ArticlePDF Available

Residue of bensulfuron methyl in soil and rice following its pre- and post-emergence application

Authors:

Abstract and Figures

Bensulfuron methyl (BSM) is applied in rice to control a wide range of weeds due to low application rate and high efficiency. A study was conducted to evaluate residues of BSM in soil and rice plant at different doses as pre- and post-emergence application in transplanted rice. The quick easy cheap effective rugged safe (QuEChERS) method was evaluated for BSM residue extraction from different matrices. The limit of detection and limit of quantification was 0.005 and 0.01 µg/g, respectively in soil and rice plant. The average BSM recovery of 91.1, 82.8, 84.5 and 88.7% was obtained from soil, rice straw, grain and husk, respectively. Though, BSM residue was detected (0.011 to 0.017 μg/g) in soil at high dose, it was below maximum residue limit (0.01 μg/g) in rice grain at both the doses of BSM. Hence, the study revealed that the BSM can be safely applied to rice at recommended doses for weed control. © 2016, Czech Academy of Agricultural Sciences. All rights reserved.
Content may be subject to copyright.
Rice (Oryza sativa L.) is the vital staple food for
about 50% of the world’s population that lives in
Asia. Hence, a worldwide research is continuing
to enhance the production and productivity of the
rice crop. A major production constraint in rice
cultivation is the lack of timely weed management
caused by the acute labour scarcity and high manual
weeding cost. Thus, the use of herbicides for weed
control is encouraged in rice production. Asia ac-
counts for vast majority of the global rice herbicide
market and the share continues to grow, since the
multinational agrochemical companies acquire the
advantage of a more liberal trading climate in Asia
(Naylor 1996). Although the use of herbicides has
increased the crop production to cope up with food
demand, there may be an unintentional exposure
of the ecosystem to herbicides residue. Hence, an
inappropriate and indiscriminate use of herbicides
in rice cultivation could cause environmental con-
tamination through longer persistence.
During last 20 years, low application rate her-
bicides viz., sulfonyl urea, sulfonamide, and imi-
dazolinone have been developed and registered
for weed control all over the world. They act by
inhibiting the action of acetolactate synthase, a
key plant enzyme essential for the synthesis of
branched-chain amino acids (Moberg and Cross
1990, Stidham 1991). Among the above classes,
sulfonyl ureas are registered largely for chemical
weed management in rice either as pre- or post-
emergence herbicide for controlling the grasses,
broad leaved weeds and sedges (Russel et al. 2002).
Bensulfuron methyl (BSM) (methyl 2-[(4,6-dimeth-
oxypyrimidin-2-yl) carbamoyl sulfamoyl methyl]
benzoate), one of the environmental-friendly low
dose herbicide belonging to sulfonyl urea is used
highly for weed control in rice. It is marketed
as Londax (single herbicide) and Londax Power
(combination with pretilachlor) in India.
Because of low application rate, its concentration
is expected to be low in water and soils. However,
it depends on the herbicide molecule and for-
mulation properties, application rate, and areas
or type of soils treated with this herbicide. The
solubility of BSM depends on pH of water and
has high solubility (880 mg/L) in water at pH 8.0
Residue of bensulfuron methyl in soil and rice following
its pre- and post-emergence application
P. Janaki1, C. Nithya2, D. Kalaiyarasi2, N. Sakthivel2, N.K. Prabhakaram2,
C. Chinnusamy2
1Soil Science and Agricultural Chemistry, Tamil Nadu Agricultural University,
Tamil Nadu, India
2Department of Agronomy, Tamil Nadu Agricultural University, Tamil Nadu, India
ABSTRACT
Bensulfuron methyl (BSM) is applied in rice to control a wide range of weeds due to low application rate and high
efficiency. A study was conducted to evaluate residues of BSM in soil and rice plant at different doses as pre- and
post-emergence application in transplanted rice. e quick easy cheap effective rugged safe (QuEChERS) method
was evaluated for BSM residue extraction from different matrices. e limit of detection and limit of quantifica-
tion was 0.005 and 0.01 µg/g, respectively in soil and rice plant. e average BSM recovery of 91.1, 82.8, 84.5 and
88.7% was obtained from soil, rice straw, grain and husk, respectively. ough, BSM residue was detected (0.011 to
0.017 g/g) in soil at high dose, it was below maximum residue limit (0.01 g/g) in rice grain at both the doses of
BSM. Hence, the study revealed that the BSM can be safely applied to rice at recommended doses for weed control.
Keywords: Oryza sativa L.; weed management; herbicide; contamination; persistence
428
Vol. 62, 2016, No. 9: 428–434 Plant Soil Environ.
doi: 10.17221/294/2016-PSE
(25°C) and very low solubility (1.1 mg/L) at pH 5.0
(Roberts et al. 1998, Vencill 2002). It is highly stable
under slightly alkaline aqueous solutions (pH 8)
and degrades slowly under acidic conditions. It is
stable to direct photolysis in sterile buffer solu-
tions and degrades rapidly in natural water under
sunshine radiation by cleavage of the sulfonylu-
rea linkage to methyl 2-(sulfomethyl) benzoate
and (4,6-dimethoxypyrimidin2-yl) urea, with a
degradation half life (DT50) of 3–4 days (Roberts
et al. 1998). Bensulfuron methyl degrades also
through chemical hydrolysis and microbial pro-
cesses in moist soils and is classified as immobile
to moderately mobile depending on the soil organic
matter and pH (Roberts et al. 1998). The major
degradation product under aerobic metabolism is
CO2; however, under sterile conditions the non-
volatile compounds were also produced, namely
sulfonamide and pyrimidine amine [2-amino-4,6-
dimethoxypyrimidine]. This showed that the mi-
crobial degradation is not obligatory for BSM
degradation (CDFA 1989).
Since sulfonyl ureas act upon a specific plant
enzyme acetolactate synthase that is not found
in mammals or other animals, their toxicities to
animals is very low (Brown 1990). As they are very
active at low concentrations, residual phytotox-
icity of sulfonyl ureas to rotation crops such as
corn, sunflowers, sugar beets, and dry beans has
been already reported in literature (Anderson and
Humburg 1987, Curran et al. 1991). The rice plant
metabolizes the BSM rapidly with a DT50 of 4–6 h.
The crop selectivity is due to the slower rate of
translocation from roots to shoots and an increased
rate of metabolism in rice (Takeda et al. 1986).
The available literature on BSM is limited to bioef-
ficacy, laboratory dissipation studies (Langeland and
Lorache 1994) and very few under field conditions
(CDFA 1989, EFSA 2008). Similarly, the influence
of time of application on the persistence of BSM
has not been reported. In view of these facts, the
present study was undertaken to study the persis-
tence of BSM in soil and its terminal residues in
rice as influenced by the time of application under
subtropical arid agro-climatic conditions.
MATERIAL AND METHODS
Experimental details. Field experiments
were conducted during the kharif season (June
to September) of 2012 and 2013 at a wet land
farm of the Tamil Nadu Agricultural University,
Coimbatore with rice (cv. ADT 48) as a test
crop. As herbicide bensulfuron methyl (60%
dry flowable (DF)) was applied at two different
doses (100 and 200 g active ingredient (ai)/ha)
as pre- and post-emergence along with control
(no herbicide). The experiment was conducted in
randomized block design and the treatments were
replicated thrice. The pre- and post-emergence
application of bensulfuron methyl was done on 3rd
and 14th day after transplanting, respectively, using
flat fan nozzle and knapsack sprayer with the spray
volume of 400 L/ha. All the management practices
were followed as prescribed for rice cultivation
in the crop production guide. The experimental
farm is located at 77oE, 11oN latitude and 426 m
a.s.l. Experimental field soil was clay loam in tex-
ture and belongs to Typic Chromusterts group
and Noyyal soil series. The experimental field
soil was medium in organic carbon status (0.60%)
with the available nutrient status of low nitrogen
(226 kg/ha), medium phosphorus (18.3 kg/ha)
and high potassium (458 kg/ha) and has alkaline
soil reaction (8.27) and the electrical content of
below 0.43 dS/m.
One week after rice harvest, the green gram (Vigna
mungo) cv. CO 6 was grown in strips without dis-
turbing the layout and observation on germination
percent, plant height, number of pods per plant and
seed yield at harvest were recorded to assess the car-
ryover effect of BSM applied to the main crop (rice).
Collection of samples. e soil samples were col-
lected for residue analysis at the time of harvest from
0–15 cm depth. Five core soils were randomly taken
using soil auger from each treated and untreated plot
avoiding outer 20 cm fringes of plot. e core soils
were pooled from each plot, air dried, powdered and
sieved through 2 mm sieve and stored for analyses.
e paddy grain and straw samples were collected
from BSM treated and untreated plots at harvest.
Grains of rice were removed from husk and crushed
using a mechanical blender, while straw was chopped
into small pieces using knife.
Meteorogical conditions. Weather parameters
prevailing during the cropping period were record-
ed (Figure 1) at weekly intervals. It was observed
that the maximum and minimum temperatures
ranged from 18.4–33.3°C during kharif 2012 and
from 21.6–32.8°C during kharif 2013, respectively.
The total rainfall per day recorded ranged from
429
Plant Soil Environ. Vol. 62, 2016, No. 9: 428–434
doi: 10.17221/294/2016-PSE
6.1–165.2 mm during kharif 2012 and 0.6–1.8 mm
during kharif 2013.
Determination of bensulfuron methyl residues
in soil and plant samples. The BSM residue was
extracted by quick easy cheap effective rugged
safe (QuEChERS) technique (Janaki et al. 2015)
using 1% HOAc in MeCN + dichloromethane and
anhydrous MgSO4 and NaOAc · 3 H2O in vertex
mixer. An aliquot was cleaned up using the Bond
Elut C18 SPE cartridge with acetonitrile as eluent
and then concentrated in rotary vacuum evapora-
tor for LC-DAD analysis.
Validity of the method was tested by conducting
the recovery experiments using spiked samples of
soil and rice plant parts from the control treatment
with the known BSM standards concentration of
0.01, 0.05, 0.1 and 0.5 µg/g in three replicates.
After spiking, the residue of BSM was extracted
and cleaned up as described for samples above. The
concentration of BSM was determined by compar-
ing the peak area of the samples and calibration
curves of five levels of standards. A quantification
limit of 0.01 g/g was used for the calculation. The
blank soil and rice plant parts reference was used to
establish the limit of quantification. A calibration
curve was prepared by plotting concentrations of
BSM on x-axis against the average peak area on
y-axis (Figure 2).
Instrumentation. The BSM residues were de-
termined using Agilent HPLC (1200 series) with
Diode Array Detector (DAD) detector, binary pump
and auto sampler having rheodyne injection sys-
tem. The compound was separated using Agilent
Eclipse XDB-C 18, 5 µm, 4.6 × 150 mm column
kept in thermos-stated oven maintained at 25°C.
The instrument was connected to a computer that
recorded the response in terms of peak area and
height using the EZChrom software (USA). The
acetonitrile:water (50:50% v/v) with orthophos-
phoric acid (pH 3.0) was used as a mobile phase
for the separation of BSM with the flow rate of
0.5 mL/min. The injection volume of sample was
20 µL. Detection was performed at 234 nm for all
the standards and unknown samples.
RESULTS AND DISCUSSION
Recoveries and detection limit. Equations of
analytical calibration graphs, obtained by plotting
peak areas on the y axis against concentrations of
BSM on the x axis showed good linearity (Figure 2)
with the correlation coefficient of 0.979. The re-
tention time of BSM standards and samples was
4.50 ± 0.2 min (Figures 3 and 4) under the given
instrumental conditions of HPLC-DAD. The av-
0
20
40
60
80
100
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Weather parameters
Kharif 2012
0
5
10
15
20
25
30
35
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Weather parameters
Crop growing period (in weeks)
Kharif 2013
Maximum temperature C) Minimum temperature C)
Sunshine (h/day) total rain fall (mm)
Figure 1. Distribution of
weather parameters dur-
ing the growing period of
rice
0
20
40
60
80
100
12345678910 11 12 13 14 15 16
Weather parameters
Kharif 2012
0
5
10
15
20
25
30
35
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Weather parameters
Crop growing period (in weeks)
Kharif 2013
Maximum temperature (°C) Minimum temperature C)
Sunshine (h/day) total rain fall (mm)
0
20
40
60
80
100
12345678910 11 12 13 14 15 16
Weather parameters
Kharif 2012
0
5
10
15
20
25
30
35
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Weather parameters
Crop growing period (in weeks)
Kharif 2013
Maximum temperature C) Minimum temperature (°C)
Sunshine (h/day) total rain fall (mm)
430
Vol. 62, 2016, No. 9: 428–434 Plant Soil Environ.
doi: 10.17221/294/2016-PSE
erage recovery of BSM obtained from soil, rice
straw, grain and husk was 91.1, 82.8, 84.5 and
88.7%, respectively. The results of the study are
presented in Table 1. The limit of detection and
quantification of BSM in all the matrices was found
to be 0.005 and 0.01 µg/g, respectively, with the
signal to noise ratio of 3:1. The recovery of BSM
from different matrices was found to be accept-
able up to fortification level of 0.01 g/g (Table 1)
and was also found satisfactory at different con-
centration levels. Niell et al. (2010) also reported
the BSM recovery of 91% from rice grain by modi-
fied QuEChERS extraction using 1% acetic acid in
acetonitrile, MgSO4 and sodium acetate.
Terminal residues of bensulfuron methyl in
rice grain, husk and straw. Irrespective of time
(pre- or post-emergence) and dose (100 or 200 g/ha)
of application, the BSM residue was below the quan-
tification limit of 0.01 µg/g in the rice straw, grain
and husk at the time of harvest in both years (Kharif
2012, 2013). This could be due to the selectivity
of rice crop to the BSM residue by slower rate of
translocation from roots to shoots and through an
increased rate of metabolism. Takeda et al. (1986)
reported that the rice shoots metabolized BSM with
a DT50 of 4–6 h while sensitive broad-leaved and
sedge weeds did not degrade BSM (DT50 > 50 h).
According to Priester (1985), the metabolism pro-
ceeds initially through -demethylation of the
pyrimidine ring to yield methyl α-(4-hydroxy-6-
methoxypyrimidin-2-ylcarbamoyl sulfamoyl)-o-
toluate and the hydrolysis of the parent methyl
ester to produce bensulfuron (Usui et al. 1993).
Similar results of less than 0.03 mg/kg of BSM
residue in rough rice, husk and rice straw were
also reported by Wu et al. (2000).
The residue of BSM in rice grain, husk and straw
was below the maximum residue limit (MRL) in
rice parts set by the EFSA (0.02 mg/kg) for the
European union (EFSA 2008), FSSAI (0.01 mg/kg)
for India (FSSAI 2015), FSCA (0.02 mg/kg in rice
and 0.05 mg/kg in rice bran) for Australia (FSCA
2014) and Japan (0.1 mg/kg) by Clever and Sato
(2011). Wei and Chen (1995) also reported a below
MRL residue of BSM (0.0116 g/g) in rice after 98
days of application. The present results suggest
that the normal rate (100 g/ha) of BSM applica-
tion is environmentally safe as its residues were
found below the MRL in rice grain and straw sets
by different agencies of various countries.
Terminal residues of bensulfuron methyl in
field soil. The sulfonyl ureas degrade in soil pri-
marily by the microbial or chemical metabolism
and conversion through photochemical occurs
only in the presence of UV light, which is a minor
decomposition mechanism (Singh et al. 2010).
The dissipation of sulfonyl ureas in soil is gener-
ally influenced by pH and other minor factors
are temperature, moisture and organic matter.
Though the residue of BSM was not detected in
the rice straw, husk and grain at doubled dose of
200 g/ha, its residue was detected in field soil at
harvest during kharif 2012 under both the pre-
and post-emergence application. It was below
y = 1E+06x + 51950
R² = 0.9796
0
200 000
400 000
600 000
800 000
1 000 000
1 200 000
1 400 000
1 600 000
0.0 0.2 0.4 0.6 0.8 1.0
Area (mAU)
Concentration of bensulfuron methyl standards
(mg/L)
Figure 2. Calibration of bensulfuron methyl standard
at concentration levels of 0.01–5.0 g/mL
Figure 3. Bensufuron methyl standard 0.01 µg/mL de-
tected by high-performance liquid chromatography
with diode-array detection (HPLC-DAD)
(min)
0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0
(mAU)
22
20
18
16
14
12
10
8
6
4
2
0
–2
–4
–6
–8
Bensulfuron methyl 4.527 28138
DAD: Signal A,
234 nm/
Bw: 4 nm
– Bensulfuron methyl
0.01 ppm
BSM std. dat
431
Plant Soil Environ. Vol. 62, 2016, No. 9: 428–434
doi: 10.17221/294/2016-PSE
the quantification limit of 0.01 µg/g during kharif
2013 (Table 2 and Figure 5). The quantity of BSM
residue detected in soil at post-emergence applica-
tion is higher (0.017 µg/g) than at pre-emergence
(0.011 µg/g) application. The presence of bensul-
furon residues in soil even at the time of harvest
suggests that being a weak acid (pKa of 5.2), it would
be relatively stable in the observed high pH of the
present experimental field soils (Langeland and
Lorache 1994) and hence it persists until harvest.
The presence of residue in soil during kharif 2013
below the detectable level might be the results of
variation in rainfall and other weather variables
(Figure 5). The amount of rainfall received on the
first month of bensulfuron application during kha-
rif 2012 was low when compared to kharif 2013.
This might have enhanced BSM sorption to soil
and reduced its leaching or runoff. Hence, it was
detected above the quantification limit in soil at
harvest. The enhanced dissipation of BSM residue
during kharif 2013 from soil could be the result
of enhanced hydrolysis and microbial degradation
by even distribution of rainfall throughout the
cropping period (Figure 1). Afyuni et al. (1997)
indicated that 1.1% to 2.3% of applied sulfonyl
ureas was lost in runoff during a simulated rainfall
event 24 h after herbicide application.
Bensulfuron methyl had medium organic carbon
normalized distribution coefficient (Koc) values
(205–567), which indicates that it is moderately
mobile (EFSA 2008) and more adsorption could
be anticipated for clay soils depending on the soil
organic matter and pH. The chemical hydrolysis
of BSM to ortho carbo methoxy group and further
microbial breakdown is the important route of its
degradation in soil (Langeland and Lorache 1994).
However, in the present study the possibility for
chemical hydrolysis is low due to the alkaline pH
of the field water (> 8.5) in which it is highly stable.
Hence in the present study, the BSM degradation
might have been enhanced by the increased sun-
shine radiation and day temperature independent
of pH as suggested by CDFA (1989) through the
formation of photo degradation products like beta-
lactic acid via the complete breakdown of phenyl
ring at low concentrations. Yordy (1987) reported
that in moist field soils, BSM has degraded via
chemical hydrolytic degradation and microbial pro-
cesses with a DT50 of 3–4 weeks by producing the
compounds like methyl 2-(aminosulfonyl methyl)
benzoate and 4,6-dimethoxy-2-aminopyrimidine
and [1H-2,3-benzothiazin-4(3H)-one 2,2-dioxide]
and CO2. Wei and Chen (1995) reported the half-life
of 28 days for bensulfuron in soil with the residue
Figure 4. Bensufuron methyl standard 1.0 µg/mL de-
tected by high-performance liquid chromatography
with diode-array detection (HPLC-DAD)
Table 1. Recovery (%) of bensulfuron methyl from soil, rice grain, husk and straw
Fortified concentration
(g/g) Soil Rice grain Rice husk Rice straw
0.01 79.0 ± 2.10 78.0 ± 1.19 87.0 ± 2.92 72.0 ± 2.87
0.05 91.6 ± 2.36 78.8 ± 2.45 89.2 ± 3.01 71.2 ± 4.17
0.10 97.4 ± 2.89 86.7 ± 1.79 87.4 ± 3.12 90.8 ± 3.16
0.50 96.2 ± 3.19 94.4 ± 2.91 91.0 ± 3.94 97.2 ± 4.27
Average recovery (%) 91.1 84.5 88.7 82.8
Mean of three replications; Values ± are % standard deviation
(min)
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
35
30
25
20
15
10
5
0
–5
–10
–15
DAD: Signal A,
234 nm/
Bw: 4 nm
– Bensulfuron methyl
1.0 ppm
BSM std. dat
(mAU)
Bensulfuron methyl
4.547 132 5294
432
Vol. 62, 2016, No. 9: 428–434 Plant Soil Environ.
doi: 10.17221/294/2016-PSE
of 0.0015 g/g in postharvest soil when it was ap-
plied at normal rate. Hence, the continuous and
indiscriminate use of BSM may be circumvented to
overcome the biomagnifications of its residues in
soil environment due to its medium to low mobil-
ity nature in soil with organic carbon normalized
freundlich coefficient (Kfoc) of 205–561 mL/g
(EFSA 2008).
Obrigawitch et al. (1998) found that the presence
of 0.1 g/ha active ingredient of sulfonyl ureas is
the threshold dose in soil and would be enough
to decrease the yields of even the most sensitive
non-target plants. Hence, in the present study, sen-
sitive indicator species namely green gram (Vigna
mungo) was grown as succeeding crop to assess
the phytotoxicity of BSM residue in soil. During
both years of study, the residual phytotoxicity was
not observed on green gram and the parameters
viz., germination percent, plant height, number
of pods per plant and seed yield were found to be
unaffected by the time and dose of BSM application.
The present results suggest that the risk of BSM
carryover to succeeding crops planted following
pre- or post-emergence application to rice is low in
tropical environment at the normal rate as single
application. However, the effect of repeated and
continuous application needs to be investigated
in wetland rice growing environment.
In conclusion, it can be concluded that the
QuEChERS method could be well applied for the
BSM extraction from soil and rice parts to the
level of below MRL prescribed by different coun-
tries. It is observed that the rice grown in soil of
neutral to alkaline pH, high sorption capacity and
even distribution of high rainfall might lead to
less persistence of BSM residues after its pre- or
post-emergence single application at normal rate
of 100 g/ha. The terminal BSM residue in rice
grain, husk and straw was found to be below the
maximum residue limit set by the Japan, FSSAI
and some European countries when it is applied
at the normal rate ranged from 100–200 g/ha.
However, the effect of repeated and continuous
application on its dissipation in soil and residues
in rice needs to be investigated in wetland rice
growing environment as there is a chance for the
biomagnifications of herbicide residues in soil
and crop.
Acknowledgements
The authors are grateful to the Directorate of
Weed Research, Jabalpur and the Tamil Nadu
Agricultural University, India for providing nec-
essary research facilities to carry out the work.
The authors declare that they have no conflict of
interest regarding this study.
Table 2. Residue of bensulfuron methyl (µg/g) in soil at harvest
Method of
application
Kharif 2012 Kharif 2013
pre-emergence post-emergence pre-emergence post-emergence
Dose 100 200 100 200 100 200 100 200
(g ai/ha)
Soil < LOQ 0.011 ± 7.17 < LOQ 0.017 ± 6.89 < LOQ < LOQ < LOQ < LOQ
LOQ – limit of quantification (0.01 µg/g); Values ± are % standard deviation
Figure 5. Detection of bensulfuron methyl residue in
post-harvest soil during Kharif 2012 by high-perfor-
mance liquid chromatography with diode-array detec-
tion (HPLC-DAD)
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
(min)
260
240
220
200
180
160
140
120
100
80
60
40
20
0
–20
(mAU)
Bensulfuron methyl 4.460 15382
DAD: Signal A,
234 nm/
Bw: 4 nm
– Bensulfuron methyl
POE - soil - R3T4.dat
433
Plant Soil Environ. Vol. 62, 2016, No. 9: 428–434
doi: 10.17221/294/2016-PSE
REFERENCES
Afyuni M.M., Wagger M.G., Leidy R.B. (1997): Runoff of two
sulfonylurea herbicides in relation to tillage system and rainfall
intensity. Journal of Environmental Quality, 26: 1318–1326.
Anderson R.L., Humburg N.E. (1987): Field duration of chlor-
sulfuron bioactivity in the central great plains. Journal of
Environmental Quality, 16: 263–266.
Brown H.M. (1990): Mode of action, crop selectivity, and soil
relations of the sulfonylurea herbicides. Pest Management
Science, 29: 263–281.
CDFA (1989): Bensulfuron Methyl (Londax). California, California
Department of Food and Agriculture.
Clever J., Sato J. (2011): Japan proposes new MRLs for bensul-
furon-methyl cyclanilide and Spinosad. Gain Report Number:
JA1044. Dt. 11/1/2011. Available at: www. http://gain.fas.
usda.gov/
Curran W.S., Knake E.L., Liebl R.A . (1991): Corn (Zea mays)
injury following use of clomazone, chlorimuron, imazaquin,
and imazethapyr. Weed Technology, 5: 539–544.
EFSA (2008): Conclusion on pesticide peer review: Conclusion
regarding the peer review of the pesticide risk assessment of
the active substance bensulfuron. EFSA scientific report 178:
1–102. Available at: http://www.efsa.europa.eu/en/efsajournal/
doc/178r.pdf
FSCA (Food Standards Code of Australia) (2014): Available at:
http://www.foodstandards.gov.au /code/ code-revision/Docu-
ments/Schedules/78%20-%20Schedule%2020%20.pdf
FSSAI (2015): Available at: http://www.fssai.gov.in/Portals/0/
Pdf/Draft_WTO_Notification_ Pesticides_ 23_11_2015.pdf
Janaki P., Sundaram K.M., Chinnusamy C., Sakthivel N. (2015):
Determination of residues of metribuzin in soil and sugarcane
by QuEChERS. Asian Journal of Chemistry, 27: 3692–3696.
Langeland K.A., Lorache F.B. (1994): Persistence of bensulfuron
methyl and control of hydrilla in shallow ponds. Journal of
Aquatic Plant Management, 32: 12–14.
Moberg W.K., Cross B. (1990): Herbicides inhibiting branched-
chain amino acid biosynthesis. Pest Management Science, 29:
241–246.
Naylor R. (1996): Herbicides in Asian Rice: Transitions in Weed
Management. Palo Alto (California): Institute for International
Studies, Stanford University and Manila (Philippines): Inter-
national Rice Research Institute, 270.
Niell S.L., Pareja L., Geis Asteggiante M.V., Cesio H.H. (2010):
Comparison of extraction solvents and conditions for herbicide
residues in milled rice with liquid chromatography-diode array
detection analysis (LCDAD). Food Additives and Contaminants:
Part A: Chemistry, Analysis, Control, Exposure and amp. Risk
Assessment, 27: 206–211.
Obrigawitch T.T., Cook G., Wetherington J. (1998): Assessment
of effects on non-target plants from sulfonylurea herbicides
using field approaches. Pest Management Science, 52: 199–217.
Priester T. (1985): Metabolism of [2-14C-Pyrimidine] DPX-F5384
in Indica rice grown in a rice paddy ecosystem. DuPont Internal
Report, AMR 315–385.
Roberts T.R., Hutson D.H., Lee P.W., Nicholls P.H., Plimmer J.R.
(1998): Metabolic Pathways of Agrochemicals: Part 1: Herbi-
cides and Plant Growth Regulators. Cambridge, Royal Society
of Chemistry, Thomas Graham House, 483–487.
Russel M.H., Saladini J.L., Lichtner F. (2002): Sulfonylurea Herbi-
cides. Cambridge, Royal Society of 3814 Chemistry – Pesticide
Outlook, 166–173.
Singh R.G., Singh S., Singh V., Gupta R.K. (2010): Efficacy of
azimsulfuron applied alone and tank mixed with metsulfuron +
chlorimuron (almix) in dry direct seeded rice. Indian Journal
of Weed Science, 42: 168–172.
Stidham M.A. (1991): Herbicides that inhibit acetohydroxyacid
synthase. Weed Science, 39: 428–434.
Takeda S., Erbes D.L., Sweetser P.B., Hay J.V., Yuyama T. (1986):
Mode of herbicidal and selective action of DPX-F5384 between
rice and weeds. Weed Research, 31: 157–163.
Usui K., Suwanwong S., Ohno A., Ishizka K. (1993): Zasso Ken-
kyu, 38: 7–11.
Vencill W.K. (2002): Bensulfuron methyl. In: Herbicide. 8th Ed.
Lawrence, Weed Science Society of America.
Wei C., Chen Z.Y. (1995): Residue of 14C-bensulfuron (Londax)
in rice. Journal of Ecology and Rural Environment, 11: 17–18.
Wu L.Q., Hu L.Y., Xu H., Zhu Y. (2000): Residue study of bensulfuron
methyl on rice. Acta Agriculturae Zhejiangensis, 12: 393–396.
Yordy J.R. (1987): Aerobic soil metabolism of DPX-F5384. Dupont
Internal Report, AMR 782–787.
Received on April 16, 2016
Accepted on August 8, 2016
Corresponding author:
Assist. Prof. Dr. P. Janaki, Ph.D., Tamil Nadu Agricultural University, Department of Agronomy, 641 003 Coimbatore,
India; e-mails: janakibalamurugan@rediffmail.com; janaki.p@tnau.ac.in
434
Vol. 62, 2016, No. 9: 428–434 Plant Soil Environ.
doi: 10.17221/294/2016-PSE
... Literature review reveals that the most popular technique for sulfonylurea herbicides extraction from plant matrices was QuEChERS. [63,[89][90][91][92][93][94][95] Single literature reports considered MSPD [97] and SPE [96] techniques. Final analysis was mostly conducted by LC with MS, [63,89,90,92,94] DAD [93,95,97] detector. ...
... [63,[89][90][91][92][93][94][95] Single literature reports considered MSPD [97] and SPE [96] techniques. Final analysis was mostly conducted by LC with MS, [63,89,90,92,94] DAD [93,95,97] detector. Also, other techniques that are rarely used in pesticides analysis, such as capillary zone electrophoresis (CZE) with ultraviolet [96] and MS/MS, [91] were applied. ...
... Also, other techniques that are rarely used in pesticides analysis, such as capillary zone electrophoresis (CZE) with ultraviolet [96] and MS/MS, [91] were applied. In the QuEChERS procedures, the most commonly used sorbent was C18 [89,[94][95][96] and PSA [91] in mixture with C18 [63,92,94] or GCB. [63,90] Effective in clean-up of these herbicide extracts was also Z-Sep (modified silica with zirconium oxide). ...
Article
The aim of this article is to present the trends in extraction techniques applied for the isolation of pesticides from plant matrix. To fully compare the effectiveness of different extraction techniques, it was required to analyze compounds with possibly wide spectrum of physicochemical properties. Hence, compounds representing neonicotinoids, pyrethroids, sulfonylureas and phenylamides were selected. Based on literature studies, it may be concluded that there are three main approaches to make the analytical procedures for pesticides determination more effective: (i) the optimization of extraction conditions, however, according to ANOVA conducted on the collected literature data, not all parameters influence the extraction process equally; chemometric studies based on literature reports may lead to the conclusion that the most favorable conditions (criterion: analyte recovery, repeatability) for neonicotinoid, pyrethroid and sulfonylurea herbicide extraction from plant tissues are provided by QuEChERS – extraction with acetonitrile, while the mixtures of PSA and GCB (for neonicotinoids), and PSA, GCB, C18 (for pyrethroids) should be used in d-SPE step. For sulfonylurea compounds and metalaxyl it was impossible to identify a sorbent(s) that cleans up the extract more effectively than the others; (ii) to develop a new generation of sorbents; however, the range of their applicability is limited, mainly due to difficulties in their synthesis; (iii) to develop the new extraction techniques with as few “trouble spots” as possible.
... Although the half lethal doses(LD50) of four herbicides in rats are all over 5000 mg/kg according to the Pesticide Manual [19], the four herbicides are extensively used in paddy fields in consideration of good selectivity, high efficiency, and low toxicity. Previous studies primarily concentrated on herbicidal activity and residue levels of four herbicides in rice cropping systems [10,20,21]. The results indicated that pyrazosulfuron-ethyl, bensulfuron-methyl, acetochlor and butachlor were safe and the residues in rice were lower than the maximum residue limit (MRL) at the recommended dosage [19,20]. ...
... Previous studies primarily concentrated on herbicidal activity and residue levels of four herbicides in rice cropping systems [10,20,21]. The results indicated that pyrazosulfuron-ethyl, bensulfuron-methyl, acetochlor and butachlor were safe and the residues in rice were lower than the maximum residue limit (MRL) at the recommended dosage [19,20]. However, weed resistance to these four herbicides increased gradually due to widespread use in paddy fields, which inevitably increased the dosage of these four paddy herbicides [22][23][24][25]. ...
Article
Full-text available
The dissipation dynamics and residue of pyrazosulfuron-ethyl, bensulfuron-methyl, acetochlor, and butachlor in paddy fields at Good Agricultural Practices (GAP) condition were carefully investigated in this study. The four herbicides’ residues were determined based on a quick, easy, cheap, rugged, safe (QuEChERS) method coupled with HPLC-MS/MS and GC-MS. The limit of detection (LOD) for pyrazosulfuron-ethyl, bensulfuron-methyl, acetochlor, and butachlor in all matrices ranged from 0.04–1.0 ng. The limit of quantification (LOQ) of the four herbicides ranged from 0.01–0.1 mg/kg. Moreover, the average recoveries of the four herbicides ranged from 78.9–108% with relative standard deviations (RSDs) less than 15% at three different fortified levels for different matrices. The dissipation results indicated that the average half-lives (t1/2) of the four herbicides in soil were in the range of 3.5–17.8 days, and more than 95% of the four herbicides dissipated within 5 days in water. Furthermore, the final residues of the four herbicides were all below the LOQ at harvest time. Such results highlight the dissipation dynamics and residue of the four herbicides in a rice cropping system and contribute to risk assessment as well as scientific guidance on the proper and safe application of herbicides in paddy fields.
... Glyphosate was extracted and analysed, as detailed by Brindhavani et al. (2020), using HPLC. The extraction and detection method was validated by performing recovery studies before analysing unknown samples, as described by Janaki et al. (2016) for bensulfuron methyl. ...
Article
Full-text available
While nutrients are administered through various sources and combinations, herbicides are used for pre-plant and pre-emergence weed control in the tomato fields. Therefore, a study was conducted to understand the effects of nutrient fertilisation through inorganic and organic sources along with farmers practice and the application of three herbicides, namely glyphosate, pendimethalin, and metribuzin, individually or in combinations, on earthworm activity in tomato crops. The herbicides caused a significant reduction in earthworm biomass and escapement to the lower 15 cm depth. Application of pendimethalin and metribuzin to soils inorganically fertilised with major and micronutrients resulted in low survival rates and a high ecological risk quotient. The effect was attenuated when farmyard manure was applied. The study suggests that when 2 to 3 herbicides are applied in succession and combined with inorganic fertilisers as nutrient sources, stringent measures like adding organic nutrient sources, applying the correct herbicide combination, etc., must be followed to reduce their toxicity to earthworms. This helps to protect and sustain earthworm activity and biodiversity in the soil.
... PXL and BSM treatments significantly affected the growth of watermelon, squash, and bottle gourd at 10 ppb after one week, suggesting the susceptibility of the cucurbits species to a low concentration of PXL and BSM [3]. They are known to control weeds by inhibiting acetolactate synthase (ALS), an essential enzyme in the biosynthesis of the branched-chain amino acids valine, leucine, and isoleucine [4][5][6][7]. ...
Article
Full-text available
Crop rotation is a very useful technique that allows the production of various crop on a fixed area. However, the use of herbicide in rice rotation has negative impact on the growth and production of watermelon. In this work, we provided evidences to support the use of strigolactone (SL) to alleviate herbicide toxicity. The findings of this work provide valuable information underlying the regulatory effects of SL on herbicide tolerance in Cucurbitaceae crops.
... PXL and BSM treatments significantly affected the growth of watermelon, squash, and bottle gourd at 10 ppb after one week, suggesting the susceptibility of the cucurbits species to a low concentration of PXL and BSM [3]. They are known to control weeds by inhibiting acetolactate synthase (ALS), an essential enzyme in the biosynthesis of the branched-chain amino acids valine, leucine, and isoleucine [4][5][6][7]. ...
Article
Full-text available
Strigolactone (SL) plays essential roles in regulating plant growth, development, and stress response. This study was conducted to evaluate the effect of exogenous SL on watermelon resistance against herbicides penoxsulam (PXL) and bensulfuron-methyl (BSM). These herbicides were found to negatively impact watermelon root growth and photosynthetic pigments, probably due to the ultrastructural damage and cell death in leaf and root tissues under herbicide stresses. The activation of SL-related gene expression suggested that the SL pathway may mitigate herbicide toxicity in watermelon. The exogenous SL dose-dependently reversed the PXL- or BSM-induced antioxidant activity, suggesting that SL may participate in maintaining antioxidant enzyme homeostasis under herbicide stresses. The up-regulation of herbicide metabolization and detoxification-related genes (Cytochrome P450 and Acetolactate Synthase) by exogenous SL also in part explained how this phytohormone alleviates herbicide toxicity in watermelon. Our findings will provide valuable information underlying the regulatory effects of SL on herbicide tolerance in Cucurbitaceae crops.
... Bensulfuron-methyl (BSM), a sulfonylurea herbicide with chemical name, Methyl α-(4,6-dimethoxypyrimidin-2-ylcarbamoylsulfamoyl)-o-toluate, is a white to light yellow, fine crystalline powder. Although applied in low quantities, BSM has high application rates [17]. It targets on weeds like Cyperaceae, Butomaceae and Alismataceae growing in the rice fields [18,19]. ...
Article
Full-text available
Detoxification and management of environmental contaminants is an exigent issue of current times. Sulfonylurea herbicide, Bensulfuron-methyl was investigated for its degradation demeanour in soils, through biotic and abiotic modes (biodegradation and hydrolysis). Solid-liquid extraction of the herbicide was followed by GC-MS and UV-visible spectrophotometry analysis. The main metabolites observed were pyrimidinamine [149 m/z] and benzylsulfonamide [182 m/z]. The rate of biodegradation achieved by Aspergillus niger and Penicillium chrysogenum was 95% and 71%, respectively. The maximal decline in Bensulfuron-methyl concentration through hydrolysis was 48%. Furthermore, hydrolytic elimination was also evaluated based on time and pH. Both these parameters had a strong influence on the rate of transformation. Soils with lower pH exhibited an increased rate of degradation while a temperature of 27±2°C gave ideal conditions for herbicide decomposition. Percentage degradation and rate constant (k) followed first order reaction kinetics. Non-inoculated soils displayed less amounts of degradation. Furthermore, relative standard deviations were calculated for the residuals extracted in all soils. Analysis of variance (ANOVA) provided a p value < 0.05 for both strains with R2 closer to 1 signifying the significance of the results. Both fungal strains proved their potential for Bensulfuron-methyl remediation in soils.
... Bensulfuron-methyl is present in agricultural soil and rice plant after applying herbicides [9] and can remain in the soil for a prolonged period [10]. The chemical may also be converted to bound residues in soil [11,12] and then released causing phytotoxicity. ...
Article
Full-text available
Bensulfuron-methyl is an herbicide widely used for weed control although its residues cause damage to other crops during crop rotations. In this study, the biodegrading activity of bensulfuron-methyl by a plant growth-promoting bacterial strain was carried out. Methylopila sp. DKT isolated from soil was determined for bensulfuron-methyl degradation and phosphate solubilization in the liquid media and soil. Moreover, the effects of the herbicide on peanut development and the role of Methylopila sp. DKT on the growth promotion of peanut were investigated. The results showed that the isolate effectively utilized the compound as a sole carbon source and solubilized low soluble inorganic phosphates. Methylopila sp. DKT also utilized 2-amino-4,6-dimethoxypyrimidine, a metabolite of bensulfuron-methyl degradation, as a sole carbon and energy source, and released ammonium and nitrate. The supplementation with Methylopila sp. DKT in soil increased the peanut biomass and the phosphorus content in the plant. In addition, the inoculation with Methylopila sp. DKT in soil and peanut cultivation increased the bensulfuron-methyl degradation by 57.7% for 1 month, which suggests that both plants and the bacterial isolate play a key role in herbicide degradation. These results indicate that the studied strain has a high potential for soil remediation and peanut growth promotion.
... It can be uptaken by weeds and translocated to target sites to inhibit the acetolactate synthase, block biosynthesis of branched-chain amino acids, and finally result in growth inhibition, chlorosis, and necrosis 3,4 . The solubility of BSM is very low (12-120 mg L −1 in water at pH 6.0-7.0 and 25 °C) 5,6 . It was reported that the herbicide was immobile to moderately mobile in soils and easy to be adsorbed by soil particles [6][7][8] . ...
Article
Full-text available
In this work, the inclusion complexes of hydrophobic herbicide bensulfuron-methyl (BSM) with β-cyclodextrin (β-CD) and (2-hydroxypropyl)-β-CD (2-HP-β-CD) were prepared and characterized. Phase solubility study showed that both β-CD and 2-HP-β-CD increased the solubility of BSM. Three-dimensional structures of the inclusion complexes were simulated by the molecular docking method. The docking results indicated that guest BSM could enter into the cavities of host CDs, folded, and centrally aligned inside the inclusion complexes. The benzene ring of the guest molecule was close to the wide rim of the host molecules; the pyrimidine ring and side chains of the guest molecule were oriented toward the narrow rim of the host molecule. The inclusion complexes were successfully prepared by the coprecipitation method. The physiochemical characterization data of 1H NMR, FT-IR, XRD, and DSC showed that the guest and host molecules were well included. BSM had lower soil adsorption and higher herbicidal activity in the complexation form with β-CD or 2-HP-β-CD than in the pure form. The present study provides an approach to develop a novel CDs-based formulation for hydrophobic herbicides.
... Atrazine residues were detected up to 10 days under zero tillage practice and 30 days under conventional tillage practices. On 0 day, the atrazine residue deposited was three times higher (315-326 μg/kg) in conventional tillage plots than the no-till (118-133 μg/kg) plots (Janaki et al. 2016b) (Table 17). ...
Chapter
Full-text available
Herbicides use has increased to 25% of total pesticides during the last 10 years in India. Currently, the herbicide consumption in Tamil Nadu contributes about 7% of the total market. Among total registered herbicides in India, 12 are prominently used in rice, maize, sugarcane, cotton, pulses, oilseeds, and non-cropped areas. Though herbicides are playing an important role for controlling weeds and increasing food production, their potential harm to human beings, soil health and environment is a cause of concern. We reviewed herbicide residue research in the state of Tamil Nadu and reported in this chapter. In long-term studies involving rice, butachlor, pretilachlor and 2,4-D dissipated by first order kinetics with half-life of 6.0–11.8, 9.9–10.8 and 6.5–11.4 days, respectively. Terminal residues were below detection limit (BDL) or maximum residue limit (MRL) in rice grain. Traizines, dinitroanilines, acetoanilide, imidazolinones, diethyl-ether, pyrimidinyloxybenzoic acid, phenoxy and propionic acid herbicides persisted in soil from 60 to 150 days. Soil half-lives of atrazine, pendimethalin, and metolachlor ranged from 31.8 to 44.9, 12.2 to 30.1, 16.0 to 38.5 days, respectively with persistence order: atrazine > pendimethalin > metolachlor > metribuzin > imazethapyr > alachlor > butachlor > pretilachlor > oxyfluorfen > imazethapyr. The herbicide residues in 99.2% crop samples were below MRL. The residues affected soil microbial populations until 30 days, and the chemical properties like organic C, texture, pH and cation exchange capacity had significant influence on herbicides persistence, mobility and sorption. About 80% residues dissipated under aquatic system on 30th day with <15% fish mortality. It was concluded that contamination of soil, crops and water by herbicides was at very low levels in Tamil Nadu.
Article
Background: Bensulfuron-methyl has recently attracted attention given its widespread use as an herbicide in crops, especially its transdermal safety. However, no dermal toxicity study of this pesticide to mammals has been reported. The present study aims to investigate subacute dermal toxicity of bensulfuron-methyl following repeated doses exposure. Methods: Forty-eight Sprague-Dawley rats were randomly divided into four groups: a normal control group and bensulfuron-methyl groups of different concentrations (250, 500, 1000 mg/kg.bw/day). The rats were topically applied with the substance dermally for 6 h per day for 28 days. At the end of the experiment, all rats were monitored for any changes in their hematological, biochemical parameters, and pathological and histological sections. Results: There were no statistically significant differences (P ≥ 0.05) in the hematological parameters and biochemical parameters. The pathological histological results of rats in the control and the highest concentration group showed no significant abnormalities. The NOAEL of subacute dermal toxicity study was found to be 1000 mg/kg.bw/day in both female and male rats. Conclusion: The result indicated that bensulfuron-methyl is probably safe for humans as a pesticide.
Article
This study examined the response of corn to clomazone, chlorimuron, imazaquin, and imazethapyr the year following their application to soybeans. Herbicides were surface-applied from one-half to three times the labeled application rates. Soybeans were planted the year of application and crop tolerance was evaluated. Corn was planted in rotation the following season. Soybeans were tolerant of all four herbicides. The highest rates of clomazone, chlorimuron, and imazaquin injured the corn early in the season. Imazethapyr did not influence corn growth. Visual estimates of clomazone injury were as high as 39% chlorosis. Seedling dry weight reductions at the highest chlorimuron and imazaquin rates were 32% and 24%, respectively. Although corn was injured by higher rates of clomazone, imazaquin, and chlorimuron at the 3-leaf stage, none of the herbicides significantly reduced grain yield. This study suggests that these herbicides can carry over and injure corn, especially if labeled application rates are exceeded. However, low to moderate early season injury to corn may not affect grain yield.
Article
Acetohydroxyacid synthase was discovered as the site of action of imidazolinone and sulfonylurea herbicides over 6 yr ago. In recent years, advances have been made in the understanding of this enzyme as a herbicide target site. Derivatives of both imidazolinones and sulfonylureas have yielded new herbicide chemistry. AH of the herbicides display unusual “slow-binding” behavior with the enzyme, and this behavior may help explain efficacy of the herbicides. Resistance to these herbicides has been developed through a number of different procedures, and the mechanism of resistance is through changes in sensitivity of the enzyme to the herbicides. The changes are either selective to only one class of chemistry, or broad to a number of classes of chemistry. These data support the idea that binding sites for the herbicides on the enzyme are only partially overlapping. Progress in purification of AHAS from corn includes discovery of the existence of the enzyme in monomer and oligomer aggregation states. The interaction of the enzyme with the herbicides is affected by enzyme aggregation state.
Article
This paper aims to apply a modified QuEChERS method of mertribuzin extraction from sugarcane plant parts, soils. The quantification of metribuzin was carried out by high performance liquid chromatography coupled with diode array detector. The method was validated by evaluating the analytical curve, linearity, limits of detection and limits of quantification, precision (repeatability) and accuracy (recovery), after the optimization of extraction parameters for the determination of metribuzin. The method provided limits of quantification in the range of 0.05-0.10 mg kg–1 for different matrices with the method detection limit of 0.03 mg kg–1 for all matrices. The developed method was applied to study the metribuzin persistence in sugarcane plant parts and soil from the field experimentation which received two different doses of metribuzin along with control. It was found that the residue of metribuzin in soil and sugarcane parts were below the detection limit (0.03 mg kg–1) except at the higher dose of application. - See more at: http://www.asianjournalofchemistry.co.in/user/journal/viewarticle.aspx?ArticleID=27_10_36#sthash.BKHIHuGZ.dpuf
Article
DPX-F5384 is the active ingredient in the rice herbicide Londax®. Growth inhibition promoted by DPX-F5384 in pea roots can be reversed by addition of valine and isoleucine, suggesting inhibition of the pathway for branched-chain amino acid biosynthesis. This compound readily inhibits an enzyme in this pathway, acetolactate synthase. Although acetolactate synthase from rice is also inhibited by DPX-F5384, the rice plant is tolerant due to its ability to metabolically demethylate DPX-F5384 to a relatively weak inhibitor of its acetolactate synthase.
Article
The use of sulfonylurea herbicides is increasing because they are used at very low rates and offer greater weed management flexibility in pre- and post-emergence programs. Information is needed on the off-site movement of these compounds under different tillage systems. Our objective was to evaluate the runoff potential of chlorimuron ethyl [(2-[[[[(4-chloro-6- methoxy-pyrimidin-2-yl] amino]carbonyl]-amino]sulfonyl)benzoic acid)], nicosulfuron [(2[[4,6-dimethoxypyrimidin-2-yl] aminocarbonyl]aminosulfonyl]- N, N-dimethyl-3-pyridinecarboxamide]), and bromide in conventional (CT) and no-tillage (NT) production systems under simulated rainfall. The soil types were a Norfolk sandy loam (fine-loamy, siliceous, thermic Typic Paleudult) at a Coastal Plain location and a Pacolet sandy clay loam (clayey, kaolinitic, thermic Typic Kanhapludult) at a Piedmont locations. The first simulated rainfall event (30 min) consisted of a low (1.27 cm h-1) or high (5.08 cm h-1) rate applied 24 h after herbicide and bromide application. One week later, the high rainfall rate was repeated on all plots. Consistently greater runoff occurred with CT compared to NT. Runoff differences between tillage systems were more pronounced for the second rainfall simulation and also with the Piedmont soil. The average herbicide losses from the initial high rainfall simulation were 1.2 and 2.2% of the total application (0.014 kg ha-1) for CT and NT, respectively, at the Coastal Plain location. Similar results were obtained at the Piedmont location. Mean herbicide losses during the second rainfall simulation represented <0.2% of the applied rate. There were no detectable concentrations of either herbicide found in the sediment phase of runoff. On the basis of these results, it is not expected that chlorimuron or nicosulfuron pose a significant threat to the environment via transport in runoff.
Article
Chlorsulfuron was applied at 35 and 70 g ha-1 to Platner sandy loam and Weld (Aridic Paleustolls) loam, Roxbury (Cumulic Haplustoll) silt loam, and Rosebud (Aridic Argiustolls) sandy loam. Soil pH, organic matter level, number of precipitation events >0.25 cm, and leaching were the major environmental factors influencing chlorosulfuron duration of bioactivity. A chlorsulfuron rate X soil X year interaction occurred, but this interaction was not consistent over all soils. Sunflower Helianthus annuus did not differ from corn Zea mays in sensitivity to chlorsulfuron residues in soil. -from Authors
Article
Mark H. Russell, John L. Saladini and Frank Lichtner of DuPont Crop Protection review some of the benefits and some of the stewardship issues relating to this versatile class of herbicide.
Article
Since their introduction in the early 1980s, there have been a number of field studies conducted to assess the effects of sulfonylurea herbicides on non-target plants (i.e. plants not labeled for use). In these studies a wide variety of plant response assessment techniques have been used to measure effects on non-target plants. This paper examines the relationship of short-term plant response measurements to plant productivity measurements such as yield or quality. Whether short-term plant response measurements have a practical degree of accuracy and precision appropriate for hazard assessment on non-target plants from sulfonylureas is discussed. A comprehensive review of published literature and unpublished field studies of the effects of sulfonylureas on the yield and quality of non-target plant species is reported. When this information is coupled with exposure factors and environmental fate characteristics, the risks to non-target plants from sulfonylureas are similar to those from other herbicides used at higher application rates. © 1998 SCI.