Next Article in Journal
Research on Spatial and Temporal Evolution Trends and Driving Factors of Green Residences in China Based on Weighted Standard Deviational Ellipse and Panel Tobit Model
Previous Article in Journal
The Standardization Method and Application of the BIM Model for Interchanges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Glyphosate-Based Herbicide and Aminomethylphosphonic Acid (AMPA): Environmental and Health Impacts

by
Babatunde Solomon Ojelade
1,*,
Olatunde Samod Durowoju
2,
Peter Oluremi Adesoye
3,
Stuart W. Gibb
4 and
Georges-Ivo Ekosse
5
1
Department of Geography and Environmental Science, University of Venda, Thohoyandou 0950, South Africa
2
Department of Earth Sciences, University of Venda, Thohoyandou 0950, South Africa
3
Department of Forestry, University of Venda, Thohoyandou 0950, South Africa
4
Environmental Research Institute, North Highland College, University of Highlands and Islands, Inverness IV3 5SQ, UK
5
Directorate of Research and Innovation, University of Venda, Thohoyandou 0950, South Africa
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(17), 8789; https://doi.org/10.3390/app12178789
Submission received: 4 June 2022 / Revised: 22 July 2022 / Accepted: 7 August 2022 / Published: 1 September 2022
(This article belongs to the Special Issue Sustainable Environmental Remediation Technologies)

Abstract

:
The use of synthetic molecules to achieve specific goals is steadily increasing in the environment, and these molecules adversely impact human health and ecosystem services. Considering the adverse effects, a better understanding of how these molecules behave in the environment and their associated risks is necessary to keep their use acceptably limited. To meet the demands of farmers and combat weed problems, woodlands and farmlands are sprayed with agrochemicals, primarily glyphosate-based herbicides. Farmers increasingly embrace these herbicides containing glyphosate. Glyphosate and aminomethylphosphonic acid (AMPA), a key metabolite of glyphosate, have been reported as toxicological concerns when they become more prevalent in the food chain. The chemical glyphosate has been linked to various health issues in humans and other living organisms, including endocrine disruption, reproductive issues, tumours, non-Hodgkin lymphomas, and liver, heart, and blood problems. Therefore, the current review aims to compile data on glyphosate-based herbicide use in the environment, potential risks to human and ecological health, and various maximum residual limits for crops as suggested by international organizations. As a result, regulatory agencies can advise glyphosate users on safe usage practices and synthesize herbicides more efficiently.

1. Introduction

Increasing numbers of synthetic molecules are being released into the environment to achieve specific outcomes [1]. Those molecules may adversely affect human health and ecosystem services [1,2]. An in-depth understanding of how those molecules behave in the natural world combined with an estimate of their complexity can help regulate their use and enable users to take precautionary measures to protect human health [3]. Although regulations have set the highest points for known pollutants found in water supplies or drinking water [4], and food [5], there are none for soil residue [3]. Indeed, the United Nations’ Food and Agriculture Organization recently released a report that exposes the unseen truth of soil degradation [6]. Agrochemicals, primarily herbicides, are used on agricultural lands to suit farmers’ demands and overcome weed resistance [7]. Formulation, approval, use, and monitoring of these herbicides, especially glyphosate-based herbicides (GBHs) involve numerous stakeholders [3].
Following World War II, food scarcity was a problem around the world. As a result, today farmers across the globe use several herbicides that are synthetic to manage pests and weeds. However, the formulation of glyphosate has been considered as the most important herbicide in that area [8]. Glyphosate-based herbicides (GBHs) come in a variety of commercial formulations, including broad-spectrum, non-selective, post-emergent, and synthetic herbicides [9]. The first glyphosate was developed in 1950, and its herbicidal properties were only discovered in 1970 when GBHs were resynthesized and tested [10]. The herbicide Roundup contains ‘GLY’ as an active component, which was introduced and commercialized by Monsanto Corporation in 1974 [11]. Agricultural weed control with glyphosate quickly became popular with farmers, who gained the ability to eliminate weeds without causing crop damage [8]. According to Zhang et al. [12], the glyphosate hinders the synthesis of amino acids including tyrosine, phenylalanine, and tryptophan, thereby killing weeds without destroying the agricultural crops. Its use has progressively increased in non-agricultural and agricultural settings, and it is now the most common herbicide worldwide [13,14]. In addition, many agrochemical companies market GLY formulations in various strengths and with various adjuvants, because they have already been reviewed and registered by regulatory organizations [11,15]. It was decided by the European Union Council on 27 November 2017 that glyphosate would be permitted to be used for five more years with the majority of 18 member nations voting in favor of allowing the use [16]. As a result, GLY can be used as a component of plant protection products (PPPs) through the end of 2022 [17]. Agencies in Europe, such as the European Food Safety Authority and the European Chemicals Agency, conducted a thorough assessment of GLY in recent years based on concerns over its environmental and human health consequences [18]. Based on current scientific data, the European Chemicals Agency (ECHA) concluded that glyphosate does not meet the criteria for a carcinogenicity hazard classification, and cannot be categorized as a carcinogen, reproductively harmful substance or mutagen [16,19]. As part of EFSA’s risk-assessment process, the scientific committee has been asked to create guidelines on how to describe, record, and justify uncertainty [20]. In order to continue the renewal as an element in PPPs, GLY must not have a negative impact on the environment or human or animal health, according to European laws [21].
Glyphosate’s widespread use stems from its effectiveness in weed control at a reasonable price, its presumed low toxicity, fast uptake by plants, and gradual weed resistance development to glyphosate [22]. Due to the accumulation of residues in the food chain, glyphosate and aminomethylphosphonic acid (AMPA), one of its main metabolites, are reported to be toxicologically problematic [23]. Several environmental conditions can affect the degradation of glyphosate, depending on its structural affinity with certain transformations [24]. Despite AMPA’s longer half-life in soil, with 23–958 days compared to glyphosate’s (1–197 days), most studies examined only glyphosate [25,26,27,28]. In the long term, contaminants with a long half-life and slow degradation can damage the environment [29]. This occurs most frequently in agricultural and forestry settings where repeated applications are common. Due to its extremely low vapour pressure, it cannot be volatilized significantly even if it undergoes mineralization, immobilization, or leaching once applied [29,30]. As a result of glyphosate mineralization, AMPA, methyl phosphonic acid, sarcosine and glycine are produced [30,31]. After that, AMPA is mineralized to methylamine and phosphate, which when decomposed to produce carbon dioxide and ammonium [29,32]. In research, the presence of glyphosate in the environment was established. As a result, potential health risks associated with glyphosate must be assessed.
The present review is therefore aimed at consolidating information on glyphosate-based herbicides in the environment, the potential threat this type of herbicide poses to ecological and human health, and various maximum residual limits (MRLs) proposed by international agencies on agricultural crops. Consequently, regulatory organizations and other authorities can provide glyphosate users with the necessary precautions and guidelines for future usage and to more effectively formulate herbicides by using safer surfactants.

2. Factors Affecting the Degradation Mechanisms of Glyphosate Pollution

Factors and mechanisms of degrading glyphosate pollution in soil include mineralization, immobilization and leaching having physical and chemical properties that influence the mechanism.

2.1. Mineralization

In some circumstances, biochemical properties of a soil can result in glyphosate and AMPA mineralization occurring very quickly [30]. Increased phosphate content, soil pH, and low Fe and Cu content accelerate glyphosate mineralization, driven primarily by increased microbial mineralization [30,33,34]. Adsorption of glyphosate to organic carbon (C) may provide environmental benefits such as delaying leaching, promoting soil degradation, and slowing the release of the herbicide. If the glyphosate use continues, the organic C system may eventually become saturated. Thus, soil biochemical properties, microbial diversity and activities are all factors in glyphosate degradation [23].

2.2. Immobilization and Leaching

The high adsorption of glyphosate results in its rapid immobilization in most natural situations after application [26,35]. Influential factors in the immobilization of glyphosate include minerals, soil organic matter, and clay. It has also been reported by Shushkova et al. [36] that adsorption to soil occurs within 3 h of the application when about 20% of glyphosate quantity is initially applied. High levels of clay, organic matter, iron, and aluminium are required for high adsorption, soils with low pH and phosphate concentrations, and high levels of clay [35,36,37,38]. Contrarily, soils with high levels of phosphate, high pH, and low levels of organic matter, Fe, and Al are more prone to glyphosate and AMPA losses because of a reduced capacity for adsorption and a larger propensity for leaching [36,38]. According to Bai and Ogbourne [23], the leaching of glyphosate and resulting contamination of water sources is increasingly due to the recurrent finding of glyphosate and AMPA residues in the water.

3. Environmental Hazards Posed by Glyphosate and AMPA Residues

A wide range of environmental risks has been created due to the relatively persistence of GLY and AMPA in the environment. There is not much information regarding the toxicity, health, or safety of glyphosate and AMPA on frequent and prolonged exposure, so it is difficult to predict their consequence and magnitude. Several issues surround these compounds’ ecotoxicological and toxicological assessments, which may contribute heavily to the toxicological properties of formulated herbicides [39,40]. Because it is extremely difficult to assess safe, marketable products (as these products have properties that are only known to their manufacturers and are partially unknowable to regulatory agencies and research scientists), it can be suggested that all herbicide formulation ingredients must be declared and regulated. As a result, this review focuses on the existence of glyphosate and AMPA residues in soil and water bodies and the risks to human and animal health.
Roundup formulations are among the most extensively utilized GBH products that consist of other surfactants and chemical adjuvants. The active ingredients in Roundup are IPA-salt, polyethoxylated tallow amine (POEA) and other constituents [41]. These adjuvants can sometimes be even more toxic than glyphosate [42,43]. A thorough examination of surfactant co-formulants in glyphosate-based herbicides is urgently needed. There are several classes of POEA molecules with common structural characteristics [44]. Over many decades, ethoxylated amines, also known as POEA, have been the most common surfactants used in GBH formulations [45]. For instance, according to Guilherme et al. [46], in a study on the Roundup Ultra formulation, POEA was detected at a concentration of 16%.
Nonetheless, the labeled Roundup Ultra in Portugal (MON 52276) contains neither POEA nor propoxylated quaternary ammonium surfactants [47]. There are many instances in which authors cite the brand without citing the source country of the formulation. The co-formulants found in a formulation called Roundup Ultra vary depending on the country of sale. For example, Roundup Ultra is sold under the MON 76473 label in Ireland, whereas it has the MON 52256 label in Germany, the MON 79351 label in Greece, and the MON 77360 label in the United States [47]. It is not surprising that the same assay used to test the same GBH brand yields different results in different laboratories worldwide because it is difficult to identify substances across studies.
Mesnage et al. [48] confirmed that formulated herbicides are possibly more hazardous than the active substances alone, as evidenced by studies using glyphosate-based herbicides including a variety of other active components. It has become apparent that glyphosate has a wide range of harmful consequences [49], and co-formulants in Roundup have endocrine-disrupting effects in human cells [50]. As a result of these impacts, agrochemicals like GBHs will affect agricultural products and the environment, notably as chemical residues in goods produced by agriculture and as adverse effects on nontarget organisms [51]. Various studies on glyphosate and AMPA in several countries is described by Gillezeau et al. [52] in Table 1 given below. Even though the values of GLY and AMPA (in the table) may or may not be harmful, accumulating them over time will result in various health problems.
Table 1. Description of glyphosate and AMPA studies in several countries.
Table 1. Description of glyphosate and AMPA studies in several countries.
CountrySubjectsType of SampleYear of SamplingLab MethodsGlyphosate
LOD
Effects of GlyphosateAMPA
LOD
Effects of AMPAReferences
EXPOSURE IN THE WORKPLACE
United States (South Carolina, Minnesota)Forty-eight farm families (farmers and wives, with 79 children aged between 4 and 18) on the day of application and three days later.UrineUnreportedHPLC1 µg per litreOn application day, farmers’ geometric mean ± SD: 3.2 ± 6.4 μg/L (range < 1–233); while on day 3, 1.0 ± 3.6 (<1–68) μg per litre. Less than 25% of the wives or kids displayed values that could be identified.UnreportedUnreported[53]
FinlandFive forest workers sprayed 8% Roundup solution 6 h a day for one weekUrine1988GC with a 63 Ni-electron capture detector100 µg per litreSamples of urine remained below LOD for glyphosate. A subsequent quantified urine sample contained 85 µg per litre of glyphosate.50 µg per litreFor AMPA, urine samples remained below the limit of detection[54]
FranceHerbicides based on glyphosate are used by a farmer and his family (5 in the household)UrineUnreportedLC-MS1 µg per litreA 9.5 µg per litre concentration was found on the farmer following spraying and about 2 µg per litre after 2 days; After 2 days, 2 µg per litre was also measured in one child. No measurable levels were present in the mother or the other two kids.UnreportedUnreported[55]
IrelandAmenity horticulturalists (17 males and 1 female), prior to and following spraying.Urine2015LC MS-MS0.5 µg per litrePrior to spraying, mean ± SD: 0.71 ± 0.92.
After spraying: 1.35 ± 2.18 μg per litre
UnreportedUnreported[56]
Mexico76 farmersUrineUnreportedELISA0.05 µg per litre (in water)In farming areas (Mean ± SD): 0.26 ± 0.23 μg per litre, (median: 0.28)UnreportedUnreported[57]
Sri LankaTen farmers in good health from regions where chronic renal disorders are endemicUrineUnreportedELISA0.6 µg per litreRanged between 40.2–>80 μg per litre, (Median: 73.5)UnreportedUnreported[58]
GENERAL POPULATION WITHOUT DIRECT CONTACT
Sri LankaTen healthy non-farmers from regions without a long-standing endemic kidney illness.UrineUnreportedELISA0.6 μg per litreRanged between 1.2–5.5 μg per litre, (Median: 3.3)UnreportedUnreported[58]
US (Iowa)Households who do not farm (23 fathers, 24 mothers, 51 children)Urine2001FCMIA0.9 μg per litreFor the non-farm fathers, djusted geometric mean was 1.5 μg per litreUnreportedUnreported[59]
US (Washington and Idaho)A total of 41 lactating women of greater than 18 years oldMilk (41), Urine (40)2014–2015LC-MS1 μg per litre in Milk;
0.02 μg per litre in Urine
For milk, glyphosate is below LOD.
For urine, the glyphosate mean is 0.28 ± 0.38 μg per litre. Glyphosate detectable in 37/40 of the urine samples. There are no statistically significant differences between consuming conventional or organic food or living in an urban or suburban region.
In milk: 1 μg per litre; while 0.03 μg per litre was detected in UrineIn milk, AMPA is below the LOD.
In urine: AMPA mean is 0.30 ± 0.33 μg per litre
[60]
CanadaSimilar in age and BMI of pregnant (30) and non-pregnant women (39), and 30 umbilical cordsMaternal and umbilical cord serumunreportedGC-MS15 μg per litreNo glyphosate found in the umbilical cord or in pregnant women. Mean of glyphosate found in non-pregnant women is 73.6 ± 28.2 μg per litre10 μg per litreIn none of the samples was AMPA found.[61]
US (Indiana)Pregnant women (71) between the ages of 18 and 39Urine and drinking water2015-2016LC MS-MSIn urine: 0.1 μg per litre; while in water: 0.2 μg per litre.Glyphosate found in the urine: mean (SD) 3.40 (±1.24) μg per litre. No glyphosate was found in the drinking water.UnreportedUnreported[62]
IrelandFifty Irish persons over the age of 18 who do not have any special dietary preferences, with no pesticide usage in their line of worksUrine2017LC-MS-MS0.5 μg per litre.A total of 47 samples were examined, and their urinary creatinine levels ranged from 3.0 to 30 nmol/L. 20% of the samples had Glyphosate levels above LOD. Glyphosate levels in samples with medians above the LOD (Range): 0.87 (0.80–1.35) μg per litre.UnreportedUnreported[63]
DenmarkA total of 13 mothers and 14 children (6–11 years old) in rural and urban communitiesUrine2011-2012ELISA2.5 ppbFor children, the mean was 1.96 (range: 0.85–3.31) μg per litre. For mothers, the mean was 1.28 (range: 0.49–3.22) μg per litreUnreportedUnreported[64]
Germany399 individuals aged 20 – 29 yearsUrine2001-2015GC-MS-MSLOQ: 0.1 μg per litreA total of 31.8% of the samples (127 samples) were found to have glyphosate level above LOD. The highest levels were found in males.LOQ: 0.1 μg per litreAMPA: 160 (40.1%) > LOD.[65]
18 European countries182 volunteersUrine2013GC-MS-MSLOQ: 0.15 μg per litre44% of the samples (of about 80 samples) were found to have glyphosate level above LOQ; Latvia had the highest glyphosate concentration, which is 1.8 μg per litre.LOQ: 0.15 μg per litre36% > LOQ AMPA;) highest AMPA concentration: 2.6 μg per litre (Croatia)[66]
ColombiaA total of 112 people who live in locations where glyphosate was applied aeriallyUrine2006GC with a detector for electron micro-capture0.5 μg per litreFor the glyphosate (mean ± SD): 7.6 ± 18.6 μg per litre, ranged from 0–130 μg per litre. There were quantifiable AMPA levels in 4/42 participants with quantifiable glyphosate levels:
Mean glyphosate: 58.8 μg per litre (range: 28–130 μg per litre).
1.0 μg per litreAMPA: 1.6 to 8.4 μg per litre (range: 0–56 μg per litre)[67]
ThailandA total of 82 expectant women, aged 19 to 35, who gave birth in a participating hospitalUmbilical cord and maternal serum2011HPLC0.4 μg per litreMedian for maternal serum: 17.5 μg per litre (range 0.2–189.1), while for the umbilical cord serum: 0.2 μg per litre (range 0.2–94.9). 50.7% of the umbilical cord serum samples are below LOD, 46.3% maternal serum samples are below LOD.UUnreportedUnreported[68]
Note: AMPA, aminomethylphosphonic acid; ELISA, enzyme-linked immunosorbent assay; FCMIA, fluorescence covalent microbead immunoassay; GC, gas chromatography; HPLC, high-performance liquid chromatography; LC, liquid chromatography; LOD, limit of detection; LOQ, limit of quantification; MS, mass spectrometry; MS/ MS tandem, mass spectrometry.

3.1. Soil with Glyphosate

Given the widespread usage of glyphosate, understanding how it interacts with the soil ecosystem is critical for environmental safety assessment and practical application. Despite not being sprayed directly on the ground, glyphosate-based herbicides can contaminate the soil in and surrounding treated areas due to spray drift during application and leaf surfaces that have been washed away by rain. Mineralization, degradation, immobilization, and leaching are all factors in glyphosate’s fate in soil. Several kinds of research have been published in recent years, attempting to discover and comprehend the processes that determine how chemicals behave in the environment and produce pollution, particularly in soil and water [12]. Mesnage et al. [69] analyzed the most prevalent surfactants used as co-formulants in glyphosate-based herbicides. They looked at how adding surfactants (such as Triton CG-110) would impact the soil’s glyphosate adsorption, mineralization and leaching processes.
Soil composition, physicochemical factors, biological properties, chemical properties of the individual pesticide, and timing of precipitation and pesticide application all influence the fate of glyphosate [25,26,70,71,72]. Total organic carbon, pH, and temperature fluctuations in German soil were recently discovered to affect glyphosate mineralization kinetics, level of bio-NER formation, and the amount of recoverable glyphosate over time [73].
In most soils, glyphosate degrades quite quickly, with an estimated half-life of 7 to 130 days on-site [74]. Because glyphosate degrades relatively quickly, it has a low impact on the environment, particularly water and soil resources. On the other hand, its metabolites, AMPA and/or sarcosine, may boost the pollution risk. According to Grandcoin et al. [14], the herbicide molecule can be degraded in two ways (Figure 1). The first technique relies on the dissolution of the carbon-nitrogen bond, which forms AMPA (glyphosate’s main metabolite) via the enzyme glyphosate oxidoreductase, which is broken down to carbon dioxide. In contrast, the second process relies on the broken carbon–phosphorus (C–P) bond, which is accomplished through the C–P lyase enzyme and results in the synthesis of glycine and sarcosine [75,76,77]. However, as an aminopolyphosphonate photodegradation product in water, AMPA can also be found in the natural world [78].
Glyphosate is a three-polar functional group amphoteric chemical (amine, carboxymethyl, and phosphonomethyl) in its structure, linearly arranged. It is an ionic compound that is highly polar and soluble in water (10.5 g L1 at 20 °C) due to the existence in its structure of those groups [79]. Even when glyphosate dissolves in water, it can sometimes attach to soil particles, especially in clays. It has been found in several laboratory experiments that the molecule has a high absorption constant in soil, ranging between 8 and 377 dm3/kg. The characteristics of the soil, temperature, and soil moisture all influence glyphosate adsorption and subsequent release from the soil. As a result, it may wash away rapidly in clay-rich soils for more than a year or in sandy soils. Even when attached to soil particles, in the presence of phosphates, in some cases, it can be broken down into soil water. Additionally, glyphosate can form compounds with metal ions, which could alter the soil’s nutritional availability [3]. Because glyphosate adsorbs to soil particles suspended in the runoff, it risks surface water contamination due to erosion [80,81].
Concerns regarding the existence of environmental levels of GLY and AMPA have grown as the use of herbicides containing glyphosate has grown. Argentina [82,83,84,85], Canada [86], the United States [87], Mexico [88], and Portugal [89], as well as Spain [90], New Zealand [91], Austria [92], and France [93], have received a lot of attention.
Even though GLY is the most widely used herbicide, an investigation into the presence and amounts of glyphosate residues in soils and analytical methods for this purpose is relatively limited, especially given the scale of its use [94,95,96]. Furthermore, some polluted soils were found in locations prone to water and wind erosion [97]. Therefore, it is necessary to establish soil residue threshold values to identify potential risks to soil health, as well as off-site consequences from wind and water erosion [3].
Studies on soil microbial diversity and composition do not necessarily support the notion that glyphosate and AMPA are non-toxic to soil microorganisms [98,99]. For example, studies indicated earthworms as an essential biomarker for soil health, and following glyphosate application to the soils, the soils’ biomass was reduced [100,101]. According to García-Pérez et al. [101], soil earthworm biomass was considerably lower in coffee plantations subjected to continuous glyphosate spraying for 22 years compared to those not treated in the previous seven years. Other research found that glyphosate had no direct effect on earthworms [102,103,104]. In contrast, others found that although earthworms may survive after glyphosate treatment, it can disrupt cocoon hatching, resulting in lower earthworm numbers in the soil [105,106]. According to another study, sub-lethal glyphosate spraying can alter soil chemistry, affecting water quality and other soil dwellers [107]. Moreover, due to glyphosate application, a change in the makeup of the soil community was seen, altering the availability of soil nutrients and nutrient balance [108]. However, there have been studies that disagree on whether or not applying glyphosate or glyphosate resistance species can cause nutritional imbalances [109].

3.2. Water with Glyphosate

A metabolite of glyphosate, AMPA, and its residues are increasingly discovered in water sources, with runoff being one source of water contamination [110]. Glyphosate concentrations of more than 400 g/L harm some aquatic animals, including amphibians and fish [111,112]. According to Mercurio et al. [113], glyphosate has been reported in the marine ecosystem, and its persistence in saltwater is now being investigated. Table 2 showed the occurrence and concentrations of glyphosate in various water bodies across several countries in America and Europe. Although, they proved to be safe according to their respective guidelines, persistent exposure to glyphosate can pose a health threat.
Regarding risks posed to human health, the maximum concentration level (MCL) of glyphosate in the United States of America [23] and Australia is 700 µg L−1 and 1000 µg L−1 respectively. In Europe, glyphosate concentrations in drinking water are permitted to be less than 0.1 g/L, but 77 g/L are considered tolerable, according to reports by Horth and Blackmore [114]. According to European criteria, glyphosate residue in human drinking water must be reduced; however, glyphosate water treatment is expensive. Although these remedies have little influence on the presence of glyphosate in the water supply, the long-term impacts of glyphosate remain a worry [23]. Saunders and Pezeshki [81] urged that correct management measures, such as lower application rates and vegetation buffers, be used to limit glyphosate’s eco-toxicity hazards.

3.3. Glyphosate in Nontarget Plant Species

In spite of the specified waiting period in harvested crops, glyphosate and AMPA residues are observed in unintended plant species after weed spraying [23]. Glyphosate residues in tree foliage that are unusually high (e.g., 1000 mg/kg) may be attributable to direct absorption into tree leaves due to airborne herbicide drift contamination [115]. In addition to the possible health problems associated with food contamination, glyphosate exposure can have phytotoxic effects. Reduced absorption of vital nutrients is one way phytotoxicity affects plant performance [116], nutritional imbalances, reduced yield, and poor food quality [117,118]. Various studies have reported that about 50% of plant biomass being reduced following glyphosate contamination in some nontarget plant species [116,119]. Following the application of GBH to crops, residual GLY and AMPA may remain in harvested crops and processed foods [120]. According to testing conducted by the UK Food Standard Agency, 27 out of 109 samples of bread had glyphosate residues of at least 0.2 mg/kg. The US Department of Agriculture Tests in 2011 revealed that 90.3% of 300 samples of soybeans contained glyphosate and 95.7% of which included AMPA with concentrations of 1.9 and 2.3 ppm, respectively [120].
Consumers are exposed to more glyphosate residues through their food, so this exposure should also be considered [51]. By drifting, leaching, and surface runoff, biologically active herbicide interacts with biomass and is absorbed by soil and water [121]. Among other places, glyphosate contamination is found in human urine, animal urine, ground water, and human milk and meat from farm animals [32,122,123,124,125]. Therefore, interactions with other stressors should be investigated in a more realistic situation when interacting with biological systems or the environment [126,127].
Table 2. Investigation of the presence and levels of glyphosate in surface and ground water samples collected from several countries in Europe, South America, and North America.
Table 2. Investigation of the presence and levels of glyphosate in surface and ground water samples collected from several countries in Europe, South America, and North America.
CountryYearPresence of Glyphosate/ConcentrationReferences
United State
  (Midwest)2002Of the sampled streams, 36% were positive/about 8.7 g/L.[128]
  (Midwest)2013Of the sampled streams, 44% were positive/about 27.8 µg/L.[129]
  (Washington, Maryland, Iowa, Wyoming)2005–2006Positive outcomes in all of the sampled streams/about 328 µg/L.[130]
  (Iowa, Indiana, Mississippi)2004–2008Positive results for most of the sampled rivers/about 430 µg/L after a storm.[131]
Spain2007–2010Of the groundwater samples, 41% were positive/about 2.5 µg/L.[90]
Canada2002Of the samples, 22% were positive/about 6.07 µg/L.[132]
Argentina2012Of the surface water samples, 35% were positive/0.1–7.6 µg/L.[133]
Switzerland2016Positive results for most of the stream samples/about 2.1 µg/L.[134]
France2003–2004Positive results from 91% of stream samples/up to 165 µg/L.[135]
Hungary2010–2011Positive results for most rivers and groundwater samples/about 0.001 µg/L.[136]
Denmark1999–2009Of the surface water samples, 25% were positive/about 31 µg/L.
Of the groundwater samples, 4% were positive/up to 0.67 µg/L.
[137]
Mexico2015The groundwater samples were all positive/about 1.42 µg/L.[57]
Germany1998Only a few positive samples in two tributaries of the Ruhr River/with concentrations up to 0.59 µg/L.[138]

4. Toxicological Effects of Glyphosate and AMPA

Glyphosate inhibits the route of shikimic acid in weeds by blocking the synthase of enolpyruvylshikimic phosphate (EPSP), which prevents aromatic amino acid production, including tryptophan, tyrosine, and phenylalanine [139]. Herbicide exposure results in green colouration disappearance, leaf wrinkling or deformation, stunted growth, and tissue damage and the plant will eventually die after 7–21 days [81,139]. Plants, fungi, and some microorganisms are the only species that synthesize aromatic amino acids through shikimic acid [81]. Animals do not synthesize shikimic acid, so it is required to supplement their diet with aromatic amino acids. As a result, the low toxicity of glyphosate in animals is due to the absence of this pathway. Despite this, some adverse effects are associated with exposure to high doses for extended periods [32,81,139,140]. Due to its minimal acute toxicity, glyphosate appears to have little or no impact on microorganism populations and processes [98,141]. However, as shown in Table 3 below, the harmful effects of glyphosate-based formulations on aquatic and terrestrial nontarget creatures were found to be distinct [42,142,143,144,145]. The observed toxicity of Roundup formulations can be attributed to surfactants, including POEA [31,139,145,146,147]. Acute toxicity could not be as serious as reproductive, chronic, and sub-chronic harm, according to new findings on glyphosate contamination in the environment [23]. Figure 2 illustrates various pathways of glyphosate formulations used in agricultural or non-agricultural settings, liable for causing potential environmental and human health risks.

4.1. Acute Toxicity

In spite of the most severe cases of glyphosate poisoning (125 µg kg−1 day−1) and AMPA poisoning (5 µg kg−1 day−1) documented in adult humans, it was concluded that these substances do not pose a risk to humans [149]. Approximately 3.2% of patients died of acute poisoning, and the pathophysiology of these deaths remains unclear. This happened after evolving cardiorespiratory toxicity over a period of many hours and no proof that increased accessibility to intensive care units or laboratory services would have made a difference in these outcomes [150]. The use of various glyphosate formulations could lead to variations in reported cases. Surfactants are typically used in commercial glyphosate formulations to help the active component penetrate more easily and effectively. It is worth noting that neat glyphosate exhibits the lowest in vitro toxicity (approx. 2 g L−1) while Roundup 450 and 400 exhibit the highest levels (approx. 0.001 g L−1) [151].

4.2. Toxicity, Both Chronic and Subchronic

A dose of 560 mg kg−1 day−1 is recommended as the highest dose for male rats and 671 mg kg−1 day−1 for female rats, that do not lead to chronic toxicity [152]. Bai and Ogbourne [23] found that even relatively low amounts of glyphosate are enough to alter cell activities and produce cytotoxicity. Sub-agricultural concentrations of glyphosate and Roundup 400 in living organisms can impact the endocrine system at 0.5 ppm, estrogen receptor transcriptional activity at 2 ppm, and cytotoxicity at 10 ppm [151].

4.3. Genotoxicity

Several studies have questioned and mostly rejected claims that glyphosate is genotoxic [149,152]. Most previous DNA damage studies used excessive glyphosate doses [23]. Human cells have experienced DNA damage in other experiments utilizing sub-agricultural levels of glyphosate and Roundup [151,153,154]. According to Koller et al. [154], evidence of DNA damage was recorded after exposure of buccal epithelial cells to Roundup and glyphosate at quantities of 10 to 20 mg/L lower than recommended agricultural rates. In addition, DNA of caiman embryos was also damaged when exposed to Roundup at various sub-lethal doses [155]. However, Kier and Kirkland [156] reported that DNA damage from glyphosate is due to cytotoxicity, rather than genotoxicity. Then Bai and Ogbourne [23] concluded that, regardless of the cause, DNA damage can still happen at relatively low glyphosate levels.

4.4. Toxicity of Reproductive System

Glyphosate is unlikely to be toxic to reproductive systems [149] with no-observed-adverse-effect-level (NOAEL) between 300 mg/kg/day and 50 mg/kg/day [152]. Other studies reported that exposure to NOAEL concentrations of glyphosate may affect reproductive function in offspring [157,158]. Rates between 50 mg/kg and 450 mg/kg of Roundup were applied to rats during pregnancy without adverse effects; however, male offspring were affected [157]. These results highlight the need for more research to be carried out in order to fully comprehend the impacts of glyphosate-based herbicide use.

4.5. Carcinogenicity

When exposure is within the allowable NOAEL, studies have shown that glyphosate is not carcinogenic [23]. According to George et al. [159] and Thongprakaisang et al. [160], They reported glyphosate could promote tumour growth in skin cells (in vivo mouse) and breast cell proliferation (in vitro human models). In Thongprakaisang’s [160] study, glyphosate residue concentration in drinking water as low as 10−12 M stimulated hormone-induced breast cancer. The International Agency for Research on Cancer recently determined that glyphosate is most likely human carcinogenic [161]. It has been reported that non-Hodgkin lymphoma (NHL), cancer derived from lymphocytes of the regular human immune system, is caused by GBH [162]. Different subtypes of NHL exist depending on where the cell originates, for example, in T and B cells, or natural killer cells. The distinct stages of lymphocyte development are related to the different NHL subtypes. Typically, lymph nodes and the spleen are enlarged, as well as other tissues such as the blood and bone marrow are affected [162]. A variety of infection agents, chemical agents, and deficiencies in the immune system are among the causes of NHL [162]. Zhang et al. [163] discovered that being exposed to GBH raises the risk of NHL among workers that are highly exposed to glyphosate. In 2020, NHL accounted for approximately 77,000 new cases and caused almost 20,000 deaths in the United States alone [164]. There were many lawsuits in court by the time Bayer bought Monsanto in 2018, alleging that Roundup is linked to NHL [165,166,167]. Researchers from the University of Washington found that Roundup exposure increases NHL risk by 41%, a clear link between the herbicide and cancer [166].
Even though glyphosate and glyphosate-based herbicides are valuable and significant weed-management tools for agricultural and forestry practices, long-term appropriate procedures and control of glyphosate usage are still needed to ensure efficacy, minimize pollution, and avoid adverse human health effects.
Table 3. The toxicity of glyphosate-based herbicide formulations to a variety of living organisms.
Table 3. The toxicity of glyphosate-based herbicide formulations to a variety of living organisms.
Living OrganismsGLY ConcentrationToxic EffectsReferences
Human0.5–10 mg L−1Endocrine disruption[151]
>20 mg/L to >80 mg L−1DNA damage[154]
<1000 mg L−1Effects on oxidative balance[168]
4000 mg L−1Neurological disorder[169,170]
85–1690 mg L−1Blood disorder[171]
1–3 mg L−1Human cell toxicity[147]
36–178 mg L−1Effects on cell physiology[172]
Earthworm1–10 mg L−1Growth and reproductive defects[106,173]
Snails0.1–10 mg L−1Body growth[174,175,176]
Wasps960 g ha−1Parasite eggs are harmed when exposed to this substance[177]
Honeybee2.5–10 mg L−1Adaptation in agricultural environments[175,177,178,179]
Amphibians1.5–684 mg L−1Chronic toxicity[41,42,180,181,182,183,184]
Daphnia1.4–250 mg L−1Chronic toxicity[39,185]
Zebrafish10 mg L−1Toxicity in the reproductive system[186]
BirdsWetland with glyphosate aerial spray coverage of 50–90 percentReduced natural habitats and bird populations, as well as affected male genital organs[187,188,189]
Goldfish32 µg L−1Effects on acetylcholinesterase (AChE), malondialdehyde (MDA), and OH in the liver[183]
Reptiles144 mg L−1Increase in heterophils and total protein content DNA damaged, physiological stress, decrease in WBC[145,190,191]
Tilapia fish108–540 mg L−1Toxic to the point of death[192]
Wistar rat14.4–375 mg kg−1Consequences for physiology and reproduction[193,194]
Silver catfish0–5 mg L−1Enzymatic activity, leukocytes, vacuolization, melanomacrophages and cytoplasm all show changes[195]
Pig41% of IPAG and 15% surfactantEffects on the cardiovascular system[196]

5. Health and Immunological Impacts of GBH

Microorganisms break down glyphosate into its most active metabolite, AMPA, and methyl phosphonic acid, once it reaches the environment [149]. They are found in water, soil, plants, and food, among other places [11,197,198]. Human urine, blood, and breast milk have all been shown to contain glyphosate. The urine levels in exposed workers ranged from 0.26 to 73.5 g/L, whereas those in the general population ranged from 0.16 to 7.6 g/L [199]. Most likely, the skin, mouth, and through the lungs are where glyphosate is absorbed [149,200]. Although skin absorption is the most widely mentioned route of ingestion in affected farmers, it only accounts for about 2% of total absorption [201]. According to Williams et al. [149], glyphosate appears to build up in the small intestine, liver, kidneys, and colon before excreting in the urine and faeces within two days. Nonetheless, an increasing body of knowledge exists about the human health impacts of GBHs and glyphosate [202].
Glyphosate has been evaluated by the International Agency for Research on Cancer (IARC) as “probably carcinogenic” in humans after more than 40 years of widespread usage [165]. Other bodies, such as Health Canada [203] and the European Commission [204] have renewed their authorizations for glyphosate use based on scientific findings as of 2017. However, with the “Monsanto Papers” revelation in 2017, the validity of several studies was called into question [205]. Monsanto is accused of interfering with the disclosure of important glyphosate toxicity data and ghostwriting studies confirming the safety of the herbicide [205]. Consequently, according to Peilex and Pelletier [206], it seems appropriate to provide a summary of the observed impacts of glyphosate and GBHs on humans’ immune systems and cellular to systemic animal health consequences. Animal data is valuable in and of itself, not simply due to glyphosate contamination affecting a wide range of species in the environment, which might be transferred down the food chain. Given that numerous animals are common laboratory models and share largely conserved immunological systems and defensive processes with humans, they can provide essential knowledge when human data is restricted [207].

6. Health Effects and Toxicity of GLY and GBHs

6.1. Cellular Impact

In a rat cardiac cell model, no substantial harm was seen at glyphosate doses that are significant to the environment. However, toxicity was found to be dose-dependent with increasing doses of TN20 surfactant at a constant glyphosate dosage [208], indicating that GBH components, including glyphosate, may be harmful. Additionally, this study discovered that this combination of glyphosate and TN20 induced the proliferation of tumour cells. Apoptosis and necrosis caused their deaths equally; however, when TN20 levels were increased, apoptosis was the predominant mode of cell death [208]. Rats exposed to a mixture of 12 chemicals, including glyphosate, for 18 months showed some cytotoxic effects in testis and kidney cells [209]. However, determining glyphosate’s role in this cytotoxicity is difficult [206].
On the one hand, investigations of glyphosate alone on human cells, both fibrosarcoma and healthy cells, found no substantial cytotoxicity at environmentally relevant amounts [210] whereby GBHs show toxicity that is dose-dependent, even when producing at the sub-agricultural level. Roundup, like another GBH product called glyphogan, killed human Sertoli cells, with glyphogan causing the most substantial harm [211]. The necessity of investigating GBHs rather than glyphosate alone is highly required.
The glyphosate appeared to be highly genotoxic in Nile tilapia erythrocytes [209], but only affects both fibrosarcoma and healthy human cells at high doses [210]. These findings show that GLY and GBHs have the potential to be genotoxic, cytotoxic, or both and that the effects vary depending on the cell type.

6.2. Impacts on Reproduction, Hormones, and Teratogenicity

It has been demonstrated that glyphosate and GBHs interfere with the human estrogen pathway [206]. Roundup inhibits the production of estrogen by the aromatase enzyme in the placental and embryonic cell lines of humans [212,213]. For glyphosate, it has been proposed that the enzyme be inhibited competitively, an effect facilitated by adjuvants that increase glyphosate solubilization and activity in herbicide formulations [214].
Mesnage et al. [215] discovered the impact of GLY on estrogen receptors (ER) in breast cancer cell lines, most likely through an indirect mechanism because it is structurally incapable of binding ERα. Because ER antagonists impede the glyphosate effect, Thongprakaisang et al. [160] theorized that glyphosate has estrogen-like properties interacting with the ERα and ERβ. Furthermore, glyphosate may trigger up to 50% of the estrogen response and cause breast cancer cell proliferation. Upon exposure, both ERs were expressed after 6 h and just ERα after 24 h. In addition, endogenous estrogen had an antagonistic effect when it was present. As a result, GLY could disrupt estrogen pathways and cause endocrine disruption [206].
Male offspring of rats that were given glyphosate exposure in an acceptable amount have lower testosterone production [158]. This endocrine imbalance influenced reproductive behaviour, resulting in changes in sexual choice and a long period in the females’ first mount starting to reproduce [158]. Young et al. [216] reported that Roundup has been shown to suppress progesterone synthesis in human placental cells effectively, but only at greater doses compared to those that kill cells, implying that the progesterone impact is most likely due to cytotoxicity. GBHs also had an effect on human pregnancies, as shown by the association between GBH exposure and an increase in the frequency of premature births and miscarriages [217]. Several studies theorized that glyphosate cytotoxicity may be responsible for this behavior in light of its effects on embryonic, placental, and human cell lines [160,218].
GBHs have also been demonstrated to cause teratogenicity in fish (primarily in the form of heart abnormalities), mice, chickens, and rats, with brain and bone deformities [219,220,221]. According to Campana et al. [222], GBH exposure before a child’s birth has been associated with a higher prevalence of malformations such as Down’s syndrome and cleft lip in humans. Then Peillex and Pelletier [206] concluded that based on the evidence shown above, glyphosate and GBHs appear to have the ability to disrupt the system of reproduction at several stages, such as teratogenicity and hormonal pathway disruption.

6.3. Neurological Impact

As seen by acute poisoning instances resulting in neurological changes, some components of GBH can penetrate the blood–brain barrier. For example, aseptic meningitis with measurable glyphosate levels was found in cerebrospinal fluid of a 58-year-old woman who attempted suicide by consuming a significant dose of the glyphosate surfactant herbicide [223]. Vasculitic neuropathy struck a 70-year-old man after a month following a large-scale Roundup application with no protection, which was linked to the herbicide [224]. As a result, acute GBH poisoning or long-term exposure could hasten the onset of neurological illness.
When pregnant rats were given various doses of Roundup in their drinking water, their progeny had impaired movement [225]. Several studies have also indicated that in Parkinson’s disease, cerebral functions (both movement and learning ability) are both hindered. Therefore, it’s not surprising that increasing cases imply a link between GBH exposure and the start of Parkinson’s disease [226,227,228,229].

6.4. Digestive Impact

The shikimate pathway that glyphosate targets in plants is also found in microorganisms exposed to GBHs in their guts whether accidentally or not, through oral ingestion. When rumen of cattle was subjected to various doses of glyphosate, specific phyla of microorganisms decreased in favour of others, and the dysbiosis benefited pathogens [230]. The glyphosate prevents the growth of Enterococcus spp., which can reduce Clostridium botulinum in cattle rumens [124]. Compared with cattle not exposed to glyphosate, mycobiota in the rumen of dairy cows exposed to glyphosate differed [231].
There is considerable concern about glyphosate and the effects of GBHs on liver, an essential organ in detoxifying xenobiotics [206]. When rainbow trout were exposed to higher concentrations of GBH, they developed various liver pathologies, such as fibrosis and mild changes [232]. Pandey et al. [233] also detected mild changes and fibrosis in rats that exhibited accumulation of collagen plus increased liver weights and varying glycogen levels following oral administration of Roundup. A hepatic progenitor cell line from a human showed that glyphosate on its own had a negligible impact on the metabolome, as well as a decrease in polyunsaturated and long-chain fatty acids [234]. Only the lowest glyphosate concentration examined resulted in a substantial reduction, demonstrating a concentration-independent impact [206].
Other digestive system components appear unaffected by Roundup in weaned piglets [235], nor was there a substantial connection between GBH exposure and diabetes in GBH applicators [236]. There was no indication of kidney damage among children, despite glyphosate presence in their urine [237]. Overall, these findings show GBH causes damage to the liver at the cellular and histological level; however, they do not show any effects on the rest of the digestive organs [206].

6.5. Cardiovascular Impact

Numerous GBHs and glyphosate effects on cardiovascular system have been observed. After consumption of glyphosate-contaminated drinking water for 72 weeks, mice hemoglobin levels were significantly lower, causing non-significant anemia [238]. The aortic rings of rats treated with 1% glyphosate displayed an insignificant vasorelaxation, corresponding to 20% of a standard response. The atria of a rat heart in isolation were stimulated like controls, but they failed to undergo spasmodic spontaneous contractions when 1% glyphosate was added [239]. In human cases of acute poisoning, it was reported that four out of ten poisoned subjects developed heart arrhythmias [240]. Therefore, glyphosate alone appears to affect the cardiovascular system, and GBHs may have similar impacts, at the very least in situations of severe poisoning [206].

6.6. Impact on Carcinogenesis

The carcinogenic potential of glyphosate is still being discussed. George et al. [159] found mice treated topically with glyphosate not to develop tumors, but did experience tumor-propagating effects in a skin cancer test.

7. Regulations Currently in Effect

According to the EPA’s glyphosate Interim Registration Review Decision Case number 0178 of January 2020, glyphosate exposure poses no harm to human health. Although projected to be limited to the application area or nearby areas, the agency acknowledged potential ecological concerns for birds and mammals. In line with glyphosate’s usage as an herbicide, the EPA identified a possible risk from off-site spray drift to terrestrial and aquatic vegetation. When glyphosate is used according to label directions, the EPA believes the benefits outweigh the potential environmental concerns [241]. However, because glyphosate is a component of several GBHs, there are concerns about the GBHs’ regulation and management. Xu et al. [242] reported that the EPA regularly reviews the maximum allowable quantities in retail GBHs, including the permission granted to farmers using them for a specific reason for feed and food purposes. The Food and Drug Administration is therefore in charge of ensuring that imported and domestic goods sold in stores do not violate the EPA’s standards. The EPA issued glyphosate residue tolerances and determined the maximum allowable residue for all group 15 cereals at 30 mg/kg, except maize and rice, which are 5 mg/kg and 0.1 mg/kg, respectively; likewise, all oilseeds in group 20 to be 40 mg/kg, apart from canola, which is 20 mg/kg [243]. In Table 4, you can find a summary of the details of the glyphosate maximum residue limits (MRLs) set by different agencies for common cereals and grains
Furthermore, in July 2017, the California Office of Environmental Health Hazard Assessment (OEHHA) listed glyphosate as a known carcinogen under Proposition 65 Law and Regulations [166]. This law ensures that California is informed about substances that could cause congenital disabilities, reproductive harm, or cancer. Meanwhile, the OEHHA claims that glyphosate exposure below 1100 µg/day has no significant risk level for cancer [244].
Glyphosate residues in drinking water are regulated differently in different countries. In the United States, for example, glyphosate has a maximum contamination limit (MCL) of 700 µg/L [245] and 1000 μg/L in Australia [246].
The use of GBHs was recently extended by the European Commission (EC) until 2022 when it will be reviewed. Based on EFSA [247], the EC established the MRL for sunflower seed, barley, sorghum, soybean, and oat at 20 mg/kg; 10 mg/kg for wheat, rye, mustard seed, linseed, lentils, peas, lupin, cotton seed and rapeseed; 1.0 mg/kg for corn; 2.0 mg/kg for beans, and 0.1 mg/kg for unnamed grains and cereals.
Health Canada regulates herbicide and pesticide maximum residue limits (MRLs) [242]. It establishes a maximum allowable residue of 20 mg/kg for soybean and rapeseeds, 5 mg/kg for wheat, 15 mg/kg for oats, 10 mg/kg for barley, and maize at 3 mg/kg for glyphosate-treated crops.
Globally, the World Trade Organization (WTO) plays a significant role in regulating GLY through its Agreement on the Application of Sanitary and Phytosanitary Measures, which all of the nations that make up the organization have in common. The Codex Alimentarius (18-3485), which outlines worldwide recognized accepted standards and recommendations, is updated by the WTO and FAO and reviews GBHs.
Table 4. Maximum permissible glyphosate residue (mg/kg) in grains placed by different agencies.
Table 4. Maximum permissible glyphosate residue (mg/kg) in grains placed by different agencies.
Food SourcesEuropean CommissionHealth CanadaUSEPAWHO/FAO
Cereals
Barley20103030
Corn1355
Not AvailableNot Available0.1 (popcorn) and 3.5 (Sweet corn)3 (Sweet corn)
Not AvailableNot Available
Millet0.1Not Available3030
Oat20153030
Rice0.1Not Available0.1NA
Rye10Not Available3030
Sorghum20Not Available3030
TeffNot AvailableNot Available530
Wheat1053030
20 (bran)
Oilseeds
Cotton seeds10404040
Hemp seeds0.1Not Available40Not Available
Linseed10340Not Available
Mustard seeds101040Not Available
Peanuts0.1Not Available0.1Not Available
Pumpkin seeds0.1Not Available40Not Available
Rapeseeds10202030
Safflower seeds0.1Not Available40Not Available
Sesame seeds0.11040Not Available
Soybeans20202020
Sunflower seeds20Not Available407
Pseudocereals
Buckwheat0.1Not Available3030
QuinoaNot AvailableNot Available530
Pulses
Beans2452
Lentils10455
Lupins10Not Available5Not Available
Peas10585
Sources: [74,241,247,248].

8. Conclusions

With respect to the literature that focuses on the effects of GBHs and the current reality in agriculture and the environment, it is clear that GBHs, with their weed-controlling effect are becoming more commonly used. Glyphosate will likely continue to be used, regardless of all the damage it can do to the environment. However, glyphosate’s accumulation in soil, water, and indirectly in the humans and animals that consume agricultural products should not be disregarded. The majority of agricultural production today relies heavily on herbicides, and there are no environmentally friendly or commercially viable alternatives. Even though glyphosate is mineralized, under certain conditions, glyphosate and its metabolites have long half-lives. In some circumstances, plants, soil, and water may still contain glyphosate and AMPA residues. Studies have shown that water, soil, and a variety of foods are contaminated at rates that could endanger the environment. Regardless, according to most scientific studies, the contamination rate does not prove harmful to most organisms and poses no risks to the environment provided recommended application rates are adhered to, and reapplications are avoided. However, a health concern can arise from the long-term buildup of these compounds in humans, animals, and the environment as glyphosate affects the body in an insidious way, slowly and over time.
In agricultural practices and home gardeners, glyphosate is an important weed control tool. Nevertheless, new research indicates that it is imperative to determine the most environmentally and toxicologically sensitive scenarios to guide glyphosate use in the future for it to continue to be useful, thereby assuring minimum contamination of the environment and no negative health consequences. Therefore, further interdisciplinary research regarding prolonged herbicide exposure at low levels, microbial community alterations, growth of antibiotic resistance, and increase in disease outbreaks in humans, plants, and animals is recommended. Considering all possible health risks, independent studies are required to revisit glyphosate residue tolerance criteria in food, animal feed, soil, and water.
Based on literature findings, around 30% of global cropland was contaminated with glyphosate at low levels, while 93% of worldwide cropland was contaminated with AMPA. Much information is lacking about the ecotoxic effects of AMPA, the most persistent and recalcitrant metabolite, i.e., complex for soil microbes to degrade. Therefore, to better assess the risk of contamination by AMPA, it is necessary to elucidate the ecotoxicity of this metabolite, as well as its biodegradation pathways and kinetics. Research like this could lead to the standardization of glyphosate rules among organizations and provide a definitive answer to glyphosate toxicity among regulators.

Author Contributions

Conceptualization, B.S.O. and O.S.D.; data collection, B.S.O.; writing—original draft preparation, B.S.O. and O.S.D.; writing—review and editing, B.S.O., O.S.D., P.O.A., S.W.G. and G.-I.E.; supervision, O.S.D., P.O.A. and S.W.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by University of Venda, South Africa through Research, and Innovation Committee of the University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge with thanks the SA-UK University Staff Doctoral Program (SA-UK USDP) for its support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Landrigan, P.J.; Fuller, R.; Acosta, N.J.; Adeyi, O.; Arnold, R.; Baldé, A.B.; Bertollini, R.; Bose-O’Reilly, S.; Boufford, J.I.; Breysse, P.N. The Lancet Commission on pollution and health. Lancet 2018, 391, 462–512. [Google Scholar] [CrossRef]
  2. Rose, M.T.; Cavagnaro, T.R.; Scanlan, C.A.; Rose, T.J.; Vancov, T.; Kimber, S.; Kennedy, I.R.; Kookana, R.S.; Van Zwieten, L. Impact of herbicides on soil biology and function. Adv. Agron. 2016, 136, 133–220. [Google Scholar]
  3. la Cecilia, D.; Maggi, F. Influential sources of uncertainty in glyphosate biochemical degradation in soil. Math. Comput. Simul. 2020, 175, 121–139. [Google Scholar] [CrossRef]
  4. USEPA. 2018 Edition of the Drinking Water Standards and Health Advisories Tables; USEPA: Washington, DC, USA, 2018.
  5. Liang, C.-P.; Sack, C.; McGrath, S.; Cao, Y.; Thompson, C.J.; Robin, L.P. US Food and Drug Administration regulatory pesticide residue monitoring of human foods: 2009–2017. Food Addit. Contam. Part A 2021, 38, 1520–1538. [Google Scholar] [CrossRef]
  6. Rodríguez-Eugenio, N.; McLaughlin, M.; Pennock, D. Soil Pollution: A Hidden Reality; FAO: Rome, Italy, 2018. [Google Scholar]
  7. Maggi, F.; Tang, F.H.; la Cecilia, D.; McBratney, A. PEST-CHEMGRIDS, global gridded maps of the top 20 crop-specific pesticide application rates from 2015 to 2025. Sci. Data 2019, 6, 1–20. [Google Scholar] [CrossRef]
  8. Franz, J.E. The Herbicide Glyphosate; Butterworths: London, UK, 1985; pp. 3–17. [Google Scholar]
  9. Blair, A.; Fritschi, L.; McLaughlin, J.; Sergi, C.; Calaf, G.; Curieux, F. IARC Monographs Volume 112: Evaluation of five organophosphate insecticides and herbicides. In International Agency for Research on Cancer; World Health Organization: Lyon, France, 2015. [Google Scholar]
  10. Temple, W. Review of the evidence relating to glyphosate and carcinogenicity. Environ. Prot. Agency 2016, 1–9. [Google Scholar]
  11. Tzanetou, E.; Karasali, H. Glyphosate residues in soil and air: An integrated review. Pests Weeds Dis. Agric. Crop Anim. Husb. Prod. 2020. [Google Scholar] [CrossRef]
  12. Zhang, C.; Hu, X.; Luo, J.; Wu, Z.; Wang, L.; Li, B.; Wang, Y.; Sun, G. Degradation dynamics of glyphosate in different types of citrus orchard soils in China. Molecules 2015, 20, 1161–1175. [Google Scholar] [CrossRef]
  13. Benbrook, C.M. Trends in glyphosate herbicide use in the United States and globally. Environ. Sci. Eur. 2016, 28, 3. [Google Scholar] [CrossRef]
  14. Grandcoin, A.; Piel, S.; Baures, E. AminoMethylPhosphonic acid (AMPA) in natural waters: Its sources, behavior and environmental fate. Water Res. 2017, 117, 187–197. [Google Scholar] [CrossRef]
  15. Franz, J.E.; Mao, M.K.; Sikorski, J.A. Glyphosate: A Unique and Global Herbcide; ACS Monograph No. 189; American Chemical Society: Washington, DC, USA, 1997; p. 653. [Google Scholar]
  16. European Chemicals Agency. Glyphosate Not Classified as a Carcinogen by ECHA. ECHA/PR/17/06; 15 March 2017. Available online: https://echa.europa.eu/-/glyphosate-not-classified-as-a-carcinogen-by-echa (accessed on 12 July 2022).
  17. Székács, A.; Darvas, B. Re-registration challenges of glyphosate in the European Union. Front. Environ. Sci. 2018, 78. [Google Scholar] [CrossRef]
  18. EFSA. Conclusion on the peer review of the pesticide risk assessment of the active substance glyphosate. EFSA J. 2015, 13, 4302. [Google Scholar]
  19. EU Votes for Five More Years Usage of Herbicide Glyphosate’. NRC Handelsblad. 28 November 2017. Available online: https://www.nrc.nl/nieuws/2017/11/28/roundup-omstreden-bestrijdingsmiddel-glyfosaat-blijft-nog-zeker-vijf-jaar-te-gebruiken-in-eu-14259800-a1582828 (accessed on 12 July 2022). (In Dutch).
  20. CLH-O-0000001412-86-149/F; Opinion Proposing Harmonised Classification and Labeling at EU Level of Glyphosate (ISO); N-(phosphonomethyl)Glycine. European Chemicals Agency: Helsinki, Finland, 2017. Available online: https://echa.europa.eu/documents/10162/2d3a87cc-5ca1-31d6-8967-9f124f1ab7ae (accessed on 12 July 2022).
  21. European-Regulation. Regulation of the European Parliament and of the Council of 21 October 2009 Concerning the Placing of Plant Protection Products on the Market and Repealing Council Directives 79/117/ EEC and 91/414/EEC; European-Regulation-1107/2009; Official Journal of the European Union: Maastricht, The Netherlands, 2009; pp. 1–50. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=celex%3A32009R1107 (accessed on 12 July 2022).
  22. Duke, S.O.; Powles, S.B. Glyphosate: A once-in-a-century herbicide. Pest Manag. Sci. Former. Pestic. Sci. 2008, 64, 319–325. [Google Scholar] [CrossRef]
  23. Bai, S.H.; Ogbourne, S.M. Glyphosate: Environmental contamination, toxicity and potential risks to human health via food contamination. Environ. Sci. Pollut. Res. 2016, 23, 18988–19001. [Google Scholar] [CrossRef]
  24. Fenner, K.; Canonica, S.; Wackett, L.P.; Elsner, M. Evaluating pesticide degradation in the environment: Blind spots and emerging opportunities. Science 2013, 341, 752–758. [Google Scholar] [CrossRef]
  25. Laitinen, P.; Siimes, K.; Eronen, L.; Rämö, S.; Welling, L.; Oinonen, S.; Mattsoff, L.; Ruohonen-Lehto, M. Fate of the herbicides glyphosate, glufosinate-ammonium, phenmedipham, ethofumesate and metamitron in two Finnish arable soils. Pest Manag. Sci. Former. Pestic. Sci. 2006, 62, 473–491. [Google Scholar] [CrossRef]
  26. Bergström, L.; Börjesson, E.; Stenström, J. Laboratory and lysimeter studies of glyphosate and aminomethylphosphonic acid in a sand and a clay soil. J. Environ. Qual. 2011, 40, 98–108. [Google Scholar] [CrossRef]
  27. Sihtmäe, M.; Blinova, I.; Künnis-Beres, K.; Kanarbik, L.; Heinlaan, M.; Kahru, A. Ecotoxicological effects of different glyphosate formulations. Appl. Soil Ecol. 2013, 72, 215–224. [Google Scholar] [CrossRef]
  28. Yang, X.; Wang, F.; Bento, C.P.; Meng, L.; van Dam, R.; Mol, H.; Liu, G.; Ritsema, C.J.; Geissen, V. Decay characteristics and erosion-related transport of glyphosate in Chinese loess soil under field conditions. Sci. Total Environ. 2015, 530, 87–95. [Google Scholar] [CrossRef]
  29. Al-Rajab, A.J.; Schiavon, M. Degradation of 14C-glyphosate and aminomethylphosphonic acid (AMPA) in three agricultural soils. J. Environ. Sci. 2010, 22, 1374–1380. [Google Scholar] [CrossRef]
  30. Mamy, L.; Barriuso, E.; Gabrielle, B. Environmental fate of herbicides trifluralin, metazachlor, metamitron and sulcotrione compared with that of glyphosate, a substitute broad spectrum herbicide for different glyphosate-resistant crops. Pest Manag. Sci. Former. Pestic. Sci. 2005, 61, 905–916. [Google Scholar] [CrossRef] [PubMed]
  31. Kwiatkowska, M.; Huras, B.; Bukowska, B. The effect of metabolites and impurities of glyphosate on human erythrocytes (in vitro). Pestic. Biochem. Physiol. 2014, 109, 34–43. [Google Scholar] [CrossRef]
  32. Borggaard, O.K.; Gimsing, A.L. Fate of glyphosate in soil and the possibility of leaching to ground and surface waters: A review. Pest Manag. Sci. Former. Pestic. Sci. 2008, 64, 441–456. [Google Scholar] [CrossRef] [PubMed]
  33. Morillo, E.; Undabeytia, T.; Maqueda, C.; Ramos, A. Glyphosate adsorption on soils of different characteristics.: Influence of copper addition. Chemosphere 2000, 40, 103–107. [Google Scholar] [CrossRef]
  34. Ghafoor, A.; Jarvis, N.; Thierfelder, T.; Stenström, J. Measurements and modeling of pesticide persistence in soil at the catchment scale. Sci. Total Environ. 2011, 409, 1900–1908. [Google Scholar] [CrossRef]
  35. Syan, H.S.; Prasher, S.O.; Pageau, D.; Singh, J. Dissipation and persistence of major herbicides applied in transgenic and non-transgenic canola production in Quebec. Eur. J. Soil Biol. 2014, 63, 21–27. [Google Scholar] [CrossRef]
  36. Shushkova, T.; Vasilieva, G.; Ermakova, I.; Leontievsky, A. Sorption and microbial degradation of glyphosate in soil suspensions. Appl. Biochem. Microbiol. 2009, 45, 599–603. [Google Scholar] [CrossRef]
  37. Gimsing, A.; Borggaard, O.; Bang, M. Influence of soil composition on adsorption of glyphosate and phosphate by contrasting Danish surface soils. Eur. J. Soil Sci. 2004, 55, 183–191. [Google Scholar] [CrossRef]
  38. Laitinen, P.; Rämö, S.; Nikunen, U.; Jauhiainen, L.; Siimes, K.; Turtola, E. Glyphosate and phosphorus leaching and residues in boreal sandy soil. Plant Soil 2009, 323, 267–283. [Google Scholar] [CrossRef]
  39. Cuhra, M.; Traavik, T.; Bøhn, T. Clone-and age-dependent toxicity of a glyphosate commercial formulation and its active ingredient in Daphnia magna. Ecotoxicology 2013, 22, 251–262. [Google Scholar] [CrossRef]
  40. Cuhra, M. Glyphosate Non-Toxicity: The Genesis of A Scientific Fact. J. Biol. Phys. Chem. 2015, 15, 89–96. [Google Scholar] [CrossRef]
  41. Moore, L.J.; Fuentes, L.; Rodgers Jr, J.H.; Bowerman, W.W.; Yarrow, G.K.; Chao, W.Y.; Bridges Jr, W.C. Relative toxicity of the components of the original formulation of Roundup® to five North American anurans. Ecotoxicol. Environ. Saf. 2012, 78, 128–133. [Google Scholar] [CrossRef]
  42. Howe, C.M.; Berrill, M.; Pauli, B.D.; Helbing, C.C.; Werry, K.; Veldhoen, N. Toxicity of glyphosate-based pesticides to four North American frog species. Environ. Toxicol. Chem. Int. J. 2004, 23, 1928–1938. [Google Scholar] [CrossRef] [PubMed]
  43. Peixoto, F. Comparative effects of the Roundup and glyphosate on mitochondrial oxidative phosphorylation. Chemosphere 2005, 61, 1115–1122. [Google Scholar] [CrossRef] [PubMed]
  44. Van Os, N.M. (Ed.) Nonionic Surfactants: Organic Chemistry; CRC Press: Boca Raton, FL, USA, 1997. [Google Scholar]
  45. European Food Safety Authority (EFSA). Request for the evaluation of the toxicological assessment of the co-formulant POE-tallowamine. EFSA J. 2015, 13, 4303. [Google Scholar]
  46. Guilherme, S.; Santos, M.A.; Barroso, C.; Gaivão, I.; Pacheco, M. Differential genotoxicity of Roundup® formulation and its constituents in blood cells of fish (Anguilla anguilla): Considerations on chemical interactions and DNA damaging mechanisms. Ecotoxicology 2012, 21, 1381–1390. [Google Scholar] [CrossRef]
  47. Monsanto. Monsanto Product Safety Assistance Website. 2019. Available online: https://www.sdslibrary.monsanto.com/Pages/Default.aspx (accessed on 12 July 2022).
  48. Mesnage, R.; Defarge, N.; Spiroux de Vendômois, J.; Séralini, G.-E. Major pesticides are more toxic to human cells than their declared active principles. BioMed Res. Int. 2014, 2014, 179691. [Google Scholar] [CrossRef]
  49. Samsel, A.; Seneff, S. Glyphosate, pathways to modern diseases III: Manganese, neurological diseases, and associated pathologies. Surg. Neurol. Int. 2015, 6. [Google Scholar] [CrossRef]
  50. Defarge, N.; Takács, E.; Lozano, V.L.; Mesnage, R.; Spiroux de Vendômois, J.; Séralini, G.-E.; Székács, A. Co-formulants in glyphosate-based herbicides disrupt aromatase activity in human cells below toxic levels. Int. J. Environ. Res. Public Health 2016, 13, 264. [Google Scholar] [CrossRef]
  51. Cuhra, M.; Bøhn, T.; Cuhra, P. Glyphosate: Too much of a good thing? Front. Environ. Sci. 2016, 4, 28. [Google Scholar] [CrossRef]
  52. Gillezeau, C.; van Gerwen, M.; Shaffer, R.M.; Rana, I.; Zhang, L.; Sheppard, L.; Taioli, E. The evidence of human exposure to glyphosate: A review. Environ. Health 2019, 18, 1–4. [Google Scholar] [CrossRef]
  53. Acquavella, J.F.; Alexander, B.H.; Mandel, J.S.; Gustin, C.; Baker, B.; Chapman, P.; Bleeke, M. Glyphosate biomonitoring for farmers and their families: Results from the Farm Family Exposure Study. Environ. Health Perspect. 2004, 112, 321–326. [Google Scholar] [CrossRef]
  54. Jauhiainen, A.; Räsänen, K.; Sarantila, R.; Nuutinen, J.; Kangas, J. Occupational exposure of forest workers to glyphosate during brush saw spraying work. Am. Ind. Hyg. Assoc. J. 1991, 52, 61–64. [Google Scholar] [CrossRef] [PubMed]
  55. Mesnage, R.; Moesch, C.; Grand, R.; Lauthier, G.; Vendômois, J.; Gress, S.; Séralini, G. Glyphosate exposure in a farmer’s family. J. Environ. Prot. 2012, 3, 1001. [Google Scholar] [CrossRef] [Green Version]
  56. Connolly, A.; Jones, K.; Galea, K.S.; Basinas, I.; Kenny, L.; McGowan, P.; Coggins, M. Exposure assessment using human biomonitoring for glyphosate and fluroxypyr users in amenity horticulture. Int. J. Hyg. Environ. Health 2017, 220, 1064–1073. [Google Scholar] [CrossRef]
  57. Rendón-von Osten, J.; Dzul-Caamal, R. Glyphosate residues in groundwater, drinking water and urine of subsistence farmers from intensive agriculture localities: A survey in Hopelchén, Campeche, Mexico. Int. J. Environ. Res. Public Health 2017, 14, 595. [Google Scholar] [CrossRef]
  58. Jayasumana, C.; Gunatilake, S.; Siribaddana, S. Simultaneous exposure to multiple heavy metals and glyphosate may contribute to Sri Lankan agricultural nephropathy. BMC Nephrol. 2015, 16, 1–8. [Google Scholar] [CrossRef] [PubMed]
  59. Curwin, B.D.; Hein, M.J.; Sanderson, W.T.; Striley, C.; Heederik, D.; Kromhout, H.; Reynolds, S.J.; Alavanja, M.C. Urinary pesticide concentrations among children, mothers and fathers living in farm and non-farm households in Iowa. Ann. Occup. Hyg. 2007, 51, 53–65. [Google Scholar] [CrossRef]
  60. McGuire, M.K.; McGuire, M.A.; Price, W.J.; Shafii, B.; Carrothers, J.M.; Lackey, K.A.; Goldstein, D.A.; Jensen, P.K.; Vicini, J.L. Glyphosate and aminomethylphosphonic acid are not detectable in human milk. Am. J. Clin. Nutr. 2016, 103, 1285–1290. [Google Scholar] [CrossRef]
  61. Aris, A.; Leblanc, S. Maternal and fetal exposure to pesticides associated to genetically modified foods in Eastern Townships of Quebec, Canada. Reprod. Toxicol. 2011, 31, 528–533. [Google Scholar] [CrossRef] [PubMed]
  62. Parvez, S.; Gerona, R.; Proctor, C.; Friesen, M.; Ashby, J.; Reiter, J.; Lui, Z.; Winchester, P. Glyphosate exposure in pregnancy and shortened gestational length: A prospective Indiana birth cohort study. Environ. Health 2018, 17, 1–12. [Google Scholar] [CrossRef] [PubMed]
  63. Connolly, A.; Basinas, I.; Jones, K.; Galea, K.S.; Kenny, L.; McGowan, P.; Coggins, M.A. Characterising glyphosate exposures among amenity horticulturists using multiple spot urine samples. Int. J. Hyg. Environ. Health 2018, 221, 1012–1022. [Google Scholar] [CrossRef]
  64. Knudsen, L.E.; Hansen, P.W.; Mizrak, S.; Hansen, H.K.; Mørck, T.A.; Nielsen, F.; Siersma, V.; Mathiesen, L. Biomonitoring of Danish school children and mothers including biomarkers of PBDE and glyphosate. Rev. Environ. Health 2017, 32, 279–290. [Google Scholar] [CrossRef]
  65. Conrad, A.; Schröter-Kermani, C.; Hoppe, H.-W.; Rüther, M.; Pieper, S.; Kolossa-Gehring, M. Glyphosate in German adults–Time trend (2001 to 2015) of human exposure to a widely used herbicide. Int. J. Hyg. Environ. Health 2017, 220, 8–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Hoppe, H.W. Determination of Glyphosate Residues in Human Urine Samples from 18 European Countries; Medical Laboratory Bremen: Bremen, Germany, 2013. [Google Scholar]
  67. Varona, M.; Henao, G.L.; Díaz, S.; Lancheros, A.; Murcia, Á.; Rodríguez, N.; Álvarez, V.H. Effects of aerial applications of the herbicide, glyphosate and insecticides on human health. Biomedica 2009, 29, 456–475. [Google Scholar] [CrossRef]
  68. Kongtip, P.; Nankongnab, N.; Phupancharoensuk, R.; Palarach, C.; Sujirarat, D.; Sangprasert, S.; Sermsuk, M.; Sawattrakool, N.; Woskie, S.R. Glyphosate and paraquat in maternal and fetal serums in Thai women. J. Agromed. 2017, 22, 282–289. [Google Scholar] [CrossRef] [PubMed]
  69. Mesnage, R.; Benbrook, C.; Antoniou, M.N. Insight into the confusion over surfactant co-formulants in glyphosate-based herbicides. Food Chem. Toxicol. 2019, 128, 137–145. [Google Scholar] [CrossRef]
  70. Sørensen, S.R.; Schultz, A.; Jacobsen, O.S.; Aamand, J. Sorption, desorption and mineralisation of the herbicides glyphosate and MCPA in samples from two Danish soil and subsurface profiles. Environ. Pollut. 2006, 141, 184–194. [Google Scholar] [CrossRef]
  71. Herath, H.; Moldrup, P.; de Jonge, L.W.; Nicolaisen, M.; Norgaard, T.; Arthur, E.; Paradelo, M. Clay-to-carbon ratio controls the effect of herbicide application on soil bacterial richness and diversity in a loamy field. Water Air Soil Pollut. 2017, 228, 1–12. [Google Scholar] [CrossRef]
  72. Gimsing, A.; Szilas, C.; Borggaard, O. Sorption of glyphosate and phosphate by variable-charge tropical soils from Tanzania. Geoderma 2007, 138, 127–132. [Google Scholar] [CrossRef]
  73. Muskus, A.M.; Krauss, M.; Miltner, A.; Hamer, U.; Nowak, K.M. Effect of temperature, pH and total organic carbon variations on microbial turnover of 13C315N-glyphosate in agricultural soil. Sci. Total Environ. 2019, 658, 697–707. [Google Scholar] [CrossRef]
  74. FAO. Pesticide Detail, CODEXALIMENTARIUS FAO-WHO. Available online: https://www.fao.org/fao-who-codexalimentarius/codex-texts/dbs/pestres/pesticide-detail/en/?p_id=158 (accessed on 19 June 2022).
  75. Bandana, B.; Sharma, N.; Joshi, R.; Gulati, A.; Sondhia, S. Dissipation kinetics of glyphosate in tea and tea-field under northwestern mid-hill conditions of India. J. Pestic. Sci. 2015, D14-085. [Google Scholar] [CrossRef]
  76. Okada, E.; Costa, J.L.; Bedmar, F. Adsorption and mobility of glyphosate in different soils under no-till and conventional tillage. Geoderma 2016, 263, 78–85. [Google Scholar] [CrossRef]
  77. Wang, M.; Zhang, G.; Qiu, G.; Cai, D.; Wu, Z. Degradation of herbicide (glyphosate) using sunlight-sensitive MnO2/C catalyst immediately fabricated by high energy electron beam. Chem. Eng. J. 2016, 306, 693–703. [Google Scholar] [CrossRef]
  78. Lesueur, C.; Pfeffer, M.; Fuerhacker, M. Photodegradation of phosphonates in water. Chemosphere 2005, 59, 685–691. [Google Scholar] [CrossRef] [PubMed]
  79. Sidoli, P.; Baran, N.; Angulo-Jaramillo, R. Glyphosate and AMPA adsorption in soils: Laboratory experiments and pedotransfer rules. Environ. Sci. Pollut. Res. 2016, 23, 5733–5742. [Google Scholar] [CrossRef]
  80. Richards, B.K.; Pacenka, S.; Meyer, M.T.; Dietze, J.E.; Schatz, A.L.; Teuffer, K.; Aristilde, L.; Steenhuis, T.S. Antecedent and post-application rain events trigger glyphosate transport from runoff-prone soils. Environ. Sci. Technol. Lett. 2018, 5, 249–254. [Google Scholar] [CrossRef]
  81. Saunders, L.E.; Pezeshki, R. Glyphosate in runoff waters and in the root-zone: A review. Toxics 2015, 3, 462–480. [Google Scholar] [CrossRef]
  82. Primost, J.E.; Marino, D.J.; Aparicio, V.C.; Costa, J.L.; Carriquiriborde, P. Glyphosate and AMPA, “pseudo-persistent” pollutants under real-world agricultural management practices in the Mesopotamic Pampas agroecosystem, Argentina. Environ. Pollut. 2017, 229, 771–779. [Google Scholar] [CrossRef] [PubMed]
  83. Alonso, L.L.; Demetrio, P.M.; Etchegoyen, M.A.; Marino, D.J. Glyphosate and atrazine in rainfall and soils in agroproductive areas of the pampas region in Argentina. Sci. Total Environ. 2018, 645, 89–96. [Google Scholar] [CrossRef]
  84. Berman, M.C.; Marino, D.J.G.; Quiroga, M.V.; Zagarese, H. Occurrence and levels of glyphosate and AMPA in shallow lakes from the Pampean and Patagonian regions of Argentina. Chemosphere 2018, 200, 513–522. [Google Scholar] [CrossRef] [PubMed]
  85. Bonansea, R.I.; Filippi, I.; Wunderlin, D.A.; Marino, D.J.G.; Amé, M.V. The fate of glyphosate and AMPA in a freshwater endorheic basin: An ecotoxicological risk assessment. Toxics 2017, 6, 3. [Google Scholar] [CrossRef] [PubMed]
  86. Van Stempvoort, D.R.; Spoelstra, J.; Senger, N.D.; Brown, S.J.; Post, R.; Struger, J. Glyphosate residues in rural groundwater, Nottawasaga River watershed, Ontario, Canada. Pest Manag. Sci. 2016, 72, 1862–1872. [Google Scholar] [CrossRef]
  87. Battaglin, W.A.; Meyer, M.; Kuivila, K.; Dietze, J. Glyphosate and its degradation product AMPA occur frequently and widely in US soils, surface water, groundwater, and precipitation. JAWRA J. Am. Water Resour. Assoc. 2014, 50, 275–290. [Google Scholar] [CrossRef]
  88. Ruiz-Toledo, J.; Castro, R.; Rivero-Pérez, N.; Bello-Mendoza, R.; Sánchez, D. Occurrence of glyphosate in water bodies derived from intensive agriculture in a tropical region of southern Mexico. Bull. Environ. Contam. Toxicol. 2014, 93, 289–293. [Google Scholar] [CrossRef]
  89. Abrantes, N.; Pereira, R.; Gonçalves, F. Occurrence of pesticides in water, sediments, and fish tissues in a lake surrounded by agricultural lands: Concerning risks to humans and ecological receptors. Water Air Soil Pollut. 2010, 212, 77–88. [Google Scholar] [CrossRef]
  90. Sanchís, J.; Kantiani, L.; Llorca, M.; Rubio, F.; Ginebreda, A.; Fraile, J.; Garrido, T.; Farré, M. Determination of glyphosate in groundwater samples using an ultrasensitive immunoassay and confirmation by on-line solid-phase extraction followed by liquid chromatography coupled to tandem mass spectrometry. Anal. Bioanal. Chem. 2012, 402, 2335–2345. [Google Scholar] [CrossRef] [PubMed]
  91. Stewart, M.; Olsen, G.; Hickey, C.W.; Ferreira, B.; Jelić, A.; Petrović, M.; Barcelo, D.A. Survey of emerging contaminants in the estuarine receiving environment around Auckland, New Zealand. Sci. Total Environ. 2014, 468, 202–210. [Google Scholar] [CrossRef] [PubMed]
  92. Todorovic, G.R.; Mentler, A.; Popp, M.; Hann, S.; Köllensperger, G.; Rampazzo, N.; Blum, W.E. Determination of glyphosate and AMPA in three representative agricultural Austrian soils with a HPLC-MS/MS method. Soil Sediment Contam. Int. J. 2013, 22, 332–350. [Google Scholar] [CrossRef]
  93. Zgheib, S.; Moilleron, R.; Chebbo, G. Priority pollutants in urban stormwater: Part 1–Case of separate storm sewers. Water Res. 2012, 46, 6683–6692. [Google Scholar] [CrossRef]
  94. Grunewald, K.; Schmidt, W.; Unger, C.; Hanschmann, G. Behavior of glyphosate and aminomethylphosphonic acid (AMPA) in soils and water of reservoir Radeburg II catchment (Saxony/Germany). J. Plant Nutr. Soil Sci. 2001, 164, 65–70. [Google Scholar] [CrossRef]
  95. Koskinen, W.C.; Marek, L.J.; Hall, K.E. Analysis of glyphosate and aminomethylphosphonic acid in water, plant materials and soil. Pest Manag. Sci. 2016, 72, 423–432. [Google Scholar] [CrossRef] [PubMed]
  96. Valle, A.; Mello, F.; Alves-Balvedi, R.; Rodrigues, L.; Goulart, L. Glyphosate detection: Methods, needs and challenges. Environ. Chem. Lett. 2019, 17, 291–317. [Google Scholar] [CrossRef]
  97. Silva, V.; Montanarella, L.; Jones, A.; Fernández-Ugalde, O.; Mol, H.G.; Ritsema, C.J.; Geissen, V. Distribution of glyphosate and aminomethylphosphonic acid (AMPA) in agricultural topsoils of the European Union. Sci. Total Environ. 2018, 621, 1352–1359. [Google Scholar] [CrossRef] [PubMed]
  98. Busse, M.D.; Ratcliff, A.W.; Shestak, C.J.; Powers, R.F. Glyphosate toxicity and the effects of long-term vegetation control on soil microbial communities. Soil Biol. Biochem. 2001, 33, 1777–1789. [Google Scholar] [CrossRef]
  99. Araújo, A.D.; Monteiro, R.T.R.; Abarkeli, R. Effect of glyphosate on the microbial activity of two Brazilian soils. Chemosphere 2003, 52, 799–804. [Google Scholar] [CrossRef]
  100. Johnson-Maynard, J.; Lugo-Perez, J. Earthworm populations, microbial biomass and coffee production in different experimental agroforestry management systems in Costa Rica. Carib. J. Sci. 2006, 42, 397–409. [Google Scholar]
  101. García-Pérez, J.A.; Alarcón-Gutiérrez, E.; Perroni, Y.; Barois, I. Earthworm communities and soil properties in shaded coffee plantations with and without application of glyphosate. Appl. Soil Ecol. 2014, 83, 230–237. [Google Scholar] [CrossRef]
  102. Zhou, C.-F.; Wang, Y.-J.; Yu, Y.-C.; Sun, R.-J.; Zhu, X.-D.; Zhang, H.-L.; Zhou, D.-M. Does glyphosate impact on Cu uptake by, and toxicity to, the earthworm Eisenia fetida? Ecotoxicology 2012, 21, 2297–2305. [Google Scholar] [CrossRef]
  103. Fusilero, M.A.; Mangubat, J.; Ragas, R.E.; Baguinon, N.; Taya, H.; Rasco Jr, E. Weed management systems and other factors affecting the earthworm population in a banana plantation. Eur. J. Soil Biol. 2013, 56, 89–94. [Google Scholar] [CrossRef]
  104. Pereira, J.L.; Antunes, S.C.; Castro, B.B.; Marques, C.R.; Gonçalves, A.M.; Gonçalves, F.; Pereira, R. Toxicity evaluation of three pesticides on non-target aquatic and soil organisms: Commercial formulation versus active ingredient. Ecotoxicology 2009, 18, 455–463. [Google Scholar] [CrossRef] [PubMed]
  105. Pelosi, C.; Barot, S.; Capowiez, Y.; Hedde, M.; Vandenbulcke, F. Pesticides and earthworms. A review. Agron. Sustain. Dev. 2014, 34, 199–228. [Google Scholar] [CrossRef]
  106. Correia, F.; Moreira, J. Effects of glyphosate and 2, 4-D on earthworms (Eisenia foetida) in laboratory tests. Bull. Environ. Contam. Toxicol. 2010, 85, 264–268. [Google Scholar] [CrossRef]
  107. Santadino, M.; Coviella, C.; Momo, F. Glyphosate sublethal effects on the population dynamics of the earthworm Eisenia fetida (Savigny, 1826). Water Air Soil Pollut. 2014, 225, 1–8. [Google Scholar] [CrossRef]
  108. Kremer, R.J.; Means, N.E. Glyphosate and glyphosate-resistant crop interactions with rhizosphere microorganisms. Eur. J. Agron. 2009, 31, 153–161. [Google Scholar] [CrossRef]
  109. Wolmarans, K.; Swart, W.J. Influence of glyphosate, other herbicides and genetically modified herbicideresistantcrops on soil microbiota: A review. S. Afr. J. Plant Soil 2014, 31, 177–186. [Google Scholar] [CrossRef]
  110. Annett, R.; Habibi, H.R.; Hontela, A. Impact of glyphosate and glyphosate-based herbicides on the freshwater environment. J. Appl. Toxicol. 2014, 34, 458–479. [Google Scholar] [CrossRef]
  111. King, J.J.; Wagner, R.S. Toxic effects of the herbicide Roundup® Regular on Pacific Northwestern amphibians. Northwestern Nat. 2010, 91, 318–324. [Google Scholar] [CrossRef]
  112. Braz-Mota, S.; Sadauskas-Henrique, H.; Duarte, R.M.; Val, A.L.; Almeida-Val, V.M. Roundup® exposure promotes gills and liver impairments, DNA damage and inhibition of brain cholinergic activity in the Amazon teleost fish Colossoma macropomum. Chemosphere 2015, 135, 53–60. [Google Scholar] [CrossRef] [PubMed]
  113. Mercurio, P.; Flores, F.; Mueller, J.F.; Carter, S.; Negri, A.P. Glyphosate persistence in seawater. Mar. Pollut. Bull. 2014, 85, 385–390. [Google Scholar] [CrossRef]
  114. Horth, H.; Blackmore, K. Survey of glyphosate and AMPA in groundwaters and surface waters in Europe. WRC Rep. No. UC8073 2009, 2. [Google Scholar]
  115. Newton, M.; Horner, L.M.; Cowell, J.E.; White, D.E.; Cole, E.C. Dissipation of glyphosate and aminomethylphosphonic acid in North American forests. J. Agric. Food Chem. 1994, 42, 1795–1802. [Google Scholar] [CrossRef]
  116. Mateos-Naranjo, E.; Perez-Martin, A. Effects of sub-lethal glyphosate concentrations on growth and photosynthetic performance of non-target species Bolboschoenus maritimus. Chemosphere 2013, 93, 2631–2638. [Google Scholar] [CrossRef]
  117. Zobiole, L.H.; Oliveira, R.S.; Visentainer, J.V.; Kremer, R.J.; Bellaloui, N.; Yamada, T. Glyphosate affects seed composition in glyphosate-resistant soybean. J. Agric. Food Chem. 2010, 58, 4517–4522. [Google Scholar] [CrossRef]
  118. Bott, S.; Tesfamariam, T.; Candan, H.; Cakmak, I.; Römheld, V.; Neumann, G. Glyphosate-induced impairment of plant growth and micronutrient status in glyphosate-resistant soybean (Glycine max L.). Plant Soil 2008, 312, 185–194. [Google Scholar] [CrossRef]
  119. Alister, C.; Kogan, M.; Pino, I. Differential phytotoxicity of glyphosate in maize seedlings following applications to roots or shoot. Weed Res. 2005, 45, 27–32. [Google Scholar] [CrossRef]
  120. Jarrell, Z.R.; Ahammad, M.U.; Benson, A.P. Glyphosate-based herbicide formulations and reproductive toxicity in animals. Vet. Anim. Sci. 2020, 10, 100126. [Google Scholar] [CrossRef] [PubMed]
  121. Mensah, P.; Muller, W.; Palmer, C. Using growth measures in the freshwater shrimp Caridina nilotica as biomarkers of Roundup® pollution of South African freshwater systems. Phys. Chem. Earth Parts A/B/C 2012, 50, 262–268. [Google Scholar] [CrossRef]
  122. Krüger, M.; Schrödl, W.; Neuhaus, J.; Shehata, A. Field investigations of glyphosate in urine of Danish dairy cows. J. Environ. Anal. Toxicol 2013, 3, 100186. [Google Scholar]
  123. Honeycutt, Z.; Rowlands, H. Glyphosate testing report: Findings in American mothers’ breast milk, urine and water. Unpubl. Rep. 2014, 7. Available online: https://www.gentechvrij.nl/wp-content/uploads/2018/09/Gly-testing-rapport-MOMS-18.pdf (accessed on 21 July 2022).
  124. Niemann, L.; Sieke, C.; Pfeil, R.; Solecki, R. A critical review of glyphosate findings in human urine samples and comparison with the exposure of operators and consumers. J. Für Verbrauch. Und Lebensm. 2015, 10, 3–12. [Google Scholar] [CrossRef]
  125. Pedersen, I. An on farm study showing deformities, abortions, and fertility problems in pigs linked to GM soya and Roundup. J. Environ. Anal. Toxicol 2014, 4, 2161-0525.1000230. [Google Scholar]
  126. Nørgaard, K.B.; Cedergreen, N. Pesticide cocktails can interact synergistically on aquatic crustaceans. Environ. Sci. Pollut. Res. 2010, 17, 957–967. [Google Scholar] [CrossRef]
  127. Bjergager, M.-B.A.; Hanson, M.L.; Lissemore, L.; Henriquez, N.; Solomon, K.R.; Cedergreen, N. Synergy in microcosms with environmentally realistic concentrations of prochloraz and esfenvalerate. Aquat. Toxicol. 2011, 101, 412–422. [Google Scholar] [CrossRef]
  128. Battaglin, W.A.; Kolpin, D.W.; Scribner, E.A.; Kuivila, K.M.; Sandstrom, M.W. Glyphosate, other herbicides, and transformation products in midwestern streams, 20021. JAWRA J. Am. Water Resour. Assoc. 2005, 41, 323–332. [Google Scholar] [CrossRef]
  129. Mahler, B.J.; Van Metre, P.C.; Burley, T.E.; Loftin, K.A.; Meyer, M.T.; Nowell, L.H. Similarities and differences in occurrence and temporal fluctuations in glyphosate and atrazine in small Midwestern streams (USA) during the 2013 growing season. Sci. Total Environ. 2017, 579, 149–158. [Google Scholar] [CrossRef] [PubMed]
  130. Battaglin, W.A.; Rice, K.C.; Focazio, M.J.; Salmons, S.; Barry, R.X. The occurrence of glyphosate, atrazine, and other pesticides in vernal pools and adjacent streams in Washington, DC, Maryland, Iowa, and Wyoming, 2005–2006. Environ. Monit. Assess. 2009, 155, 281–307. [Google Scholar] [CrossRef] [PubMed]
  131. Coupe, R.H.; Kalkhoff, S.J.; Capel, P.D.; Gregoire, C. Fate and transport of glyphosate and aminomethylphosphonic acid in surface waters of agricultural basins. Pest Manag. Sci. 2012, 68, 16–30. [Google Scholar] [CrossRef]
  132. Humphries, D.; Byrtus, G.; Anderson, A.-M. Glyphosate Residues in Alberta’s Atmospheric Deposition, Soils and Surface Waters; Alberta Environment: Edmonton, AB, Canada, 2005. [Google Scholar]
  133. Aparicio, V.C.; De Gerónimo, E.; Marino, D.; Primost, J.; Carriquiriborde, P.; Costa, J.L. Environmental fate of glyphosate and aminomethylphosphonic acid in surface waters and soil of agricultural basins. Chemosphere 2013, 93, 1866–1873. [Google Scholar] [CrossRef]
  134. Poiger, T.; Buerge, I.J.; Bächli, A.; Müller, M.D.; Balmer, M.E. Occurrence of the herbicide glyphosate and its metabolite AMPA in surface waters in Switzerland determined with on-line solid phase extraction LC-MS/MS. Environ. Sci. Pollut. Res. 2017, 24, 1588–1596. [Google Scholar] [CrossRef] [PubMed]
  135. Rosenbom, A.; Brüsch, W.; Juhler, R.; Ernstsen, V.; Gudmundsson, L.; Plauborg, F.; Grant, R.; Olsen, P. The Danish Pesticide Leaching Assessment Programme. In Monitoring Results May 1999–June 2009; Geological Survey of Denmark and Greenland: Copenhagen, Denmark, 2010; p. 102. [Google Scholar]
  136. Villeneuve, A.; Larroudé, S.; Humbert, J.-F. Herbicide contamination of freshwater ecosystems: Impact on microbial communities. Pestic. Formul. Eff. Fate 2011, 1, 285–2311. [Google Scholar]
  137. Mörtl, M.; Németh, G.; Juracsek, J.; Darvas, B.; Kamp, L.; Rubio, F.; Székács, A. Determination of glyphosate residues in Hungarian water samples by immunoassay. Microchem. J. 2013, 107, 143–151. [Google Scholar] [CrossRef]
  138. Skark, C.; Zullei-Seibert, N.; Schöttler, U.; Schlett, C. The occurrence of glyphosate in surface water. Int. J. Environ. Anal. Chem. 1998, 70, 93–104. [Google Scholar] [CrossRef]
  139. Mesnage, R.; Defarge, N.; De Vendômois, J.S.; Séralini, G. Potential toxic effects of glyphosate and its commercial formulations below regulatory limits. Food Chem. Toxicol. 2015, 84, 133–153. [Google Scholar] [CrossRef] [PubMed]
  140. Morini, R.; Derr, J.F.; Fenner-Crisp, P. Glyphosate: Health Controversy, Benefits and Continuing Debate. 2018. Available online: https://vtechworks.lib.vt.edu/bitstream/handle/10919/85301/SPES-63.pdf (accessed on 12 July 2022).
  141. Ratcliff, A.W.; Busse, M.D.; Shestak, C.J. Changes in microbial community structure following herbicide (glyphosate) additions to forest soils. Appl. Soil Ecol. 2006, 34, 114–124. [Google Scholar] [CrossRef]
  142. Oliveira, A.G.; Telles, L.F.; Hess, R.A.; Mahecha, G.A.; Oliveira, C.A. Effects of the herbicide Roundup on the epididymal region of drakes Anas platyrhynchos. Reprod. Toxicol. 2007, 23, 182–191. [Google Scholar] [CrossRef] [PubMed]
  143. Giesy, J.P.; Dobson, S.; Solomon, K.R. Ecotoxicological risk assessment for Roundup® herbicide. Rev. Environ. Contam. Toxicol. 2000, 167, 35–120. [Google Scholar]
  144. Lugowska, K. The effects of Roundup on gametes and early development of common carp (Cyprinus carpio L.). Fish Physiol. Biochem. 2018, 44, 1109–1117. [Google Scholar] [CrossRef]
  145. Carpenter, J.K.; Monks, J.M.; Nelson, N. The effect of two glyphosate formulations on a small, diurnal lizard (Oligosoma polychroma). Ecotoxicology 2016, 25, 548–554. [Google Scholar] [CrossRef]
  146. Van Bruggen, A.H.; He, M.M.; Shin, K.; Mai, V.; Jeong, K.; Finckh, M.; Morris, J., Jr. Environmental and health effects of the herbicide glyphosate. Sci. Total Environ. 2018, 616, 255–268. [Google Scholar] [CrossRef]
  147. Mesnage, R.; Bernay, B.; Séralini, G.-E. Ethoxylated adjuvants of glyphosate-based herbicides are active principles of human cell toxicity. Toxicology 2013, 313, 122–128. [Google Scholar] [CrossRef]
  148. Meftaul, I.M.; Venkateswarlu, K.; Dharmarajan, R.; Annamalai, P.; Asaduzzaman, M.; Parven, A.; Megharaj, M. Controversies over human health and ecological impacts of glyphosate: Is it to be banned in modern agriculture? Environ. Pollut. 2020, 263, 114372. [Google Scholar] [CrossRef]
  149. Williams, G.M.; Kroes, R.; Munro, I.C. Safety evaluation and risk assessment of the herbicide Roundup and its active ingredient, glyphosate, for humans. Regul. Toxicol. Pharmacol. 2000, 31, 117–165. [Google Scholar] [CrossRef] [PubMed]
  150. Roberts, D.M.; Buckley, N.A.; Mohamed, F.; Eddleston, M.; Goldstein, D.A.; Mehrsheikh, A.; Bleeke, M.S.; Dawson, A.H. A prospective observational study of the clinical toxicology of glyphosate-containing herbicides in adults with acute self-poisoning. Clin. Toxicol. 2010, 48, 129–136. [Google Scholar] [CrossRef] [PubMed]
  151. Gasnier, C.; Dumont, C.; Benachour, N.; Clair, E.; Chagnon, M.-C.; Séralini, G.-E. Glyphosate-based herbicides are toxic and endocrine disruptors in human cell lines. Toxicology 2009, 262, 184–191. [Google Scholar] [CrossRef]
  152. Greim, H.; Saltmiras, D.; Mostert, V.; Strupp, C. Evaluation of carcinogenic potential of the herbicide glyphosate, drawing on tumor incidence data from fourteen chronic/carcinogenicity rodent studies. Crit. Rev. Toxicol. 2015, 45, 185–208. [Google Scholar] [CrossRef] [PubMed]
  153. Prasad, S.; Srivastava, S.; Singh, M.; Shukla, Y. Clastogenic effects of glyphosate in bone marrow cells of Swiss albino mice. J. Toxicol. 2009, 2009, 308985. [Google Scholar] [CrossRef]
  154. Koller, V.J.; Fürhacker, M.; Nersesyan, A.; Mišík, M.; Eisenbauer, M.; Knasmueller, S. Cytotoxic and DNA-damaging properties of glyphosate and Roundup in human-derived buccal epithelial cells. Arch. Toxicol. 2012, 86, 805–813. [Google Scholar] [CrossRef]
  155. Poletta, G.; Larriera, A.; Kleinsorge, E.; Mudry, M. Genotoxicity of the herbicide formulation Roundup® (glyphosate) in broad-snouted caiman (Caiman latirostris) evidenced by the Comet assay and the Micronucleus test. Mutat. Res. Genet. Toxicol. Environ. Mutagenesis 2009, 672, 95–102. [Google Scholar] [CrossRef]
  156. Kier, L.D.; Kirkland, D.J. Review of genotoxicity studies of glyphosate and glyphosate-based formulations. Crit. Rev. Toxicol. 2013, 43, 283–315. [Google Scholar] [CrossRef]
  157. Dallegrave, E.; Mantese, F.D.; Oliveira, R.T.; Andrade, A.J.; Dalsenter, P.R.; Langeloh, A. Pre-and postnatal toxicity of the commercial glyphosate formulation in Wistar rats. Arch. Toxicol. 2007, 81, 665–673. [Google Scholar] [CrossRef]
  158. Romano, M.A.; Romano, R.M.; Santos, L.D.; Wisniewski, P.; Campos, D.A.; de Souza, P.B.; Viau, P.; Bernardi, M.M.; Nunes, M.T.; de Oliveira, C.A. Glyphosate impairs male offspring reproductive development by disrupting gonadotropin expression. Arch. Toxicol. 2012, 86, 663–673. [Google Scholar] [CrossRef]
  159. George, J.; Prasad, S.; Mahmood, Z.; Shukla, Y. Studies on glyphosate-induced carcinogenicity in mouse skin: A proteomic approach. J. Proteom. 2010, 73, 951–964. [Google Scholar] [CrossRef] [PubMed]
  160. Thongprakaisang, S.; Thiantanawat, A.; Rangkadilok, N.; Suriyo, T.; Satayavivad, J. Glyphosate induces human breast cancer cells growth via estrogen receptors. Food Chem. Toxicol. 2013, 59, 129–136. [Google Scholar] [CrossRef] [PubMed]
  161. Fritschi, L.; McLaughlin, J.; Sergi, C.; Calaf, G.; Le Curieux, F.; Forastiere, F.; Kromhout, H.; Egeghy, P.; Jahnke, G.; Jameson, C. Carcinogenicity of tetrachlorvinphos, parathion, malathion, diazinon, and glyphosate. Red 2015, 114, 70134–70138. [Google Scholar]
  162. Hohenadel, K.; Harris, S.A.; McLaughlin, J.R.; Spinelli, J.J.; Pahwa, P.; Dosman, J.A.; Demers, P.A.; Blair, A. Exposure to multiple pesticides and risk of non-Hodgkin lymphoma in men from six Canadian provinces. Int. J. Environ. Res. Public Health 2011, 8, 2320–2330. [Google Scholar] [CrossRef]
  163. Zhang, L.; Rana, I.; Shaffer, R.M.; Taioli, E.; Sheppard, L. Exposure to glyphosate-based herbicides and risk for non-Hodgkin lymphoma: A meta-analysis and supporting evidence. Mutat. Res. Rev. Mutat. Res. 2019, 1, 186–206. [Google Scholar] [CrossRef]
  164. Watnick, V.J. The Lautenberg Chemical Safety Act of 2016: Cancer, industry pressure, and a proactive approach. Harv. Envtl. L. Rev. 2019, 43, 373. [Google Scholar]
  165. IARC. IARC Monographs on the Evaluation of Carcinogen Risks to Humans. J. Clin. Pathol. 2015, 112, 691. Available online: https://www.thelancet.com/pdfs/journals/lanonc/PIIS1470-2045%2815%2970134-8.pdf (accessed on 3 July 2022).
  166. Goguen, D. Roundup Cancer Lawsuits. Available online: https://www.nolo.com/legal-encyclopedia/roundup-cancer-lawsuits.html#:~:text=In%20May%202019%2C%20a%20California,%2455%20million%20in%20compensatory%20damages (accessed on 3 July 2022).
  167. OEHHA. Initial Statement of Reasons Title 27, California Code of Regulations Proposed Amendments to Article 6 Clear and Reasonable Warnings New Sections 25607.48 and 25607.49 Warnings for Exposures to Glyphosate from Consumer Products. 2021. Available online: https://oehha.ca.gov/media/downloads/crnr/glyphosateisor071921.pdf (accessed on 3 July 2022).
  168. Chaufan, G.; Coalova, I.; Del Carmen Ríos de Molina, M. Glyphosate commercial formulation causes cytotoxicity, oxidative effects, and apoptosis on human cells: Differences with its active ingredient. Int. J. Toxicol. 2014, 33, 29–38. [Google Scholar] [CrossRef]
  169. Colovic, M.B.; Krstic, D.Z.; Lazarevic-Pasti, T.D.; Bondzic, A.M.; Vasic, V.M. Acetylcholinesterase inhibitors: Pharmacology and toxicology. Curr. Neuropharmacol. 2013, 11, 315–335. [Google Scholar] [CrossRef] [PubMed]
  170. Coullery, R.P.; Ferrari, M.E.; Rosso, S.B. Neuronal development and axon growth are altered by glyphosate through a WNT non-canonical signaling pathway. Neurotoxicology 2016, 52, 150–161. [Google Scholar] [CrossRef]
  171. Kwiatkowska, M.; Reszka, E.; Woźniak, K.; Jabłońska, E.; Michałowicz, J.; Bukowska, B. DNA damage and methylation induced by glyphosate in human peripheral blood mononuclear cells (in vitro study). Food Chem. Toxicol. 2017, 105, 93–98. [Google Scholar] [CrossRef] [PubMed]
  172. Martini, C.N.; Gabrielli, M.; Codesido, M.M.; Del Vila, M.C. Glyphosate-based herbicides with different adjuvants are more potent inhibitors of 3T3-L1 fibroblast proliferation and differentiation to adipocytes than glyphosate alone. Comp. Clin. Pathol. 2016, 25, 607–613. [Google Scholar] [CrossRef]
  173. Springett, J.; Gray, R. Effect of repeated low doses of biocides on the earthworm Aporrectodea caliginosa in laboratory culture. Soil Biol. Biochem. 1992, 24, 1739–1744. [Google Scholar] [CrossRef]
  174. Tate, T.; Spurlock, J.; Christian, F. Effect of glyphosate on the development of Pseudosuccinea columella snails. Arch. Environ. Contam. Toxicol. 1997, 33, 286–289. [Google Scholar] [CrossRef]
  175. Balbuena, M.S.; Tison, L.; Hahn, M.-L.; Greggers, U.; Menzel, R.; Farina, W.M. Effects of sublethal doses of glyphosate on honeybee navigation. J. Exp. Biol. 2015, 218, 2799–2805. [Google Scholar] [CrossRef] [PubMed]
  176. Druart, C.; Millet, M.; Scheifler, R.; Delhomme, O.; De Vaufleury, A. Glyphosate and glufosinate-based herbicides: Fate in soil, transfer to, and effects on land snails. J. Soils Sediments 2011, 11, 1373–1384. [Google Scholar] [CrossRef]
  177. de Freitas Bueno, A.; de Freitas Bueno, R.C.O.; Parra, J.R.P.; Vieira, S.S. Effects of pesticides used in soybean crops to the egg parasitoid Trichogramma pretiosum. Ciência Rural. 2008, 38, 1495–1503. [Google Scholar] [CrossRef]
  178. Boily, M.; Sarrasin, B.; DeBlois, C.; Aras, P.; Chagnon, M. Acetylcholinesterase in honey bees (Apis mellifera) exposed to neonicotinoids, atrazine and glyphosate: Laboratory and field experiments. Environ. Sci. Pollut. Res. 2013, 20, 5603–5614. [Google Scholar] [CrossRef]
  179. Herbert, L.T.; Vázquez, D.E.; Arenas, A.; Farina, W.M. Effects of field-realistic doses of glyphosate on honeybee appetitive behaviour. J. Exp. Biol. 2014, 217, 3457–3464. [Google Scholar] [CrossRef] [PubMed]
  180. Yadav, S.S.; Giri, S.; Singha, U.; Boro, F.; Giri, A. Toxic and genotoxic effects of Roundup on tadpoles of the Indian skittering frog (Euflictis cyanophlyctis) in the presence and absence of predator stress. Aquat. Toxicol. 2013, 132–133, 1–8. [Google Scholar] [CrossRef] [PubMed]
  181. Lajmanovich, R.C.; Attademo, A.M.; Simoniello, M.F.; Poletta, G.L.; Junges, C.M.; Peltzer, P.M.; Grenón, P.; Cabagna-Zenklusen, M.C. Harmful effects of the dermal intake of commercial formulations containing chlorpyrifos, 2,4-D, and glyphosate on the common toad Rhinella arenarum (Anura: Bufonidae). Water Air Soil Pollut. 2015, 226, 1–12. [Google Scholar] [CrossRef]
  182. Dornelles, M.; Oliveira, G. Toxicity of atrazine, glyphosate, and quinclorac in bullfrog tadpoles exposed to concentrations below legal limits. Environ. Sci. Pollut. Res. 2016, 23, 1610–1620. [Google Scholar] [CrossRef] [PubMed]
  183. Soloneski, S.; Ruiz de Arcaute, C.; Larramendy, M.L. Genotoxic effect of a binary mixture of dicamba-and glyphosate-based commercial herbicide formulations on Rhinella arenarum (Hensel, 1867) (Anura, Bufonidae) late-stage larvae. Environ. Sci. Pollut. Res. 2016, 23, 17811–17821. [Google Scholar] [CrossRef]
  184. Pérez-Iglesias, J.M.; Franco-Belussi, L.; Moreno, L.; Tripole, S.; de Oliveira, C.; Natale, G.S. Effects of glyphosate on hepatic tissue evaluating melanomacrophages and erythrocytes responses in neotropical anuran Leptodactylus latinasus. Environ. Sci. Pollut. Res. 2016, 23, 9852–9861. [Google Scholar] [CrossRef]
  185. Alberdi, J.; Sáenz, M.; Di Marzio, W.; Tortorelli, M. Comparative acute toxicity of two herbicides, paraquat and glyphosate, to Daphnia magna and D. spinulata. Bull. Environ. Contam. Toxicol. 1996, 57, 229–235. [Google Scholar] [CrossRef]
  186. Uren Webster, T.M.; Laing, L.V.; Florance, H.; Santos, E.M. Effects of glyphosate and its formulation, roundup, on reproduction in zebrafish (Danio rerio). Environ. Sci. Technol. 2014, 48, 1271–1279. [Google Scholar] [CrossRef]
  187. Santillo, D.J.; Brown, P.W.; Leslie, D.M., Jr. Response of songbirds to glyphosate-induced habitat changes on clearcuts. J. Wildl. Manag. 1989, 53, 64–71. [Google Scholar] [CrossRef]
  188. Linz, G.M.; Blixt, D.C.; Bergman, D.L.; Bleier, W.J. Responses of red-winged blackbirds, yellow-headed blackbirds and marsh wrens to glyphosate-induced alterations in cattail density (Respuesta de Agelaius phoeniceus, Xanthocephalus xanthocephalus y Cistothorus palustris a Alteración en la Densidad de Eneas Tratadas con Yerbicidas. J. Field Ornithol. 1996, 67, 167–176. [Google Scholar]
  189. Fan, J.; Geng, J.; Ren, H.; Wang, X. Time-effect relationship of toxicity induced by Roundup® and its main constituents in liver of Carassius auratus. Comput. Water Energy Environ. Eng. 2013, 2, 20. [Google Scholar] [CrossRef]
  190. Schaumburg, L.G.; Siroski, P.A.; Poletta, G.L.; Mudry, M.D. Genotoxicity induced by Roundup®(Glyphosate) in tegu lizard (Salvator merianae) embryos. Pestic. Biochem. Physiol. 2016, 130, 71–78. [Google Scholar] [CrossRef] [PubMed]
  191. Siroski, P.A.; Poletta, G.L.; Latorre, M.A.; Merchant, M.E.; Ortega, H.H.; Mudry, M.D. Immunotoxicity of commercial-mixed glyphosate in broad snouted caiman (Caiman latirostris). Chem. Biol. Interact. 2016, 244, 64–70. [Google Scholar]
  192. Nwani, C.; Ibiam, U.; Ibiam, O.; Nworie, O.; Onyishi, G.; Atama, C. Investigation on Acute toxicity and behavioral changes in Tilapia zillii due to glyphosate-based herbicide, forceup. J. Anim. Plant Sci. 2013, 23, 888–892. [Google Scholar]
  193. Tizhe, E.V.; Ibrahim, N.D.-G.; Fatihu, M.Y.; Igbokwe, I.O.; George, B.-D.J.; Ambali, S.F.; Shallangwa, J.M. Serum biochemical assessment of hepatic and renal functions of rats during oral exposure to glyphosate with zinc. Comp. Clin. Pathol. 2014, 23, 1043–1050. [Google Scholar] [CrossRef] [PubMed]
  194. Tizhe, E.V.; Ibrahim, N.D.-G.; Fatihu, M.Y.; Onyebuchi, I.I.; George, B.D.J.; Ambali, S.F.; Shallangwa, J.M. Influence of zinc supplementation on histopathological changes in the stomach, liver, kidney, brain, pancreas and spleen during subchronic exposure of Wistar rats to glyphosate. Comp. Clin. Pathol. 2014, 23, 1535–1543. [Google Scholar] [CrossRef]
  195. Murussi, C.R.; Costa, M.D.; Leitemperger, J.W.; Guerra, L.; Rodrigues, C.C.; Menezes, C.C.; Severo, E.S.; Flores-Lopes, F.; Salbego, J.; Loro, V.L. Exposure to different glyphosate formulations on the oxidative and histological status of Rhamdia quelen. Fish Physiol. Biochem. 2016, 42, 445–455. [Google Scholar]
  196. Lee, H.-L.; Kan, C.-D.; Tsai, C.-L.; Liou, M.-J.; Guo, H.-R. Comparative effects of the formulation of glyphosate-surfactant herbicides on hemodynamics in swine. Clin. Toxicol. 2009, 47, 651–658. [Google Scholar] [CrossRef]
  197. Alferness, P.; Wiebe, L. Determination of mesotrione residues and metabolites in crops, soil, and water by liquid chromatography with fluorescence detection. J. Agric. Food Chem. 2002, 50, 3926–3934. [Google Scholar] [CrossRef]
  198. Caloni, F.; Cortinovis, C.; Rivolta, M.; Davanzo, F. Suspected poisoning of domestic animals by pesticides. Sci. Total Environ. 2016, 539, 331–336. [Google Scholar] [CrossRef]
  199. Gillezeau, C.; Lieberman-Cribbin, W.; Taioli, E. Update on human exposure to glyphosate, with a complete review of exposure in children. Environ. Health 2020, 19, 1–8. [Google Scholar] [CrossRef] [PubMed]
  200. Martinez, T.; Long, W.; Hiller, R. Comparison of the toxicology of the herbicide Roundup by oral and pulmonary routes of exposure. Proc. West. Pharmacol. Soc. 1990, 33, 193–197. [Google Scholar]
  201. Connolly, A.; Coggins, M.A.; Galea, K.S.; Jones, K.; Kenny, L.; McGowan, P.; Basinas, I. Evaluating glyphosate exposure routes and their contribution to total body burden: A study among amenity horticulturalists. Ann. Work. Expo. Health 2019, 63, 133–147. [Google Scholar] [CrossRef] [PubMed]
  202. Vandenberg, L.N.; Blumberg, B.; Antoniou, M.N.; Benbrook, C.M.; Carroll, L.; Colborn, T.; Everett, L.G.; Hansen, M.; Landrigan, P.J.; Lanphear, B.P. Is it time to reassess current safety standards for glyphosate-based herbicides? J. Epidemiol. Community Health 2017, 71, 613–618. [Google Scholar] [CrossRef]
  203. Health Canada; Canadian Pest Management Regulatory Agency. Statement from Health Canada—Final Re-Evaluation Decision on Glyphosate. 28 April 2017. Available online: https://www.canada.ca/en/health-canada/news/2017/04/statement_from_healthcanadafinalre-evaluationdecisiononglyphosat.html?wbdisable=true (accessed on 21 July 2022).
  204. EFSA Scientific Committee; Benford, D.; Halldorsson, T.; Jeger, M.J.; Knutsen, H.K.; More, S.; Naegeli, H.; Noteborn, H.; Ockleford, C.; Ricci, A.; et al. Guidance on uncertainty analysis in scientific assessments. EFSA J. 2018, 16, e05123. [Google Scholar]
  205. McHenry, L.B. The Monsanto Papers: Poisoning the scientific well. Int. J. Risk Saf. Med. 2018, 29, 193–205. [Google Scholar] [CrossRef]
  206. Peillex, C.; Pelletier, M. The impact and toxicity of glyphosate and glyphosate-based herbicides on health and immunity. J. Immunotoxicol. 2020, 17, 163–174. [Google Scholar] [CrossRef] [PubMed]
  207. Secombes, C.; Wang, T. The innate and adaptive immune system of fish. In Infectious Disease in Aquaculture; Elsevier: Amsterdam, The Netherlands, 2012; pp. 3–68. [Google Scholar]
  208. Kim, Y.-h.; Hong, J.-r.; Gil, H.-w.; Song, H.-y.; Hong, S.-y. Mixtures of glyphosate and surfactant TN20 accelerate cell death via mitochondrial damage-induced apoptosis and necrosis. Toxicol. Vitr. 2013, 27, 191–197. [Google Scholar] [CrossRef] [PubMed]
  209. Tsatsakis, A.; Docea, A.O.; Constantin, C.; Calina, D.; Zlatian, O.; Nikolouzakis, T.K.; Stivaktakis, P.D.; Kalogeraki, A.; Liesivuori, J.; Tzanakakis, G. Genotoxic, cytotoxic, and cytopathological effects in rats exposed for 18 months to a mixture of 13 chemicals in doses below NOAEL levels. Toxicol. Lett. 2019, 316, 154–170. [Google Scholar] [CrossRef]
  210. Monroy, C.M.; Cortés, A.C.; Sicard, D.M.; de Restrepo, H.G. Cytotoxicity and genotoxicity of human cells exposed in vitro to glyphosate. Biomedica 2005, 25, 335–345. [Google Scholar] [CrossRef]
  211. Vanlaeys, A.; Dubuisson, F.; Seralini, G.-E.; Travert, C. Formulants of glyphosate-based herbicides have more deleterious impact than glyphosate on TM4 Sertoli cells. Toxicol. Vitr. 2018, 52, 14–22. [Google Scholar] [CrossRef] [PubMed]
  212. Alvarez-Moya, C.; Reynoso Silva, M.; Valdez Ramírez, C.; Gómez Gallardo, D.; León Sánchez, R.; Canales Aguirre, A.; Feria Velasco, A. Comparison of the in vivo and in vitro genotoxicity of glyphosate isopropylamine salt in three different organisms. Genet. Mol. Biol. 2014, 37, 105–110. [Google Scholar] [CrossRef] [PubMed]
  213. Richard, S.; Moslemi, S.; Sipahutar, H.; Benachour, N.; Seralini, G.-E. Differential effects of glyphosate and roundup on human placental cells and aromatase. Environ. Health Perspect. 2005, 113, 716–720. [Google Scholar] [CrossRef] [PubMed]
  214. Benachour, N.; Sipahutar, H.; Moslemi, S.; Gasnier, C.; Travert, C.; Seralini, G. Time-and dose-dependent effects of roundup on human embryonic and placental cells. Arch. Environ. Contam. Toxicol. 2007, 53, 126–133. [Google Scholar] [CrossRef] [PubMed]
  215. Mesnage, R.; Phedonos, A.; Biserni, M.; Arno, M.; Balu, S.; Corton, J.C.; Ugarte, R.; Antoniou, M.N. Evaluation of estrogen receptor alpha activation by glyphosate-based herbicide constituents. Food Chem. Toxicol. 2017, 108, 30–42. [Google Scholar] [CrossRef]
  216. Young, F.; Ho, D.; Glynn, D.; Edwards, V. Endocrine disruption and cytotoxicity of glyphosate and roundup in human JAr cells in vitro. Synthesis 2015, 14, 17. [Google Scholar]
  217. Arbuckle, T.E.; Lin, Z.; Mery, L.S. An exploratory analysis of the effect of pesticide exposure on the risk of spontaneous abortion in an Ontario farm population. Environ. Health Perspect. 2001, 109, 851–857. [Google Scholar] [CrossRef]
  218. Benachour, N.; Séralini, G.-E. Glyphosate formulations induce apoptosis and necrosis in human umbilical, embryonic, and placental cells. Chem. Res. Toxicol. 2009, 22, 97–105. [Google Scholar] [CrossRef]
  219. Dallegrave, E.; Mantese, F.D.; Coelho, R.S.; Pereira, J.D.; Dalsenter, P.R.; Langeloh, A. The teratogenic potential of the herbicide glyphosate-Roundup® in Wistar rats. Toxicol. Lett. 2003, 142, 45–52. [Google Scholar] [CrossRef]
  220. Antoniou, M.; Habib, M.; Howard, C.; Jennings, R.; Leifert, C.; Nodari, R.; Robinson, C.; Fagan, J. Teratogenic effects of glyphosate-based herbicides: Divergence of regulatory decisions from scientific evidence. J. Environ. Anal. Toxicol. S 2012, 4, 2161-0525. [Google Scholar]
  221. Roy, N.M.; Ochs, J.; Zambrzycka, E.; Anderson, A. Glyphosate induces cardiovascular toxicity in Danio rerio. Environ. Toxicol. Pharmacol. 2016, 46, 292–300. [Google Scholar] [CrossRef] [PubMed]
  222. Campana, H.; Pawluk, M.S.; Camelo, J.S.L. Births prevalence of 27 selected congenital anomalies in 7 geographic regions of Argentina. Arch. Argent. De Pediatr. 2010, 108, 409–417. [Google Scholar]
  223. Sato, C.; Kamijo, Y.; Yoshimura, K.; Ide, T. Aseptic meningitis in association with glyphosate-surfactant herbicide poisoning. Clin. Toxicol. 2011, 49, 118–120. [Google Scholar] [CrossRef] [PubMed]
  224. Kawagashira, Y.; Koike, H.; Kawabata, K.; Takahashi, M.; Ohyama, K.; Hashimoto, R.; Iijima, M.; Katsuno, M.; Sobue, G. Vasculitic neuropathy following exposure to a glyphosate-based herbicide. Intern. Med. 2017, 56, 1431–1434. [Google Scholar] [CrossRef] [PubMed]
  225. Gallegos, C.E.; Bartos, M.; Bras, C.; Gumilar, F.; Antonelli, M.C.; Minetti, A. Exposure to a glyphosate-based herbicide during pregnancy and lactation induces neurobehavioral alterations in rat offspring. Neurotoxicology 2016, 53, 20–28. [Google Scholar] [CrossRef]
  226. Barbosa, E.R.; Leiros da Costa, M.D.; Bacheschi, L.A.; Scaff, M.; Leite, C.C. Parkinsonism after glycine-derivate exposure. Mov. Disord. Off. J. Mov. Disord. Soc. 2001, 16, 565–568. [Google Scholar] [CrossRef]
  227. Kamel, F.; Tanner, C.; Umbach, D.; Hoppin, J.; Alavanja, M.; Blair, A.; Comyns, K.; Goldman, S.; Korell, M.; Langston, J. Pesticide exposure and self-reported Parkinson’s disease in the agricultural health study. Am. J. Epidemiol. 2007, 165, 364–374. [Google Scholar] [CrossRef]
  228. Zheng, Q.; Yin, J.; Zhu, L.; Jiao, L.; Xu, Z. Reversible Parkinsonism induced by acute exposure glyphosate. Parkinsonism Relat. Disord. 2018, 50, 121. [Google Scholar] [CrossRef]
  229. Eriguchi, M.; Iida, K.; Ikeda, S.; Osoegawa, M.; Nishioka, K.; Hattori, N.; Nagayama, H.; Hara, H. Parkinsonism relating to intoxication with glyphosate: A case report. Intern. Med. 2019, 58, 1935–1938. [Google Scholar] [CrossRef]
  230. Ackermann, W.; Coenen, M.; Schrödl, W.; Shehata, A.A.; Krüger, M. The influence of glyphosate on the microbiota and production of botulinum neurotoxin during ruminal fermentation. Curr. Microbiol. 2015, 70, 374–382. [Google Scholar] [CrossRef]
  231. Schrödl, W.; Krüger, S.; Konstantinova-Müller, T.; Shehata, A.A.; Rulff, R.; Krüger, M. Possible effects of glyphosate on Mucorales abundance in the rumen of dairy cows in Germany. Curr. Microbiol. 2014, 69, 817–823. [Google Scholar] [CrossRef] [PubMed]
  232. Topal, A.; Atamanalp, M.; Uçar, A.; Oruç, E.; Kocaman, E.M.; Sulukan, E.; Akdemir, F.; Beydemir, Ş.; Kılınç, N.; Erdoğan, O. Effects of glyphosate on juvenile rainbow trout (Oncorhynchus mykiss): Transcriptional and enzymatic analyses of antioxidant defence system, histopathological liver damage and swimming performance. Ecotoxicol. Environ. Saf. 2015, 111, 206–214. [Google Scholar] [CrossRef] [PubMed]
  233. Pandey, A.; Dabhade, P.; Kumarasamy, A. Inflammatory effects of subacute exposure of Roundup in rat liver and adipose tissue. Dose-Response 2019, 17, 1559325819843380. [Google Scholar] [CrossRef] [Green Version]
  234. Mesnage, R.; Antoniou, M.N. Ignoring adjuvant toxicity falsifies the safety profile of commercial pesticides. Front. Public Health 2018, 5, 361. [Google Scholar] [CrossRef] [PubMed]
  235. Qiu, S.; Fu, H.; Zhou, R.; Yang, Z.; Bai, G.; Shi, B. Toxic effects of glyphosate on intestinal morphology, antioxidant capacity and barrier function in weaned piglets. Ecotoxicol. Environ. Saf. 2020, 187, 109846. [Google Scholar] [CrossRef]
  236. Montgomery, M.; Kamel, F.; Saldana, T.; Alavanja, M.; Sandler, D. Incident diabetes and pesticide exposure among licensed pesticide applicators: Agricultural Health Study, 1993–2003. Am. J. Epidemiol. 2008, 167, 1235–1246. [Google Scholar] [CrossRef]
  237. Trasande, L.; Aldana, S.I.; Trachtman, H.; Kannan, K.; Morrison, D.; Christakis, D.A.; Whitlock, K.; Messito, M.J.; Gross, R.S.; Karthikraj, R. Glyphosate exposures and kidney injury biomarkers in infants and young children. Environ. Pollut. 2020, 256, 113334. [Google Scholar] [CrossRef]
  238. Wang, L.; Deng, Q.; Hu, H.; Liu, M.; Gong, Z.; Zhang, S.; Xu-Monette, Z.Y.; Lu, Z.; Young, K.H.; Ma, X. Glyphosate induces benign monoclonal gammopathy and promotes multiple myeloma progression in mice. J. Hematol. Oncol. 2019, 12, 1–11. [Google Scholar] [CrossRef]
  239. Chan, Y.-C.; Chang, S.-C.; Hsuan, S.-L.; Chien, M.-S.; Lee, W.-C.; Kang, J.-J.; Wang, S.-C.; Liao, J.-W. Cardiovascular effects of herbicides and formulated adjuvants on isolated rat aorta and heart. Toxicol. Vitr. 2007, 21, 595–603. [Google Scholar] [CrossRef]
  240. Zouaoui, K.; Dulaurent, S.; Gaulier, J.; Moesch, C.; Lachâtre, G. Determination of glyphosate and AMPA in blood and urine from humans: About 13 cases of acute intoxication. Forensic Sci. Int. 2013, 226, e20–e25. [Google Scholar] [CrossRef]
  241. USEPA. Glyphosate Interim Registration Review Decision Case Number 0178, January 2020. Docket Number EPA-HQ-OPP-2009-0361. Available online: https://www.epa.gov/sites/default/files/2020-01/documents/glyphosate-interim-reg-review-decision-case-num-0178.pdf (accessed on 12 July 2022).
  242. Xu, J.; Smith, S.; Smith, G.; Wang, W.; Li, Y. Glyphosate contamination in grains and foods: An overview. Food Control 2019, 106, 106710. [Google Scholar] [CrossRef]
  243. Code of Federal Regulations. Title 40, 180.364 Glyphosate: Tolerances for Residu. 2022. Available online: https://www.ecfr.gov/current/title-40/chapter-I/subchapter-E/part-180/subpart-C/section-180.364 (accessed on 19 June 2022).
  244. Analytica, O. US appeals court ruling is new setback for herbicide. Emerald Expert Brief. 2021. [Google Scholar] [CrossRef]
  245. Cotruvo, J.A.; Regelski, M. National primary drinking water regulations for volatile organic chemicals. In Safe Drinking Water Act (1989); CRC: Boca Raton, FL, USA, 2017; Volume 29. [Google Scholar]
  246. Nhmrc, N. Australian Drinking Water Guidelines Paper 6 National Water Quality Management Strategy; National Health and Medical Research Council, National Resource Management Ministerial Council, Commonwealth of Australia: Canberra, Australia, 2011; pp. 5–7. [Google Scholar]
  247. European Food Safety Authority (EFSA). Review of the existing maximum residue levels for glyphosate according to Article 12 of Regulation (EC) No 396/2005–revised version to take into account omitted data. EFSA J. 2019, 17, e05862. [Google Scholar]
  248. Health Canada. Proposed Maximum Residue Limit PMRL2021-10, Glyphosate. 2021. Available online: https://www.canada.ca/en/health-canada/services/consumer-product-safety/pesticides-pest-management/public/consultations/proposed-maximum-residue-limit/2021/glyphosate/document.html (accessed on 19 June 2022).
Figure 1. Biodegradation pathways of glyphosate in the environment [14].
Figure 1. Biodegradation pathways of glyphosate in the environment [14].
Applsci 12 08789 g001
Figure 2. Pathways of glyphosate-based formulations in human health and the environment [148].
Figure 2. Pathways of glyphosate-based formulations in human health and the environment [148].
Applsci 12 08789 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ojelade, B.S.; Durowoju, O.S.; Adesoye, P.O.; Gibb, S.W.; Ekosse, G.-I. Review of Glyphosate-Based Herbicide and Aminomethylphosphonic Acid (AMPA): Environmental and Health Impacts. Appl. Sci. 2022, 12, 8789. https://doi.org/10.3390/app12178789

AMA Style

Ojelade BS, Durowoju OS, Adesoye PO, Gibb SW, Ekosse G-I. Review of Glyphosate-Based Herbicide and Aminomethylphosphonic Acid (AMPA): Environmental and Health Impacts. Applied Sciences. 2022; 12(17):8789. https://doi.org/10.3390/app12178789

Chicago/Turabian Style

Ojelade, Babatunde Solomon, Olatunde Samod Durowoju, Peter Oluremi Adesoye, Stuart W. Gibb, and Georges-Ivo Ekosse. 2022. "Review of Glyphosate-Based Herbicide and Aminomethylphosphonic Acid (AMPA): Environmental and Health Impacts" Applied Sciences 12, no. 17: 8789. https://doi.org/10.3390/app12178789

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