Download as pdf or txt
Download as pdf or txt
You are on page 1of 131

ANALYTICAL ANALYSIS OF HERBICIDE THERMAL STABILITY AND

ENVIRONMENTAL FATE IN SOIL AND MULCH PRODUCTION SYSTEMS

by

KAYLA MARIE EASON

(Under the Direction of TIMOTHY L. GREY)

ABSTRACT

Weed control is essential for viable profits for production agriculture including row

crops, vegetables, and perennial crops. The extent of residual and contact herbicide persistence

in Georgia soils and on plastic mulch has not been fully evaluated in the literature and requires

further research to better understand their long-term effects. Research was conducted to

evaluate indaziflam and flumioxazin soil dissipation over time from pecan groves, determine

the thermal stability and activation energy of indaziflam and flumioxazin, and determine the

concentration of glyphosate, glufosinate, halosulfuron-methyl, S-metolachlor, and acetochlor

wash-off within the transplant hole using a simulated vegetable bed. This research will help

determine the best production practices that correspond to the dissipation of each herbicide for

Georgia producers.

Indaziflam and flumioxazin half-life was equivalent to 78-96 d and 27 d in Georgia

pecan. Both herbicides were determined to have a groundwater ubiquity score that indicated a

low leachability potential. Sorption of indaziflam increased with increasing clay content.

Indaziflam degradation increased with increasing activity or size of the soil microbial
population. There was an inverse correlation between flumioxazin concentration in soil,

rainfall, and solar radiation. There was no direct correlation between flumioxazin

concentration and soil temperature. Indaziflam and flumioxazin displayed varying kinetic

behavior in aqueous solution. Temperature did not influence indaziflam stability. Flumioxazin

was determined to have an activation energy of 58.4 (± 1.2) kJ mol-1.

Using a simulated vegetable bed covered with TIF, after irrigating 0.63-cm less than 2%

of halosulfuron, glufosinate, and glyphosate remained on the surface of the plastic mulch. In

contrast, 91% and 15% of acetochlor remained on the mulch while S-metolachlor remained on

the mulch at 17% and 3% after 0.63-cm and 1.27-cm irrigation volumes. All herbicide

concentration was detected below 1.0 mg ai or ae in the transplant hole area despite irrigation

amount. However, for halosulfuron, glyphosate, and glufosinate, these concentrations were equal

to a 1.3X-8.9X field rate washing into the transplant hole. Acetochlor and S-metolachlor

concentrations in the transplant hole were equivalent to 0.1X-0.7X field rates, respectively. The

order of concentration detected in the transplant hole area was directly related to water-solubility.

INDEX WORDS: Herbicide Dissipation, LC/MS, Thermal Stability, Degradation Kinetics.


ANALYTICAL ANALYSIS OF HERBICIDE THERMAL STABILITY AND

ENVIRONMENTAL FATE IN SOIL AND MULCH PRODUCTION SYSTEMS

by

KAYLA MARIE EASON

B.S., Abraham Baldwin Agricultural College, 2015

M.S., The University of Georgia, 2018

A Dissertation Submitted to the Graduate Faculty of The University of Georgia in Partial

Fulfillment of the Requirements for the Degree

DOCTOR OF PHILOSOPHY

ATHENS, GEORGIA

2021
© 2021

KAYLA MARIE EASON

All Rights Reserved


ANALYTICAL ANALYSIS OF HERBICIDE THERMAL STABILITY AND

ENVIRONMENTAL FATE IN SOIL AND MULCH PRODUCTION SYSTEMS

by

KAYLA MARIE EASON

Major Professor: Timothy Grey


Committee: Stanley Culpepper
Miguel Cabrera
Nicholas Basinger
Timothy Coolong

Electronic Version Approved:

Ron Walcott
Vice Provost for Graduate Education and Dean of the Graduate School
The University of Georgia
May 2021
iv

DEDICATION

In honor of my late grandparents, William & Edith Hoffman –

Your unconditional love & support are the reason I’m here today.

“In that fair homeland,


We'll know no parting,
Beyond the sunset for evermore”
v

ACKNOWLEDGEMENTS

I would like to start by thanking my major professor, Dr. Timothy Grey, for his

unwavering support and guidance. There truly are no words to adequately describe my eternal

gratitude. As my major professor and mentor, he has dedicated countless hours answering any

random question I could think of, consistently reminding me that I do belong here, and

celebrating our wins (no matter how big or small). He taught me what a true mentor and boss

should act like – by being willing to do any job and help in any way possible, from field work to

lab work, no matter the time of day. Dr. Grey pushed me to succeed, allowed me to explore any

avenue of research, and always gave me the tools I needed to complete any task. I am who I am

today, in academics and life, because of who he helped me become. I am excited to be able to

continue working together in the future.

I would like to express my gratitude to my committee members, Dr. Stanley Culpepper,

Dr. Nicholas Basinger, and Dr. Tim Coolong, for providing me with new perspectives,

dedicating their time toward my studies, and mentoring me along the way. I would not have the

skills and mind-set I have today without all of you.

I would like to individually thank another one of my committee members, Dr. Miguel

Cabrera. Throughout my academic career, Dr. Cabrera has been a patient, kind, and steady

support system for me and I consider him to be one of my greatest mentors. I will continue to use

the tools he instilled in me as I move into the next phase of my career and life.

Finally, I would like to thank my fellow graduate students, friends, and family. All of the

long days and late nights would never have been possible without being surrounded by such a
vi

great group of graduate students. To my friends – thank you for all the laughter, love, and

support. To my family – I could never have done this without each and every one of you.

To my parents – we did it!


vii

TABLE OF CONTENTS

Page

ACKNOWLEDGEMENTS......................................................................................................... v

LIST OF TABLES ..................................................................................................................... ix

LIST OF FIGURES ................................................................................................................... xi

CHAPTER

1 INTRODUCTION ..................................................................................................... 1

Residual and Contact Herbicide Use in Georgia ................................................... 1

Overview of Herbicides ....................................................................................... 3

Behavior of Herbicides......................................................................................... 8

Objectives .......................................................................................................... 10

References ......................................................................................................... 12

2 QUANTIFYING INDAZIFLAM SOIL DISSIPATION AND THERMAL

STABILITY USING LC/MS ................................................................................... 17

Abstract ............................................................................................................. 18

Introduction ....................................................................................................... 19

Materials and Methods ....................................................................................... 22

Results ............................................................................................................... 27

Conclusions ....................................................................................................... 33

References ......................................................................................................... 35
viii

3 ASSESSMENT OF FLUMIOXAZIN SOIL BEHAVIOR AND THERMAL

STABILITY IN AQUEOUS SOLUTION ................................................................ 51

Abstract ............................................................................................................. 52

Introduction ....................................................................................................... 53

Materials and Methods ....................................................................................... 54

Results and Discussion ....................................................................................... 61

Conclusions ....................................................................................................... 68

References ......................................................................................................... 70

4 QUANTIFYING HERBICIDE MOVEMENT FROM PLASTIC MULCH INTO

THE PLANT HOLE USING SIMULATED VEGETABLE BEDS .......................... 82

Abstract ............................................................................................................. 83

Introduction ....................................................................................................... 84

Materials and Methods ....................................................................................... 86

Results ............................................................................................................... 91

Discussion.......................................................................................................... 95

References ......................................................................................................... 98

5 CONCLUSIONS ................................................................................................... 113


ix

LIST OF TABLES

Page

Table 1.1: Physiochemical properties of various herbicides ....................................................... 11

Table 2.1: Environmental measures recorded during the course of the Webster County

experiment from the University of Georgia weather station (20 km from site)................ 41

Table 2.2: Environmental measures recorded during the course of the Sumter County experiment

from the University of Georgia weather station (20 km from site) .................................. 42

Table 2.3: ETHOS X microwave extraction program parameters............................................... 43

Table 2.4: Chromatographic and mass spectrometer instrument parameters. .............................. 44

Table 2.5: Parameter estimates for the dissipation of indaziflam in soil from experiments

conducted from 2016-2018 in Webster and Sumter counties, Georgia. ........................... 45

Table 2.6: Parameter estimates for indaziflam stability in aqueous solution at various

temperatures over time when evaluated on a thermal gradient table.. ............................. 46

Table 3.1: Environmental measures recorded during the course of the experiment from the

University of Georgia weather station.. .......................................................................... 76

Table 3.2: Correlation values of flumioxazin concentration over time and environmental

measures recorded during the course of the experiment from the University of Georgia

weather station.. ............................................................................................................. 77

Table 3.3: Parameter estimates for flumioxazin stability in aqueous solution (pH 4.5) at various

temperatures over time when evaluated on a thermal gradient table... ............................ 78

Table 4.1: Detailed descriptions of the seven herbicide treatments... ........................................ 102
x

Table 4.2: Environmental measurements recorded at the time of each herbicide application.... 103

Table 4.3: Herbicide concentration (mg ai or ae m-2 and %) on plastic mulch as affected by

sample timing.... .......................................................................................................... 104

Table 4.4: Concentration (mg ai m-2 or %) of acetochlor and S-metolachlor on plastic mulch as

affected by irrigation amount and sample timing... ....................................................... 105

Table 4.5: Herbicide concentration in the transplant hole area as affected by irrigation

amount......................................................................................................................... 106
xi

LIST OF FIGURES

Page

Figure 2.1: Indaziflam persistence (t = day) in (a) Webster County and (b) Sumter County,

Georgia using the exponential decay equation. ............................................................... 48

Figure 2.2: Indaziflam persistence (t = growing degree day (GDD)) in (a) Webster County and

(b) Sumter County, Georgia using the exponential decay equation. ................................ 49

Figure 2.3: Time course for degradation of indaziflam in solution at various temperatures (20 to

70°C) on a thermal gradient. ........................................................................................... 50

Figure 3.1: Flumioxazin soil persistence (t = day) using the exponential decay equation

combined over four site-years. ....................................................................................... 79

Figure 3.2: Time course for degradation of flumioxazin in solution (pH 4.5) at various

temperatures (11 to 40°C) on a thermal gradient. ............................................................ 80

Figure 3.3: Arrhenius plot of flumioxazin degradation............................................................... 81

Figure 4.1: Simulated bed construction without carpet or plastic mulch covering. ................... 107

Figure 4.2: Bottom view of the constructed simulated bed with 14 glass jars attached. ............ 108

Figure 4.3: The concentration (mg ai or ae) of halosulfuron-methyl, glufosinate, glyphosate, and

acetochlor, and S-metolachlor detected on the plastic surface as affected by sample

timing (before or after irrigation was applied) and irrigation volume: (A) 0.63 cm; (B)

1.27-cm.. ..................................................................................................................... 109

Figure 4.4: The average amount of water (mL) caught in a jar (crop transplant hole) from the

0.63-cm and 1.27-cm irrigation event........................................................................... 110


xii

Figure 4.5: . The concentration (mg ai) of halosulfuron-methyl, glufosinate, glyphosate, and

acetochlor, and S-metolachlor detected in the transplant hole.. ..................................... 111

Figure 4.6: The concentration (mg ai or ae) of halosulfuron-methyl, glufosinate, glyphosate, and

acetochlor, and S-metolachlor detected in the transplant hole area and water (mL) caught

in the transplant hole area at each irrigation amount: (A) 0.63 cm; (B) 1.27-cm.. ......... 112
1

CHAPTER 1

INTRODUCTION

1. RESIDUAL AND CONTACT HERBICIDE USE IN GEORGIA.

Weed control is essential for viable profits for production agriculture including row

crops, vegetables, and perennial, orchard crops. In the southeastern United States, a virtual

plethora of weed species are encountered due to the temperate climate of the region. Pecan

(Carya illinoinensis (Wangenh.) K. Koch) and vegetable cropping systems typically rely heavily

on residual and postemergence (POST) herbicides to provide adequate season-long weed control.

However, the extent of residual and contact herbicide persistence in Georgia soils and on plastic

mulch has not been fully evaluated in the literature and requires further research to better

understand their long-term effects.

1.1. Pecan.

Pecan production continues to increase in the United States as exports to foreign markets

stimulate the crop’s value. Georgia remains the largest pecan producing state in the United

States, accounting for over 30% of all planted hectares (USDA-NASS 2020). It is forecast that

production in Georgia could exceed 133,000 Mg of harvested nuts annually by 2025 with newly

planted trees and renovated groves (Wells 2014).

With increased demand and value, growers in Georgia interplant older groves with new

cultivars. Hurricane Michael caused a significant increase of interplanting trees due to its

devastation of orchards across the highest production area of Southwest Georgia in 2018. This
2

management tactic can expose newly planted trees to herbicides that have been applied in

previous seasons and are currently not labeled for non-bearing trees (Wells 2014).

Weed competition can reduce growth in perennial crops by more than 50% (Wells 2014).

Newly planted trees and bushes are especially sensitive to competition for sunlight, moisture,

and nutrients (Smith 2011). In established perennial crop areas, weeds also serve as inoculum for

diseases and alternate hosts for insects (Faircloth et al. 2007). Establishing weed-free strips by

applying herbicides between perennial species areas can increase survival, water use efficiency,

and growth. Ideally, weeds are controlled in perennial cropping systems during establishment

and throughout the lifespan of the orchard. Herbicide-resistant and herbicide-tolerant weeds,

such as Italian ryegrass (Lolium perenne L. ssp. multiforum (Lam.) Husnot) and Palmer amaranth

(Amaranthus palmeri S. Wats.), infest Georgia pecan orchards, therefore developing a sound

management program while maintaining profitability is essential (Grey et al. 2014).

1.2. Vegetables.

Plasticulture is one management technique used in vegetable production that has gained

favorability among fruiting vegetable growers in Georgia. Plastic mulch systems utilize low-

density polyethylene (LDPE) mulch or totally impermeable films (TIF) stretched over

aggressively tilled soils that have been formed into raised beds that range in width and height

(Lamont 1996; Grey et al. 2007). These systems may be used for 3 to 5 (or more) different

vegetable crops (transplanted and seeded) throughout multiple growing seasons, generally over 2

years. Spreading the cost of production for the equipment, labor, and mulch over multiple crops

improves overall grower sustainability (Nyoike and Liburd 2014; Culpepper et al. 2009).

To maintain the integrity of the mulch intact across the multiple crops, weed suppression

and termination of the previous crop are crucial. Nutsedge species (Cyperus spp.) are especially
3

troublesome for fruiting vegetables grown on mulch, with the potential to cause significant

damage to the bed (Webster 2002). To control nutsedge and other weed species from damaging

the mulch, producers can apply herbicides on top of and under the mulch.

Fomesafen (Potter et al. 2016), oxyfluorfen (Boyd 2015), and sulfentrazone (Grey et al.

2007) are used as soil-applied herbicides for their residual activity. S-metolachlor (Cornelius et

al. 1985), glufosinate (Sharpe and Boyd 2019), and paraquat, glyphosate, and halosulfuron-

methyl (Grey et al. 2009) are all herbicides that could potentially be used on plastic mulch as

over-the-top applications to mitigate weed species populations between crops in vegetable

production. However, herbicide persistence in the soil and on the plastic mulch is a major

concern for growers due to the potential injury of a subsequent crop or transplant. (Grey et al.

2002; Grey et al. 2009).

2. OVERVIEW OF HERBICIDES.

2.1. Indaziflam.

Indaziflam (N-[(1R,2S)-2,3-dihydro-2,6-dimethyl-1H-inden-1-yl]-6-(1-fluoroethyl)-

1,3,5-triazine-2,4-diamine) is a broad-spectrum herbicide that inhibits cellulose biosynthesis and

is commonly used in pecan groves for preemergence control of grass and broadleaf weeds

(Shaner 2014). Indaziflam is a weak acid that dissipates by biotic degradation and leaching and is

applied at relatively low rates, providing good residual activity with its persistent nature

(Gonzalez-Delgado et al. 2015). Indaziflam provides preemergence control for up to 6 months of

control in pistachio (Pistacia vera L.), stone fruit (Prunus armeniaca L. and Prunus domestica

L.), and citrus (Citrus spp.) (Allen 2011). Desorption of the herbicide is reported to be hysteric

for multiple soils, indicating a decrease potential for mobility in the soil profile (Alonso et al.
4

2015). Indaziflam has a reported half-life of 30-150 days (Gonzalez-Delgado and Shukla 2017;

USEPA 2010). With such a wide range in half-life, further research is warranted to develop more

specific values for Georgia.

2.2. Halosulfuron-Methyl.

Halosulfuron-methyl (methyl 3-chloro-5-[[[[(4,6-dimethoxy-2

pyrimidinyl)amino]carbonyl]amino]sulfonyl]-1-methyl-1H-pyrazole-4-carboxylate), is a

sulfonylurea herbicide that inhibits acetolactate synthase (ALS), which in turn inhibits branched-

chain amino acid production (Shaner 2014). Halosulfuron-methyl is known to control Cyperus

spp. in a variety of cropping systems. Halosulfuron when applied to LDPE mulch was found to

have a DT50 of 3 h when irrigated and 18 h in dry situations (Grey et al. 2009). The sulfonylurea

herbicides are all weak acids with relatively low disassociation constants (Shaner 2014). This

group of herbicides is typically applied at low use rates, however they still show residual soil

activity. Sulfonylurea persistence is affected by soil pH, with higher persistence rates with

neutral pH values (Grey and McCullough 2012), similar to the pH values found in commercial

vegetable production beds (Kelley and Boyhan 2017).

2.3. Fomesafen, Flumioxazin, Oxyfluorfen, and Sulfentrazone.

Fomesafen (5-[2-chloro-4-(trifluoromethyl)phenoxy]-N-(methylsulfonyl)-2-

nitrobenzamide) is a protoporphyrinogen oxidase (PPO)-inhibiting diphenylether herbicide. PPO

inhibition leads to the buildup of protoporphyrin IX, resulting in the prevention of properly

functioning chloroplast (Shaner 2014). Fomesafen is the active ingredient (as the sodium salt) in

Reflex, a common vegetable herbicide. Fomesafen has a wide range of rotation restrictions,

including up to an 18-month planting date from the time of application. Fomesafen is used to

control multiple weed species including the suppression of yellow nutsedge (Cyperus esculentus
5

L.). The behavior of fomesafen in soil is environment dependent (Shaner 2014). With a typical

production soil pH, fomesafen would exist in anionic form, becoming more available as soil pH

drops (Weber 1993). The solubility of the formulated sodium salt of fomesafen is 600,000 mg/L,

while the acid is only 50mg/L. It is a weak acid, with a pka of 2.7.

Flumioxazin (2-[7-fluoro-3,4-dihydro-3-oxo-4-(2-propynyl)- 2H-1,4-benzoxazin-6-yl]-

4,5,6,7-tetrahydro-1H-isoindole-1,3(2H)-dione) is N-phenylpththalimide herbicide that can be

applied to control a broad spectrum of weeds in a variety of cropping systems (Grey et al. 2014;

Shaner 2014). Similar to fomesafen, flumioxazin inhibits protoporphyrinogen oxidase, which

disrupts proper chloroplast formation and subsequently causing lipid peroxidation and cell

membrane damage (Moreland 1999). The main physiochemical properties that impact

flumioxazin soil movement and availability include an average half-life of 12-18 days, low water

solubility (1.79 mg L-1), and being non-ionizable (Shaner 2014).

Oxyfluorfen (2-chloro-1-(3-ethoxy-4-nitrophenoxy)-4-(trifluoromethyl) benzene) is

another diphenylether herbicide, similar to fomesafen. Oxyfluorfen is relatively non-ionizable

and therefore unaffected by pH, however, volatility is increased with higher temperatures

(similar to those in Georgia) (Yen et al. 2003; Shaner 2014). The average half-life of oxyfluorfen

can range from 5-58 d, depending on soil type (Shaner 2014). Higher clay content and organic

matter provide more adsorption area for oxyfluorfen.

Sulfentrazone (N-2,4-dichloro-5-[4-difluoromethyl)-4,5-dihydro-3-methyl-5-oxo-1H-

1,2,4-triazo-1-yl]phenyl]methanesulfonamide) is a PPO-inhibitor from the triazinone herbicide

family. Sulfentrazone provides control of both purple (Cyperus rotundus L.) and yellow

(Cyperus esculentus L.) nutsedge (Shaner 2014). Sulfentrazone is a weak acid (pKa of 6.56) and

availability increases as soil conditions become more alkaline (Mueller et al. 2014). It has an
6

estimated half-life of 16 d in Georgia soils and can reach 111 d under controlled conditions

(Grey et al. 2007; Meuller et al. 2014).

2.4. Terbacil.

Terbacil (5-chloro-3-(1,1-dimethylethyl)-6-methyl-2,4-(1H,3H)-pyrimidinedione) is a

photosystem II (PSII) inhibiting herbicide. Terbacil provides partial control of nutsedge and is

primarily used for broadleaf and grass weed control (Shaner 2014). Terbacil is weak base, with a

pKa of 9, therefore the amount of organic matter in the soil can influence its behavior (Rahman

1977). Terbacil has weak absorptivity but still has an average half-life of 120 d (Shaner 2014).

2.5. Acetochlor and S-Metolachlor.

Acetochlor (2-chloro-N-(ethoxymethyl)-N-(2-ethyl-6-methylphenyl)acetamide) and S-

metolachlor (2-chloro-N-(2-ethyl-6-methylphenyl)-N-[(1S)-2-methoxy-1-methylethyl]acetamide)

are chloroacetamide herbicides that inhibit very long chain fatty acid biosynthesis. The

chloroacetamide herbicides are known for controlling yellow nutsedge, annual grassed, and

several broadleaf weeds (mainly small-seeded) (Shaner 2014). S-metolachlor is relatively soluble

in water (480 mg L-1) while acetochlor is comparatively less water soluble (230 mg L -1)

(Wolejko et al. 2017). The chloroacetamide herbicides have a reported half-life of 12 to 25 d

with microbial degradation being the major breakdown pathway (Mueller et al. 1999; Shaner

2014).

2.6. Glyphosate, Glufosinate, and Paraquat.

Glyphosate (N-phosphonomethyl)glycine) and glufosinate (2-amino-4-

(hydroxymethylphosphinyl)butanoic acid) are non-selective, contact herbicides labeled for use in

a wide range of crop and non-crop systems (Anonymous 2020). Glyphosate inhibits the synthase

of enzyme 5-enolpyruvylshikimate-3-phosphate (EPSP), which disrupts the shikimic acid


7

pathway, therefore leading to a reduction in aromatic amino acids and eventual plant death

(Henderson et al. 2010). Glyphosate is relied on as an over-the-top herbicide application for

several vegetable crops grown on plastic mulch, with a required 1.25 cm of water applied, due to

its negligible photodegradation losses and high water solubility allowing for movement off of the

plastic mulch (Grey and Vencill 2011; Culpepper et al. 2009; Shaner 2014). Grey et al. (2009)

found that glyphosate was still available in efficacious amounts on LDPE mulch out to 120 h

after treatment when no water was applied. While glyphosate may be a viable option for control,

growers still have to be conscious of potential injury.

Glufosinate is a glutamine synthetase inhibitor, causing a buildup of ammonium in

tissues and subsequently inhibiting photosynthesis and photorespiration, which leads to plant

death (Sellers et al. 2004). Glufosinate is relatively water soluble and can therefore has potential

for use in plasticulture systems with the proper amount of irrigation before transplant (Isaac et al.

2006; Shaner 2014). However, there is still concern for injury if the herbicide adheres to the

plastic mulch.

Paraquat (1,1’-dimethyl-4,4’-bipyridnium dichloride) is a bipyridylium herbicide that

inhibits photosystem I, which creates reactive and toxic radicals within the plant (Lehoczki et al.

1992). The actual cause of tissue damage in a plant is from a mixture of the rapid cycling

between the paraquat ion and paraquat radical and the large number of electrons flowing through

photosystem I. Symptoms progress to necrosis within one or two days from application. Paraquat

does not stop electron flow within photosystem I, but merely captures electrons and converts

them into free radicals (Hess 2000). Paraquat is currently labeled for a wide variety of crops and

non-agricultural uses for its control of most annual broadleaf weeds and suppression of many

perennial weeds (Anonymous 2019). Another characteristic of paraquat is its rapid adsorption in
8

soil. Its adsorption characteristic allows for wide use as a burndown application or early POST

treatment (Tucker et al. 1969).

3. BEHAVIOR OF HERBICIDES.

Herbicide dissipation varies greatly between soils, compounds, and environments. Soils

high in organic matter tend to decrease herbicide dissipation while increasing adsorption. Soil

moisture near field capacity (33 kpa) shows more rapid dissipation rates. Herbicide dissipation in

dry soils is described by a low k-value while dissipation in wet soil is described by high a k-

value. This k-value, a first-order rate constant, can be used to predict half-life information and

therefore availability of herbicides (Nash 1988).

Microbial degradation is another large factor in understanding herbicide dissipation and

is the major factor in degradation for many herbicides (Gonzalez-Delgado and Shukla 2017;

Shaner 2014). Microbial degradation is influenced by soil temperature, soil moisture, and soil

microbial biomass. As soil temperature and soil moisture decrease (cold, dry soil), microbial

degradation also decreases. As soil temperature and soil moisture increase (very hot, saturated

soil), microbial biomasses cannot complete their breakdown processes which can cease

degradation (Nash 1988; Pertile et al. 2020). When microbial degradation decreases, herbicide

dissipation also decreases (Nash 1988).

The mobility of herbicides in the soil profile can have profound effects on their efficacy,

effectiveness as a preemergent herbicide, or potential to contaminate water sources. The

movement of herbicides in soil is done by two major processes, mass flow and hydrodynamic

dispersion. Water is the main factor in transporting herbicides in soil (Rao et al. 1988).

Physiochemical properties of herbicides typically found in groundwater include solubility of 30


9

mg/L and a half-life (DT50) longer than 30 days (Shaner 2014). The movement of herbicides out

of soil and into surface waters can be described by the relative contamination potential. This is

divided into two major indices, the retardation factor (RF) and the attenuation factor (AF). The

RF is based on the time it takes the herbicide to move past the root zone, while the AF is based

on the amount of pesticide leaching past the root zone. The difference in leaching rates (RF and

AF) can be described by the herbicide’s octanol/water coefficient (Kow) (Roa et al. 1988). The

octanol-water partition coefficient is a way to express lipophilicity of a compound and can be

used to estimate the solubility of compounds. Higher Kow values show more hydrophobicity

while low Kow values are usually indicative that the compound is polar and more lipophilic.

Sorption kinetics (adsorption and desorption) are useful in understanding and quantifying

the biological availability of certain herbicides. The rate of adsorption is directly related to the

amount of clay minerals and organic matter present in the soil. The surfaces of the clay minerals

are negatively charged, therefore exchange cations are present (Hance 1988). While most

herbicides are uncharged, there are numerous and widely used weak acids and weak bases. The

dissociation constant pKa can be used to describe the strength of an acid or base.

Table 1 describes the physiochemical properties of the various herbicides. Halosulfuron

is highly soluble in water and therefore highly mobile. Halosulfuron has exhibited very poor

sorption in soils with organic matter. Oxyfluorfen is in stark contrast to the sulfonylurea

herbicides, being low in water solubility and very strongly adsorbed to soils with organic matter

with desorption being irreversible in some soils (Shaner 2014). Similar to oxyfluorfen,

indaziflam has low water solubility and is strongly sorbed to organic matter and clay content in

soils (Gonzalez-Delgado et al. 2015; Alonso et al. 2015; Scheider et al. 2015). In the middle of

these, sulfentrazone is moderately soluble in water and binds poorly to most soils (Shaner 2014).
10

Ordering the residual herbicides from longest to shortest half-life: indaziflam > terbacil >

fomesafen > halosulfuron-methyl > sulfentrazone > oxyflurofen > S-metolachlor > flumioxazin

> acetochlor. Indaziflam, sulfentrazone, and terbacil are similar in degradation, with the main

pathway being microbial degradation and negligible photodegradation (Santiago-Mora et al.

2005; Shaner 2014). Fomesafen is the contrast of these herbicides, with little degradation

occurring microbially and the major pathway being photodecomposition.

4. OBJECTIVES.

Herbicide dissipation will be evaluated in multiple cropping and plasticulture systems.

Research was conducted to 1) evaluate indaziflam and flumioxazin soil dissipation over time

from pecan groves; 2) determine the thermal stability and activation energy of indaziflam and

flumioxazin; 3) determine the concentration of glyphosate, glufosinate, halosulfuron-methyl, S-

metolachlor, and acetochlor wash-off within the transplant hole using a simulated vegetable bed.

4) determine the fate of fomesafen, oxyfluorfen, sulfentrazone, and terbacil in soil covered by

LDPE mulch and bare soil over time; 5) compare the dissipation of glyphosate, glufosinate,

halosulfuron-methyl, S-metolachlor, and paraquat between plasticulture systems utilizing LDPE

and TIF mulch types. This research will help determine the best production practices that

correspond to the dissipation of each herbicide for Georgia producers.


11

Table 1.1. Physiochemical Properties of Various Herbicides (Shaner 2014).


Average Acid Octanol/Water
Molecular Soil Water Organic
Field Dissociation Partition
Herbicide Weight Sorption Solubility Sorption
Half-Life Constant Coefficient
(g/mol) (Kd) (mg/L) (Koc)
(days) (pKa) (log Kow)

Indaziflam 301.36 150 4.9 to 12.8 3.50 2.8 2.80 434 to 1000

600,000
Fomesafen 438.76 100 - 2.7 2.89 60
(salt)

Flumioxazin 354.34 12 to 18 - Non-ionizable 1.79 2.55 500 to 13000

Oxyfluorfen 361.79 5 to 58 99.4 Non-ionizable 0.116 4.68 12233

Sulfentrazone 387.19 32 < 1.0 6.56 780 0.99 43

Terbacil 216.67 120 - 9.00 710 1.89 55

Halosulfuron-
434.81 51 0.36 to 1.6 3.5 1650 -0.02 124
methyl

S-Metolachlor 283.80 15 to 25 2.16 Non-ionizable 488 794 200

Acetochlor 269.77 12 0.4 to 2.7 Non-ionizable 223 300 314

15700 0.0006 to
Glyphosate 169.07 (acid) > 25 324 to 600 2.6 (acid) 24000
(acid) 0.0017

Glufosinate 181.13 (acid) >7 - <2 1370000 - 100


12

REFERENCES

Allen R (2011) Alion: A new preemergent herbicide for the TNV market. Calif Weed Sci Soc J

7:4-5

Alonso DG, Rubem OS, Hall KE, Koskinen WC, Jamil C, Suresh M (2015) Changes in sorption

of indaziflam and three transformation products in soil with aging. Geoderma 239-

240:250-256

Anonymous (2019) Gramoxone® SL 2.0 product label. Greensboro, NC: Syngenta Crop

Protection LLC

Anonymous (2020) Roundup PowerMax® Herbicide product label. St. Louis, MO: Bayer

CropScience LP

Boyd NS (2015) Evaluation of preemergence herbicides for purple nutsedge (Cyperus rotundus)

control in tomato. Weed Technol 29:480-487

Cornelius AJ, Meggitt WF, Penner D (1985) Metolachlor effects on peanut growth and

development. Peanut Sci 15:57-60

Culpepper AS, Grey TL, Webster TM (2009) Vegetable response to herbicides applied to low-

density polyethylene mulch prior to transplant. Weed Technol 23:444-449

Henderson AM, Gervais JA, Luukinen B, Buhl K, Stone D, Strid A, Cross A, Jenkins J (2010)

Glyphosate technical fact sheet. National Pesticide Information Ceneter, OR State Univ

Ext.

Faircloth WH, Patterson MG, Foshee WG, Nesbitt ML, Goff WD (2007) Comparison of

preemergence and postemergence weed control systems in newly established pecan.

Weed Technol 21:972-976

Gonzalez-Delgado AM, Ashigh J, Shukla MK, Perkins R (2015) Mobility of indaziflam


13

influenced by soil properties in a semi-arid area. PloS One 10(5):1-12

Gonzalez-Delgado AM, Shukla MK (2017) Effect of indaziflam application and soil

manipulations on pecan evapotranspiration and gas exchange parameters. Hort Sci

52:910-915

Grey TL, Bridges DC, NeSmith DS (2002) Transplanted pepper (Capsicum annum) tolerance to

selected herbicides and method of application. J Veg Crop Prod 8:27

Grey TL, Vencill WK, Mantripagada N, Culpepper AS (2007) Residual herbicide dissipation

from soil with low density polyethylene mulch or left bare. Weed Sci 55:638-64

Grey TL, Vencill WK, Webster TM, Culpepper AS (2009) Herbicide dissipation from low

density polyethylene mulch. Weed Sci 57:351-356

Grey TL, Vencill WK (2011) Residual herbicide dissipation in Larramendy M, ed. Vegetable

Production, Herbicides, Theory, and Applications. Croatia: InTech

Grey TL, McCullough PE (2012) Sulfonylurea herbicides’ fate in soil: dissipation, mobility, and

other processes. Weed Technol 26:579-581

Grey TL, Turpin FS, Wells L, Webster TM (2014) A survey of weeds and herbicides in

Georgia pecan. Weed Technol 28: 552-559

Hess FD (2000) Light-Dependent Herbicides: An Overview. Weed Sci 48:160-70

Hance RJ (1988) Adsorption and bioavailability in Grover R, ed. Environmental

Chemistry of Herbicides. Florida: CRC Press pp 1-19

Isaac WA, Brathwaite RA, Cohen JE, Bekele I (2006) Effects of alternative weed management

strategies on Commelina diffusa Burm. Infestations in fairtrade banana in St. Vincent and

the grenadines. Crop Prot 26:1219-1225


14

Kelley WT, Boyhan GE (2017) Lime and Fertilizer Management in Commercial Tomato

Production Handbook. Univ of GA Ext Bull 1312 p 17

Lamont WJ (1996) What are the components of a plasticulture system? Hort Technol 6:150-154

Lehoczki E, Laskay G, Gaal I, and Szigeti Z (1992) Mode of Action of Paraquat in Leaves of

Paraquat-Resistant Conyza Canadensis (L.) Cronq. Plant Cell Env 15: 531-39

Moreland, DE (1999) Biochemical mechanisms of action of herbicides and the impact of

biotechnology on the development of herbicides. J Pestic Sci 24: 299-307

Mueller TC, Shaw DR, Witt WW (1999) Relative dissipation of acetochlor, alachlor,

metolachlor, and SAN 582 from three surface soils. Weed Technol 13:341-346

Mueller TC, Boswell BW, Mueller SS, Steckel LE (2014) Dissipation of fomesafen,

saflufenacil, sulfentrazone, and flumioxazin from a Tennessee soil under field conditions.

Weed Sci 62: 664-671

Nash RG (1988) Dissipation from soil. Pages 131-169 in Grover R, ed. Environmental

Chemistry of Herbicides. Florida: CRC Press

Nyoike TW, Liburd OE (2014) Reusing plastic mulch for a second strawberry crop: effects on

arthropod pests, weeds, diseases, and strawberry yields. FL Ent 97:928-936

Pertile M, Lopes Antunes JE, Araujo FF, Mendes LW, Van den Brink PJ, Araujo ASF (2020)

Responses of soil microbial biomass and enzyme activity to herbicidies imazethapyr and

flumioxazin. Sci Rep 10: 7694

Potter TL, Bosch DD, Strickland TC (2016) Field and laboratory dissipation of the herbicide

fomesafen in the southern Atlantic Coastal Plain (USA). J Agric Food Chem 64:5156-

5163

Rahman A (1977) Persistance of terbacil and trifluralin under different soil and climatic
15

conditions. Weed Res 17:145-152

Rao PSC, Jessup RE, Davidson JM (1988) Mass flow and dispersion. Pages 21-43 in Grover R,

ed. Environmental Chemistry of Herbicides. Florida: CRC Press

Santiago-Mora R, Martin-Laurent F, de Prado R, Franco AR (2005) Degradation of simazine by

microorganisms isolated from soils of Spanish olive fields. Pest Manag Sci 61:917-921

Schneider JG, Haguewood JB, Song E, Pan X, Rutledge JM, Monke BJ, Myers DF, Anderson

SH, Ellersieck MR, Xiong X (2015) Indaziflam effect on bermudagrass (Cynodon

dactylon L. Pers.) shoot growth and root initiation as influenced by soil texture and

organic matter. Crop Sci 55:429-436

Sellers BA, Smeda RJ, Li J (2004) Glutamine synthetase activity and ammonium accumulation

is influenced by time of glufosinate application. Pest Bio Phys 78:9-20

Shaner DL, ed (2014) Herbicide handbook. 10th ed. Lawrence, KS: Weed Science Society of

America

Sharpe SM, Boyd NS (2019) Utility of glufosinate in postemergence row middle weed control in

Florida plasticulture production. Weed Technol 33:495-502

Smith MW (2011) Pecan production increased by larger vegetation-free area surrounding the

tree. Scientia Hort 130:211-213

Tucker BV, Pack DE, Ospenson JN, Omid A, and Thomas WD (1969) Paraquat soil

bonding and plant response. Weed Sci 17:448-51

USDA-NASS (2020) Noncitrus Fruits and Nuts: 2019 Summary. Natl Agr Stat Serv,

Washington DC

USEPA (2010) Pesticide Fact Sheet: Indaziflam.


16

https://www3.epa.gov/pesticides/chem_search/reg_actions/registration/fs_PC-

080818_26-Jul-10.pdf [accessed December 14 2020]

Weber JB (1993) Ionization and sorption of fomesafen and atrazine by soils and soil

constituents. J Pest Sci 39:31-38

Webster TM (2002) Weed survey-southern states: vegetable, fruit and nut crops subsection.

Proc South Weed Sci Soc 55:237-258

Wells L (2014) Pecan planting trends in Georgia. Hort Tech 24:475-479

Wolejko E, Kaczynski P, Lozowicka B, Wydro U, Borusiewicz A, Hrynko I, Konecki R,

Snarska K, Dec D, Malinowski P (2017) Dissipation of S-metolachlor in plant and soil

and effect on enzymatic activities. Environ Monit Assess 189:355

Yen J, Sheu W, Wang Y (2003) Dissipation of the herbicide oxyfluorfen in subtropical soils and

its potential to contaminate groundwater. Ecotox Environ Safe 54:151-156


17

CHAPTER 2

QUANTIFYING INDAZIFLAM SOIL DISSIPATION AND THERMAL STABILITY USING

LC/MS 1

________________________________________________________
1
Eason, K. M., Basinger, N. T., Cabrera, M. L., Rucker, K. S., and Grey, T. L. Accepted to Pest
Management Science, 4/6/2021.
18

ABSTRACT

BACKGROUND: As a nonselective cellulose biosynthesis inhibitor, indaziflam has a niche for

broad-spectrum weed control with long residual activity in various perennial cropping systems.

The extent of indaziflam soil persistence and chemical behavior at various temperatures has not

been fully evaluated, therefore the objectives of these experiments were: (i) quantify indaziflam

soil persistence in two common Georgia soils and (ii) evaluate indaziflam molecular stability as

affected by temperature and time.

RESULTS: Indaziflam soil dissipation followed first-order kinetics and was adequately

described by the exponential decay equation. Indaziflam half-life in Greenville sandy clay loam

and Faceville loamy sand were 96 and 78 days or 770 and 693 growing degree days,

respectively. Indaziflam half-life and soil clay content (%) had a direct relationship, while

indaziflam half-life and microbial biomass had an inverse relationship. GUS index ranging from

1.89-1.98 were determined. Aqueous solutions containing 0.33 µmol L-1 of indaziflam were

exposed to temperatures that ranged from 20 to 70°C for up to 672 hours, with results indicating

that temperature had no influence on indaziflam molecular stability over time.

CONCLUSIONS: Sorption of indaziflam increased with increasing clay content. Indaziflam

degradation increased with increasing activity or size of the soil microbial population. The GUS

index values ranked indaziflam as a low to transitional leacher. There was no indaziflam

degradation by hydrolysis at any temperature. Relating the indaziflam stability laboratory

experiments to the field dissipation studies indicates that indaziflam persistence in soil over time

is not influenced by changes in soil temperature.

Keywords: crop protection; environmental impact; half-life; herbicide; LC/MS; perennial crops;

soil behavior
19

1 INTRODUCTION

Indaziflam (N-[(1R,2S)-2,3-dihydro-2,6-dimethyl-1H-inden-1-yl]-6-(1-fluoroethyl)-

1,3,5-triazine-2,4-diamine) is a preemergent, broad-spectrum herbicide that is classified as a

cellulose biosynthesis inhibitor.1,2 Indaziflam controls multiple broadleaf and annual grass weeds

in perennial crops, turfgrass, forestry sites, along with non-crop, residential and non-residential

areas.3 Indaziflam is currently registered for use across the United States (US) in multiple

cropping systems including citrus (Citrus L.), apple (Malus Mill.), pear (Pyrus L.), stone (Prunus

L.) and pome (Punica L.) fruits, grape (Vitis L.), olive (Olea europaea L. ssp. Europaea),

pistachio (Pistacia vera L.), almond (Prunus dulcis (Mill.) D. A. Webb.), walnut (Juglans regia

L.), pecan (Carya illinoinensis (Wangenh.) K. Koch), blueberries (Vaccinium spp.), blackberries

(Rubus spp.), and bermudagrass (Cynodon dactylon (L.) Pers.).4,5,6 The use of indaziflam in

various permanent cropping systems provides unique long-term control with no reported

resistant weeds.7,2

Indaziflam has a niche in Georgia for broad-spectrum control of grass and broadleaf

weeds in various permanent tree cropping systems, such as olive and pecan, to maintain bare

ground around the tree rows.8,4 Interest in olive production as a specialty crop in the southeastern

US has increased with demand for locally produced virgin olive oil. Indaziflam is currently

registered for olive trees that have been established for at least 3 years. 9,10 Grey et al.9 reported

that olive trees in Georgia were tolerant of multiple applications of indaziflam over time,

exhibiting no injury symptoms in the form of leaf chlorosis or necrosis.

Similar to olive, pecan production continues to increase in the US as exports to foreign

markets stimulate the crop’s value. Pecan production in Georgia increased from 46,266 MT in

201111,12 to 61,235 MT in 202013 resulting in a 32% increase13,11. Georgia remains the largest
20

pecan producing state in the US, accounting for over 30% of all planted hectares. 14 Weed

competition can reduce growth in perennial crops by more than 50%. 15 Newly planted trees and

bushes are especially sensitive to competition for sunlight, moisture, and nutrients. 16 Weeds are

controlled in perennial cropping systems during establishment and throughout the life-span of the

orchard. In established perennial crop areas, weeds can interfere with cultural practices5,

irrigation equipment 17, harvesting operations18, as well as serve as inoculum for diseases and

alternate hosts for insects19. Weed control programs for perennial crops typically rely on residual

herbicides to provide adequate season-long control by establishing weed-free strips between

perennial species areas.16,19

The most common and troublesome weeds for Georgia pecan include wild radish

(Raphanus raphanistrum L.), Palmer amaranth (Amaranthus palmeri S. Wats.), Italian ryegrass

(Lolium perenne L. ssp. multiflorum (Lam.) Husnot), bermudagrass, crabgrass spp. (Digitaria

spp.), bahiagrass (Paspalum notatum Fluegge), Florida pusley (Richardia scabra L.), purslane

spp. (Portulaca spp.), morningglory spp. (Ipomoae spp.), and curly dock (Rumex crispus L.).

Herbicide-resistant and herbicide-tolerant weeds infest Georgia pecan orchards as they are

widespread in this region and have been reported by growers. 12 Indaziflam controls a variety of

these weed species, including Italian ryegrass, crabgrass, and Amaranthus species along with

other weeds common across the region including annual bluegrass (Poa annua L.), goosegrass

(Eleusine indica L.), and cutleaf evening primrose (Oenothera laciniate Hill).9,20 Previous

research reports residual control from indaziflam for 28 weeks in turfgrass 21,22, 3 to 4 months in

Florida citrus23, and up to 6 months in pistachio, pome fruit, and stone fruit 8. In California

orchards and vineyards, chemical weed management programs that relied on indaziflam as a

preemergence (PRE) herbicide application provided adequate residual control, including control
21

of glyphosate resistant weeds.5,24,9 Weed control in perennial crops is a year-round management

concern for states from Georgia to California.5

In 2012 indaziflam, as Alion™ (Bayer CropScience, Research Triangle Park, NC), was

registered for use in pecan orchards in the US. Several months after application, growers in New

Mexico and Arizona began to report varying herbicide injury symptoms on trees where

indaziflam was applied.3 Gonzalez-Delgado et al.3 attributed the indaziflam injury symptoms to

variations in the soil physical properties and estimated that pecan roots were being exposed to

indaziflam concentrations above than the recommended field rate. Gonzalez-Delgado et al.25

further attributed the injury to the movement of indaziflam with water, consistent with the flood-

irrigation practices used in those states. Sparks26 also reported that pecan trees are highly

sensitive to poor soil drainage and their productivity decreases in areas under prolonged flooding

conditions. In Georgia, where flood-irrigation is not used as a mechanism of irrigation, contrary

findings to Gonzalez-Delgado et al.3 were noted.4 Grey et al.4 reported that newly planted pecan

trees, watered through microjet irrigation, were tolerant to indaziflam when applied as a spring

application over time, which has also been observed in other indaziflam-treated perennial crops,

such as olive9 and woody ornamentals27.

Herbicide dissipation varies greatly among soils, compounds, and environments.

Microbial degradation is a large factor in understanding herbicide dissipation and is the major

mechanism of degradation for many herbicides, including indaziflam. 1,10,28 Microbial

degradation is influenced by soil temperature, moisture, and microbial biomass. As soil

temperature and moisture decrease below certain levels (cold, dry soil), microbial degradation

virtually stops.29 As soil temperatures and moisture increase above certain levels (very hot,

saturated soil), microbial biomasses cannot complete their breakdown processes, also ceasing
22

degradation.29 Sorption kinetics are also useful in understanding and quantifying the biological

availability of certain herbicides. The main physiochemical properties that impact indaziflam soil

movement and availability are the dissociation constant (3.5; weak acid) 3, an average half-life of

30-150 days10,28 (>2 years in cold, dry areas of the Midwestern US30), and low water solubility1.

Sorption of indaziflam is moderate for most soil types but increases as soil organic matter

increases, showing a positive correlation between sorption and organic matter. 30 Desorption of

indaziflam is known to show hysteresis, indicating a decreased potential for mobility in the soil

profile.31 While indaziflam is relatively immobile, the long residual half-life and moderate

adsorption properties suggests that indaziflam could move throughout the soil profile. 32

Even though indaziflam is already a commonly used residual herbicide in Georgia pecan

orchards for preemergence control of grass and broadleaf weeds1, the extent of indaziflam soil

persistence and behavior at various temperatures has not been fully evaluated. Therefore, the

objectives of these experiments were: (i) quantify indaziflam soil persistence in Georgia soils

from pecan orchards and (ii) evaluate indaziflam stability as affected by temperature and time

using laboratory techniques.

2 MATERIALS AND METHODS

2.1 Field experiment

Experiments were conducted at two pecan orchards in Georgia, located in Webster

(31°56’34” N 84°34’23” W) and Sumter (31°56’54” N 83°58’52” W) counties with the cultivars

Byrd and Pawnee, respectively. The Webster county soil was Greenville sandy clay loam (fine,

kaolinitic, and thermic Rhodic Kandiudults) with 61.9% sand, 16.1% silt, 22.0% clay, 2.1%

organic matter, and pH 5.3 (water:soil, 2:1). The soil at Sumter county location was Faceville
23

loamy sand (fine, kaolinitic, and thermic Typic Kandiudults) with 77.9% sand, 16.1% silt, 6.0%

clay, 3.9% organic matter, and pH 5.9 (water:soil, 2:1). The soil microbial biomass (microbial

biomass carbon based on 2% of the soil carbon levels33,34) was 266 and 686 mg C kg -1 for

Webster and Sumter county, respectively. The Webster county trial began in April 2016 and

concluded 603 days after application while the Sumter county trial began in March 2017 and

concluded 478 days after application (Tables 1 and 2). Indaziflam was applied (0.25 µmol L -1)

using a four nozzle (1.8 m) tractor-mounted boom, calibrated to deliver 140 L ha -1 to either side

of the tree row. All management practices were common to all trees during the growing seasons

and directed by the growers.

Soil sampling and preparation procedures outlined by Grey et al.35 were followed. One

soil core was taken from each replication at each sample date (3 soil samples per location per

sample date). Soil cores were collected using an aluminum cylinder (7.62 cm diameter x 7.62 cm

height) that was hammered into the soil until flush with the soil surface. The cylinder was

removed from the soil and the contents were individually wrapped in aluminum foil. Individual

samples were placed into a sealable plastic bag, which was subsequently placed into cold storage

for transportation and then frozen at -10°C until analysis.

Rainfall, solar radiation, soil temperature (10 cm) and daily maximum and minimum air

temperature data were collected at a University of Georgia Weather Monitoring Network station,

located 20 km from the sites, and were used to calculate growing degree day (GDD) for both

locations (Table 1 and 2). GDD accumulation was calculated according to the following

equation:

tn = Σ[((Tmax + Tmin)/2) – Tb] (1)


24

where tn is the sum of GDDs (GDD), Tmax is the daily maximum air temperature (°C), T min is the

daily minimum air temperature (°C), and T b is the base temperature for pecan4 (15.5°C). GDDs

were calculated only for the months when trees were actively growing with leaves (April-

October) to give a more biologically accurate measure of tree growth.

2.2 Thermal gradient table experiment

To further understand the implications of long-term residual herbicides (such as

indaziflam), stability as affected by temperature and time was evaluated using a thermal gradient

table. The thermal gradient table36 is constructed from solid aluminum blocks (2.4 m x 0.9 m x

7.6 cm) with a warming or cooling unit (Anova Model A40, Anova Industries Inc., Stafford, TX)

on each side, pushing a 1:10 ethylene glycol:water solution across the table at a rate of 3.8 L min -
1
. Thermocouples (Omega Engineering, Stamford, CT) are within 5 mm of the table surface in

10-cm intervals, recording temperatures every 30 min using a Graphtec data logger

(MicroDAQ.com Ltd., Contoocook, NH).

A solution of indaziflam (0.33 µmol L -1) was prepared using formulated product (Alion®

Bayer CropScience LP, Research Triangle Park, NC) in HPLC water, followed by stirring with a

stir bar until all indaziflam was dissolved. This solution was transferred to flasks (100 mL flask -
1
) that were sealed with parafilm and placed onto the thermal gradient table. Flasks were placed

on cells across 11 temperatures, ranging from 20 to 70°C with three replications per temperature,

and then the experiment was repeated in time. Duplicate 1-mL aliquots were transferred from the

flasks to HPLC vials at 0, 1, 6, 12, 24, 48, 72, 96, 168, 240, 336, 408, 504, 576, 672 hr after trial

initiation. These vials were then placed into storage and frozen immediately after sampling at -

10°C until analysis.


25

2.3 Analysis

Soil samples were prepared for extraction by allowing them to acclimate to room

temperature, sieving to remove foreign material (>2 mm), and then weighing out 15 g into a

microwave glass tube (Milestone Srl, Sorisole, Italy). Then 30 mL of an 80:20 (v:v)

acetonitrile:water solution was added to each tube, which was then capped with a rubber stopper

and vigorously shaken on a vortex mixer to break up any soil aggregates. The rubber stoppers

were removed and glass tubes were placed into slots evenly spaced on the microwave carousel

assembly. Samples were subjected to microwave assisted extraction3 using an ETHOS X

(Milestone Srl, Sorisole, Italy) 37 following the parameters listed in Table 3. After microwave

extraction, a 1.5-mL supernatant aliquot was pipetted into a 2.0 mL microcentrifuge tubes

(Fisher Scientific International, Waltham, MA). This was then centrifuged for 5 min (12500

RPM) using an Eppendorf MiniSpin® (Eppendorf AG, Germany). Then 1.0 ml of the aliquot

from each sample was transferred into HPLC vials (Fisher Scientific International, Waltham,

MA) for analysis.

All samples were analyzed using a Waters Acquity Ultra-High Performance Liquid

Chromatography (LC) coupled with a Waters 2998 PDA and Waters QDa Mass Spectrometry

(MS) Detector (LC/MS) (Waters Corporation, Milford, MA). The LC separation was performed

on a C18 reversed-phase column (Symmetry C18, 4.6 x 75 mm, 3.5 µm). A water and methanol

gradient was used (Table 4). Table 4 summarizes the LC/MS instrument parameters used for

quantification. Indaziflam amounts (µmol L -1) were quantified by correlating peak area detected

to those of analytical grade standard solutions (Fisher Scientific International, Waltham, MA) of

various known concentrations (0.1 to 1.0 µmol L -1 ).


26

Soil dissipation data were analyzed separately for each experiment, to account for trees at

different locations having varying planting dates and multiple measures taken over time.

Indaziflam dissipation data from the field experiment were subjected to regression analysis using

SAS nonlinear regression (SAS Institute, Cary NC) to determine whether the response could be

described by the following exponential decay equation36:

y = B0e-B1(t) (2)

where y is the measured indaziflam concentration (µmol L -1), B0 is the initial indaziflam

concentration (µmol L-1) when time t is 0, B1 is the dissipation rate of indaziflam (slope), and t is

time or energy elapsed after indaziflam application (day or GDD). After regression against time

or energy, the output included the first-order dissipation rate constant (k)38. Indaziflam field

persistence (DT50) was then determined using the half-life equation36:

DT50 = ln(2)/k (3)

where DT50 is the half-life of indaziflam (day or GDD) and k is the first-order dissipation rate

constant. Data were then graphed in SigmaPlot 14.0 (Systat Software, San Jose, CA). After

determining the net effect of degradation, the groundwater ubiquity score (GUS) index39 was

used to assess the leaching potential of indaziflam in the given soil types by combining the

calculated persistence and mobility measurements40:

GUS = log10(DT50)*(4-log10(Koc)) (4)

where DT50 is the half-life in days (determined from equation 3) and the Koc for indaziflam1 is

approximately 1000 mL g -1. A GUS index score of <1.8 represents a non-leaching herbicide, 1.8-

2.8 is a transitional herbicide that may leach, and >2.8 is considered a leaching herbicide. 39

Samples from the thermal gradient table were analyzed for indaziflam concentrations

over temperature and time. Data were subjected to linear regression (PROC REG) using SAS 9.4
27

to determine if there were any effects of temperature over time on indaziflam concentration. Data

were then graphed in SigmaPlot 14.0.

3 RESULTS AND DISCUSSION

3.1 Field experiment

Indaziflam soil dissipation data were not combined across locations due to differing

planting dates of trees and therefore are presented by location. Overall, the exponential decay

equation (eq. 2) adequately described indaziflam dissipation at both locations, with R 2 values of

0.56-0.62. Table 5 describes the parameters of indaziflam dissipation in soil at both locations,

including first-order dissipation rate constants (k), half-lives (DT50), and R2 values.

Indaziflam dissipation reported in Figures 1 and 2 is for soil 0-7.62 cm deep, which

accounts for all of the indaziflam applied. 41 Half-life (DT50) values for the Webster and Sumter

county location were 96 d (Figure 1a) and 78 d (Figure 1b), respectively (Table 5). Gonzalez-

Delgado et al.25 reported indaziflam half-life in a sandy loam soil field study of 53-63 days (0-15

cm of soil) and consistently detected 135 days after application, while Golzalez-Delgado and

Shukla41 reported half-life values of 63-99 days in greenhouse experiments using the same soil

type. Lower half-life values from field studies than those from greenhouse studies could be the

result of greater degradation from microbial biomass under field conditions. 41 Most of the

determined indaziflam half-lives 25,41 in field research are lower than those reported by the

registrant of 150 days.10 This difference suggests indaziflam dissipation is influenced by soil

type, experimental and environmental conditions, and movement of herbicide with irrigation or

rainfall.
28

Indaziflam typically persists in soil for a longer time period when compared to other

residual herbicides. Pendimethalin is recommended for use with indaziflam in sequential

application herbicide programs for California tree nut crops. 5 Pendimethalin has a much lower

half-life value, 24 days, than indaziflam.42 In a similar Georgia soil (Tifton loamy sand),

halosulfuron and flumioxazin soil dissipation were rapid with an average half-life of 9 days35 and

13-14 days43, respectively.

Indaziflam dissipation presented as growing degree days (GDDs) also allows for a more

biologically accurate measure of time when the trees are actively growing (April-October).36

Therefore, half-life (DT50) values for the Webster and Sumter county location were 770 GDDs

(Figure 2a) and 693 GDDs (Figure 2b), respectively (Table 5). Indaziflam half-life and the

cumulative number of GDDs at each location had a direct relationship indicating longer

persistence in warmer years. Atrazine, which has an inverse half-life to GDD relationship (base

temperature of 10°C), indicated faster dissipation rates in warmer years, which was attributed to

increased microbial degradation.40,44 This would also be expected with indaziflam, due to

microbial metabolism being a major degradation pathway1, however our results did not correlate

with an inverse half-life to GDD relationship. This anomaly may be attributed to the GDDs only

slightly varying between locations and other differing soil properties, including the stark contrast

in microbial biomass between the two soils, with Sumter county having 158% greater microbial

biomass (mg kg-1) than Webster county.

Microbial activity varies greatly for many herbicides in the soil. Indaziflam is not

expected to negatively impact the population or activity of soil microorganisms, since it is a

cellulose biosynthesis inhibitor.3 Microbial biomass and indaziflam half-life had an inverse

relationship, indicating that indaziflam degradation was dependent on the population size or
29

activity of the soil microorganisms. Walker et al. 44 reported that alachlor degradation in the soil

is also positively correlated to microbial biomass.

For this research, no trees died during these experiments (2016-2018). According to Jhal

et al.45, rainfall was sufficient in these studies (Table 1 and 2) to leach indaziflam into pecan tree

rooting zones, similar to rainfall amounts reported by Grey et al. 4, where no visual injury

symptoms on newly planted pecan trees was noted. This is supported by Gonzalez-Delgado and

Shukla46,41, who noted no phytotoxic injury on greenhouse-grown trees from rates as high as 150

g ai ha-1. Coupled together, these data do not help to explain the injury symptoms reported by

Gonzalez-Delgado et al.3 however organic matter content for these soils (0.60%) was much

lower when compared to our Georgia soils (2.1-3.9%).

Gonzalez-Delgado et al.3, Alonso et al.31, and Schneider et al.47 all reported a positive

correlation between indaziflam sorption and organic matter content, which indicates increased

sorption of indaziflam with increasing organic matter content. Similarly, sulfonylurea herbicides

are noted to have increasing herbicide sorption with increasing soil organic matter content. 48

Indaziflam sorption and half-life also have a positive relationship.3 Jones et al.49 noted a decrease

in indaziflam injury to bermudagrass with increasing soil organic matter content. Gonzalez-

Delgado et al.25, indicated a higher mass recovery (%) of indaziflam in areas with higher organic

matter content and reduced weed control (%) in areas with lower organic matter.

Our results indicated an inverse relationship between organic matter content and

indaziflam half-life. Alachlor half-life and soil organic matter is also reported to be negatively

correlated.44 While the organic matter content was greater in Sumter county (3.9%) than in

Webster county (2.1%), clay content was 73% lower and microbial biomass was 158% greater in

Sumter county. The discrepancy between our results and previous literature 3,31 could be
30

explained by these other factors influencing indaziflam sorption and degradation, which may

have affected the lower half-life determined for Sumter county.

Dissipation data from Sumter and Webster county indicated that indaziflam half-life and

clay content (%) were directly related. This is supported by Alonso et al. 50, who noted a positive

correlation between sorption of indaziflam and clay content, and Gonzalez-Delgado et al.3 who

noted faster dissipation rates of indaziflam for soils with lower clay content. Schneider et al. 47

reported decreased indaziflam injury to hybrid bermudagrass with increasing clay content and

more injury when applied to sandy soil. Similarly, atrazine is known to remain in soils with

greater clay content for a longer-period.40 Flumioxazin half-life in a Greenville sandy clay loam

soil, with higher clay content, was greater (16-18 days) when compared to a Tifton loamy sand

soil (13-14 days).43

Gonzalez-Delgado et al.3 reported a more rapid dissipation rate of indaziflam in soil with

higher sand content (77%) and drainage capacity. Similarly, indaziflam dissipation was faster in

Sumter county, which had higher sand content (78%) than Webster county (62%). Jefferies and

Gannon32 reported that sandy soils may allow for downward movement of indaziflam in the

profile and plant injury may increase with increasing soil coarseness. Soil tillage practices

influence soil hydraulic properties46, with higher macropore flow expected in no-tillage

situations51 which is the most common practice for Georgia pecan orchards. Macropores play an

important role in the movement and distribution of water through the soil profile. 51 Faster

movement of solutes and water in soil is influenced by macropores that promote their

preferential flow. Increased preferential flow could enhance deeper than expected herbicide

leaching, ultimately leading to increased herbicide injury. 41


31

Groundwater ubiquity scores (GUS)39 ranging from 1.89-1.98 were determined for

Webster and Sumter county. These values indicate that indaziflam has a transitional potential to

leach in Greenville sandy clay loam (61.9% sand, 16.1% silt, 22.0% clay, and 2.1% organic

matter) and Faceville loamy sand (77.9% sand, 16.1% silt, 6.0% clay, and 3.9% organic matter)

soils. Alonso et al.50 ranked indaziflam as transitional to leacher with GUS index values of 1.84-

3.00 in oxisol and 2.66-2.83 in mollisol soils from Brazil and the US. These authors used the

published DT50 value for indaziflam of 150 days10 to determine the GUS index values, which

could be the explanation for our GUS index values being lower, due to our determined half-lives

being less than 150 days. Jhala and Singh52 reported no leaching of indaziflam beyond 30 cm,

further indicating limited mobility of the herbicide. Similar to indaziflam, other commonly used

residual herbicides in permeant cropping systems are ranked as having a low to transitional

potential to leach. Halosulfuron-methyl, rimsulfuron, and diuron were ranked as transitional

leachers (GUS index of 2.25-2.6553), while flazasulfuron and flumioxazin were ranked as low

leachability (GUS index of 1.31-1.5553).

3.2 Thermal gradient table experiment

There is no previous information in the literature explaining the influence of temperature

on indaziflam degradation, so the focus of this research was to evaluate the stability of

indaziflam at various temperatures over time. Figure 3 represents the time course for degradation

of indaziflam in solution at various temperatures (20 to 70°C) on a thermal gradient. Table 6

describes the parameters of indaziflam stability evaluated on a thermal gradient table, including

degradation rate constants (k), half-lives (DT50), and R2 values. The parameter estimates did not

differ between time and indaziflam concentration across all temperatures (p=0.4563).
32

Indaziflam is known to be stable in anerobic aquatic environments10, however there was

no information available on that stability at increasing temperatures. Data from the thermal

gradient table indicated that indaziflam is hydrolytic stable at various temperatures (20 to 70°C)

(Figure 3). Consequently, half-life information could not be determined for any temperature.

With the magnitude of the activation energy, corresponding to the effect of temperature on

hydrolysis rate constants36, these data indicated that the activation energy of indaziflam could not

be described using the Arrhenius equation (Table 6).

Comparing indaziflam to other commonly used residual herbicides in pecan shows

varying results. Similar to indaziflam, pendimethalin is known to be stable in aqueous solutions

when subjected to temperatures up to 120°C54 with an average field half-life of 24-34 days55. In

contrast, halosulfuron-methyl is influenced by temperature over time, with a reported activation

energy of 86 kJ mol-1 and a half-life of 2.6 days36,56, while another sulfonylurea herbicide,

rimsulfuron, has a reported activation energy of 92 kJ mol-1 with an average half-life of 0.06 to

12 days57,58 depending on the pH of the solution (2 to 8.5). Diuron is reported to have an

activation energy of 127 kJ mol-1 with a half-life of 4 months59. The trends for halosulfuron-

methyl, rimsulfuron, and diruon indicate that as temperature increase, degradation also increases.

One of the most commonly persistent pesticides, DDT (1,1,1-trichloro-2,2’bis(p-

chlorophenyl) ethane), is typically stable in aqueous solutions and has a half-life ranging from 2

to 15 years.60 While DDT is stable in aqueous solutions, Qin et al. 60 was able to determine an

activation energy of 72.3 kJ mol-1 using a peroxymonosulfate activated by cobalt (PMS/CO(II))

system, which can also be used in the degradation of 2,4-D. Similar methods could be explored

for use in solution with indaziflam to possibly quantitate its activation energy.
33

4 CONCLUSIONS

Despite wide use across the US in various permanent cropping systems, there is still little

information available on the parameters affecting the soil behavior of indaziflam. In conclusion,

indaziflam soil dissipation followed first-order kinetics and was adequately described by the

exponential decay equation at both locations. Indaziflam half-life in Greenville sandy clay loam

(61.9% sand, 16.1% silt, 22.0% clay, 2.1% organic matter, and pH 5.3) and Faceville loamy sand

(77.9% sand, 16.1% silt, 6.0% clay, 3.9% organic matter, and pH 5.9) were 96 and 78 days,

respectively. Quantifying pesticide dissipation in terms of time after application in days may not

provide the most accurate model outputs when indaziflam is used in permanent crop production,

rather GDDs may be a more logical method of estimating the half-life of indaziflam in soil, as it

can be quantified over time from weather stations. Indaziflam half-life values for Webster and

Sumter county were 770 and 693 GDDs, respectively. Indaziflam dissipation presented as

growing degree days also allows for a more biologically accurate measure of time when the trees

are actively growing (April-October).

The combination of environmental and soil properties creates a dynamic system greatly

impacting herbicide behavior and persistence. Indaziflam half-life and clay content (%) had a

direct relationship, indicating sorption of indaziflam increases with increasing clay content.

Indaziflam half-life and microbial biomass had an inverse relationship, indicating increased

degradation of indaziflam with increasing activity or population size soil microorganisms.

Groundwater ubiquity scores ranging from 1.89-1.98 were determined for indaziflam, with these

values indicating a transitional potential to leach. As indaziflam persistence increases, leaching

potential could also increase.


34

When solutions containing 0.33 µmol L-1 of indaziflam were exposed to temperatures

that ranged from 20 to 70°C for up to 672 hours, data indicated that temperature had no influence

on molecular stability over time. There was no indaziflam degradation by hydrolysis at any

temperature (Figure 3). Relating indaziflam molecular stability laboratory experiments to the

field dissipation studies indicates that indaziflam persistence in soil over time is not directly

influenced by diurnal and seasonal changes in soil temperature (Tables 1 and 2). However,

indaziflam dissipation did slightly slow down between approximately 240 to 340 days after

application (Figure 1 and 2) which corresponded to decreasing soil temperatures (Table 1 and 2),

indicating that changes in soil temperature directly affected microbial activity which in turn

affected indaziflam dissipation. This is a unique and important aspect as there have been no

previous reports relating how temperature is not a direct factor concerning indaziflam

degradation. This has many positive implications for providing extended residual weed control

as indaziflam is used in multiple facets including perennial crops, forages, industrial sites, and

turf.

As a nonselective cellulose biosynthesis inhibitor, indaziflam has a niche for broad-

spectrum weed control with long residual activity in Georgia pecan orchards to maintain bare

ground in the tree row. Since Webster and Sumter counties located in the largest pecan growing

region of Georgia, most growers can expect their indaziflam to provide long-term residual weed

control with a half-life of 78-96 days. Georgia growers utilizing indaziflam in their residual

weed control programs can now understand the longevity of this chemistry in their specific soil

types.
35

REFERENCES

1 Shaner DL (ed.), Herbicide handbook, 10th edition. Weed Science Society of America,

Lawrence, KS, p. 513 (2014)

2 Brabham C, Lei L, Gu Y, Stork J, Barrett M, DeBolt S, Indaziflam herbicidal action: a potent

cellulose biosynthesis inhibitor. Plant Phys 166:1177-1185 (2014)

3 Gonzalez-Delgado AM, Ashigh J, Shukla MK, Perkins R, Mobility of indaziflam influenced

by soil properties in a semi-arid area. PLOS ONE 10:1-12 (2015)

4 Grey TL, Rucker K, Wells L, Luo X, Response of young pecan trees to repeated applications

of indaziflam and halosulfuron. Hort Sci 53:313-317 (2018)

5 Brunharo CACG, Watkins S, Hanson BD, Season-long weed control with sequential herbicide

programs in California tree nut crops. Weed Tech 34:834-842 (2020)

6 Grey TL, Hurdle NL, Rucker K, Basinger NT, Blueberry and blackberry are tolerant to

repeated indaziflam applications. Weed Tech 35:Accepted (2021)

7 Heap I, The International Herbicide-Resistant Weed Database. www.weedscience.org

[accessed September 24 2020]

8 Allen R, Alion: A new preemergent herbicide for the TNV market. Calif Weed Sci Soc J 7:4-5

(2011)

9 Grey TL, Rucker K, Webster TM, Luo X, High-density plantings of olive trees are tolerant to

repeated applications of indaziflam. Weed Sci 64:766-771 (2016)

10 USEPA, Pesticide Fact Sheet: Indaziflam.

https://www3.epa.gov/pesticides/chem_search/reg_actions/registration/fs_PC-

080818_26-Jul-10.pdf [accessed September 24 2020] (2010)

11 USDA-NASS, Crop Production. Natl Agr Stat Serv, Washington DC (2012)


36

12 Grey TL, Turpin FS, Wells L, Webster TM, A survey of weeds and herbicides in Georgia

pecan. Weed Tech 28:552-559 (2014)

13 USDA-NASS, Pecan Production. Natl Agr Stat Serv, Washington DC (2020)

14 USDA-NASS, Noncitrus Fruits and Nuts: 2019 Summary. Natl Agr Stat Serv, Washington

DC (2020)

15 Wells L, Pecan planting trends in Georgia. Hort Tech 24:475-479 (2014)

16 Smith MW, Pecan production increased by larger vegetation-free area surrounding the tree.

Scientia Hort 130:211-213 (2011)

17 Belding RD, Majek BA, Lokaj GRW, Hammerstedt J, Ayen AO, Orchard floor management

influence on summer annual weeds and young peach tree performance. Weed Tech

18:215-222 (2004)

18 Company RSI, Gradziel TM, Almonds: botany, production, and uses. CABI, Boston, MA. P

510 (2017)

19 Faircloth WH, Patterson MG, Foshee WG, Nesbitt ML, Goff WD, Comparison of

preemergence and postemergence weed control systems in newly established pecan.

Weed Tech 21:972-976 (2007)

20 Brosnan JT, McCullough PE, Breeden GK, Smooth crabgrass control with indaziflam at

various spring timings. Weed Tech 25:363-366 (2011)

21 Perry DH, McElroy JS, Doroh MC, Walker RH, Indaziflam utilization for control of

problematic turfgrass weeds. Appl Turf Sci (2011)

22 McCullough PE, Yu J, de Barreda D, Efficacy of preemergence herbicides for controlling

dinitroaniline-resistant goosegrass in Georgia. Weed Tech 27:639-644 (2013)


37

23 Singh M, Ramirez AM, Edenfield M, Indaziflam: a new preemergence herbicide for citrus, in

Proc of the 51st Weed Sci Soc of America Ann Conf, Oregon, p. 21 (2011)

24 Jhala AJ, Hanson BD, Summer weed control with glyphosate tank-mixed with indaziflam or

penoxsulam in California orchards and vineyards, in Proc of the 51st Weed Sci Soc of

America Ann Conf, Oregon, p. 21 (2011)

25 Gonzalez-Delgado AM, Shukla MK, Ashigh J, Perkins R, Effect of application rate and

irrigation on the movement and dissipation of indaziflam. J Env Sci 51:111-119 (2016)

26 Sparks D, Adaptability of pecan as a species. Hort Sci 40:1175-1189 (2005)

27 Parker A, Myers D, Weed control and ornamental tolerance with indaziflam. Proc South

Weed Sci Soc 64:103 (2011)

28 Gonzalez-Delgado AM, Shukla MK, Effect of indaziflam application and soil manipulations

on pecan evapotranspiration and gas exchange parameters. Hort Sci 52:910-915 (2017)

29 Nash RG, Dissipation from soil, in Environmental Chemistry of Herbicides, Florida: CRC

Press, p 131-169 (2016)

30 Sebastian DJ, Fleming MB, Patterson EL, Sebastian JR, Nissen SJ, Indaziflam: a new

cellulose-biosynthesis-inhibiting herbicide provides long-term of invasive winter annual

grasses. Pest Manag Sci 73:2149-2162 (2017)

31 Alonso DG, Rubem OS, Hall KE, Koskinen WC, Jamil C, Suresh M, Changes in sorption of

indaziflam and three transformation products in soil with aging. Geoderma 239-240:250-

256 (2015)

32 Jefferies MD, Gannon TW, Soil organic matter content and volumetric water content affect

indaziflam soil bioavailability. Weed Sci 64:1-12 (2016)


38

33 Sparling GP, The soil biomass, in Soil Organic Matter and Biological Activity, Dordrecht, p

223 (1985)

34 McGonigle TP, Turner WG, Grasslands and croplands have different microbial biomass

carbon levels per unit of soil organic carbon. Agric 57:1-8 (2017)

35 Grey TL, Vencill WK, Mantripagada N, Culpepper AS, Residual herbicide dissipation from

soil covered with low-density polyethylene mulch or left bare. Weed Sci 55:638-643

(2007)

36 Grey TL, Culpepper AS, Li X, and Vencill WK, Halosulfuron-methyl degradation from the

surface of low-density polyethylene mulch using analytical and bioassay techniques.

Weed Sci 66:15-24 (2018)

37 Mao X, Wan T, Yan A, Shen M, Wei Y, Simultaneous determination of organophosphorus,

organochlorine, pyrethroid and carbamate pesticides in Radix astragali by microwave-

assisted extraction/dispersive-solid phase extraction coupled with GS-MS. Talanta

97:131-141 (2012)

38 Ohmes GA, Mueller TC, Hayes RM, Sulfentrazone dissipation in a Tennessee soil. Weed

Tech 14:100-105 (2000)

39 Gustafson DI, Groundwater ubiquity score: a simple method for assessing pesticide

leachability. Env Toxic Chem 8:339-357 (1989)

40 Clay SA, Dowdy RH, Lamb JA, Anderson JL, Lowery B, Knighton RE, Clay DE, Herbicide

movement and dissipation at four midwestern sites. J Env Sci Health 35:259-278

41 Gonzalez-Delgado AM, Shukla MK, Mobility, degradation, and uptake of indaziflam under

greenhouse conditions. Hort Sci 55:1216-1221 (2020)


39

42 Kocarek M, Artikov H, Vorisek K, Boruvka L, Pendimethalin degradation in soil and its

interactions with soil microorganisms. Soil Water Res 226:1-7 (2015-SWR)

43 Ferrell JA, Vencill WK, Flumioxazin soil persistence and mineralization in laboratory

experiments. J Agric Food Chem 51:4719-4721 (2003)

44 Walker A, Moon YH, Welch SJ, Influence of temperature, soil moisture, and soil

characteristics on the persistence of alachlor. Pest Sci 35:109-116 (1992)

45 Jhala AJ, Ramirez AHM, Leaching of indaziflam applied at two rates under different rainfall

situations in florida candler soil. Bull Env Cont Toxicol 88:326-332 (2012)

46 Gonzalez-Delgado AM, Shukla MK, Effect of indaziflam application and soil manipulations

on pecan evapotranspiration and gas exchange parameters. Hort Sci 52:910-915 (2017)

47 Schneider JG, Haguewood JB, Song E, Pan X, Rutledge JM, Monke BJ, Myers DF, Anderson

SH, Ellersieck MR, Xiong X, Indaziflam effect on bermudagrass (Cynodon dactylon L.

Pers.) shoot growth and root initiation as influenced by soil texture and organic matter.

Crop Sci 55:429-436 (2015)

48 Grey TL, McCullough PE, Sulfonylurea herbicides’ fate in soil: dissipation, mobility, and

other processes. Weed Tech 26:579-581 (2012)

49 Jones PA, Brosnan JT, Kopsell GK, Breeden GK, Soil type and rooting depth affect hybrid

bermudagrass injury with preemergence herbicide. Crop Sci 53:660-665 (2013)

50 Alonso DG, Koskinen WC, Oliveira Jr RS, Constantin J, Mislankar S, Sorption-desorption of

indaziflam in selected agricultural soils. J Agric Food Chem 59:13096-13101 (2011)

51 Shipilato MJ, Dick WA, Edwards WM, Conservation tillage and macropore factors that affect

water movement and the fate of chemicals. Soil Tillage Res 53:167-183 (2000)
40

52 Jhala AJ, Singh M, Leaching of indaziflam compared with residual herbicides commonly

used in Florida citrus. Weed Tech 26:602-607 (2012)

53 Lewis KA, Tzilivakis J, Warner D, Green A, An international database for pesticide risk

assessments and management. Human Eco Risk Assess: An Int J 22:1050-1064 (2016)

54 European Food Safety Authority, Conclusion on the peer review of the pesticide risk

assessment of the active substance pendimethalin. ESFA J 14:12-69 (2016)

55 Kocarek M, Karel V, Boruvka L, Artikov H, Pendimethalin degradation in soil and its

interaction with soil microorganisms. Soil Water Res 11:1-7 (2016)

56 Sarmah AK, Sabadie J, Hydrolysis of sulfonylurea herbicides in soils and aqueous solutions:

a review. J Agric Food Chem 50:6253-6265 (2002)

57 Dinelli G, Vicari A, Bonetti A, Catizone P, Hydrolytic dissipation of four sulfonylurea

herbicides. J Agric Food Chem 45:1940-1945 (1997)

58 Vicari A, Zimdhal RL, Cranmer BK, Dinelli G, Primisulfuron and rimsulfuron degradation in

aqueous solution and adsorption in six Colorado soils. Weed Sci 44:672-677 (1996)

59 Salvestrini S, Di Cerbo P, Capasso S, Kinetics of the chemical degradation of diuron.

Chemosphere 48:69-73 (2002)

60 Qin W, Fang G, Wang Y, Wu T, Zhu C, Efficient transformation of DDT by

peroxymonosulfate activated with cobalt in aqueous systems: kinectics, products, and

reactive species identification. Chemosphere 148:68-76 (2016)


41

Tables

Table 2.1. Environmental measures† recorded during the course of the Webster County

experiment from the University of Georgia weather station (20 km from site).

Sample Date DAA‡ R (mm) GDD ST (°C; 10 cm)


4/19/2016 0 0 2 23
4/22/2016 3 20 14 22
4/26/2016 7 20 36 27
5/5/2016 16 23 97 24
5/13/2016 24 23 142 29
5/19/2016 30 33 178 26
6/1/2016 43 43 284 34
7/7/2016 79 146 698 33
8/3/2016 106 204 1032 29
9/23/2016 157 359 1609 31
12/12/2016 237 479 1856 14
3/14/2017 329 897 1856 11
3/27/2017 342 898 1856 23
4/11/2017 357 994 1886 24
5/15/2017 391 1093 2075 29
6/19/2017 426 1293 2376 30
7/21/2017 459 1475 2724 33
8/29/2017 498 1569 3155 31
12/13/2017 603 1890 3624 9

Cumulative values are reported from initial herbicide application date for DAA, R, SR, and
GDD. ST values are reported for each sample date.

Abbreviations: DAA, days after application; R, rainfall; GDD, growing degree days; ST, soil
temperature (10 cm).
42

Table 2.2. Environmental measures† recorded during the course of the Sumter County

experiment from the University of Georgia weather station (20 km from site).

Sample Date DAA‡ R (mm) GDD ST (°C; 10 cm)


3/6/2017 0 0 0 17
3/9/2017 3 4 0 18
3/13/2017 7 20 0 11
3/20/2017 14 22 0 17
4/4/2017 29 52 20 22
5/4/2017 45 181 164 24
5/24/2017 65 301 296 24
6/19/2017 90 418 520 30
7/21/2017 122 600 869 33
8/22/2017 154 694 1225 33
12/13/2017 267 1015 1768 9
2/28/2018 344 1203 1842 18
3/7/2018 351 1228 1848 15
3/20/2018 364 1264 1863 21
4/26/2018 397 1421 1914 21
5/24/2018 428 1522 2105 27
7/13/2018 478 1785 2641 34

Cumulative values are reported from initial herbicide application date for DAA, R, SR, and
GDD. ST values are reported for each sample date.

Abbreviations: DAA, days after application; R, rainfall; GDD, growing degree days; ST, soil
temperature (10 cm).
43

Table 2.3. ETHOS X microwave extraction program parameters.

Step Time (min) Temperature (°C) Power Limit (W) Comments


1 10 60 < 350 Ramp from ambient to 60 °C
2 10 60 < 350 Maintain at 60 °C
3 5 - - Cool down to ambient
44

Table 2.4. Chromatographic and mass spectrometer instrument parameters.

Instrumentation Waters ACQUITY Arc LC system


Waters QDa MS system
Waters Empower 3 data software
Column Symmetry C18
4.6 x 75 mm, 3.5 µm
Column Temperature ambient
Injection Volume 50 µL
Run Time 7.0 min
Mobile Phase A – LC/MS water + 0.05% formic acid
B – LC/MS methanol + 0.05% formic acid
Flow Rate 0.8 mL min-1
Gradient Time (min) A (%) B (%)
0.00 60 40
0.08 60 40
0.40 35 65
1.59 30 70
2.38 20 80
2.58 5 95
3.93 5 95
3.97 60 40
Scan Type Multiple Reaction Monitoring
Polarity Positive
Cone Voltage 17 V
45

Table 2.5. Parameter estimates for the dissipation of indaziflam in soil from experiments

conducted from 2016-2018 in Webster and Sumter counties, Georgia.

In Time (days) Cumulative Radiation (GDDs)


2
Location k †
DT50 ‡ R k DT50 R2
Webster 0.0072 96 0.62 0.0009 770 0.62
Sumter 0.0089 78 0.56 0.0010 693 0.56

First-order dissipation rate constants (k) were calculated by nonlinear regression of the
herbicide quantity with respect to time and solar radiation.

Half-lives (DT50) were determined using the k-values and the equation: DT50 = ln(2)/k.
46

Table 2.6. Parameter estimates for indaziflam stability in aqueous solution at various

temperatures over time when evaluated on a thermal gradient table. †

Temperature (°C) k‡ Standard Error (±) DT50§


20 2.1 x 10-5 A 5.2 x 10-6 stable
-5 -6
25 1.7 x 10 A 5.7 x 10 stable
-5 -6
30 2.2 x 10 A 5.7 x 10 stable
-5 -6
35 2.1 x 10 A 5.2 x 10 stable
-6
40 1.6 x 10-5 A 5.9 x 10 stable
-5 -6
45 2.1 x 10 A 5.4 x 10 stable
-5 -6
50 1.6 x 10 A 5.6 x 10 stable
-5 -6
55 1.7 x 10 A 5.3 x 10 stable
-5 -6
60 1.9 x 10 A 5.4 x 10 stable
-5 -6
65 1.8 x 10 A 5.4 x 10 stable
-5 -6
70 1.6 x 10 A 5.6 x 10 stable

Each value is the average of three replicates per experiment, with experiments conducted
twice over time and combined for presentation. Values for each rate constant within a column
followed by the same letter are not significantly different at p<0.05 probability level.

Degradation rate constants (k) were calculated by linear regression of the herbicide quantity
with respect to time (0 to 672 hr).
§
Half-lives (DT50) were determined using the k-values and the equation: DT50 = ln(2)/k.
47

Figure Legend

Figure 2.1. Indaziflam persistence (t = day) in (a) Webster County and (b) Sumter County,

Georgia using the exponential decay equation. Non-linear regression was applied. Model shows

that data can be described by first-order kinetics. Lines represent the first-order regression

equation. Data points indicate the means of three replications. Error bars represent the standard

error of each mean (SEM). Parameter estimates: (a) Webster County: y = 0.2040e(-0.0072*t), R2:

0.62; (b) Sumter County: y = 0.1806e(-0.0089*t), R2: 0.56.

Figure 2.2. Indaziflam persistence (t = growing degree day (GDD)) in (a) Webster County and

(b) Sumter County, Georgia using the exponential decay equation. Non-linear regression was

applied. Model shows that data can be described by first-order kinetics. Lines represent the first-

order regression equation. Data points indicate the means of three replications. Error bars

represent the standard error of each mean (SEM). Parameter estimates: (a) Webster County: y =

0.2010e(-0.0009*t), R2: 0.62; (b) Sumter County: y = 0.1548e(-0.0010*t), R2: 0.56.

Figure 2.3. Time course for degradation of indaziflam in solution at various temperatures (20 to

70°C) on a thermal gradient. Data points indicate the means of three replications, with

experiments conducted twice and combined. Error bars represent the standard error of each mean

(SEM).
48

Figure 2.1:

0.3
(A)
Indaziflam Concentration (µmol L-1)

0.2

0.1

0.0
0 100 200 300 400 500 600
Days After Treatment (days)

0.3
(B)
Indaziflam Concentration (µmol L-1)

0.2

0.1

0.0
0 100 200 300 400 500
Days After Treatment (days)
49

Figure 2.2:

0.3
(A)
Indaziflam Concentration (µmol L-1)

0.2

0.1

0.0
0 1000 2000 3000 4000
Growing Degree Days (Tbase = 15.5°C)

0.3
(B)
Indaziflam Concentration (µmol L-1)

0.2

0.1

0.0
0 1000 2000 3000
Growing Degree Days (Tbase = 15.5°C)
50

Figure 2.3:

0.35
Indaziflam Concentration (µmol L-1)

0.30

20°C 35°C 50°C 65°C


25°C 40°C 55°C 70°C
30°C 45°C 60°C

0.25
0 1 6 12 24 48 72 96 168 240 336 408 504 576 672
Incubation Time (hour)
51

CHAPTER 3

ASSESSMENT OF FLUMIOXAZIN SOIL BEHAVIOR AND THERMAL STABILITY IN

AQUEOUS SOLUTION 1

________________________________________________________
1
Eason, K. M., Basinger, N. T., Cabrera, M. L., and Grey, T. L. To be submitted to
Chemosphere.
52

ABSTRACT

Flumioxazin is a preemergence, N-phenylpththalimide herbicide that can be applied to

control a broad spectrum of weeds in a variety of cropping systems. Currently, limited

information exists concerning the environmental fate of flumioxazin, therefore the present

studies investigated the kinetic behavior of flumioxazin in soil and aqueous solution using field

and analytical techniques. Flumioxazin half-life in a Greenville sandy clay loam and Faceville

loamy sand was 26.6 d. Flumioxazin was determined to have a groundwater ubiquity score of

1.79, indicating a low leachability potential. There was an inverse correlation between

flumioxazin concentration in soil, rainfall, and solar radiation. There was no direct correlation

between flumioxazin concentration and soil temperature. Flumioxazin was determined to have an

activation energy of 58.4 (± 1.2) kJ mol-1 and a Q10 value of 2.2. Even at the lowest amount of

solar radiation and soil temperature, the energy from these environmental measures exceeded the

activation energy needed for flumioxazin degradation. Flumioxazin stability in solution and field

dissipation indicate that, with the input of thermal energy, degradation can be rapid.

Keywords: Flumioxazin, Soil, Environmental Behaviors, Thermal Degradation.


53

1. Introduction

Georgia remains the largest pecan (Carya illinoinensis (Wangenh.) K. Koch) producing

state in the US, accounting for over 30% of all planted hectares (USDA-NASS, 2020). Newly

planted trees and bushes are especially sensitive to competition for sunlight, moisture, and

nutrients (Smith, 2011). In established perennial crop areas, weeds can interfere with cultural

practices, irrigation equipment, harvesting operations, as well as serve as inoculum for diseases

and alternate hosts for insects (Belding et al, 2004; Faircloth et al., 2007; Company and Gradziel,

2017). Weed competition can reduce growth in perennial crops by more than 50% and are

controlled during establishment and throughout the life-span of the orchard (Wells. 2014). Weed

control programs for perennial crops typically rely on residual herbicides to provide adequate

season-long control by establishing weed-free strips between perennial species areas (Faircloth et

al., 2007; Smith, 2011).

Herbicide-resistant and herbicide-tolerant weeds infest Georgia pecan orchards, therefore

identifying new ways to control weeds and utilizing herbicides products so that growers can

maintain profitability is essential (Grey et al. 2014). WSSA group 14 herbicides are essential in

management of glyphosate-resistant weeds. Flumioxazin (2-[7-fluoro-3,4-dihydro-3-oxo-4-(2-

propynyl)- 2H-1,4-benzoxazin-6-yl]-4,5,6,7-tetrahydro-1H-isoindole-1,3(2H)-dione) is a

preemergence, WSSA group 14, N-phenylpththalimide herbicide that can be applied to control a

broad spectrum of weeds in a variety of cropping systems, such as peanut, plasticulture

strawberries, and orchard crops (Boyd et al., 2020; Hurdle et al., 2020; Grey et al., 2014; Shaner,

2014). Flumioxazin inhibits protoporphyrinogen oxidase, which disrupts proper chloroplast

formation and subsequently causing lipid peroxidation and cell membrane damage (Moreland,

1999; Matringe et al., 1989). The main physiochemical properties that impact flumioxazin soil
54

movement and availability include an average half-life of 12-18 days, low water solubility (1.79

mg L-1), and it is non-ionizable (Shaner, 2014).

Herbicide dissipation varies greatly among soils, compounds, and environments.

Microbial degradation is a large factor in understanding herbicide dissipation and is the major

mechanism of degradation for many herbicides and is influenced by soil temperature, moisture,

and microbial biomass (Shaner, 2014; Soulas and Lagacherie, 2001). Sorption kinetics are also

useful in understanding and quantifying the biological availability of certain herbicides.

Currently, limited information exists concerning the environmental fate of flumioxazin.

The present studies investigated the kinetic behavior of flumioxazin in soil and aqueous

solution using field and analytical techniques. The main objectives of these experiments were to:

(1) quantify flumioxazin soil persistence in Georgia soils, (2) determine the temperature effect of

flumioxazin in aqueous solution, and (3) determine the activation energy of flumioxazin.

2. Materials and Methods

2.1. Chemicals

Water (H20; Optima™ for HPLC), water (Optima™ LC-MS Grade), acetonitrile

(CH3CN; Optima™ for HPLC; > 99.0% purity), acetonitrile (Optima™ LC-MS Grade; ≥ 99.9%

purity), O-Phosphoric Acid (≥ 85% purity), and formic acid (Optima™ LC-MS Grade; ≥ 99.0%

purity) were purchased from Fisher Chemical™ (Fisher Scientific, Pittsburgh, PA 15275).

PESTANAL™ analytical standard of flumioxazin (≥ 95.7% purity) was purchased from Sigma-

Aldrich (St. Louis, MO 63103). Chateau® herbicide (water dispersible granule; 51%

flumioxazin) was obtained from Valent (Valent ®, Walnut Creek, CA 94596). Standard buffer
55

solutions used to calibrate the pH meter electrode were reagent grade or better and obtained from

Fisher Scientific.

2.2. Field Experimental Procedure

Field experiments were conducted at two pecan orchards in Webster (31°56’34” N

84°34’23” W) and Sumter (31°56’54” N 83°58’52” W) counties Georgia, with the cultivars Byrd

and Pawnee, respectively. Plots (3m x 15m) at each location consisted of 5 trees with three

replications total. Rainfall (mm), solar radiation (MJ/m2), and soil temperature (°C at 10 cm)

data, throughout the entirety of the experiments, were collected at a University of Georgia

Weather Monitoring Network station located 20 km from the field sites (Table 1).

The Webster county soil was Greenville sandy clay loam (fine, kaolinitic, and thermic

Rhodic Kandiudults) with 61.9% sand, 16.1% silt, 22.0% clay, 2.1% organic matter, bulk density

of 1.63 g/cm3, and pH 5.3 (water:soil, 2:1). The soil at Sumter county location was Faceville

loamy sand (fine, kaolinitic, and thermic Typic Kandiudults) with 77.9% sand, 16.1% silt, 6.0%

clay, 3.9% organic matter, bulk density of 1.54 g/cm3, and pH 5.9 (water:soil, 2:1). The soil

microbial biomass (microbial biomass carbon based on 2% of the soil carbon levels (McGonigle

and Turner, 2017; Sparling, 1985)) was 266 and 686 mg C kg -1 for Webster and Sumter counties,

respectively.

The Webster county trial began in 2016, with the first application made on 19 April 2016

and the second application made on 21 March 2017 (Table 1). The Sumter county trial began in

2017, with the first application made on 6 March 2017 and the second application made on 27

February 2018 (Table 1). Flumioxazin was applied at 0.22 kg ai ha -1 (3320 µmol L-1) using a

four nozzle (1.8 m) tractor-mounted boom (AIXR11002 nozzles), calibrated to deliver 189 L ha-1

to either side of the tree row. Tractor speed was maintained at 4.8 kph, with the boom held 45.7
56

cm above the soil surface. All management practices were common to all trees during the

growing seasons and directed by the growers.

One soil core was taken from each replication at each sample date (3 soil samples per

location per sample date). Soil cores were collected using an aluminum cylinder (7.62 cm

diameter x 7.62 cm height) that was hammered into the soil until flush with the soil surface. The

cylinder was removed from the soil and the contents were individually wrapped in aluminum

foil. Individual samples were placed into a sealable plastic bag, which was subsequently placed

into cold storage for transportation and then frozen at -10°C until analysis. Field methodologies

follow those described in Mueller and Senseman (2015) and the soil sampling and preparation

procedures used were outlined by Grey et al. (2018).

2.2.1. Herbicide Extraction

Methods for herbicide extraction were adapted from previous literature (Eskilsson and

Bjorklund, 2000; Grahovac et al., 2017; Mueller et al., 2014; Vryzas and Papadopoulou-

Mourkidou, 2002). Microwave-assisted extraction (MAE) using an ETHOS X (Milestone Srl)

was used to extract flumioxazin from the respective soil samples. MAE was used for its

significant advantages of reducing organic solvent consumption and extraction time when

compared to traditional extraction techniques (Eskilsson and Bjorklund, 2000; Mueller et al.,

2014). This MAE method utilized acetonitrile as the microwave-absorbing solvent within the

closed vessels.

Soil samples were acclimated to room temperature before processing for extraction.

Samples were homogenized and sieved (2 mm) before weighing out 15 g of each samples into

glass tubes (Milestone Srl, Sorisole, Italy). A volume of 30mL of an acetonitrile:water solution

(90:10 v/v) was added to each glass tube and then capped with a rubber stopper. Tubes were
57

shaken vigorously to break up any soil aggregates and then evenly placed across the microwave

carousel with the rotor mechanism turned on to continuously turn the samples during extraction.

MAE parameters were set to a 10-min ramp up to 60°C, held for 10 min at 60°C, and then

subjected to a cooling period of at least 5 min (or until samples reach ambient temperature). The

power limit was set to < 350 W. Following MAE, 1.5 mL of the supernatant was pipetted into

microcentrifuge tubes (Fisher Scientific) and centrifuged using an Eppendorf MiniSpin

(Eppendorf AG, Hamburg) for 5 min at 12,500 RPM. Once completed, an aliquot from each

sample was transferred into HPLC vials (Fisher Scientific) for analysis.

2.3. Thermal Stability Experiments

O-phosphoric acid was added to HPLC grade water in 1 µL-increments until a pH of 4.0

was achieved. The pH values of all solutions were monitored and confirmed using a

Fisherbrand™ accumet™ AB150 pH meter (Fisher Scientific). A solution of flumioxazin (0.28

µmol L-1) was prepared using formulated product (Chateau®, 51% ai) in the HPLC-grade water

(pH 4.5). The stock solution was mixed using a stirring bar until all flumioxazin was dissolved.

Concentration of the stock solution was verified by correlating peak area detected to those of

analytical grade standard solutions at known concentrations.

Flasks (250 mL) were individually filled with 100 mL of the aqueous flumioxazin

solution, sealed with parafilm, and placed onto the thermal gradient table. Specifically, flasks

were placed on cells across 7 temperatures, ranging from 11 to 40°C, with four replications per

temperature. The experiment was repeated in time. Duplicate 1 mL-aliquots were transferred

from the flasks to HPLC vials at 0, 1, 6, 12, 24, 48, 72, 96, 168, 240, and 336 h after trial

initiation. After sampling, these vials were immediately placed into storage and frozen at -10°C

until analysis.
58

The thermal gradient table is constructed from solid aluminum blocks (2.4 m x 0.9 m x

7.6 cm) with a warming or cooling unit (Anova Model A40, Anova Industries Inc., Stafford, TX)

on each side, pushing a 1:10 ethylene glycol:water solution across the table at a rate of 3.8 L min-
1
(Grey et al., 2018). Thermocouples (Omega Engineering, Stamford, CT) are within 5 mm of the

table surface in 10-cm intervals, recording temperatures every 30 min using a Graphtec data

logger (MicroDAQ.com Ltd., Contoocook, NH).

2.4. Analytical Method

Flumioxazin concentration was analyzed by a Waters Acquity ultra-high performance

liquid chromatography (UHPLC) coupled with a Waters 2998 PDA and Waters QDa mass

spectrometry (MS) detector (Waters Corporation, Milford, MA). The LC separation was

performed on a Symmetry C18 reverse-phase column (4.6 mm x 75 mm, 3.5 µm; Waters

Corporation, Milford, MA) sustained at ambient temperature. The mobile phase was acetonitrile

+ 0.1% formic acid (A) and HPLC grade water + 0.1% formic acid (B). Formic acid was used to

maintain ion strength. The mobile phase was held isocratic at 70% A and 30% B at a flow rate of

0.75 mL min-1 for 2.7 min total. The injection volume was 10 µL. The MS was run in ESI

positive (+) mode using multiple reaction monitoring (MRM) from 50 to 600 Da and single ion

recording (SIR) at 355 Da. Each analysis was performed in at least duplicate. Flumioxazin

amounts were quantified by correlating peak area detected to those of analytical grade standard

solutions of various known concentrations (0.01 to 0.5 µmol L-1). The limit of detection for

flumioxazin was approximately 3.0 (± 2.0) µg L-1.

2.5. Kinetic Evaluation of Herbicide Dissipation

Since pesticide dissipation can be described by either a linear or nonlinear method,

FOCUS (2014) work group guidelines were followed when selecting the kinetic model that best
59

described the data. Flumioxazin dissipation data from the field experiment were subjected to

regression analysis using SAS nonlinear regression (SAS Institute, Cary, NC) to determine

whether the response could be described by the following exponential decay equation (Grey et

al., 2018):

y = B0e-k(t) (1)

where y is the measured flumioxazin concentration (µmol L-1), B0 is the initial flumioxazin

concentration (µmol L-1) when time t is 0, k is the dissipation rate of flumioxazin (slope), and t is

time elapsed after flumioxazin application (day).

After regression against time, the output included the first-order dissipation rate constant

(k) (Ohmes et al., 2000). Flumioxazin field persistence (DT50) was then determined using the

half-life equation (Grey et al., 2018):

DT50 = ln(2)/k (2)

where DT50 is the half-life of flumioxazin (day) and k is the first-order dissipation rate constant.

After determining the net effect of dissipation, the groundwater ubiquity score (GUS)

index was used to assess the leaching potential of flumioxazin for these experiments by

combining the calculated persistence and mobility measurements (Gustafson, 1989):

GUS = log10(DT50)*(4-log10(Koc)) (3)

where DT50 is the half-life in days (determined from equation 2) and the Koc for flumioxazin is

approximately 557 mL g-1 (Shaner, 2014). A GUS index score of <1.8 represents a non-leaching

herbicide, 1.8-2.8 is a transitional herbicide that may leach, and >2.8 is considered a leaching

herbicide (Gustafson, 1989).

2.6. Kinetic Evaluation of Herbicide Thermal Stability in Solution


60

Differences in flumioxazin stability between temperatures were evaluated using

regression models that fit flumioxazin concentration (µmol L-1) to incubation time (h) for each

temperature to determine whether the response could be described by the exponential decay

equation (Eq. 1). The slope of the line for a given temperature represents the rate of change, or

first order rate constant, in flumioxazin concentration. Half-life (DT50) was then determined for

each temperature using Eq. 2.

To accurately characterize temperature effects on flumioxazin stability on the thermal

gradient table, the Arrhenius equation, expressed as a natural logarithm, was used:

ln k = -Ea/RT + ln A (4)

where Ea is the activation energy of flumioxazin (kJ mol-1), R is the universal gas constant

(8.314 J K-1 mol-1), T is the absolute temperature (K), and A is the pre-exponential factor.

To further describe the dependence of the thermal degradation rate on temperature, thermal

gradient table data were subjected to the Q10-value equation:

Q10 = exp(EaΔT/RT1T2) (5)

where Ea is the assumed activation energy (kJ mol-1), ΔT represents a 10° change in temperature

(K), R is the universal gas constant (0.008314 kJ K-1 mol-1), and T1 is a temperature that is 10°

lower than T2 (K) (Cessna et al., 2017).

2.7. Statistical Analysis

Field data were first analyzed separated by location and application timing. Parameter

estimates from these outputs were then tested for interactions. Data were then analyzed using

SAS 9.4 nonlinear regression and graphed in SigmaPlot 14.0 (Systat Software, San Jose, CA).

Environmental measurements, described in Table 1, and flumioxazin concentration over time


61

were subjected to correlation analysis using PROC CORR in SAS 9.4 to determine the

relationship between flumioxazin concentration over time and the environmental conditions.

Samples from the thermal gradient table were analyzed for flumioxazin concentrations

over temperature and time. Data were subjected to nonlinear regression using SAS 9.4 to

determine if there were any effects of temperature over-time on flumioxazin concentration. Data

were then graphed in SigmaPlot 14.0.

3. Results and Discussion

3.1. Dissipation of Flumioxazin in Soil

For Webster and Sumter counties, the parameter estimates did not differ between

application timing and flumioxazin concentration (p=0.18 and p=0.21, respectively). Data were

then combined across application timing and tested for interactions between locations. The

regression coefficients did not differ between location and flumioxazin concentration over time

(p=0.72), therefore data were combined across location and application timing.

3.1.1. Exponential Decay Model

The exponential decay equation represents the decline in the concentration of a reactant

as a function of time. Overall, the exponential decay equation (Eq. 1) adequately described

flumioxazin dissipation in soil (Figure 1). From the first-order regression analysis, several

parameters were determined. The average value for B0 was 0.49 (± 0.03) µmol kg soil-1, which

represents the concentration of flumioxazin at the time of application. The field rate applied was

3320 µmol L-1 or 0.51 (± 0.014) µmol kg soil-1. The average first-order rate constant (k)

determined was 0.026, similar to the k-value reported for flumioxazin dissipation in soil by

Ferrell and Vencill (0.038-0.053) and by Mueller et al. (0.032) (2003; 2014).
62

The determined half-life (DT50) for flumioxazin in soil was 26.6 d (Figure 1). Mueller et

al. reported a field half-life of 21.4 d for flumioxazin in Sequatchie loam (36% sand, 44% silt,

20% clay, and 1.9% OM) (2014). Under field conditions across four Chilean soils, flumioxazin

had a half-life of 10.6-32.1 d (Alister et al., 2008). In comparison to other residual herbicides,

such as fomesafen (DT50 = 46.0 d), indaziflam (DT50 = 53.0-63.0 d), and sulfentrazone (DT50 =

71.0 d), flumioxazin has a relatively short half-life (Mueller et al., 2014; Gonzalez-Delgado et

al., 2016). However, when compared to residual herbicides such as saflufenacil (DT50 = 21.1 d)

and pendimethalin (DT50 = 24.0 d), flumioxazin provides longer residual weed control (Mueller

et al., 2014; Kocarek et al., 2016).

Previous studies have reported positive correlations between clay content and the

degradation rates of pesticides in soil (Kah et al., 2007). However, greater sorption by soil

organic matter occurs for herbicides with low water solubility, because of the hydrophobic

surface of organic matter when compared to clay surfaces. Since flumioxazin has a water

solubility of 1.78 mg L-1, adsorption likely occurs to the organic matter fraction in soil (Ferrell et

al., 2005; Shaner, 2014). Ferrell et al. (2005) further supported this by determining that

flumioxazin adsorption depended on the percentage of organic matter present in the soil and

favored anionic exchangers over clay minerals as sorption sites. The preference of flumioxazin to

organic matter is also supported by the varying half-lives reported by Alister et al (2008). The

soil in which flumioxazin had the longest persistence (DT50 = 32.1 d) had an organic matter

content of 2.4%, while the soil showing the shortest persistence (DT 50 = 10.6 d) had an organic

matter content of 1.2% (Alister et al., 2008).

Ferrell and Vencill (2003), in laboratory experiments using similar soil types, reported

flumioxazin half-life values in Greenville sandy clay loam (58% sand, 10% silt, and 32% clay;
63

soil temperature: 25°C) of 17.9 d and in Tifton loamy sand (94% sand, 4% silt, and 2% clay; soil

temperature: 25°C) of 13.6 d. Varying half-life values from field and laboratory experiments

could be the result of increased degradation by microbial biomass under controlled conditions or

greater impacts by environmental conditions in field studies (Gonzalez-Delgado and Shukla,

2020).

3.1.2. Groundwater Ubiquity Score

Using the reported Koc of 557, our data indicated a GUS index of 1.79 for flumioxazin

(Shaner, 2014). This value ranks flumioxazin as a low-potential leacher in Greenville sandy clay

loam and Faceville loamy sand. Lewis et al. (2016) also ranked flumioxazin as having low

leachability potential and reported a GUS index of 1.55. Under rainfall conditions greater than

180 mm in 90 d after application the GUS index for flumioxazin was reported as 1.64 (low

leachability potential) (Kogan et al. 2007). Given the correlation between flumioxazin

concentration over time and cumulative rainfall, it can be inferred that instead of leaching out of

the soil profile and becoming a groundwater contaminant, flumioxazin likely moves downward

and stays in the soil solution.

Similar to flumioxazin, other commonly used residual herbicides are ranked as low to

transitional leachers. Flazasulfuron, pendimethalin, and oxyfluorfen are ranked as having low

leachability potential (0.50-1.33) while simazine, halosulfuron-methyl, rimsulfuron, and diuron

were ranked as transitional leachers (2.05-2.65) (Kogan et al., 2007; Lewis et al., 2016). An

example of a herbicide on the opposite end of the scale is atrazine, which was ranked at 3.16-

3.65 (easily leachable) (Liu et al., 2021).

3.1.3. Flumioxazin Degradation


64

Table 2 describes the correlations between flumioxazin concentration over time and the

environmental measures from Table 1. Correlation analysis showed that flumioxazin

concentration was negatively correlated with cumulative rainfall (r= -0.69, P < 0.001) and

cumulative solar radiation (r= -0.72, P < 0.001).

Flumioxazin can become readily available in soil solution with an increase in soil water

content from rainfall or irrigation events. This allows flumioxazin to move downward in the soil

profile, with most of the herbicide weakly adsorbing to soil colloids or staying in solution

(depending on the amount and occurrence of rainfall or irrigation), and consequently decreasing

the half-life (Alister et al., 2008; Ferrell et al., 2005). This supports the inverse relationship

determined from these experiments between flumioxazin concentration over time and cumulative

rainfall, which was also reported by Alister et al. (2008).

The main factors impacting flumioxazin concentration in solution are light and pH, which

lead to degradation by hydrolysis and photolysis (Kwon et al., 2004; Shibata et al., 2011). Alister

et al. reported an inverse relationship between pH and flumioxazin half-life due to rapid

hydrolysis as pH increases from pH 5 to pH 7 (2008). Shibata et al. determined the hydrolytic

degradation pathway of flumioxazin in aqueous solution to start with an opening of the imide

ring to form N-[7-fluoro-3-oxo-4-(2-propynyl)-2H-1,4-benzoxazin-6-yl]-3,4,5,6-

tetrahydrophthalamic acid, then a cleavage of the amide linkage to form 3,4,5,6-

tetrahydrophthalic acid and 6-amino-7-fluoro-4-(2-propynyl)-2H-1, 4-benzoxazin-3(4H)-one

(2014). These authors attributed a 1.1-1.7 times faster degradation of flumioxazin to the

degradation products in water with soil sediment when compared to an aqueous solution only to

microbial degradation and organic matter content in the sediment (Shibata et al., 2014).
65

The mobility of flumioxazin down into the rooting zone with water helps reduce the

likelihood of the compound to be degraded by photodegradation (Mueller et al., 2014). Previous

literature describes the degradation of flumioxazin solely by photolysis resulted in a half-life of

4.9-42.0 hr (pH 5-9), which supports the strong inverse relationship determined between

flumioxazin concentration over time and cumulative solar radiation (Kwon et al., 2004). Kown et

al. reported the photolytic degradation of flumioxazin at a pH of 7 as having as half-life 10 times

lower than at a pH of 5, which was attributed to an opening of the imide ring followed by

cleavage of an amide linkage (2004). Shibata et al. determined the half-life of flumioxazin in an

aqueous solution with and without sediment to be 0.2-1.5 and 0.1-0.3 d (2014). These authors

further contributed direct photolysis to degrade flumioxazin into two products, N-(2-propynyl)-4-

[4-carboxy-3- fluoro-2-(3,4,5,6-tetrahydrophthalimido)-2-butenylidene]azetidine-2-one and N-

(2- propynyl)-4-[4-carboxy-3-fluoro-2-(2-carboxy-1-cyclohexenecarbonylamino)-2-

butenylidene]azetidine-2-one (Shibata et al., 2014).

Flumioxazin concentration over time was not correlated to soil temperature (P=0.29)

(Table 2). Ferrell and Vencill also determined that soil temperature (15-25°C) had little to no

impact on flumioxazin degradation in soil (2003). Herbicides with half-lives less than 30 d, such

as flumioxazin and metolachlor, are typically known to have little impacts on degradation rates

from soil temperature (Lehmann et al., 1993).

3.2. Thermal Stability of Flumioxazin

Figure 2 represents the time course for degradation of flumioxazin in solution (pH 4.5) at

various temperatures (11 to 40°C) and Table 3 describes the parameter estimates from Figure 2,

including degradation rate constants (k), half-lives (DT50), and r2 values on a thermal gradient.

Figure 3 represents the linear Arrhenius plot of ln k against 1/T, which gives a slope equal to -
66

Ea/R (r2 = 0.87). The Arrhenius equation can be used to describe a pesticide’s dependence of

degradation on temperature through the activation energy (Grey et al., 2018). The Arrhenius

equation represents the magnitude of the rate constant as a function of activation energy over the

average kinetic energy (-Ea/RT). The activation energy is the minimum amount of kinetic energy

required to generate a reaction at the molecular level (Ma et al., 2017).

Specific to flumioxazin, 58.4 (± 1.2) kJ mol-1 is the threshold energy, determined from

the present studies (pH 4.5), required to reach the transition state where degradation begins

(Table 3). The amount of energy, based on the determined activation energy, for degradation to

start equaled 1.63 x 10-6 kJ cylinder -1 (cylinder volume = 348 cm3) for the field experiment. From

Table 1, the lowest recorded amount of solar radiation was equal to 31.9 kJ cylinder -1 (7 MJ m-2)

and the highest amount was equal to 118.6 kJ cylinder-1 (26 MJ m-2). Also from Table 1, the

lowest recorded soil temperature was equal to 13.0 kJ cylinder-1 (9°C) and the highest amount

was equal to 33.3 kJ cylinder -1 (23°C). All of these energy values exceed the activation energy

needed for flumioxazin degradation and is supported by the immediate field dissipation shown in

Figure 1. This is also supported by the lack of correlation between flumioxazin concentration

over time and soil temperature since even at the lowest recorded temperature there was enough

energy to cause degradation (Table 2). Using the lowest temperature (11°C) of the thermal

gradient table experiment, the energy from one cell of the thermal gradient table (described in

2.3.) equaled 7.5 x 108 kJ mol-1. This value also exceeds the activation energy needed for

flumioxazin degradation and is supported by the immediate degradation across all temperatures

(Figure 2).

Previous literature has reported activation energies for various herbicides, include

glufosinate (77.0 kJ mol-), metribuzin (54.6 kJ mol-1), mesosulfuron-methyl (79.1 kJ mol-1),


67

metsulfuron-methyl (54.7-82.4 kJ mol-1), and 2,4-D (96.1-190.4 kJ mol-1) (Parker and Doxtader,

1983; PPR-Panel, 2007; Wang et al., 2010). Comparatively, flumioxazin has lower activation

energy than the previously listed herbicides, indicating that flumioxazin is less sensitive to

temperature. Cessna et al. reported metsulfuron-methyl and 2,4-D to have field half-lives of 10-

38 d and 4-7 d, respectively (2017). Flumioxazin has a similar activation energy (58.4 (± 1.2) kJ

mol-1) and field half-life (26.6 d) to metsulfuron-methyl with 2,4-D having higher activation

energy and shorter half-life than flumioxazin.

Halosulfuron-methyl, another commonly used residual herbicide, is also influenced by

temperature over time, with a reported activation energy of 86 kJ mol-1, while another

sulfonylurea herbicide, rimsulfuron, has a reported activation energy of 65.1-92.0 kJ mol-1

(Sarmah and Sabadie, 2002; Grey et al., 2018; Vicari et al., 1996). Diuron is reported to have an

activation energy of 127 kJ mol-1 with a field half-life of 4 months (Salvestrini et al., 2002). In

contrast, pendimethalin is known to be stable in aqueous solutions when subjected to

temperatures up to 120°C with an average field half-life of 24-34 days (EFSA, 2016; Kocarek et

al., 2016).

The greater the activation energy, the more sensitive the dissipation of the active

ingredient is to temperature. This temperature dependence can also be described using Q 10, the

temperature coefficient. The European Food Safety Authority (EFSA) has a default Q 10 value of

2.2, which correlates to an activation energy of 54 kJ mol-1, however, the EFSA has

recommended using the activation energy to calculate pesticide-specific Q10 values (2016;

Marin-Benito et al., 2019).

The Q10 value determined from the present studies for flumioxazin was 2.2 (averaged

over temperatures 20-40°C). This value, coupled with an activation energy of 58.4 (± 1.2) kJ
68

mol-1, fits well within the EFSA default values (2016). Mamy et al. reported glyphosate,

trifluralin, and metolachlor in a clay loam soil to have Q10 values of 1.7-2.3 and Cessna et al.

reported Q10 values of 1.22 and 3.47 for metsulfuron-methyl and 2,4-D, respectively (2008;

2017). Other herbicides, such as glufosinate, metribuzin, mesosulfuron-methyl, and rimsulfuron,

have reported Q10 values of 2.2-3.2 (PPR-Panel, 2007).

4. Conclusions

In conclusion, flumioxazin soil dissipation followed first-order kinetics and was

adequately described by the exponential decay equation across all four herbicide applications in

Georgia pecan groves. Flumioxazin half-life in Greenville sandy clay loam (61.9% sand, 16.1%

silt, 22.0% clay, 2.1% organic matter, bulk density of 1.63 g/cm3, and pH 5.3) and Faceville

loamy sand (77.9% sand, 16.1% silt, 6.0% clay, 3.9% organic matter, bulk density of 1.54 g/cm3,

and pH 5.9) was 26.6 d. The relatively short half-life of flumioxazin in soil (26.6 d) is consistent

with the amount of residual control that should be expected. Flumioxazin is used as a residual

herbicide to control sensitive weed species without concern for long-term rotational effects for

most crops grown in the Southeastern US.

The combination of soil and environmental properties creates a dynamic system that

directly impacts herbicide behavior and persistence. Flumioxazin was determined to have a

groundwater ubiquity score of 1.79, indicating flumioxazin has a low leachability potential.

There was an inverse correlation between flumioxazin concentration, rainfall, and solar radiation.

Flumioxazin is known to stay in soil solution, which can lead to degradation through hydrolysis

or photolysis. While there was no direct correlation between flumioxazin concentration and soil

temperature, flumioxazin dissipation is known to be directly influenced by soil microbial


69

activity. Microbial degradation is influenced by soil temperature, with increases in temperature

decreasing the rate of degradation (Nash, 2016).

Both the temperature coefficient (Q10) and activation energy for a reaction provides a

measurement of how the rate constant (k) for that reaction will be affected by a change in

temperature. Flumioxazin was determined to have an activation energy of 58.4 (± 1.2) kJ mol-1

and a Q10 value of 2.2. Even at the lowest amount of solar radiation and soil temperature, the

energy from these environmental measures exceeded the activation energy needed for

flumioxazin degradation and is shown in Figures 1 and 2 by the immediate degradation. Overall,

flumioxazin stability in aqueous solution related to field dissipation indicated that as the input of

thermal energy increases, flumioxazin degradation increases.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.
70

References

Alister, C., Rojas, S., Gomez, P., Kogan, M., 2008. Dissipation and movement of flumioxazin in

soil at four field sites in Chile. Pest Manag. Sci. 64, 579-583.

Belding, R. D., Majek, B. A., Lokaj, G. R. W., Hammerstedt, J., Aven, A. O., 2004. Orchard

floor management influence on summer annual weeds and young peach tree performance.

Weed Technol. 18, 215-222.

Boyd, N. S., Sharpe, S. M., Kanissery, R., 2020. Flumioxazin soil persistence under plastic

mulch and effects of pretransplant applications on strawberry. Weed Technol., 1-5.

Cessna, A. J., Knight, J. D., Ngombe D., Wolf T. M., 2017. Effect of temperature on the

dissipation of seven herbicides in a biobed matrix. Can. J. Soil Sci. 97, 717-731.

Company, R. S. I., Gradziel, T. M., 2017. Almonds: botany, production, and uses. CABI,

Boston, MA. pp 510.

Eskilsson, C., S., Bjorklund, E., 2000. Analytical-scale microwave-assisted extraction. J. Chrom.

A. 902, 227-250.

European Food Safety Authority [EFSA], 2016. Conclusion on the peer review of the pesticide

risk assessment of the active substance pendimethalin. EFSA J. 14, 12-69.

Faircloth, W. H., Patterson, M. G., Foshee, W. G., Nesbitt, M. L., Goff, W. D., 2007.

Comparison of preemergence and postemergence weed control systems in newly

established pecan. Weed Technol. 21, 972-976.

Ferrell, J. A., Vencill, W. K., 2003. Flumioxazin soil persistence and mineralization in laboratory

experiments. J. of Ag. and Food Chem. 51, 4719-4721.

Ferrell, J. A., Vencill, W. K., Xia, K., Grey, T. L., 2005. Sorption and desorption of flumioxazin

to soil, clay minerals, and ion-exchange resin. Pest Manag. Sci. 61, 40-46.
71

FOCUS, 2014. Generic Guidance for Estimating Persistence and Degradation Kinetics from

Environmental Fate Studies on Pesticides in EU Regulation. Report version 1.1.

Gonzalez-Delgado, A. M., Shukla, M. K., Ashigh, J., Perkins, R., 2016. Effect of application rate

and irrigation on the movement and dissipation of indaziflam. J. Env. Sci. 51, 111-119.

Gonzalez-Delgado, A. M., Shukla, M. K., 2020. Mobility, degradation, and uptake of indaziflam

under greenhouse conditions. Hort Sci. 55, 1216-1221.

Grey, T. L., Turpin, F. S., Wells, L., Webster, T. M., 2014. A survey of weeds and herbicides in

Georgia pecan. Weed Technol. 28, 552-559.

Grey T. L., Culpepper, A. S., Li X., Vencill, W. K., 2018. Halosulfuron-methyl degradation from

the surface of low-density polyethylene mulch using analytical and bioassay techniques.

Weed Sci. 66, 15-24.

Gustafson, D. I., 1989. Groundwater ubiquity score: a simple method for assessing pesticide

leachability. Env. Toxic. Chem. 8, 339-357.

Grahovac, N. L., Stokanovic, Z. S., Kravic, S. Z., Orcic, D. Z., Suturovic, Z. J., Kondic-Spika, A.

D., Vasin, J. R., Sunjka, D. B., Jaksic, S. P., Rajkovic, M. M., Grahovac, N. M., 2017.

Determination of residues of sulfonylurea herbicides in soil by using microwave-assited

extraction and high performance liquid chromatographic method. Hem. Ind. 71, 289-298.

Hurdle, N. L., Grey, T. L., Pilon, C., Monfort, W. S., Prostko, E. P., 2020. Peanut seed

germination and radicle development response to direct exposure of flumioxazin across

multiple temperatures. Peanut Sci. 47, 89-93.

Kah, M., Beulke, S., Brown, C. D., 2007. Factors influencing degradation of pesticides in soil. J.

Agric. Food Chem. 55, 4487-4492.


72

Kocarek, M., Karel, V., Boruvka, L., Artikov, H., 2016. Pendimethalin degradation in soil and its

interaction with soil microorganisms. Soil Water Res. 11, 1-7.

Kogan, M., Rojas, S., Gomez, P., Suarez, F., Munoz, J. F., Alister, C., 2007. Evaluation of six

pesticides leaching indexes using field data of herbicide application in Casablanca

Valley, Chile. Water Sci. & Tech. 56, 169-178.

Kwon, J. W., Armnrust, K. L., Grey, T. L., 2004. Hydrolysis and photolysis of flumioxazin in

aqueous buffer solutions. Pest Manag. Sci. 60, 939-943.

Lehmann, R. G., Fontaine, D. D., Olberding, L. L., 1993. Soil degradation of flumetsulam at

different soil temperatures in the laboratory and field. Weed Res. 33, 189-195.

Liu, J., Zhou, J. H., Guo, Q. N., Ma, L. Y., Yang, H., 2021. Physiochemical assessment of

environmental behaviors of herbicide atrazine in soils associated with its degradation and

bioavailability to weeds. Chemosphere 262, 127830.

Lewis, K. A., Tzilivakis, J., Warner, D., Green, A., 2016. An international database for pesticide

risk assessments and management. Human Eco. Risk Assess.: An Int. J. 22, 1050-1064.

Ma, J., Li, H., Chi, L., Chen, H., Chen, C., 2017. Changes in activation energy and kinetics of

heat-activated persulfate oxidation of phenol in response to changes in pH and

temperature. Chemosphere 189, 86-93.

Mamy, L., Gabrielle, B., Barriuso, E., 2008. Measurement and modeling of glyphosate fate

compared with that of herbicides replaced as a result of the introduction of glyphosate-

resistant oilseed rape. Pest Manag. Sci. 64, 262-275.

Marin-Benito, J. M., Carpio, M. J., Sanchez-Martin, M. J., Rodriguez-Cruz, M. S., 2019.

Previous degradation study of two herbicides to simulate their fate in a sandy loam soil:

effect of the temperature and the organic amendments. Sci. Total Env. 653, 1301-1310.
73

Matringe, M., Camadro, J. M., Labbe, P., Scalla, R., 1989. Protoporphyrinigen oxidase as a

molecular target for diphenyl ether herbicides. Biochem. J. 260, 231-235.

McGonigle, T. P., Turner, W. G., 2017. Grasslands and croplands have different microbial

biomass carbon levels per unit of soil organic carbon. Agric. 57, 1-8.

Moreland, D. E., 1999. Biochemical mechanisms of action of herbicides and the impact of

biotechnology on the development of herbicides. J. Pestic. Sci. 24, 299-307.

Mueller, T. C., Boswell, B. W., Mueller, S. S., Steckel, L. E., 2014. Dissipation of fomesafen,

saflufenacil, sulfentrazone, and flumioxazin from a Tennessee soil under field conditions.

Weed Sci. 62, 664-671.

Mueller, T. C., Senseman, S. C., 2015. Methods related to herbicide dissipation or degradation

under field or laboratory conditions. Weed Sci. Special Issue, 133-139.

Nash, R. G., 2016. Dissipation from soil, in: Environmental Chemistry of Herbicides. Florida:

CRC Press, pp. 131-169.

Ohmes, G. A., Mueller, T. C., Hayes, R. M., Sulfentrazone dissipation in a Tennessee soil. Weed

Tech. 14, 100-105.

Panel on Plant Protection Products and Their Residues [PPR-Panel], 2007. Opinion on a request

from EFSA related to the default Q10 value used to describe the temperature effect on

transformation rates of pesticides in soil. EFSA J. 622, 1-32.

Parker, L. W., Doxtader, K. G., 1983. Kinectics of the microbial degradation of 2,4-D in soil:

effects of temperature and moisture. J. Env. Qual. 12, 553-558.

Salvestrini, S., Carbo P. D., Capasso, S., 2002. Kinetics of the chemical degradation of diuron.

Chemosphere 48, 69-73.


74

Sarmah, A. K., Sabadie, J., 2002. Hydrolysis of sulfonylurea herbicides in soils and aqueous

solutions: a review. J. Argic. Food Chem. 50, 6253-6265.

Shaner, D. L., 2014. Herbicide Handbook, tenth ed. Lawrence, Kansas, pp. 212-213.

Shibata, A., Kodaka, R., Fujisawa, T., Katagi, T., 2011. Degradation of flumioxazin in

illuminated water-sediment systems. J. of Ag. and Food Chem. 59, 11186-11195.

Smith, M. W., 2011. Pecan production increased by larger vegetation free area surrounding the

tree. Scientia Hort. 130, 211-213.

Sparling, G. P., 1985. The soil biomass, in: Soil Organic Matter and Biological Activity.

Dordrecht, pp. 223-224.

Soulas, G., Lagacherie, B., 2001. Modelling of microbial degradation of pesticides in soil. Biol.

Fert. Soils 33, 551-557.

United States Department of Agriculture-National Agricultural Statistics Service [USDA-

NASS], 2020. Noncitrus Fruits and Nuts: 2019 Summary. Natl. Agr. Stat. Serv.,

Washington D.C.

Vicari, A., Zimdhal, R. L., Cranmer, B. K., Dinelli, G., 1996. Primisulfuron and rimsulfuron

degradation in aqueous solution and adsorption in six Colorado soils. Weed Sci. 44, 672-

677.

Vryzas, Z., Papadopoulou-Mourkidou, E., 2002. Determination of triazine and chloroacetanilide

herbicides in soils by microwave-assisted extraction (MAE) coupled with gas

chromatographic analysis with either GC-NPD or GC-MS. J. Agric. Food Chem. 50,

5026-5033.

Wang, H., Xu, J., Yates, S. R., Zhang, J., Gan, J., Ma, J., Wu, J., Xuan, R., 2010. Mineralization

of metsulfuron-methyl in Chinese paddy soils. Chemosphere 78, 335-341.


75

Wells, L., 2014. Pecan planting trends in Georgia. Hort. Tech. 24, 475-479.
76

Table 3.1. Environmental measures recorded during the course of the experiment from the

University of Georgia weather station. a

Applicationb Sample Date DAAc R (mm) SR (MJ/m2) ST (°C; 10 cm)


Webster 1: 4/19/2016 0 0 26 23
4/22/2016 3 20 74 22
4/26/2016 7 20 171 27
5/5/2016 16 23 360 24
5/13/2016 24 23 555 29
5/19/2016 30 33 676 26
6/1/2016 43 43 1001 34
7/7/2016 79 146 1821 33
8/3/2016 106 204 2417 29
9/23/2016 157 359 3423 31
12/12/2016 237 479 4593 14
3/14/2017 329 897 5624 11

Webster 2: 3/21/2017 0 0 22 22
3/27/2017 6 1 136 23
4/11/2017 21 97 451 24
5/15/2017 55 196 1235 29
6/19/2017 90 396 1944 30
7/21/2017 122 578 2595 33
8/29/2017 161 672 3400 31
12/13/2017 267 993 4898 9

Sumter 1: 3/6/2017 0 0 15 17
3/9/2017 3 4 63 18
3/13/2017 7 20 125 11
3/20/2017 14 22 244 17
4/4/2017 29 52 529 22
5/4/2017 45 181 1219 24
5/24/2017 65 301 1674 24
6/19/2017 90 418 2189 30
7/21/2017 122 600 2840 33
8/22/2017 154 694 3493 33
12/13/2017 282 1015 5142 9

Sumter 2: 2/28/2018 0 0 7 18
3/7/2018 7 25 108 15
3/20/2018 20 61 325 21
4/26/2018 57 218 1007 21
5/24/2018 85 319 1597 27
7/13/2018 135 582 2613 34
a
Cumulative values are reported from initial herbicide application date for DAA, SR, and R.
ST values are reported for each sample date. UGA weather station is approximately 20 km
from field sites.
b
Application dates: Webster 1, 4/19/2016; Webster 2, 3/2/2017; Sumter, 3/6/2017; Sumter 2,
2/27/2018.
c
Abbreviations: DAA, days after application; R, rainfall; SR, solar radiation; ST, soil
temperature (10 cm).
77

Table 3.2. Correlation values of flumioxazin concentration over time and environmental

measures recorded during the course of the experiment from the University of Georgia weather

station.a

DAAb CONC R SR ST
1 -0.69 0.95 0.99 -0.16 CC
DAA
- < 0.001 < 0.001 < 0.001 NS p-value
-0.69 1 -0.69 -0.72 -0.18 CC
CONC
< 0.001 - < 0.001 < 0.001 NS p-value
0.95 -0.69 1 0.94 -0.07 CC
R
< 0.001 < 0.001 - < 0.001 NS p-value
0.99 -0.72 0.94 1 -0.07 CC
SR
< 0.001 < 0.001 < 0.001 - NS p-value
-0.16 -0.18 -0.07 -0.07 1 CC
ST
NS NS NS NS - p-value
a
Each value is the average of four replicates per experiment, with experiments conducted twice
over time and combined for presentation. Correlation analysis was performed at α=0.05.
b
Abbreviations: DAA, days after application; CONC, flumioxazin concentration; R,
cumulative rainfall; SR, cumulative solar radiation; ST, soil temperature at each respective
sample date; CC, correlation coefficient; NS, non-significant.
78

Table 3.3. Parameter estimates for flumioxazin stability in aqueous solution (pH 4.5) at various

temperatures over time when evaluated on a thermal gradient table. a

Temperature (°C) kb Standard Error (±) DT50c (hr) r2


11 0.0106 1.1 x 10-3 65.4 0.71
15 0.0124 1.5 x 10-3 55.9 0.70
20 0.0114 1.3 x 10-3 60.8 0.69
25 0.0167 1.9 x 10-3 41.5 0.73
30 0.0327 3.9 x 10-3 21.2 0.75
35 0.0816 9.1 x 10-3 8.5 0.81
40 0.0784 9.5 x 10-3 8.8 0.76
Activation Energyd: 58.4 (± 1.2) kJ mol-1
a
Each value is the average of four replicates per experiment, with experiments conducted twice
over time and combined for presentation.
b
Degradation rate constants (k) were calculated by non-linear regression of the herbicide
quantity with respect to time (0 to 336 hr).
c
Half-lives (DT50) were determined using the k-values and the equation: DT50 = ln(2)/k.
d
Activation energy was determined using the Arrhenius equation expressed as a natural
logarithm: ln k = -Ea/RT + ln A.
79

0.8
Flumioxazin Concentration (μmol kg soil-1)

0.6

0.4

0.2

0.0
0 100 200 300
DaysAfter Application (day)

Figure 3.1. Flumioxazin soil persistence (t = day) using the exponential decay equation

combined over four site-years (Greenville sandy clay loam with 61.9% sand, 16.1% silt, 22.0%

clay, 2.1% organic matter, bulk density of 1.63 g/cm3, pH 5.3, and soil microbial biomass of 266

C kg-1; Faceville loamy sand with 77.9% sand, 16.1% silt, 6.0% clay, 3.9% organic matter, bulk

density of 1.54 g/cm3, pH 5.9, and soil microbial biomass of 686 C kg-1). Non-linear regression

was applied. Model shows that data can be described by first-order kinetics. Lines represent the

first-order regression equation. Data points indicate the means of replications. Error bars

represent the standard error of each mean (SEM). Initial flumioxazin concentration applied (0.49

µmol kg soil-1) is represented per cylinder volume (347.5 g cm-3). Parameter estimates: y =

0.4929e(-0.0261*t), r2: 0.67.


80

11C 15C 20C 25C 30C 35C 40C


Flumioxazin Concentration (umol/L)

0.3

0.2

0.1

0.0
0 48 96 144 192 240 288 336
Incubation Time (hr)

Figure 3.2. Time course for degradation of flumioxazin in solution (pH 4.5) at various

temperatures (11 to 40°C) on a thermal gradient using the exponential decay equation (y = B0e-
B1(t)
). Data points indicate the means of four replications, with experiments conducted twice and

combined. Error bars represent the standard error of each mean (SEM).
81

-2

-3
ln k

-4

-5
3.1e-3 3.2e-3 3.3e-3 3.4e-3 3.5e-3 3.6e-3

1/T x 10-3

Figure 3.3. Arrhenius plot of flumioxazin degradation. The slope of the line is equal to -Ea/R,

where Ea is the activation energy (kJ mol-1) and R is the universal gas constant (0.00831 kJ mol-
1
). From this plot, the determined activation energy was 58.4 (± 1.2) kJ mol-1 (r2 = 0.87).
82

CHAPTER 4

QUANTIFYING HERBICIDE MOVEMENT FROM PLASTIC MULCH INTO THE PLANT

HOLE USING SIMULATED VEGETABLE BEDS 1

________________________________________________________
1
Eason, K. M., Grey, T. L., Culpepper, A. S., Hurdle, N. L., de Souza Rodrigues, J., and
Coolong, T. To be submitted to Agronomy.
83

Abstract:

Vegetable production on plastic mulch in Georgia includes over 10 crops with a farm

gate value exceeding $596 million. These mulched systems often combine the use of fumigation,

drip tape, raised beds, and plastic mulch where 3 to 5 high-value crops are produced over a two-

year period. With the loss of methyl bromide, herbicides applied over plastic mulch prior to

planting a crop have become essential to maintain weeds, but care must be taken to avoid crop

damage. Experiments using simulated vegetable beds were conducted to quantify the

concentration of halosulfuron-methyl, glufosinate, glyphosate, S-metolachlor, and acetochlor

remaining on the mulch and to quantify the amount moving into the crop transplant hole with

irrigation was applied.

When irrigating at least 0.63-cm, less than 2% of halosulfuron, glufosinate, and

glyphosate remained on the surface of the plastic mulch. In contrast, 91% and 15% of acetochlor

remained on the mulch after irrigating with 0.63-cm and 1.27-cm, respectively. S-metolachlor

remained on the surface of the plastic mulch at 17% and 3% after the aforementioned irrigation

volumes, respectively. The order of concentration detected in the transplant hole area was

equivalent to ranking the herbicides by water-solubility. All herbicide concentration was detected

below 1.0 mg ai or ae in the transplant hole area despite irrigation amount. For halosulfuron,

glyphosate, and glufosinate, these concentrations were equal to a 1.3X-8.9X field rate washing

into the transplant hole. Acetochlor and S-metolachlor concentrations in the transplant hole were

equivalent to 0.1X-0.7X field rates, respectively.

Keywords: Plasticulture; Herbicide Wash-Off; Transplant Hole


84

1. Introduction

Plasticulture is a management technique that offers growers improved plant growth,

quality, and weed control through the use of fumigation, drip tape, raised beds, and plastic mulch

(Sanders et al. 1996; Dickerson 2007). For these reasons, mulch production has gained

favorability among Georgia vegetable growers producing fruiting, cucurbit, and cole crop

vegetables. Plastic mulch systems utilize low-density polyethylene (LDPE) mulch or totally

impermeable films (TIF) stretched over aggressively tilled soils that have been formed into

raised beds that range in width and height (Lamont 1996). These systems are often used for 3 to

5 different vegetable crops (transplanted and seeded) over the course of 2 years, spreading the

cost of production for the plastic mulch, drip tape, and fumigation over multiple crops (Nyoike

and Liburd 2014; Culpepper et al. 2009; SEVEW 2020).

To keep the integrity of the mulch intact across the multiple crops and over time, weed

control and termination of the previous crop is crucial. Fumigation is the most effective approach

to control weeds and terminate the previous crop in this system, but with the loss of methyl

bromide escapes occur and additional measures are required (Gilreath and Santos 2004; Webster

et al. 2005). Thus, herbicides have become an integral component in both controlling escapes of

both weeds and the previous crop. Nutsedge species (Cyperus spp.) are especially troublesome

for crops grown on mulch, as it is the only weed present that penetrates through the mulch, with

the potential to cause significant damage (Webster 2002; Johnson and Mullinix 2017). However

other weeds, such as Amaranthus, Portulaca, Ipomoea, and annual grasses are problematic as the

commonly infest row middles, crop plant holes punched in the mulch from planting, and holes in

the mulch from degradation or animal damage (Wychen 2019). These weeds, even at low

densities, can be extremely competitive and costly in vegetables.


85

To control nutsedge and other weed species from damaging the mulch, producers can

apply herbicides on-top of and under the mulch. Previous literature indicates that halosulfuron-

methyl (methyl 3-chloro-5-[[[[(4,6-dimethoxy-2 pyrimidinyl)amino]carbonyl]amino]sulfonyl]-1-

methyl-1H-pyrazole-4-carboxylate), glyphosate (N-phosphonomethyl)glycine) (Grey et al.

2009), glufosinate (2-amino-4-(hydroxymethylphosphinyl)butanoic acid) (Sharpe and Boyd

2019), and S-metolachlor (2-chloro-N-(2-ethyl-6-methylphenyl)-N-[(1S)-2-methoxy-1-

methylethyl]acetamide) (Cornelius et al. 1985) are all herbicides that are used or have the

potential to be used over-the-top of plastic mulch to mitigate previous crop and weed escapes.

Halosulfuron-methyl is a sulfonylurea herbicide that is relied on for control of Cyperus

spp. in a variety of cropping systems, including plasticulture vegetable production (Anonymous

2015; Shaner 2014). Glyphosate and glufosinate are non-selective, contact herbicides that control

a wide range of weed species (Anonymous 2017a; Anonymous 2017b; Shaner 2014). Acetochlor

(2-chloro-N-(ethoxymethyl)-N-(2-ethyl-6-methylphenyl)acetamide) and S-metolachlor are

chloroacetamide herbicides known for controlling annual grasses and several broadleaf weeds,

including Palmer amaranth (Anonymous 2020a; Anonymous 2020b; Shaner 2014). Since these

herbicides are applied broadcast over mulch and row middles to help manage these troublesome

weeds, herbicide persistence on plastic mulch and its movement into holes punched for crop

transplants must be better understood to avoid crop injury (Grey et al. 2002; Grey et al. 2009).

Determining the movement of herbicides from plastic mulch into the transplant crop hole

has not been studied and there is need for additional research determining the influence of

rainfall or irrigation in removing herbicides from plastic mulch. Thus, the objective of this

experiment was to determine the concentration of five herbicides remaining on the surface of

plastic mulch used for vegetable production and into a transplant hole after irrigation.
86

Quantification of herbicides was conducted 1) on the plastic mulch prior to irrigation; 2) on the

plastic mulch after irrigation; and 3) in water accumulated in the transplant hole.

2. Materials and Methods

2.1. Simulated Vegetable Bed Design

Figure 4.1 depicts a simulated bell pepper (Capsicum annuum) production bed (2.44-m

wide by 0.76-long, equaling 1.86-m2). A slope of 6.7% from the center of the bed to each outside

edge was included to prevent water pooling, following standard field protocol. Soil was

mimicked by adding a layer of carpet padding (Leggett & Platt®, Carthage, MO) on-top of the

plywood. Holes (64-mm diameter) were cut into the 0.63-cm thick plywood. Seven holes were

placed on each side of the bed (14 total holes) with a spacing of 30.5-cm down the bed and 20.3-

cm across the bed, following standard field production procedures. TIF (Guardian Agro Plastics,

Tampa, FL) was secured to the simulated bed using staples down each side of the bed with

excess removed. Transplant holes were punched into the plastic mulch using a hand-made 5-cm

x 5-cm wooden wedge, which is a standard size used for field production.

Jars (0.95 L) were placed on the under-side of each transplant hole using screw-on metal

lids, with the flat lid piece removed. These were attached using screws, being careful not to

penetrate the carpet or mulch (Figure 4.2.). A piece of plywood (0.76 x 0.76 m) was placed at the

end of each bed and covered with the same TIF that was used for the simulated beds. This

section was used to sample the plastic mulch after herbicide application but prior to irrigation.

For irrigation, a stationary system was constructed using PVC pipe and a single sprinkler

(Rainbird® Surepop Sprinkler, Azusa, CA) equipped with a high-efficiency variable arc nozzle

with a 4.5-m radius. The simulated irrigation was maintained at 140 kPa using a pressure
87

regulator and fed by a city-water faucet source. Multiple tests were conducted to ensure

uniformity of irrigation across the surface of each bed and even water distribution, prior to

conducting the herbicide experiment (±0.7%).

2.2. Experimental Design

An experiment was conducted at the Coastal Plains Experiment Station (31.475, -83.527;

The University of Georgia, Tifton, GA 31794) in 2020. The experiment was arranged as a 2 by 5

factorial (two irrigation volumes and five herbicide options) in a randomized complete block

design with three replications and was repeated twice in time. Halosulfuron-methyl, glufosinate,

glyphosate, acetochlor, and S-metolachlor were applied over-the-top of the simulated mulch beds

at field use rates for vegetable production (Table 4.1) (Culpepper 2020). Applications of

halosulfuron-methyl included non-ionic surfactant (0.25% v/v). All herbicide treatments were

applied using a CO2-pressurized backpack sprayer, calibrated to deliver 140 L ha -1 at 150 kPa,

held approximately 46-cm above the mulch. The spray boom was equipped with three TTI11002

nozzles with a 46-cm spacing (Teejet Technologies, Wheaton, IL). After application, the surface

of the mulch was allowed to completely dry (1-4 h) prior to initiating irrigation (either 0.63-cm

or 1.27-cm).

2.3. Data Collection

Various environmental measurements, including air temperature (°C), wind speed (kph),

relative humidity (%) at a height of 2 m, and cloud cover (%), were recorded at the time of each

herbicide application by sensors located on site (Table 4.2).

Mulch samples after herbicide application but before irrigation were taken from the

mulch covered plywood (0.76-m x 0.76-m) adjacent to each bed, while samples after herbicide

application and after irrigation were taken from the center of each bed. Plastic mulch sampling
88

procedures were adapted from previous literature (Grey et al. 2009, Grey et al. 2018). Mulch

samples were collected using an open-faced square frame (0.1m2) and box-cutting knife. Needle-

nose pliers are used to place the individual mulch samples into sealable plastic bags, by only

touching the under-side of the mulch, to prevent contamination between samples and were then

immediately frozen upon collection at -10°C until analysis. Used mulch was removed after the

data collection was complete for each treatment and replaced with new plastic mulch.

After herbicide application and irrigation, the 14 jars were removed from the beds. The

amount of water in each jar was recorded using a graduated cylinder. One cylinder was used per

simulated bed and individual cylinders were rinsed once with organic solvent and then triple

rinsed with water to ensure no cross-contamination between measurements. Approximately 10

mL of sample was taken from each jar and transferred into a 20-mL glass vial with a screw-on

cap (14 subsamples per replication). Samples from the jars were collected to determine the

respective herbicide concentration. The 20-mL vials were frozen immediately after collection at -

10°C until analysis. The site of the experiment was 0.46 km from the freezers where samples

were stored until analysis.

2.4. Sample Preparation

2.4.1. Plastic Samples

Methods for herbicide extraction from plastic mulch were adapted from previous

literature (Grey et al. 2018, Grey et al. 2009, Hand et al. 2021, Shaner 2014). Sample integrity

was maintained throughout sample collection, preparation, and analysis (Mueller and Senseman

2015). Plastic mulch samples were removed from the freezer and allowed to equilibrate to room

temperature before processing. Samples were placed into individual 125 mL volumetric flasks

containing 100 mL of a methanol:water solution (50:50, 90:10, and 20:80 v/v for halosulfuron-
89

methyl, acetochlor and S-metolachlor, and glyphosate and glufosinate, respectively), and capped

with a rubber stopper. Flasks were then placed on a reciprocating shaker for 2 hr at 200 rpm.

Once completed, an aliquot from each sample was transferred into HPLC vials (Fisher Scientific,

Waltham, MA) for analysis.

2.4.2. Water Samples

Individual water samples in the 20-mL vials were removed from the freezer and allowed

to acclimate to room temperature before analysis. Since soil was not a factor in these

experiments, samples did not warrant further cleanup or filtering. An aliquot (1.0 mL) was then

transferred into a HPLC-vial (Fisher Scientific).

2.5. Analytical Methods

Herbicide concentration was analyzed by a Waters Acquity Arc ultra-high performance

liquid chromatography (UHPLC) system coupled with a Waters 2998 PDA and Waters QDa

mass spectrometry (MS) detector (Waters Corporation, Milford, MA). The LC separation was

performed on a Cortecs® C18 reverse-phase column (4.6 mm x 50 mm, 2.7 µm; Waters

Corporation) for halosulfuron-methyl, an Anionic Polar Pesticide column (2.1 mm x 100 mm, 5

µm; Waters Corporation) for glyphosate and glufosinate, and a Symmetry C18 reverse-phase

column (4.6 mm x 75 mm, 3.5 µm; Waters Corporation) for acetochlor and S-metolachlor. Each

analysis was duplicated. The various herbicide amounts were quantified by correlating peak area

detected to those of analytical grade standard solutions of various known concentrations. The

limit of detection across herbicides was approximately 1.0-3.0 (± 2.0) µmol L-1. Selectivity was

tested by using blank samples with no interfering peaks detected.

For halosulfuron-methyl, the mobile phase was water plus 0.1% formic acid (A) and

acetonitrile plus 0.1% formic acid (B). Formic acid was used to maintain ion strength. The
90

mobile phase followed a gradient, starting at 70% A, at 0.8 min was 10% A and held for 1.2 min,

then increased to 70% A at 2.3 min and held for 1.0 min. Flow rate was maintained at 1.0 mL

min-1 for 3.2 min, with an injection volume of 8.0 µL. The MS was run in ESI positive (+) mode

using multiple reaction monitoring (MRM) from 50 to 600 Da and single ion recording (SIR) at

435 Da. The column was sustained at ambient temperature.

For glyphosate and glufosinate, the mobile phase was water plus 0.9% formic acid (A)

and acetonitrile plus 0.9% formic acid (B). The mobile phase gradient started at 10% A, ramped

up to 60% at 2.0 min, increased to 90% A at 4.0 min. Flow rate was maintained at 0.75 mL min-

1 for 3.5 min, with an injection volume of 7.5 µL. The column was sustained at 40°C. The MS

was run in ESI negative (-) mode using multiple reaction monitoring (MRM) from 50 to 600 Da.

Single ion recording (SIR) at 168 and 180 Da for glyphosate and glufosinate, respectively.

For acetochlor and S-metolachlor, the mobile phase was water plus 0.1% formic acid (A)

and acetonitrile plus 0.1% formic acid (B). The mobile phase followed a gradient, starting at

90% A, decreasing to 10% at 2.1 min and held for 3.0 min, increasing to 90% A at 5.1 min and

held for 2.0 min. Flow rate was maintained at 1.37 mL min-1 for 7.0 min, with an injection

volume of 200 µL. The column was sustained at 25°C. The MS was run in ESI positive (+) mode

using multiple reaction monitoring (MRM) from 50 to 600 Da. Single ion recording (SIR) at 270

and 284 Da for acetochlor and S-metolachlor, respectively.

2.6. Statistical Analysis

Data for herbicide concentration on the plastic mulch, herbicide concentration in the plant

hole, and the amount of water in the plant hole were subjected to ANOVA using PROC MIXED

in SAS 9.4 (SAS Institute, Cary, NC). There were no differences between jars on the same bed

(subsamples), so data were combined (P=0.32). As there were interactions between herbicide and
91

irrigation amount (P=0.015), data were analyzed separately. Data from the plastic samples were

further analyzed for interactions between sample timings and irrigation amount. All data were

then graphed using SigmaPlot 14.0 (Systat Software, San Jose, CA).

3. Results and Discussion

3.1. Plastic Wash-Off

The ANOVA indicated interactions between herbicide and irrigation amount (0.63 and

1.27-cm) on the plastic mulch (% and mg ai or ae) ) (P=0.01) and that irrigation volume did not

influence the concentration of halosulfuron (P=0.99), glufosinate (P=0.35), and glyphosate

(P=0.85) removed from the mulch. After irrigation, halosulfuron, glufosinate, and glyphosate

concentration was 1.78, 0.11, and 0.30% (respectively) of what was applied (Table 4.3.). Figure

4.3 also depicts the effectiveness of irrigation at 0.63-cm and 1.27-cm in removing glyphosate,

glufosinate, and halosulfuron from the surface of TIF.

Previous research has shown that glyphosate can be removed from the surface of plastic

mulch with at least 1-cm of water (Grey et al. 2009). Grey et al. (2018) determined that after this

amount of irrigation the concentration of glyphosate remaining on the surface of plastic mulch

was 2.66 mg ae m-2, which is greater than our determined concentration of glyphosate remaining

on the plastic mulch (0.02 mg ae m-2) after irrigation. Hand et al. (2021) detected glyphosate at

1/1000th of a typical field rate applied over mulch after 3.5 cm of irrigation was applied.

Similarly, Culpepper et al. (2009) reported that tomato and squash could safely be planted into

plastic mulch beds treated with glyphosate, after at least 1-cm of irrigation was applied.

However, Grey et al. (2009) reported that glyphosate was still available in efficacious amounts

on LDPE mulch out to 120 h after treatment when no water was applied.
92

The relationship of glufosinate and plastic mulch is less understood than glyphosate,

however, previous research suggests the herbicide can be removed from mulch with irrigation.

Smith et al. (2017) noted that applying glufosinate over-the-top of mulch at 670 g ai ha -1 and

1,340 g ai ha-1 and transplanting bell pepper, tomato, watermelon, squash, and cucumber without

applying irrigation resulted in injury ranging from 5-75%, respectively; while the addition of an

0.83-cm irrigation event after application but prior to planting eliminated crop injury.

Grey et al. (2018) determined that after 1 cm of irrigation, the concentration of

halosulfuron remaining on the surface of plastic mulch to be 5.86 mg ai m-2. Randell et al. (2020)

noted injury and yield reductions from halosulfuron applied over mulch 1-21 d before

transplanting various vegetable crops. Additionally, the authors determined that halosulfuron can

remain active while binding to the mulch surface and slowly release when rainfall or irrigation is

applied. The amount of halosulfuron required to reduce the growth of squash varieties commonly

grown in Georgia ranges from 8.2 to 45.0 g ai ha -1 (equivalent to 0.82 to 4.5 mg ai m-2). While

the concentration of halosulfuron on the mulch before irrigation was enough to cause injury to

squash, once irrigation was applied concentration was detected well below this range (0.04 mg ai

m-2) (Table 4.3.). The behavior of halosulfuron on mulch, as determined from this experiment, is

not supported by the aforementioned previous literature. A major difference in experimental

design between studies includes the length of time between herbicide application and irrigation

event (Grey et al. 2018; Randell et al. 2020). While halosulfuron was allowed to completely dry

on the mulch, further research is warranted to explore the impact of this drying-time on the

removal of halosulfuron from the surface of plastic mulch.

Acetochlor and S-metolachlor were more difficult to remove from the mulch. The volume

of irrigation did influence herbicide concentration remaining on the mulch (P=0.04 and P<.001,
93

respectively) (Table 4.4.). Acetochlor had the highest concentration remaining on the plastic

when compared to the other herbicides, with 90.89% (143.10 mg ai m-2) and 15.16% (24.66 mg

ai m-2) after 0.63-cm and 1.27-cm of irrigation was applied, respectively. For acetochlor, the

concentration detected on the plastic mulch after 0.63-cm of irrigation was not different from the

concentration detected on the plastic mulch before irrigation was applied (Table 4.4). The

quantity of S-metolachlor remaining on the mulch after 0.63-cm and 1.27-cm of irrigation was

17.09% (24.66 mg ai m-2) and 2.84% (3.03 mg ai m-2), respectively. While there is no

information in previous literature on the movement of the chloroacetamide herbicides from the

surface of plastic mulch, injury to various crops and behavior in soil when applied directly to soil

has been studied (Grey and Vencill 2011; Ferebee et al. 2019; Mueller et al. 1999; Song et al.

2006).

The movement of the herbicides from the surface of the plastic mulch was directly related

to water solubility (Table 4.1.). Glyphosate and glufosinate removal from the surface of the

much was as expected, given that both herbicides are very water soluble and experience

negligible photodegradation losses (Isaac et al. 2006; Grey and Vencill 2011; Culpepper et al.

2009; Culpepper 2020). The concentration of halosulfuron removed from the surface of plastic

mulch was less than glyphosate and glufosinate, but greater than the chloroacetamide herbicides.

Halosulfuron is less water soluble than the contact herbicides, with a water solubility of 1650 mg

L-1 (Shaner 2014). S-metolachlor removal was greater than acetochlor but less than the

aforementioned herbicides. Acetochlor has the lowest water solubility (223 mg L -1) and

subsequently had the lowest concentration removed from the surface of the plastic mulch

(Wolejko et al. 2017). Previous research has demonstrated that damage could occur to
94

subsequent crops from contact with herbicide residue left on plastic mulch (Culpepper et al.

2009; Gilreath et al. 2006).

3.2. Wash-Off Into Transplant Hole

The average amount of water that ran into each individual jar, or transplant hole, was

equivalent to 62 (± 2.03 mL) and 134 mL (± 6.81 mL) for the 0.63-cm and 1.27-cm irrigation

treatments, respectively (Figure 4.4). There was uniformity of irrigation and even distribution of

water across replications and experiment runs.

The ANOVA indicated interactions between herbicide and irrigation amount (0.63-cm

and 1.27-cm) for total concentration in the transplant hole (P=0.03). All herbicides were detected

below 1.0 mg ai or ae in the transplant hole, despite irrigation amount (Table 4.5; Figure 4.5;

Figure 4.6).

While 1.0 mg ai or ae in the transplant hole seems low, correlating these values to g ai or

ae ha-1 highlighted how herbicide movement into the transplant hole can cause concentrations to

be greater than the applied field rate. Table 4.5 represents the total amount of herbicide that

washed into the transplant hole compared to the field rate applied (as described in Table 4.1).

After 0.63-cm and 1.27-cm irrigation volume was applied, halosulfuron was detected at 8.9X and

5.3X field rates. The extremely concentrated amount of halosulfuron that washed into the

transplant hole supports the earlier notion that halosulfuron was not given adequate time to dry

and subsequently adsorb onto the surface of the mulch, allowing it to move with a single

irrigation event into the transplant hole. The amount of glyphosate that washed into the

transplant hole after 0.63-cm and 1.27-cm irrigation was equivalent to a 3.3X and 1.9X field rate.

Glufosinate was the only herbicide that the total amount of herbicide in the transplant hole was

not different between the 0.63-cm and 1.27-cm irrigation volumes (Table 4.5; Figure 4.5).
95

Concentrations of glufosinate in the transplant hole were equivalent to a 1.3X field rate or

approximately 844 g ai ha-1. The amount of S-metolachlor in the transplant hole ranged from

0.10-0.17 mg ai, which is equivalent to a 0.39X-0.66X field rate. Acetochlor had the lowest

concentration despite irrigation volume, with 0.01-0.03 mg ai or a 0.03-0.09X field rate washing

into each transplant hole. The low amount of acetochlor washing into the transplant hole is

promising, however, further research is needed to determine the potential for use in vegetable

production.

Figure 4.6 represents the average herbicide concentration in the transplant hole area after

0.63-cm (a) and 1.27-cm (b) of irrigation was applied, along with the average amount of water

caught per jar for each respective herbicide. In Figure 4.6, the amount of water caught per jar is

similar (within the same irrigation volume) while herbicide concentrations vary. This could be

attributed to each herbicide’s ability to move from the plastic mulch surface with the irrigation

water, since the more water-soluble herbicides were detected in greater amounts (Table 4.1;

Table 4.5; Figure 4.5; Figure 4.6).

4. Conclusion

To quantitate herbicide concentration moving from a treated plastic mulch into vegetable

transplant holes when rainfall or irrigation occurs, simulated vegetable beds were designed and

constructed. The simulated vegetable bed design adds a level of mobility and controllability that

could not be achieved in field experiments and allows for rapid testing of numerous herbicides.

Halosulfuron, glufosinate, and glyphosate, on the surface of the plastic mulch, were

detected below 2% after irrigation volumes of 0.63-cm or 1.27-cm were implemented. Irrigation

was less effective in removing the chloroacetamide herbicides from the surface of the plastic
96

mulch. Acetochlor was the most stable with 91% and 15% of the herbicide remaining on the

mulch after 0.63-cm and 1.27-cm, respectively, were implemented. For S-metolachlor, 17% and

3% remained on the mulch at the aforementioned irrigation volumes, respectively.

All herbicides were detected below 1.0 mg ai or ae in the transplant hole area despite

irrigation amount. Correlating the total amount of herbicide that washed into the transplant hole

(mg ai or ae) to the field rate applied highlighted how herbicide movement into the transplant

hole can cause concentrations to be greater than the applied field rate.

The amount of water caught per jar was similar while herbicide concentrations varied,

which could be correlated to water-solubility, since the more water-soluble herbicides were

detected in greater amounts. Ranking the herbicides based on the amount of herbicide in the

transplant hole across irrigation amount, as well as water-solubility, gives: glyphosate >

glufosinate > halosulfuron and S-metolachlor > acetochlor. However, ranking the herbicides

based on the field rate detected in the transplant hole gives: halosulfuron > glyphosate >

glufosinate > S-metolachlor > acetochlor. Further research is needed to replicate these results in

field and determine the implications of injury from the determined field rates detected in the

transplant hole.
97

Acknowledgements: The authors would like to thank Sidney Cromer and Samantha Bowen for

their technical support.

Funding: This research received no external funding

Conflicts of Interest: The authors declare no conflict of interest.


98

References

Anonymous. Dual Magnum® Herbicide product label. Syngenta Crop Protection: Greensboro,

NC, 2015

Anonymous. Sandea® Herbicide product label. Gowan Company: Yuma, AZ, 2017a

Anonymous. Liberty® 280 SL Herbicide product label. Bayer CropScience: St. Louis, MO,

2017b

Anonymous. Roundup PowerMax® Herbicide product label. Bayer CropScience: St. Louis,

MO, 2020a

Anonymous. Warrant Herbicide product label. Bayer CropScience: St. Louis, MO, 2020b

Culpepper, A.S.; Grey, T.L.; Webster, T.M. Vegetable response to herbicides applied to low-

density polyethylene mulch prior to transplant. Weed Technol 2009, 23:444-449

Culpepper, A.S. Commercial Vegetable Production. In Georgia Pest Management

Handbook; Univ of Georgia Ext Sp Bull: Athens, GA, 2020, 28:428-497

Dickerson, G.W. Commercial vegetable production with plastic mulches. New Mexico State

Univ Coop Ext Serv Guide: Las Cruces, NM, 2007, H-245: 1-4

Ferebee, J.H.; Cahoon, C.W.; Besancon, T.E.; Flessner, M.L.; Langston, D.B.; Hines, T.E.;

Blake, H.B.; Askew, M.C. Fluridone and acetochlor cause unnaceptable injury to pumpkin.

Weed Technol 2019, 33:748-756

Gilreath, J.P.; Santos, B.M. Herbicide dose and incorporation depth in combination with 1,3-

dichloroprpene plus chloropicrin for Cyperus rotundus control in tomato and pepper.

Crop Prot 2004, 23:205-210

Gilreath, J.P.; Bielinksi, M.S.; Duranceau, S.J. Seasonal varation of paraquat photodegradation

rate on polyethylene mulch. Weet Technol 2006, 20:315-318


99

Grey, T.L.; Bridges, D.C.; NeSmith, D.S. Transplanted pepper (Capsicum annum) tolerance to

selected herbicides and method of application. J Veg Crop Prod 2002, 8:27

Grey, T.L.; Vencill, W.K.; Mantripagada, N.; Culpepper, A.S. Residual herbicide dissipation

from soil with low density polyethylene mulch or left bare. Weed Sci 2007, 55:638-645

Grey, T.L.; Vencill, W.K.; Webster, T.M.; Culpepper, A.S. Herbicide dissipation from low

density polyethylene mulch. Weed Sci 2009, 57:351-356

Grey T.L.; Vencill, W.K. Residual Herbicide Dissipation. In Vegetable Production, Herbicides,

Theory, and Applications; Larramendy, M., Ed; InTech: Croatia, 2011

Grey, T.L.; Culpepper, A.S.; Li, X.; Vencill, W.K. Halosulfuron-methyl degradation from the

surface of low-density polyethylene mulch using analytical and bioassay techniques.

Weed Sci 2018, 66, 15-24

Hand, L.C.; Eason, K.M.; Randell, T.M.; Grey, T.L.; Richburg, J.S.; Culpepper, A.S.

Quantifying glyphosate plus 2,4-D or dicamba removal from the surface of totally

impermeable film using analytical and bioassay techniques. Weed Technol 2021, doi:

10.1017/wet.2021.2

Isaac, W.A.; Brathwaite, R.A.; Cohen, J.E.; Bekele, I. Effects of alternative weed management

strategies on Commelina diffusa Burm. Infestations in fairtrade banana in St. Vincent and

the grenadines. Crop Prot 2006, 26:1219-1225

Johnson, W.C.; Mullinix, B.G. Weed management in watermelon (Citrullus lanatus) and

cantaloupe (Cucmis melo) transplanted on ployethylene covered seedbeds. Weed Technol

2002, 16:860-866

Lamont, W.J. What are the components of a plasticulture system? Hort Technol 1996, 6:150-154

Mueller, T. C.; Senseman, S. C. Methods related to herbicide dissipation or degradation


100

under field or laboratory conditions. Weed Sci Sp Iss 2015, 133-139

Mueller, T.C.; Shaw, D.R.; Witt, W.W. Relative dissipation of acetochlor, alachlor, metolachlor,

and SAN 582 from three surface soils. Weed Technol 1999, 13:341-346

Nyoike, T.W.; Liburd, O.E. Reusing plastic mulch for a second strawberry crop: effects on

arthropod pests, weeds, diseases, and strawberry yields. FL Ent 2014, 97:928-936

Randell, T.M.; Vance, J.C.; Culpepper, A.S. Broccoli, cabbage, squash and watermelon

response to halosulfuron preplant over plastic mulch. Weed Technol 2020, 34:202-207

Sanders, D.; Granberry, D.; Cook, W.P. Plasticulture for commercial vegetables. NC State Ext

Bull: Raleigh, NC, 1996, AG-489

Shaner, D.L. Herbicide Handbook, 10th ed.; Weed Science Society of America: Lawrence,

KS, 2014

Sharpe, S.M.; Boyd, N.S. Utility of glufosinate in postemergence row middle weed control in

Florida plasticulture production. Weed Technol 2019, 33:495-502

Smith, J.C.; Culpepper, A.S.; Stewart, K.; Rucker, K. Vegetable response to glufosinate applied

preplant over mulch or applied in row middles. Proc South Weed Sci Soc 2017, 70: 198

Song, C.; Teng, C.; Tian, L.; Ma, H.; Tao, B. Seedling growth tolerance of cucurbit crops to

herbicides stomp and acetochlor. Gen Appl Plant Phys 2006, 32:165-174

Southeastern Vegetable Extension Workers (SEVEW). Mulches and Row Covers. In

Southeastern U.S. Vegetable Crop Handbook. Kemble, J.M., Ed; 2020, 21-23

Vencill, W.K.; Richburg, J.S.; Wilcut, J.W.; Hawf, L.R. Effect of MON-12037 on purple

(Cyperus rotundus) and yellow (Cyperus esculentus) nutsedge. Weed Technol 1995,

9:148-152

Webster, T.M. Weed survey-southern states: vegetable, fruit and nut crops subsection.
101

Proc South Weed Sci Soc 2002, 55:237-258

Webster, T.M.; Culpepper, A.S. Halosulfuron has a variable effect on cucurbit growth and yield.

Hort Sci 2005, 40:1796-1800

Whychen, V.L. 2019 survey of the most common and troublesome weeds in broadleaf crops,

fruits, and vegetables in the United States and Canada. Weed Sci Soc of America Ntl

Weed Survey Dataset 2019

Wolejko, E.; Kaczynski, P.; Lozowicka, B.; Wydro, U.; Borusiewicz, A.; Hrynko, I.; Konecki,

R.; Snarska, K.; Dec, D.; Malinowski, P. Dissipation of S-metolachlor in plant and soil

and effect on enzymatic activities. Env Monit Assess 2017, 189:355


102

Tables:

Table 4.1. Detailed descriptions of the five herbicide treatments. 1

Common Name Rate Trade Name Manufacturer Water Solubility (mg L-1)
halosulfuron-methyl 53 g ai ha-1 Sandea® Gowan Company 1650
glufosinate 656 g ai ha-1 Liberty® 280 SL BASF Corporation 1370000
glyphosate 867 g ae ha-1 Roundup Powermax® Bayer Crop Science 15700 (acid); > 900000 (salt)
acetochlor 1051 g ai ha-1 Warrant® Bayer Crop Science 223
S-metolachlor 799 g ai ha-1 Dual Magnum® Syngenta Crop Protection, LLC 488
1
References: Anonymous 2015; Anonymous 2017a; Anonymous 2017b; Anonymous 2020a; Anonymous 2020b.
103

Table 4.2. Environmental measurements recorded at the time of each herbicide application.

Run 1 Run 2
Herbicide Irr1 (cm) Temp (°C) WS (kph) RH (%) CC (%) Temp (°C) WS (kph) RH (%) CC (%)

Halosulfuron 0.63 35.4 1.8 59.2 0.0 32.0 1.3 73.7 0.0

Glufosinate 0.63 33.5 1.1 58.3 50.0 29.8 1.1 69.8 0.0

Glyphosate 0.63 28.9 2.3 83.7 0.0 35.9 1.6 53.6 10.0

Acetochlor 0.63 28.5 2.9 77.1 0.0 28.5 1.8 84.3 100.0

S-metolachlor 0.63 30.3 1.8 70.0 0.0 34.6 1.8 64.2 10.0

Halosulfuron 1.27 36.4 1.1 46.7 10.0 27.8 1.4 86.4 5.0

Glufosinate 1.27 32.2 1.6 74.0 0.0 34.2 1.3 82.5 100.0

Glyphosate 1.27 33.9 2.9 78.0 10.0 32.6 1.8 60.4 5.0

Acetochlor 1.27 33.3 1.4 39.9 20.0 34.9 3.1 54.3 30.0

S-metolachlor 1.27 36.1 1.8 55.5 10.0 31.4 1.9 73.5 60.0

1
Abbreviations: Irr, irrigation; Temp, temperature; WS, wind speed; RH, relative humidity; CC, cloud cover.
104

Table 4.3. Herbicide concentration (mg ai or ae m-2 and %) on plastic mulch as affected by

sample timing.1

Halosulfuron Glufosinate Glyphosate


– (mg ai or ae m-2) –

Before Irrigation 26.04 a2 40.36 a 73.73 a

After Irrigation 0.51 b 0.04 b 0.02 b

– (%) –

Before Irrigation 99.873 a2 100.03 a 99.99 a

After Irrigation 1.78 b 0.11 b 0.03 b


1
Concentrations represents the mean of two irrigation amounts (0.63 and 1.27-cm) with the
experiments conducted twice and combined.
2
Values for each herbicide within the same column followed by the same letter are not
significantly different at P<0.05.
3
Values, as %, were based on the average amount (mg ai m-2) on the plastic mulch before
irrigation for each herbicide at each irrigation amount.
105

Table 4.4. Concentration (mg ai m-2 or %) of acetochlor and S-metolachlor on plastic mulch as

affected by irrigation amount and sample timing. 1

Irrigation (cm) Acetochlor S-metolachlor


– (mg ai m-2) –
0.63 157.44 a3 114.23 a
Before Irrigation
1.27 162.63 a 106.37 a

0.63 143.10 a 24.65 b


After Irrigation
1.27 24.66 b 3.03 c

– (%) –
0.63 100.002 a3 100.00 a
Before Irrigation
1.27 99.99 a 100.00 a

0.63 90.89 a 17.09 b


After Irrigation
1.27 15.16 b 2.84 c
1
Sample timings are before or after irrigation was applied to the simulated vegetable beds.
Irrigation amounts were 0.63 and 1.27-cm. Concentrations represents the mean of two
irrigation amounts, with the experiment conducted twice and combined.
2
Values, as %, were based on the average amount (mg ai m-2) on the plastic mulch before
irrigation for each herbicide at each irrigation amount.
3
Values for each herbicide within the same column followed by the same letter are not
significantly different at P<0.05.
106

Table 4.5. Herbicide concentration in the transplant hole area as affected by irrigation amount. 1

Irrigation Halosulfuron Glufosinate Glyphosate Acetochlor S-Metolachlor


– (cm) – – (mg ai or ae per transplant hole) –
0.63 0.152 a3 0.28 a 0.92 a 0.03 a 0.17 a
1.27 0.09 b 0.26 a 0.52 b 0.01 b 0.10 b

– (cm) – – (g ai or ae ha-1) –

0.63 4694 a 875 a 2875 a 94 a 531 a


1.27 281 b 812 a 1625 b 31 b 313 b

– (cm) – – (field rate of X) –

0.63 8.855 a 1.33 a 3.32 a 0.09 a 0.66 a


1.27 5.30 b 1.24 a 1.87 b 0.03 b 0.39 b
1
Concentrations represent the mean of 14 jars per bed and three replications per treatment, with the experiment conducted
twice and combined.
2
Concentrations were individually adjusted for the volume of water (mL) caught in the jar before statistical analysis.
3
Values for each herbicide within the same column followed by the same letter are not significantly different at P<0.05.
4
Concentrations as g ai or ae ha-1 are on a surface basis; Assumptions: hole diameter = 64-mm; hole surface area = 32-cm2.
5
Field rate (g ai or ae ha-1) is based on X of: halosulfuron, 53; glufosinate, 656; glyphosate, 867; acetochlor, 1051; S-
metolachlor, 799.
107

Figures:

Figure 4.1. Bed construction without carpet or plastic mulch covering. Design of the simulated

bed was created to mimic a soil bed used in pepper production, including a slope of 6.7% from

the middle to each side of the bed, in comparison to an actual field bed.
108

Figure 4.2. Bottom view of the constructed simulated bed with 14 glass jars attached

representing a transplant crop hole. Top row are jars 1-7 (left side of bed) and the bottom row are

jars 8-14 (right side of jars), with jars being opposite going down the length of the bed (i.e., jars

1 and 8 are across from each other, etc.). Holes were spaced 30.5 cm apart from the center of one

hole to the center of another and 20.3 cm from the center of one hole to the edge of the plywood
109

(A)

(B)

Figure 4.3. The concentration (mg ai or ae) of halosulfuron-methyl, glufosinate, glyphosate, and

acetochlor, and S-metolachlor detected on the plastic surface as affected by sample timing

(before or after irrigation was applied) and irrigation volume: (A) 0.63 cm; (B) 1.27-cm. Bars

represent the respective herbicide concentration remaining on the plastic surface (mg ai or ae m -
2
); averaged over three replications and combined over two runs. Error bars represent the

standard errors of the means (P<0.05).


110

200
Amount of Water (mL Jar-1)

150

100

50

0
0.63 1.27
Irrigation Amount (cm)

Figure 4.4. The average amount of water (mL) caught in a jar (crop transplant hole) from the

0.63-cm and 1.27-cm irrigation event. Bars represent the average mean of 14 jars per bed, three

replications, and two runs. Error bars represent the standard errors of the means (P<0.05).
111

Herbicide Concentration in Hole Area (mg ai hole-1)


1.2
0.63 cm
1.27 cm

0.8

0.4

0.0
Halo Gluf Glyph Acet S-Met
Herbicide

Figure 4.5. The concentration (mg ai) of halosulfuron-methyl, glufosinate, glyphosate, and

acetochlor, and S-metolachlor detected in the transplant hole. Concentration is based on the

amount of water (L) recorded in each respective jar. The bars represent the two irrigation

amounts (0.63 and 1.27-cm), the mean of 14 jars per replication over three replications. The

experiment was conducted twice and combined. Error bars represent the standard errors of the

means (P<0.05).
112

(A)

(B)

Figure 4.6. The concentration (mg ai or ae) of halosulfuron-methyl, glufosinate, glyphosate, and

acetochlor, and S-metolachlor detected in the transplant hole area and water (mL) caught in the

transplant hole area at each irrigation amount: (A) 0.63 cm; (B) 1.27-cm. Concentration is based

on the amount of water (L) recorded in each respective jar. Round data points represent the

average amount of water (mL) caught in the hole area. Both the bars and round data points

represent the mean of 14 jars per replication over three replications. The experiment was

conducted twice and combined. Error bars represent the standard errors of the means (P<0.05).
113

CHAPTER 5

CONCLUSIONS

In the southeastern United States, a virtual plethora of weed species are encountered due

to the temperate climate of the region, making weed control essential for achieving viable profits

in production agriculture. However, the extent of residual and contact herbicide persistence in

Georgia soils and on plastic mulch has not been fully evaluated in the literature and required

further research to better understand their long-term effects.

Georgia remains the largest pecan producing state, accounting for over 30% of all planted

hectares. Weed competition can reduce growth in perennial crops by more than 50% and

interfere with cultural practices, irrigation equipment, and harvest operations. Weed control

programs in perennial crops typically rely on residual herbicides to achieve season-long control

by establishing weed-free strips between perennial species. Indaziflam and flumioxazin are two

commonly used, residual herbicides in Georgia pecan. Indaziflam has a niche for broad-spectrum

weed control with long residual activity and no reported resistant weeds. Flumioxazin is also

used to control a range of weed species, including glyphosate-resistant weeds. Despite wide use

across the US in various permanent cropping systems, there is still little information available on

the parameters affecting the soil behavior of indaziflam and flumioxazin.

The behavior of various herbicides in soil is influenced by multiple processes, including

application losses, photodegradation, adsorption, plant uptake, chemical degradation, enzymatic

degradation, microbial degradation, leaching, and mechanical removal. Soil type, organic matter
114

content, soil moisture, and pH are the most influential factors when determining herbicide fate.

When evaluating herbicide behavior in the environment and soil, it is important to remember that

not only is the parent compound dissipating, but the products of the degradation processes are

also dissipating and moving in the soil. Most of the described factors and processes are

intertwined and impact each other. The combination of environmental and soil properties creates

a dynamic system that can greatly impact herbicide behavior and persistence.

In conclusion, indaziflam and flumioxazin soil dissipation followed first-order kinetics

and was adequately described by the exponential decay equation. Indaziflam half-life in

Greenville sandy clay loam and Faceville loamy sand were 96 and 78 days, respectively.

Flumioxazin half-life in Greenville sandy clay loam and Faceville loamy was 26.6 d. Indaziflam

half-life and clay content (%) had a direct relationship, indicating sorption of indaziflam

increases with increasing clay content. There was an inverse correlation between flumioxazin

concentration, rainfall, and solar radiation. Flumioxazin is known to stay in soil solution, which

can lead to degradation through hydrolysis or photolysis. Indaziflam half-life and microbial

biomass had an inverse relationship, indicating increased degradation of indaziflam with

increasing activity or population size soil microorganisms. While there was no direct correlation

between flumioxazin concentration and soil temperature, flumioxazin dissipation is known to be

directly influenced by soil microbial activity. Microbial degradation is influenced by soil

temperature, with increases in temperature decreasing the rate of degradation.

The activation energy for a reaction provides a measurement of how the rate constant (k)

for that reaction will be affected by a change in temperature. When indaziflam in solution was

exposed to temperatures ranging from 20 to 70°C, for up to 672 hours, data indicated that

temperature had no influence on molecular stability over time. Relating indaziflam molecular
115

stability laboratory experiments to the field dissipation studies indicates that indaziflam

persistence in soil over time is not directly influenced by diurnal and seasonal changes in soil

temperature. Flumioxazin was determined to have an activation energy of 58.4 (± 1.2) kJ mol-1.

Even at the lowest amount of solar radiation and soil temperature, the energy from these

environmental measures exceeded the activation energy needed for flumioxazin degradation.

Overall, flumioxazin stability in aqueous solution related to field dissipation indicated that as the

input of thermal energy increases, flumioxazin degradation increases.

Vegetable production on plastic mulch in Georgia includes over 10 crops with a farm

gate value exceeding $596 million. These mulched systems often combine the use of fumigation,

drip tape, raised beds, and plastic mulch where 3 to 5 high-value crops are produced over a two-

year period. With the loss of methyl bromide, herbicides applied over plastic mulch prior to

planting a crop have become essential to maintain weeds, but care must be taken to avoid crop

damage. To keep the integrity of the mulch intact across the multiple crops and over time, weed

control and termination of the previous crop is crucial. Herbicides have become an integral

component in both controlling escapes of both weeds and the previous crop.

To control nutsedge and other weed species from damaging the mulch, producers

can apply herbicides on-top of and under the mulch. Halosulfuron-methyl, glufosinate,

glyphosate, acetochlor and S-metolachlor are all herbicides that are used or have the potential to

be used over-the-top of plastic mulch to mitigate previous crop and weed escapes. When

irrigating at least 0.63-cm, less than 2% of halosulfuron, glufosinate, and glyphosate remained on

the surface of the plastic mulch. In contrast, 91% and 15% of acetochlor remained on the mulch

after irrigating with 0.63-cm and 1.27-cm, respectively. S-metolachlor remained on the surface of

the plastic mulch at 17% and 3% after the aforementioned irrigation volumes, respectively. All
116

herbicide concentration was detected below 1.0 mg ai or ae in the transplant hole area despite

irrigation amount. For halosulfuron, glyphosate, and glufosinate, these concentrations were equal

to a 1.3X-8.9X field rate washing into the transplant hole. Acetochlor and S-metolachlor

concentrations in the transplant hole were equivalent to a 0.1X-0.7X field rate, respectively.

Ranking the herbicides based on the amount of herbicide in the transplant hole across irrigation

amount, as well as water-solubility, gives: glyphosate > glufosinate > halosulfuron and S-

metolachlor > acetochlor. However, ranking the herbicides based on the field rate detected in the

transplant hole gives: halosulfuron > glyphosate > glufosinate > S-metolachlor > acetochlor.
ProQuest Number: 28411930

INFORMATION TO ALL USERS


The quality and completeness of this reproduction is dependent on the quality
and completeness of the copy made available to ProQuest.

Distributed by ProQuest LLC ( 2021 ).


Copyright of the Dissertation is held by the Author unless otherwise noted.

This work may be used in accordance with the terms of the Creative Commons license
or other rights statement, as indicated in the copyright statement or in the metadata
associated with this work. Unless otherwise specified in the copyright statement
or the metadata, all rights are reserved by the copyright holder.

This work is protected against unauthorized copying under Title 17,


United States Code and other applicable copyright laws.

Microform Edition where available © ProQuest LLC. No reproduction or digitization


of the Microform Edition is authorized without permission of ProQuest LLC.

ProQuest LLC
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346 USA

You might also like