Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-26T21:53:04.679Z Has data issue: false hasContentIssue false

4-Hydroxyphenylpyruvate dioxygenase (HPPD)-inhibiting herbicides: past, present, and future

Published online by Cambridge University Press:  21 October 2022

Amit J. Jhala*
Affiliation:
Associate Professor, Department of Agronomy and Horticulture, University of Nebraska-Lincoln, Lincoln, NE, USA
Vipan Kumar
Affiliation:
Assistant Professor, Agricultural Research Center at Hays, Kansas State University, Hays, KS, USA
Ramawatar Yadav
Affiliation:
Postdoctoral Research Associate, Department of Agronomy, Iowa State University, Ames, IA, USA
Prashant Jha
Affiliation:
Professor, Department of Agronomy, Iowa State University, Ames, IA, USA
Mithila Jugulam
Affiliation:
Professor, Department of Agronomy, Kansas State University, Manhattan, KS, USA
Martin M. Williams II
Affiliation:
Ecologist, Global Change and Photosynthesis Research, U.S. Department of Agriculture–Agricultural Research Service, Urbana, IL, USA
Nicholas E. Hausman
Affiliation:
Agricultural Science Technician, Global Change and Photosynthesis Research, U.S. Department of Agriculture–Agricultural Research Service, Urbana, IL, USA
Franck E. Dayan
Affiliation:
Professor, Department of Soil and Crop Sciences, Department of Biological Sciences and Pest Management, Colorado State University, Fort Collins, CO, USA
Paul M. Burton
Affiliation:
Herbicide Chemistry and Senior Leader, Syngenta Crop Protection, Syngenta Jealott’s Hill International Research Center, Warfield, Bracknell, United Kingdom
Richard P. Dale
Affiliation:
Herbicide Molecular Sciences Team Leader, Syngenta Jealott’s Hill International Research Center, Warfield, Bracknell, United Kingdom
Jason K. Norsworthy
Affiliation:
Distinguished Professor and Elms Farming Chair of Weed Science, Department of Crop, Soil, and Environmental Sciences, University of Arkansas, Fayetteville, AR, USA
*
Author for correspondence: Amit J. Jhala, Department of Agronomy and Horticulture, University of Nebraska–Lincoln, 279 Plant Science Hall, P.O. Box 830915, Lincoln, NE 68583. Email: Amit.Jhala@unl.edu
Rights & Permissions [Opens in a new window]

Abstract

The herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD) are primarily used for weed control in corn, barley, oat, rice, sorghum, sugarcane, and wheat production fields in the United States. The objectives of this review were to summarize 1) the history of HPPD-inhibitor herbicides and their use in the United States; 2) HPPD-inhibitor resistant weeds, their mechanism of resistance, and management; 3) interaction of HPPD-inhibitor herbicides with other herbicides; and 4) the future of HPPD-inhibitor-resistant crops. As of 2022, three broadleaf weeds (Palmer amaranth, waterhemp, and wild radish) have evolved resistance to the HPPD inhibitor. The predominance of metabolic resistance to HPPD inhibitor was found in aforementioned three weed species. Management of HPPD-inhibitor-resistant weeds can be accomplished using alternate herbicides such as glyphosate, glufosinate, 2,4-D, or dicamba; however, metabolic resistance poses a serious challenge, because the weeds may be cross-resistant to other herbicide sites of action, leading to limited herbicide options. An HPPD-inhibitor herbicide is commonly applied with a photosystem II (PS II) inhibitor to increase efficacy and weed control spectrum. The synergism with an HPPD inhibitor arises from depletion of plastoquinones, which allows increased binding of a PS II inhibitor to the D1 protein. New HPPD inhibitors from the azole carboxamides class are in development and expected to be available in the near future. HPPD-inhibitor-resistant crops have been developed through overexpression of a resistant bacterial HPPD enzyme in plants and the overexpression of transgenes for HPPD and a microbial gene that enhances the production of the HPPD substrate. Isoxaflutole-resistant soybean is commercially available, and it is expected that soybean resistant to other HPPD inhibitor herbicides such as mesotrione, stacked with resistance to other herbicides, will be available in the near future.

Type
Review
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, provided the original article is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of the Weed Science Society of America

Introduction

Herbicides are used for managing weeds in diverse cropping systems in many countries (Jhala et al. Reference Jhala, Knezevic, Ganie, Singh, Chauhan and Mahajan2014a). The 4-hydroxyphenylpyruvate dioxygenase (HPPD)-inhibiting herbicides represent one of the latest discoveries of a new herbicide site of action that was introduced in the late 1990s (Mitchell et al. Reference Mitchell, Bartlett, Fraser, Hawkes, Holt, Townson and Wichert2001). Based on the site of action, HPPD inhibitor has been classified as Group 27 herbicides by the Weed Science Society of America and Herbicide Resistance Action Committee (Mallory-Smith and Retzinger Reference Mallory-Smith and Retzinger2017). The HPPD-inhibiting herbicides are broadly classified into chemical families: isoxazole (e.g., isoxaflutole), pyrazolone (e.g., pyrasulfotole, tolpyralate, topramezone), triketone (e.g., bicyclopyrone, mesotrione, and tembotrione; Figure 1), and isoxazolidinone (e.g., clomazone; Lee et al. Reference Lee, Prisbylla, Cromartie, Dagarin, Howard, Provan, Ellis, Fraser and Mutter1997). An additional class, azole carboxamides, has emerged in the patent literature, but these molecules have not been commercialized as of 2022 (Figure 1).

Figure 1. Chemical structures of some herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD).

These herbicides inhibit the HPPD enzyme found in plants and animals that is essential for the synthesis of plastoquinone and tocopherols (Liu and Lu Reference Liu and Lu2016). The plastoquinone is in turn a co-factor in the formation of carotenoids, which protect chlorophyll in plants. Because of the lack of plastoquinone, tocopherols and carotenoid synthesis due to HPPD inhibition, sensitive plants suffer oxidative damage and chlorophyll destruction, turn white without deformation, and eventually die (Mitchell et al. Reference Mitchell, Bartlett, Fraser, Hawkes, Holt, Townson and Wichert2001). The HPPD-inhibiting herbicides are mainly used to control annual grass and broadleaf weeds, including herbicide-resistant biotypes primarily in grass crops such as sugarcane and corn (Grossman and Ehrhardt Reference Grossmann and Ehrhardt2007; Pallett et al. Reference Pallett, Cramp, Little, Veerasekaran, Crudace and Slater2001). After the evolution and widespread occurrence of weeds that became resistant to the acetolactate synthase (ALS) inhibitor, atrazine and glyphosate, the HPPD-inhibiting herbicides, played a key role for their management in agronomic crops, primarily in corn (Ganie and Jhala Reference Ganie and Jhala2017).

The HPPD-inhibitor herbicides are used primarily for selective preemergence (PRE) and postemergence (POST) use as a weed control mechanism primarily in corn, barley, oat, rice, sorghum, sugarcane, and wheat. Mesotrione is labeled for PRE use in weed control in sorghum and sugarcane. The HPPD- and photosystem (PS) II-inhibiting herbicides are applied in a mixture, because certain herbicides belonging to both sites of action interact synergistically and provide higher efficacy compared with being applied alone (Fluttert et al. Reference Fluttert, Soltani, Galla, Hooker, Robinson and Sikkema2022). For example, field experiments conducted in Nebraska reported that Palmer amaranth (Amaranthus palmeri L.) that was resistant to atrazine and an HPPD inhibitor was effectively controlled by their mix, even applied at labeled rates (Chahal and Jhala Reference Chahal and Jhala2018a).

The scientific literature is not available to cover the past, present, and future of the HPPD inhibitor. Therefore, the objectives of this review were to 1) summarize the history of HPPD-inhibitor herbicides and their use in the United States; 2) summarize HPPD-inhibitor resistant weeds, their mechanism of resistance, and management; 3) highlight the interactions of HPPD-inhibitor herbicides with other herbicides; and 4) summarize the future of HPPD-inhibitor herbicides, including products in the pipeline and HPPD-inhibitor-resistant crops.

History of HPPD-Inhibiting Herbicides

Inhibitors of HPPD (HPPD, EC 1.13.11.27) were the results of several concurrent industry research programs. Pyrazolones were first commercialized by the Sankyo company in 1980 with pyrazolynate in the United States (Figure 2). Pyrazoxyfen by Ishihara followed in 1985, benzofenap by Mitsubishi and Rhône-Poulenc in 1987, topramezone by BASF in 2006, and pyrasulfotole by Bayer Crop Science in 2007 (Figure 3). The Ishihara company has commercialized its corn herbicide tolpyralate, which was first registered in 2017 (Tsukamoto et al. Reference Tsukamoto, Kikugawa, Nagayama, Suganuma, Okita and Miyamoto2021).

Figure 2. Chemical structures of pyrazolone herbicides, a chemical family of herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD).

Figure 3. Timeline of commercialization of herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD), their respective chemical classes, and manufacturer.

KingAgroot launched four new pyrazolone herbicides in China in 2020: cyprafluone, bipyrazone, fenpyrozone, and tripyrasulfone. Cyprafluone became KingAgroot’s first active ingredient to launch outside of China when granted registration in Pakistan in 2021 (KingAgroot 2021), with a plan for registration in other countries. Cyprafluone controls grass weeds such as Japanese foxtail (Alopecurus japonicus Steud.) and littleseed canarygrass (Phalaris minor Retz.) in wheat (Triticum aestivum L.). Wang et al. (Reference Wang, Liu, Jin, Peng, Zhang and Wang2020) reported that bipyrazone applied POST has a potential for broadleaf weed control in wheat in China.

Concurrently, a research group at Stauffer, a legacy company of ICI and now Syngenta, discovered the triketone-type HPPD inhibitor in 1982 through a chemical ecology approach. Researchers at Reed Gray observed that few weeds grew under crimson bottlebrush [Callistemon citrinus (Curtis) Skeels] in the California chaparral. Bioassay-guided isolation of crimson bottlebrush extracts led to the identification of fractions that can induce bleaching in developing seedlings. The active fractions contained the natural product leptospermone, a natural inhibitor of HPPD (Dayan et al. Reference Dayan, Duke, Sauldubois, Singh, McCurdy and Cantrell2007; Owens et al. Reference Owens, Nanayakkara and Dayan2013). The herbicidal activity of leptospermone and a series of synthetic triketone analogues were patented in 1980 (Figure 4). Structure-activity relationship studies characterized the chemical toxophore that is responsible for inhibiting HPPD (Ahrens et al. Reference Ahrens, Lange, Müller, Rosinger, Willms and van Almsick2013).

Figure 4. Chemical structures of triketone herbicides, a chemical family of herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD).

Triketones were first introduced to growers in 1991 by Zeneca (now Syngenta) with sulcotrione. Since then, a steady stream of triketones have been launched: benzobicyclon (by SDS Biotech in 2001), mesotrione (by Syngenta in 2002), tembotrione (by Bayer Crop Science in 2007), tefuryltrione (by Bayer in 2009), and bicyclopyrone (by Syngenta in 2015; Beaudegnies et al. Reference Beaudegnies, Edmunds, Fraser, Hall, Hawkes, Mitchell, Schaetzer, Wendeborn and Wibley2009). Kumiai Chemical registered fenquinotrione (Figure 4), trademarked as Effeeda, to control broadleaf weeds and sedges in rice in Japan. The molecule’s 4-methoxyphenyl group confers resistance to rice while maintaining weed control efficacy via selective metabolism (Yamamoto et al. Reference Yamamoto, Tanetani, Uchiyama, Nagamatsu, Kobayashi, Ikeda and Kawai2021). Kumiai collaborated with Certis to register fenquinotrione in the European Union for weed control in cereals and rice. Certis submitted a registration package in 2021. If successful, it is expected that the product will be available around 2025. Ishihara launched lancotrione-sodium in 2019 for control of broadleaf weeds and sedges in rice, including weeds that are resistant to sulfonylurea herbicides. Originally invented at Central China Normal University, benquitrione is the first in a series of quinazoline-2,4-diones triketone herbicides from Guang-Fu Yang’s laboratory, co-developed with Shandong Cynda (Figure 4; Wang et al. Reference Wang, Lin, Cao, Ming, Chen, Hao, Yang and Yang2015).

The first HPPD inhibitor was discovered serendipitously by Japanese companies, with pyrazolynate discovered by Sankyo and commercialized in 1980, and pyrazoxyfen discovered by Ishihara in 1985 (Figure 4). Both were commercialized for weed control in rice before their site of action was understood. It is now known that these compounds were pro-herbicides that are bio-activated into free hydroxypyrazole active pharmacophore, which inhibits HPPD.

Studies carried out by Rhone-Poulenc (now Bayer Crop Science) in the late 1980s led to the discovery of isofluxatole, an isoxazole heterocyclic proherbicide that is bio-activated to a diketonitrile by soil and plant enzymes (Pallett et al. Reference Pallett, Cramp, Little, Veerasekaran, Crudace and Slater2001; Figure 4). The bleaching caused by the HPPD inhibitor is similar to that observed with an inhibitor of phytoene desaturase (PDS), but the mechanism by which this bleaching occurred eluded researchers. The link between triketone molecules and their inhibition of HPPD was first elucidated using mammalian systems related to tyrosine metabolism. Subsequent investigations in plant systems established that HPPD catalyzes a key step in plastoquinone and tocopherol synthesis (Schultz et al. Reference Schultz, Soll, Fiedler and Schultze-Siebert1985), and further studies demonstrated that plastoquinone was an essential co-factor for phytoene desaturase (Norris et al. Reference Norris, Barrette and DellaPenna1995). This established the link between inhibition of HPPD and the bleaching symptoms that can be observed in the foliage of treated plants. In brief, plants treated with an HPPD inhibitor accumulate tyrosine and are depleted in plastoquinone. Without plastoquinone, PDS cannot function, which halts carotenoid biosynthesis, resulting in bleaching of the new growth, which is known as the “triketone effect” (Lee et al. Reference Lee, Prisbylla, Cromartie, Dagarin, Howard, Provan, Ellis, Fraser and Mutter1997).

Use of HPPD Inhibitor in the United States

In a survey conducted by the United States Department of Agriculture–National Agricultural Statistics Service (USDA-NASS) in 2018, the use of HPPD-inhibiting herbicides, including isoxaflutole, tembotrione, mesotrione, bicyclopyrone, and topramezone, was estimated at about 193,000, 214,000, 1,898,000, 102,000, and 41,000 kg, respectively (Figure 5).

Figure 5. Annual use of major herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD) in corn production in the United States in 2018 (Source: USDA-NASS 2018).

Mesotrione. Mesotrione belongs to the triketone family of HPPD-inhibitor herbicides and represents one of the most used active ingredients in corn (applied to about 42% of planted corn in 2018) with an average of one application (75 to 150 g ha−1) per year (Figure 6; USDA-NASS 2018). The Midwestern states, including Iowa, Illinois, Kansas, Minnesota, and Nebraska, led the use of mesotrione with an average annual use of >10,000 kg in 2018 at the rate of >1.27 kg mesotrione per square mile in each state (Figure 6). Increased use of mesotrione in recent years is primarily attributed to increasing demand for controlling glyphosate-resistant weeds (Chahal and Jhala Reference Chahal and Jhala2018b; Ganie et al. Reference Ganie, Stratman and Jhala2015; Ganie and Jhala Reference Ganie and Jhala2017). Mesotrione is a systemic herbicide applied alone or in mixture for selective PRE and POST control of grass and broadleaf weeds in field corn, seed corn, yellow popcorn, sweet corn, and grain sorghum (Abit et al. Reference Abit, Al-Khatib, Currie, Stahlman, Geier, Gordon, Olson, Claassen and Regehr2010; Armel et al. Reference Armel, Wilson, Richardson and Hines2003; Currie and Geier Reference Currie and Geier2018; Janak and Grichar Reference Janak and Grichar2016; Stephenson et al. Reference Stephenson, Bond, Walker, Bararpour and Oliver2004; Williams et al. Reference Williams, Pataky, Nordby, Riechers, Sprague and Masiunas2005).

Figure 6. Mesotrione use in agricultural land across the United States in 2018 (adapted from the U.S. Geological Survey by the U.S. Department of the Interior).

Mesotrione in a pre-mixture or tank-mixture with other herbicides can provide effective control of ALS-, PS II-, and glyphosate-resistant weeds (Chahal and Jhala Reference Chahal and Jhala2018a; Ganie et al. Reference Ganie, Stratman and Jhala2015). In addition to corn and grain sorghum, the use of mesotrione applied PRE in spring cereals, including barley (Hordeum vulgare L.), durum wheat, oats (Avena sativa L.), and spring wheat, has been found to be safe and provides adequate selective control of broadleaf weeds, including common lambsquarters (Chenopodium album L.), common ragweed (Ambrosia artemisiifolia L.), and wild buckwheat (Polygonum convolvulus L.) in a study conducted in Ontario, Canada (Soltani et al. Reference Soltani, Shropshire and Sikkema2011, Reference Soltani, Shropshire, Cowan and Sikkema2014); however, mesotrione use in those crops is limited (Walsh et al. Reference Walsh, Newman and Chatfield2021).

Tembotrione

Tembotrione is a member of the triketone family of HPPD inhibitors that is used for selective POST control of grass and broadleaf weeds in corn (Stephenson et al. Reference Stephenson, Bond, Landry and Edwards2015). Tembotrione is the second-highest used HPPD inhibitor in the United States, with an annual use rate of >200,000 kg (Figure 7; USDA-NASS 2018). Minnesota, Illinois, Nebraska, Indiana, and Iowa were leading states for annual use (>20,000 kg) of tembotrione in corn production in 2018 (Figure 7). Tembotrione (Laudis; Bayer Crop Science, St Louis, MO) is applied alone or in a mixture from field corn emergence to the V8 growth stage or V7 (sweet corn). More recently, metabolic-based resistance (CYP-mediated metabolism) to tembotrione has been identified in several grain sorghum lines (Pandian et al. Reference Pandian, Varanasi, Vannapusa, Sathishraj, Lin, Zhao, Tunnell, Tesso, Liu, Prasad and Jugulam2020), indicating a future increase in tembotrione use in other crops.

Figure 7. Tembotrione use in various corn-producing states in the United States (Source: USDA-NASS 2018).

Isoxaflutole

Isoxaflutole was the first member of the isoxazole class of HPPD inhibitor. Common brand names include Balance Flexx™ and Corvus™, and were the first HPPD-inhibiting herbicides introduced in North America in 1996 (Figure 1; Pallett et al. Reference Pallett, Little, Sheekey and Veerasekaran1998). It is a selective herbicide primarily used for PRE control of grass and broadleaf weeds in corn, and recently in isoxaflutole-resistant soybean (Mausbach et al. Reference Mausbach, Irmak, Sarangi, Lindquist and Jhala2021). Isoxaflutole is commonly mixed with PS II-inhibiting herbicides (e.g., atrazine) to improve weed control efficacy and spectrum (Benoit et al. Reference Benoit, Soltani, Hooker, Robinson and Sikkema2019; Chahal and Jhala Reference Chahal and Jhala2018a; Fluttert et al. Reference Fluttert, Soltani, Galla, Hooker, Robinson and Sikkema2022). According to the survey conducted by the USDA-NASS (2018), isoxaflutole was the third-most used HPPD inhibitor in corn (used in about 8% of planted corn) with an average of one application (72 g ha−1) per year (Figure 8). Iowa, Illinois, and Nebraska were the leading states for isoxaflutole use among various corn-producing states in 2018 (Figure 8). Isoxaflutole has been widely used as a part of herbicide-resistant weed management strategies (including ALS-, PS II-, and glyphosate-resistant) in corn (Benoit et al. Reference Benoit, Soltani, Hooker, Robinson and Sikkema2019; Chahal et al. Reference Chahal, Aulakh, Jugulam, Jhala and Price2015; Stephenson and Bond Reference Stephenson and Bond2012).

Figure 8. Isoxaflutole use in major corn-producing states in the United States (Source: USDA-NASS 2018).

Isoxaflutole can also be used for weed control in fallow fields depending on the subsequent rotational crop (Currie and Geier Reference Currie and Geier2016; Kumar and Jha Reference Kumar and Jha2015). In this context, isoxaflutole-resistant soybean has recently been developed in which isoxaflutole can be used as a part of an herbicide strategy to control ALS- and glyphosate-resistant weeds, including Palmer amaranth, waterhemp, and Canada fleabane (Erigeron canadensis L.; Ditschun et al. Reference Ditschun, Soltani, Robinson, Tardif, Kaastra and Sikkema2016; Mausbach et al. Reference Mausbach, Irmak, Sarangi, Lindquist and Jhala2021; Smith et al. Reference Smith, Soltani, Kaastra, Hooker, Robinson and Sikkema2019a).

Bicyclopyrone and Topramezone

Bicyclopyrone and topramezone are two other HPPD-inhibitor herbicides (via an active ingredient of an individual product or various premixtures) that are commonly used for grass and broadleaf weed control in field corn, seed corn, silage corn, yellow popcorn, sweet popcorn, and sugarcane (Sarangi and Jhala Reference Sarangi and Jhala2018). Topramezone belongs to the pyrazolone family with an annual use of >41,000 kg, whereas bicyclopyrone belongs to the triketone family with an annual use of >100,000 kg in corn production (Figure 9).

Figure 9. Bicyclopyrone and topramezone use in various corn-producing states in the United States (Source: USDA-NASS 2018).

According to the USDA-NASS (2018) survey, Illinois, Iowa, Kansas, Missouri, Nebraska, and Wisconsin were the leading states for annual use of bicyclopyrone with an estimate of >5,000 kg in corn crops, whereas Illinois and Iowa were the top states in annual use of topramezone (>5,000 kg) for weed control in corn (Figure 9). In addition to corn, topramezone and bicyclopyrone are known to provide effective weed control in other crops, including turf, sweet potato [Ipomoea batatas (L.) Lam.], wheat, chickpea (Cicer arietinum L.), and rice (Brosnan and Breeden Reference Brosnan and Breeden2013; Lindley et al. Reference Lindley, Jennings, Monks, Chaudhari, Schultheis, Waldschmidt and Brownie2020; Moore 2019).

Pyrasulfotole and Tolpyralate

Pyrasulfotole is a member of the pyrazolone family of HPPD inhibitors and is registered for use on cereal grains, including wheat, barley, rye, triticale, and grain sorghum (Kumar et al. Reference Kumar, Jha and Reichard2014; Reddy et al. Reference Reddy, Stahlman, Geier, Thompson, Currie, Schlegel, Olson and Lally2013; Torbiak et al. Reference Torbiak, Blackshaw, Brandt, Hamman and Geddes2021). Pyrasulfotole is an active ingredient of Huskie™ (a premixture of pyrasulfotole and bromoxynil; Bayer Crop Science, Saint Louis, MO) that is used for broadleaf weed control in sorghum. In contrast, tolpyralate is a new HPPD inhibitor that came to the market in 2020 and controls several annual grass and broadleaf weed species with a low use rate (30 to 50 g ha−1) in corn (Tsukamoto et al. Reference Tsukamoto, Kikugawa, Nagayama, Suganuma, Okita and Miyamoto2021; Willemse et al. Reference Willemse, Soltani, Metzger, Hooker, Jhala, Robinson and Sikkema2021c).

Benzobicyclon

Benzobicyclon, a pro-herbicide, is a member of the triketone family of HPPD inhibitors. It was first registered for use in rice crops in 2021 under the tradename Rogue® (Gowan Company, Yuma, AZ), mainly for control of aquatic weeds, sprangletop species (Leptochloa spp.), and suppression of weedy rice biotypes when applied post-flood. This is the only HPPD-inhibiting herbicide available for use rice production in the United States.

HPPD-Inhibitor-Resistant Weeds and Their Mechanisms of Resistance

Although HPPD-inhibitor herbicides have been in use for more than two decades, the evolution of HPPD-inhibitor-resistant weeds is relatively less widespread and slower than some other herbicide sites of action (Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b; Kaundun Reference Kaundun2021). As of 2022, three broadleaf weeds (Palmer amaranth, waterhemp, and wild radish) have evolved resistance to HPPD-inhibitor herbicides across the globe (Heap Reference Heap2022). Several populations of HPPD-inhibitor-resistant Palmer amaranth and waterhemp have evolved across the Midwestern United States (Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b; Hausman et al. Reference Hausman, Singh, Tranel, Riechers, Kaundun, Polge, Thomas and Hager2011), while HPPD-inhibitor-resistant wild radish has been documented in Western Australia. The first case of resistance to these herbicides was reported in a population of waterhemp from a corn field in Illinois that had a history of repeated HPPD-inhibitor use (Hausman et al. Reference Hausman, Singh, Tranel, Riechers, Kaundun, Polge, Thomas and Hager2011). The resistance in Palmer amaranth (Thompson et al. Reference Thompson, Peterson and Nathan2012) and wild radish (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020) were not selected with HPPD-inhibitor herbicides; rather, these populations exhibited cross-resistance with the mechanisms that bestow resistance to different herbicide sites of action (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020; Nakka et al. Reference Nakka, Godar, Wani, Thompson, Peterson, Roelofs and Jugulam2017).

Mechanisms of Resistance to HPPD Inhibitors

Palmer Amaranth

The first case of Palmer amaranth resistance to the HPPD inhibitor (also found to be resistant to PS II- and ALS-inhibitor formulas) was confirmed in a field in central Kansas where there was no history of HPPD inhibitor use, though there was a long history of use of herbicides that PS II and ALS (Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b; Thompson et al. Reference Thompson, Peterson and Nathan2012). This population was originally found to be resistant to Huskie®, a premix of pyrasulfotole (an HPPD inhibitor) and bromoxynil (a PS II inhibitor). Furthermore, this Palmer amaranth biotype was resistant to several HPPD inhibitor herbicides, including mesotrione, tembotrione, and topramezone (Thompson et al. Reference Thompson, Peterson and Nathan2012). Later, a Palmer amaranth population from a corn field in Nebraska that had a history of HPPD inhibitor use was found to be resistant to these herbicides (Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b). Populations of Palmer amaranth in Kansas and Nebraska exhibited up to 18-fold resistance to mesotrione, tembotrione, or topramezone (Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b; Nakka et al. Reference Nakka, Godar, Wani, Thompson, Peterson, Roelofs and Jugulam2017; Thompson et al. Reference Thompson, Peterson and Nathan2012). In both populations, the mechanism of resistance to the HPPD inhibitor was due to neither differential herbicide uptake/translocation nor mutations or amplification of the HPPD gene (Küpper et al. Reference Küpper, Peter, Zöllner, Lorentz, Tranel, Beffa and Gaines2018; Nakka et al. Reference Nakka, Godar, Wani, Thompson, Peterson, Roelofs and Jugulam2017). The Kansas Palmer amaranth biotype metabolized more than 90% of mesotrione at 24 h after treatment compared with sensitive plants (Nakka et al. Reference Nakka, Godar, Wani, Thompson, Peterson, Roelofs and Jugulam2017). Additionally, a 4-fold to 14-fold higher HPPD gene expression was found in this population (Nakka et al. Reference Nakka, Godar, Wani, Thompson, Peterson, Roelofs and Jugulam2017). Similarly, the rapid metabolism of tembotrione was attributed to the resistance in Palmer amaranth population from Nebraska (Küpper et al. Reference Küpper, Peter, Zöllner, Lorentz, Tranel, Beffa and Gaines2018). Although 4-hydroxylation of tembotrione followed by glycosylation was identified in both resistant and sensitive plants, the time taken to form metabolites was shorter in resistant plants compared with sensitive plants (Küpper et al. Reference Küpper, Peter, Zöllner, Lorentz, Tranel, Beffa and Gaines2018). More recently, a population of Palmer amaranth from Kansas (Riley County) was resistant to mesotrione and tembotrione (Shyam et al. Reference Shyam, Borgato, Peterson, Dille and Jugulam2021). The mechanism of resistance in this population is being investigated.

Waterhemp

Resistance to HPPD inhibitor herbicides has been documented in several populations of waterhemp across the Midwestern United States, including Illinois, Iowa, and Nebraska (Heap Reference Heap2022). A biotype of waterhemp known as MCR (for McLean County resistant) from Illinois was the first reported case of resistance to an HPPD inhibitor (Hausman et al. Reference Hausman, Singh, Tranel, Riechers, Kaundun, Polge, Thomas and Hager2011). This biotype was previously confirmed to be resistant to atrazine and ALS-inhibiting herbicides. MCR waterhemp had 10-fold and 35-fold resistance to mesotrione compared with two susceptible populations from Illinois (Hausman et al. Reference Hausman, Singh, Tranel, Riechers, Kaundun, Polge, Thomas and Hager2011). The mechanism of mesotrione resistance in MCR waterhemp was not due to reduced herbicide absorption/translocation nor because of alterations in the HPPD gene sequence or expression. However, compared with sensitive plants, MCR waterhemp rapidly metabolized mesotrione via hydroxylation of the cyclohexanedione ring of mesotrione (Ma et al. Reference Ma, Kaundun, Tranel, Riggins, McGinness, Hager, Hawkes, McIndoe and Riechers2013). Importantly, the time required to metabolize 50% of the absorbed mesotrione was ∼11.7 h in MCR compared with 25.4 to 27.8 h in the susceptible plants. Application of the cytochrome P450 inhibitor (malathion) increased the susceptibility of MCR plants to mesotrione, suggesting that the metabolism of mesotrione was mediated via P450 activity in this population (Ma et al. Reference Ma, Kaundun, Tranel, Riggins, McGinness, Hager, Hawkes, McIndoe and Riechers2013).

The HPPD-inhibitor-resistant waterhemp from Nebraska known as NEB showed a 2.4-fold and 45-fold level of resistance to mesotrione applied PRE and POST, respectively, compared with a known susceptible population (Kaundun et al. Reference Kaundun, Hutchings, Dale, Howell, Morris, Kramer, Shivrain and McIndoe2017). Similar to MCR waterhemp, mesotrione resistance in the Nebraska population was primarily due to higher levels of mesotrione metabolism via 4-hydroxylation (Kaundan et al., Reference Kaundun, Hutchings, Dale, Howell, Morris, Kramer, Shivrain and McIndoe2017). Furthermore, the metabolites of mesotrione were identified as 4-hydroxymesotrione and AMBA [2-amino-4-(methylsulfonyl) benzoic acid; (Kaundan et al. Reference Kaundun, Hutchings, Dale, Howell, Morris, Kramer, Shivrain and McIndoe2017)]. No duplication, alteration, or over-expression of the HPPD gene that can confer resistance was found in this population (Kaundan et al. Reference Kaundun, Hutchings, Dale, Howell, Morris, Kramer, Shivrain and McIndoe2017). Moreover, mesotrione-resistant waterhemp population from Illinois and Nebraska were also resistant to topramezone, which belongs to the pyrazolone subfamily of HPPD inhibitors. Both populations rapidly metabolized topramezone, and the metabolic profiles indicated two different putative hydroxylated forms of topramezone (hydroxytopramezone-1 and hydroxytopramezone-2), although hydroxytopramezone-1 was more abundant in the Illinois waterhemp population (Lygin et al. Reference Lygin, Kaundun, Morris, Mcindoe, Hamilton and Riechers2018). When metabolic profiles at 48 h after treatment were compared with naturally tolerant corn, the waterhemp population from Illinois had more hydroxylated metabolites, whereas corn plants produced desmethyl and benzoic acid metabolites of topramezone, suggesting that waterhemp initially metabolizes topramezone differently than corn (Lygin et al. Reference Lygin, Kaundun, Morris, Mcindoe, Hamilton and Riechers2018).

More recently, the mechanism of resistance to syncarpic acid-3, a nonselective, noncommercial HPPD inhibitor, was investigated in an Illinois population of waterhemp (Concepcion et al. Reference Concepcion, Kaundun, Morris, Hutchings, Strom, Lygin and Riechers2021). Although the Phase I metabolite, likely produced due to P450-mediated hydroxylation was detected in this population, this metabolite was not found to be responsible for resistance; rather, the glutathione-syncarpic acid conjugate that formed as a result of Phase II metabolism was associated with resistance to syncarpic acid in the waterhemp population (Concepcion et al. Reference Concepcion, Kaundun, Morris, Hutchings, Strom, Lygin and Riechers2021).

Wild Radish. A population of wild radish from a Western Australian grain field with no prior history of HPPD inhibitor use is resistant to these herbicides (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020). This population is also resistant to other herbicides such as PS II inhibitor, ALS inhibitor, and synthetic auxin (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020). This wild radish population exhibited 4-fold to 6.5-fold resistance to mesotrione, tembotrione, and isoxaflutole (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020). The resistant plants were able to metabolize mesotrione approximately 8-fold faster than the sensitive plants (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020). It was also confirmed that the resistance was not due to reduced uptake/translocation of mesotrione, and no target site alterations were detected (Lu et al. Reference Lu, Yu, Han, Owen and Powles2020).

Although resistance to HPPD inhibitor herbicides has currently been reported in three weed species across the globe, if selection pressure continues, new cases of resistance evolution to HPPD inhibitor will increase. More importantly, the predominance of metabolic resistance to HPPD inhibitor herbicides, was found in all three weed species (Jugulam and Shyam Reference Jugulam and Shyam2019; Yu and Powles Reference Yu and Powles2014). Therefore, prudent strategies, including integration of nonchemical methods, need to be designed for sustainable weed management.

Management of HPPD-Inhibitor-Resistant Weeds

As of 2022, Palmer amaranth and waterhemp are the only two weed species in the United States that have evolved resistance to HPPD-inhibitor herbicides (Heap Reference Heap2022; Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b). Therefore, strategies described here focus on the management of HPPD-inhibitor-resistant Palmer amaranth and waterhemp primarily in corn and soybean production systems. Although the evolution of herbicide resistance in weed species cannot be completely averted, it can possibly be delayed by implementing diversified weed management practices (Norsworthy et al. Reference Norsworthy, Ward, Shaw, Llewellyn, Nichols, Webster, Bradley, Frisvold, Powles, Burgos, Witt and Barrett2012). Mixing herbicides that have different sites of action is often recommended to delay the evolution of herbicide-resistant weeds (Diggle et al. Reference Diggle, Neve and Smith2003; Evans et al. Reference Evans, Tranel, Hager, Schutte, Wu, Chatham and Davis2016). In corn, HPPD-inhibiting herbicides are mixed with PS II-inhibiting herbicides due to their synergistic activity for controlling triazine-resistant weeds (Chahal et al. Reference Chahal, Jugulam and Jhala2019; Hugie et al. Reference Hugie, Bollero, Tranel and Riechers2008; Woodyard et al. Reference Woodyard, Hugie and Riechers2009c). However, continued use of this mixture to control atrazine-resistant weeds in corn has resulted in the evolution and widespread occurrence of Palmer amaranth and waterhemp populations resistant to PS II and HPPD inhibitors (Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b).

Herbicide options to control Palmer amaranth and waterhemp in corn and soybean crops include inhibitors of ALS, PS II, HPPD, protoporphyrinogen oxidase (PPO), very long chain fatty acid (VLCFA), glyphosate, glufosinate, and synthetic auxins. However, Palmer amaranth and waterhemp populations with multiple resistance to ALS, PS II, HPPD, and PPO inhibitors, and glyphosate are increasingly common in the Midwestern United States (Heap Reference Heap2022; Jhala et al. Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b; Schultz et al. Reference Schultz, Chatham, Riggins, Tranel and Bradley2015; Shyam et al. Reference Shyam, Borgato, Peterson, Dille and Jugulam2021; Varanasi et al. Reference Varanasi, Brabham, Norsworthy, Nie, Young, Houston, Barber and Scott2018). This has reduced herbicide options to control weeds in corn and soybean production systems (Sarangi et al. Reference Sarangi, Stephens, Barker, Patterson, Gaines and Jhala2019). Therefore, herbicide programs containing diverse herbicide sites of action and detoxification pathways are required to manage HPPD-inhibitor-resistant Palmer amaranth and waterhemp.

Fortunately, waterhemp biotypes with metabolic resistance to atrazine, a resistance mechanism present in a majority of populations in the Midwest (Tranel Reference Tranel2021), are sensitive to other PS II-inhibitor herbicides such as metribuzin (Jacobs et al. Reference Jacobs, Butts-Wilmsmeyer, Ma, O’Brien and Riechers2020). Therefore, metribuzin mixed with an HPPD-inhibitor can control populations that are resistant to an HPPD inhibitor and atrazine. For example, atrazine at 4.48 kg ha−1 applied PRE provided 26% control of PS II-inhibitor-resistant and HPPD-inhibitor-resistant waterhemp at 4 wk after treatment (WAT), whereas metribuzin 560 g ha−1 provided 95% control of the same population (Evans et al. Reference Evans, Strom, Riechers, Davis, Tranel and Hager2019). Similarly, in a greenhouse study, waterhemp that was resistant to PS II and HPPD inhibitors exhibited a synergistic response to metribuzin at 191 g ha−1 + mesotrione at 53 g ha−1 applied POST, indicating that this may be a viable option for controlling atrazine- and HPPD-inhibitor-resistant waterhemp in corn (O’Brien et al. Reference O’Brien, Davis and Riechers2018).

PRE herbicides serve as a foundation for herbicide-resistant weed management; however, using a PRE and a POST herbicide from the same site of action is not a recommended strategy because it can potentially lead to an increase in the frequency of resistance over time (Hausman et al. Reference Hausman, Tranel, Riechers, Maxwell, Gonzini and Hager2013; Wuerffel et al. Reference Wuerffel, Young, Tranel and Young2015). Therefore, herbicides from alternative sites of action should be included in herbicide programs when HPPD-inhibitor-resistant weeds are present in the field (Chahal and Jhala Reference Chahal and Jhala2018b). For example, a PRE application of isoxaflutole at 105 g ha−1 plus mesotrione 210 g ha−1 provided <65% control of HPPD-inhibitor-resistant waterhemp at 4 WAT compared with >85% control by using acetochlor 1,680 g ha−1 applied PRE in corn (Hausman et al. Reference Hausman, Tranel, Riechers, Maxwell, Gonzini and Hager2013). In a similar study conducted in soybean, flumioxazin at 70 g ha−1, sulfentrazone 280 g ha−1, metribuzin 420 g ha−1, or pyroxasulfone 210 g ha−1 applied PRE provided >85% control of HPPD-inhibitor-resistant waterhemp (Hausman et al. Reference Hausman, Tranel, Riechers, Maxwell, Gonzini and Hager2013). This strategy would potentially reduce the number of survivors, thereby delaying the selection of resistance alleles in the population (Wuerffel et al. Reference Wuerffel, Young, Tranel and Young2015).

Relatively fewer POST herbicide options are available to control HPPD-inhibitor-resistant Palmer amaranth and waterhemp in corn and soybean crops (Jhala et al. Reference Jhala, Knezevic, Ganie, Singh, Chauhan and Mahajan2014a). Glufosinate, 2,4-D, or dicamba are among the few POST herbicides that provide more than 80% control; for example, Jhala et al. (Reference Jhala, Sandell, Rana, Kruger and Knezevic2014b) reported that glufosinate (450 g ha−1), 2,4-D ester (560 g ha−1), or dicamba (560 g ha−1) provided >80% control of HPPD-inhibitor-resistant Palmer amaranth 3 WAT. Similarly, 740 g ha−1 of glufosinate or 280 g ha−1 of dicamba provided >90% control of HPPD inhibitor-resistant waterhemp at 3 WAT (Sarangi et al. Reference Sarangi, Stephens, Barker, Patterson, Gaines and Jhala2019). Glufosinate used at 595 g ha−1 applied alone or mixed with dicamba at 280 g ha−1 provided ≥79% control of HPPD-inhibitor-resistant Palmer amaranth 4 WAT in corn (Chahal and Jhala Reference Chahal and Jhala2018a). Additionally, Oliveira et al. (Reference Oliveira, Jhala, Gaines, Irmak, Amundsen, Scott and Knezevic2017) reported that mixing metribuzin (210 g ha−1) to a premix of mesotrione + atrazine (650 g ha−1) applied POST in corn, controlled HPPD-inhibitor-resistant waterhemp by >90% at 3 WAT.

Although herbicides such as metribuzin, pyroxasulfone, glufosinate, or dicamba applied alone can control early- to mid-season cohorts of Palmer amaranth and waterhemp, a season-long control of these weed species is rarely achieved due to their extended period of emergence (Hager et al. Reference Hager, Wax, Simmons and Stoller1997; Keeley et al. Reference Keeley, Carter and Thullen1987). Therefore, multiple herbicide applications (PRE followed by POST), specifically overlapping residual herbicides, are recommended to achieve a season-long control of HPPD-inhibitor-resistant weeds. For example, a premix of acetochlor + clopyralid + flumetsulam (1,190 g ha−1), or saflufenacil + dimethenamid-P (780 g ha−1) applied PRE provided >90% control of HPPD-inhibitor-resistant Palmer amaranth for 3 wk; however, control was reduced to <70% later in the season (Chahal and Jhala Reference Chahal and Jhala2018b). In the same study, glyphosate (870 g ha−1) + dicamba (280 g ha−1) was needed to obtain >96% control. Similarly, overlapping residual herbicide programs, including pyroxasulfone (110 g ha−1) + saflufenacil (75 g ha−1), or saflufenacil + dimethenamid-P (586 g ha−1) applied PRE followed by glyphosate (870 g ha−1) + diflufenzopyr + dicamba (157 g ha−1) + pyroxasulfone (91 g ha−1), or glyphosate + dicamba + diflufenzopyr (157 g ha−1) + pendimethalin (1,060 g ha−1) applied POST controlled HPPD-inhibitor-resistant Palmer amaranth >95% at corn harvest (Chahal et al. Reference Chahal, Ganie and Jhala2018a). In a study conducted in a conventional corn crop, herbicide programs including acetochlor (2,130 g ha−1), or mesotrione + S-metolachlor + atrazine (2,780 g ha−1) applied PRE, followed by dicamba + diflufenzopyr (196 g ha−1) applied POST, controlled HPPD-inhibitor-resistant Palmer amaranth >95% (Chahal et al. Reference Chahal, Irmak, Gaines, Amundsen, Jugulam, Jha, Travlos and Jhala2018b). Similarly, dicamba + thiencarbazone-methyl + atrazine or dicamba + ABMS (acetochlor + bicyclopyrone + mesotrione + S-metolachlor) applied PRE followed by ABMS alone or in a mixture with atrazine, S-metolachlor, or mesotrione applied early POST provided 85% to 6% control of glyphosate and mesotrione-resistant Palmer amaranth at 2 wk after early POST and 2 and 7 wk after late POST in glyphosate/glufosinate-resistant corn in central Kansas (Liu et al. Reference Liu, Kumar, Jhala, Jha and Stahlman2021).

While sequential applications of PRE followed by POST herbicides with multiple sites of action can control HPPD-inhibitor-resistant Palmer amaranth and waterhemp, relying on a single control tactic would potentially enhance selection pressure for the evolution of multiple-herbicide-resistant weeds. Therefore, diversified weed management strategies, including cultural, biological, mechanical, and chemical weed management (with multiple sites of action), are needed to manage herbicide-resistant weed seed banks. More specifically, multi-tactic strategies that target multiple life stages of the weed, including understanding reproductive biology and potential for pollen-mediated gene flow, are required (Jhala et al. Reference Jhala, Beckie, Mallory-Smith, Jasieniuk, Busi, Norsworthy, Bagavathiannan, Tidemann and Geddes2021a, Reference Jhala, Norsworthy, Ganie, Sosnoskie, Beckie, Mallory-Smith, Liu, Wei, Wang and Stoltenberg2021b). This can be accomplished by using an effective multiple-sites-of-action herbicide program, using cover crops, planting corn or soybean in narrow rows, using a harvest weed seed control method, and adopting diversified crop rotations (Mohler et al. Reference Mohler, Teasdale and DiTommaso2021; Striegel and Jhala Reference Striegel and Jhala2022). The increasing use of HPPD-inhibitor herbicides in agronomic crops requires research on herbicide interactions and alternative herbicides or methods for controlling multiple herbicide-resistant weeds.

Interactions of HPPD-Inhibitor Herbicides with Other Herbicides

The HPPD-inhibiting herbicides are commonly mixed with other herbicides, particularly PS II-inhibiting herbicides, to increase weed control and spectrum. The assumption of an herbicide combination is that each herbicide acts independently when applied together (i.e., additive); however, that is not always the case. Weed control from a mixture of two herbicides may be greater (synergistic) or less than (antagonistic) the combined effect of the herbicides applied alone (Colby Reference Colby1967; Hatzios and Penner Reference Hatzios and Penner1985).

Efficacy

The HPPD-inhibiting herbicides applied PRE with a PS II-inhibitor herbicide can have both additive and synergistic effects. In greenhouse studies, atrazine + mesotrione applied PRE were additive for control of velvetleaf (Abutilon theophrasti Medik.) and ivyleaf morningglory (Ipomoea hederacea Jacq.); however, several rate combinations indicated synergistic control (Bollman et al. Reference Bollman, Kells and Penner2006). In field experiments, isoxaflutole + metribuzin applied PRE exhibited additive or synergistic control of Canada fleabane, common lambsquarters, Amaranthus spp., common ragweed, velvetleaf, Setaria spp., barnyardgrass [Echinochloa crus-galli (L.) Beauv.], and fall panicum (Panicum dichotomiflorum Michx.; Ditschun et al. Reference Ditschun, Soltani, Robinson, Tardif, Kaastra and Sikkema2016; Smith et al. Reference Smith, Soltani, Kaastra, Hooker, Robinson and Sikkema2019b). In contrast, control of a population of HPPD-inhibitor-resistant and PS II-inhibitor-resistant Palmer amaranth with mesotrione or topramezone applied PRE with atrazine was additive (Chahal and Jhala Reference Chahal and Jhala2018a).

The literature is replete with observations of additive or synergistic weed control when HPPD-inhibiting herbicides are applied POST with a PS II inhibitor. Mesotrione + atrazine (Abendroth et al. Reference Abendroth, Martin and Roeth2006; Armel et al. Reference Armel, Hall, Wilson and Cullen2005; Creech et al. Reference Creech, Monaco and Evans2004), mesotrione + bromoxynil or metribuzin (Abendroth et al. Reference Abendroth, Martin and Roeth2006) were consistent for controlling several annual weeds as well as Canada thistle [Cirsium arvense (L.) Scop.] compared to mesotrione applied alone. Mixing atrazine with tolpyralate improved control (Metzger et al. Reference Metzger, Soltani, Raeder, Hooker, Robinson and Sikkema2018) or reduced the biologically effective dose of tolpyralate for control of weeds commonly found in corn production fields in Nebraska (Osipitan et al. Reference Osipitan, Scott and Knezevic2018). In research plots throughout North America, mixing atrazine with tembotrione reduced variability in weed control and sweet corn yield variation (Williams et al. Reference Williams, Boydston, Peachey and Robinson2011a, Reference Williams, Boydston, Peachey and Robinson2011b). Similar findings were observed with atrazine + isoxaflutole, mesotrione, topramezone, tembotrione, or tolpyralate for waterhemp control in field corn (Willemse et al. Reference Willemse, Soltani, Benoit, Jhala, Hooker, Robinson and Sikkema2021a). Furthermore, atrazine improved the efficacy of pyrasulfotole + bromoxynil for weed control in grain sorghum (Reddy et al. Reference Reddy, Stahlman, Geier, Thompson, Currie, Schlegel, Olson and Lally2013).

Synergism between an HPPD inhibitor and a PS II inhibitor applied POST can be observed in herbicide-resistant weed populations. Synergistic control with mesotrione + atrazine has been observed in PS II-inhibitor-resistant redroot pigweed (Amaranthus retroflexus L.) (Hugie et al. Reference Hugie, Bollero, Tranel and Riechers2008; Sutton et al. Reference Sutton, Richards, Buren and Glasgow2002) and PS II-inhibitor-resistant wild radish (Walsh et al. Reference Walsh, Stratford, Stone and Powles2012), including temporally separated herbicide applications (e.g., atrazine PRE followed by mesotrione POST; Woodyard et al. Reference Woodyard, Hugie and Riechers2009a). Palmer amaranth, including a PS II-inhibitor-resistant population, exhibited synergistic control with atrazine + mesotrione or tembotrione, but not atrazine with tolpyralate or topramezone (Kohrt and Sprague Reference Kohrt and Sprague2017). Synergistic control of multiple-herbicide-resistant waterhemp was observed with mesotrione + bromoxynil or bentazon; and tolpyralate + bromoxynil (Willemse et al. Reference Willemse, Soltani, David, Jhala, Robinson and Sikkema2021b). In contrast, activity of isoxaflutole or mesotrione applied POST with metribuzin on waterhemp populations varying in herbicide resistance traits was mostly additive (O’Brien et al. Reference O’Brien, Davis and Riechers2018).

Mixing an HPPD inhibitor with a PS II inhibitor does not always result in synergistic weed control. Volunteer potato (Solanum tuberosum L.) control with mesotrione, tembotrione, or topramezone applied POST was not improved when mixed with atrazine, bentazon, or bromoxynil (Koepke-Hill et al. Reference Koepke-Hill, Armel, Wilson, Hines and Vargas2010). Advances have been made in understanding the mechanisms that account for synergistic weed control from mixing HPPD and PS II inhibitors. Armel et al. (Reference Armel, Hall, Wilson and Cullen2005) reported that uptake, translocation, and metabolism of mesotrione did not account for improved control of Canada thistle with mesotrione + atrazine. Mesotrione absorption in Palmer amaranth increased when it was mixed with atrazine, partially accounting for observed synergism (Chahal et al. Reference Chahal, Jugulam and Jhala2019). PS II inhibitors compete with plastoquinones for the D1 protein binding site, disrupting electron transfer in PS II. The inability to transfer electrons creates triplet chlorophyll and singlet oxygen that destroy plant membranes (Hess Reference Hess2000). Armel et al. (Reference Armel, Rardon, McComrick and Ferry2007) showed that carotenoid biosynthesis inhibitor increased the binding efficiency and efficacy of PS II inhibitor by reducing the reformation of the D1 protein following initiation of photo inhibition.

Several factors influence the synergism of an HPPD inhibitor applied in a mixture with a PS II inhibitor. For both PRE and POST applications, the herbicide rate influences the extent of synergistic weed control (Bollman et al. Reference Bollman, Kells and Penner2006; Hugie et al. Reference Hugie, Bollero, Tranel and Riechers2008). In addition to the application rate of the HPPD inhibitor, synergistic weed control was observed more frequently with triketone herbicides (mesotrione and tembotrione) compared to pyrazolone herbicides (topramezone and tolpyralate; Kohrt and Sprague Reference Kohrt and Sprague2017). Adverse environmental conditions (e.g., inadequate rainfall) influence plant response to a mixture of HPPD and PS II inhibitors applied PRE (Smith et al. Reference Smith, Soltani, Kaastra, Hooker, Robinson and Sikkema2019b) and POST (Woodyard et al. Reference Woodyard, Bollero and Riechers2009b). Moreover, the time of POST herbicide application can influence the plant response to HPPD and PS II inhibitors (O’Brien et al. Reference O’Brien, Davis and Riechers2018).

HPPD-inhibitor herbicides can interact with herbicides other than PS II inhibitors. The synthetic auxin triclopyr improved foliar uptake of mesotrione and control of smooth crabgrass (Yu and McCullough Reference Yu and McCullough2016). Conversely, a mixture of an HPPD inhibitor and an ALS inhibitor can be antagonistic. For example, reduced efficacy of sulfonylurea herbicides applied with mesotrione + atrazine for control of Setaria spp. (Schuster et al. Reference Schuster, Al-Khatib and Dille2008) was due to decreased absorption, and in some cases reduced translocation of nicosulfuron (Schuster et al. Reference Schuster, Al-Khatib and Dille2007). Not only can an HPPD inhibitor antagonize an ALS inhibitor for annual grass control, but an ALS inhibitor can also antagonize the HPPD inhibitor (Kaastra et al. Reference Kaastra, Swanton, Tardif and Sikkema2008).

Crop Tolerance

Field corn production systems rely extensively on a mixture of HPPD and PS II inhibitors. Considerable field research demonstrates excellent crop tolerance with their mixtures (Johnson et al. Reference Johnson, Young and Matthews2002; Osipitan et al. Reference Osipitan, Scott and Knezevic2018; Stephenson et al. Reference Stephenson, Bond, Walker, Bararpour and Oliver2004; Whaley et al. Reference Whaley, Armel, Wilson and Hines2006; Willemse et al. Reference Willemse, Soltani, Benoit, Jhala, Hooker, Robinson and Sikkema2021a). Additional research shows that the synergistic effect of HPPD and PS II inhibitors for weed control was not observed on sweet corn response (Choe et al. Reference Choe, Williams, Boydston, Huber, Huber and Pataky2014). Sweet corn injury from tembotrione was influenced by the safener isoxadifen and the genotypic class at a P450 locus (Nsf1; Williams and Pataky Reference Williams and Pataky2010). The extent to which crops other than corn respond to a mixture of HPPD inhibitor and other herbicides has been studied. In an herbicide carryover study, a mixture of atrazine and mesotrione accentuated crop injury and yield losses in broccoli (Brassica oleracea var. italica), carrot (Daucus carota L.), cucumber (Cucumis sativus L.), and onion (Alium cepa L.; Robinson Reference Robinson2008). Grain sorghum was not injured by pyrasulfotole + bromoxynil applied alone or with synthetic auxin; however, the mixing of carfentrazone, a PPO inhibitor, increased phytotoxicity (Besançon et al. Reference Besançon, Riar, Heiniger, Weisz and Everman2016). Isoxaflutole + metribuzin applied PRE on isoxaflutole-resistant soybean injured the crop in environments with the most rainfall, and injury was often synergistic (Smith et al. Reference Smith, Soltani, Kaastra, Hooker, Robinson and Sikkema2019b). Sugarcane displayed transient injury symptoms when topramezone was mixed with ametryn or metribuzin compared with topramezone applied alone (Negrisoli et al. Reference Negrisoli, Odero, MacDonald, Sellers and Laughinghouse2020).

Future of HPPD-Inhibiting Herbicides

HPPD-inhibitor herbicides continue to be researched, patented, and commercialized by agrochemical companies. Benquitrione is currently in development for use in sorghum. A new class of HPPD inhibitor, azole carboxamides, were first disclosed by Bayer Crop Science in 2011 (Koehn et al. Reference Koehn, Tiebes, Van Almsick, Ahrens, Heinemann, Braun, Schmitt, Willms, Feucht and Rosinger2011). While no molecules have yet been commercialized, azole carboxamides have come to dominate the HPPD-inhibitor patent literature, with contributions from Syngenta, BASF, Nissan, KingAgroot, Nippon Soda, and SSARD, in addition to follow-up inventions from Bayer Crop Science. There have been more than 120 international patent applications for azole carboxamides, which represent applications of more than 50% of all HPPD-inhibiting herbicides since 2012. Herbicides from this class are in development that are expected to be available commercially around 2030. Azole carboxamides have different physical properties compared to the previously described classes with a different metabolism, which could potentially overcome non-target-site resistance (Jugulam and Shyam Reference Jugulam and Shyam2019). Patent applications that describe new herbicidal active ingredients will typically include thousands of compounds. However, companies will typically include a low number, often one, of these molecules in additional patent applications for use in mixtures with other active ingredients, for use in herbicide-resistant crops, or for inventions in the synthetic process. These additional patents hint to these specific molecules being of particular interest and likely candidates for further development (Figure 10).

Figure 10. Examples of recently submitted patents for herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD) from Bayer Crop Science, Nissan, and KingAgroot.

Future of HPPD-Inhibitor-Resistant Crops

Research and development of HPPD-inhibitor-resistant crop traits began in the early 2000s. Traits were initially created that were useful in PRE weed control programs. Due to the commercial success of glyphosate-resistant crops, the impact of the early HPPD-inhibitor-resistant crop traits were not as large as anticipated. In the last decade, however, interest in developing HPPD-inhibitor-resistant crop traits has re-emerged and is being driven by the impact of wide-scale occurrence of glyphosate-resistant weeds. This has led to the need for alternatives, and new HPPD-inhibitor crop traits are actively being developed across the crop protection industry, primarily for use in soybean and cotton.

Certain grass crop species such as corn are resistant to most HPPD-inhibitor herbicides (Mitchell et al. Reference Mitchell, Bartlett, Fraser, Hawkes, Holt, Townson and Wichert2001); therefore, HPPD inhibitors such as mesotrione can be applied PRE and POST in corn, POST in oats and sugarcane, but it is labeled only for PRE weed control in sorghum. Nonetheless, HPPD-inhibitor-resistant lines in dicot species [e.g., tobacco (Nicotiana tabacum L.) and soybean that would be otherwise highly sensitive to these herbicides] have been developed. Tobacco transformed with an HPPD gene from wheat showed resistance to mesotrione (Hawkes et al. Reference Hawkes, Holt, Andrews, Thomas, Langford, Hollingworth and Mitchell2001, Reference Hawkes, Langford, Viner, Blain, Callaghan, Mackay, Hogg, Singh and Dale2019). Siehl et al. (Reference Siehl, Tao, Albert, Dong, Heckert, Madrigal, Lincoln-Cabatu, Lu, Fenwick, Bermudez, Sandoval, Horn, Green, Hale, Pagano, Clark, Udranszky, Rizzo, Bourett, Howard, Johnson, Vogt, Akinsola and Castle2014) developed transgenic soybean that is resistant to isoxaflutole, mesotrione, and tembotrione with increased selectivity and a wide spectrum of weed control. In addition, isoxaflutole-resistant soybean has been developed and is available for commercial cultivation in the United States; however, its adoption is limited due to restriction in use of isoxaflutole (Alite 27) in certain counties in states such as Nebraska (Mausbach et al. Reference Mausbach, Irmak, Sarangi, Lindquist and Jhala2021).

Bayer Crop Science has a long history of involvement in the development of HPPD-inhibitor-resistant crop traits, with efforts mainly focused on the expression of an herbicide-insensitive bacterial HPPD enzyme from Pseudomonas fluorescens. A mutated form of the gene that carries a mutation at amino acid position G336W had increased tolerance to isoxaflutole. This gene is used in the FG72 soybean in commercial use (Matringe et al. Reference Matringe, Sailland, Pelissier, Rolland and Zink2005). Work initiated by the former Monsanto business (now part of Bayer Crop Science) also focused on development of an HPPD trait and a planned launch of this trait (HT4) in soybean is expected in the late 2020s. Details of the trait gene used are unknown at the time of writing. This trait is expected to be stacked with 2,4-D, glyphosate, glufosinate, and dicamba. A further development known as HT5 is expected to launch later, adding resistance to PPO-inhibiting herbicides (Reither Reference Reither2021).

BASF acquired much of Bayer Crop Science’s HPPD-inhibitor-resistant crop technology during the crop protection industry consolidation period in the mid-2010s. As such, BASF is now bringing to market products containing the HPPD trait from FG72 soybean into other crops such as cotton (Steadman Reference Steadman2021). A further development is the HPPD trait known as pfHPPD-4, which is expressed in the GMB151 soybean line. This is a Pseudomonas HPPD gene that carries four mutations compared to a single mutation that occurs in the FG72 trait (Olson and Weeks Reference Olson and Weeks2020).

Syngenta’s involvement with HPPD inhibitor-resistant crop traits dates to 2000, when several HPPD target genes, including the wild oat (Avena sativa L.) gene were characterized. The Avena sativa HPPD gene was later used to develop a soybean (SYT-0H2) that is resistant to mesotrione applied PRE (Hawkes et al. Reference Hawkes, Langford, Viner, Blain, Callaghan, Mackay, Hogg, Singh and Dale2019). Syngenta has recently disclosed the invention of a series of further evolved Avena HPPD target site genes that have much enhanced resistance to a broad range of HPPD inhibitors including mesotrione and bicyclopyrone (Hawkes et al. Reference Hawkes, Langford, Viner, Blain, Callaghan, Mackay, Hogg, Singh and Dale2019). These genes are capable of providing resistance to POST applications of these herbicides (Hawkes et al. Reference Hawkes, Langford, Viner, Blain, Callaghan, Mackay, Hogg, Singh and Dale2019). Plant Arc Bio has applied for an exemption for an HPPD-inhibitor trait based on a fungal (Trichoderma harzianum spp.) HPPD gene. The target crops are soybean and cotton (PlantArcBio 2022; Shatlin Reference Shatlin2021), although the spectrum of herbicide resistance of this trait is unknown as of 2022. An alternative technology has been described by the NARO Institute in Japan, which involves the metabolic degradation of certain HPPD-inhibitor formulas such as mesotrione. The His-1 gene was discovered as a part of a project to study the differences in herbicide sensitivity among rice cultivars. The metabolic nature of the gene means that this trait is likely to be narrower in the resistance spectrum compared with target-site-based approaches (Maeda et al. Reference Maeda, Murata, Sakuma, Takei, Yamazaki, Karim, Kawata, Hirose, Kawagishi-Kobayashi, Taniguchi, Suzuki, Sekino, Ohshima, Kato, Yoshida and Tozawa2019). Given the renewed investment in the development of HPPD-inhibitor-resistant crops, it is expected that HPPD-inhibiting herbicides will play an important role in weed control programs in soybean and cotton from the late 2020s onward. If such traits can be combined with the next generation of HPPD-inhibitors, their usefulness will likely be extended into the future. Management of multiple herbicide–resistant crop volunteers might be challenging, and future research should focus on this topic (Jhala et al. Reference Jhala, Beckie, Peters, Culpepper and Norsworthy2021c).

Acknowledgments

This project was partially supported by the Nebraska Agricultural Experiment Station with funding from the Hatch Act through the U.S. Department of Agriculture (USDA) National Institute of Food and Agriculture Project No. NEB-22-396. This project was also supported by the USDA–National Institute of Food and Agriculture Nebraska Extension Implementation Program. No conflicts of interest have been declared.

Footnotes

Associate Editor: William Johnson, Purdue University

References

Abendroth, JA, Martin, AR, Roeth, FW (2006) Plant response to combinations of mesotrione and photosystem II inhibitors. Weed Technol 20:267274 CrossRefGoogle Scholar
Abit, MJ, Al-Khatib, K, Currie, RS, Stahlman, PW, Geier, PW, Gordon, BW, Olson, BL, Claassen, MM, Regehr, DL (2010) Effect of postemergence mesotrione application timing on grain sorghum. Weed Technol 24:8590 CrossRefGoogle Scholar
Ahrens, H, Lange, G, Müller, T, Rosinger, C, Willms, L, van Almsick, A (2013) 4-Hydroxyphenylpyruvate dioxygenase inhibitors in combination with safeners: Solutions for modern and sustainable agriculture. Angew Chem Int Ed 52:93889398 CrossRefGoogle ScholarPubMed
Armel, GR, Wilson, HP, Richardson, RJ, Hines, TE (2003) Mesotrione combinations in no-till corn (Zea mays). Weed Technol 17:111116 CrossRefGoogle Scholar
Armel, GR, Hall, GJ, Wilson, HP, Cullen, N (2005) Mesotrione plus atrazine mixtures for control of Canada thistle (Cirsium arvense). Weed Sci 53:202211 CrossRefGoogle Scholar
Armel, GR, Rardon, PL, McComrick, MC, Ferry, NM (2007) Differential response of several carotenoid biosynthesis inhibitors in mixtures with atrazine. Weed Technol 21:947953 CrossRefGoogle Scholar
Beaudegnies, R, Edmunds, AJ, Fraser, TE, Hall, RG, Hawkes, TR, Mitchell, G, Schaetzer, J, Wendeborn, S, Wibley, J. (2009) Herbicidal 4-hydroxyphenylpyruvate dioxygenase inhibitors—a review of the triketone chemistry story from a Syngenta perspective. Bioorg Med Chem 17:41344152 CrossRefGoogle ScholarPubMed
Benoit, L, Soltani, N, Hooker, DC, Robinson, DE, Sikkema, PH (2019) Efficacy of HPPD-inhibiting herbicides applied preemergence or postemergence for control of multiple herbicide resistant waterhemp [Amaranthus tuberculatus (Moq.) Sauer]. Can J Plant Sci 99:379383 CrossRefGoogle Scholar
Besançon, TE, Riar, R, Heiniger, RW, Weisz, R, Everman, WJ (2016) Weed control and crop tolerance with pyrasulfotole plus bromoxynil combinations in grain sorghum. Int J Agron Crop Sci 1:1427 Google Scholar
Bollman, SL, Kells, JJ, Penner, D (2006). Weed response to mesotrione and atrazine applied alone and in combination preemergence. Weed Technol 20:903907 CrossRefGoogle Scholar
Brosnan, JT, Breeden, GK (2013) Bermudagrass (Cynodon dactylon) control with topramezone and triclopyr. Weed Technol 27:138142 CrossRefGoogle Scholar
Chahal, PS, Aulakh, JS, Jugulam, M, Jhala, AJ (2015) Herbicide resistant Palmer amaranth (Amaranthus palmeri S. Wats.) in the United States—Mechanisms of resistance, impact, and management. In Price, A, ed., Herbicides, agronomic crops, and weed biology. Rijeka, Croatia: In Tech Google Scholar
Chahal, PS, Ganie, ZA, Jhala, AJ (2018a) Overlapping residual herbicides for control of photosystem (PS) II- and 4-hydroxyphenylpyruvate dioxygenase (HPPD)-inhibitor-resistant Palmer amaranth (Amaranthus palmeri S. Watson) in glyphosate-resistant maize. Front Plant Sci 8: 2231 CrossRefGoogle Scholar
Chahal, PS, Irmak, S, Gaines, T, Amundsen, K, Jugulam, M, Jha, P, Travlos, IS, Jhala, AJ (2018b) Control of photosystem II–and 4-hydroxyphenylpyruvate dioxygenase inhibitor–resistant Palmer amaranth (Amaranthus palmeri) in conventional corn. Weed Technol 32:326335 CrossRefGoogle Scholar
Chahal, PS, Jhala, AJ (2018a) Interaction of PS II- and HPPD-inhibiting herbicides for control of Palmer amaranth resistant to both herbicide sites of action. Agron J 110:24962506 CrossRefGoogle Scholar
Chahal, PS, Jhala, AJ (2018b) Economics of management of photosystem II and HPPD inhibitor-resistant Palmer amaranth in corn. Agron J 110:19051914 CrossRefGoogle Scholar
Chahal, PS, Jugulam, M, Jhala, AJ (2019) Basis of atrazine and mesotrione synergism for controlling atrazine- and HPPD inhibitor-resistant Palmer amaranth. Agron J 111:32653273 CrossRefGoogle Scholar
Choe, E, Williams, MM II, Boydston, RA, Huber, JL, Huber, SC, Pataky, JK (2014) Photosystem II-inhibitors play a limited role in sweet corn response to 4-hydroxyphenyl pyruvate dioxygenase-inhibiting herbicides. Agron J 106:13171323 CrossRefGoogle Scholar
Colby, SR (1967) Calculating synergistic and antagonistic responses of herbicide combinations. Weeds 15:2022 CrossRefGoogle Scholar
Concepcion, JC, Kaundun, SS, Morris, JA, Hutchings, S-J, Strom, SA, Lygin, AV, Riechers, DE (2021) Resistance to a nonselective 4-hydroxyphenylpyruvate dioxygenase-inhibiting herbicide via novel reduction–dehydration–glutathione conjugation in Amaranthus tuberculatus . New Phytol 232:20892105 CrossRefGoogle ScholarPubMed
Creech, JE, Monaco, TA, Evans, JO (2004) Photosynthetic and growth responses of Zea mays L and four weed species following post-emergence treatments with mesotrione and atrazine. Pest Manag Sci 60:10791084 CrossRefGoogle Scholar
Currie, RS, Geier, P (2016) Fallow weed control with preemergence applications of Balance Pro, Corvus, Banvel, Atrazine, and Authority MTZ. Kansas Agricultural Experiment Station Research Reports, Vol 2(7). DOI: 0.4148/2378-5977.1263 Google Scholar
Currie, RS, Geier, P (2018) Efficacy of mesotrione-based tank mixtures and application timings compared to standards in irrigated corn. Kansas Agricultural Experiment Station Research Reports: Vol 4(8). DOI: 10.4148/2378-5977.7640 CrossRefGoogle Scholar
Dayan, FE, Duke, SO, Sauldubois, A, Singh, N, McCurdy, CR, Cantrell, CL (2007) p-Hydroxyphenylpyruvate dioxygenase is a herbicidal target site for β-triketones from Leptospermum scoparium . Phytochemistry 68:20042014 CrossRefGoogle ScholarPubMed
Diggle, AJ, Neve, PB, Smith, FP (2003) Herbicides used in combination can reduce the probability of herbicide resistance in finite weed populations. Weed Res 43:371382 CrossRefGoogle Scholar
Ditschun, S, Soltani, N, Robinson, DE, Tardif, FJ, Kaastra, AC, Sikkema, PH (2016) Control of glyphosate-resistant Canada fleabane (Conyza canadensis L. Cronq.) with isoxaflutole and metribuzin tank mix. Can J Plant Sci 96:7280 CrossRefGoogle Scholar
Evans, CM, Strom, SA, Riechers, DE, Davis, AS, Tranel, PJ, Hager, AG (2019) Characterization of a waterhemp (Amaranthus tuberculatus) population from Illinois resistant to herbicides from five site-of-action groups. Weed Technol 33: 400410 CrossRefGoogle Scholar
Evans, JA, Tranel, PJ, Hager, AG, Schutte, B, Wu, C, Chatham, LA, Davis, AS (2016) Managing the evolution of herbicide resistance. Pest Manag Sci 72:7480 CrossRefGoogle ScholarPubMed
Fluttert, JC, Soltani, N, Galla, M, Hooker, DC, Robinson, DE, Sikkema, PH (2022) Additive and synergistic interactions of 4-hydroxyphenylpyruvate dioxygenase (HPPD)- and photosystem II (PSII)-inhibitors for the control of glyphosate-resistant horseweed (Conyza canadensis) in corn. Weed Sci 70:319327 CrossRefGoogle Scholar
Ganie, ZA, Jhala, AJ (2017) Confirmation of glyphosate-resistant common ragweed (Ambrosia artemisiifolia) in Nebraska and response to postemergence corn and soybean herbicides. Weed Technol 31:225237 CrossRefGoogle Scholar
Ganie, ZA, Stratman, G, Jhala, AJ (2015) Response of selected glyphosate-resistant broadleaved weeds to premix of fluthiacet-methyl and mesotrione (SolsticeTM) applied at two growth stages. Can J Plant Sci 95:861869 CrossRefGoogle Scholar
Grossmann, 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 63:429439 CrossRefGoogle ScholarPubMed
Hager, AG, Wax, LM, Simmons, FW, Stoller, EW (1997) Waterhemp management in agronomic crops. Champaign: University of Illinois Bulletin, X855. 12 pGoogle Scholar
Hatzios, KK, Penner, D (1985) Interactions of herbicides with other agrochemicals in higher plants. Rev Weed Sci 1:163 Google Scholar
Hausman, NE, Singh, S, Tranel, PJ, Riechers, DE, Kaundun, SS, Polge, ND, Thomas, DA, Hager, AG (2011) Resistance to HPPD-inhibiting herbicides in a population of waterhemp (Amaranthus tuberculatus) from Illinois, United States. Pest Manag Sci 67:258261 CrossRefGoogle Scholar
Hausman, NE, Tranel, PJ, Riechers, DE, Maxwell, DJ, Gonzini, LC, Hager, AG (2013) Responses of an HPPD inhibitor-resistant waterhemp (Amaranthus tuberculatus) population to soil residual herbicides. Weed Technol 27:704711 CrossRefGoogle Scholar
Hawkes, TR, Langford, MP, Viner, R, Blain, RE, Callaghan, FM, Mackay, EA, Hogg, BV, Singh, S, Dale, RP (2019) Characterization of 4-hydroxyphenylpyruvate dioxygenases, inhibition by herbicides and engineering for herbicide tolerance in crops. Pestic Biochem Physiol 156:928 CrossRefGoogle ScholarPubMed
Hawkes, TR, Holt, DC, Andrews, CJ, Thomas, PG, Langford, MP, Hollingworth, S, Mitchell, G (2001) Mesotrione: mechanism of herbicidal activity and selectivity in corn. Vol 2, pages 563–568 in The BCPC Proceedings - Weeds, British Crop Protection Council. Brighton, UK, November 12–15, 2001Google Scholar
Heap, I (2022) The International Herbicide-Resistant Weed Database. http://www.weedscience.org. Accessed: March 11, 2022Google Scholar
Hess, FD (2000) Light-dependent herbicides: an overview. Weed Sci 48:160170 Google Scholar
Hugie, JA, Bollero, GA, Tranel, PJ, Riechers, DE (2008) Defining the rate requirements for synergism between mesotrione and atrazine in redroot pigweed (Amaranthus retroflexus). Weed Sci 56:265270 CrossRefGoogle Scholar
Jacobs, KE Jr, Butts-Wilmsmeyer, CJ, Ma, R, O’Brien, SR, Riechers, DE (2020) Association between metabolic resistances to atrazine and mesotrione in a multiple-resistant waterhemp (Amaranthus tuberculatus) population. Weed Sci 68:358366 CrossRefGoogle Scholar
Janak, TW, Grichar, WJ (2016) Weed control in corn (Zea mays L.) as influenced by preemergence herbicides. Int J Agron https://doi.org/10.1155/2016/2607671. Accessed: May 2, 2022CrossRefGoogle Scholar
Jhala, AJ, Beckie, HJ, Mallory-Smith, C, Jasieniuk, M, Busi, R, Norsworthy, JK, Bagavathiannan, MV, Tidemann, BD, Geddes, CM (2021a) Transfer of resistance alleles from herbicide-resistant to susceptible grass weeds via pollen-mediated gene flow. Weed Technol 35:869885 CrossRefGoogle Scholar
Jhala, AJ, Beckie, HJ, Peters, T, Culpepper, S, Norsworthy, J (2021c) Interference and management of herbicide-resistant crop volunteers. Weed Sci 69:257273 CrossRefGoogle Scholar
Jhala, AJ, Knezevic, SZ, Ganie, ZA, Singh, M (2014a) Integrated weed management in corn (Zea mays L.). Pages 177196 in Chauhan, B, Mahajan, G, eds. Recent Advances in Weed Management. New York: Springer CrossRefGoogle Scholar
Jhala, AJ, Norsworthy, JK, Ganie, ZA, Sosnoskie, LM, Beckie, HJ, Mallory-Smith, CA, Liu, J, Wei, W, Wang, J, Stoltenberg, DE (2021b) Pollen-mediated gene flow and transfer of resistance alleles from herbicide-resistant broadleaf weeds. Weed Technol 35:173187 CrossRefGoogle Scholar
Jhala, AJ, Sandell, LD, Rana, N, Kruger, GR, Knezevic, SZ (2014b) Confirmation and control of triazine and 4-hydroxyphenylpyruvate dioxygenase-inhibiting herbicide-resistant Palmer amaranth (Amaranthus palmeri) in Nebraska. Weed Technol 28:2838 CrossRefGoogle Scholar
Johnson, BC, Young, BG, Matthews, JL (2002) Effect of postemergence application rate and timing of mesotrione on corn (Zea mays) response and weed control. Weed Technol 16:414420 CrossRefGoogle Scholar
Jugulam, M, Shyam, C (2019) Non-target site resistance to herbicides: Recent developments. Plants 8:417. https://doi.org/10.3390/plants8100417 Accessed: June 04, 2022CrossRefGoogle ScholarPubMed
Kaastra, AC, Swanton, CJ, Tardif, FJ, Sikkema, PH (2008) Two-way performance interactions among ρ-hydroxyphenylpyruvate dioxygenase- and acetolactate synthase-inhibiting herbicides. Weed Sci 56:841851 CrossRefGoogle Scholar
Kaundun, SS (2021) Syngenta’s contribution to herbicide resistance research and management. Pest Manag Sci 77:15641571 CrossRefGoogle ScholarPubMed
Kaundun, SS, Hutchings, S-J, Dale, RP, Howell, A, Morris, JA, Kramer, VC, Shivrain, VK, McIndoe, E (2017) Mechanism of resistance to mesotrione in an Amaranthus tuberculatus population from Nebraska, USA. PLOS ONE 12(6): e0180095 CrossRefGoogle Scholar
Keeley, PE, Carter, CH, Thullen, RJ (1987) Influence of planting date on growth of Palmer amaranth (Amaranthus palmeri). Weed Sci 35:199204 CrossRefGoogle Scholar
KingAgroot (2021) Cypyrafluone was approved and registered by Pakistan, KingAgroot officially entered the overseas market. https://www.kingagroot.com/En/html/news-21Dec07.html. Accessed: March 3, 2022Google Scholar
Koepke-Hill, RM, Armel, GR, Wilson, HP, Hines, TE, Vargas, JJ (2010) Herbicide combinations for control of volunteer potato. Weed Technol 24:9194 Google Scholar
Kohrt, JR, Sprague, CL (2017) Response of a multiple-resistant Palmer amaranth (Amaranthus palmeri) population to four HPPD-inhibiting herbicides applied alone and with atrazine. Weed Sci 65:534545 CrossRefGoogle Scholar
Kumar, V, Jha, P, Reichard, N (2014) Occurrence and characterization of kochia (Kochia scoparia) accessions with resistance to glyphosate in Montana. Weed Technol 28:122130 CrossRefGoogle Scholar
Kumar, V, Jha, P (2015) Effective preemergence and postemergence herbicide programs for kochia control. Weed Technol 29:2434 CrossRefGoogle Scholar
Koehn, A, Tiebes, J, Van Almsick, A, Ahrens, H, Heinemann, I, Braun, R, Schmitt, MH, Willms, L, Feucht, D, Rosinger, CH (2011) Preparation of N-(1,2,5-oxadiazol-3-yl) benzamides as herbicides. World Intellectual Property Organization. International Patent Application # WO2011035874. https://patentscope.wipo.int/search/en/detail.jsf?docId=WO2011035874&_cid=P20-L3ZT1G-82733-1. Accessed: June 4, 2022Google Scholar
Küpper, A, Peter, F, Zöllner, P, Lorentz, L, Tranel, PJ, Beffa, R, Gaines, TA (2018) Tembotrione detoxification in 4-hydroxyphenylpyruvate dioxygenase (HPPD) inhibitor-resistant Palmer amaranth (Amaranthus palmeri S. Wats.). Pest Manag Sci 74: 23252334 CrossRefGoogle Scholar
Lee, DL, Prisbylla, MP, Cromartie, TH, Dagarin, DP, Howard, SW, Provan, WM, Ellis, MK, Fraser, T, Mutter, LC (1997) The discovery and structural requirements of inhibitors of p-hydroxyphenylpyruvate dioxygenase. Weed Sci 45:601609 CrossRefGoogle Scholar
Lindley, JJ, Jennings, KM, Monks, DW, Chaudhari, S, Schultheis, JR, Waldschmidt, M, Brownie, C (2020) Effect of bicyclopyrone herbicide on sweetpotato and Palmer amaranth (Amaranthus palmeri). Weed Technol 34:552559 CrossRefGoogle Scholar
Liu, M, Lu, S (2016) Plastoquinone and ubiquinone in plants: Biosynthesis, physiological function, and metabolic engineering. Front Plant Sci 7:1898 CrossRefGoogle ScholarPubMed
Liu, R, Kumar, V, Jhala, A, Jha, P, Stahlman, PW (2021) Control of glyphosate- and mesotrione-resistant Palmer amaranth in glyphosate- and glufosinate-resistant corn. Agron J 113:53625372 CrossRefGoogle Scholar
Lu, H, Yu, Q, Han, H, Owen, MJ, Powles, SB (2020) Evolution of resistance to HPPD-inhibiting herbicides in a wild radish population via enhanced herbicide metabolism. Pest Manag Sci 76:19291937 CrossRefGoogle Scholar
Lygin, AV, Kaundun, SS, Morris, JA, Mcindoe, E, Hamilton, AR, Riechers, DE (2018) Metabolic pathway of topramezone in multiple-resistant waterhemp (Amaranthus tuberculatus) differs from naturally tolerant maize. Front Plant Sci 9:1644 CrossRefGoogle ScholarPubMed
Ma, R, Kaundun, SS, Tranel, PJ, Riggins, CW, McGinness, DL, Hager, AG, Hawkes, T, McIndoe, E, Riechers, DE (2013) Distinct detoxification mechanisms confer resistance to mesotrione and atrazine in a population of waterhemp. Plant Physiol 163:363377 CrossRefGoogle Scholar
Maeda, H, Murata, K, Sakuma, N, Takei, S, Yamazaki, A, Karim, MR, Kawata, M, Hirose, S, Kawagishi-Kobayashi, M, Taniguchi, Y, Suzuki, S, Sekino, K, Ohshima, M, Kato, H, Yoshida, H, Tozawa, Y (2019) A rice gene that confers broad-spectrum resistance to β-triketone herbicides. Science 365:393396 CrossRefGoogle Scholar
Mallory-Smith, C, Retzinger, EJ (2017) Revised classification of herbicides by site of action for weed resistance management strategies. Weed Technol 17:605619 Google Scholar
Matringe, M, Sailland, A, Pelissier, B, Rolland, A, Zink, O (2005) Hydroxyphenylpyruvate dioxygenase inhibitor-resistant plants. Pest Manag Sci 61:269276 CrossRefGoogle ScholarPubMed
Mausbach, J, Irmak, S, Sarangi, D, Lindquist, J, Jhala, AJ (2021) Control of acetolactate synthase inhibitor/glyphosate-resistant Palmer amaranth (Amaranthus palmeri) in isoxaflutole/glufosinate/glyphosate-resistant soybean. Weed Technol 35:779785 CrossRefGoogle Scholar
Metzger, BA, Soltani, N, Raeder, AJ, Hooker, DC, Robinson, DE, Sikkema, PH (2018) Tolpyralate efficacy: Part 1. Biologically effective dose of tolpyralate for control of annual grass and broadleaf weeds in corn. Weed Technol 32:698706 CrossRefGoogle Scholar
Mitchell, G, Bartlett, DW, Fraser, TE, Hawkes, TR, Holt, DC, Townson, JK, Wichert, RA (2001) Mesotrione: a new selective herbicide for use in maize. Pest Manag Sci 57:120128 3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Mohler, CL, Teasdale, JR, DiTommaso, A (2021) Manage weed on your farm- A guide to ecological strategies. SARE handbook series # 16. https://www.sare.org/resources/manage-weeds-on-your-farm/. Accessed: April 4, 2022Google Scholar
Moore (2019) Evaluation of topramezone for use in rice (Oryza sativa L.) production. Master’s thesis. Fayetteville: University of Arkansas. https://scholarworks.uark.edu/etd/3436 Accessed: April 14, 2022Google Scholar
Nakka, S, Godar, AS, Wani, PS, Thompson, CR, Peterson, DE, Roelofs, J, Jugulam, M (2017) Physiological and molecular characterization of hydroxyphenylpyruvate dioxygenase (HPPD)-inhibitor resistance in Palmer amaranth (Amaranthus palmeri S. Wats.). Frontiers Plant Sci 8:555 CrossRefGoogle Scholar
Negrisoli, RM, Odero, DC, MacDonald, GE, Sellers, BA, Laughinghouse, HD (2020) Sugarcane response and fall panicum (Panicum dichotomiflorum) control with topramezone and triazine herbicides. Weed Technol 34:241249 CrossRefGoogle Scholar
Norris, SR, Barrette, TR, DellaPenna, D (1995) Genetic dissection of carotenoid synthesis in Arabidopsis defines plastoquinone as an essential component of phytoene desaturation. Plant Cell 7:21392149 Google ScholarPubMed
Norsworthy, JK, Ward, SM, Shaw, DR, Llewellyn, RS, Nichols, RL, Webster, TM, Bradley, KW, Frisvold, G, Powles, SB, Burgos, NR, Witt, WW, Barrett, M (2012) Reducing the risks of herbicide resistance: Best management practices and recommendations. Weed Sci 60(SPI):3162 CrossRefGoogle Scholar
O’Brien, SR, Davis, AS, Riechers, DE (2018) Quantifying resistance to isoxaflutole and mesotrione and investigating their interactions with metribuzin POST in waterhemp (Amaranthus tuberculatus). Weed Sci 66:586594 CrossRefGoogle Scholar
Oliveira, MC, Jhala, AJ, Gaines, T, Irmak, S, Amundsen, K, Scott, JE, Knezevic, SZ (2017) Confirmation and control of HPPD-inhibiting herbicide–resistant waterhemp (Amaranthus tuberculatus) in Nebraska. Weed Technol 31:6779 CrossRefGoogle Scholar
Olson, A, Weeks, M (2020) Petition for a determination of nonregulated status for a plant-parasitic nematode-protected and herbicide tolerant GMB151 soybean. https://www.aphis.usda.gov/brs/aphisdocs/19_31701p.pdf Accessed: May 18, 2022Google Scholar
Osipitan, OA, Scott, JE, Knezevic, SZ (2018) Tolpyralate applied alone and with atrazine for weed control in corn. J Agric Sci 10:3239 Google Scholar
Owens, DK, Nanayakkara, NPD, Dayan, FE (2013) In planta mechanism of action of leptospermone: Impact of its physico-chemical properties on uptake, translocation, and metabolism. J Chem Ecol 39:262270 CrossRefGoogle ScholarPubMed
Pallett, KE, Cramp, SM, Little, JP, Veerasekaran, P, Crudace, AJ, Slater, AE (2001) Isoxaflutole: the background to its discovery and the basis of its herbicidal properties. Pest Manag Sci 57:133142 3.0.CO;2-0>CrossRefGoogle ScholarPubMed
Pallett, KE, Little, JP, Sheekey, M, Veerasekaran, P (1998) The mode of action of isoxaflutole: I. physiological effects, metabolism, and selectivity. Pest Biochem Physiol 62:113124 CrossRefGoogle Scholar
Pandian, BA, Varanasi, A, Vannapusa, AR, Sathishraj, R, Lin, G, Zhao, M, Tunnell, M, Tesso, T, Liu, S, Prasad, PV, Jugulam, M (2020) Characterization, genetic analyses, and identification of QTLs conferring metabolic resistance to a 4-hydroxyphenylpyruvate dioxygenase inhibitor in sorghum (Sorghum bicolor). Front Plant Sci 11:596581 CrossRefGoogle ScholarPubMed
PlantArcBio (2022) PlantArcBio Receives USDA/APHIS Regulatory Exemption in the US for its Proprietary HPPD Herbicide-Tolerant Gene for Soybean and Cotton. CISION PR Newswire. https://www.prnewswire.com/il/news-releases/plantarcbio-receives-usdaaphis-regulatory-exemption-in-the-us-for-its-proprietary-hppd-herbicide-tolerant-gene-for-soybean-and-cotton-301454516.html. Accessed: July 20, 2022Google Scholar
Reddy, SS, Stahlman, PW, Geier, PW, Thompson, CR, Currie, RS, Schlegel, AJ, Olson, BL, Lally, NG (2013) Weed control and crop safety with premixed pyrasulfotole and bromoxynil in grain sorghum. Weed Technol 27:664670 CrossRefGoogle Scholar
Reither, B (2021) R&D pipeline update: The beginning of what’s next. https://www.bayer.com/sites/default/files/BayerCMD2021_CropScience_RandD_Presentation.pdf. Accessed: July 20, 2022Google Scholar
Robinson, DE (2008) Atrazine accentuates carryover injury from mesotrione in vegetable crops. Weed Technol 22:641645 CrossRefGoogle Scholar
Sarangi, D, Jhala, AJ (2018) Comparison of a premix of atrazine, bicyclopyrone, mesotrione, and s-metolachlor with other preemergence herbicides for weed control and corn yield in no-tillage and reduced-tillage production systems in Nebraska, USA. Soil Tillage Res 178:8291 Google Scholar
Sarangi, D, Stephens, T, Barker, AL, Patterson, EL, Gaines, TA, Jhala, AJ (2019) Protoporphyrinogen oxidase (PPO) inhibitor–resistant waterhemp (Amaranthus tuberculatus) from Nebraska is multiple herbicide resistant: confirmation, mechanism of resistance, and management. Weed Sci 67:510520 CrossRefGoogle Scholar
Schultz, JL, Chatham, LA, Riggins, CW, Tranel, PJ, Bradley, KW (2015) Distribution of herbicide resistances and molecular mechanisms confer-ring resistance in Missouri waterhemp (Amaranthus rudis Sauer) populations. Weed Sci 63:336345 CrossRefGoogle Scholar
Schuster, CL, Al-Khatib, K, Dille, JA (2007) Mechanism of antagonism of mesotrione on sulfonylurea herbicides. Weed Sci 55:429434 CrossRefGoogle Scholar
Schuster, CL, Al-Khatib, K, Dille, JA (2008) Efficacy of sulfonylurea herbicides when tank mixed with mesotrione. Weed Technol 22:222230 CrossRefGoogle Scholar
Schultz, C, Soll, J, Fiedler, E, Schultze-Siebert, D (1985) Synthesis of prenylquinones in chloroplasts. Physiol Plantarum 64:123129 CrossRefGoogle Scholar
Shatlin, D (2021) Request 21-264-01cr: 7 CFR § 340.1(e) Request for confirmation that PlantArcBio’s HPPD herbicide tolerant soy is exempt from regulation under §340.1(c)(2). https://www.aphis.usda.gov/biotechnology/downloads/confirmation-response/21-264-01cr-a1.pdf. Accessed: July 20, 2022Google Scholar
Shyam, C, Borgato, EA, Peterson, DE, Dille, JA, Jugulam, M (2021) Predominance of metabolic resistance in a six-way-resistant Palmer amaranth (Amaranthus palmeri) population. Front Plant Sci 11:614618 CrossRefGoogle Scholar
Siehl, DL, Tao, Y, Albert, H, Dong, Y, Heckert, M, Madrigal, A, Lincoln-Cabatu, B, Lu, J, Fenwick, T, Bermudez, E, Sandoval, M, Horn, C, Green, JM, Hale, T, Pagano, P, Clark, J, Udranszky, IA, Rizzo, N, Bourett, T, Howard, RJ, Johnson, DH, Vogt, M, Akinsola, G, Castle, LA (2014) Broad 4-hydroxyphenylpyruvate dioxygenase inhibitor herbicide tolerance in soybean with an optimized enzyme and expression cassette. Plant Physiol 166:11621176 CrossRefGoogle ScholarPubMed
Smith, A, Soltani, N, Kaastra, AJ, Hooker, DC, Robinson, DE, Sikkema, PH (2019a) Annual weed management in isoxaflutole-resistant soybean using a two-pass weed control strategy. Weed Technol 33:411425 CrossRefGoogle Scholar
Smith, A, Soltani, N, Kaastra, AC, Hooker, DC, Robinson, DE, Sikkema, PH (2019b) Isoxaflutole and metribuzin interactions in isoxaflutole-resistant soybean. Weed Sci 67:485496 CrossRefGoogle Scholar
Soltani, N, Shropshire, C, Sikkema, PH (2011) Response of spring planted barley (Hordeum vulgare L.), oats (Avena sativa L.) and wheat (Triticum aestivum L.) to mesotrione. Crop Prot 30:849853 CrossRefGoogle Scholar
Soltani, N, Shropshire, C, Cowan, T, Sikkema, PH (2014) Weed management in spring planted cereals with mesotrione. Am J Plant Sci DOI: 10.4236/ajps.2014.51020 CrossRefGoogle Scholar
Steadman, J (2021) Inside cotton’s pipeline: What’s coming from BASF. Cotton Grower. https://www.cottongrower.com/crop-inputs/inside-cottons-pipeline-whats-coming-from-basf/. Accessed: July 20, 2022Google Scholar
Stephenson, DO, Bond, JA, Walker, ER, Bararpour, MT, Oliver, LR (2004) Evaluation of mesotrione in Mississippi Delta corn production. Weed Technol 18:11111116 CrossRefGoogle Scholar
Stephenson, DO, Bond, JA (2012) Evaluation of thiencarbazone-methyl– and isoxaflutole-based herbicide programs in corn. Weed Technol 26:3742 CrossRefGoogle Scholar
Stephenson, DO, Bond, JA, Landry, RL, Edwards, HM (2015) Weed management in corn with postemergence applications of tembotrione or thiencarbazone: tembotrione. Weed Technol 29:350358 CrossRefGoogle Scholar
Striegel, A, Jhala, AJ (2022) Economics of reducing Palmer amaranth (Amaranthus palmeri S. Watson) seed production in dicamba/glufosinate/glyphosate-resistant soybean. Agron J 114:25182540 CrossRefGoogle Scholar
Sutton, P, Richards, C, Buren, L, Glasgow, L (2002) Activity of mesotrione on resistant weeds in maize. Pest Manag Sci 58:981984 CrossRefGoogle ScholarPubMed
Thompson, CR, Peterson, ED, Nathan, GL (2012) Characterization of HPPD-resistant Palmer amaranth. Abstract No. 413 in Proceedings of the Weed Science Society of America annual meeting. Waikoloa, Hawaii, February 6–9, 2012Google Scholar
Torbiak, AT, Blackshaw, RE, Brandt, RN, Hamman, B, Geddes, CM (2021) Herbicide strategies for managing glyphosate-resistant and -susceptible kochia (Bassia scoparia) in spring wheat. Can J Plant Sci 101:607620 CrossRefGoogle Scholar
Tranel, PJ (2021) Herbicide resistance in Amaranthus tuberculatus . Pest Manag Sci 77:4354 CrossRefGoogle ScholarPubMed
Tsukamoto, M, Kikugawa, H, Nagayama, S, Suganuma, T, Okita, T, Miyamoto, H (2021) Discovery and structure optimization of a novel corn herbicide, tolpyralate. J Pestic Sci 46:152159 CrossRefGoogle ScholarPubMed
[USDA-NASS] United States Department of Agriculture–National Agricultural Statistics Service (2018) Agricultural chemical use. https://quickstats.nass.usda.gov. Accessed: June 13, 2022Google Scholar
Varanasi, VK, Brabham, C, Norsworthy, JK, Nie, H, Young, BG, Houston, M, Barber, T, Scott, RC (2018) A statewide survey of PPO-inhibitor resistance and the prevalent target-site mechanisms in Palmer amaranth (Amaranthus palmeri) accessions from Arkansas. Weed Sci 66:149158 CrossRefGoogle Scholar
Walsh, MJ, Stratford, K, Stone, K, Powles, SB (2012) Synergistic effects of atrazine and mesotrione on susceptible and resistant wild radish (Raphanus raphanistrum) populations and the potential for overcoming resistance to triazine herbicides. Weed Technol 26:341347 CrossRefGoogle Scholar
Walsh, MJ, Newman, P, Chatfield, P (2021) Mesotrione: a new preemergence herbicide option for wild radish (Raphanus raphanistrum) control in wheat. Weed Technol 35:924931 CrossRefGoogle Scholar
Wang, DW, Lin, HY, Cao, RJ, Ming, ZZ, Chen, T, Hao, GF, Yang, WC, Yang, GF (2015) Design, synthesis and herbicidal activity of novel quinazoline-2,4-diones as 4-hydroxyphenylpyruvate dioxygenase inhibitors. Pest Manag Sci 71:11221132 CrossRefGoogle Scholar
Wang, H, Liu, W, Jin, T, Peng, X, Zhang, L, Wang, J (2020) Bipyrazone: a new HPPD inhibiting herbicide in wheat. Sci Rep 10:5521 CrossRefGoogle ScholarPubMed
Whaley, CM, Armel, GR, Wilson, HP, Hines, TE (2006) Comparison of mesotrione combinations with standard weed control programs in corn. Weed Technol 20:605611 CrossRefGoogle Scholar
Willemse, C, Soltani, N, Benoit, L, Jhala, AJ, Hooker, DC, Robinson, DE, Sikkema, PH (2021a) Is there a benefit of adding atrazine to HPPD-inhibiting herbicides for control of multiple-herbicide-resistant, including group 5-resistant, waterhemp in corn? J Agr Sci 13:2131 Google Scholar
Willemse, C, Soltani, N, David, CH, Jhala, AJ, Robinson, DE, Sikkema, PH (2021b) Interaction of 4-hydroxyphenylpyruvate dioxygenase (HPPD) and atrazine alternative photosystem II (PS II) inhibitors for control of multiple herbicide–resistant waterhemp (Amaranthus tuberculatus) in corn. Weed Sci 69:492503 CrossRefGoogle Scholar
Willemse, C, Soltani, N, Metzger, B, Hooker, DC, Jhala, AJ, Robinson, DE, Sikkema, PH (2021c) Biologically-effective-dose of tolpyralate and tolpyralate plus atrazine for control of multiple-herbicide-resistant waterhemp [Amaranthus tuberculatus (Moq.) J. D. Sauer] in corn. Agr Sci 12:424443 Google Scholar
Williams, MM II, Pataky, JK, Nordby, JN, Riechers, DE, Sprague, CL, Masiunas, JB (2005) Cross-sensitivity in sweet corn to nicosulfuron and mesotrione applied postemergence. HortScience 40:18011805 CrossRefGoogle Scholar
Williams, MM II, Pataky, JK (2010) Factors affecting differential sensitivity of sweet corn to HPPD-inhibiting herbicides. Weed Sci 58:289294 CrossRefGoogle Scholar
Williams, MM II, Boydston, RA, Peachey, RE, Robinson, D (2011a) Performance consistency of reduced atrazine use in sweet corn. Field Crops Res 121:96104 CrossRefGoogle Scholar
Williams, MM II, Boydston, RA, Peachey, RE, Robinson, D (2011b) Significance of atrazine as a tank-mix partner with tembotrione. Weed Technol 25:299302 CrossRefGoogle Scholar
Woodyard, AJ, Hugie, JA, Riechers, DE (2009a) Interactions of mesotrione and atrazine in two weed species with different mechanisms for atrazine resistance. Weed Sci 57:369378 Google Scholar
Woodyard, AJ, Bollero, GA, Riechers, DE (2009b) Broadleaf weed management in corn utilizing synergistic postemergence herbicide combinations. Weed Technol 23:513518 CrossRefGoogle Scholar
Woodyard, AJ, Hugie, JA, Riechers, DE (2009c) Interactions of mesotrione and atrazine in two weed species with different mechanisms for atrazine resistance. Weed Sci 57:369378 CrossRefGoogle Scholar
Wuerffel, RJ, Young, JM, Tranel, PJ, Young, BG (2015) Soil-residual protoporphyrinogen oxidase-inhibiting herbicides influence the frequency of associated resistance in waterhemp (Amaranthus tuberculatus). Weed Sci 63:529538 CrossRefGoogle Scholar
Yamamoto, S, Tanetani, Y, Uchiyama, C, Nagamatsu, A, Kobayashi, M, Ikeda, M, Kawai, K (2021) Mechanism of action and selectivity of a novel herbicide, fenquinotrione. J Pestic Sci 46:249257 CrossRefGoogle ScholarPubMed
Yu, J, McCullough, PE (2016) Triclopyr reduces foliar bleaching from mesotrione and enhances efficacy for smooth crabgrass control by altering uptake and translocation. Weed Technol 30:516523 CrossRefGoogle Scholar
Yu, Q, Powles, S (2014) Metabolism-based herbicide resistance and cross resistance in crop weeds: A threat to herbicide sustainability and global crop production. Plant Physiol 166:11061118 CrossRefGoogle Scholar
Figure 0

Figure 1. Chemical structures of some herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD).

Figure 1

Figure 2. Chemical structures of pyrazolone herbicides, a chemical family of herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD).

Figure 2

Figure 3. Timeline of commercialization of herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD), their respective chemical classes, and manufacturer.

Figure 3

Figure 4. Chemical structures of triketone herbicides, a chemical family of herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD).

Figure 4

Figure 5. Annual use of major herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD) in corn production in the United States in 2018 (Source: USDA-NASS 2018).

Figure 5

Figure 6. Mesotrione use in agricultural land across the United States in 2018 (adapted from the U.S. Geological Survey by the U.S. Department of the Interior).

Figure 6

Figure 7. Tembotrione use in various corn-producing states in the United States (Source: USDA-NASS 2018).

Figure 7

Figure 8. Isoxaflutole use in major corn-producing states in the United States (Source: USDA-NASS 2018).

Figure 8

Figure 9. Bicyclopyrone and topramezone use in various corn-producing states in the United States (Source: USDA-NASS 2018).

Figure 9

Figure 10. Examples of recently submitted patents for herbicides that inhibit 4-hydroxyphenylpyruvate dioxygenase (HPPD) from Bayer Crop Science, Nissan, and KingAgroot.