ArticlePDF Available

Toxicity of the herbicides used on herbicide-tolerant crops, and societal consequences of their use in France

Authors:

Abstract

In France, the implementation of mutant herbicide-tolerant crops and the use of the related herbicides – sulfonylureas and imidazolinones – have triggered a strong societal reaction illustrated by the intervening actions of environmentalist groups illegally mowing such crops. Trials are in progress, and therefore should be addressed the questions of the environmental risks and the toxicity of these herbicides for the animals and humans consuming the products derived from these plants. Regulatory authorities have allowed these mutant and herbicide-tolerant plants arguing that the herbicides against which they resist only target an enzyme found in ‘weeds’ (the acetolactate synthase, ALS), and that therefore all organisms lacking this enzyme would be endowed with immunity to these herbicides. The toxicological literature does not match with this argument: 1) Even in organisms displaying the enzyme ALS, these herbicides impact other molecular targets than ALS; 2) These herbicides are toxic for animals, organisms that do not possess the enzyme ALS, and especially invertebrates, amphibians and fish. In humans, epidemiological studies have shown that the use and handling of these toxins are associated with a significantly increased risk of colon and bladder cancers, and miscarriages. In agricultural soils, these herbicides have a persistence of up to several months, and water samples have concentrations of some of these herbicides above the limit value in drinking water.
!
1!
Toxicity of the herbicides used on herbicide-tolerant crops, and societal
consequences of their use in France.
Jean-Paul Bourdineauda,b
Published in Drug and Chemical Toxicology, accepted the 2nd of May 2020
DOI: 10.1080/01480545.2020.1770781
To link to this article: https://doi.org/10.1080/01480545.2020.1770781
Published online: 16 Jun 2020.
aUniversity of Bordeaux, CNRS, Laboratory of Fundamental Microbiology and
Pathogenicity, European Institute of Chemistry and Biology, 2 rue Robert
Escarpit, 33607 Pessac, France; bCRIIGEN, Paris, France.
(jean-paul.bourdineaud@u-bordeaux.fr).
!
2!
Abstract
In France, the implementation of mutant herbicide-tolerant crops and the
use of the related herbicides sulfonylureas and imidazolinones - have triggered
a strong societal reaction illustrated by the intervening actions of
environmentalist groups illegally mowing such crops. Trials are in progress, and
therefore should be addressed the questions of the environmental risks and the
toxicity of these herbicides for the animals and humans consuming the products
derived from these plants.
Regulatory authorities have allowed these mutant and herbicide-tolerant
plants arguing that the herbicides against which they resist only target an
enzyme found in "weeds" (the acetolactate synthase, ALS), and that therefore all
organisms lacking this enzyme would be endowed with immunity to these
herbicides. The toxicological literature does not match with this argument: 1 /
Even in organisms displaying the enzyme ALS, these herbicides impact other
molecular targets than ALS; 2 / These herbicides are toxic for animals,
organisms that do not possess the enzyme ALS, and especially invertebrates,
amphibians and fish. In humans, epidemiological studies have shown that the
use and handling of these toxins are associated with a significantly increased risk
of colon and bladder cancers, and miscarriages. In agricultural soils, these
herbicides have a persistence of up to several months, and water samples have
concentrations of some of these herbicides above the limit value in drinking
water.
Keywords : herbicides ; sulfonylurea; imidazolinone; herbicide-tolerant plants;
chemical mutagenesis; acetolactate synthase
!
3!
Introduction
The molecules of the imidazolinone (IMI) and sulfonylurea (SUL) chemical
families are used as herbicides. Their primary site of action is the enzyme
acetolactate synthase (ALS), also known as acetohydroxyacid synthase (AHAS),
of which they are inhibitors (the molecules of the triazolopyrimidine sulfonamide,
pyrimidinylthiobenzoate, and sulfonylamino-carbonyltriazolinone chemical
families are also inhibitors of ALS: Mallory-Smith and Retzinger 2003). This
enzyme is involved in the synthesis of the amino acids valine, leucine and
isoleucine (Zhou et al. 2007). The inhibition of this enzyme by this type of
herbicides consequently causes a toxic accumulation of α-ketoglutarate along
with a deficit of protein synthesis in the polluted plants, and in the long term the
inability to grow or death. Bacteria, fungi and yeasts also have this enzyme,
making all of the species represented by these branches of life potential targets
(Boldt and Jacobsen 1998; Desai et al. 2009; Duggleby et al. 2003; Rachedi et
al. 2018). Aquatic plants such as the duckweed Lemna minor are also affected by
these herbicides (Peterson et al. 1994).
The continuous overuse of a single herbicide multiple times in a growing
season is leading to anaturallyacquired resistance to this herbicide. As for the
ALS gene, point mutations making the enzyme, and consequently the plant,
resistant to ALS-inhibiting herbicides are known and detailed not only in many
weeds but also in crops (Kolkman et al. 2004; Powles and Yu 2010; Shimizu et
al. 2005). In 2014, 404 unique cases of herbicide resistant weeds were known.
ALS inhibitor-resistant weeds accounted for about a third of all cases (Heap
2014). Ian Heap has established the most extensive database in the world
displaying the herbicide-resistant weeds to ALS-inhibiting herbicides that can be
consulted in weedscience.org. Ironically, these herbicides are used as an
!
4!
alternative to glyphosate in crops where weeds have become resistant to
Roundup. Sunflowers made resistant to these herbicides were created by
chemical mutagenesis and first patented by Du Pont De Nemours (Gabard and
Huby 2001). Sunflower, rapeseed, soy, wheat, rice and maize resistant to these
herbicides are now marketed and planted (Tan et al. 2005).
The International Maize and Wheat Improvement Center (CIMMYT)
developed through a partnership herbicide tolerant maize lines based on a
natural mutation in maize (Gressel 2009). However, environmentalists consider
that the “naturally” acquired mutations found in such crops are the result of a
selection pressure exerted by sprayed herbicides, so that these field mutants can
be considered as the products of a field-scale chemical mutagenesis. Beside ALS
mutations, plants can to some extent tolerate ALS-inhibiting herbicides through
the activation of detoxification genes or induction of related enzymes such as
cytochrome P450 monooxygenases (Barrett 1995; Pan et al. 2006; Persans and
Schuler 1995; Powles and Yu 2010; Saika et al. 2014; Duhoux and Délye 2013;
Liu et al. 2015; Domínguez-Mendez et al 2017; Yang et al. 2018), glutathione S-
transferases (Liu et al. 2015; Balabanova et al. 2018; Sun et al. 2017 and
2018), glycosyltransferase genes (Liu et al. 2015) and the multidrug resistance-
associated protein MRP1 (Liu et al. 2015; Sun et al. 2018). For instance, a hybrid
maize line is resisting to the SUL foramsulfuron through the intervening action of
a P450 gene (Paporisch and Rubin 2017). However, the metabolization of the
herbicide is not possible without the combined use of another chemical called
“safener” (isoxadifen in the case of SUL herbicides) that protects the plant from
the toxic effects of the herbicide through the activation of a cognate P450 gene
or enzyme (Hatzios and Burgos 2004; Paporisch and Rubin 2017; Yun et al.
2001).
!
5!
In France, an environmentalist protest movement has emerged, and during
militant actions have been illegally mown mutant rapeseed, soya and maize
cultures made resistant to ALS inhibitory herbicides, and these actions strongly
remind the struggle of luddites of the 18th and 19
th centuries (Chevassus-au-
Louis 2006; Thompson 1966).
Although these plants are genetic mutants selected through an artificial
chemical mutagenesis and therefore true genetically modified organisms (GMOs)
on a biological ground, an institutional trick is that they are not legally
recognized as such (since a foreign and transgenic gene has not been inserted in
their genome and also because the technological process used is leading to the
same type of mutations than those arising through a “naturally” acquired
resistance). Consequently, these plants are excluded from the scope of the
European directive on GMOs, which exempts them from any evaluation,
traceability and labeling. Militant environmentalists are contesting this
institutional position and point of view and are worrying about the environmental
and possible health consequences of the use of herbicides of the IMI and SUL
families. Trials are in progress and in this context the aims of the present review
are first to address the question of the risks, potential toxicity and health
consequences of the use of such mutant plants regularly sprayed with herbicides
not only for the environment but also for the animals and humans consuming the
products derived from these plants. And second to discuss the flaws of the
official regulation during the marketing approval process of these mutant plants
and the corresponding herbicides.
1/ The toxicity of these herbicides to living organisms goes far beyond
the recognized molecular target ALS/AHAS enzyme.
!
6!
Organisms carrying the genuine ALS enzyme are not only facing the
inhibitory effects of IMI and SUL herbicides on ALS: many other intracellular and
molecular actors are targeted by these herbicides.
1.1. Genotoxic outcomes
The IMI imazethapyr causes chromosomal abnormalities during the mitosis
process, including an increased frequency of micronuclei in root cells of durum
wheat (Triticum durum) and bean (Vicia faba) (El-Nahas 2000; Rad et al. 2011).
In onions (Allium cepa), imazethapyr causes DNA damage (Liman et al. 2015),
and micronuclei and chromosomal aberrations after 4 days of exposure to
concentrations of 10 and 1 µg/L, respectively (Magdaleno et al. 2015).
1.2. Deregulation of the expression of genes at the whole genome scale
A concentration as low as 10 µg/L of imazethapyr in the culture medium of
A. thaliana affects sexual organ development and reproductive success through
the repression of genes involved in the synthesis of anthers and pollen (Qian et
al. 2015a). Indeed, the exposure of plants to ALS-inhibiting herbicides modifies
the expression of hundreds to thousands of genes in various plants (Table 1). In
rice, genes involved in photosynthesis and carbon fixation are repressed, as well
as genes for glucose consumption and starch mobilization, and amino acids
synthesis. Genes encoding enzymes of the Krebs cycles are up-regulated, along
with stress response and defense genes such as glutathione S-transferase (GST).
In maize, up-regulated genes are those encoding proteins involved in
translation, ribosomal structure and biogenesis, DNA replication, recombination
and repair. Down regulated genes are mainly involved in translation, ribosomal
!
7!
structure and biogenesis, posttranslational modification, protein turnover, and
chaperones.
In rapeseed, in the anthers of large buds, up-regulated genes are involved
in degradation of carbohydrates, lipids, proteins, nucleic acids and cell wall
degradation, and down-regulated genes are involved in the synthesis of
carbohydrates, lipids, proteins, nucleic acids, and cell wall synthesis.
In wheat, down-regulated genes are encoding proteins involved in stress
response and defense, carbohydrate metabolism and transport, and DNA
methylation.
In Arabidopsis thaliana, up-regulated genes are those encoding enzymes
involved in the metabolization of herbicides such as cytochrome P450 (CYP), GST
and UDP glycosyl transferases (UGT), and allowing detoxification such as ATP-
binding cassette (ABC) transporters and multidrug resistance (MRP) and toxin
extrusion (MATE) protein families, and the iron superoxide dismutase Fe-SOD
genes (FSD). Also up-regulated genes comprise those encoding ribosome
associated proteins or protein synthesis initiation factors (eIF4A, eIF4E and
eIF5), and some involved in amino acid synthesis. The mitochondrial function is
also impacted since the genes encoding mitochondrial genes alternative oxidases
ATAOX1a and ATAOX1b, along with two mitochondrial NADH dehydrogenases,
NDB2 and NDB4 are up-regulated. Down-regulated genes are those involved in
cell wall biosynthesis, along with genes involved in the neutralization of the
oxidative stress such as those encoding copper-zinc superoxide dismutases
Cu/ZnSOD (CSD), ascorbate peroxidases (APX) and glutathione peroxidases
(GPX).
In keeping with these gene deregulations, the SUL tribenuron-methyl
triggers the increase of the release rate of the plant hormone ethylene in the
!
8!
flower buds of rapeseed (Lian et al. 2019). This herbicide causes starch or
soluble sugar accumulation in rice (Qian et al. 2011b), Arabidopsis thaliana (Qian
et al. 2011a) and peas (Gaston et al. 2002; Royuela et al. 2000). Also, ALS-
inhibiting herbicides are strong promoters of an oxidative stress with an
important increase of lipid peroxidation, and of the concentrations in superoxide
radicals and hydrogen peroxide, and most often an anti-adaptive decrease of the
activity of the oxidative stress-neutralizing enzymes (Table 2).
1.3. Deregulation of the expression of proteins at the whole cell scale
Upon exposure of rapeseed seedlings to 1.5 µg of monosulfuron sprayed on
leaves, a proteomic analysis performed on whole tissues after one week showed
a total of 131 differentially expressed proteins between the treated and control
plants (Cheng et al. 2013): 34 proteins up-regulated and 97 down-regulated.
Up-regulated proteins comprised the metabolizing enzymes GSTs, UGTs and a
peroxidase, and proteins involved in photosynthesis and energy production.
Down-regulated proteins comprised those involved in cell defense and protein
synthesis and degradation, in carbohydrate, cell wall, energy metabolisms and
cytoskeleton dynamics. In another study rapeseed plants were foliar-sprayed
with a single quantity of 15 µg of the sulfonylurea SX-1 each. Compared to
control, 87, 25, 74 and 60 proteins were found to be up-regulated over two-fold
at young buds, short, mid and late anther stages respectively, and the decreased
ones were 130, 220, 329 and 366, respectively. These proteins were involved in
protein, energy, carbohydrate and amino acid metabolisms, and defense (Ning et
al. 2018).
1.4. Chloroplasts destruction and induction of male sterility
!
9!
In rapeseed, at a dose so little as 1 µg sprayed on leaves of a single plant,
amidosulfuron causes chloroplast destruction as soon as 3 days after treatment
(Liu et al. 2017), chlorsulfuron causes six hours later thylakoids swelling and
chloroplast disorganization (Kim and Vanden Born 1997), and exposure of A.
thaliana to 2.5 µg/L of imazethapyr for 4 weeks also causes a 2-fold decrease in
the number of chloroplasts (Qian et al. 2011a).
In rapeseed, monosulfuron is a male gametocide inducing male sterility
(Cheng et al. 2015), and transcriptomic analysis showed this is linked to
disruptions in pollen wall formation, chloroplast structure, cell cycle and tissue
autophagy (Ning et al. 2018; Lian et al. 2019). Tribenuron-methyl causes male
sterility not only in rapeseed but also in 17 species of cruciferous plants (Yu et al.
2017) through the onset of an autophagic cell death in anthers (Zhao et al.
2015). Recently, it has been shown that 18 SULs and 2 IMIs could induce male
sterility in rapeseed (Yu et al. 2020).
1.5. Poisoning of the mitochondrial respiration
Sulfonylureas and imidazolinone herbicides poison the mitochondrial
respiration in sycamore cultured cells (Aubert et al. 1997). Sulfometuron at a
dose so little as 1 nM completely blocked the mitochondrial respiratory chain
after 5 days of exposure leading to cell death. The mitochondrial alternative
oxidase AOX activity was 6-fold induced after three days of exposure to 1 nM
sulfometuron, 1 nm chlorsulfuron, 1 µM of the IMI Scepter, but not when
exposed even for 5 days to 1 mM glyphosate. AOX dissipates as heat the
electron flow that is not used to yield ATP (only the complex I can contribute to
the proton gradient and ATP synthesis when electrons flow through AOX). The
herbicides-linked increased activity of this enzyme was paralleled by an up-
!
10!
regulation of the Aox1 gene and an increased mitochondrial concentration of the
AOX enzyme.
1.6. Effects on non-target aquatic microorganisms
The photosynthetic potential of the cyanobacteria Microcystis aeruginosa
suffered 29 and 14 % decreases after 16 h exposure to 67 µg/L of the IMI
imazethapyr and 3 µg/L of the SUL metsulfuron-methyl (MSM), respectively, and
that of the algae Selenastrum capricornutum presented a 27 % decrease after
exposure to 3 µg/L of MSM (Peterson et al. 1994). The SUL nicosulfuron had an
impact on algal community diversity (it inhibited diatoms more than
chlorophytes) and on the photosynthetic yield at a concentration of 30 µg/L
(Seguin et al. 2001). SUL herbicides affected growth and reproduction of
Chlorella fusca at concentrations between 0.08 and 1.2 ppm (Fahl et al. 1995),
and those of Scenedesmus acutus at about 0.1 ppm (Grossmann et al. 1992;
Sabater and Carrasco 1997). SUL herbicides could inhibit the incorporation of
adenine and thymidine of periphyton communities from natural creeks for
concentration as little as 3 nM (Nyström et al. 1999).
If these herbicides target other molecular actors than the ALS enzyme, then
could organisms without ALS be sensitive to the deleterious effects of these
chemicals as well?
2/ These herbicides are toxic, including organisms that do not have the
enzyme ALS/AHAS.
The pesticide lobby and several French institutions such as the national
agronomic research institute INRA, the national federation of agricultural unions
!
11!
FNSEA, and the Ministry of Agriculture all pretend that the only molecular target
of IMI and SUL herbicides would be the enzyme ALS, and that consequently all
the organisms deprived of this enzyme would be tolerant to these herbicides.
This belief does not match with the literature.
2.1. Invertebrate species
In rice paddy water, the half-lives (time required for 50 % dissipation of the
herbicide) of the IMIs imazethapyr and imazapic were equal to 4.5 and 11.9
days, respectively (Reimche et al. 2015). In such artificial environment, a
combined treatment of imazethapyr and imazapic at concentrations equal to 17.4
and 9.5 µg/L at time 0 resulted in a very significant decrease in the density of
several rotifer species. At day 56, 3.4-, 10- and 15-times lower densities
compared to unexposed control were observed for the genera Keratella,
Polyarthra and Trichocerca, respectively. At that time, the calculated residual
concentrations could be calculated to be 0.24 and 20.0 ng/L (0.8 and 73 pM) for
imazethapyr and imazapic, respectively, meaning that these IMIs can still slow
down animal growth at the range of pmol/L. The recovery of densities similar to
those of the control experiment was observed after 88 days (Reimche et al.
2015). The SUL metsulfuron-methyl combined with its adjuvant Assist proved to
decrease the reproductive success of Enchytraeus crypticus oligochaeta worms
and Proisotoma minuta collembolans (de Santo et al. 2019) and to trigger an
avoidance behavior in earthworms Eisenia andrei and Folsomia candida
collembolans (de Santo et al. 2018).
IMIs imazamox Sweeper and imazaquin Scepter 70 DG from BASF are
genotoxic for Drosophila (Fragiorge et al. 2008). Genotoxicity has also been
proven in Drosophila melanogaster for the SUL triasulfuron Amber 75WG from
!
12!
Novartis when the feed was containing the herbicide or an aqueous extract of
wheat previously exposed to the herbicide (Heres-Pulido et al. 2008).
The bioaccumulation of IMI herbicides in Eisenia fetida earthworms (the
ratio of the herbicide concentration in earthworms to that in soil) is depending
upon the richness of the soil in its organic matter. When soils were spiked with 2
mg herbicide/kg, earthworm concentrations up to 170, 140 and 240 µg herbicide
per kg were recorded for both enantiomers of imazapic, imazamox and
imazethapyr, respectively, and after 6 weeks of exposure bioaccumulation
factors were varying according to the quality of soil with values ranging between
0.2 and 0.45 for imazapic, 0.25 and 1 for imazamox, and 0.4 to 0.8 for
imazethapyr (Hu et al. 2018).!
In the earthworm species Dendrobaena veneta, the SUL nicosulfuron
caused after 28 days of exposure to a soil concentration of 0.3 mg/kg, a relevant
environmental one, a 20 % increase in lipid peroxidation compared to the control
animals, and a 60 % drop in the activity of the glutathione-S-transferase enzyme
(GST, involved in the neutralization of oxidative radical species) (Hackenberger
et al. 2018). In the earthworm Eisenia fetida, the SUL tribenuron-methyl at a
concentration of 52 mg/kg of soil (5 % of the LC50 on day 7) triggered after 3
days of exposure the decreased activity of SOD and cellulase enzymes, the latter
playing an important role in the processing of organic matter in worms’ gut (- 14
and 61 %, respectively), whereas that of catalase enzyme was increased by 41
% (Chen et al. 2018). The activities were fully recovered after 14 days probably
due to the degradation of the herbicide and its adsorption on soil particles
thereby decreasing its the bioavailability as exemplified for IMIs in earthworm-
soil microcosms (Hu et al. 2018).
!
!
13!
2.2. Amphibians
IMI imazethapyr in its Pivot formulation (10.6%) is very toxic to Hypsiboas
pulchellus tadpoles (LD50 at 96 hrs equal to 1.55 ± 0.05 mg/L) altering the
swimming behavior of animals and displaying a genotoxicity (increased
frequency of micronuclei and lobed nuclei in erythrocytes) as soon as 0.39 mg/L
after 48 hours (Pérez-Iglesias et al. 2015). This concentration at 48 h also
caused mutations of the DNA (by oxidation of the nucleic bases) in the blood
cells (Pérez-Iglesias et al. 2017). Boana pulchella tadpoles exposed for 4 days at
this concentration had oral morphological abnormalities (loss of rows of
keratodonts), which persisted 21 days after transfer to a healthy medium. In
addition, another type of caudal morphological anomaly (lateral torsion)
appeared after 14 and 21 consecutive days of transfer to a healthy medium
(Pérez-Iglesias et al. 2018). DNA damage were observed in blood erythrocytes of
Rhinella arenarum and Leptodactylus latinasus tadpoles exposed for 96 h to a
concentration of Pivot imazethapyr equal to 0.05 mg/L (5 % of the LC50 at 96 h)
and 0.07 mg/L, respectively (Carvalho et al. 2019; Pérez-Iglesias et al. 2020). In
L. latinasus tadpoles, this herbicide also caused irregular swimming (LOAEL
48h
equal to 0.81 mg/L), loss of keratodonts (LOAEL48h equal to 0.81 mg/L), and
nuclear abnormalities in blood erythrocytes as soon as 48h after exposure to
0.07 mg/L (Pérez-Iglesias et al. 2020). The SULs sulfometuron methyl and
nicosulfuron proved to be teratogenic in embryos of the frog Xenopus laevis, and
the median teratogenic concentrations EC50 were equal to 4.2 and 3.1 mg/L,
respectively (Fort et al. 1999).!
!
2.3. Fish
!
14!
The toxicity of ALS-inhibiting herbicides is well documented in the
toxicological literature in freshwater fish (Table 3). Recorded are a decrease in
the activity of several enzymes including catalase, GST, and acetylcholine
esterase (AchE) in several tissues, changes in swimming behavior, increased
oxidative stress-related processes such as lipid peroxidation and protein
carbonylation, increased blood glucose, decreased muscle glycogen
concentration, and decreased total protein concentration in several tissues.
Histological alterations and morphological degeneration of mitochondria were
demonstrated by electron microscopy. The toxicity of these herbicides is also
accompanied by a change in the homeostasis of several amino acids and of that
of neurotransmitters glutamate, taurine and glycine (the corresponding
references are collected in Table 3).!
2.4. Mammalian cells
Chinese hamster ovary cells (CHO-K1) exposed for 90 min to the IMI
imazethapyr exhibited DNA damage at the lowest concentration tested (100
µg/L), which were five times more numerous than in unexposed control cells.
Damage were essentially related to the oxidation of the DNA bases, and caused
by the herbicide's only molecule, because the commercial Pivot formulation
tested at the IMI equivalent concentration caused the same increase in DNA
damage as the IMI alone (Soloneski et al. 2017).
The IMIs iodosulfuron-methyl-sodium, mesosulfuron-methyl, and
metsulfuron-methyl proved to be agonists of the soluble AhR dioxin receptor
(aromatic hydrocarbon receptor) for a concentration of 0.1 mM applied for 4 h to
human hepatocytes Hepa 1.12cR (Ghisari et al. 2015). The AhR-controlled
pathway is responsible for most of the toxic and biological effects of polycyclic
and halogenated aromatic hydrocarbons (Puga et al. 2009).
!
15!
A panoptic figure is summarizing the main toxic effects of ALS-inhibiting
herbicides on plants and animals (Figure 1).
3/ The problem of testing herbicides out of their commercial
formulation.
Official regulatory agencies - pushed by industrial lobbies - consider that the
toxicology of pesticides must be conducted on the active ingredient or molecule,
and not on the total formulation, the same used in fields. However, the
toxicological literature undeniably shows that the formulation including the active
molecule is always significantly more toxic than the latter used alone. The
examples developed below in the case of ALS inhibitory herbicides are illustrating
that matter of fact.
In the onion, after 5 days of exposure, no effect on root growth was
observed until the highest dose tested of 80 mg imazethapyr/L (Liman et al.
2015). However, in lettuce, the Verosil formulation of Agrofina (10.6%
imazethapyr) showed a 34 % inhibition of root growth after 5 days of exposure
for the smallest tested dose of 10 mg imazethapyr/L (Magdaleno et al. 2015).
In the onion, after 4 days of exposure to the Verosil formulation, this
herbicide caused the appearance of micronuclei and chromosomal aberrations at
concentrations equal to 10 and 1 µg imazethapyr/L, respectively (Magdaleno et
al. 2015); on the other hand, when added pure in the medium, a concentration
2000-times higher (20 mg/L) was required to observe DNA damage after 4 days
of exposure (Liman et al. 2015).
The Verosil formulation was significantly more toxic than 2ASU whereas it
contains half as much active compound (10.6 and 22.2% imazethapyr,
!
16!
respectively), which highlights the important role of adjuvants in the overall
toxicity (Table 4).
Metsulfuron-methyl is a herbicide belonging to the family of SULs, and is
sold in Brazil under the brand Ally by Du Pont. The adjuvant (mineral oil)
associated is marketed by BASF under the brand Assist. Toxic avoidance tests in
earthworms Eisenia andrei and in collembolans Folsomia candida (pancrustacean
arthropod) showed that the SUL Ally was significantly more toxic in combination
with its adjuvant Assist (de Santo et al. 2018). The observed LOAELs (smaller
doses for which a significant effect is observed) showed that the adjuvant greatly
magnified the toxicity of the herbicide in these animals (Table 5). Also, the
reproductive success of Enchytraeus crypticus oligochaeta worms and Proisotoma
minuta collembolans was significantly reduced (Table 5) when the exposure to
the SUL Ally was combined with its adjuvant Assist (de Santo et al. 2019).
Using the marine bacterium Vibrio fischeri as a model organism, it has been
observed that the growth inhibitory concentration (IC50) of the SUL nicosulfuron
was reduced from 168 down to 4 mg/L when comparing the active ingredient
alone or the Milagro (Syngenta) formulation, respectively (Joly et al., 2013).
The IMI imazapyr is marketed under the brand name Arsenal 250 NA by
BASF. This formulation contains 25 g/L of imazapyr, 186 g/L of ammonium
hydroxide, and 18 g/L of nonylphenol ethoxylate in water. Imazapyr is
significantly more toxic in its Arsenal formulation (Grisolia et al. 2004). In three
different animals (aquatic snail, vinegar fly, and mouse), the LD50 were
significantly lower after exposure to the Arsenal formulation; and the onion
exhibited chromosome separation aberrations during mitosis for an equivalent
imazapyr concentration 8,000 times lower in its Arsenal formulation (Table 6).
!
17!
In addition to the presence of adjuvants, another herbicide may be added to
the commercial formulation of a product. This is the case for BASF's Cleravis and
Cleranda formulations used on Clearfield plants, which contain 17.5 g/L of the
IMI imazamox and 375 g/L of metazachlor. Metazachlor is a chloroacetamide for
which mutant plants tolerant to SULs and IMIs have no tolerance or resistance.
However, the toxicity of these compounds is not assessed in combination or even
in their commercial formulation in the presence of all the adjuvants used in the
composition. However, the use of metazachlor has caused an environmental
disaster in Luxembourg, with the regulatory concentration limit for drinking
water being exceeded for several months, and the main drinking water reservoir
in Luxembourg being contaminated (Karier et al. 2017). In France, an episode of
drinking water pollution by metazachlor led to a five-week ban on drinking water
consumption (from February 17 to March 21, 2016) in several villages in
Burgundy (l’Yonne Républicaine 2016; a local newspaper).
4/ Epidemiological and pharmacological data on human populations.
Regulatory authorities minimize the toxicity of ALS inhibitory herbicides
because SULs and IMIs are rapidly excreted from mammalian organisms,
although intestinal absorption is also recognized high and rapid. Thus, in animals
receiving an oral dose of the SUL MSM, this toxic and its metabolites are found 3
to 4 days later up to 71 to 95 % in urine and 5 to 13 % in excrements (Health
Canada 2008, page 10). In the rat, almost 100 % of the radioactive IMI
imazethapyr administered was found in the following 4 days at 89 to 95 % in the
urine and 6 to 11 % in the excrement (Toxnet 2012).
However, IMIs have a wide variety of pharmacological effects:
anticonvulsant, sedative, antihistaminic, fungicidal, bactericidal, anti-
!
18!
inflammatory, monoamine oxidase (thus anti-depressive), and antihypertensive
(Desai et al. 2009). It is therefore unreasonable to imagine that these herbicide
molecules once transferred to human organisms would be devoid of activity.
When herbicide molecules have entered our body by air or intestinal tract,
human serum albumin binds these toxins and transports them into the blood,
distributing these toxins to tissues (Ding et al. 2010). This process is very
effective since human serum albumin is the most abundant protein in serum with
a concentration of 35 to 50 g/L of serum.
Cases of acute poisoning with the IMI imazapyr (Arsenal) have been
reported and the toxic panel resulting from an ingestion of over 100 mL of
Arsenal consisted of hypotension, pulmonary dysfunction, oral mucosal and
gastrointestinal irritation, and transient liver and renal dysfunction (Lee et al.
1999). In one case a cardiac arrest has been observed and this correlated with a
serum imazapyr concentration of 0.71 mg/mL (Wu et al. 2019).
Epidemiological studies performed on a cohort of 20,646 IMI herbicide
applicators have shown that the use and handling of these toxins is associated
with a significantly increased risk of colon and bladder cancer (Lee et al. 2007;
Koutros et al. 2009 and 2016; Weichenthal et al. 2010; Alexander et al. 2012).
In regions of rural Minnesota where soybean is cultivated and imazethapyr used
in combination with other pesticides, increased mortality rates by cancer have
been observed in women (nasopharynx and non-Hodgkin’s lymphoma)
(Schreinemachers et al. 1999).
Inhibitors of ALS are known to impact the human reproduction since the use
of herbicides containing sulfonylurea and imidizolinone by male applicators was
correlated with increased miscarriage risk in the spring, at the time herbicides
are applied in fields (Garry et al. 2002; Colborn and Carroll 2007). Conceptions
!
19!
in the spring also led to significantly more birth defects among children of rural
Minnesota (Garry et al. 1996). In Brazil also, some soybean farmers exposed to
33 different pesticides reported that their wives had experienced spontaneous
abortion (Bernieri et al. 2019).
SULs are also endocrine disrupters and this effect has been harnessed for
therapeutic strategies. Indeed, SULs are also used in human therapy as
hypoglycemic agents in the treatment of type II diabetes. Their action depends
on their binding to the pancreatic β-cell SUL receptor SUR1 (SUR1 is the
regulatory subunit of the ATP-sensitive potassium channel, KATP, the closure of
which triggers the secretion of insulin). The SUL depuration metabolism is
catalyzed by the cytochrome P450 CYP2C9. The CYP2C9*2 and CYP2C9*3 alleles
are associated with hypoglycemic events in diabetic patients taking Glimepiride
or Gliclazide SULs (Ragia et al. 2009), since these molecules are no longer
effectively detoxified. Also, deleterious drug interactions may occur: voriconazole
and fluconazole antifungals, for example in the treatment of fungal pneumonias,
are also CYP2C9 inhibitors, and may lead to hypoglycemia in diabetic patients
treated with SULs (Gunaratne et al. 2018).
Cohort studies show that the risk of cancer is greater in diabetic patients
using SULs rather than the metformin hypoglycemic agent (Chen et al. 2017).
Given the great similarity of the SULs used as a herbicide with those used as a
hypoglycemic agent, it would be unreasonable to imagine that this type of
herbicides do not bind to the SUR-1 SUL receptor, from which it can be
formulated the hypothesis that this class of herbicides plays a role in blood
glucose after accumulation in our bodies. For instances, fish exposed for 90 days
in rice field to Ally herbicides (MSM, 5.8 µg/L) or Only (at concentrations of 14.8
µg/L of imazethapyr and 3.1 µg/L of imazapic) present, among other effects (see
!
20!
Table 1), 15 and 80 % increases in blood glucose levels compared to unexposed
animals (Pretto et al. 2011; Moraes et al. 2011). This strongly suggests that
these herbicide molecules trigger the opening of KATP channels in β-cells of the
pancreas in these fish, and therefore mimic the effects of diazoxide, a SUR-1
receptor ligand (Emfinger et al. 2017), the active principle of the hyperglycemic
drug Proglicem proposed to hyperinsulinic patients.
5/ Impact of ALS-inhibiting herbicides on biodiversity and their fate in
the environment.
5.1. Loss of biodiversity
ALS-inhibiting enzymes are contributing among several other causes and
chemicals to the loss of biodiversity. The representative tree species of the
Brazilian forest Dipteryx alata, ranked by the International Union for
Conservation of Nature among the vulnerable species, has been shown to be
sensitive to recommended dose of the SUL nicosulfuron, with the onset of an
oxidative stress and the alteration of the photosynthetic potential (Silva et al.
2020). The ALS-inhibiting herbicide bispyribac at the recommended dose of 35
g/ha triggered a decrease of the soil microbial population and of microbial
diversity, and a decrease of microbial enzyme activities such as dehydrogenases
and urease (Kumar et al. 2020). In soils from fields treated with 30 g of the SUL
chlorimuron-ethyl/ha for 5 and 10 years, in which soybean had been cropped
continuously, the residual herbicide concentrations were 3.4 and 8.5 µg/kg, and
the soil microbial population presented a 33 and 51 % decrease, with a shift of
the soil bacterial and fungal community structure. In particular, after 10 years of
!
21!
treatment, the actinomycete titer decreased 7.5-fold whereas the fungal titer
increased 2.5-fold (Zhang et al. 2011).
5.2. Onset of ALS-inhibiting herbicides resistant weeds, and mutant ALS
gene transfer
The Clearfield technology has allowed the emergence of herbicide-tolerant
weeds, and this has been described in the case of the rice Clearfield where
resistant weedy red rice evolved both from selection pressure due to continuous
exposure to herbicide and gene transfer from Clearfield rice (Sudianto et al.
2013). The Clearfield plants and those harboring a mutated and tolerant form of
the ALS enzyme are resisting to concentrations much higher than their sensitive
counterparts (Table 7). After only one year of Clearfield rice cultivation, the ALS
resistant allele transfer was made evident and the outcrosses amounted to a
mean of 0.17 % (Zhang et al. 2006). The gene transfer from Clearfield rice to
weedy rice has been reported in several places around the world (Villa et al.
2006; Shivrain et al. 2007; Burgos et al. 2008; Shivrain et al. 2009a and 2009b;
Busconi et al. 2012; Dauer et al. 2018). Weeds sampled from fields with a
history of Clearfield rice had at least 20 % resistant offspring (Burgos et al.
2014). Resistant gene transfer also occurred from Clearfield rapeseed to the wild
turnip (Brassica rapa) weed (Ureta et al. 2017).
5.3. Persistence of ALS-inhibiting herbicides in soils and in the
environment
SUL and IMI herbicides are degraded in soils by microbes, and both bacteria
and fungi participate to this process (Goetz et al. 1990; Flint and Witt 1997;
Kraemer et al. 2009a; Singh and Singh 2014; Arya et al. 2016). The application
!
22!
rates in fields of ALS-inhibiting herbicides, along with their half-lives and
environmental recorded concentrations have been collected (Table 8). The half-
life of the IMI imazamox depends on the type of soil and varies from 2 to more
than 120 days. The rate of degradation is faster in neutral rather than acidic soils
(Buerge et al. 2019). In the aquatic environment, the half-life of the MSM SUL
was 8 days at an initial concentration of 5.8 µg/L (Pretto et al. 2011), and those
of imazethapyr and imazapic IMIs were equal to 6 days for initial concentrations
of 14.8 and 3.1 µg/L, respectively (Moraes et al. 2011). Although the persistence
of these molecules in aquatic environments was low, fish exposed to these
molecules showed obvious signs of toxicity after 90 days of exposure to a single
dose added the first day in rice field, that is after 15 half-lives of the compounds
and residual calculated concentrations of 0.45 and 0.1 ng/L (1.6 and 0.35 pM,
respectively) (Table 3, Moraes et al. 2011; Pretto et al. 2011). The fish are
therefore permanently intoxicated by low doses to which they were exposed
weeks earlier and commensurate with the picomolar range. SUL herbicides such
as chlorsulfuron can present high persistence in soil: recorded half-lives range
from 2 weeks (Fredrickson and Shea 1986) to 7.5 months (Thirunarayanan et al.
1985).
In agricultural soils and wetland sediments, the three most persistent
environmental SULs, ethametsulfuron-methyl, sulfosulfuron, and MSM, were
detected at mean concentrations ranging from 1.2 to 10 µg/kg (Degenhardt et
al. 2010), and thus already toxic levels for organisms. Nicosulfuron has been
detected in freshwater rivers and reservoirs in the United States at median and
maximum concentrations of 10 and 270 ng/L (Battaglin et al. 2000). The half-
lives measured in grasslands for thifensulfuron-methyl, ethametsulfuron-methyl,
and MSM were 16, 30, and 84 days, respectively (Cessna et al. 2006). MSM
!
23!
showed a dissipation rate of only 34 % at 20°C after 35 days of incubation in an
experimental biobed matrix (Cessna et al. 2017). In pools adjacent to farms, the
measured half-lives of nicosulfuron, sulfosulfuron, and rimsulfuron were 75, 44,
and 10 days, respectively (Cessna et al. 2015). The leaching of IMIs from fields
during rainfall and irrigation has been estimated to range between 0.7 to 3.1 and
0.6 to 2.8 % of the total amount initially applied for imazethapyr and imazamox,
respectively (Cessna et al. 2012). In a 1998 study, in 129 water samples from
US Midwestern streams and rivers, 108 tested positive for the sum of 16
herbicides (comprising 1 sulfonamide, 3 IMIs and 12 SULs) with median and
maximum concentrations of 137 ng/L and 2100 ng/L, respectively (Battaglin et
al. 2000). A study extending from 2009 to 2011 in the Saint-François Bay,
adjacent to the Saint Lawrence River, showed that among 70 water samples, the
detection frequencies (> 10 ng/L) of the SULs ethametsulfuron-methyl, MSM,
nicosulfuron and sulfosulfuron reached 4.3, 10.0, 8.6 and 7.1 %, respectively (de
Lafontaine et al. 2014). Maximum concentrations equal to 148 and 223 ng/L
were sporadically above the baseline level (100 ng/L) for aquatic plant toxicity,
meaning potential toxic stress to flora in the streams (de Lafontaine et al. 2014).
The IMI imazethapyr was detected in 71 % of water samples collected from 75
sites in the United States (Upper Mississippi, Missouri and Ohio River basins) with
median and maximum concentrations of 30 and 700 ng/L (Battaglin et al. 2000).
By way of comparison, the European Community has set a limit value of
pesticides in drinking water at 100 ng/L (quoted by Ding et al. 2010). This IMI is
degraded in soils with the following measured half-lives: 2.6 to 10.6 months
(Goetz et al. 1990), 23.5 to 62 days (Ramezani et al. 2010), and 73 to 257 days
(Vischetti 1995).
!
24!
5.4. Carry-over of herbicides and impact on rotational crops
Another problem beside resistance gene transfer is the persistence in soils
of ALS-inhibiting herbicides linked to repeated application and the poisoning of
the next rotational crops. The concentrations of SUL herbicides necessary to
inhibit the growth of some plants by 10 % has been recorded: 340, 180, 90 and
30 ng/kg of soil for barley, corn, sugar beet and oilseed rape, respectively,
exposed to tribenuron-methyl; 140, 70, 50, and 40 ng/kg of soil for barley,
sugar beet, corn and oilseed rape, respectively, exposed to sulfosulfuron
(Mehdizadeh et al. 2016). Given the application rates commonly used and
recommended by manufacturers, the half-lives in soils of ALS-inhibiting
herbicides, and the concentrations found in agricultural soils (Table 8), it is not
surprising that carry-over effects can be observed on rotational crops. In rice
paddy soil imazethapyr and imazapic (Only formulation, BASF) used with
Clearfield rice still remained up to 20 cm of depth 540 days after the last
application with surface concentration equal to 1281 µg imazethapyr/m2
(equivalent to 12.8 g/ha) (Kraemer et al. 2009b). After use of a mixture of
imazapic and imazapyr on an IMI-tolerant corn, the growth ability of non-tolerant
rotational crops, such as corn, cucumber, wheat, sugar beet, chili pepper,
tomato, melon, barley, oat, pea, and edible bean was impacted by the carried-
over herbicides (Alister and Kogan 2005; Ulbrich et al. 2005). The yield of non-
tolerant rice was reduced up to 55 % when planted 705 days after application of
the IMI herbicide Only formulation (Marchesan et al. 2010). Also the yields of
sorghum and of ryegrass, the most widely used pasture grass in Brazil, were
deeply impacted after being planted in rotation with Clearfield rice treated with
the Only formulation (Pinto et al. 2009a and 2009b). Carry-over of imazaquin
caused corn and sunflower injury and yields reduction of corn and cotton (Barnes
!
25!
et al. 1989; Loux and Reese 1993; Fleck and Vidal 1994; Marsh and Lloyd 1996).
Imazethapyr application proved to cause yield losses of sorghum seeded 120
days later (da Silva et al. 1999), flax, corn, mustard, sunflower and wheat
seeded one year later; canola seeded up to two years later, and sugarbeet and
potato seeded up to three years later (Moyer and Esau 1996). Imazamox applied
to IMI-tolerant wheat damaged barley and canola grown one year after
imazamox treatment (Ball et al. 2003). Application of the SUL sulfosulfuron
inhibited nine months later the growth of seeded sunflower (Alonso-Prados et al.
2002), and caused 40 % of reduction in yield to winter rape seeded in crop
rotation (Adamczewski and Paradowski 2004). The mixture of the SULs
amidosulfuron and tribenuron-methyl also decreased the yields of sugar beet,
beans and winter rape seeded in crop rotation (Adamczewski et al. 2004), and
MSM affected growth of lettuce and sugar beet sown one year later (Walker and
Welch 1989). The SULs chlorsulfuron, triasulfuron and MSM affected sunflower,
lentil and sugar beet sown 8 months after initial treatment of wheat (Kotoula-
Syka et al. 1993). Triasulfuron at 22 g/ha reduced the growth of alfalfa, canola,
corn, lentil, pea, potato, and sugarbeet the year after application, and at 11 g/ha
increased injury on alfalfa, lentil and sugar beet one year after treatment (Moyer
1995; Shinn et al. 1998). Sulfosulfuron treatment of wheat reduced one year
later the growth of barley, pea and canola, for an application rate of 36, 18 and
18 g/ha, respectively (Shinn et al. 1998).
Finally, although these herbicides are intended to increase yields by
reducing competition from weeds, sometimes yields or benefits may be lower
than those obtained without the herbicide with hand weeding. For example, after
treating potatoes with sulfosulfuron at a rate of 25 g/ha, total yield was 24 t/ha
compared to 29.6 t/ha with hand weeding. Operating profits (difference between
!
26!
selling price and operating costs) amounted to 59,000 Indian rupees per ha
compared to 81,000/ha with hand weeding. Increasing the herbicide application
rate to 50 or 75 g/ha did not improve yields or benefits (Singh et al. 2017).
Lower yields after treatment with the IMIs imazaquin or imazethapyr than with
hand weeding have also been observed with soybean (Mills and Witt 1989) and
corn (Curran et al. 1992).
6/ The official regulation and its shortcomings.
If we take as examples an IMI molecule, imazethapyr, and a SUL one, MSM,
it can be found that the toxic potential of these inhibitors of ALS is minimized by
the regulatory authorities in each country. Thus, it is said that MSM has no toxic
effects on terrestrial invertebrates, birds or mammals through food and no
effects on reproduction; the same applies for freshwater invertebrates and fish in
acute exposure mode, so that chronic effects can not be predicted (Health-
Canada 2008, p. 19-20), in contradiction with all published results in the
literature and explained above in this short review. SUL-type herbicides are
applied in Canadian fields at a concentration of 3 to 40 g per hectare to wheat,
barley, oats, rapeseed, corn, soybean, and forage crops (Saskatchewan Ministry
of Agriculture, cited by Cessna et al. 2015). However, concentrations of these
herbicides applied in fields planted with herbicide-tolerant crops or labeled
Clearfield technology are much higher (Table 8). Tolerable doses of MSM and its
metabolites in food and cereals are as follows: barley and wheat kernels: 0.1
mg/kg; sugar cane: 0.05 mg/kg; meat (cattle, sheep, pigs, horse): 0.1 mg/kg;
milk: 0.1 mg/kg (Toxnet 2006). The tolerable daily intake (TDI) is the dose at
which an individual can be exposed during his life without damaging his health.
MSM TDI equals 83.3 µg/kg body weight/day (Health-Canada 2008, p. 16). For a
!
27!
person of mass equal to 60 kg, this represents 5 mg/d. However, a diabetic
patient who takes as hypoglycemic drug the SUL glimepiride will ingest 1 to 6 mg
of this compound per day (the common posology), and if that person takes
gliclazide, then it will be 30 to 120 mg/d (the common posology). It is obvious
that any addition of this type of molecule by food will contribute to the exceeding
of the TDI (this is already the case for those treated with gliclazide).
Imazethapyr is classified as a weakly toxic compound (Class III) by the EPA
(Environmental Protection Agency) in the United States (2002), and claimed
unlikely to be hazardous by the WHO (Pesticide Action Network), in contradiction
with the published literature quoted above in this short review. Tolerable doses
of imazethapyr and its metabolites in food and cereals are: rapeseed and soya
beans, corn kernels, vegetables, peanuts: 0.1 mg/kg; meat (cattle, sheep, pigs,
horse): 0.1 mg/kg; crayfish: 0.15 mg/kg; grains of rice: 0.3 mg/kg (Toxnet
2012). The TDI of imazethapyr is 0.56 mg/kg body weight/day. However,
exposure of the general population through food and drink is 0.050 mg/kg bw/d
(8.9% of TDI). In children 1 to 2 years of age, exposure is greater: 0.24 mg/kg
bw/day (43% of TDI) (Health-Canada 2010, p. 45). What will happen when
these children will consume food from plants resistant to ALS-inhibiting
herbicides? With the use of such mutant plants resistant to this herbicide, it can
be feared that the doses and applications will increase with the result that
children will exceed the TDI.
Other limits and shortcomings of the regulation: there is no obligation
pending on agrochemical companies offering the mutant plants to the market to
publish analytical dosages of the IMI and SUL herbicides in the tissues, grains or
seeds of these mutant plants after treatment and harvest at the doses they are
recommending. In the literature, no data are available on the IMIs and SULs
!
28!
accumulated in seeds and cereals produced and consumed by humans. An
experimental study could give the remaining concentrations of the IMI
imazosulfuron applied on rice at a dose rate of 60 g/ha: 60 days after treatment
the soil contained 10 ng/g and the plant tissues 64 ng/g (32 ng/g 90 d after
treatment). At harvest the rice grains and straw contained 9 and 39 ng/g,
respectively (Sondhia 2008). Because of the resistance of these mutant plants,
the concentration is necessarily much higher than in non-resistant plants for
which the intensity of treatment must be limited so as not to kill the producing
plant. Similarly, there is no requirement for companies or independent
laboratories to submit a study in which animals are fed for life long with food
derived from crops resistant to ALS-inhibiting herbicides, with an analysis of the
transfer of toxic molecules into animal tissues. The few weeks or less frequently
the three months usually used for this type of studies are far too short in relation
to the longevity of human beings whose chronic exposure spread over decades
(nobody has contracted lung cancer after only three months of smoking, even
exaggerated).
7. The paradigmatic case of the marketing authorization for Zeneca
mutant maize.
Another problem, the regulatory authorities do not take into consideration
that the toxicity can reside in the amounts of pesticides present in the tissues
and seeds of the mutant plants consumed. Thus, and to take an example, when
Zeneca deposited its new EXP1910IT maize hybrid made resistant to the IMI
imazethapyr by chemical mutagenesis, Health Canada stated in its information
bulletin that after having reviewed the information presented in support of the
food use of imazethapyr tolerant corn hybrid EXP1910IT, it was concluded that
!
29!
this corn does not raise concerns related to human food safety. Products from
EXP1910IT corn are as safe and nutritious as those available from current
commercial field corn varieties” (Health Canada 1999). The regulatory authorities
are indeed backing their recommendations, and as amazing as it may seem, on
the data and studies put forward by the company applying for the marketing
authorization, and not on studies independent from the manufacturer led by
laboratories that do not maintain any link with the industry and, as a result,
would escape the conflict of interest. Moreover, if one reads why Health Canada
accepts Zeneca corn and declares it to be harmless, one can only be struck by
what can only be described as denial of reality (denial of the presence of
herbicide and adjuvant molecules in the sold maize after applications of the
recommended treatments). First, Health Canada recognizes that the chemical
composition of Zeneca mutant maize is identical to natural corn in terms of
protein, fat, fiber, starch and minerals. This is the principle of equivalence in
substances. However, herbicide and adjuvant concentrations in the mutant maize
have not been measured after treatments. If they had been, the principle of
equivalence in substances would have been reduced to nothingness. Rather than
conducting these herbicide and adjuvants analysis, Zeneca insisted that "the
amino acid sequence of the mutant form of ALS present in EXP1910IT corn is
identical, except for a single amino acid substitution, with the wild-type form of
this enzyme (Health Canada 1999). As a result, “the mutated form of this
enzyme is not judged to have any potential for human toxicity (but who claimed
the opposite?) In addition, " the imazethapyr tolerant form of ALS is extremely
unlikely to be allergenic » (Health Canada 1999). It is clear that in the mind of
the manufacturer and experts of the regulatory institution, the only difference
that deserves attention is the mutated amino acid of the mutant form of the
!
30!
enzyme, whereas the health problem obviously concerns the quantities of
herbicides and adjuvants present after harvest in the plant material and which
must necessarily be much higher than in non-resistant maize since the mutant
plant can resist up to 0.5 mM of the herbicide.
In fact, it has been observed that the treatment of maize with the SUL
nicosulfuron in combination with the “safener” isoxadifen-ethyl altered the
nutritional content and the taste of the edible crop (Cutulle et al. 2018). In the
case of the Clearfield canola, lepidopteran animals can make the difference
between three canola varieties: Clearfield, Roundup Ready and a common one.
Even without herbicide or pesticide treatment, most eggs were laid on the
common canola. In addition, for the same amount of fertilizer used, the common
variety presented the highest potassium content, and a higher calcium and
magnesium content than the Clearfield variety (Weeraddana and Evenden 2018).
8/ The precautionary principle and the Dijon trial
A social movement protesting against these new mutant plants has
emerged in France these last recent years. Illegal mowings have taken place,
and lawsuits are ongoing. In particular, by November 28, 2016, 67 volunteer
mowers have neutralized Clearfield rapeseed (BASF) test plots used with the
herbicides Cleranda and Cleravis. These activists wanted to demonstrate that
GMOs are grown in France (despite the moratorium applied in France), and to
trigger the constitutional principle of precaution.
Indeed, the Environmental Chart promulgated on March 1
st, 2005 by the
French Republic, which has constitutional value, specifies that "everyone has the
right to live in an equilibrated environment that respects health". It also
establishes the precautionary principle with regard to potentially serious and
!
31!
irreversible damage to the environment, even if it is uncertain in the light of
current scientific knowledge. However, pesticides are by definition toxic to living
things, and even when they target organisms other than animals, they can still
be harmful to the health of animals. The accumulated toxicological and
epidemiological data on ALS-inhibiting herbicides, synthesized in this review, are
sufficient to claim that the state of scientific knowledge is not uncertain but more
than likely as to their dangerousness. However, the ANSES for France (French
Agency for Food, Environmental and Occupational Health and Safety) and the
EFSA for Europe (European Food Safety Authority) do not require any
independent studies from industry or independent laboratories to quantify the
concentrations of herbicides contained in foodstuffs, and to control the
toxicological effects of the lifetime intake of these foods by animals. Launched in
2007 by the French Ministry of the Environment, the Ecophyto plan provided for
a 50 % reduction in pesticide use over ten years. However, the opposite has
happened: a 20% increase in pesticides between 2007 and 2017 (Nicolino and
Veillerette 2018). These mutant plants will necessarily accelerate this process.
The 67 activists were tried in Dijon (Burgundy) on 15 November 2018, and
the French edition of this toxicological review was sent to the defence lawyer and
then to the president of the court, at their request. The judgment was handed
down on 17 January 2019, and all the activists were released.
Declaration of interest statement
The author declares that there are no potential conflicts of interest.
!
32!
References
Adamczewski K, Paradowski A, Krawczyk R. 1998. Carryover effects of
amidosulfuron and tribenuron-methyl mixture on yielding of faba bean, sugar
beet and winter rape seeded in crop rotation. J Plant Prot Res, 38(1):89-92.
Adamczewski K., Paradowski A. 2004. Effect of adjuvants on biological efficacy of
sulfosulfuron and propoxycarbazone-sodium for weed control in winter wheat and
carryover effects. J Plant Prot Res, 44(4):347-363.
Aichele TM, Penner D. 2005. Adsorption, desorption, and degradation of
imidazolinones in soil. Weed Technol, 19(1):154-159.
Alexander DD, Weed DL, Mink PJ, Mitchell ME. 2012. A weight-of-evidence
review of colorectal cancer in pesticide applicators: the agricultural health study
and other epidemiologic studies. Int Arch Occup Environ Health, 85(7):715-745.
Alister C, Kogan M. 2005. Efficacy of imidazolinone herbicides applied to
imidazolinone-resistant maize and their carryover effect on rotational crops. Crop
Prot, 24:375-379.
Alonso-Prados JL, Hernández-Sevillano E, Llanos S, Villarroya M, García-Baudín
JM. 2002. Effects of sulfosulfuron soil residues on barley (Hordeum vulgare),
sunflower (Helianthus annuus) and common vetch (Vicia sativa). Crop Protection,
21:10611066.
Arya R, Mishra NK, Sharma AK. 2016. Brevibacillus borstelensis and
Streptomyces albogriseolus have roles to play in degradation of herbicide,
sulfosulfuron. 3 Biotech, 6(2):246.
Aubert S, Bligny R, Day DA, Whelan J, Douce R. 1997. Induction of alternative
oxidase synthesis by herbicides inhibiting branched-chain amino acid synthesis.
Plant J, 11(4):649-657.
Balabanova D, Remans T, Vassilev A, Cuypers A, Vangronsveld J. 2018. Possible
involvement of glutathione S-transferases in imazamox detoxification in an
imidazolinone-resistant sunflower hybrid. J Plant Physiol, 221:62-65.
Ball DA, Yenish JP, Alby T. 2003. Effect of imazamox soil persistence on dryland
rotational crops. Weed Technol, 17(1):161-165.
Barnes CJ, Goetz AJ, Lavy TL. 1989. Effects of imazaquin residues on cotton
(Gossypium hirsitum). Weed Sci, 37(6):820-824.
Barrett M. 1995. Metabolism of herbicides by cytochrome P450 in corn. Drug
Metabol Drug Interact, 12:299-315.
Battaglin WA, Furlong ET, Burkhardt MR, Peter CJ. 2000. Occurrence of
sulfonylurea, sulfonamide, imidazolinone, and other herbicides in rivers,
reservoirs and ground water in the Midwestern United States, 1998. Sci Total
Environ, 248(2-3):123-133.
!
33!
Battaglin WA, Rice KC, Focazio MJ, Salmons S, Barry RX. 2009. 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, 155(1-4):281-307.
Bedmar F, Perdigon JA, Monterubbianesi MG. 2006. Residual phytotoxicity and
persistence of chlorimuron and metsulfuron in soil of Argentina. J Env Biol,
27(2):175-179.
Bernieri T, Rodrigues D, Randon Barbosa I, Perassolo MS, Grolli Ardenghi P,
Basso da Silva L. 2019. Effect of pesticide exposure on total antioxidant capacity
and biochemical parameters in Brazilian soybean farmers. Drug Chem Toxicol,
doi: 10.1080/01480545.2019.1566353. [Epub ahead of print].
Bi Y, Han L, Song S, Yao W, Qin F, Xu Y, Wu Q. 2020. Method validation, storage
stability and field trial for residues of florasulam and pyroxsulam in cereal by
liquid chromatography with tandem mass spectrometry. Food Addit Contam Part
A, 2020 Feb 20:1-11.
Boldt TS, Jacobsen CS. 1998. Different toxic effects of the sulfonylurea
herbicides metsulfuron methyl, chlorsulfuron and thifensulfuron methyl on
fluorescent pseudomonads isolated from an agricultural soil. FEMS Microbiol Lett,
161:29-35.
Börjesson E, Torstensson L, Stenström J. 2004. The fate of imazapyr in a
Swedish railway embankment. Pest Manag Sci, 60(6):544-549.
Bretaud S, Toutant JP, Saglio P. 2000. Effects of carbofuran, diuron and
nicosulfuron on acetylcholinesterase activity in goldfish (Carassius auratus).
Ecotoxicol Environ Saf, 47:117124.
Buerge IJ, Bächli A, Kasteel R, Portmann R, López-Cabeza R, Schwab LF, Poiger
T. 2019. Behavior of the chiral herbicide imazamox in soils: pH-dependent,
enantioselective degradation, formation and degradation of several chiral
metabolites. Environ Sci Technol, 21;53(10):5725-5732.
Burgos N.R., Norsworthy J.K., Scott R.C., Smith K.L. 2008. Red rice status after
five years of Clearfield rice technology in Arkansas. Weed Technol, 22:200208.
Burgos NR, Singh V, Tseng TM, Black H, Young ND, Huang Z, Hyma KE, Gealy
DR, Caicedo AL. 2014. The impact of herbicide-resistant rice technology on
phenotypic diversity and population structure of United States weedy rice. Plant
Physiol, 166(3):1208-1220.
Busconi M, Rossi D, Lorenzoni C, Baldi G, Fogher C. 2012. Spread of herbicide-
resistant weedy rice (red rice, Oryza sativa L.) after 5 years of Clearfield rice
cultivation in Italy. Plant Biol, 14:751-759.
Carvalho WF, Ruiz de Arcaute C, Pérez-Iglesias JM, Laborde MRR, Soloneski S,
Larramendy ML. 2019. DNA damage exerted by mixtures of commercial
formulations of glyphosate and imazethapyr herbicides in Rhinella arenarum
(Anura, Bufonidae) tadpoles. Ecotoxicology, 28(3):367-377.
!
34!
Cessna AJ, Donald DB, Bailey J, Waiser M, Headley JV. 2006. Persistence of the
sulfonylurea herbicides thifensulfuron-methyl, ethametsulfuron-methyl and
metsulfuron-methyl in farm dugouts (ponds). J Environ Qual, 35:2395-2401.
Cessna AJ, Elliott JA, Bailey J. 2012. Leaching of three imidazolinone herbicides
during sprinkler irrigation. J Environ Qual, 41(3):882-892.
Cessna AJ, Donald DB, Bailey J, Waiser M. 2015. Persistence of the sulfonylurea
herbicides sulfosulfuron, rimsulfuron, and nicosulfuron in farm dugouts (ponds). J
Environ Qual, 44(6):1948-1955.
Cessna AJ, Knight JD, Ngombe D, Wolf TM. 2017. Effect of temperature on the
dissipation of seven herbicides in a biobed matrix. Can. J. Soil Sci, 97:717-731
Chen Y, Du L, Li L, Ma J, Geng X, Yao X, Liu G, Sun X. 2017. Cancer risk of
sulfonylureas in patients with type 2 diabetes mellitus: A systematic review. J
Diabetes, 9(5):482-494.
Chen J, Saleem M, Wang C, Liang W, Zhang Q. 2018. Individual and combined
effects of herbicide tribenuron-methyl and fungicide tebuconazole on soil
earthworm Eisenia fetida. Sci Rep, 8(1):2967.
Cheng Y, Wang Q, Li Z, Cui J, Hu S, Zhao H, Chen M. 2013. Cytological and
comparative proteomic analyses on male sterility in Brassica napus L. induced by
the chemical hybridization agent monosulphuron ester sodium. PLoS ONE,
8(11):e80191.
Cheng Y, Cui J, Li Z, Hu Z, Xing Z, Wang J, Zhao H, Hu S. 2015. The
sulphonylurea herbicide monosulphuron ester sodium as a special male
gametocide in Brassica napus L. Czech J Genet Plant Breed, 51(1):1622.
Colborn T, Carroll LE. 2007. Pesticides, sexual development, reproduction, and
fertility: current perspective and future direction. Hum Ecol Risk Assess,
13:10781110.
Chevassus-auLouis N. 2006. Les briseurs de machines. De Ned Ludd à José
Bové. Éditions du Seuil, collection « Science ouverte ». Paris, France.
Cranmer BK, Westra P, Zimdahl RL. 1999. Sorption, dissipation, and leaching of
metsulfuron in Colorado soils. Weed Sci, 47(3):353-359.
Curran, WC, Liebl RA, Simmons FW. 1992. Effects of tillage and application
method on clomazone, imazaquin, and imazethapyr persistence. Weed Sci,
40(3):482-489.
Cutulle MA, Armel GR, Kopsell DA, Wilson HP, Brosnan JT, Vargas JJ, Hines TE,
Koepke-Hill RM. 2018. Several pesticides influence the nutritional content of
sweet corn. J Agric Food Chem, 66(12):3086-3092.
Das M, Reichman JR, Haberer G, Welzl G, Aceituno FF, Mader MT, Watrud LS,
Pfleeger TG, Gutiérrez RA, Schäffner AR, Olszyk DM. 2010. A composite
!
35!
transcriptional signature differentiates responses towards closely related
herbicides in Arabidopsis thaliana and Brassica napus. Plant Mol Biol, 72(4-
5):545-556.
da Costa Marinho MI, Costa AIG, Vieira NM, Paiva MCG, de Freitas FCL, da Silva
AA. 2019. Validation and application of a QuEChERS based method for estimation
of half-life of imidazolinone herbicides in soils by LC-ESI-MS/MS. Ecotoxicol
Environ Saf, 167:212-217.
da Silva AA, de Oliveira Jr RS, Costa ER, Ferreira LR. 1999. Soil residual effect of
the herbicides imazamox and imazethapyr to corn and sorghum crops. Planta
Daninha, 17(3):345-354.
Dauer J, Hulting A, Carlson D, Mankin L, Harden J, Mallory-Smith C. 2018. Gene
flow from single and stacked herbicide-resistant rice (Oryza sativa): modeling
occurrence of multiple herbicide-resistant weedy rice. Pest Manag Sci,
74(2):348-355.
Degenhardt D, Cessna AJ, Raina R, Pennock DJ, Farenhorst A. 2010. Trace level
determination of selected sulfonylurea herbicides in wetland sediment by liquid
chromatography electrospray tandem mass spectrometry. J Environ Sci Health B,
45(1):11-24.
de Lafontaine Y, Beauvais C, Cessna AJ, Gagnon P, Hudon C, Poissant L. 2014.
Sulfonylurea herbicides in an agricultural catchment basin and its adjacent
wetland in the St. Lawrence River basin. Sci Total Environ, 479-480:1-10.
Desai NC, Bhavsar AM, Baldaniya BB. 2009. Synthesis and antimicrobial activity
of 5-imidazolinone derivatives. Indian J Pharm Sci, 71(1):90-94.
de Santo FB, Ramos GA, Ricardo Filho AM, Marchioro CA, Niemeyer JC. 2018.
Screening effects of metsulfuron-methyl to collembolans and earthworms: the
role of adjuvant addition on ecotoxicity. Environ Sci Pollut Res Int,
25(24):24143-24149.
de Santo FB, Guerra N, Vianna MS, Torres JPM, Marchioro CA, Niemeyer JC.
2019. Laboratory and field tests for risk assessment of metsulfuron-methyl-
based herbicides for soil fauna. Chemosphere, 222:645-655.
Ding F, Liu W, Zhang X, Zhang L, Sun Y. 2010. Fluorescence and circular
dichroism studies of conjugates between metsulfuron-methyl and human serum
albumin. Colloids Surf B Biointerfaces, 76(2):441-448.
Domínguez-Mendez R, Alcántara-de la Cruz R, Rojano-Delgado AM, Fernández-
Moreno PT, Aponte R, De Prado R. 2017. Multiple mechanisms are involved in
new imazamox-resistant varieties of durum and soft wheat. Sci Rep, 7(1):14839.
Donald DB, Cessna AJ, Sverko E, Glozier NE. 2007. Pesticides in surface
drinking-water supplies of the northern Great Plains. Environ Health Perspect,
115(8):1183-1191.
!
36!
dos Santos Miron D, Crestani M, Rosa Shettinger M, Maria Morsch V,
Baldisserotto B, Angel Tierno M, Moraes G, Vieira VL. 2005. Effects of the
herbicides clomazone, quinclorac, and metsulfuron methyl on
acetylcholinesterase activity in the silver catfish (Rhamdia quelen)
(Heptapteridae). Ecotoxicol Environ Saf, 61(3):398-403.
Duggleby RG, Pang SS, Yu H, Guddat LW. 2003. Systematic characterization of
mutations in yeast acetohydroxyacid synthase. Interpretation of herbicide-
resistance data. Eur J Biochem, 270(13):2895-2904.
Duhoux A, Délye C (2013) Reference genes to study herbicide stress response in
Lolium sp.: up-regulation of P450 genes in plants resistant to acetolactate-
synthase inhibitors. PLoS ONE 8:e63576
Duhoux A, Carrere S, Gouzy J, Bonin L, Delye C. 2015. RNA-Seq analysis of rye-
grass transcriptome response to an herbicide inhibiting acetolactate-synthase
identifies transcripts linked to non-target- site-based resistance. Plant Mol Biol,
87:473487.
Eizenberg H, Goldwasser Y, Achdary G, Hershenhorn J. 2003. The potential of
sulfosulfuron to control troublesome weeds in tomato. Weed Technology,
17(1):133-137.
El-Nahas AI. 2000. Mutagenic potential of imazethapyr herbicide (pursuit) on
Vicia faba in the presence of urea fertilizer. Pak J Biol Sci, 3:900905.
Emfinger CH, Welscher A, Yan Z, Wang Y, Conway H, Moss JB, Moss LG, Remedi
MS, Nichols CG. 2017. Expression and function of ATP-dependent potassium
channels in zebrafish islet β-cells. R Soc Open Sci, 4(2):160808.
Fahl GM, Kreft L, Altenburger R, Faust M, Boedeker W, Grimme LH. 1995. pH-
Dependent sorption, bioconcentration and algal toxicity of sulfonylurea
herbicides. Aquat Toxicol, 31:175-187.
Fathy M, Mohamed IA, Farghal AIA, Temerak SAH, Sayed AEH. 2019. Hemotoxic
effects of some herbicides on juvenile of Nile tilapia Oreochromis niloticus.
Environ Sci Pollut Res Int, 26(30):30857-30865.
Fleck NG, Vidal RA. 1994. Potential sunflower injury by soil applied herbicides. III
- Imazaquin and imazethapyr. Planta Daninha, 12(1):39-43.
Flint JL, Witt WW. 1997. Microbial degradation of imazaquin and imazethapyr.
Weed Sci, 45(4):586-591.
Fort DJ, Rogers R, Copley H, Bruning L, Stover EL, Rapaport D. 1999. Effect of
sulfometuron methyl and nicosulfuron on development and metamorphosis in
Xenopus laevis: impact of purity. Environ Toxicol Chem, 18(12): 2934-2940.
Fragiorge EJ, Rezende AA, Graf U, Spanó MA. 2008. Comparative genotoxicity
evaluation of imidazolinone herbicides in somatic cells of Drosophila
melanogaster. Food Chem Toxicol, 46(1):393-401.
!
37!
Fredrickson DR, Shea PJ. 1986. Effect of soil pH on degradation, movement and
plant uptake of chlorsulfuron. Weed Sci, 34:328-332.
Gabard J, Huby JP. 2001. Sulfonylurea-tolerant sunflower plants. World
Intellectual Property Organization. International publication number WO
01/65922 A2. Publication date: 13th September 2001.
Garry, V. F., Schreinemachers, D., Harkins, M. E., and Griffith, J. 1996. Pesticide
appliers, biocides and birth defects in rural Minnesota. Environ. Health Perspect,
104:394399.
Garry VF, Harkins ME, Lyubimov A, Erickson L, Long L. 2002. Reproductive
outcomes in the women of the Red River Valley of the north. I. The spouses of
pesticide applicators: pregnancy loss, age at menarche, and exposures to
pesticides. J Toxicol Environ Health A, 65:769-786.
Gaston, S.; Zabalza, A.; Gonzalez, E. M.; Arrese-Igor, C.; Aparicio-Tejo, P. M.;
Royuela, M. Imazethapyr, an inhibitor of the branched-chain amino acid
biosynthesis, induces aerobic fermentation in pea plants. Physiol. Plant. 2002,
114, 524532.
Gazziero DLP, Karan D, Voll E, Ulbrich A. 1997. Persistence of imazaquin and
imazethapyr in the soil , and their effects on corn and cucumber plants. Planta
Daninha, 15(2):162-168.
Ghosh B, Bhattacharyya A, Mukherjee S, Das S, Bhattacharyya A. 2014. Residual
fate and persistence behaviour of a mixed herbicide formulation (pyroxsulam
4.5% OD + sulfosulfuron 75% WDG) in wheat plant and field soil. J Crop Weed,
10(2):407-413.
Ghisari M, Long M, Tabbo A, Bonefeld-Jørgensen EC. 2015. Effects of currently
used pesticides and their mixtures on the function of thyroid hormone and aryl
hydrocarbon receptor in cell culture. Toxicol Appl Pharmacol, 284(3):292-303.
Gianelli VR, Bedmar F, Costa JL. 2014. Persistence and sorption of imazapyr in
three Argentinean soils. Environ Toxicol Chem, 33(1):29-34.
Glombitza S,Dubuis PH, Thulke O, Welzl G, Bovet L, Götz M, et al. 2004.
Crosstalk and differential response to abiotic and biotic stressors reflected at the
transcriptional level of effector genes from secondary metabolism. Plant Mol Biol,
54:817835.
Goetz AJ, Lavy TL, Gbur EE. 1990. Degradation and field persistence of
imazethapyr. Weed Sci, 38:421-428.
Gressel J. 2009. Crops with target-site herbicide resistance for Orobanche and
Striga control. Pest Manag Sci, 65(5):560-565.
Grisolia CK. 2002. A comparison between mouse and fish micronucleus test using
cyclophosphamide, mitomycin C and various pesticides. Mutat Res, 518:145
150.
!
38!
Grisolia CK, Bilich MR, Formigli LM. 2004. A comparative toxicologic and
genotoxic study of the herbicide arsenal, its active ingredient Imazapyr, and the
surfactant nonylphenol ethoxylate. Ecotoxicol Environ Safe, 59:123126.
Grossmann K, Berghaus R, Retzlaff G. 1992. Heterotrophic plant cell suspension
cultures for monitoring biological activity in agrochemical research. Comparison
with screens using algae, germinating seeds and whole plants. Pestic Sci,
35:283-289.
Gunaratne K, Austin E, Wu PE. 2018. Unintentional sulfonylurea toxicity due to a
drug-drug interaction: a case report. BMC Res Notes, 11(1):331.
Hackenberger DK, Stjepanović N, Lončarić Ž, Hackenberger BK. 2018. Acute and
subchronic effects of three herbicides on biomarkers and reproduction in
earthworm Dendrobaena veneta. Chemosphere, 208:722-730.
Hatzios KK, Burgos N. 2004. Metabolism-based herbicide resistance: regulation
by safeners. Weed Sci, 52:454-467.
Health Canada. 1999. Novel food information Food biotechnology. Imazethapy
tolerant corn, EXP1910IT. FD/OFB-097-189-A. October 1999. Office of Food
Biotechnology, Ottawa, Ontario.
Health Canada. Metsulfuron methyl. PRVD2008-08. 28 February 2008. Health
Canada Pest Management Regulatory Agency, Ottawa, Ontario.
Health Canada. Imazethapyr. PRVD2010-02. March 3, 2010. Pest Management
Regulatory Agency, Ottawa.
Heap I. 2014. Global perspective of herbicide-resistant weeds. Pest Manag Sci,
70(9):1306-1315.
Heres-Pulido ME, Lombera-Hernández S, Dueñas-García I, Perales-Canales I,
Castañeda-Partida L, Rocha-Ortiz C, Flores-Maya S, Durán-Díaz A, Graf U. 2008.
Genotoxicity of triasulfuron in the wing spot test of Drosophila melanogaster is
modulated by winter wheat seedlings. Mutat Res, 653(1-2):70-75.
Hu T, He S, Yang G, Zeng H, Wang G, Chen Z, Huang X. 2011. Isolation and
characterization of a rice glutathione S-transferase gene promoter regulated by
herbicides and hormones. Plant Cell Rep, 30(4):539-549.
Hu M, Liu K, Qiu J, Zhang H, Li X, Zeng D, Tan H. 2020. Behavior of
imidazolinone herbicide enantiomers in earthworm-soil microcosms: Degradation
and bioaccumulation. Sci Total Environ, 707:135476.
Jabran K, Ehsanullah, Hussain M, Farooq M, Babar M, Doğan MN, Lee DJ. 2012.
Application of bispyribac-sodium provides effective weed control in direct-planted
rice on a sandy loam soil. Weed Biology Management, 12:136-145.
James TK, Klaffenbach P, Holland PT, Rahman A. 1995. Degradation of
primisulfuron-methyl and metsulfuron-methyl in soil. Weed Res, 35(2):113-120.
!
39!
Jin J, Kurobe T, Ramírez-Duarte WF, Bolotaolo MB, Lam CH, Pandey PK, Hung
TC, Stillway ME, Zweig L, Caudill J, Lin L, Teh SJ. 2018. Sub-lethal effects of
herbicides penoxsulam, imazamox, fluridone and glyphosate on Delta Smelt
(Hypomesus transpacificus). Aquat Toxicol, 197:79-88.
Joly P, Bonnemoy F, Charvy JC, Bohatier J, Mallet C. 2013. Toxicity assessment
of the maize herbicides S-metolachlor, benoxacor, mesotrione and nicosulfuron,
and their corresponding commercial formulations, alone and in mixtures, using
the Microtox® test. Chemosphere, 93(10):2444-2450.
Jovanović-Rodovanov K. 2017. Imazethapyr persistence in sandy loam detected
using white mustard bioassay. J Environ Sci Health B, 52(10):711-718.
Kalsi NK, Kaur P. 2019. Dissipation of bispyribac sodium in aridisols: Impact of
soil type, moisture and temperature. Ecotoxicol Environ Saf, 170:375-382.
Karier P, Kraus G, Kolber I. 2017. Metazachlor traces in the main drinking water
reservoir in Luxembourg: a scientific and political discussion. Environ Sci Eur,
29(1):25.
Kemmerich M, Bernardi G, Adaime MB, Zanella R, Prestes OD. 2015. A simple
and efficient method for imidazolinone herbicides determination in soil by ultra-
high performance liquid chromatography-tandem mass spectrometry. J
Chromatogr A, 1412:82-89.
Kim S, Han S, Vanden Born WH. 1997. Effect of chlorsulfuron on assimilate
transport: ultrastructural implications. Weed Sci, 45(4):470-473.
Kolkman JM, Slabaugh MB, Bruniard JM, Berry S, Bushman BS, Olungu C, Maes
N, Abratti G, Zambelli A, Miller JF, Leon A, Knapp SJ. 2004. Acetohydroxyacid
synthase mutations conferring resistance to imidazolinone or sulfonylurea
herbicides in sunflower. Theor Appl Genet, 109(6):1147-1159.
Kotoula-Syka E, Eleftherohorinos IG, Gagianas AA, Sficas AG. 1993. Persistence
of preemergence applications of chlorsulfuron, metsulfuron, triasulfuron, and
tribenuron in three soils in Greece. Weed Sci, 41(2):246-250.
Koutros S, Lynch CF, Ma X, Lee WJ, Hoppin JA, Christensen CH, Andreotti G,
Freeman LB, Rusiecki JA, Hou L, Sandler DP, Alavanja MC. 2009. Heterocyclic
aromatic amine pesticide use and human cancer risk: results from the U.S.
Agricultural Health Study. Int J Cancer, 124(5):1206-1212.
Koutros S, Silverman DT, Alavanja MC, Andreotti G, Lerro CC, Heltshe S, Lynch
CF, Sandler DP, Blair A, Beane Freeman LE. 2016. Occupational exposure to
pesticides and bladder cancer risk. Int J Epidemiol, 45(3):792-805.
Kraemer AF, Marchesan E, de Avila LA, Machado SLO, Grohs M. 2009a.
Environmental fate of imidazolinone herbicides A review. Planta Daninha,
27(3):629-639.
Kraemer AF, Marchesan E, Grohs M, de Avila LA, Machado SLO, Zanella R,
Massoni PFS, Sartori GMS. 2009b. Imazethapyr leaching in rice paddy soil under
!
40!
two tillage systems. Cienc. Rural, 39:1660-1666.
Kumar U, Behera S, Saha S, Das D, Guru PK, Kaviraj M, Munda S, Adak T, Nayak
AK. 2020. Non-target effect of bispyribac sodium on soil microbial community in
paddy soil. Ecotoxicol Environ Saf, 189:110019.
Lee HL, Chen KW, Wu MH. 1999. Acute poisoning with a herbicide containing
imazapyr (Arsenal): a report of six cases. J Toxicol Clin Toxicol, 37(1):83-89.
Lee WJ, Sandler DP, Blair A, Samanic C, Cross AJ, Alavanja MC. 2007. Pesticide
use and colorectal cancer risk in the Agricultural Health Study. Int J Cancer,
121(2):339-346.
Li Z, Cheng Y, Cui J, Zhang P, Zhao H, Hu S. 2015. Comparative transcriptome
analysis reveals carbohydrate and lipid metabolism blocks in Brassica napus L.
male sterility induced by the chemical hybridization agent monosulfuron ester
sodium. BMC Genomics, 16:206.
Lian JL, Ren LS, Zhang C, Yu CY, Huang Z, Xu AX, Dong JG. 2019. How exposure
to ALS-inhibiting gametocide tribenuron-methyl induces male sterility in
rapeseed. BMC Plant Biol, 19(1):124.
Liman R, Ciğerci İH, Öztürk NS. 2015. Determination of genotoxic effects of
Imazethapyr herbicide in Allium cepa root cells by mitotic activity, chromosome
aberration, and comet assay. Pestic Biochem Physiol, 118:38-42.
Liu X, Xu J, Li Y, Dong F, Li J, Song W, Zheng Y. 2011. Rapid residue analysis of
four triazolopyrimidine herbicides in soil, water, and wheat by ultra-performance
liquid chromatography coupled to tandem mass spectrometry. Anal Bioanal
Chem, 399(7):2539-2547.
Liu X, Xu X, Li B, Wang X, Wang G, Li M. 2015. RNA-Seq transcriptome analysis
of maize inbred carrying nicosulfuron-tolerant and nicosulfuron-susceptible
alleles. Int J Mol Sci, 16(3):5975-5989.
Liu XQ, Yu CY, Dong JG, Hu SW, Xu AX. 2017. Acetolactate synthase-inhibiting
gametocide amidosulfuron causes chloroplast destruction, tissue autophagy, and
elevation of ethylene release in aapeseed. Front Plant Sci, 8:1625.
López-Piñeiro A, Sánchez-Llerena J, Peña D, Albarrán A, Ramírez M. 2016.
Transition from flooding to sprinkler irrigation in Mediterranean rice growing
ecosystems: effect on behaviour of bispyribac sodium. Agric Ecosyst Environ,
223:99-107.
Loux MM, Reese KD. 1993. Effect of soil type and pH on persistence and
carryover of imidazolinone herbicides. Weed Technol, 7(2):452-458.
Magdaleno A, Peralta Gavensky M, Fassiano AV, Ríos de Molina MC, Santos M,
March H, Moretton J, Juárez ÁB. 2015. Phytotoxicity and genotoxicity assessment
of imazethapyr herbicide using a battery of bioassays. Environ Sci Pollut Res Int,
22(23):19194-19202.
!
41!
Mallory-Smith CA and Retzinger EJ Jr. 2003. Revised classification of herbicides
by site of action for weed resistance management strategies. Weed Technol,
17(3):605619.
Manabe Y, Tinker N, Colville A, Miki B. 2007. CSR1, the sole target of
imidazolinone herbicide in Arabidopsis thaliana. Plant Cell Physiol, 48(9):1340-
1358.
Marchesan E, dos Santos FM, Grohs M, de Avila LA, Machado SLO, Senseman SA,
Massoni PFS, Sartori GMS. 2010. Carryover of imazethapyr and imazapic to
nontolerant rice. Weed Technol, 24:6-10.
Marsh BH, Lloyd RW. 1996. Soil pH effect on imazaquin persistence in soil. Weed
Technol, 10(2):337-340.
Mehdizadeh M, Alebrahim MT, Roushani M, Streibig JC. 2016. Evaluation of four
different crops’ sensitivity to sulfosulfuron and tribenuron methyl soil residues.
Acta Agr Scand B, 66(8):706-713.
Mehdizadeh M, Alebrahim MT, Roushani M. 2017. Determination of two
sulfonylurea herbicides residues in soil environment using HPLC and phytotoxicity
of these herbicides by lentil bioassay. Bull Environ Contam Toxicol, 99(1):93-99.
Mills JA, Witt WW. 1989. Efficacy, phytotoxicity, and persistence of imazaquin,
imazethapyr, and clomazone in non-till doublecrop soybeans (Glycine max).
Weed Sci, 37(3):353-359.
Moraes BS, Clasen B, Loro VL, Pretto A, Toni C, de Avila LA, Marchezan E, de
Oliveira Machado S, Zanella R, Reimche GB. 2011. Toxicological responses of
Cyprinus carpio after exposure to a commercial herbicide containing imazethapyr
and imazapic. Ecotoxicol Environ Saf, 74(3):328-335.
Moyer JR. 1995. Sulfonylurea herbicide effects on following crops. Weed Technol,
9(2):373-379.
Moyer JR, Esau R. 1996. Imidazolinone herbicide effects on following rotational
crops in Southern Alberta. Weed Technol, 10(1):100-106.
Müller F, Applebyki AP; Weed Control, 2. Individual Herbicides. Ullmann's
Encyclopedia of Industrial Chemistry. 7th ed. (2012). New York, NY: John Wiley
& Sons.
Nicolino F, Veillerette F. 2018. Nous voulons des coquelicots. Éditions Les Liens
qui Libèrent, Paris.
Ning L, Lin Z, Gu J, Gan L, Li Y, Wang H, Miao L, Zhang L, Wang B, Li M. 2018.
The initial deficiency of protein processing and flavonoids biosynthesis were the
main mechanisms for the male sterility induced by SX-1 in Brassica napus. BMC
Genomics, 19(1):806.
Norsworthy JK, Griffith GM, Scott RC. 2008. Imazethapyr use with and without
clomazone for weed control in furrow-irrigated, imidazolinone-tolerant rice. Weed
!
42!
Technology, 22:217221.
Nyström B, Björnsäter B, Blanck H. 1999. Effects of sulfonylurea herbicides on
non-target aquatic micro-organisms. Growth inhibition of micro-algae and short-
term inhibition of adenine and thymidine incorporation in periphyton
communities. Aquat Toxicol, 47:9-22.
Pan G, Zhang XY, Liu KD, Zhang JW, Wu XZ, Zhu J, Tu JM. 2006. Map-based
cloning of a novel rice cytochrome P450 gene CYP81A6 that confers resistance to
two different classes of herbicides. Plant Mol Biol, 61:933-943.
Paporisch A, Rubin B. 2017. Isoxadifen safening mechanism in sweet corn
genotypes with differential response to P450-metabolized herbicides. Pestic
Biochem Physiol, 138, 22-28.
Pasquer, F., Ochsner, U., Zarn, J., and Keller, B. 2006. Common and distinct
gene expression patterns induced by the herbicides 2, 4-dichlorophenoxyacetic
acid, cinidon-ethyl and tribenuron-methyl in wheat. Pest Manag Sci, 2:1155-
1167.
Pérez-Iglesias JM, Soloneski S, Nikoloff N, Natale GS, Larramendy ML. 2015.
Toxic and genotoxic effects of the imazethapyr-based herbicide formulation Pivot
H® on montevideo tree frog Hypsiboas pulchellus tadpoles (Anura, Hylidae).
Ecotoxicol Environ Saf, 119:15-24.
Pérez-Iglesias JM, Ruiz de Arcaute C, Natale GS, Soloneski S, Larramendy ML.
2017. Evaluation of imazethapyr-induced DNA oxidative damage by alkaline Endo
III- and Fpg-modified single-cell gel electrophoresis assay in Hypsiboas
pulchellus tadpoles (Anura, Hylidae). Ecotoxicol Environ Saf, 142:503-508.
Pérez-Iglesias JM, Natale GS, Soloneski S, Larramendy ML. 2018. Are the
damaging effects induced by the imazethapyr formulation Pivot
® H in Boana
pulchella (Anura) reversible upon ceasing exposure? Ecotoxicol Environ Saf,
148:1-10.
Pérez-Iglesias JM, Brodeur JC, Larramendy ML. 2020. An imazethapyr-based
herbicide formulation induces genotoxic, biochemical, and individual
organizational effects in Leptodactylus latinasus tadpoles (Anura:
Leptodactylidae). Environ Sci Pollut Res Int, 27(2):2131-2143.
Persans MW, Schuler MA. 1995. Differential induction of cytochrome P450-
mediated triasulfuron metabolism by naphthalic anhydride and triasulfuron. Plant
Physiol, 109:1483-1490.
Peterson HG, Boutin C, Martin PA, Freemark KE, Ruecker NJ, Moody MJ. 1994.
Aquatic phyto-toxicity of 23 pesticides applied at expected environmental
concentrations. Aquatic Toxicol, 28(3-4):275-292.
Pinto JJO, Noldin JA, Pinho CF, Rossi F, Galon L, Almeida GF. 2009a. Field
persistence of (imazethapyr + imazapic) to grain sorghum (Sorghum bicolor)
planted in rotation after irrigated rice. Planta Daninha, 27:1015-1024.
!
43!
Pinto JJO, Noldin JA, Rosenthal MD, Pinho CF, Rossi F, Machado A, Piveta L,
Galon L. 2009b. Residual activity of (imazethapyr + imazapic) on ryegrass
(Lolium multiflorum), following Clearfield rice. Planta Daninha, 27:609-619.
Powles SB, Yu Q. 2010. Evolution in action: plants resistant to herbicides. Annu
Rev Plant Biol, 61:317-347.
Pretto A, Loro VL, Menezes C, Moraes BS, Reimche GB, Zanella R, de Ávila LA.
2011. Commercial formulation containing quinclorac and metsulfuron-methyl
herbicides inhibit acetylcholinesterase and induce biochemical alterations in
tissues of Leporinus obtusidens. Ecotoxicol Environ Saf, 74(3):336-341.
Puga A, Ma C, Marlowe JL. 2009. The aryl hydrocarbon receptor cross-talks with
multiple signal transduction pathways. Biochem Pharmacol, 77:713-722.
Qian H, Hu H, Mao Y, Ma J, Zhang A, Liu W, Fu Z. 2009. Enantioselective
phytotoxicity of the herbicide imazethapyr in rice. Chemosphere, 76(7):885-892.
Qian H, Lu T, Peng X, Han X, Fu Z, Liu W. 2011a. Enantioselective phytotoxicity
of the herbicide imazethapyr on the response of the antioxidant system and
starch metabolism in Arabidopsis thaliana. PLoS One, 6(5):e19451.
Qian H, Wang R, Hu H, Lu T, Chen X, Ye H, Liu W, Fu Z. 2011b. Enantioselective
phytotoxicity of the herbicide imazethapyr and its effect on rice physiology and
gene transcription. Environ Sci Technol, 45(16):7036-7043.
Qian H, Li Y, Sun C, Lavoie M, Xie J, Bai X, Fu Z. 2015a. Trace concentrations of
imazethapyr (IM) affect floral organs development and reproduction in
Arabidopsis thaliana: IM-induced inhibition of key genes regulating anther and
pollen biosynthesis. Ecotoxicology, 24(1):163-171.
Qian H, Lu H, Ding H, Lavoie M, Li Y, Liu W, Fu Z. 2015b. Analyzing Arabidopsis
thaliana root proteome provides insights into the molecular bases of
enantioselective imazethapyr toxicity. Sci Rep, 5:11975.
Rachedi K, Zermane F, Tir R, Ayache F, Duran R, Lauga B, Karama S, Simon M,
Boulahrouf A. 2018. Effect of sulfonylurea tribenuron methyl herbicide on soil
Actinobacteria growth and characterization of resistant strains. Braz J Microbiol,
49(1):79-86.
Rad MH, Aivazi AA, Jagannath S. 2011. Cytogenetic and biochemical effects of
imazethapyr in wheat (Triticum durum). Turk J Biol, 35:663-670.
Ragia G, Petridis I, Tavridou A, Christakidis D, Manolopoulos VG. 2009. Presence
of CYP2C9*3 allele increases risk for hypoglycemia in Type 2 diabetic patients
treated with sulfonylureas. Pharmacogenomics, 10(11):1781-1787.
Ramesh A, Elumalai TP, Sivanandam S. 2007a. Identification of residues of
sulfosulfuron and its metabolites in subsoil-dissipation kinetics and factors
influencing the stability and degradation of residues from topsoil to subsoil under
predominant cropping conditions. Environ Monit Assess, 130:519–528.
!
44!
Ramesh A, Sathiyanarayanan S, Chandran L. 2007b. Dissipation of sulfosulfuron
in water - bioaccumulation of residues in fish - LC-MS/MS-ESI identification and
quantification of metabolites. Chemosphere, 68(3):495-500.
Ramezani MK, Oliver DP, Kookana RS, Lao W, Gill G, Preston C. 2010. Faster
degradation of herbicidally-active enantiomer of imidazolinones in soils.
Chemosphere, 79(11):1040-1045.
Ramprakash T, Madhavi M, Yakadri M, Srinivas A. 2015. Bispyribac sodium
persistence in soil, plant and grain in direct seeded rice and its effect on soil
properties. Nat Environ Poll Tech, 14(3):605-609.
Rangel PHN, Neto FPM, Fagundes PRR, Magalhaes AM Jr, de Morais OP, Schmidt
AB, Mendonca JA, Santiago CM, Rengel PN, Cutrim VA, Ferreira ME. 2010.
Development of herbicide-tolerant irrigated rice cultivars. Pesq Agropecu Bras,
45:701708.
Rani D, Duhan A, Punia SS, Yadav DB, Duhan S. 2018. Behavior of pre-mix
formulation of imazethapyr and imazamox herbicides in two different soils.
Environ Monit Assess, 191(1):33.
Reimche GB, Machado SL, Oliveira MA, Zanella R, Dressler VL, Flores EM,
Gonçalves FF, Donato FF, Nunes MA. 2015. Imazethapyr and imazapic,
bispyribac-sodium and penoxsulam: zooplankton and dissipation in subtropical
rice paddy water. Sci Total Environ, 514:68-76.
Rojano-Delgado AM, João M. Portugal JM, Palma-Bautista C, Alcántara-de la Cruz
R, Torra J, Alcántara E, De Prado R. 2019. Target site as the main mechanism of
resistance to imazamox in a Euphorbia heterophylla biotype. Sci Rep, 9:15423.
Rouchaud J, Neus O, Cools K, Bulcke R. 1999. Metsulfuron-methyl soil
persistence and mobility in winter wheat and following green manure crops.
Toxicol Environ Chem, 71:369-381.
Royuela M, Gonzalez A, Gonzalez EM, Arrese-Igor C, Aparicio-Tejo PM, Gonzalez-
Murua C. 2000. Physiological consequences of continuous, sublethal imazethapyr
supply to pea plants. J. Plant Physiol, 157(3):345-354.
Sabater C, Carrasco JM. 1997. Effects of chlorsulfuron on growth of three
freshwater species of phytoplankton. Bull Environ Contam Toxicol, 58:807-813.
Sada Y, Ikeda H, Yamato S, Kizawa S. 2003. Characterization of sulfonylurea-
resistant Schoenoplectus juncoides having a target-site Asp376Glu mutation in the
acetolactate synthase. Pestic Biochem Physiol, 107(1):106-111.
Saglio P, Olsén KH, Bretaud S. 2001. Behavioral and olfactory responses to
prochloraz, bentazone, and nicosulfuron-contaminated flows in goldfish. Arch
Environ Contam Toxicol, 41(2):192-200.
Saglio P, Bretaud S, Rivot E, Olsén KH. 2003. Chemobehavioral changes induced
by short-term exposures to prochloraz, nicosulfuron, and carbofuran in goldfish.
Arch Environ Contam Toxicol, 45:515-524.
!
45!
Saha S, Roy S, Das TK, Bhattacharyya A. 2016. Dissipation kinetics of a new
mixture formulation of bispyribac-sodium and metamifop in rice. J Crop Weed,
12(1):129-134.
Saika H, Horita J, Taguchi-Shiobara F, Nonaka S, Nishizawa-Yokoi A, Iwakami S,
Hori K, Matsumoto T, Tanaka T, Itoh T, Yano M, Kaku K, Shimizu T, Toki S. 2014.
A novel rice cytochrome P450 gene, CYP72A31, confers tolerance to acetolactate
synthase-inhibiting herbicides in rice and Arabidopsis. Plant Physiol,
166(3):1232-1240.
Samanta P, Bandyopadhyay N, Pal S, Mukherjee AK, Ghosh AR. 2015.
Histopathological and ultramicroscopical changes in gill, liver and kidney of
Anabas testudineus (Bloch) after chronic intoxication of almix (metsulfuron
methyl 10.1% + chlorimuron ethyl 10.1%) herbicide. Ecotoxicol Environ Saf,
122:360-367.
Santos FM, Marchesan E, Machado SLO, Avila LA, Zanella R, Gonçalves FF. 2008.
Imazethapyr and clomazone persistence in rice paddy water. Planta Daninha,
26(4):875-881.
Schreinemachers DM, Creason JP, Garry VF. 1999. Cancer mortality in
agricultural regions of Minnesota. Environ Health Perspect, 107(3):205-211.
Seguin F, Leboulanger C, Rimet F, Druart JC, Bérard A. 2001. Effects of atrazine
and nicosulfuron on phytoplankton in systems of increasing complexity. Arch
Environ Contam Toxicol, 40(2):198-208.
Shimizu T, Kaku K, Kawai K, Miyazawa T, Tanaka Y. 2005. Molecular
characterization of acetolactate synthase in resistant weeds and crops. In JM
Clark, H Ohkawa, eds, Environmental fate and safety management of
agrochemicals. ACS Symposium Series, Vol 899. American Chemical Society,
Washington, DC, pp. 255-271.
Shinn SL, Thill DC, Price WJ, Ball DA. 1998. Response of downy brome (Bromus
tectorum) and rotational crops to MON 37500. Weed Technol, 12(4):690698.
Shivrain V.K., Burgos N.R., Rajguru S.N., Anders M.M., Moore J.W., Sales M.A.
2007. Gene flow between Clearfield rice and red rice. Crop Prot, 26:349356.
Shivrain VK, Burgos NR, Sales MA, Mauromoustakos A, Gealy DR, Smith KL,
Black HL, Jia M. 2009a. Factors affecting the outcrossing rate between Clearfield
rice (Oryza sativa L.) and red rice. Weed Sci, 57:394-403.
Shivrain VK, Burgos NR, Gealy DR, Sales MA, Smith KL. 2009b. Gene flow from
weedy red rice (Oryza sativa L.) to cultivated rice and fitness of hybrids. Pest
Manag Sci, 65:1124-1129.
Si Y, Wang S, Zhou J, Hua R, Zhou D. 2005. Leaching and degradation of
ethametsulfuron-methyl in soil. Chemosphere, 60(5):601-609.
!
46!
Silva FB, Costa AC, Müller C, Nascimento KT, Batista PF, Vital RG, Megguer CA,
Jakelaitis A, Domingos M. 2020. Dipteryx alata, a tree native to the Brazilian
Cerrado, is sensitive to the herbicide nicosulfuron. Ecotoxicology, 29(2):217-225.
Simoni F, Victoria Filho R, San Martin HAM, Salvador FL, Alves ASR, Bremer Neto
H. 2006. Efficacy of imazapic and sulfentrazone applied on Cyperus rotundus
under different rain and sugarcane straw conditions. Planta daninha, 24(4):769-
778.
Singh B, Singh K. 2016. Microbial degradation of herbicides. Crit Rev Microbiol,
42(2):245-261.
Singh SP, Rawal S, Dua VK, Sharma SK. 2017. Weed control efficiency of
herbicide sulfosulfuron in potato crop. Potato J, 44(2):110-116.
Soloneski S, Ruiz de Arcaute C, Nikoloff N, Larramendy ML. 2017. Genotoxicity of
the herbicide imazethapyr in mammalian cells by oxidative DNA damage
evaluation using the Endo III and FPG alkaline comet assays. Environ Sci Pollut
Res Int, 24(11):10292-10300.
Sondhia S. 2008. Determination of imazosulfuron persistence in rice crop and
soil. Environ Monit Assess, 137(1-3):205-211.
Sondhia S. 2008b. Dissipation of sulfosulfuron from wheat field and detection of
its residues in wheat grains and straw. Indian J Weed Sci, 40(3&4, supp):192-
194.
Struger J, Grabuski J, Cagampan S, Rondeau M, Sverko E, Marvin C. 2011.
Occurrence and distribution of sulfonylurea and related herbicides in central
Canadian surface waters 20062008. Bull Environ Contam Toxicol, 87:420-425.
Sudianto E, Song BK, Neik TX, Saldain NE, Scott RC, Burgos NR. 2013. Clearfield
rice: its development, success, and key challenges on a global perspective. Crop
Prot, 49:4051.
Sun L, Wu R, Su W, Gao Z, Lu C. 2017. Physiological basis for isoxadifen-ethyl
induction of nicosulfuron detoxification in maize hybrids. PLoS One,
12:e0173502.
Sun L, Xu H, Su W, Xue F, An S, Lu C, Wu R. 2018. The expression of
detoxification genes in two maize cultivars by interaction of isoxadifen-ethyl and
nicosulfuron. Plant Physiol Biochem, 129:101-108.
Tan S, Evans RR, Dahmer ML, Singh BK, Shaner DL. 2005. Imidazolinone-
tolerant crops: history, current status and future. Pest Manag Sci, 61(3):246-
257.
Tatum VL, Borton DL, Streblow WR, Louch J, Shepard JP. 2012. Acute toxicity of
commonly used forestry herbicide mixtures to Ceriodaphnia dubia and
Pimephales promelas. Environ Toxicol, 27(12):671-684.
!
47!
Thirunarayanan K, Zimdahl RL, Smika DE. 1985. Chlorsulfuron adsorption and
degradation in soil. Weed Sci, 33:558-563.
Thompson EP. 1966. The making of the English working class. Vintage books.
Alfred A. Knopf and Random House, New-York.
Toxnet: Toxicology data network. 2006. Metsulfuron methyl. 1st December 2006.
US National Library of Medicine, National Institutes of Health, Bethesda,
Maryland.
Toxnet: Toxicology data network. 2012. Imazethapyr. 6th January 2012. US
National Library of Medicine, National Institutes of Health, Bethesda, Maryland.
Ulbrich, A.V., Souza, R.P., Shaner, D., 2005. Persistence and carryover effect of
imazapic and imazapyr in Brazilian cropping systems. Weed Technol, 19:986-
991.
Ureta MS, Torres Carbonell F, Pandolfo C, Presotto AD, Cantamutto MA, Poverene
M. 2017. IMI resistance associated to crop-weed hybridization in a natural
Brassica rapa population: characterization and fate. Environ Monit Assess,
189(3):101.
US EPA. 2008, Sept 24. Environmental fate and ecological risk assessment.
Registration of New Use Imazamox for the proposed new use for the control of
weeds in Clearfield rice (page 14). United States Environmental Protection
Agency, PC Code: 129171. 91 pages, Washington DC.
Villa, S.C.C., Marchezan, E., Avila, L.A., Massoni, P.F.S., Telo, G.M., Machado,
S.L.O., Camargo, E.R., 2006. Imidazolinone-tolerant rice: red rice control, out-
crossing and herbicide carryover to non-tolerant crops. Planta Daninha,
24(4):761-768.
Vischetti C. 1995. Measured and simulated persistence of imazethapyr in soil.
Bull Environ Contam Toxicol, 54(3):420-427.
Walker A, Welch SJ. 1989. The relative movement and persistence in soil of
chlorsulfuron and metsulfuron methyl and triasulfuron. Weed Res, 29(5):375-
383.
Wang X, Wang H, Fan D. 2006. Degradation and metabolism of imazapyr in soils
under aerobic and anaerobic conditions. Int J Environ Anal Chem, 86(8):541-
551.
Wang J, Zhong X, Li F, Shi Z. 2018a. Effects of nicosulfuron on growth, oxidative
damage, and the ascorbate-glutathione pathway in paired nearly isogenic lines of
waxy maize (Zea mays L.). Pestic Biochem Physiol, 145:108-117.
Wang J, Zhong X, Zhu K, Lv J, Lv X, Li F, Shi Z. 2018b. Reactive oxygen species,
antioxidant enzyme activity, and gene expression patterns in a pair of nearly
isogenic lines of nicosulfuron-exposed waxy maize (Zea mays L.). Environ Sci
Pollut Res Int, 25(19):19012-19027.
!
48!
Weeraddana CS, Evenden ML. 2018. Canola nutrition and variety affect
oviposition and offspring performance in the generalist herbivore, Mamestra
configurata (Lepidoptera: Noctuidae). J Econ Entomol, 111(4):1702-1710.
Weichenthal S, Moase C, Chan P. 2010. A review of pesticide exposure and
cancer incidence in the Agricultural Health Study cohort. Environ Health Perspect,
118(8):1117-1125.
Wu Q, Chen X, Xu Y, Han L. 2010. Dissipation and residues of nicosulfuron in
corn and soil under field conditions. Bull Environ Contam Toxicol, 85(1):79-82.
Wu YH, Liu KT, Shih YL, Yeh IJ. 2019. Chronological serum concentration of
imazapyr in a case of life threatening intoxication. Clin Toxicol, 57(9):824-825.
Wright AA, Sasidharan R, Koski L, Rodriguez-Carres M, Peterson DG, Nandula
VK, Ray JD, Bond JA, Shaw DR. 2018. Transcriptomic changes in Echinochloa
colona in response to treatment with the herbicide imazamox. Planta,
247(2):369-379.
Xu J, Guo L, Dong F, Liu X, Wu X, Sheng Y, Zhang Y, Zheng Y. 2013. Response
of the soil microbial community to imazethapyr application in a soybean field. J
Environ Sci Health B, 48(6):505-511.
Yang Q, Li J, Shen J, Xu Y, Liu H, Deng W, Li X, Zheng M. 2018. Metabolic
resistance to acetolactate synthase inhibiting herbicide tribenuron-methyl in
Descurainia sophia L. mediated by cytochrome P450 enzymes. J Agric Food
Chem, 66(17):4319-4327.
Yousefi M, Kamkar B, Gherekhloo J, Faez R. 2016. Sulfosulfuron persistence in
soil under different cultivation systems of wheat (Triticum aestivum).
Pedosphere, 26(5):666-675.
Yu CY, Dong JG, Hu SW, Xu AX. 2017. Exposure to trace amounts of sulfonylurea
herbicide tribenuron-methyl causes male sterility in 17 species or subspecies of
cruciferous plants. BMC Plant Biol, 17(1):95.
Yu CY, Lian JL, Gong Q, Ren LS, Huang Z, Xu AX, Dong JG. 2020. Sublethal
application of various sulfonylurea and imidazolinone herbicides favors
outcrossing and hybrid seed production in oilseed rape. BMC Plant Biol, 20(1):69.
Yun MS, Shim IS, Usui K. 2001. Involvement of cytochrome P450 enzyme
activity in the selectivity and safening action of pyrazosulfuron-ethyl. Pest Manag
Sci, 57:283-288.
Zhang W.Q., Linscombe S.D., Webster E., Tan S.Y., Oard J. 2006. Risk
assessment of the transfer of imazethapyr herbicide resistance from Clearfield
rice to red rice (Oryza sativa). Euphytica, 152:7586.
Zhang C, Xu J, Liu X, Dong F, Kong Z, Sheng Y, Zheng Y. 2010. Impact of
imazethapyr on the microbial community structure in agricultural soils.
Chemosphere, 81(6):800-806.
!
49!
Zhang X, Li X, Zhang C, Li X, Zhang H. 2011. Ecological risk of long-term
chlorimuron-ethyl application to soil microbial community: an in situ
investigation in a continuously cropped soybean field in Northeast China. Environ
Sci Pollut Res Int, 18(3):407-415.
Zhang Q, Zhao Y, Fan S, Bai A, Li X, Pan C. 2013. Dissipation and residues of
bispyribac-sodium in rice and environment. Environ Monit Assess, 185(12):9743-
9749.
Zhang H, Zhao L. 2017. Influence of sublethal doses of acetamiprid and
halosulfuron-methyl on metabolites of zebra fish (Brachydanio rerio). Aquat
Toxicol, 191:85-94.
Zhao L, Jing X, Chen L, Liu Y, Su Y, Liu T, Gao C, Yi B, Wen J, Ma C, Tu J, Zou J,
Fu T, Shen J. 2015. Tribenuron-methyl induces male sterility through anther-
specific inhibition of acetolactate synthase leading to autophagic cell death. Mol
Plant, 8(12):1710-1724.
Zhou Q, Liu W, Zhang Y, Liu KK. 2007. Action mechanisms of acetolactate
synthase-inhibiting herbicides. Pestic Biochem Phys, 89:8996.
!
50!
Legend to figure
Figure 1. Synthetic panoptic of the toxicological effects of ALS-inhibiting
herbicides. The solid line arrows indicate the different effects of herbicides used
against herbicide-resistant plants. Dotted line arrows indicate possible
interactions between effects. For example, the onset of an oxidative stress in
turn leads to genotoxicity, and thus cancer and teratogenicity.
!
Herbicides
Oxidative stress
(earthworms, fish)
Genotoxicity
(drosophila,
tadpoles,
human cells)
Teratogenicity
(tadpoles)
Pharmacological effects
(various models)
SUR receptor
(fish, humans)
Mortality
(rotifers,
phytoplankton)
Avoidance strategy
(springtails,
earthworms)
Neurotransmitters
(fish)
Impact on
reproduction
(humans, plants
invertebrates)
Altered glycaemia
(fish, humans)
AhR receptor
(human cells)
Amino acids
(fish)
Cancers
(humans)
Deregulation of gene
and protein
expression (plants)
Chloroplast
destruction
(plants)
!
1!
Tables
Table 1. Effects of ALS-inhibiting herbicides on the transcriptome and the
regulation of gene expression in plants.
Species
Treatment
Reference
Arabidopsis
thaliana
weed
Arabidopsis plants
were sprayed with
750 µg/m2
primisulfuron-
methyl, and 1.5
µg/m2 prosulfuron.
Whole leaves were
harvested after 24
h.
Microarray containing
genes encoding for 109
glycosyltransferases, 63
glutathione enzymes, 62
CYP and 26 drug resistance
transporters.
Up-regulated genes by both
herbicides: 3 CYPs, 4 GSTs,
6 UGTs, 1 GR, MRP3 and
PDR8. a
Glombitza
et al.
2004
Wheat
Triticum
aestivum
Spraying
tribenuron-methyl
at 25 g ai/ha.
One week of
exposure.
Microchips containing 600
genes.
Genes differentially
expressed by at least a
factor 1.9 were 15 down-
regulated and were
encoding proteins involved
in stress response and
defense, carbohydrate
metabolism and transport,
and DNA methylation.
Pasquer et
al. 2006
Arabidopsis
thaliana
weed
Seedlings were
exposed to 100
µg/L imazapyr for
48 h and shoots
harvested.
Microarray.
1001 genes were up-
regulated and 597 genes
were down-regulated.
Early-response genes
included those encoding
members of the UGT, GST,
CYP, ABC, MRP and MATE
protein families, and
mitochondrial alternative
oxidase (ATOX1a) and
NADH dehydrogenase
NDB2. Later stages of the
imazapyr response involved
regulation of genes
participating in biosynthesis
of amino acids, secondary
metabolites and tRNA.
Manabe et
al. 2007
Rice
Oryza
sativa L.
6 days of exposure
to 50 µg/L R(-)-
imazethapyr
qPCR.
In roots: OSPC b gene: 1.5-
fold repressed
Qian et al.
2009
!
2!
japonica
cv. Xiushui
63
In seeds:
OSPC gene: 7.2-fold
repressed
ALS gene: 1.5-fold
repressed
α-amy c: 1.4-fold repressed
β-amy d: 3.9-fold repressed
In shoots:
ALS gene: 7-fold repressed
β-amy: 5.6-fold repressed
Arabidopsis
thaliana
weed
Plants have been
sprayed with four
ALS-inhibiting
herbicides at their
EC50
concentrations:
0.13 g/ha for
sulfometuron-
methyl, 0.16 g/ha
for cloransulam-
methyl, 0.40 g/ha
for imazapyr, and
0.59 g/ha for
primisulfuron-
methyl.
Comparison with
Roundup
(glyphosate at 10.6
g ai/ha). Leaves
were harvested 24 h
later.
Microarray.
16 genes upregulated by
Roundup: genes allowing
detoxification genes
encoding multidrug
transporters, and
metabolizing enzymes
(CYP71A13, GST and UGT).
478 genes affected by
the four herbicides: 82
were down-regulated and
396 were up-regulated.
Up-regulated genes were
encoding ribosome-
associated proteins, CYP
enzymes linked to the
biosynthesis of the
defensive alkaloid
camalexin, and genes
involved in protein
synthesis initiation, and
amino acid synthesis.
Highly up-regulated genes
were 5 CYPs, 3 GSTs, and 3
UGTs.
Genes encoding
mitochondrial alternative
oxidases ATAOX1a and
ATAOX1b, along with two
mitochondrial NADH
dehydrogenases, NDB2 and
NDB4 were up-regulated.
Cell wall biosynthesis genes
were down regulated.
Das et al.
2010
Rice
Oryza
sativa ssp.
japonica
cv.
Seedlings exposed
in water to 0.2 g/L
of chlorsulfuron
for 6 h
qPCR.
The glutathione S-
transferase OsGSTL2 gene
was up-regulated 3-times
Hu et al.
2011
!
3!
Zhonghua
11
Arabidopsis
thaliana
weed
Exposure to 2.5
µg/L R(-)
imazethapyr
enantiomer for 4
weeks
Genes selected were
encoding 3 FSD, 3
CSD, one
manganese
superoxide
dismutase, one
catalase, 6 APX and
8 GPX.!
qPCR.
The following gene
expressions were observed:
Down-regulation:
CSD1: 2 x
CSD2: 3 x
APX1: 2 x
APX5: 1.5 x
GPX1: 1.7 x
GPX2: 1.9 x
GPX5: 1.5 x
GPX7: 2.4 x
Up-regulation:
FSD1: 3 x
FSD2: 3 x
Qian et al.
2011a
Rice
Oryza
sativa L.
japonica
cv. Xiushui
63
0.05 mg/L of R(-)-
imazethapyr for 7
days.
Microarray.
In the leaves 2507 genes
are overexpressed and
2607 repressed, and in the
roots 1882 genes are
overexpressed and 2012
repressed. In particular,
genes involved in
photosynthesis and carbon
fixation are repressed, as
well as genes for glucose
consumption and starch
mobilization, and amino
acids synthesis. Genes
encoding enzymes of the
Krebs cycles were up-
regulated.
Qian et al.
2011b
Rye-grass
Lolium sp.
Pyroxsulam (Abak,
7.5 %; Dow
AgroScience) at
18.75 g/ha for 48h.
RNA sequencing 48 h after
treatment.
15590 transcripts were up-
regulated and 13759 down-
regulated. Major up-
regulated genes were
involved in gene expression
regulation, amino-acid
metabolism and
detoxification (CYP450,
glycosyl- and GSH-S-
transferases, multidrug
resistance and drug
transporters). Major down-
regulated genes were
Duhoux et
al. 2015
!
4!
involved in photosynthesis
and gene expression
regulation.
Rapeseed
Brassica
napus L.
Leaves of each plant
were sprayed with
15 mL of a solution
containing 1.5 µg of
monosulfuron
(approximately 1%
of the concentration
required for its
herbicide action in
wheat field to
control broadleaf
weeds) during the
entire flowering
period (about one
week), making 0.24
g/ha MES.
Microarray.
In the anthers of large
buds, 415 up-regulated
genes involved in
degradation of
carbohydrates, lipids,
proteins, nucleic acids and
cell wall degradation, and
1238 down-regulated genes
involved in the synthesis of
carbohydrates, lipids,
proteins, nucleic acids, and
cell wall synthesis.
Li et al.
2015
Maize
Zea mays
Seedlings treated
with 60 g ai/ha
nicosulfuron for 24
h. RNA extracted
from leaves.
RNA-seq: There were 2100
genes were differentially
expressed: 1391
upregulated and 709 down-
regulated.
Up-regulated genes were
those encoding proteins
involved in translation,
ribosomal structure and
biogenesis, DNA replication,
recombination and repair.
Down regulated genes were
mainly involved in
translation, ribosomal
structure and biogenesis,
posttranslational
modification, protein
turnover, and chaperones.
Liu et al.
2015
Rapeseed
Brassica
napus L.
Plants were foliar-
sprayed with a
single quantity of 1
µg of
amidosulfuron
each (in presence of
of 1 µl of the Bayer
Biopower
alkylethersulfate
surfactant).
RNA sequencing 5 days
after treatment.
142 up-regulated and 201
down-regulated differential
expression transcripts in
young flower buds. Affected
genes showed functional
abnormalities linked to cell
cycle, cell wall formation,
chloroplast structure, and
tissue autophagy. Ethylene-
Liu et al.
2017
!
5!
responsive transcription
factor RAP2-11-like was
up-regulated in the flower
buds paralleling the
increased ethylene release
rate.
Maize
Zea mays
Exposure to 80 mg
nicosulfuron per
kg of soil for 7 days.
qPCR.
Isogenic nicosulfuron-
resistant and sensitive lines
have been compared as for
the expression of genes
encoding antioxidant
enzymes.
In the resistant maize,
nicosulfuron treatment
increased the transcript
levels of most of the
ascorbate peroxidase APX
genes along with those of
SOD9, MDHAR e, DHAR f,
and GR a as compared to
the sensitive line.
Wang et
al. 2018b
Rapeseed
Brassica
napus L.
Plants were foliar-
sprayed with a
single quantity of 15
µg of the
sulfonylurea SX-1
each.
RNA sequencing at various
anther development stages.
998, 2194, 2428 and
10,627 up-regulated genes
and 1177, 2488, 827 and
9745 down-regulated genes
at young buds, short, mid
and late anther stages,
respectively.
Affected genes showed
altered protein processing
in ER, pollen wall
development, and
flavonoids biosynthesis at
early stage. At later stages
hormone signal
transduction, biosynthesis
of amino acids, fatty acids
and steroid were
down-regulated.
Ning et al.
2018
Junglerice
Echinochloa
colona
The treated plants
were sprayed with
imazamox
(Beyond, BASF) at a
rate of 53 g/ha.
RNA were isolated
from leaves 1 h
RNA sequencing.
136 differentially expressed
transcripts: 82 upregulated
and 54 downregulated.
These transcripts are
encoding proteins involved
in metabolism,
Wright et
al. 2018
!
6!
after spraying.
transcription, protein
modification, signaling, cell
wall modification, and
transport.
Rapeseed
Brassica
napus L.
Plants were foliar-
sprayed with a
single quantity of 1
µg of tribenuron-
methyl each
(Express, equivalent
to 0.15 g/ha), in
presence of 1 µl of
the Bayer Biopower
alkylethersulfate
surfactant.
RNA sequencing 5 days
after treatment.
200 up-regulated and 163
down-regulated differential
expression transcripts in
young flower buds. The top
50 up-regulated genes
were involved in stress
response and
detoxification, including 4
MRP and 1 MDR efflux
transporters, 4
mitochondrial heat shock
proteins, 1 GST, 5
sulfotransferases, and 5
UGTs. The top 50 down-
regulated genes were
involved in carbohydrates,
lipid and flavonoids
synthesis.
Lian et al.
2019
a GR: glutathione reductase gene; MRP3: multidrug resistance associated protein 3
gene; PDR8: pleiotropic drug resistance proteins.!
b The OSPC gene is encoding a GRAS transcription factor involved in plant
development.
c α-amy: α-amylase gene.
d β-amy: β-amylase gene.
e MDHAR: monodehydroascorbate reductase gene.
f DHAR: dehydroascorbate reductase gene.!
!
7!
Table 2. Oxidative damage caused by ALS-inhibiting herbicides on plants.
Species
Treatment
Reference
Rice
Oryza
sativa L.
japonica
cv. Xiushui
63
6 days of exposure
to 50 µg/L R(-)-
imazethapyr.
In rice seedlings:
glutathione peroxidase
enzyme activity: 3.4-fold
increase.
MDA a content: 2.2-fold
increase.
Qian et al.
2009
Arabidopsis
thaliana
weed
Exposure to 2.5
µg/L R(-)
imazethapyr
enantiomer for 4
weeks.
MDA content:
x 6 with R(-)-IM
x 3 with racemate-IM
superoxide radical O2.-
concentration:
+ 50 % with R(-)-IM
CAT b activity: reduced 3.6-
fold with R(-)-IM
SOD activity: reduced 2.4-
fold with R(-)-IM
Qian et al.
2011a
Maize
Zea mays
Exposure to 80 mg
nicosulfuron per
kg of soil for 7 days.
O2.- production: x 2.2
H2O2 production: x 1.6
MDA concentration: x 3.2
Dehydroascorbate content:
x 0.48
Ascorbate peroxidase APX
activity: - 33 %
Glutathione peroxidase GPX
activity: x 0.45
Wang et
al. 2018a
a MDA: malondialdehyde, a biomarker of lipid peroxidation.!
b CAT: catalase enzyme.!
!
8!
Table 3. Toxicity of ALS inhibitory herbicides in fish.
Goldfish Carassus auratus exposed for 48 h to nicosulfuron (Bretaud et
al. 2000).
Treatment
Activity of the enzyme acetylcholine esterase (AchE)
50 µg/L
In brain, decrease between 9.2 et 9.8 %. No effect in
muscles.
100 µg/L
In brain, decrease between 11.4 et 12.8 %. No effect in
muscles.
Goldfish Carassus auratus exposed for 15 minutes to nicosulfuron
(Saglio et al. 2001).
Swimming behavior
LOAELa
Increase of the number of feeding attempts
10 µg/L
Increase of the number of social groupings
10 mg/L
aLOAEL : lowest observed adverse effect level
Tilapia Tilapia rendalli given a dose of imazapyr injected abdominally.
Erythrocytes have been sampled from gills 4 days later (Grisolia 2000).
Result: at a dose of 80 mg/kg, the frequency of appearance of micronuclei in
erythrocytes increased 10-fold, a sign of genotoxicity.
Goldfish Carassus auratus exposed to nicosulfuron (Saglio et al. 2003).
Duration of treatment
(h)
Swimming behavior
2
Number of swimming bursts: x 7 at 50 µg/L ; x 9,5
at 100 µg/L
4
Number of social groupings: x 2 at 100 µg/L
6
Number of social groupings: x 2 at 25 µg/L
Silver catfish Rhamdia quelen exposed for 96 h to 400 mg/L of Ally
herbicide (metsulfuron-methyl without adjuvant) (dos Santos et al.
2005).
Activity of AchE
Brain
95 % increase
Muscle
56 % decrease
Fish Leporinus obtusidens exposed for 90 days in rice field at 5.8 µg/L
of Ally herbicide (metsulfuron-methyl) (Pretto et al. 2011).
Biochemical parameter
Effect
AchE activity
50 % decrease in the brain and 60 %
in muscles
Catalase activity (CAT)
50 % decrease in the liver
Lipid peroxidation
Four-fold decrease in muscles, 2.5-
fold in the brain and 3-fold in the liver
Muscle glycogen
36 % decrease
Lactate
15 % decrease in muscles; 2.5-fold
increase in plasma, increase of 19 %
in liver
!
9!
Plasma glucose
15 % increase
Muscle proteins
7 % decrease
Carp Cyprinus carpio exposed for 90 d to BASF herbicide Only at 14.8
µg/L of imazethapyr and 3.1 µg/L of imazapic (Moraes et al. 2011).
Biochemical parameter
Effect
Activity of muscular AchE
40 % decrease
Activity of hepatic GST
45 % decrease
Protein carbonylation
91 % increase in liver
Lipid peroxidation
4- and 2.1-fold increases in brain
and muscles, respectively
Plasmatic glucose
80 % increase
Lactate
30 % increase in muscles
Proteins
29 % and 19 % decreases in liver
and plasma, respectively
Ammonium ion
2.9-fold decrease in liver
Amino acids
3.1-fold decrease in muscles
Climbing perch Anabas testudineus exposed in rice field for 30 d to
Almix 20WP herbicide (DuPont) at a surface concentration of 2 mg/m2.
Almix is composed of 10.1 % metsulfuron-methyl, 10.1% chlorimuron-
ethyl and 79.8 % adjuvants (Samanta et al. 2015).
Tissues
Histopathological and ultrastructural effects
Gills
Hypertrophy and oedema. Degenerative changes in
mitochondria
Liver
Lipid deposits. Dilatation and swelling of mitochondria
Kidneys
Degenerative changes in mitochondria
Zebra fish Danio rerio exposed for 4 d to a concentration of 3.5 mg/L of
halosulfuron-methyl herbicide (LD50/10) (Zhang and Zhao 2017). By
gas chromatography coupled with mass spectrometry, the
concentrations of 51 metabolites have been measured.
Results: in the head, serum and liver the concentrations of several metabolites
were increased:
- the amino acids leucine, valine, serine, glycine, proline, and alanine
- the Krebs’ cycle metabolites malate and fumarate
- the neurotransmitters glutamate, taurine, and glycine
Delta smelt Hypomesus transpacificus (critically endangered species,
IUCN) exposed for 6 h to penoxsulam or imazamox (Jin et al. 2018).
In the brain, 20 to 33 % decrease in the activity of AchE enzyme:
- as soon as 0.01 and 0.53 µM penoxsulam in females and males,
respectively.
- as soon as 0.036 µM imazamox for both females and males.
Nile tilapia exposed to bispyribac-sodium, bensulfuron-methyl and
halosulfuron-methyl at 0.8, 2.5, and 1.275 mg/L, respectively for 96 h
(Fathy et al. 2019).
Bispyribac-sodium: counts of red blood cells (RBC) and platelets decreased; the
concentration of aspartate aminotransferase (AST) and cholesterol increased.
Halosulfuron-methyl: the concentration of AST and cholesterol increased.
!
10!
Bensulfuron-methyl: the count of RBC decreased; the concentration of AST and
cholesterol increased.
Table 5. Decisive influence of the adjuvant Assist in the toxicity of the herbicide
metsulfuron-methyl.
Animal avoidance strategy for metsulfuron-methyl herbicide.a
(de Santo et al. 2018).
Ally
Ally + Assist
Collembola Folsomia candida
> 300
0.07
Earthworms Eisenia andrei
150
0.59
Impact of metsulfuron-methyl herbicide on animal reproduction.a
(de Santo et al. 2019).
Ally
Ally + Assist
Collembola Proisotoma minuta
> 10
0.003 (0-0.010)
Oligochaeta worms Enchytraeus crypticus
> 33.3
5.7 (1.7-13.2)
a After 48 hours of exposure to varying doses of the herbicide, LOAELs were
recorded (mg metsulfuron-methyl/kg soil). The Assist adjuvant was added to
dilutions of the herbicide in water at a concentration of 5 mL/L of water.
Table 4. Impact of the SUL imazethapyr herbicide on the growth of several
organisms exposed for 4 days to two commercial formulations.
Species
Formulation
IC50 (mg/L) a
Reference
Green algae Pseudokirchneriella
subcapitata
V b
1.05
Magdaleno et al.
2015
P. subcapitata
2ASU c
22.4
Health Canada
2010
Cyanobacterium Anabaena flos-
aquae
2ASU
4.8
Health Canada
2010
Diatom Navicula pelliculosa
2ASU
22.9
Health Canada
2010
a (IC50: concentration inhibiting growth at 50 %); b V: Verosil Agrofina,
containing 10.6 % of the active ingredient imazethapyr; c 2ASU: 22.2 % of the
active ingredient imazethapyr.
!
11!
Table 6. The Arsenal herbicide is significantly more toxic than its active
ingredient imazapyr.
Species
Effect
Imazapyr
NPE a
Arsenal
Freshwater snail
Biomphalaria
tenagophila
LC50 at 3 days
(mg/L)
46
12.6
20.1
Vinegar fly
Drosophila
melanogaster
LD50 at 5 days
(oral) (mg/L)
> 2000
0.043
0.174
Mouse
LD50 at 7 days
(i.p.) (mg/kg)
1498
76
262
Onion Allium cepa
Mitosis after a
48 h exposure
(LOAEL in µl/L)
2000
10
0.25
a Nonylphenol ethoxylate, one of the Arsenal adjuvants (Grisolia et al. 2004).
!
12!
Table 7. Compared growth of ALS-mutated plants and their sensitive counterparts:
a few examples.
Plant
Herbicide
Growth phenotype
Reference
Resistant
Sensitive
Arabidopsis
thaliana
Imazapyr
Upper limit in concentration
allowing growth
Manabe et
al. 2007
Mutant csr1-
2D: 1000
µg/L
Wild-type:
25 µg/L
Junglerice
Schoenoplectus
juncoides
Resistance factors a
Sada et al.
2013
Imazosulfuron
Bensulfuron-
methyl
Metsulfuron-
methyl
Bispyribac-
sodium
Imazaquin
176
40
14
5.2
1.5
1
1
1
1
1
Maize Zea
mays
Nicosulfuron
at 60 g ai/ha
Normal
growth
No growth
allowed
Liu et al.
2015
Wheat
Imazamox
Resistance factors a
Dominguez-
Mendez et
al. 2017
Triticum
aestivum
Triticum durum
Clearfield
Rafalín: 93.7
Clearfield
Antoñín:
43.7
Gazul: 1
Simeto: 1
Maize Zea
mays
Nicosulfuron
at 80 mg ai/kg
of soil
SN509-R:
normal
growth
SN509-S:
no growth
allowed
Wang et al.
2018a
Weed
Euphorbia
heterophylla
Imazamox
GR50: 1250
g/ha
GR50: 7.4
g/ha
Rojano-
Delgado et
al. 2019
a Resistance factors between resistant and sensitive plants are calculated as the
ratio of the concentration inhibiting 50 % of the growth (GR50) in the resistant
variety over that in the sensitive one.
Table 8. Application rates, half-lives and environmental persistence of the most used ALS-inhibiting herbicides.
Herbicide
Manufacturer
Brand
Recommended dose
Reported use
Half-life
Environmental persistence
Imidazolinones
Imazamox
BASF
BASF
BASF
Pulsar 40 (4 %)
Beyond (12.1 %)
Cleranda: 17.5 g/L
imazamox and 375 g/l
metazachlor
Cleravis:
17.5 g/l imazamox, 375
g/l metazachlor, and
100 g/L quinmerac
50 g/ha in combination
with the Dash adjuvant
(1.25 L/ha).
Intended for soybean
and Clearfield
sunflower.
35-53 g/ha after Prowl
treatment
(pendimethalin
herbicide).
Intended for
Clearfield canola,
lentil, wheat,
sunflower, and
Clearfield Plus
wheat and
sunflower.
2 L/ha + Dash HC 1
l/ha
Intended for
Clearfield rapeseed
2 L/ha + Dash HC 1
l/ha
Intended for
Clearfield rapeseed
50 and 100 g/ha on
soybean (da Silva
et al. 1999)
45 g/ha at fall and
90 g/ha at spring
on wheat (Ball et
al. 2003)
70 and 140 g/ha
(Rani et al. 2019)
In soil: 9.8 d (Aichele
and Penner 2005); 25
d (Müller and
Applebyki 2012); at
140 g/ha (t0): 29.5
and 34.3 d in sandy
and clay loam soils,
respectively (Rani et
al. 2019)
Terrestrial field
dissipation: 15 to 130
d (US EPA 2008); 1.8
to > 120 d (Buerge et
al. 2019)
In water-sediment:
half-life of 142 d
(Müller and Applebyki
2012)
< LOQa (5 µg/kg) in all of 22
Brazilian soils (Kemmerich et al.
2015).
Imazapic
BASF
Plateau DG
28.5-85 g/ha
Intended for prairie
grass
147 g/ha on
sugarcane (Simoni
et al. 2006)
With Clearfield
maize: 17.5%
imazapyr + 52.5%
imazapic mixture,
at two doses of 80
and 160 g/ha, with
a 31% hydrocarbon
petroleum (Dash)
at 0.125 and 0.250
L/ha, respectively
(Alister and Kogan
2005)
32-66 d at 52.5 g/ha
(t0)(Ulbrich et al.
2005); 120 d (Müller
and Applebyki 2012);
40 to 64 d (da Costa
Marinho et al. 2019)
In rice paddy water:
8.1 to 11.9 d at 25
g/ha (t0) (Reimche et
al. 2005)
In 2 out of 22 Brazilian soils, 5.8
and 12.1 µg/kg, respectively
(Kemmerich et al. 2015)
Imazapyr
BASF
Arsenal Power (240
g/L); also marketed as
Assault, Chopper,
Contain, and Pivot.
504-2016 g/ha
Intended for sugarcane
With Clearfield
corn: 52.5 + 17.5,
and 105 + 35 g/ha
for imazapic and
imazapyr,
respectively
(Ulbrich et al.
2005)
In soil:
43-53 d at 17.5 g/ha
(t0) (Ulbrich et al.
2005); 26-44 d (Wang
et al. 2006); 11 d
(Müller and Applebyki
2012); 37 to 121 d
(Gianelli et al. 2014)
In water:
30 d (Müller and
Applebyki 2012)
In 1 out of 22 Brazilian soils, 5.3
µg/kg (Kemmerich et al. 2015).
In soil: 5.3 µg/kg; in water: 50
ng/L (Börjesson et al. 2004)
Imazaquin
BASF
BASF
Image 70 DG
Scepter 70DG
440-582 g/ha
Intended for turf grass
71.5-143 g/ha
Intended for soybean
140 g/ha (Barnes
et al. 1989); 140
and 280 g/ha on
maize (Loux et al.
1993); 150 and
300 g/ha on maize
(Gazziero et al.
1997)
In fields:
39 to 150 d (Curran et
al. 1992); 43 d (Mills
et al. 1989)
In soil:
1337 d (Aichele and
Penner 2005); 60 d
(Müller and Applebyki
2012)
In water-sediment:
516 d (Müller and
Applebyki 2012)
In fields:
4 µg/kg both in 1985 and 1986
(Mills and Witt 1989)
In ground water:
in 2 out of 25 samples (LOQ: <
10 ng/L), max. conc.: 24 ng/L
(Battaglin et al. 2000)
Imazethapyr
BASF
BASF
BASF
Only (75 g/L
imazethapyr, 25 g/L
imazapic)
Newpath
Pursuit
100-200 g/ha
Intended for
Clearfield rice
67-100 g/ha
Intended for
Clearfield rice
50-100 g/ha
Intended for alfalfa,
clover, beans and
peas, peanut, and
soybean
75 g/ha
imazethapyr + 25
g/ha imazapic
(Marchesan et al.
2010)
1.8 L Only/ha plus
surfactant Dash at
0.5% v/v to select
tolerant plants,
then 1 L/ha plus
Dash for Clearfield
rice (Rangel et al.
2010)
Imazethapyr was
applied up to 105
g/ha on Clearfield
rice (Norsworthy et
al. 2008)
With Clearfield
maize: 17.5%
In soils:
silty-clay: 192 to 318
d; silty-loam: 78 to
270 d (Goetz et al.
1990); 73 to 257 d
(Vischetti 1995); 112 d
(Aichele and Penner
2005); 63 to 112 d
(Zhang et al. 2010);
90 d (Müller and
Applebyki 2012); 36 to
98 d (da Costa Marinho
et al. 2019); at 140
g/ha (t0): 39.8 and
43.3 d in sandy and
clay loam soils,
respectively (Rani et
al. 2019)
In fields:
49 to 122 d (Curran et
al. 1992); 60 d (Mills
and Witt 1989); 30 to
In 2 out of 22 Brazilian soils, 24
and 37.7 µg/kg, respectively
(Kemmerich et al. 2015)
In fields:
13 and 5 µg/kg in 1985 and
1986, respectively (Mills and Witt
1989)
In paddy field: mean of 4.5 µg/kg
up to 15 cm of depth, 540 days
after the last treatment (Kraemer
et al. 2009b)
In ground water:
in 4 out of 25 samples (LOQ: <
10 ng/L), max. conc.: 59 ng/L
(Battaglin et al. 2000)
imazapyr + 52.5%
imazethapyr
mixture, at two
doses of 80 and
160 g/ha, with a
31% hydrocarbon
petroleum (Dash)
at 0.125 and 0.250
L/ha, respectively
(Alister and Kogan
2005)
43 d (Xu et al. 2013);
28 to 42 d (Jovanović-
Rodovanov 2017)
In rice paddy water:
1.6 to 5.2 d at 75 g/ha
(t0) (Santos et al.
2008); 4.5 to 5.3 d at
75 g/ha (t0) (Reimche
et al. 2005)
Sulfonylureas
Ethametsulfuron-
methyl
DuPont
DuPont
Muster Toss-N-Go (75
%)
Salsa (75 %)
20 to 30 g/ha
+ surfactant Agral 90,
AG-Surf ou Citowett+
25 g/ha + surfactant
Trend 90
Intended for rapeseed
Not documented
In soil:
13 to 67 d
(Si et al. 2005); 70 d
(Müller and Applebyki
2012)
In ponds :
30 d (Cessna et al.
2006)
Drinking water in Canada (2003-
2005), detected in 35 % of
samples, max. conc.: 80.4 ng/L;
mean of 8.5 ng/L in July 2005
(Donald et al. 2007)
Wetland sediment: max. conc.:
10 µg/kg; > LOQ in 39 % of
samples (Degenhardt et al. 2010)
In Canadian streams (2009-
2011):
17 % of samples above the limit
of detection; max. conc.: 148
ng/L (de Lafontaine et al. 2014)
Metsulfuron
methyl
DuPont
Ally (600g/kg)
5 to 7 g/ha
Intended for barley,
rye, triticale, wheat,
linseed, safflower!
6 g/ha on winter
wheat (Rouchaud
et al. 1999)
10, 20 and 40 g/ha
on wheat (Kotoula-
Syka et al. 1993)
In soils:
from 33 d at pH 5.8 to
120 d at pH 7.4 for 32
g/ha (t0) (Walker and
Welch 1989); 8 to 36 d
at pH 5.7 (James et al.
1995); from 12 to 28 d
in Colorado soils
(Cranmer et al. 1999);
67 to 78 d at 6 g/ha
(t0) (Rouchaud et al.
1999); from 51 d at pH
5.9 to 82 d at pH 6.8
for 4.6 ng/g soil (t0)
(Bedmar et al. 2006);
89.3, 84, and 67.6 d at
5, 13, and 20 °C,
respectively (Cessna et
al. 2017)
Drinking water in Canada (2003-
2005), detected in 2 % of
samples, max. con.: 2.1 ng/L
(Donald et al. 2007)
Wetland sediment: max. conc.:
1.4 µg/kg; > LOQ in 11 % of
samples (Degenhardt et al. 2010)
In Ontario streams (2006-2008):
2 % of samples above the limit of
detection; max. conc.: 57 ng/L
(Struger et al. 2011)
In Canadian streams (2009-
2011):
18.5 % of samples above the
limit of detection; max. conc.: 48
ng/L (de Lafontaine et al. 2014)
In ponds :
84 d (Cessna et al.
2006)
In water-sediment:
140 d (Müller and
Applebyki 2012)
Nicosulfuron
DuPont
DuPont
DuPont
Syngenta
Accent Q (54 %)
Zest (75 %)
Pastora (56 %
nicosulfuron + 15 %
metsulfuron-methyl)
Milagro (4 %)
36 to 72 g/ha
Intended for corn
37 to 73 g/ha
Intended for
herbicide-tolerant
INZEN sorghum
52 to 130 g/ha
Intended for wheat,
barley, oat, soybean,
corn, alfalfa, clover,
ryegrass
60 g/ha
60 g/ha on maize
(Liu et al. 2015)
In field:
13.6 d at 202 g/ha (t0)
(Wu et al. 2010); 26 d
(Müller and Applebyki
2012)
In corn plants: 0.53 to
0.73 d at 202 g/ha (t0)
(Wu et al. 2010)
In water-sediment:
41.5 d (Müller and
Applebyki 2012)
In ponds :
75 d (Cessna et al.
2015)
US streams and rivers: maximal
concentration in 2005-2006:
3290 ng/L (Battaglin et al. 2009)
In ground water:
in 2 out of 25 samples (LOQ: <
10 ng/L), max. conc.: 16 ng/L
(Battaglin et al. 2000)
In Ontario streams (2006-2008):
11 % of samples above the limit
of detection; max. conc.: 525
ng/L (Struger et al. 2011)
In Canadian streams (2009-
2011):
10 % of samples above the limit
of detection; max. conc.: 84 ng/L
(de Lafontaine et al. 2014)
Sulfosulfuron
Monsanto
Monsanto
Apyros (75% WG)
Monitor 75% WG
20-35 g/ha + adjuvant
Genamin.
Intended for wheat
and potato
40 g/ha+ surfactant
Villa 51
Intended for wheat
and tomato
20 and 40 g/ha on
winter wheat + the
adjuvant Genamin
(Alonso-Prados et
al. 2002)
37.5, 75 and 112.5
g/ha on tomato +
surfactant DX
(Eizenberg et al.
2003)
26.5 g/ha on winter
wheat
(Adamczewski and
Paradowski 2004)
50 and 100 g/ha on
wheat crop
(Ramesh et al.
2007a)
20 g/ha on wheat
In field:
105 to 125 d (from 30
to 60 cm of depth) at
50 g/ha (t0) (Ramesh
et al. 2007a); 14.4 d
at 25 g/ha (t0)
(Sondhia 2008b); 3.4
to 3.6 d at 34 g/ha (t0)
(Ghosh et al. 2014);
4.6 to 11.6 d at 20
g/ha (t0) (Yousefi et al.
2016); 5.4 to 10.8 d at
26 g/ha (t0)
(Mehdizadeh et al.
2017)
In soil:
24 d (Müller and
Applebyki 2012)
In water: 67 to 76 d
(Ramesh et al. 2007b);
168 d (Müller and
Drinking water in Canada (2003-
2005), detected in 10 % of
samples, max. conc.: 36.1 ng/L
(Donald et al. 2007)
Wetland sediment: max. conc.:
7.1 µg/kg; > LOQ in 39 % of
samples (Degenhardt et al. 2010)
In Canadian streams (2009-
2011):
31 % of samples above the limit
of detection; max. conc.: 28 ng/L
(de Lafontaine et al. 2014)
Sulfosulfuron
Monsanto
Monsanto
Apyros (75% WG)
Monitor 75% WG
20-35 g/ha + adjuvant
Genamin.
Intended for wheat
and potato
40 g/ha+ surfactant
Villa 51
Intended for wheat
and tomato
20 and 40 g/ha on
winter wheat + the
adjuvant Genamin
(Alonso-Prados et
al. 2002)
37.5, 75 and 112.5
g/ha on tomato +
surfactant DX
(Eizenberg et al.
2003)
26.5 g/ha on winter
wheat
(Adamczewski and
Paradowski 2004)
50 and 100 g/ha on
wheat crop
(Ramesh et al.
2007a)
20 g/ha on wheat
(Yousefi et al.
2016)
In field:
105 to 125 d (from 30
to 60 cm of depth) at
50 g/ha (t0) (Ramesh
et al. 2007a); 14.4 d
at 25 g/ha (t0)
(Sondhia 2008b); 3.4
to 3.6 d at 34 g/ha (t0)
(Ghosh et al. 2014);
4.6 to 11.6 d at 20
g/ha (t0) (Yousefi et al.
2016); 5.4 to 10.8 d at
26 g/ha (t0)
(Mehdizadeh et al.
2017)
In soil:
24 d (Müller and
Applebyki 2012)
In water: 67 to 76 d
(Ramesh et al. 2007b);
168 d (Müller and
Applebyki 2012)
In water-sediment:
26 d (Müller and
Applebyki 2012)
In ponds :
44 d (Cessna et al.
2015)
In wheat plant: 2.5 to
3.0 d at 34 g/ha (t0)
(Ghosh et al. 2014)
Drinking water in Canada (2003-
2005), detected in 10 % of
samples, max. conc.: 36.1 ng/L
(Donald et al. 2007)
Wetland sediment: max. conc.:
7.1 µg/kg; > LOQ in 39 % of
samples (Degenhardt et al. 2010)
In Canadian streams (2009-
2011):
31 % of samples above the limit
of detection; max. conc.: 28 ng/L
(de Lafontaine et al. 2014)
Tribenuron-
methyl
DuPont
DuPont
Express
(50 %)
Express SX
(50 %)
9.1 to 18.25 g/ha
Intended for wheat,
barley, triticale, fallow
45 to 60 g/ha +
surfactant Trend 90
Intended for sunflower
25 g/ha on spring
wheat (Pasquer et
al. 2006)
10, 20 and 40 g/ha
on wheat (Kotoula-
Syka et al. 1993)
In field: 3.2 to 5.7 d at
15 g/ha (t0)
(Mehdizadeh et al.
2017)
In soil:
14 d (Müller and
Applebyki 2012); 56.4,
26.3, and 10.3 d at 5,
13, and 20 °C,
respectively (Cessna et
al. 2017)
In water:
16 d (Müller and
Applebyki 2012)
In water-sediment:
26 d (Müller and
Applebyki 2012)
Drinking water in Canada (2003-
2005), detected in 20 % of
samples, max. conc.: 30.1 ng/L;
mean of 4.9 ng/L in July 2005
(Donald et al. 2007)
Pyrimidinyl
carboxy
compound
Bispyribac-
sodium
Arysta
LifeScience
Nominee Gold 10SC
37.5 g/ha
Intended for rice
In paddy field: 25
g/ha (Jabran et al.
2012); From 37.5
to 75 g/ha in paddy
field (Zhang et al.
2013)
In rice paddy water:
8.9 to 15.5 d at 50
g/ha (t0) (Reimche et
al. 2005)
In water-sediment:
35 d (Müller and
Applebyki 2012)
In soil: 45-131 d and
30-51 d under
anaerobic and aerobic
conditions (pez-
Piñeiro et al. 2011); 13
d (Müller and
Applebyki 2012); 13.1,
10.2 and 9.9 d at 40,
25 and 20 g/ha (t0),
respectively
(Ramprakash et al.
2015)
In paddy soil: 1.4 to
5.6 d at 75 g/ha (t0)
(Zhang et al. 2013);
17.7 and 23.2 d at 70
and 140 g/ha (t0),
respectively (Saha et
al. 2016)
In loamy sand: 27.3 d
at 50 g/ha (t0), and
41.0 d under
submerged condition
(Kalsi and Kaur 2019)
In paddy soil: 0.18 to 0.2 µg/kg
at harvest for a treatment of 37.5
g/ha (t0) (Zhang et al. 2013)
a: LOQ: limit of quantification.
... 1996-2010 yılları arasında herbisit tolerant bitkilerin kullanımı, sürdürebilir tarıma katkı sağlayan, daha uygun ve esnek ürün yönetimi, daha yüksek üretkenlik veya net getiri/hektar ve çevresel anlamda daha güvenli, herbisitlerin kullanımının azaltılması gibi önemli faydalar sunan bir yöntem olarak küresel anlamda benimsenmiştir (Naidu ve Ranganath, 2011). Ancak, son yıllarda yapılan bir çok araştırma raporuna göre herbisit tolerantlı kültür bitkilerinin kullanımıyla herbisit kullanımında artış (Peerzada ve ark.,2019) ve buna bağlı olarak çevreye olumsuz etkiler, biyoçeşitlilik üzerinde değişim, akraba türler arasında gen kaçışı nedeniyle dirençli yabancı otların ve kendiliğinden gelen bitkilerin gelişimi gibi sorunlar bildirilmiştir (Chen ve ark., 2004;Sudianto ve ark., 2013;Bourdineaud, 2020;Prakash ve ark., 2020). ...
... İmi-tolerant kültür bitkilerinin kullanımı ile ortaya çıkabilecek olası negatif etkiler aşağıda listelenmiştir (Prakash ve ark., 2020)  Artan herbisit kullanıma bağlı çevre ve insan sağlığına olumsuz etkileri (Goldburg, 1992;Bourdineaud, 2020).  Ekotoksisite (toprak mikroorganizmaları, tarımsal fauna ve flora üzerinde yan etkiler)  Herbisite dayanıklı yabancı ot ve kendiliğinden gelen kültür bitkilerinin gelişmesi (Senior ve Dale, 2002;Sudianto ve ark., 2013). ...
Article
Full-text available
Bitki ekosistemi içerisinde bitkilerin çimlenme, gelişme ve dağılımında etkili olan faktörler mevcuttur. Bu faktörlerin bitkinin istediği optimum koşullarda gerçekleşmesi durumunda, sağlıklı bir büyüme gerçekleşir. Bu faktörlerin en önemlilerinden birisi de sıcaklıktır. Bu kaynak ihtiyacını karşılamada yaşanacak aksaklık, bitkilerde olumsuz etkilere neden olmaktadır. Küresel iklim değişikliğine bağlı olarak artan sıcaklık faktörünün bitkiler üzerindeki etkilerinin bilinmesi ve bitkinin buna karşı oluşturduğu cevap mekanizmalarının iyi anlaşılması oldukça önemlidir. Bununla birlikte artan sıcaklık stresinin bitkilerde morfolojik, fizyolojik ve verimsel açıdan meydana getirdiği etkileri anlamak ve bitkilerin buna karşı verdikleri tepkileri öğrenmek, sıcaklığa toleranslı genotiplerin geliştirilmesine yönelik adımları atmada önem arz etmektedir. Bu derlemede, sıcaklığın bitkiler üzerindeki etkileri ve yüksek sıcaklık stresine karşı verdiği cevaplar tartışılmıştır.
... 1996-2010 yılları arasında herbisit tolerant bitkilerin kullanımı, sürdürebilir tarıma katkı sağlayan, daha uygun ve esnek ürün yönetimi, daha yüksek üretkenlik veya net getiri/hektar ve çevresel anlamda daha güvenli, herbisitlerin kullanımının azaltılması gibi önemli faydalar sunan bir yöntem olarak küresel anlamda benimsenmiştir (Naidu ve Ranganath, 2011). Ancak, son yıllarda yapılan bir çok araştırma raporuna göre herbisit tolerantlı kültür bitkilerinin kullanımıyla herbisit kullanımında artış (Peerzada ve ark.,2019) ve buna bağlı olarak çevreye olumsuz etkiler, biyoçeşitlilik üzerinde değişim, akraba türler arasında gen kaçışı nedeniyle dirençli yabancı otların ve kendiliğinden gelen bitkilerin gelişimi gibi sorunlar bildirilmiştir (Chen ve ark., 2004;Sudianto ve ark., 2013;Bourdineaud, 2020;Prakash ve ark., 2020). ...
... İmi-tolerant kültür bitkilerinin kullanımı ile ortaya çıkabilecek olası negatif etkiler aşağıda listelenmiştir (Prakash ve ark., 2020)  Artan herbisit kullanıma bağlı çevre ve insan sağlığına olumsuz etkileri (Goldburg, 1992;Bourdineaud, 2020).  Ekotoksisite (toprak mikroorganizmaları, tarımsal fauna ve flora üzerinde yan etkiler)  Herbisite dayanıklı yabancı ot ve kendiliğinden gelen kültür bitkilerinin gelişmesi (Senior ve Dale, 2002;Sudianto ve ark., 2013). ...
Article
Full-text available
Weeds cause damage by competing with crops for resources such as water, nutrients, light and space. In its control, chemical control methods are preferred due to its advantages such as short-term results and ease of use. However, the problem of resistance, which arises as a result of intensive use of herbicides, has increased with the development of selective herbicides. In particular, the problem of resistance to ALS (Acetolactate synthase) inhibitor herbicides is in the first place. Imidazolinones, one of the active substances in the sulfonylurea class from the ALS inhibitor herbicide group, are widely used against narrow and broad-leaved weeds. Imidazolinones have a very broad weed spectrum that is effective even at low doses, and are herbicides with a relatively environmental profile, with low toxicity to animals, birds, fish and invertebrates. These features have enabled the development of imidazolinone-tolerant (IMI-tolerant) crops with traditional production (non-transgenic) techniques. IMI-tolerant crops mean that are tolerant to imidazolinone group herbicides. While controlling weeds in the production areas where these crops are used, it is aimed to prevent the product from being damaged. It has advantages such as low cost, better and more effective control of weeds, increasing the yield in agricultural production. However, there may be some disadvantages such as negative environmental effects due to intensive use of herbicides in the areas where IMI-tolerant plants are cultivated, gene escape to weed species related to the cultivated plant and the emergence of resistant weeds accordingly. In this review, the advantages and disadvantages of using IMI-tolerant plants in weed control are discussed.
... Известно, что для эффективного разложения остатков гербицидов в почве необходимо соблюдение некоторых почвенно-климатических условий: достаточная влажность почвы, необходимый уровень рН, а также повышенная температура. В засушливых условиях, а также в условиях низких температур скорость деградации гербицидов почвенной микробиотой может быть значительно снижена, что приводит к накоплению гербицидов и их сохранности в почве до следующего сезона [59], к изменению почвенного биоразнообразия [60], развитию устойчивых сорняков и самопроизвольному появлению новых растений из-за утечки генов среди родственных видов [34,[61][62][63]. Например, при перекрестном опылении может происходить обмен генов (перенос трансгенного признака) между растениями одного или разных видов. ...
Article
Currently, in industrial agricultural production, the best results are demonstrated by technological schemes for weed control based on the use of complexes of non-selective and selective herbicides in crop rotations. Herbicides of the imidazolinone group are well-known representatives of selective herbicides class. Commercial drugs based on them are widely represented on the market. At the same time, the negative consequences of using herbicides of the imidazolinone group are well known. They lead to disruption of natural soil biodiversity, inhibition of growth and decreasing in the yield of subsequent crops in crop rotations and the emergence of resistant weeds. The manuscript discusses the mechanisms of action of herbicides of the imidazolinone group, the current state of methods of their analysis in the soil, as well as approaches to reducing the negative effects of their usage.
... These herbicides inhibit the acetolactate synthase (ALS) enzyme, which is needed for producing certain branched chain amino acids required for plant growth, and they are thought to be nontoxic to nontarget organisms that do not have ALS (e.g. mammals, birds, reptiles, and insects, Tu et al. 2001, but see Bourdineaud 2022). However, because bacteria and fungi also have the ALS enzyme, these herbicides can directly affect the soil microbial community composition and the processes they mediate (El-Ghamry et al. 2001;Maruyama et al. 2016;Thiour-Mauprivez et al. 2019), though these effects appear to be short-lived (Busse et al. 2004;Souza et al. 2013). ...
Article
Community‐type conversions, such as replacement of perennials by exotic annual grasses in semiarid desert communities, are occurring due to plant invasions that often create positive plant–soil feedbacks, which favor invaders and make restoration of native perennials difficult. Exotic annual grass control measures, such as pre‐emergent herbicides, can also alter soil ecosystems directly or indirectly (i.e. via the plant community), yet there are few studies on the topic in natural, non‐cropped landscapes. We asked how spray treatments applied to soil post‐fire with the intention of inhibiting invasive annual grasses (such as Bromus tectorum L.) and releasing existing native perennial grasses affected soil resources, a microbial process, and invertebrates in three climatically varied sagebrush steppe sites. Spray treatments included chemical herbicides (imazapic and rimsulfuron) that strongly affected plant communities and a bioherbicide ( Pseudomonas fluorescens strain D7) that did not. Chemical herbicides increased soil mineral nitrogen in proportion to their negative effects on plant cover for 2 years after treatments in all sites and increased soil water and net N mineralization (measured at one site) but did not affect total carbon, nitrogen, or organic matter. Invertebrate responses to herbicides varied by site, and invertebrates increased with chemical herbicides at the highest, wettest site. We show that herbicide treatments can exacerbate pulses of mineral nutrients, which previous studies have shown can weaken ecosystem resistance to invasion. Thus, restoration strategies that increase the likelihood that desired plants can capture mineralized nutrients after herbicide application will likely be more successful.
... Since the 1990s, in most regions of the world as the United States, Brazil, Argentina, China, and India (Masuda & Goldsmith 2009;Araújo et al. 2019), has occurred the change of perennial crops and grasses toward annual crops, mainly soybean (Glycine max). These crops are genetically modified to be tolerant to herbicides (Webster and Sosnoskie 2010;Leguizamón 2020;Bourdineaud 2022) such as glyphosate and other herbicides. This situation is accompanied by no-till technological packages (Montoya et al. 2019). ...
Article
Full-text available
Surface water and unconfined and semiconfined (SAC B) aquifers of Marcos Juárez plain, in the central Pampa of Argentina, were studied to assess hydrogeological features and the presence and distribution of herbicides in water. The studied area, located near a regional hydrological discharge area shows streams and channels of high salinity fed by the groundwater with low values of herbicides as a consequence of hydrodynamic dispersion. The unconfined and semiconfined aquifers (SAC B) are impacted by the presence of herbicides although the percentage of positive samples is low. The unconfined aquifer, directly connected to the atmospheric phase, has a shallow water table being more likely to be reached by effective water infiltration. Despite their high Koc and low GUS index, glyphosate and AMPA were present in the highest concentrations in the unconfined aquifer. It is assumed that this situation is the result of its widespread use, its high solubility in water, and its high persistence. The chemical characteristics of metolachlor and imazethapyr, such as their relatively low Koc, moderate and high GUS index, and high persistence have allowed also their leaching into groundwater. Herbicide migration from the unconfined aquifer to the semiconfined aquifer layer in the loessical plain can occur due to the differences in their hydraulic potential. This situation facilitates the mixing of water, a process that is interpreted through fractures or cracks in the confining units (aquitards) or due to the presence of broken old pipes. The multivariate statistical analysis of the unconfined groundwater chemistry leads us to the conclusion that approximately 53% of the variance is explained by different levels of salinization resulting from natural geochemical processes (such as cation exchange, dissolution of salts, etc.), while the remaining 47% is attributed to contamination processes (arrival of herbicides and nitrates). The second and third positions of herbicides in the PCA analysis demonstrate the impact of agricultural activity, especially in the unconfined aquifer.
... However, continuous overuse of a single herbicide, in a growing season, will provoke an acquired resistance of weeds to that herbicide [3]. A clear example is Glyphosate, considered the most widely used herbicide in the world [4]. ...
... The introduction of novel technologies of growing of genetically modified (GM) crops, which drastically changed the US agriculture, is dated back to 1996. Since then, the areas planted to GM crops keep on increasing, and so do the concerns of some researchers about safety of such approaches (Zimdahl, 2018;Nishimoto, 2019;Clark, Maselko, 2020;Bourdineaud, 2022). Without going into details of this criticism, the very fact of the development of these technologies is undoubtfully a remarkable milestone in plant protection from weeds (Brookes, 2014;Gosavi et al., 2022;Brunharo et al., 2022). ...
Article
Full-text available
Changes in herbicides recommended for the use in Russian Federation between 2000 and 2022 are analyzed. The main directions of iimproving chemical control of weeds are identified based on the integration of domestic market with the world market. Only a limited number of active ingredients was introduced in Russia during the last decade, including pinoxaden, thiencarbazone-methyl, piroxulam, sodium flucarbazone, topramezone, diclosulam, tembotrione, and metamifop. Improved formulations of herbicides such as colloidal solution concentrate with increased penetrability due to the particle size reduced by an order of magnitude became widely available. Premix herbicides were developed based on tribenuronmethyl, metsulfuron-methyl, florasulam, clopiralid, picloram, imazamox, imazapyr, imazethapyr, etc.Parameters for herbicide application were optimized to consider phenology of weeds. Novel technologies were implemented, such as growing hybrids resistant to certain active ingredients to allow their application during crop vegetation.
... The factor like a group of herbicides, time of application, the active ingredient, dose, and path of intake of herbicides decide their impact on human health. The Hou et al. (2006) Triazine (atrazine, simazine, and propazine) Breast and ovarian cancer in female, Sanderson et al. (2013) Imidazolinone and sulfonylurea Bladder and colon cancer, miscarriages in pregnant lady Bourdineaud (2020) metribuzin in drinking water can damage the kidney of human beings. Some of the experiments documented that the Triazine group of herbicide causes the carcinogenic effect in some animals (Jowa and Howd, 2011) and reproductive failure in humans (Gammon et al., 2005). ...
Chapter
Full-text available
Food and nutritional security, environmental sustainability, mitigating climatic vulnerability, shifting of weed flora, weed developed resistance against the herbicide, high capital investment through manual weed management, and increasing the requirement for energy input in the agriculture sector are the major issues in crop production in the coming years. It is no doubt that the introduction of herbicide in the agriculture sector increases the income of farmers, which boost the economy of the nation, but its improper uses create several problems. The consumption of herbicide in the world during 2018 was 1.30Mt. The excess uses of herbicide in agriculture pose several consequences such as environmental pollution, increasing demand for energy in the industrial sector, increase resistance in different weed species, appearing novel weed flora in the cropping system, and incurred higher cost of cultivation in crop production. Sustainable food production is one of the important tools in maintaining ecological balance and soil health. In this circumstance, integrating legumes into cropping systems provides several ecosystem services which fulfill the objectives of ecological weed management. Sustainable intensification is fulfilling the demand for food and ensuring nutritional security in a sustainable manner while maintaining biodiversity and providing many ecosystem services. In a cropping system or single crop production weeds are poses a serious loss by reducing crop growth, yield, quality, depletes fertility status of soil, and act as an alternate host for several insects, pest, and diseases. The yields reduction in direct-seeded rice due to weeds was reported up to 90%. Globally, more than US$ 100 billion was a loss due to infestation of weed in annual crops. The weed seed of Argemone mexicana crushed mustard seed and the oil feed by human beings causes glaucoma or dropsy. The weed green Amaranthus (Amaranthus viridis) can accumulate about 3% N in its biomass and causes severe depletion of nitrogen (N) economy in soil. The three solanaceous weeds such as Solanum nigrum, Datura stramonium, and Datura ferox are act as an alternate host for tomato leaf minor. The application of herbicides during the crop production causes adverse effects on the environment, soil ecosystem, pollute ground water, damage ecological diversity, and affects human health. Besides, the use of herbicide for weed management incurred about US$ 25 billion annually across the globe. Therefore, to tackle such issues of weed the integration of legumes in the different crop production systems as cover crop, relay crop, green manure crop, brown manuring crop play a key role in providing many ecosystem services such as suppressing weed species by smothering or by allelopathy effect, break the life cycle of disease and pest, increasing carbon (C) and N pool in soil, enhancing soil organic matter content, enhance soil health by improving physical, chemical and biological properties of soil. In intercropping system, legumes have better suppression on weed flora by reducing their density and biomass. Further, legumes fulfill the requirement of N of the component crop. Legumes in the crop rotation system break the infestation of frequently occurrence weeds due to its allelopathic effects or smothering effects on the weed seed bank. Based on the diverse benefits of legumes, it is ensured that legumes either in the cropping system or alone as crop residue plays a key role in driving sustainable intensification.
Thesis
Full-text available
Common ragweed (Ambrosia artemisiifolia L.), a particularly troublesome and allergenic weed, is mainly controlled in agricultural fields using ALS inhibitor herbicides. Recent cases of herbicide resistance have been reported in France and are jeopardising the efficacy of this mode of action. Both target site resistance (TSR, structural mutation in ALS gene) and non target site resistance (NTSR, regulatory and/or structural mutations in secondary metabolism) are involved. The fundamental aim of this work was to identify the genetic determinisms of resistance to ALS inhibitors that have evolved in common ragweed populations in France. As an applied objective, this work also aimed to prepare the development of a high- throughput molecular diagnostic tool that would ensure a rapid detection of resistance. We first assessed the situation of common ragweed resistance in France and identified the mechanisms involved and their modalities of evolution. Using herbicide sensitivity bioassays coupled with ALS gene sequencing, we showed that ragweed resistance to two active substances, imazamox and tribenuron, is emerging in France and is mainly due to NTSR mechanisms. The observed resistance patterns suggest that a diversity of NTSR mechanisms are evolving in France. Furthermore, we demonstrated that TSR evolved locally, through multiple and independent appearance of mutations in the ALS gene. Thanks to the innovative application of high- throughput sequencing for the diagnosis of TSR on a national scale, we identified several foci of TSR emergence, as well as an unsuspected diversity of mutations in the ALS gene. We then studied the genetic determinisms of NTSR. A transcriptomic approach (RNASeq) associated with an analysis of nucleotide polymorphisms was conducted, based on the hypothesis that genes and/or markers of NTSR differed by their expression level and/or by sequence polymorphisms between plants resistant or sensitive to ALS inhibitors. For the first time, this approach was conducted directly on field plant material, i.e. six populations with distinct geographical origins and/or resistance profiles. Constitutive expression differences between resistant and sensitive plants were identified, especially in genes from families known to be involved in herbicide metabolism (cytochromes P450, transferase enzymes, transporters, etc.), but also in genes that may be involved in regulatory cascades activated by the herbicide. Validation of their relative expression levels and their ability to predict NTSR was performed on a massive sampling of plants. Taken together, the results indicate that a very high diversity of mechanisms is involved in RNLC within and between populations, highlighting the highly polygenic nature of RNLC in ragweed. In addition, assessment of the early response of plants to herbicide application showed that genes involved in secondary plant metabolism are specifically induced by treatment in resistant plants from different populations. Finally, sequence variants potentially correlated with NTSR were identified. Their validity as resistance markers remains to be confirmed. The diversity of resistance mechanisms identified within each population renders the development of a molecular diagnosis tool complex. On the other hand, it opens exciting perspectives for the study of the evolutionary dynamics of the adaptation of an invasive species subjected to a particularly intense anthropogenic selection pressure.
Article
Full-text available
A supervised field trial on rice was conducted at the experimental farm, BCKV, West Bengal where the mixture formulation of bispyribac sodium 4% + metamifop 10% was applied at 14 days after transplanting of the rice seedlings at three doses viz. recommended dose i.e. @ 70 g a.i. ha-1 and double dose i.e. @ 140 g a.i. ha-1and another is untreated control The residue was extracted by QuEChERS method and quantified by liquid chromatography electrospray ionization tandem mass spectrometry. The limit of quantification of both the compound was 0.02mg kg-1in all the substrates (rice green plant, straw, grain, husk, and soil).The half-life (T1/2) value of Bbispyribac sodium in soil ranged from 17.71 -23.16 days where as the T1/2 value of metamifop in soil and rice green plant ranged from 12.54 - 15.84 and 1.91-2.71 days respectively. No residue of bispyribac sodium and metamifop was quantified in harvested rice straw, grain, husk and soil. Keywords : Bispyribac sodium, half-life, metamifop, persistence, rice, soil
Article
Full-text available
Background: Acetolactate synthase (ALS)-inhibiting herbicides from the chemical families of sulfonylureas and imidazolinones are used worldwide. However, drift or sprayer contamination from some sulfonylurea herbicides causes a high level of male sterility in cruciferous species, especially oilseed rape (OSR). In this paper, we evaluated the gametocidal effects of 27 ALS-inhibiting herbicides that were sprayed on OSR plants at the bolting stage. Results: OSR anther development was very sensitive to sublethal exposure to most ALS-inhibiting herbicides. The application of 18 out of the 20 tested sulfonylureas (except ethametsulfuron and ethoxysulfuron), two imidazolinones (imazethapyr and imazamox), and one sulfonylamino-carbonyltriazolinone (flucarbazone-sodium) at suitable rates could induce male sterility. Eight of the herbicides, including chlorsulfuron (at application rates of 60-120 mg/ha), halosulfuron-methyl (300-600 mg/ha), sulfosulfuron (400-600 mg/ha), triflusulfuron-methyl (500-750 mg/ha), pyrazosulfuron-ethyl (150-225 mg/ha), nicosulfuron (200-300 mg/ha), imazethapyr (750-1125 mg/ha), and imazamox (400-800 mg/ha), could induce over 90% male sterility and over 60% relative outcrossed seed set in six cultivars with different origins. These eight chemicals could be used as new gametocides for hybrid seed production. This study also examined the possibility of external application of these gametocides on several unstable Polima cytoplasmic male sterile and thermosensitive genic male sterile lines. Although the outcrossed seed set of the treated lines was slightly reduced, the gametocide application significantly increased the seed purity of the resulting hybrid. Conclusion: The finding of the gametocidal effects of most sulfonylureas and imidazolinones are of great importance for developing new functions for ALS-inhibiting herbicides. The application of gametocides will also greatly promote the safe utilization of environment-sensitive male sterility in hybrid seed production. Unexpectedly, the application of three triazolopyrimidines (florasulam, flumetsulam, and penoxsulam) and one pyrimidinylthiobenzoate (bispyribac-sodium) did not cause male sterility, although these herbicides obviously inhibited the activity of ALS and plant growth. This result suggests that inhibition of ALS activity does not always lead to male sterility in plants, and these gametocides may also inhibit other biological functions vital for microspore development.
Article
Full-text available
The expansion of land use for agricultural interests and the excessive use of herbicides are among the causes of biodiversity losses in the Brazilian Cerrado biome. Therefore, we aimed to test the hypothesis that Dipteryx alata Vogel, a common species in this biome, is sensitive to nicosulfuron because of its high phytotoxicity. We evaluated physiological, biochemical and morphological responses in D. alata plants exposed to increasing doses of the herbicide. Young plants were transplanted to 10 L pots containing substrate composed of soil and sand (2:1) after fertilization. After an acclimation period, the following doses of nicosulfuron were applied: 0 (control), 6, 12, 24, 48, and 60 g a.e. ha−1. The experiment was conducted in a randomized block design factorial scheme with six doses of nicosulfuron, three evaluation times, and five replicates per treatment. The effects of the herbicide were assessed by measuring gas exchange, chlorophyll a fluorescence, photosynthetic pigments, membrane permeability, antioxidant enzymes and acetolactate synthase. Nicosulfuron altered the photosynthetic machinery and enzymatic metabolism of D. alata. Reductions in physiological traits, increased catalase and ascorbate peroxidase activities, enhanced malondialdehyde concentrations rate of electrolyte leakage and decreased acetolactate synthase activity in response to nicosulfuron all suggest that D. alata is sensitive to this herbicide. As part of the Springer Nature SharedIt initiative, we can now publicly share a full-text view-only version of our paper using the link: https://rdcu.be/b1kuh
Article
Full-text available
Genotoxic, biochemical and individual organizational effects on Leptodactylus latinasus tadpoles were evaluated after exposure to an imazethapyr (IMZT)-based commercial herbicide formulation, Pivot® H (10.59% IMZT). A determination of the value of the lethal concentration (LC50) was determined as a toxicological endpoint. Alterations in animal behaviour and morphological abnormalities as well as cholinesterase (ChE), catalase (CAT) and glutathione S-transferase (GST) activities were employed as individual sublethal endpoints. Micronuclei frequencies (MNs), binucleated cells (BNs), blebbed nuclei (BLs), lobed nuclei (LBs), notched nuclei (NTs), erythroplastids (EPs), and evaluation of DNA strand breaks were employed as genotoxic endpoints. All biomarkers were evaluated after 48 and 96 h of exposure to concentrations of IMZT within 0.07-4.89 mg/L. LC5096h values of 1.01 and 0.29 mg/L IMZT were obtained for Gosner stages 25 and 36, respectively. Irregular swimming, diamond body shape and decreased frequency of keratodonts were detected at both sampling times. Results showed that IMZT increased GST activity and MNs frequency at 48 and 96 h of exposure. Other nuclear abnormalities were also observed in the circulating erythrocytes of tadpoles, i.e., NTs and BLs values after 48 h, and LNs, BLs and EPs values after 96 h. Finally, results showed that IMZT within 0.07-0.22 mg/L increased the genetic damage index in tadpoles exposed for both exposure times (48 and 96 h). This study is the first to report the sublethal biochemical effects of IMZT in anurans and is also the first report using L. latinasus tadpoles as a bioindicator for ecotoxicological studies.
Article
Full-text available
Euphorbia heterophylla is a weed species that invades extensive crop areas in subtropical regions of Brazil. This species was previously controlled by imazamox, but the continuous use of this herbicide has selected for resistant biotypes. Two biotypes of E. heterophylla from southern Brazil, one resistant (R) and one susceptible (S) to imazamox, were compared. The resistance of the R biotype was confirmed by dose-response assays since it required 1250.2 g ai ha-1 to reduce the fresh weight by 50 % versus 7.4 g ai ha-1 for the S biotype. The acetolactate synthase (ALS) enzyme activity was studied using ALS-inhibiting herbicides from five different chemical families. The R biotype required the highest concentrations to reduce this enzyme activity by 50 %. A Ser653Asn mutation was found in the ALS gene of the R biotype. The experiments carried out showed that imazamox absorption and metabolism were not involved in resistance. However, greater 14C-imazamox root exudation was found in the R biotype (~70 % of the total absorbed imazamox). Target site mutation in the ALS gene is the principal mechanism that explains the imazamox resistance of the R biotype, but root exudation seems to also contribute to the resistance of this biotype.
Article
Full-text available
Recently, the residues of some common and widely used herbicides (acetochlor, bispyribac-sodium, bentazon, bensulfuron-methyl, halosulfuron-methyl, and quinclorac) were detected in the surface water, soil, sediments, and fish tissues as the agricultural drainage problems. In this study, juveniles of Nile tilapia Oreochromis niloticus were exposed to sub-lethal concentrations of these herbicides as 2.625, 0.800, 36.00, 2.50, 1.275, and 11.250 mg/l for acetochlor, bispyribac-sodium, bentazon, bensulfuron-methyl, halosulfuron-methyl, and quinclorac respectively for 96 h. Some hemato-biochemical parameters were evaluated. In comparison with the control group, sub-lethal concentrations of all tested herbicides induced alterations in the shape of erythrocytes. Also, in all tested herbicides, hematological parameters of exposed fish exhibited a significant decrease in red blood cell count except bentazon. However, all tested herbicides showed an insignificant reduction in mean corpuscular hemoglobin concentration and total white blood cells except bensulfuron-methyl. For biochemical parameters, most tested herbicides induced a significant increase in levels of cholesterol, albumin, globulin, albumin/globulin ratio, activity of alkaline phosphatase (ALP), alanine aminotransferase (ALT), aspartate aminotransferase (AST), and total plasma protein (only with acetochlor), urea, and creatinine (except bentazon and halosulfuron-methyl that exhibited non-significant decrease in creatinine level) compared with the control. In conclusion, the fish blood profiles can be used as good biomarkers for laboratory study to assess the toxicity of the tested rice herbicides at a sub-acute level especially acetochlor on O. niloticus. Graphical Abstract
Article
A QuEChERS method with liquid chromatography-tandem mass spectrometry (LC-MS/MS) was modified and validated for the determination of florasulam and pyroxsulam residues in wheat grain and straw. The validated method was applied to cereals including oat, millet, corn and rice. Average recoveries were 76-113% with RSDs 2-15%. The limits of quantitation (LOQ) were 0.005 mg/kg for wheat grain and 0.01 mg/kg for wheat straw and four cereals. Ion suppression for florasulam (−28% to −76%) was observed in all the matrices except corn, whereas ion enhancement were shown for pyroxsulam (44% to 83%). Degradation rates of florasulam and pyroxsulam were 6% and 23%, respectively, in wheat grain and straw after eight-week storage at −20°C. The ultimate residues in field trials in ten regions were all ≤0.05 mg/kg, and long term dietary risk assessment indicated that hazard quotients were 0.02% and 0.001% for florasulam and pyroxsulam, respectively, which shows that it is safe to spray the two herbicides on wheat.
Article
Bispyribac sodium is frequently used herbicide in the rice field. Though, it has been targeted to kill rice weeds, but its non-target effect on soil microbes in paddy soil was largely unknown. Therefore, in the present study, an attempt was made to assess the non-target effect of bispyribac sodium on alteration of functional variation of soil microbial community and their correlation with microbial biomass carbon (MBC) and soil enzymes. A microcosm experiment set up was made comprising three treatments viz., control (CON) (without application of bispyribac sodium), recommended dose of bispyribac sodium (35 g ha-1) (BS), and double the dose of BS (70 g ha-1) (DBS). Results indicated that the MBC and soil enzyme activities (dehydrogenase, alkaline phosphatase and urease) in BS and DBS-treated soil were significantly (p < 0.05) declined from 1st to 30th day after application as compared to CON. Counts of heterotrophic bacteria, actinomycetes and fungal population were also decreased in BS and DBS-treated soil. The average well color development (AWCD) values derived from Biolog®ecoplates followed the order of DBS ˂ BS ˂ CON. Shannon index value was high (p ≤ 0.05) in CON compared to soil-treated with BS and DBS. Principal component analysis (PCA) showed a clear distinction of the cluster of treatments between CON, BS and DBS. Biplot analysis and heatmap suggested that carboxylic compounds and amino acids showed positive response towards BS-treated soil, whereas phenolic compounds had positive correlation with DBS-treated soil. PCA analysis indicated that oligotrophs was rich in BS-treated paddy soil, whereas copiotrophs and asymbiotic nitrogen fixers were richer in DBS treatment. Overall, the present study revealed that application of recommended dose of BS and its double dose alter the soil microbial population, enzyme activities and functional microbial diversity in paddy soil.
Article
Imidazolinone herbicides are a group of chiral herbicides that are widely used to control weeds in crops. Despite their wide use, few studies on the behavior of enantiomers in terrestrial systems have been reported. In this study, the bioaccumulation of imazamox, imazapic, and imazethapyr enantiomers in earthworm and their degradation in soils were assessed using earthworm-soil microcosms. The bioaccumulation of the three herbicides in earthworm was not significantly enantioselective. Imazamox and imazethapyr did not significant stereoselective degradation in soil (p > 0.05), while the enantioselectivity of the degradation of imazapic was significant (p < 0.05). Furthermore, biota to soil accumulation factor (BSAF) values were also calculated for three herbicides. Relationships between BSAF values and organic matter content of soil and log KOW of herbicides were investigated. The BSAFs values were negatively correlated with the log KOW of herbicides, and were positively correlated with organic matter content of soil in earthworm-soil microcosms. These relationships indicated that chemical hydrophobicity (Kow) and organic matter content of soil were good predictors to estimate the bioavailability of imidazolinone herbicides to earthworm.
Article
Field and laboratory studies were conducted in 1987 and 1988 on several Arkansas soils to determine the effects of residual levels of imazaquin on cotton. Imazaquin concentrations ranging from 0.007 to 0.024 μg/g and 0.00 to 0.015 μg/g were found in soil samples taken 12 and 24 months, respectively, following preplant-incorporated applications at rates of 140 g/ha. Cotton yields were reduced from 7 to 42% in 1987 as the soil concentration of imazaquin increased from 0.007 to 0.024 μg/g. Imazaquin persistence was greater in clay soils than loams or sandy loams. Imazaquin was weakly adsorbed to a clay, loam, and sandy loam soil in laboratory studies; the adsorption isotherms did not differ significantly by soil texture. A bioassay utilizing cotton as the indicator species was developed to determine imazaquin concentrations in soils ranging from 0 to 0.06 μg/g. Imazaquin concentrations determined by the bioassay method were found to be 75, 77, and 80% of those determined by chemical extractions for a clay, loam, and sandy loam soil, respectively.