Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-19T01:18:05.441Z Has data issue: false hasContentIssue false

Spray deposition, adjuvants, and physiochemical properties affect benzobicyclon efficacy

Published online by Cambridge University Press:  19 March 2019

Chad Brabham*
Affiliation:
Postdoctoral Research Associate, Department of Crop, Soil, and Environmental Sciences, University of Arkansas, Fayetteville, AR, USA
Jason K. Norsworthy
Affiliation:
Professor and Elms Farming Chair of Weed Science, Department of Crop, Soil, and Environmental Sciences, University of Arkansas, Fayetteville, AR, USA
Craig A. Sandoski
Affiliation:
Southern Region Field Development, Gowan USA, Collierville, TN, USA
Vijay K. Varanasi
Affiliation:
Postdoctoral Research Associate, Department of Crop, Soil, and Environmental Sciences, University of Arkansas, Fayetteville, AR, USA
Lauren M. Schwartz-Lazaro
Affiliation:
Assistant Professor, Department of Plant, Environmental and Soil Sciences, Louisiana State University, Baton Rouge, LA, USA
*
Author for correspondence: Chad Brabham, Email: cbrabham@uark.edu
Rights & Permissions [Opens in a new window]

Abstract

Benzobicyclon is a new pro-herbicide being evaluated in the Midsouth United States as a post-flood weed control option in rice. Applications of benzobicyclon to flooded rice are necessary for efficacious herbicide activity, but why this is so remains unknown. Two greenhouse experiments were conducted to explore how herbicide placement (foliage only, flood water only, foliage and flood water simultaneously) and adjuvants (nonionic surfactant, crop oil concentrate, and methylated seed oil [MSO]) affect herbicide activity. The first experiment focused on importance of herbicide placement. Little to no herbicidal activity (<18% visual control) was observed on two- to four-leaf barnyardgrass, Amazon sprangletop, and benzobicyclon-susceptible weedy rice with benzobicyclon treatments applied to weed foliage only. In contrast, applications made only to the flood water accounted for >82% of the weed control and biomass reduction achieved when benzobicyclon was applied to flood water and foliage simultaneously. The second experiment concentrated on adjuvant type and benzobicyclon efficacy when applied to foliage and flood water simultaneously. At 28 days after treatment, benzobicyclon alone at 371 g ai ha−1 provided 29% and 67% control of three- to five-leaf barnyardgrass and Amazon sprangletop, respectively. The inclusion of any adjuvant significantly increased control, with MSO providing near-complete control of barnyardgrass and Amazon sprangletop. Furthermore, we used the physiochemical properties of benzobicyclon and benzobicyclon hydrolysate to derive theories to explain the complex activity of benzobicyclon observed in our study and in field trials. Benzobicyclon applications should contain an oil-based adjuvant and must be applied to flooded rice fields for optimal activity.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© Weed Science Society of America, 2019

Introduction

In the United States from 2006 to 2017, rice was planted on an average of 1.16 million ha, 75% of which was located in the Midsouth (Arkansas, Louisiana, Missouri, and Mississippi) (USDA-NASS 2018). The majority of rice in the Midsouth is grown in a delayed-flood rice system (Hardke Reference Hardke2014; Saichuk Reference Saichuk2014). Here, rice is planted into a dry seedbed and once rice reaches the four- to six-leaf stage, a permanent flood is established. In contrast, California implements a water-seeded rice system and uses a continuous flood or establishes a flood once rice has pegged (Strand et al. Reference Strand, Espino, Fischer, Godfrey, Greer, Hill, Marsh and Mutters2013). In both systems, the establishment of a permanent flood is an important cultural weed management practice; however, herbicides are still needed for effective season-long weed control. The prevalent weeds in both systems are Echinochloa spp., Leptochloa spp., Cyperaceae species, ducksalad [Heteranthera limosa (Sw.) Willd.], weedy rice, jointvetches (Aeschynomene spp.), hemp sesbania [Sesbania herbacea (Mill.) McVaugh], and more aquatic-adapted weeds in the water-seeded systems (Norsworthy et al. Reference Norsworthy, Bond and Scott2013; Saichuk Reference Saichuk2014; Strand et al. Reference Strand, Espino, Fischer, Godfrey, Greer, Hill, Marsh and Mutters2013). Although the weed species may differ between water- and dry-seeded rice, herbicide resistance is an ever-increasing problem in both systems, and new herbicide sites of action (SOA) are needed.

Benzobicyclon (3-[2-chloro-4-(methylsulfonyl)benzoyl]-4-(phenylthio)bicyclo[3.2.1]oct-3-en-2-one) is a new herbicide in the United States and will soon be marketed as a post-flood weed control option (Sandoski et al. Reference Sandoski, Brazzle, Holmes and Takahashi2014). Benzobicyclon is a pro-herbicide; in a nonenzymatic hydrolytic reaction, it is converted to the active ingredient benzobicyclon hydrolysate (Sekino Reference Sekino2002; Williams and Tjeerdema Reference Williams and Tjeerdema2016). Benzobicyclon hydrolysate, a triketone, is a potent inhibitor of 4-hydroxyphenylpyruvate dioxygenase (HPPD; WSSA Group 27) (Kakidani and Hirai Reference Kakidani and Hirai2003; Sekino Reference Sekino2002) and is a new SOA in U.S. rice. Benzobicyclon has been commercially available in Asia since 2001, where it is used at rates from 200 to 300 g ai ha−1 for PRE to early POST weed control in transplanted and water-seeded rice (Komatsubara et al. Reference Komatsubara, Sekino, Yamada, Koyanagi and Nakahara2009). In the United States, benzobicylon plus halosulfuron (Butte®, Gowan, Yuma, AZ) was recently registered in California as a slow-release granular for use in flooded rice from seeding to four-leaf rice, with benzobicyclon applied at 257 to 297 g ha−1 (Anonymous 2017b). In the Midsouth, benzobicyclon is expected to be formulated as a soluble concentrate (Rogue®, Gowan) or in a water-dispersible granular premix with halosulfuron (Rogue Plus®, Gowan) at a proposed benzobicyclon use rate of 247 to 371 g ha−1 (Sandoski et al. Reference Sandoski, Brazzle, Holmes and Takahashi2014).

Benzobicyclon has POST activity on several problematic and acetolactate synthase inhibitor–resistant grass, sedge, broadleaf, and aquatic weeds commonly found in rice fields, and it is especially effective on sprangletops, annual sedges, and ducksalad (Komatsubara et al. Reference Komatsubara, Sekino, Yamada, Koyanagi and Nakahara2009; McKnight Reference McKnight2017; Young et al. Reference Young, Norsworthy and Scott2018b). Flood establishment in relation to benzobicyclon application timing and flood depth is an important factor in maximizing benzobicyclon efficacy (McKnight Reference McKnight2017; Norsworthy et al. Reference Norsworthy, Sandoski and Scott2014; Young Reference Young2017). For example, in a greenhouse trial, Young et al. (Reference Young, Norsworthy and Scott2018b) showed control of one- to two-leaf Amazon sprangletop, barnyardgrass, rice flatsedge (Cyperus iria L.), and yellow nutsedge (C. esculentus L.) was improved an average of 47 percentage points to 92% when applied to a 5- or a 15-cm flood versus applications to a saturated field soil (45% control). In addition, McKnight (Reference McKnight2017) and Norsworthy et al. (Reference Norsworthy, Sandoski and Scott2014) found benzobicyclon was most effective when applied 24 h after the establishment of a continuous flood at pegging rice or five-leaf rice, respectively. The half-life (i.e., hydrolysis rate) of benzobicyclon in water at a pH of 7 and a temperature of 25 C is 15 h; under these conditions, the flood would need to be held for >4 d for benzobicyclon to undergo complete hydrolysis (Williams and Tjeerdema Reference Williams and Tjeerdema2016). This further highlights the importance of the flood holding period in relation to application timing. Together, these trials revealed the necessity of a flood for benzobicyclon activity. However, applications in the Midsouth will be made later in the growing season; thus, there is a higher probability for benzobicyclon-containing spray droplets to come into direct contact with foliage. Therefore, research is needed to understand the uptake and translocation of benzobicyclon and the potential for benzobicyclon to be converted to the hydrolysate form at the leaf surface or within cellular content.

For POST-applied herbicides to reach their cellular target site, they must first cross multiple barriers (i.e., cuticle, cell wall, plasma membrane, and organelle membranes) and, if necessary, they must be translocated to the SOA. Most herbicides cross membranes by simple diffusion (i.e., down a concentration gradient); however, the physicochemical properties of an herbicide can significantly affect the rate at which this occurs (Sterling Reference Sterling1994). Two important physicochemical properties that affect herbicide uptake and translocation are the octanol/water partitioning coefficient (log Kow) and the acid dissociation constant (pKa) (Hsu and Kleier Reference Hsu and Kleier1996; Hsu et al. Reference Hsu, Marxmiller and Yang1990; Zhang et al. Reference Zhang, Lorsbach, Castetter, Lambert, Kister, Wang, Klittich, Roth, Sparks and Loso2018). The log Kow represents the lipophilic nature of a compound, and for most herbicides, the log Kow value ranges from −4 to 4 (more lipophilic). In general, lipophilic herbicides diffuse across lipid barriers more freely than do more hydrophilic herbicides. Furthermore, the pKa for herbicides with an ionizable group can alter the log Kow (Kleier Reference Kleier1988). The pKa is the pH at which half of the herbicide molecules are protonated and half are negatively charged. At a pH below the pKa, more of the herbicide molecules are protonated (neutral charge) and, thus, are more lipophilic. In addition to herbicide physicochemical properties, product formulations and adjuvants can improve translaminar movement (Kirkwood Reference Kirkwood1999). The objective of this investigation was 2-fold: (1) to determine the importance of benzobicyclon absorption as spray droplets versus uptake via flood water and (2) to select the appropriate adjuvant needed for optimum benzobicyclon activity.

Materials and methods

Two separate experiments were conducted in the greenhouses located at the Altheimer Laboratory at the University of Arkansas, Fayetteville. Before the experiments, seeds of barnyardgrass, Amazon sprangletop, and weedy rice were germinated in plastic trays filled with potting mix (Sunshine premix No. 1®; Sun Gro Horticulture, Bellevue, WA). The weedy rice accessions were only used in Experiment 1. Barnyardgrass seeds were obtained from Azlin Seed Service (Leland, MS); Amazon sprangletop seeds were collected from field plots at the Rice Research and Extension Center near Stuttgart, AR; and the weedy rice seed was a subsample of multiple accessions known to be susceptible to benzobicyclon (Young et al. Reference Young, Norsworthy, Scott, Bond and Heiser2018a). Four to five seedlings at the one-leaf stage were transplanted into 7.5-L buckets (24-cm diameter) three-quarters filled with a Pembroke silt loam (fine-silty, mixed, active, mesic Mollic Paleudalfs) soil. Plant growth was maintained under of 32/22 C day/night temperature regimen and a 16-h photoperiod consisting of natural lighting supplemented with a halide (Experiment 1) or light-emitting diode (Experiment 2) lighting system. In both experiments, seedlings were flooded 1 d before the application of treatments. A 5-cm flood was established, and buckets were checked daily to ensure the appropriate flood level was maintained for 28 d after treatment (DAT). The water used in all experiments had a pH of 8.06 and an electric conductivity of 216 μS cm−1.

Benzobicyclon placement experiment

In the first experiment, benzobicyclon, at the expected field use rate of 371 g ai ha−1, was applied to weed foliage only, flood water only, and both foliage and flood water. In addition, adjuvant type was tested with foliage-only treatments. The adjuvant treatments were none, nonionic surfactant (NIS; Induce, Helena Agri-Enterprises, LLC, Collierville, TN), crop oil concentrate (COC; Superb HC, WinField Solutions, LLC, Shoreview, MN), and a methylated seed oil (MSO; Leci-Tech, Loveland Products, Loveland, CO). Adjuvants were added at 1% v/v. Treatments were applied using a research track sprayer equipped with flat-fan spray nozzles calibrated to deliver 187 L ha−1 at 1.6 km h−1.

The experiment was conducted as a randomized complete block design with three replications and was repeated once. The weed species and size at application were barnyardgrass with two-to three-leaves (8- to 12-cm tall), Amazon sprangletop with three-to four-leaves (6- to 10-cm tall), and two-leaf weedy rice (10- to 15-cm tall). For the foliage-only treatments, activated charcoal at 0.6 g L−1 was added to each bucket and stirred before and after treatment to bind benzobicyclon in the flood water. In addition, aluminum foil was used to cover large swaths of buckets with no plants to prevent droplets from reaching the flood water. In a preliminary experiment, activated charcoal in the flood water had no adverse effects on plant growth. For the flood water–only treatments, plant foliage was wrapped in aluminum foil. At 28 DAT, percent control was recorded, and plant tissue was harvested for dry weight. Dry weights represented the average grams per plant of four to five plants from each bucket and are expressed here as percent dry weight reduction from nontreated plants.

Benzobicyclon adjuvant experiment

The second experiment focused on the efficacy of benzobicyclon (371 g ha−1) applied with or without adjuvants to foliage and flood water simultaneously. The experimental setup was similar to the benzobicyclon placement study, except there were four replications per run and a total of three runs. The adjuvants used were the same as in Experiment 1 and the treatments were none, NIS at 0.25% v/v, and COC and MSO each at 1% v/v. A nontreated control was included. The weed species and size at application were barnyardgrass with three to four leaves (8 to 18 cm tall) and Amazon sprangletop with four to five leaves (8 to 12 cm tall). Weedy rice accessions were not used in this experiment.

Before analysis, data were tested for homoscedasticity using Bartlett’s test and data were pooled across runs for each species (P > 0.05). Control data were not included in analyses. For both experiments, data were subjected to ANOVA using the MIXED procedure in SAS (SAS Institute, Cary NC) and means were separated using Fisher’s protected LSD (α = 0.05).

Results and discussion

Experiment 1: Benzobicyclon efficacy as a function of spray placement

Previous research has illustrated the necessity of a flood for benzobicyclon herbicide activity, but the reason for this phenomenon is not well understood. To explore this, benzobicyclon (371 g ha−1) was selectively applied to flood water and/or plant foliage where, under our experimental setup, we assumed benzobicyclon and the hydrolysate active form were available for uptake only via the deposition method.

At 28 DAT, benzobicyclon applied simultaneously to foliage and flood water without an adjuvant provided 88%, 69%, and 54% visual control of weedy rice, Amazon sprangletop, and barnyardgrass, respectively. Similarly, plant dry weights of weedy rice, Amazon sprangletop, and barnyardgrass were reduced by 78%, 77%, and 68%, respectively, when benzobicyclon was applied simultaneously to foliage and flood water (Table 1). When benzobicyclon was sprayed on the flood water only, the percent control and biomass reduction accounted for >82% of the total efficacy observed when benzobicyclon was applied simultaneously to flood water and foliage. For example, benzobicyclon applied to foliage and flood water simultaneously provided 69% control of 6- to 10-cm–tall Amazon sprangletop, whereas the observed control with the flood water–only treatment, although significantly different, was still 59%. In contrast, all benzobicyclon treatments applied to the foliage only, with or without adjuvants, provided <18% visual control of all weed species. However, control of all species with foliage-only treatments containing MSO was either similar to or significantly better than control obtained with foliage-only treatments with no adjuvant. The data clearly illustrate the absorption of benzobicyclon, and presumably benzobicyclon hydrolysate, through flood water is necessary and sufficient to obtain the majority of efficacy from a benzobicyclon application.

Table 1. Visual control and dry weight reduction of Amazon sprangletop, barnyardgrass, and weedy rice at 28 days after treatment with benzobicyclon applied at 371 g ai ha−1 to foliage only, flood water only, and foliage and flood water simultaneously (Experiment 1).

a Abbreviations: COC, crop oil concentrate; MSO, methylated seed oil; NIS, nonionic surfactant.

b The NIS, COC, and MSO were applied at 1% v/v.

c Mean values within a column, averaged over two trials, were separated using Fisher’s protected LSD values at an α of 0.05.

d Weedy rice accessions were collected and prescreened for benzobicyclon susceptibility in Young et al. (Reference Young, Norsworthy, Scott, Bond and Heiser2018a).

Experiment 2: Benefits of adjuvants with benzobicyclon

Although little POST activity was observed in Experiment 1 with foliage-only benzobicyclon treatments, the addition of an adjuvant did appear useful. A separate experiment was conducted to explore how the addition of NIS, COC, or MSO to benzobicyclon would improve weed control when applied to foliage and flood water simultaneously. Control of four- to five-leaf Amazon sprangletop and three- to four-leaf barnyardgrass with benzobicyclon alone at 371 g ha−1 was 67% and 27%, respectively. Research has shown benzobicyclon has considerably more activity on species of sprangletop than on barnyardgrass, and our data further highlight this point (Komatsubara et al. Reference Komatsubara, Sekino, Yamada, Koyanagi and Nakahara2009; McKnight Reference McKnight2017; Young et al. Reference Young, Norsworthy and Scott2018b; CA Sandoski, personal communication). An adjuvant effect was detected in visible control estimates and dry weight reduction of Amazon sprangletop and barnyardgrass (Table 2). For Amazon sprangletop, the addition of NIS improved control to 91%, but COC and MSO were the best adjuvant options, with visual control improved to 99%. A stair-step improvement in barnyardgrass control was observed when an adjuvant was used, compared with benzobicyclon alone (27%). The addition of NIS, COC, and MSO improved barnyardgrass control to 44%, 82%, and 96%, respectively. Overall, the addition of MSO or COC to benzobicyclon provide the greatest level of control and dry weight reduction (>85%).

Table 2. Visual control estimates and dry weight reduction of Amazon sprangletop and barnyardgrass at 28 days after treatment with benzobicyclon applied at 371g ai ha−1 to foliage and flood water with or without adjuvants (Experiment 2).

a Abbreviations: COC, crop oil concentrate; MSO, methylated seed oil; NIS, nonionic surfactant.

b COC and MSO were applied at 1% v/v and the NIS was applied at 0.25% v/v.

c Mean values within a column, averaged over three trials, were separated using Fisher’s protected LSD values at an α of 0.05.

Relationship between herbicide physiochemical properties and efficacy

The translaminar movement and subsequent translocation of agrochemicals is a highly dynamic process. However, in the absence of experimentation with radiolabeled compounds, one can sufficiently predict these characteristics using the log Kow (i.e., lipophilicity [log P]), water solubility, and pKa values of agrochemicals (Hsu and Kleier Reference Hsu and Kleier1996; Hsu et al. Reference Hsu, Marxmiller and Yang1990; Klittich and Ray Reference Klittich and Ray2013; Zhang et al. Reference Zhang, Lorsbach, Castetter, Lambert, Kister, Wang, Klittich, Roth, Sparks and Loso2018). Empirical or in silico–derived physiochemical properties of benzobicyclon and benzobicyclon hydrolysate were amassed from the literature and used to discuss observed herbicidal activity (Table 3).

Table 3. Physiochemical properties of benzobicyclon and benzobicyclon hydrolysate stated in the literature or derived in silico.

a Abbreviations: Log D, log of the octanol/water partitioning coefficient plus log of the acid dissociation constant; Log Kow, octanol/water partitioning coefficient; Log Koc, soil organic carbon/water partitioning coefficient; pKa, acid dissociation constant.

b Average of values from multiple sources.

Benzobicyclon is a highly lipophilic (log Kow), nonpolar compound that is expected to readily penetrate the cuticle and may exhibit xylem, but no phloem mobility (Hsu and Kleier Reference Hsu and Kleier1996; Hsu et al. Reference Hsu, Marxmiller and Yang1990; Tomlin 2012; Zhang et al. Reference Zhang, Lorsbach, Castetter, Lambert, Kister, Wang, Klittich, Roth, Sparks and Loso2018). In theory, this indicates benzobicyclon, if it had herbicidal properties, would act similar to a contact herbicide. However, negligible HPPD-inhibitor–like activity (i.e., bleaching) was observed in our study with foliar-only treatments, indicating benzobicyclon was not sufficiently converted to the active compound benzobicyclon hydrolysate at the leaf surface or within cellular content. It should be noted, however, that weeds in the MSO-containing treatments did exhibit noticeable necrosis at the leaf tips. In addition, benzobicyclon is predicted to have poor water solubility (0.052 mg L−1) and may readily crystalize at the leaf surface, further inhibiting uptake and hydrolytic conversion. In contrast, benzobicyclon hydrolysate is expected to have excellent phloem mobility (ion trapping) (Hsu and Kleier Reference Hsu and Kleier1996) and has a predicted distribution coefficient (log D; i.e., log Kow plus pKa) value of −1.5 at pH 7.4, which is similar to that of other triketone and pyrazole HPPD-inhibiting herbicides (0.58 > log D > −1.45 at pH 7.4) (Gandy et al. Reference Gandy, Corral, Mylne and Stubbs2015). However, benzobicyclon hydrolysate is ionizable (pKa 2.89) (Williams et al. Reference Williams, Gladfelder, Quigley, Ball and Tjeerdema2017) and will be negatively charged in spray solution and flood water; thus, it will have difficulty crossing the cuticle. In this study, benzobicyclon with or without adjuvants had little activity when spray droplets came in direct contact with leaf foliage, but the addition of an oil-based adjuvant, especially MSO, was necessary to optimize the activity of benzobicyclon when spray droplets came in contact with foliage and flood water.

This phenomenon is difficult to explain, but two schools of thought exist. First, and most obvious, is simply the addition of an oil-based adjuvant improved the penetration and thus diffusion of benzobicyclon across the cuticle. The addition of COC or MSO significantly improves the translaminar movement of lipophilic compounds and formulations (Beckett et al. Reference Beckett, Stoller and Bode1992; Grossmann and Ehrhardt Reference Grossman and Ehrhardt2007; Miller Reference Miller2017; Young and Hart Reference Young and Hart1998). Furthermore, the labels of the HPPD-inhibiting herbicides mesotrione (Anonymous 2011), tembotrione (Anonymous 2016), and topramezone (Anonymous 2017a) recommend COC or MSO as the preferred adjuvant for optimizing activity. However, benzobicyclon is a unique compound because of the slow hydrolytic conversion rate to the herbicidal active compound benzobicyclon hydrolysate (half-life approximately 15 h) (Williams and Tjeerdema Reference Williams and Tjeerdema2016), and lack of systemic movement of benzobicyclon probably limits any POST activity via direct leaf uptake.

The second theory is more abstract. Benzobicyclon has poor water solubility and can readily adsorb to soil particles in soils and in flood water (log of the soil organic carbon/water partitioning coefficient approximately 3.9, Table 3). If bound to soil particles, benzobicyclon activity would be reduced because it would be unavailable for plant uptake, and conversion to benzobicyclon hydrolysate would be significantly hindered. These physicochemical properties may help explain why McKnight (Reference McKnight2017) and Norsworthy et al. (Reference Norsworthy, Sandoski and Scott2014) found significantly reduced benzobicyclon activity when benzobicyclon was applied to saturated soil versus applied to flood water. Thus, we hypothesize an oil-based adjuvant will help keep the lipophilic compound benzobicyclon in suspension when applied to flood water and, additionally, will increase the probability benzobicyclon is converted to the systemic herbicide benzobicyclon hydrolysate. However, it cannot be ruled out that both theories work together to improve the overall efficacy of benzobicyclon. Regardless, benzobicyclon applications should contain an oil-based adjuvant and must be applied to flooded rice for optimal activity. Additional research is needed to replicate these results in the field and to determine crop safety to benzobicyclon plus MSO.

Author ORCIDs

Chad Brabham https://orcid.org/0000-0002-0076-7602

Acknowledgments

We would like to thank Dr. Cammy Willett for invaluable discussions and input on this manuscript. This research was funded in part by Gowan (here we declare a potential conflict of interest) and the Arkansas Rice Research and Promotion Board.

References

Advanced Chemistry Development (2019) Percepta Predictors. https://www.acdlabs.com/products/percepta/predictors.php. Accessed: September 1, 2019Google Scholar
Anonymous (2011) Callisto herbicide product label. Syngenta Group Company Publication No. SCP1131A-L2F0409. Greensboro, NC: Syngenta Crop Protection LLC. 8 pGoogle Scholar
Anonymous (2016) Laudis herbicide product label. Bayer Publication No. US61381463B. Research Triangle Park, NC: Bayer CropScience LP. Pages 1516Google Scholar
Anonymous (2017a) Armezon herbicide product label. BASF Corporation Publication No. NVA 2015–04–216–0157. Research Triangle Park, NC: BASF Corporation. 7 pGoogle Scholar
Anonymous (2017b) Butte herbicide product label. Gowan Company Publication No. 01–R0417. Yuma, AZ: Gowan Company. 2 pGoogle Scholar
Beckett, TH, Stoller, EW, Bode, LE (1992) Quizalofop and sethoxydim activity as affected by adjuvants and ammonium fertilizers. Weed Sci 40:1219CrossRefGoogle Scholar
ChemAxon Ltd (2019) Instant Cheminformatics Solutions. https://chemicalize.com. Accessed: September 1, 2019.Google Scholar
Gandy, MN, Corral, MG, Mylne, JS, Stubbs, KA (2015) An interactive database to explore herbicide physiochemical properties. Org Biomol Chem 13:55595832CrossRefGoogle Scholar
Grossman, K, Ehrhardt, T (2007) On the mechanism of action and selectivity of the corn herbicide topramezone: a new inhibitor of 4-hydroxyphenylpyruvate dioxygenase. Pest Manag Sci 63:429439CrossRefGoogle Scholar
Hardke, JT (2014) Arkansas Rice Production Handbook. Arkansas Cooperative Extension Service Miscellaneous Publications 192. Little Rock, AR: University of ArkansasGoogle Scholar
Hsu, FC, Kleier, DA (1996) Phloem mobility of xenobiotics VII. A short review. J Exp Bot 47:12651271CrossRefGoogle Scholar
Hsu, FC, Marxmiller, RL, Yang, AYS (1990) Study of root uptake and xylem translocation of cinmethylin and related compounds in detopped soybean roots using a pressure chamber technique. Plant Physiol 93:15731578CrossRefGoogle ScholarPubMed
Kakidani, H, Hirai, K (2003) Three-dimensional modeling of plant4-hydroxyphenylpyrvate dioxygenase, a molecular target of triketone-type herbicides. J Pestic Sci 28:409415CrossRefGoogle Scholar
Kirkwood, RC (1999) Recent developments in our understanding of the plant cuticle as a barrier to the foliar uptake of pesticides. Pestic Sci 55:69773.0.CO;2-H>CrossRefGoogle Scholar
Kleier, DA (1988) Phloem mobility of xenobiotics: mathematical model unifying the weak acid and intermediate permeability theories. Plant Physiol 86:803810CrossRefGoogle ScholarPubMed
Klittich, CJR, Ray, SL (2013) Effects of physical properties on the translaminar activity of fungicides. Pestic Biochem Phys 107:351359CrossRefGoogle ScholarPubMed
Komatsubara, K, Sekino, K, Yamada, Y, Koyanagi, H, Nakahara, S (2009) Discovery and development of a new herbicide, benzobicyclon. J Pestic Sci 34:113114CrossRefGoogle Scholar
McKnight, B (2017) Activity of benzobicyclon herbicide in common Louisiana rice production practices. Ph.D dissertation. Baton Rouge, LA: Louisiana State UniversityGoogle Scholar
Miller, MR (2017) Characterization of candidate compound for control of barnyardgrass and other troublesome weeds in rice. Ph.D dissertation. Fayetteville, AR: University of Arkansas. Pages:65–67.Google Scholar
Norsworthy, JK, Bond, J, Scott, RC (2013) Weed management practices and needs in Arkansas and Mississippi rice. Weed Technol 27:623630CrossRefGoogle Scholar
Norsworthy, JK, Sandoski, CA, Scott, RC (2014) A review of benzobicyclon trials in Arkansas rice. Page 99 in Proceedings of the 35th Rice Technical Working Group. New Orleans, LA: Louisiana State University Agricultural CenterGoogle Scholar
Saichuk, J (2014) Louisiana Rice Production Handbook. Agricultural Center Publication 2321-/14 rev. Baton Rouge, LA: Louisiana State UniversityGoogle Scholar
Sandoski, CA, Brazzle, JR, Holmes, KA, Takahashi, A (2014) Benzobicyclon: a novel herbicide for U.S rice production. Page 99 in Proceedings of the 35th Rice Technical Working Group. New Orleans, LA: Louisiana State University Agricultural CenterGoogle Scholar
Sekino, K (2002) Discovery study of new herbicides form the inhibition of photosynthetic pigments biosynthesis: development of a new plastoquinone biosynthetic inhibitor, benzobicyclon as a herbicide. J Pestic Sci 27:388391CrossRefGoogle Scholar
Sterling, TM (1994) Mechanisms of herbicide absorption across plant membranes and accumulation in plant cells. Weed Sci 42:263267CrossRefGoogle Scholar
Strand, LL, Espino, LA, Fischer, AJ, Godfrey, LD, Greer, CA, Hill, JE, Marsh, RE, Mutters, RG (2013) Integrated Pest Management for Rice, 3rd ed. University of California Agriculture and Natural Resources Publication 328. Oakland, CA: University of CaliforniaGoogle Scholar
Tomlin CDS, ed (2012) The Pesticide Manual: A World Compendium. 16th ed. Hampshire, UK: British Crop Production Council. 96 pGoogle Scholar
[USDA-NASS] U.S. Department of Agriculture (2018) National Agricultural Statistics Service: Quick Stats. https://www.nass.usda.gov/Quick_Stats/index.php. Accessed: May 20, 2018Google Scholar
Williams, KL, Gladfelder, JJ, Quigley, LL, Ball, DB, Tjeerdema, RS (2017) Dissipation of the herbicide benzobicyclon hydrolysate in a model California rice field soil. J Agric Food Chem 65:92009207CrossRefGoogle Scholar
Williams, KL, Tjeerdema, RS (2016) Hydrolytic activation kinetics of the herbicide benzobicyclon in simulated aquatic systems. J Agric Food Chem 64:48384844CrossRefGoogle ScholarPubMed
Young, BG, Hart, SE (1998) Optimizing foliar activity of isoxaflutole on giant foxtail (Setaria faberi) with various adjuvants. Weed Sci 46:397402CrossRefGoogle Scholar
Young, ML (2017) Evaluation of benzobicyclon for use in midsouthern rice (Oryza sativa) systems. MS thesis. Fayetteville, AR: University of ArkansasGoogle Scholar
Young, ML, Norsworthy, JK, Scott, RC (2018b) Optimizing benzobicyclon efficacy in drill-seeded rice. Adv Crop Sci Tech 6:339Google Scholar
Young, ML, Norsworthy, JK, Scott, RC, Bond, JA, Heiser, J (2018a) Benzobicyclon as a post-flood option for weedy rice control. Weed Technol 32:371378CrossRefGoogle Scholar
Zhang, Y, Lorsbach, B, Castetter, S, Lambert, WT, Kister, J, Wang, NX, Klittich, C, Roth, J, Sparks, TC, Loso, MR (2018) Physiochemical property guidelines for modern agrochemicals. Pest Manag Sci 74:19791991CrossRefGoogle Scholar
Figure 0

Table 1. Visual control and dry weight reduction of Amazon sprangletop, barnyardgrass, and weedy rice at 28 days after treatment with benzobicyclon applied at 371 g ai ha−1 to foliage only, flood water only, and foliage and flood water simultaneously (Experiment 1).

Figure 1

Table 2. Visual control estimates and dry weight reduction of Amazon sprangletop and barnyardgrass at 28 days after treatment with benzobicyclon applied at 371g ai ha−1 to foliage and flood water with or without adjuvants (Experiment 2).

Figure 2

Table 3. Physiochemical properties of benzobicyclon and benzobicyclon hydrolysate stated in the literature or derived in silico.