Agricultural Pharmacology and Ecotoxicology, Faculty of Agricultural Development, Democritus University of Thrace, 193 Pantazidou, 68200 Orestias, Greece
2
Laboratory of Ecological Engineering and Technology, Department of Environmental Engineering, School of Engineering, Democritus University of Thrace, 67100 Xanthi, Greece
*
Author to whom correspondence should be addressed.
Environments 2024, 11(1), 18; https://doi.org/10.3390/environments11010018
Submission received: 23 December 2023 / Revised: 12 January 2024 / Accepted: 15 January 2024 / Published: 18 January 2024

Abstract

:
The use of selective herbicides is one of the best methods for weed management. However, the extensive use of herbicides can have adverse impacts on non-target organisms. The goals of this study were to assess the dissipation kinetics, leaching, and ecological risk assessment of S-metolachlor and benfluralin residues in silty loam soil planted with chickpea (Cicer arietinum L.). The experimental setup included four different layers with four replications corresponding to an experimental randomized complete block design consisting of 16 plots. The application doses of S-metolachlor and benfluralin were 1350 and 1920 g a.i./ha, respectively, according to manufacturer recommendations. Soil samples were split into four depths, 0 to 20 cm (Layer A), 20 to 40 cm (Layer B), 40 to 60 cm (Layer C), and 60 to 80 cm (Layer D), to determine the dissipation kinetics and the leaching behavior of the herbicides. Gas chromatography coupled with the electron capture detector (GC-ECD) method was developed and validated for the determination of S-metolachlor and benfluralin residues in soil. The analytes were extracted from the soil with distilled water and ethyl acetate followed by solid-phase extraction (SPE). The limit of quantification (LOQ) of the method was 0.1 μg/g, and the recoveries of S-metolachlor and benfluralin were in the ranges 81% to 97% and 88% to 101%, respectively, with relative standard deviations (RSD) of less than 9.7%. The dissipation kinetics of S-metolachlor and benfluralin in soil (0–20 cm) followed first-order kinetics with half-lives of 21.66 and 30.13 days, respectively. The results for samples obtained from the 20–80 cm soil profile showed that both benfluralin and S-metolachlor presented high leaching, following preferential flow. Also, a soil ecological risk assessment was conducted in the top 0–20 cm soil profile, estimating the toxicity–exposure ratio (TER) for four soil organisms and the risk quotient (RQ). The mean herbicide levels found at the studied soil profile at 0 days (2 h) and 60 days of the experiment were used for risk assessment. In the first case, the mean pesticide concentration (MPC) gives a worst-case scenario (ws); in the second case, a dissipation scenario (ds) is given using the respective MPC. In all cases, both TER and RQ values showed that benfluralin corresponds to a higher risk than S-metolachlor for soil organisms.

1. Introduction

Soil pollution is an international issue with both natural and anthropogenic causes. Compounds, chemicals, and chemical agents have been used as a result of urbanization, industrialization, and rising food consumption, which over time has led to the dispersion and buildup of pollutants in the environment [1]. Heavy metals, pesticides, and polycyclic aromatic hydrocarbons (PAHs) are typical soil contaminants [2]. The European Commission has already named eight soil concerns in the 2006 EU Soil Strategy: soil erosion, soil contamination, floods and landslides, reduction in soil organic matter, salinization, soil compaction, soil sealing, and soil biodiversity [3]. The newly released “EU Soil Strategy for 2030” highlights the advantages of healthy soils for people, food, nature, and climate, demonstrating how the soil will be treated in the future [4].
The environmental behavior of pesticides can be influenced by several factors and mechanisms. The run-off, adsorption, and leaching potential of pesticides is significantly influenced by the soil’s properties, including its organic carbon content, texture, pH, type of clay mineral, dissolved organic matter, and cation exchange capacity, and the physicochemical characteristics of the pesticides, such as their ionization, water solubility, volatility, octanol–water partition coefficients, and thermo-, photo-, and hydrolysis stability all have an impact on pesticide fate [5]. The intensity of rainfall and irrigation, biological processes (biodegradation), and agricultural techniques also affect pesticide fate [6].
Pesticides that have not been absorbed or broken down while moving through the soil, a process known as leaching, move down into the lower soil layers with percolating water. If appropriate filtration procedures are lacking, pesticides which have leached from the soil and are moving through the rock layer are likely to pollute groundwater. Evaluating pesticides’ leaching behavior in agricultural soils is crucial to prevent the contamination of groundwater [7].
Weed infestation raises production expenses, lowers yields, and degrades product quality. The use of selective herbicides is one of the best methods for weed management [8]. Herbicides are used to prevent weeds and crops from competing with each other. Herbicides are applied at the highest concentrations of active chemicals per hectare of all pesticides. They can efficiently suppress weeds when used alone or in mixes, but they can also cause unexpected side effects on organisms inside and beyond the treated field. The amount of herbicide that reaches the target plants is frequently less than 0.1%, with the remainder absorbed by the crop, left in the soil, or polluting the environment [9].
Both terrestrial and aquatic environments have started to demonstrate the negative impacts of herbicides’ excessive use on non-target species, either directly or indirectly. Species that play significant ecological roles in agroecosystems, such as nematodes, earthworms, collembolans and isopods, spiders, and insects, have shown the physiological and behavioral consequences of exposure at the organism level [10].
Ecological risk assessment is required for the safe usage of pesticides and the sustainable management of soil ecosystems. The key aspects of the EFSA’s risk assessment process are the evaluation of exposure and the characterization of risk. The toxicity–exposure ratio (TER), which measures the relationship between toxicity and exposure concentrations, determines the risks caused by pesticides [11]. The risk quotient (RQ) approach is frequently used in the environmental risk assessment of pesticides and other chemicals to assess risk quantitatively. A basic formula for calculating RQ is to divide a number for environmental exposure by a value for the toxicity endpoint. The RQ is the ratio of exposure to effect as a result. Risk analysts and other decisionmakers can then utilize the RQ to determine whether the value exceeds established threshold levels of concern [12].
S-metolachlor (GAS name: (2-chloro-N-(2-ethyl-6-methylphenyl)-N-(methoxy-1-methylethyl) acetamide) is regularly used on more than 70 crops and has been available for more than 60 years. S-metolachlor is a selective chloroacetamide herbicide approved for use on a variety of high-value, small-acreage crops, including several vegetables and small fruits as well as major field crops including cotton, corn, soybeans, and cotton. It performs effectively on small-seeded broadleaf plants and annual grasses. S-metolachlor’s residual activity increases its usefulness for managing weeds that are resistant to post-emergence herbicides [13]. It is classified as an inhibitor of very-long-chain fatty acid formation [14].
Dinitroaniline pesticide benfluralin (GAS name: (N-butyl-N-ethyl-2,6-dinitro-4-(trifluoromethyl) benzenamine) is used as a pre-emergence herbicide to manage grasses and broad-leaved weeds in numerous crops, such as alfalfa, clover, lettuce, bean, pea, and other crops. It disrupts the mitotic cycle of weeds, which stunts their growth [15,16]. By 12 August 2023, the member states of the EU must revoke authorizations for plant protection products that use benfluralin as an active ingredient. However, other countries with significant agricultural sectors, such as the UK, have approved its use [16]. Any grace time given to member states in accordance with Article 46 of Regulation (EC) No. 1107/2009 will expire on 12 May 2024 [17].
According to our knowledge, the majority of studies describe adsorption or degradation experiments, dissipation kinetics, and leaching behavior for S-metolachlor and benfluralin in laboratory conditions, usually using batch experiments or soil columns [10]. In our opinion, a similar experiment in field conditions could fill a research gap.
Validation of the methods for the qualification and quantification of the two herbicides is necessary to show that it is appropriate for the intended purpose and to ensure the reliability of the results, and is frequently lacking in published studies. This can be achieved through a variety of tests, which can allow researchers to pinpoint important technique and performance characteristics, such as selectivity, limits of detection and quantitation, working range, sensitivity, trueness, precision, and measurement uncertainty [18]. Similarly, pesticide risk assessment studies are usually missing, even though they are mandatory for pesticide registration.
The well-known solid-phase extraction (SPE) technique and gas chromatograph systems equipped with electron capture detectors (GC-ECD) provide satisfactory analytical methods for numerous pesticides. Surprisingly, SPE paired with the GC-ECD method for S-metolachlor and benfluralin determination in soil is missing in the literature. Therefore, the objectives of this study are (a) to validate a simple, sensitive, and effective method to determine S-metolachlor and benfluralin residues in soil under chickpea cultivation based on SPE using GC-ECD, (b) to study the persistence, dissipation, and leaching of S-metolachlor and benfluralin in soil, and (c) to determine whether herbicide residues pose a risk to soil organisms, such as earthworms (Eisenia fetida), enchytraeids (Enchytraeus crypticus), springtails (Folsomia candida), and mites (Hypoaspis aculeifer), as well as nitrogen and carbon mineralization microorganisms, according to the EFSA’s guidelines [19].

2. Materials and Methods

2.1. Reagents and Materials

S-metolachlor and benfluralin analytical standards (purity 99.5%) were purchased from Dr. Ehrenstorfer GmbH (Augsburg, Germany). Table 1 presents the physico-chemical parameters of the herbicides. Methanol and ethyl acetate of HPLC-grade purity and n-hexane of proanalysis grade were obtained from Riedel de Haen (Seelze, Germany). They were utilized for the cleanup and elution of the samples. Sodium chloride was purchased from Sigma-Aldrich (Zwijndrecht, The Netherlands) and was used during the herbicide extraction process. C18 cartridges (1000 mg, 6 mL; Thermo Scientific™, Vantaa, Finland) were used for solid-phase extraction (SPE). The commercial formulations Dual Gold 96 EC (S-metolachlor) and Bonalan 18 EC (benfluralin) were provided by Syngenta Hellas S.A. (Thessaloniki, Greece) and Gowan Crop Protection Limited (Yuma, AZ, USA), respectively.

2.2. Experimental Setup

The supervised field experiment was conducted at the Democritus University experimental farm, Orestiada, Thrace, Northern Greece (41°30′08.7″ N, 26°32′25.6″ E), to determine the persistence, behavior, and ecological risk assessment of S-metolachlor and benfluralin in soil. The farm had never previously received S-metolachlor or benfluralin treatment. The experimental setup included four different layers with four replications corresponding to an experimental randomized complete block design consisting of 16 plots. Each experimental plot had a surface area of 20 m2. A distance of one meter separated the plots and blocks from one another. Samples were randomly pulled from each plot. Soil samples were split into four depths, 0 to 20 cm (Layer A), 20 to 40 cm (Layer B), 40 to 60 cm (Layer C), and 60 to 80 cm (Layer D), to determine the dissipation kinetics and the leaching behavior of the herbicides (Figure 1). The soil (1 kg) samples were collected randomly at 0 (2 h), 7, 15, 30, 45, and 60 days after application in triplicate. The soil samples were drawn from the soil at random, using a soil auger. The collected samples were transferred to the analytical laboratory for sample preparation and instrumental analysis.
The farm was planted with chickpea Cicer arietinum L. (cv. Amorgos). During the experiment, all cultivation tasks were carried out according to good agricultural practices. Chickpea is considered a non-irrigated crop and thus irrigation was not applied. The commercial formulations Dual Gold 96 EC (S-metolachlor) and Bonalan 18 EC (benfluralin) were applied using an experimental Azo field plot sprayer (length 2.4 m) equipped with six flat fan nozzles. The spray volume was 300L/ha, and the pressure was set at 2.8 atm. Dual Gold 96 EC and Bonalan 18 EC were applied on 20 March, pre-emerged to chickpeas. The application doses of S-metolachlor and benfluralin were 1350 and 1920 g a.i./ha, respectively, according to the commercial formulation’s recommended doses. Herbicides were incorporated into the soil. The soil was silty loam with 47.5% silt, 35% clay, 17.5% sand, 1.5% organic matter (OM), and a pH of 8.5. Throughout the trial (from 20 March to 21 May), the average minimum and maximum temperatures were 11.38 °C and 21.61 °C, respectively. Average humidity ranged from 56% to 81% and average soil moisture from 33.6% to 62.9% (measured by the gravimetric method). Light average rainfall was logged during the experimental period, ranging from 20.3 to 48.71 mm.

2.3. Sample Preparation

The soil (1 kg) samples were collected in triplicate using a soil auger. Initially, large stones were removed from each soil sample, and then the soil samples were air-dried, homogenized using a mortar and a pestle (clods smashed), and sieved (small stones removed) with a 2 mm sieve. An amount of 250 g was obtained from the samples following the quartering method, and a representative amount (10 g) of each soil sample was placed in a 50 mL centrifuge tube. Then, 5 mL of ethyl acetate and 20 mL of water were added. Extraction was performed by ultrasonication for 20 min. About 5 g of sodium chloride was added, and the mixture was vigorously shaken for 1 min. Centrifugation at 3000 rpm for 5 min followed. The aliquot of the upper layer was treated with SPE [21].
For the cleanup, C18 cartridges were used. The cartridges were initially conditioned with 3 mL of ethyl acetate, 3 mL of methanol, and 2 mL of distilled water. Samples were loaded onto C18 cartridges connected to a SPE vacuum manifold, with a flow rate of about 5 mL/min. The elution was conducted with 3 mL of ethyl acetate and 1 mL of hexane. The eluent was collected, evaporated to near dryness using a vacuum rotary evaporator at 30 °C, and evaporated to complete dryness under a nitrogen stream. The pesticide residues were dissolved in 2.5 mL ethyl acetate for GC–ECD analysis [22].

2.4. Instrumental Analysis

A Shimadzu gas chromatograph GC-17A (Duisburg, Germany) equipped with an ECD detector and an HP-5 (30 m 0.32 mm 0.25 m) capillary column was used to quantify and qualify S-metolachlor and benfluralin. The injection volume was 2 μL, and the injector was run at 220 °C. The temperature of the oven was set to start at 80 °C with a holding time of 2 min, then increase to 220 °C at 20 °C/min, be held for 4 min, and finally increase to 250 °C at 10 °C/min and be held for 30 min. The carrier gas was helium, flowing at a rate of 1.5 mL/min. The detector was operated at 290 °C.

2.5. Method Validation

Mixed standard stock solutions were created separately for herbicides in ethyl acetate at 1 mg/mL. Using a serial dilution method, the calibration standard solutions were created from secondary stock solutions. The calibration standard solutions were prepared at the different concentration rates of 0.1, 0.2, 0.5, 1, 2, and 5 μg/L in ethyl acetate. To prove the linearity of the matrix-matched calibration curve, standard solutions were added to a blank soil matrix in the range 0.1–5 μg/L. The linearity of the calibration curve in ethyl acetate was also tested.
The SANTE guidelines [23] were followed in the development and validation of the quantitative analytical method, which was assessed using the validation criteria of specificity, linearity, sensitivity, trueness, precision, and matrix effect (ME). To test specificity by integrating the peaks at the target retention time for the eluted S-metolachlor and benfluralin, blank samples were spiked with S-metolachlor and benfluralin at the limit of quantification and injected into the GC-ECD instrument. The average area of the blank sample was divided by the average area of the standard at the reporting limit and then multiplied by 100 to obtain the specificity. Linearity was determined by plotting concentrations against the peak area derived from the GC-ECD chromatogram. The sensitivity of the method was assessed by reaching the limit of detection (LOD) and limit of quantitation (LOQ) by spiking the S-metolachlor and benfluralin with soil at the lowest concentration level complying with the analytical method’s specifications. The LOD was evaluated as three times the signal–background-noise ratio on each chromatogram, and the LOQ was the concentration of the lowest spiked level in samples giving reliable and reproducible results. To conduct the recovery studies, blank (control) samples of soil were spiked with standard solutions of S-metolachlor and benfluralin at concentrations of 0.1, 0.5, and 1 μg/g. Spiking levels were replicated five times. Ten grams of the soil sample were fortified by the addition onto the soil of the initial standard solution (1mg/mL). The fortification was performed by adding the initial solution volumes of 1 mL, 5 mL, and 10 mL for the levels 0.1, 0.5, and 1 μg/g, respectively. The repeatability (RSD %) of the procedure was also evaluated for each spiking level.
Calibration curves prepared with standards in ethyl acetate were compared with matrix-matched calibration curves in soil to determine a potential matrix effect. To prevent false negative or false positive reports, MEs (signal suppression and enhancement) were investigated [24]. ME was assessed utilizing the following equation:
M E = s l o p e   o f   t h e   c a l i b r a t i o n   c u r v e   i n   m a t r i x s l o p e   o f   t h e   c a l i b r a t i o n   c u r v e   i n   s o l v e n t 1 × 100 %

2.6. Ecological Risk Assessment

2.6.1. Exposure Assessment

The ecological risk assessment was conducted in the top 0–20 cm soil profile, according to Bhadari et al. [11]. The mean herbicide levels found at the studied soil profile at 0 (2 h) and 60 days of the experiment were used for risk assessment. In the first case, the concentration gives a worst-case scenario (MPCws), and, in the second case, a dissipation scenario is given (MPCds).
The no-observed-effect concentration (NOEC) and/or the lethal concentration at which 50% of investigated organisms demonstrate mortality (LC50), along with the median effective concentrations (EC50) for organisms including the earthworm (E. fetida), the enchytraeid (E. crypticus), the springtail (F. candida), the mite (H. aculeifer), and nitrogen and carbon mineralization microorganisms are used to evaluate the ecological risks of pesticides. Any study evaluating pesticides for approval by the EFSA for the European Union must take these organisms into account. The current investigation was predicated on the available NOEC endpoints, even though pesticides found in soil have LC50 and EC50 values.
The predicted no-effect concentration (PNEC) for the most vulnerable species was calculated using the NOEC value. The PNEC value was calculated as the lowest long-term NOEC divided by the assessment factor (AF), to overcome challenges like incomplete toxicity data, errors, and inaccurate results from the conservative method. To obtain the PNEC with an AF and take into account potential chronic effects, the most sensitive organism for each pesticide was chosen. The EU’s 2002 guidance document (SANCO/10329/2002 rev 2 final) served as the basis for selecting the AF, which can number between 10 and 1000: (a) when data from a long-term assay was available, an AF of 100 was used, (b) when at least one LC50 at one ecological level was available, an AF of 1000 was used, and (c) when two or more NOECs were available, AFs of 50 and 10, respectively, were used [25]. Based on the long-term NOECs that were available, we used AFs of 100, 50, and 10 in the current investigation (Table 2). The ecological risk assessment was conducted using the toxicity–exposure ratio (TER) [25] and the risk quotient (RQ) [26].

2.6.2. Toxicity–Exposure Ratio

The TER method connects exposure and toxicity based on EC [25]. Using the TER for the test organisms and the following Equation (2), the TER for each pesticide was calculated:
T E R = N O E C s M P C
where NOEC is the no-observed-effect concentration for all species and MPC is the mean pesticide concentration in the soil.
Acceptable risk for the organisms was indicated by TER values of ≥10 or ≥5, which are acceptable trigger point values for acute and chronic exposure, respectively.

2.6.3. Risk Quotient

The following Equation (3) was used to determine the risk quotient of a pesticide, which served as an index for the risk of a single pesticide:
RQ = M P C P N E C s
where PNECs is the predicted no-effect concentration for the most sensitive species.
According to Vryzas et al. [28], the risk quotient was divided into four categories: no risk (RQ < 0.01), lesser risk (0.01 ≤ RQ ≤ 0.1), moderate risk (0.1 ≤ RQ <1), and higher risk (RQ ≥ 1).

2.7. Dissipation Kinetics

The field experiment data on S-metolachlor and benfluralin residues obtained from the 0–20 cm sampling depth zone were subjected to the first-order dissipation kinetics equation Ct = Coe-kt, where Ct is the herbicide concentration at time t (day), Co is the initial concentration (μg/g), and k is the dissipation rate constant. The herbicides’ half-lives were calculated as DT50 = ln2/k [29].

3. Results and Discussion

3.1. Method Validation Results

A GC-ECD analytical method was developed and validated. The instrument parameters were optimized to detect, confirm, and quantify S-metolachlor and benfluralin residues in soil. With the set operating chromatographic conditions, S-metolachlor and benfluralin were separated and eluted. The approximate retention times of S-metolachlor and benfluralin were 13.438 and 10.236 min, respectively. The blank soil and herbicide chromatographs are shown in Figure 2.
Both matrix-matched standards and standards in ethyl acetate were used in the creation of calibration curves, which were then examined in three replicates. Calibration curves prepared with pesticide standards in a solvent were compared with matrix-matched calibration curves in soil to determine a potential ME. Both of them were linear, presenting R2 > 99%. The matrix effect of S-metolachlor was −17.3% and that of benfluralin was −4.86% (signal suppression). When the effect ranges from −20% to 20%, there is a mild ME [24] (Table 3). Therefore, the matrix calibration curves were used for the quantification of herbicides in the current study.
The LOD and LOQ were calculated according to guidelines provided by SANTE [23]. The LOD was determined as the signal corresponding to three times the background noise on each chromatogram, whereas the LOQ was considered the concentration at the lowest spiked level in samples giving accurate and precise quantified results. For S-metolachlor, the LOD and LOQ values were 0.01 and 0.1 μg/g, whereas for benfluralin they were 0.015 and 0.1 μg/g, respectively. The accuracy of the method was determined by spiking blank samples with S-metolachlor and benfluralin at three concentration levels (0.1, 0.5, and 1 μg/g) (n = 5). The average recovery rates of S-metolachlor and benfluralin in soil were from 81% to 97% and from 88% to 101%, respectively (Table 4). The RSDs (%) for S-metolachlor and benfluralin were from 6.6% to 8.5% and from 8.7% to 9.7%, respectively (Table 4). These values were acceptable and reached the conditions of analysis of pesticide residues, since SANTE [23] guidelines demand recovery ranges of 70–120% and RSDs up to 20%.
In the last three decades, various techniques have been developed to detect S-metolachlor and benfluralin residues in soil. However, the majority of them have not been presented in detail, excluding the method validation description. Table 5 summarizes the analytical methods to determine S-metolachlor and benfluralin levels in soil reported by previous studies.

3.2. Dissipation and Leaching of S-Metolachlor and Benfluralin

A field experiment was conducted at a farm to determine the persistence and behavior of S-metolachlor and benfluralin in soil. The initial deposits of S-metolachlor and benfluralin residues obtained from the 20 cm sampling depth zone were 1.003 and 2.64 μg/g at the recommended doses of 1350 and 1920 g a.i./ha, respectively. Both herbicides showed a faster dissipation rate during the initial seven days compared with that at 15 days. In particular, S-metolachlor and benfluralin showed 33.90% and 21.10% dissipation rates, respectively, from 0 to 7 days, while from 7 to 15 days the dissipation rates were 21.93% and 17.54%, respectively. From 15 days to 60 days, the dissipation rates of the herbicides did not exceed a percentage of 15%. The final S-metolachlor and benfluralin residue levels (60 days) in the soil in the first layer were 0.13 and 0.61 μg/g, which correspond to 87.04% and 76.89% of the initial concentration, respectively (Table 6 and Table 7). The dissipations of S-metolachlor and benfluralin were well-fitted in the first-order dissipation kinetics equation with high regression coefficients R2 of 0.946 and 0.963, respectively. The half-life (DT50) of benfluralin was higher (30.13 days) than the DT50 value for S-metolachlor in soil (21.66 days). The rate constant (k) for S-metolachlor dissipation was 0.032 d−1, and the rate constant for benfluralin dissipation was 0.023 d−1 (Figure 3 and Figure 4).
The leaching potential of the herbicides was estimated by collecting samples from three different depth zones. Leaching depths for S-metolachlor and benfluralin are presented in Table 8 and Table 9, respectively. S-metolachlor and benfluralin concentrations in the 20–40, 40–60, and 60–80 cm sampling zones were below the LOD of the analytical method 2 h after herbicide application. From 7 to 45 days, S-metolachlor residues in the 20–40 cm sampling zone ranged from 0.104 to 0.202 μg/g. In the sampling zone at 40–60 cm, S-metolachlor was detected only on Day 7, at a concentration equal to 0.127 μg/g. In the 60 to 80 cm soil profile, no S-metolachlor residue was detected on Day 60. From 7 to 60 days, benfluralin residues in the 20–40 cm sampling zone ranged from 0.121 to 0.230 μg/g. In the depth zone at 40–60 cm, benfluralin residues ranged from 0.111 to 0.166 μg/g, but no benfluralin residue was found on Day 60. In the 60 to 80 cm soil profile, benfluralin was detectable only on Day 7.
According to Badou-Jeremie et al. [13] and Wołejko et al. [14], dissipation is a complex process that is affected by several physicochemical and biological changes, which cause the active substance’s concentration to decline over time. Residue degradation can be affected by several factors, including the pesticide’s stability in the soil and plants, the rate and frequency of application (initial concentration), the weather (sunlight, temperature, humidity, and wind), microorganisms, the pH of the soil and water, and the species of cultivated plants. In general, processes such as leaching, runoff, volatilization, adsorption, photodegradation, plant uptake, and biodegradation can impact the dissipation of herbicides in the field [35].
In our study, during the experimental period, air temperatures were moderate and rainfall was low, which could affect the dissipation and fate of S-metolachlor and benfluralin. As the temperature rises, pesticide vapor pressures can also rise. Because the herbicides used in this study have low vapor pressures (Table 1), losses due to volatilization were low. After herbicide application, soil incorporation was conducted. As a result, the photodegradation of pesticides was also low [36].
The GUS (groundwater ubiquity score) is an experimentally estimated value that connects pesticide half-life and Koc. The GUS can be used to categorize pesticides by their possibility of moving towards groundwater [37]. The GUS was used by Hyun et al. [38] to assess the potential of certain pesticides in the soil of Jeju Island to pollute groundwater. They listed pesticides including alachlor, metolachlor, bromacil, ethoprophos, carbofuran, and metalaxyl as groundwater pollutants if their GUS index was greater than 2.8. On the other hand, in Hawaii’s soil, the transport of certain pesticides was predicted using GUS values, and pesticide leaching was investigated [39]. With GUS 1.8, it was projected that the fungicide trifloxystrobin would not be leached. S-metolachlor and benfluralin have GUS values of 2.32 and −0.62, respectively [16,20]. Consequently, a moderate leaching potential for S-metolachlor and a low leaching potential for benfluralin are expected. However, in our study, both herbicides presented high leaching.
The high mobility of herbicides in the soil profile is due to the observed preferential flow. Preferential flow can happen in dry soils when partially water-repellent layers weaken the wetting front, creating fingered flow, or when flow is concentrated via recently developed cracks. Near-positive pore water pressures in soils that are getting close to saturation can push water out of the matrix and into highly conductive macropores, making the entire flow considerably more preferential [40]. Additionally, earlier studies indicate that preferential flow has a significant influence on herbicide leaching in the studied area due to the poor adsorption capacity for the herbicides atrazine and metolachlor in the soil profile. This is due to the fact that pollutants might avoid degradation by entering the saturated zone of the aquifer by preferential flow channels, such as plant roots, shrinking clay minerals, and earthworm burrows, rather than going through chromatographic flow within the unsaturated zone [41,42]. In our experiment, a short but intensive rainfall that temporarily created flood conditions after a dry period appeared to enhance preferential flow in the soil profile. Also, the chickpea root system and earthworms created channels in the soil, which facilitated preferential flow. Figure 5 shows earthworm, root, and crack channels in the soil profile.
The soil was silty loam with 47.5% silt, 35.0% clay, 17.5% sand, 1.5% OM, and a pH of 8.5. OM content and clay percentage in soils have a key role in the rate of adsorption. High clay and OM content in the soil promotes and facilitates greater pesticide molecule adhesion [43]. The soil and the herbicides’s physicochemical properties are both essential for herbicide adsorption in the soil. Pesticides with high Koc (Koc > 1000 L/kg) and low water solubility (water solubility < 10 mg/L) exhibit strong pesticide adsorption in soil and organic matter fractions [39,40]. S-metolachlor and benfluralin have water solubility values of 480 and 0.064 mg/L and respective Koc values of 200.2 and 646 L/kg [16,20]. Therefore, moderate adsorption of S-metolachlor and benfluralin in the soil was anticipated. Milan et al. [43] demonstrated the possibility of S-metolachlor leaching at values greater than 0.25 μg/L. According to Alleto et al., S-metolachlor adsorption in soil is low to moderate, with adsorption coefficients of Kd = 1.3 to 8.7 and a mean of 3.0 L/kg, which have been recorded from soils taken from fields with various crops and tillage practices [44]. Nevertheless, the residence period of water in the critical zone is dramatically decreased by preferential flow, which minimizes the probability that dissolved compounds will bind to soil particles, leading to low adsorption [40].
A significant factor involved in the breakdown of pesticides in the soil is microorganisms. The activity of aerobic microorganisms is reduced when the soil water capacity approaches saturation, resulting in an anoxic and nearly saturated soil–water system with a slower degradation rate and a longer half-life [45]. Throughout the experiment, the average soil moisture ranged from 33.6% to 62.9%. Consequently, although S-metolachlor and benfluralin could adversely impact soil microbial communities [33,46], the dissipation of herbicides was probably affected by biodegradation. Pesticide biodegradation, however, depends on a wide range of environmental conditions and is not just dependent on the existence of bacteria with the appropriate degrading enzymes [47].
The physical and chemical characteristics of a pesticide, as well as its interactions with the soil, plant microbiome, water, and other compounds of various types in the rhizosphere, determine how the pesticide behaves inside a plant through numerous mechanisms (such as uptake and translocation). The octanol/water partition coefficients (LogKow) of pesticides have a significant impact on their fate in plant tissues with regard to phytoaccumulation and pesticide transport within the plant. Pesticides with LogKow values between 3.0 and 4.0 are more easily absorbed and translocated in plant tissues [48,49]. S-metolachlor and benfluralin have LogKow values of 3.05 and 5.19, respectively [16,20]. Also, Gikas et al. [50] investigated the uptake and translocation of S-metolachlor within constructed wetland vegetation. The results showed that S-metolachlor was detected in different plant parts (roots, stems, and leaves) of Typha latifolia and Phragmites australis, with concentrations ranging from 1.67 μg/g to 6.18 μg/g. Benfluralin residue studies in primary crops were conducted on leafy crops (lettuces), pulses, and oilseeds (alfalfa and peanuts) after soil application and cereals (wheat) after foliar application. Low levels of benfluralin (1.3%) were observed in lettuces; the majority of residues were recovered as aqueous and organo-soluble fractions (46% and 16%, respectively), while up to 47% of the residues remained unextracted, of which 17.7% was detected to be incorporated into the plants’ natural components. Alfalfa and peanuts showed a similar trend, with the extracted fractions including a large number of unknown metabolites that together accounted for less than 3% of the total [51]. Therefore, chickpea accumulated a higher amount of S-metolachlor than benfluralin, which is clarified by their different hydrophobicity levels. This is probably another parameter that led to different half-life values for S-metolachlor and benfluralin in the top 0–20 cm of the soil profile.
In the present study, the half-life was calculated for S-metolachlor in soil by the equation DT50 = ln2/k. The half-life for the first-order kinetics was 21.66 days (Table 5). Numerous studies have recorded DT50 values for S-metolachlor in soils collected from fields with various crops and tillage practices. Previously, the first-order kinetics DT50 value for S-metolachlor dissipation in soil was lower (12 days) than that reported by Caracciolo et al. [52]. Similarly, Wołejko et al. [12] reported that the estimated value of DT50 for S-metolachlor was 11.1–14.7 days in soil. On the other hand, the DT50 values in experiments by Long et al. [53] ranged from 26.3 to 40.1 days for five different soils, and were higher than in the findings of the current study. Also, the estimated half-life for S-metolachlor was in the same range as reported by Shaner et al. [54], which ranged from 18 to 27 days. There is a lack of dissipation studies for benfluralin in soil. Vischetti et al. [33] reported the same range of DT50 as the current study for benfluralin in three different soils, ranging from 11.4 to 37.9 days. In the current study, the half-life value of benfluralin was 30.13 days (Table 6).

3.3. Soil Risk Assessment

An LOQ of 0.1 μg/g was chosen for soil risk assessment given the fact that LOQ values have been reported to range from 1.58 × 10−3 to 1.63 [55]. According to Ockleford et al. [56] and Silva et al. [57], predicted environmental concentrations (PEC) of authorized pesticides are frequently used to interpret their concentrations in soil. These PEC values are utilized in the evaluation of certain active substances and are determined using worst-case scenarios. PECs are determined for the primary crops to which the substance is applied, taking into account tillage depths of 5 cm for permanent crops and 20 cm for annual crops. Chickpea is an annual crop. Therefore, we believe that the depth of 20 cm is appropriate for risk assessment. The ecological risk based on TER under the worst-case scenario (ws) and the dissipation scenario (ds) is demonstrated in Table 10. In general, for the two herbicides and four soil organisms, benfluralin showed the highest risk under both scenarios for F. candida, because of its higher half-life and initial concentration. Also, the higher risk could be due to its higher initial concentration and lower NOEC [9]. More specifically, the TERws of benfluralin was equal to 2.08, exhibiting an acute risk for F. candida, while the TERds was 9.01, showing a chronic risk. In the cases of E. fetida and H. aculeifer, the TER values were greater than 10, which means a negligible risk. The high TER values for S-metolachlor (TER > 10) indicate a negligible risk in soil. Consequently, benfluralin is riskier than S-metolachlor for soil organisms.
The risk quotient (RQ) values under the ws scenario and ds scenario for the studied herbicides are shown in Table 11. Our study presented higher risks (RQs > 1) for both herbicides for the WS scenarios, indicating higher risk for soil organisms. For the ds scenario, the RQ value for S-metolachlor was 0.48, presenting medium risk, while the RQ value for benfluralin was 1.11. Therefore, benfluralin remains highly risky for soil organisms, even 60 days after pesticide application. The higher risk of benfluralin could be due to its higher initial concentration and half-life [9].

4. Conclusions

Our study aimed to assess the soil ecological risks of S-metolachlor and benfluralin in soil. In the present study, a sensitive and efficient GC-ECD method was developed and validated for the determination of S-metolachlor and benfluralin residues in soil. The dissipation kinetics of S-metolachlor and benfluralin in soil (0–20 cm) followed first-order kinetics with half-lives of 21.66 and 30.13 days, respectively, indicating that benfluralin is more persistent than S-metolachlor in soil. The results for samples obtained from the 20–80 cm soil profile showed that both S-metolachlor and benfluralin present high leaching following preferential flow, which significantly contributed to the dissipation of these pesticides. Also, a soil ecological risk assessment was conducted in the top 0–20 cm of the soil profile, indicating the TERs for four soil organisms and the RQ. In all cases, both the TER and RQ values showed that benfluralin is riskier than S-metolachlor for soil organisms.

Author Contributions

Conceptualization, G.D.G., Z.V. and P.P.; methodology, C.A., G.D.G., Z.V., G.S.A., P.P., M.K., S.V. and K.D.; investigation, G.S.A., P.P., M.K. and K.D.; writing—original draft preparation, G.S.A. and P.P.; writing—review and editing, G.D.G. and Z.V; supervision, C.A., G.D.G. and Z.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, B.; Hu, K.; Lysenko, V.; Khan, K.Y.; Wang, Y.; Jiang, Y.; Guo, Y. A scientometric analysis of agricultural pollution by using bibliometric software VoSViewer and Histcite™. Environ. Sci. Pollut. Res. Int. 2022, 29, 37882–37893. [Google Scholar] [CrossRef]
  2. Raffa, C.M.; Chiampo, F. Bioremediation of Agricultural Soils Polluted with Pesticides: A Review. Bioengineering 2021, 8, 92. [Google Scholar] [CrossRef]
  3. Commission of the European Communities. Communication from the Commission to the Council, the European Parliament, the European Economic and Social Committee and the Committee of the Regions; Thematic Strategy for Soil Protection; COM (2006) 231 final of 22 September 2006; European Commission: Brussels, Belgium, 2006; pp. 1–12. [Google Scholar]
  4. European Commission. Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee and the Committee of the Regions, EU Soil Strategy for 2030—Reaping the Benefits of Healthy Soils for People, Food, Nature and Climate; COM (2021) 699 final of 17.11.2021; European Commission: Brussels, Belgium, 2021; pp. 1–25. [Google Scholar]
  5. Syafrudin, M.; Kristanti, R.A.; Yuniarto, A.; Hadibarata, T.; Rhee, J.; Al-onazi, W.A.; Algarni, T.S.; Almarri, A.H.; Al-Mohaimeed, A.M. Pesticides in Drinking Water—A Review. Int. J. Environ. Res. 2021, 18, 468. [Google Scholar] [CrossRef]
  6. Parlakidis, P.; Rodriguez, M.S.; Gikas, G.D.; Alexoudis, C.; Perez-Rojas, G.; Perez-Villanueva, M.; Carrera, A.P.; Fernández-Cirelli, A.; Vryzas, Z. Occurrence of Banned and Currently Used Herbicides, in Groundwater of Northern Greece: A Human Health Risk Assessment Approach. Int. J. Environ. Res. Public Health 2022, 19, 8877. [Google Scholar] [CrossRef]
  7. Park, W.P.; Chang, K.M.; Hyun, H.N.; Boo, K.H.; Koo, B.J. Sorption and leaching characteristics of pesticides in volcanic ash soils of Jeju Island, Korea. Appl. Biol. Chem. 2020, 63, 71. [Google Scholar] [CrossRef]
  8. Drouin, G.; Droz, B.; Leresche, F.; Payraudeau, S.; Masbou, J.; Imfeld, G. Direct and indirect photodegradation of atrazine and S-metolachlor in agriculturally impacted surface water and associated C and N isotope fractionation. Environ. Sci. Process. Impacts 2021, 23, 1791–1802. [Google Scholar] [CrossRef]
  9. Samadi Kalkhoran, E.; Alebrahim, M.T.; Mohammaddoust Chamn Abad, H.R.; Streibig, J.C.; Ghavidel, A.; Tseng, T.-M.P. The Survival Response of Earthworm (Eisenia fetida L.) to Individual and Binary Mixtures of Herbicides. Toxics 2022, 10, 320. [Google Scholar] [CrossRef]
  10. Giglio, A.; Vommaro, M.L. Dinitroaniline herbicides: A comprehensive review of toxicity and side effects on animal non-target organisms. Environ. Sci. Pollut. Res. Int. 2022, 29, 76687–76711. [Google Scholar] [CrossRef]
  11. Bhandari, G.; Atreya, K.; Vašíčková, J.; Yang, X.; Geissen, V. Ecological risk assessment of pesticide residues in soils from vegetable production areas: A case study in S-Nepal. Sci. Total Environ. 2021, 788, 147921. [Google Scholar] [CrossRef]
  12. Peterson, R.K. Comparing ecological risks of pesticides: The utility of a Risk Quotient ranking approach across refinements of exposure. Pest. Manag. Sci. 2006, 62, 46–56. [Google Scholar] [CrossRef]
  13. Kouame, K.B.-J.; Savin, M.C.; Willett, C.D.; Bertucci, M.B.; Butts, T.R.; Grantz, E.; Roma-Burgos, N. S-metolachlor persistence in soil as influenced by within-season and inter-annual herbicide use. Environ. Adv. 2022, 9, 100318. [Google Scholar] [CrossRef]
  14. Wołejko, E.; Kaczyński, P.; Łozowicka, B.; Wydro, U.; Borusiewicz, A.; Hrynko, I.; Konecki, R.; Snarska, K.; Dec, D.; Malinowski, P. Dissipation of S-metolachlor in plant and soil and effect on enzymatic activities. Environ. Monit. Assess. 2017, 189, 355. [Google Scholar] [CrossRef]
  15. Travlos, I.S.; Gkotsi, T.; Roussis, I.; Kontopoulou, C.K.; Kakabouki, I.; Bilalis, D.J. Effects of the herbicides benfluralin, metribuzin and propyzamide on the survival and weight of earthworms (Octodrilus complanatus). Plant. Soil Environ. 2017, 63, 117–124. [Google Scholar] [CrossRef]
  16. Lewis, K.A.; Tzilivakis, J.; Warner, D.; Green, A. An international database for pesticide risk assessments and management. Hum. Ecol. Risk Assess. 2016, 22, 1050–1064. [Google Scholar] [CrossRef]
  17. European Commission. Comission Implanting Regulation (EU) 2021/2068 of 25 November 2021 Amending Implementing Regulation (EU) No 540/2011 as Regards the Extension of the Approval Periods of the Active Substances Benfluralin, Dimoxystrobin, Fluazinam, Flutolanil, Mecoprop-P, Mepiquat, Metiram, Oxamyl and Pyraclostrobin; European Commission: Brussels, Belgium, 2021. [Google Scholar]
  18. Škufca, D.; Kovačič, A.; Bulc, T.G.; Heath, E. Determination of 18 bisphenols in aqueous and biomass phase of high rate algal ponds: Development, validation and application. Chemosphere 2021, 271, 129786. [Google Scholar] [CrossRef]
  19. European Food Safety Authority (EFSA). EFSA Guidance Document for predicting environmental concentrations of active substances of plant protection products and transformation products of these active substances in soil. EFSA J. 2017, 15, e04982. [Google Scholar] [CrossRef]
  20. Zemolin, C.R.; Avila, L.A.; Cassol, G.V.; Massey, J.H.; Camargo, E.R. Environmental Fate of S-Metolachlor: A Review. Planta Daninha 2014, 32, 655–664. [Google Scholar] [CrossRef]
  21. Li, L.; Jiang, G.; Liu, C.; Liang, H.; Sun, D.; Li, W. Clothianidin dissipation in tomato and soil, and distribution in tomato peel and flesh. Food Control 2012, 25, 265–269. [Google Scholar] [CrossRef]
  22. Vryzas, Z.; Tsaboula, A.; Papadopoulou-Mourkidou, E. Determination of alachlor, metolachlor, and their acidic metabolites in soils by microwave-assisted extraction (MAE) combined with solid phase extraction (SPE) coupled with GC-MS and HPLC-UV analysis. J. Sep. Sci. 2007, 30, 2529–2538. [Google Scholar] [CrossRef]
  23. European Commission. Guidance Document on Pesticide Analytical Methods for Risk Assessment and Post-Approval Control and Monitoring Purposes. SANTE/2020/12830, Rev.1 24 February 2021. Available online: https://food.ec.europa.eu/system/files/2021-03/pesticides_ppp_app-proc_guide_res_mrl-guidelines-2020-12830.pdf (accessed on 5 December 2023).
  24. Parlakidis, P.; Adamidis, G.; Alexoudis, C.; Pythoglou, P.; Papadopoulos, S.; Vryzas, Z. Adjuvant Effects on Pyraclostrobin and Boscalid Residues, Systemic Movement, and Dietary Risk in Garlic under Field Conditions. Agriculture 2023, 13, 1636. [Google Scholar] [CrossRef]
  25. European Commission. EC Guidance Document on Terrestrial Ecotoxicology under Council Directive 91/414/EEC. SANCO/10329/2002 rev 2 Final; European Commission: Brussels, Belgium, 2002; p. 39. [Google Scholar]
  26. Renaud, M.; Akeju, T.; Natal-da-Luz, T.; Leston, S.; Rosa, J.; Ramos, F.; Sousa, J.P.; Azevedo-Pereira, H.M.V.S. Effects of the neonicotinoids acetamiprid and thiacloprid in their commercial formulations on soil fauna. Chemosphere 2018, 194, 85–93. [Google Scholar] [CrossRef]
  27. EFSA (European Food Safety Authority). Conclusion on the peer review of the pesticide risk assessment of the active substance benfluralin. EFSA J. 2019, 17, 5842–5848. [Google Scholar] [CrossRef]
  28. Vryzas, Z.; Alexoudis, C.; Vassiliou, G.; Galanis, K.; Papadopoulou-Mourkidou, E. Determination and aquatic risk assessment of pesticide residues in riparian drainage canals in northeastern Greece. Ecotoxicol. Environ. Saf. 2011, 74, 174–181. [Google Scholar] [CrossRef]
  29. Paramasivam, M. Dissipation kinetics, dietary and ecological risk assessment of chlorantraniliprole residue in/on tomato and soil using GC-MS. J. Food Sci. Technol. 2021, 58, 604–611. [Google Scholar] [CrossRef]
  30. Dollinger, J.; Bourdat-Deschamps, M.; Pot, V.; Serre, V.; Bernet, N.; Deslarue, G.; Montes, M.; Capowiez, L.; Michel, E. Leaching and degradation of S-Metolachlor in undisturbed soil cores amended with organic wastes. Environ. Sci. Pollut. Res. 2021, 29, 20098–20111. [Google Scholar] [CrossRef]
  31. Westra, E.; Shaner, D.; Westra, P.; Chapman, P. Dissipation and Leaching of Pyroxasulfone and S-Metolachlor. Weed Technol. 2014, 28, 72–81. [Google Scholar] [CrossRef]
  32. Abdallah, O.I. Simultaneous determination of nine dinitroaniline herbicides in environmental samples using a validated vortex-assisted dispersive liquid–liquid microextraction procedure coupled with GC–MS/MS. Chem. Pap. 2020, 74, 2311–2326. [Google Scholar] [CrossRef]
  33. Vischetti, C.; Casucci, C.; Perucci, P. Relationship between changes of soil microbial biomass content and imazamox and benfluralin degradation. Biol. Fertil. Soils 2002, 35, 13–17. [Google Scholar] [CrossRef]
  34. Niemczyk, H.D.; Krause, A.A. Behaviour and mobility of preemergent herbicides in turfgrass: A field study. J. Environ. Sci. Health B 1994, 29, 507–539. [Google Scholar] [CrossRef]
  35. Tudi, M.; Daniel Ruan, H.; Wang, L.; Lyu, J.; Sadler, R.; Connell, D.; Chu, C.; Phung, D.T. Agriculture Development, Pesticide Application and Its Impact on the Environment. Int. J. Environ. Res. Public Health 2021, 18, 1112. [Google Scholar] [CrossRef]
  36. Gustafson, D.I. Groundwater ubiquity score: A simple method for assessing pesticide leachability. Environ. Toxicol. Chem. 1989, 8, 339–357. [Google Scholar] [CrossRef]
  37. Hyun, H.N.; Jang, G.M.; Oh, S.S.; Chung, J.B. Evaluation of groundwater contamination potential of pesticides using groundwater ubiquity score in Jeju Island soils. Korean J. Pestic. Sci. 2007, 11, 144–153. [Google Scholar]
  38. Dusek, J.; Sanda, M.; Loo, B.; Ray, C. Field leaching of pesticides at five test sites in Hawaii: Study description and results. Pest. Manag. Sci. 2010, 66, 596–611. [Google Scholar] [CrossRef]
  39. Ahmad, K.S. Evaluating the Adsorption Potential of Alachlor and Its Subsequent Removal from Soils via Activated Carbon. Soil Sediment Contam. 2018, 27, 249–266. [Google Scholar] [CrossRef]
  40. Radolinski, J.; Le, H.; Hilaire, S.S.; Xia, K.; Scott, D.; Stewart, R.D. A spectrum of preferential flow alters solute mobility in soils. Sci. Rep. 2022, 12, 4261. [Google Scholar] [CrossRef]
  41. Vryzas, Z.; Papadakis, E.N.; Vassiliou, G.; Papadopoulou-Mourkidou, E. Occurrence of pesticides in transboundary aquifers of North-eastern Greece. Sci. Total Environ. 2012, 441, 41–48. [Google Scholar] [CrossRef]
  42. Vryzas, Z.; Papadakis, E.N.; Papadpoulou-Mourkidou, E. Leaching of Br-, metolachlor, alachlor, atrazine, deethylatrazine and deisopropylatrazine in clayey vadoze zone: A field scale experiment in north-east Greece. Water Res. 2012, 46, 1979–1989. [Google Scholar] [CrossRef]
  43. Milan, M.; Ferrero, A.; Fogliatto, S.; Piano, S.; Vidotto, F. Leaching of S-metolachlor, terbuthylazine, desethylterbuthylazine, mesotrione, flufenacet, isoxaflutole, and diketonitrile in field lysimeters as affected by the time elapsed between spraying and first leaching event. J. Environ. Sci. Health Part B Pest. Food Contam. Agric. Wastes 2015, 50, 851–861. [Google Scholar] [CrossRef]
  44. Alletto, L.; Benoit, P.; Bolognsi, B.; Couffignal, M.; Bergheaud, V.; Dumny, V.; Longueval, C.; Barriuso, E. Sorption and mineralization of S-metolachlor in soils from fields cultivated with different conservation tillage systems. Soil Tillage. Res. 2013, 128, 97–103. [Google Scholar] [CrossRef]
  45. Cheng, Z.; Hou, Z.; Han, H.; Yu, X.; Li, J.; Zhao, Q.; Zhang, N.; Lu, Z. Adsorption, mobility, and degradation of the pesticide propaquizafop in five agricultural soils in China. Sci. Rep. 2023, 13, 5814. [Google Scholar] [CrossRef]
  46. Torabi, E.; Wiegert, C.; Guyot, B.; Vuilleumier, S.; Imfeld, G. Dissipation of S-metolachlor and butachlor in agricultural soils and responses of bacterial communities: Insights from compound-specific isotope and biomolecular analyses. J. Environ. Sci. 2020, 92, 163–175. [Google Scholar] [CrossRef] [PubMed]
  47. Rasool, S.; Rasool, T.; Gani, K.M. A review of interactions of pesticides within various interfaces of intrinsic and organic residue amended soil environment. Adv. Chem. Eng. 2022, 11, 100301. [Google Scholar] [CrossRef]
  48. Vallée, R.; Dousset, S.; Billet, D.; Benoit, M. Sorption of selected pesticides on soils, sediment and straw from a constructed agricultural drainage ditch or pond. Environ. Sci. Pollut. Res. 2014, 21, 4895–4905. [Google Scholar] [CrossRef]
  49. Parlakidis, P.; Gounari, I.; Georgiou, A.; Adamidis, G.; Vryzas, Z.; Gikas, G.D. Removal of Two Triazole Fungicides from Agricultural Wastewater in Pilot-Scale Horizontal Subsurface Flow Constructed Wetlands. Agronomy 2023, 13, 265. [Google Scholar] [CrossRef]
  50. Gikas, G.D.; Vryzas, Z.; Tsihrintzis, V.A. S-metolachlor Herbicide Removal in Pilot-scale Horizontal Subsurface Flow Constructed Wetlands. Chem. Eng. J. 2018, 339, 108–116. [Google Scholar] [CrossRef]
  51. EFSA (European Food Safety Authority). Peer review of the pesticide risk assessment of the active substance benfluralin. EFSA J. 2022, 20, 7756. [Google Scholar]
  52. Caracciolo, A.B.; Giuliano, G.; Grenni, P.; Guzzella, L.; Pozzoni, F.; Bottoni, P.; Fava, L.; Crobe, A.; Orrù, M.; Funari, E. Degradation and leaching of the herbicides metolachlor and diuron: A case study in an area of Northern Italy. Environ. Pollut. 2005, 134, 525–534. [Google Scholar] [CrossRef]
  53. Long, Y.H.; Li, R.Y.; Wu, X.M. Degradation of S-metolachlor in soil as affected by environmental factors. J. Soil Sci. Plant Nutr. 2014, 14, 189–198. [Google Scholar] [CrossRef]
  54. Shaner, D.L.; Brunk, G.; Belles, D.; Westra, P.; Nissen, S. Soil dissipation and biological activity of metolachlor and S-metolachlor in five soils. Pest. Manag. Sci. 2006, 62, 617–623. [Google Scholar] [CrossRef]
  55. Sun, X.; Liu, M.; Meng, J.; Wang, L.; Chen, X.; Peng, S.; Rong, X.; Wang, L. Residue level, occurrence characteristics and ecological risk of pesticides in typical farmland-river interlaced area of Baiyang Lake upstream, China. Sci. Rep. 2022, 12, 12049. [Google Scholar] [CrossRef]
  56. EFSA Panel on Plant Protection Products and their Residues (PPR); Ockleford, C.; Adriaanse, P.; Berny, P.; Brock, T.; Duquesne, S.; Grilli, S.; Hernandez-Jerez, A.F.; Bennekou, S.H.; Klein, M.; et al. Scientific Opinion addressing the state of the science on risk assessment of plant protection products for in-soil organisms. EFSA J. 2017, 15, e04690. [Google Scholar] [CrossRef] [PubMed]
  57. Silva, V.; Mol, H.G.J.; Zomer, P.; Tienstra, M.; Ritsema, C.J.; Geissen, V. Pesticide residues in European agricultural soils—A hidden reality unfolded. Sci. Total Environ. 2019, 653, 1532–1545. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic layout of the experimental randomized complete block design in the field. The four different sampling depths were: 0 to 20 cm (Layer A), 20 to 40 cm (Layer B), 40 to 60 cm (Layer C), and 60 to 80 cm (Layer D).
Figure 1. Schematic layout of the experimental randomized complete block design in the field. The four different sampling depths were: 0 to 20 cm (Layer A), 20 to 40 cm (Layer B), 40 to 60 cm (Layer C), and 60 to 80 cm (Layer D).
Environments 11 00018 g001
Figure 2. GC-ECD chromatograph of (a) the spiked soil matrix with S-metolachlor and benfluralin (1 μg/g) and (b) blank soil. The approximate retention times of S-metolachlor and benfluralin were 13.43 and 10.23 min, respectively. The injection volume was 2 μL, and the injector was run at 220 °C. The temperature of the oven was set to start at 80 °C with a holding time of 2 min, then increase to 220 °C at 20 °C/min and be held for 4 min, and finally increase to 250 °C at 10 °C/min and be held for 30 min. The carrier gas was helium, flowing at a rate of 1.5 mL/min. The detector was operated at 290 °C.
Figure 2. GC-ECD chromatograph of (a) the spiked soil matrix with S-metolachlor and benfluralin (1 μg/g) and (b) blank soil. The approximate retention times of S-metolachlor and benfluralin were 13.43 and 10.23 min, respectively. The injection volume was 2 μL, and the injector was run at 220 °C. The temperature of the oven was set to start at 80 °C with a holding time of 2 min, then increase to 220 °C at 20 °C/min and be held for 4 min, and finally increase to 250 °C at 10 °C/min and be held for 30 min. The carrier gas was helium, flowing at a rate of 1.5 mL/min. The detector was operated at 290 °C.
Environments 11 00018 g002
Figure 3. The dissipation curve for S-metolachlor (mean concentration) in the top (0–20 cm) soil profile.
Figure 3. The dissipation curve for S-metolachlor (mean concentration) in the top (0–20 cm) soil profile.
Environments 11 00018 g003
Figure 4. The dissipation curve for benfluralin (mean concentration) in the top (0–20 cm) soil profile.
Figure 4. The dissipation curve for benfluralin (mean concentration) in the top (0–20 cm) soil profile.
Environments 11 00018 g004
Figure 5. Preferential flow channels created by (a) roots and earthworms and (b) cracks (personal file).
Figure 5. Preferential flow channels created by (a) roots and earthworms and (b) cracks (personal file).
Environments 11 00018 g005
Table 1. Physicochemical properties and characteristics of S-metolachlor and benfluralin [16,20].
Table 1. Physicochemical properties and characteristics of S-metolachlor and benfluralin [16,20].
ParameterS-MetolachlorBenfluralin
Molecular formulaC15H22ClNO2C13H16F3N3O4
Substance groupChloroacetamideDinitroaniline
Mode of actionCell division inhibitorMicrotubule assembly inhibitor
Structural formulaEnvironments 11 00018 i001Environments 11 00018 i002
Molecular weight (g/mol)283.79387.82
Dissociation constant (pKa) at 25 °CNo dissociation−0.59
Vapour pressure at 20 °C (mPa)3.71.8
Water solubility at 20 °C (mg/L)4800.064
Henry’s law constant at 25 °C (Pa m3/mol)2.20 × 10−39.5
Octanol–water partition coefficient (25 °C) LogKow3.055.19
Adsorption coefficient Kfoc (mL/g)200.210,777
Soil degradation DT50field 123.1753
1 Half-life for field studies.
Table 2. Ecotoxicology (NOEC and NSDE in μg/g) of herbicides for E. fetida, E. crypticus, F. candida, H. aculeifer, and N and C mineralization organisms.
Table 2. Ecotoxicology (NOEC and NSDE in μg/g) of herbicides for E. fetida, E. crypticus, F. candida, H. aculeifer, and N and C mineralization organisms.
PesticideE. fetida. NOEC 1F. candida NOECE. crypticus NOECH. aculeifer. NOECNSDE 2 for N|C Mineralization MicroorganismsCritical ConcentrationPNEC 3 AF 4PNECs 5
S-metolachlor [16]26.65na 6nanana26.651000.27
Benfluralin [27]30.85.5na1000115.5100.55
1 No-observed-effect concentration. 2 No significant adverse effects. 3 Predicted no-effect concentration. 4 Assessment factor. 5 Lowest long-term PNECs of the most susceptible species/AF. 6 Information not available.
Table 3. Analytical method parameters for S-metolachlor and benfluralin.
Table 3. Analytical method parameters for S-metolachlor and benfluralin.
Analytical Method ParameterS-MetolachlorBenfluralin
Calibration curve in solventy = 306918x + 71.3 (R2 = 99.5)y = 554044 + 18304.8 (R2 = 99.7)
Calibration curve in blank soily = 253821x + 380.1 (R2 = 99.3)y = 431600 + 3753.9 (R2 = 99.1)
Matrix effect (%)−17.3−22.1
LOD (μg/g)0.010.015
LOQ (μg/g)0.10.1
LOD = limit of detection; LOQ = limit of quantitation.
Table 4. Recovery rates and RSDs for S-metolachlor and benfluralin in soil.
Table 4. Recovery rates and RSDs for S-metolachlor and benfluralin in soil.
MatrixSpiked Level (μg/g)Average Recovery (%)RSD (%) (n = 5)
S-MetolachlorBenflulralinS-MetolachlorBenflulralinS-MetolachlorBenflulralin
Soil0.181888.59.7
Soil0.590856.68.7
Soil1.0971017.99.1
Table 5. Comparison of detection methods for S-metolachlor and benfluralin in soil.
Table 5. Comparison of detection methods for S-metolachlor and benfluralin in soil.
HerbicideMethodLOQ 1 (μg/g)Retention TimeReference
S-metolachlorGC-MS/MS 2 (scan mode)na 107.31[13]
QuEChERS 3 with online SPE-UHPLC 4-MS–MS (MRM 5)nana[30]
GC-MS (scan mode)nana[31]
BenfluralinDLLME 6 with GC-MS/MS (SRM 7 mode)28.1[32]
GC/NPD 80.514[33]
GLC 9/NPDnana[34]
1 Limit of quantification. 2 Mass spectrometer. 3 Quick, easy, cheap, effective, rugged, safe. 4 Ultra-high-performance liquid chromatograph. 5 Multiple reaction monitoring. 6 Dispersive liquid–liquid microextraction. 7 Elected reaction monitoring, 8 nitrogen–phosphorus detector. 9 Gas–liquid chromatograph. 10 Information not available.
Table 6. Residues and dissipation parameters for S-metolachlor in soil at 1350 g a.i./ha in the top 0–20 cm soil profile.
Table 6. Residues and dissipation parameters for S-metolachlor in soil at 1350 g a.i./ha in the top 0–20 cm soil profile.
Days after ApplicationResidue (µg/g) 1Dissipation (%)
0 (2 h)1.003-
70.66333.90
150.44355.83
300.31768.39
450.22377.77
600.13087.04
k (rate constant)0.032-
R20.946-
DT50 (days) 221.66-
1 Mean concentration. 2 Half-life.
Table 7. Residues and dissipation parameters for benfluralin in soil at 1920 g a.i./ha in the top 0–20 cm soil profile.
Table 7. Residues and dissipation parameters for benfluralin in soil at 1920 g a.i./ha in the top 0–20 cm soil profile.
Days after ApplicationResidue (µg/g) 1Dissipation (%)
0 (2 h)2.640-
72.08321.10
151.62038.64
301.23653.18
450.88666.44
600.61076.89
k (rate constant)0.023-
R20.963-
DT50 (days) 230.13-
1 Mean concentration. 2 Half-life.
Table 8. Mobility of S-metolachlor in the 0–80 cm soil profile.
Table 8. Mobility of S-metolachlor in the 0–80 cm soil profile.
Days after ApplicationS-Metolachlor (µg/g)
0–20 cm20–40 cm40–60 cm60–80 cm
0 (2 h)1.003 1Nd 2ndnd
70.6630.2020.127nd
150.4430.185ndnd
300.3170.157ndnd
450.2230.104ndnd
600.130ndndnd
1 Mean of four replicates. 2 Not detectable.
Table 9. Mobility of benfluralin in the 0–80 cm soil profile.
Table 9. Mobility of benfluralin in the 0–80 cm soil profile.
Days after ApplicationBenfluralin (µg/g)
0–20 cm20–40 cm40–60 cm60–80 cm
0 (2 h)2.640 1nd 2ndnd
72.0830.2300.1660.112
151.6200.1870.129nd
301.2360.2010.134nd
450.8860.2300.111nd
600.6100.121ndnd
1 Mean of four replicates. 2 Not detectable.
Table 10. Toxicity–exposure ratios (TERws and TERds) for soil organisms at a 0–20 cm depth.
Table 10. Toxicity–exposure ratios (TERws and TERds) for soil organisms at a 0–20 cm depth.
PesticideMPCws 1MPCds 2E. fetidaF. candidaE. crypticusH. aculeifer
TERwsTERdsTERwsTERdsTERwsΤERdsTERwsΤERds
S-metolachlor1.00.1326.5725na 3nanananana
Benfluralin2.640.6111.6650.492.089.01naan378.781639.34
1 Mean pesticide concentration for the worst-case scenario (μg/g). 2 Mean pesticide concentration for the dissipation scenario (μg/g). 3 Not applicable.
Table 11. Risk quotients for the most sensitive species at a 0–20 cm depth.
Table 11. Risk quotients for the most sensitive species at a 0–20 cm depth.
PesticideMPCws 1MPCds 2RQwsRQds
S-metolachlor1.000.133,760.48
Benfluralin2.640.614.801.11
1 Mean pesticide concentration for the worst-case scenario (μg/g). 2 Mean pesticide concentration for the dissipation scenario (μg/g).
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

Parlakidis, P.; Adamidis, G.S.; Gikas, G.D.; Vasiliou, S.; Kissa, M.; Doitsinis, K.; Alexoudis, C.; Vryzas, Z. Dissipation Kinetics, Leaching, and Ecological Risk Assessment of S-Metolachlor and Benfluralin Residues in Soil. Environments 2024, 11, 18. https://doi.org/10.3390/environments11010018

AMA Style

Parlakidis P, Adamidis GS, Gikas GD, Vasiliou S, Kissa M, Doitsinis K, Alexoudis C, Vryzas Z. Dissipation Kinetics, Leaching, and Ecological Risk Assessment of S-Metolachlor and Benfluralin Residues in Soil. Environments. 2024; 11(1):18. https://doi.org/10.3390/environments11010018

Chicago/Turabian Style

Parlakidis, Paraskevas, George S. Adamidis, Georgios D. Gikas, Sofia Vasiliou, Melpomeni Kissa, Konstantinos Doitsinis, Christos Alexoudis, and Zisis Vryzas. 2024. "Dissipation Kinetics, Leaching, and Ecological Risk Assessment of S-Metolachlor and Benfluralin Residues in Soil" Environments 11, no. 1: 18. https://doi.org/10.3390/environments11010018

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