i BIOCHEMICAL, PHYSIOLOGICAL, AND GENETIC INVESTIGATIONS OF MULTIPLE HERBICIDE RESISTANT AVENA FATUA L. by Lucas Arlin Wright A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in Plant Sciences MONTANA STATE UNIVERSITY Bozeman, Montana December 2023 ©COPYRIGHT by Lucas Arlin Wright 2023 All Rights Reserved ii ACKNOWLEDGEMENTS I would like to express my deepest gratitude to my advisor, Dr. William Dyer, whose invaluable guidance and support made it possible for me to pursue this path. A special thank you to Dr. Barbara Keith, whose assistance was crucial throughout this journey. My utmost appreciation goes to my committee members, Dr. Jennifer Lachowiec and Dr. Jason Cook, for their instrumental role in my academic and professional growth. I am also grateful to the faculty who provided the knowledge and resources necessary for achieving this milestone. A big thank you to the many student workers whose help has been crucial in my projects. To my fellow students, thank you for the friendships and camaraderie we have shared. My family deserves my deepest thanks for their support and encouragement over the years. Lastly, a very special acknowledgment to Sally Johnson, whose unwavering support and ability to keep me grounded throughout this journey. Finally, thank you to the many organizations that funded this research, the USDA-NIFA- AFRI, the US EPA Strategic Agricultural Initiative, the Montana Wheat and Barley Committee, and the Montana Agricultural Experiment Station. iii TABLE OF CONTENTS INTRODUCTION .................................................................................................................... 1 Multiple Herbicide Resistant Avena fatua .......................................................................... 1 Volatile Organic Compounds .................................................................................. 5 Plant Pigments and Non-Photochemical Quenching .............................................. 8 Quantitative Trait Loci Analysis of MHR Avena fatua ......................................... 13 Research Objectives .............................................................................................. 15 Project Significance .............................................................................................. 16 INSECT GENERALIST PREFERENCE TO MHR AVENA FATUA ..................................... 17 Summary ........................................................................................................................... 17 Introduction ....................................................................................................................... 18 Methods............................................................................................................................. 20 Plant Material ........................................................................................................ 20 Spodoptera exigua Choice and No-choice Herbivory Experiments ..................... 20 Greenhouse Experiments .......................................................................... 20 Field Experiments ..................................................................................... 21 Measurements ........................................................................................... 22 Spodoptera exigua Olfaction Choice Assay ......................................................... 22 Statistics ................................................................................................................ 23 Results ............................................................................................................................... 24 Greenhouse Choice Experiments .......................................................................... 24 Greenhouse No-choice Experiments ..................................................................... 26 Field Choice and No-choice Experiments ............................................................ 26 Olfactory Preferences of S. exigua........................................................................ 28 Discussion ......................................................................................................................... 31 Conclusion ........................................................................................................................ 35 EXOGENOUS VOLATILE ORGANIC COMPOUNDS EFFECTS ON HERBICIDE RESISTANCE IN TWO AVENA FATUA LINES ............................................. 36 Summary ........................................................................................................................... 36 Introduction ....................................................................................................................... 37 Methods............................................................................................................................. 39 Plant Material ........................................................................................................ 39 Exogenous VOC Treatments ................................................................................. 40 Herbicide Treatments ............................................................................................ 41 Statistics ................................................................................................................ 41 Results ............................................................................................................................... 42 Open Bag Experiments ......................................................................................... 42 CO2 Concentrations .............................................................................................. 44 iv TABLE OF CONTENTS CONTINUED Discussion ......................................................................................................................... 51 Conclusion ........................................................................................................................ 57 PLANT PIGMENTS DETECTION, PHOTOSYNTHETIC AND ENERGY MANAGEMENT PARAMETERS OF MULTIPLE HERBICIDE RESISTANT AVENA FATUA ................................................................................................. 58 Summary ........................................................................................................................... 58 Introduction ....................................................................................................................... 59 Methods............................................................................................................................. 61 Plant Material ........................................................................................................ 61 Thin Layer Chromatography................................................................................. 61 Spectrophotometer Assay...................................................................................... 62 Pulse Amplitude Modulated Fluorometry ............................................................. 62 Statistics ................................................................................................................ 63 Results ............................................................................................................................... 64 Thin Layer Chromatography................................................................................. 64 Spectrophotometer Assay...................................................................................... 65 Pulse Amplitude Modulated Fluorometry Assay .................................................. 68 Discussion ......................................................................................................................... 73 Conclusion ........................................................................................................................ 77 QUANTITATIVE TRAIT LOCI ANALYSIS OF MULTIPLE HERBICIDE RESISTANT AVENA FATUA ................................................................................................. 79 Summary ........................................................................................................................... 79 Introduction ....................................................................................................................... 80 Methods............................................................................................................................. 82 Plant Material and Experimental Design .............................................................. 82 Phenotypic Data .................................................................................................... 82 Genotyping and Linkage Map Construction ......................................................... 83 QTL Analysis ........................................................................................................ 84 RNA-seq Analysis ................................................................................................. 84 Results ............................................................................................................................... 85 Genotyping and Linkage Map Construction ......................................................... 85 QTL Analysis ........................................................................................................ 86 RNA-Seq Analysis ................................................................................................ 90 Discussion ......................................................................................................................... 92 Conclusion ........................................................................................................................ 96 REFERENCES CITED ........................................................................................................... 97 v LIST OF TABLES Table Page Table 1.1: Mean A. fatua leaf areas consumed (in mm²) in greenhouse S. exigua choice experiments. Two-way comparisons were between line S2 and one of six MHR lines within the same pot. ............................................................................. 25 Table 1.2: Mean A. fatua leaf areas consumed (in mm²) in field S. exigua choice experiments. Two-way comparisons were between line S1 and one of four MHR lines in the same pot. . ........................................................................................... 27 Table 1.3: Mean A. fatua leaf areas consumed (in mm²) in field S. exigua choice experiments. Two-way comparisons were between line S2 and one of four MHR lines in the same pot. ............................................................................................ 27 Table 1.4: Mean A. fatua leaf areas consumed (in mm²) in field S. exigua no- choice experiments. Multi-way comparisons were between six A. fatua lines. ..................... 27 Table 1.5: Mean A. fatua leaf areas consumed (in mm²) in field grasshopper choice experiments. Two-way comparisons were between line S1 by 20R1 and 20R2, and S2 by 20R1 and 20R2 in the same pot. ................................................................. 28 Table 1.6: Mean A. fatua leaf areas consumed (in mm²) in field grasshopper no-choice experiments. Multi-way comparisons were between four A. fatua lines.. ....................................................................................................................................... 28 Table 2.1: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Linalool Across Herbicide Treatments in Open Bags with P-Values .......................................................................................................................... 43 Table 2.2: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Limonene Across Herbicide Treatments in Open Bags with P-Values .......................................................................................................................... 44 Table 2.3: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Linalool Across Herbicide Treatments in Closed Bags with P-Values .......................................................................................................................... 46 Table 2.4: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Limonene Across Herbicide Treatments in Closed Bags with P-Values .......................................................................................................................... 47 vi LIST OF TABLES CONTINUED Table 2.5: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for R4 Lines with Exogenous Linalool Across Herbicide Treatments in Closed Bags with P-Values .......................................................................................................................... 48 Table 2.6: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for R4 Lines with Exogenous Limonene Across Herbicide Treatments in Closed Bags with P-Values .......................................................................................................................... 49 Table 2.7: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Combined VOCs Across Herbicide Treatments in Closed Bags with P-Values ..................................................................................................... 50 Table 2.8: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for R4 Lines with Exogenous Combined VOCs Across Herbicide Treatments in Closed Bags with P-Values ..................................................................................................... 51 Table 3.1: Fluorescence Measurements Used for Photosynthetic and Energy Management Parameters of All A. fatua Lines Tested. ......................................................... 70 Table 4.1. Complete list of QTLs found in the 93 RIL population using MIM. ..................... 87 Table 4.2. Complete list of QTLs found in the 06 RIL population using MIM. ..................... 89 Table 4.2 Continued ................................................................................................................ 90 Table 4.3. DEGs underlying QTL peaks in R4 lines............................................................... 91 vii LIST OF FIGURES Figure Page Figure 1.2: Bar graph comparing the mean area consumed by S. exigua (in mm2) between A. fatua HS line S1 and MHR lines within the same pot. a) S1 v 18R, b) S1 v 20R1, c) S1 v 20R2, d) S1 v 93R, e) S1 v R3, f) S1 v R4 .............................. 25 Figure 1.3: Bar graph comparing the mean area consumed by S. exigua (in mm2) between A. fatua lines. ................................................................................................ 26 Figure 1.4 - Bar graph illustrates the olfactory choices of S. exigua regarding A. fatua lines R4 and S1. Each bar represents the count of olfactory choices made by S. exigua. .................................................................................................................. 29 Figure 1.5 – Bar graph illustrates the olfactory choices of S. exigua regarding Lactuca sativa L. and filtered airflow. Each bar represents the count of olfactory choices made by S. exigua. ...................................................................................... 30 Figure 1.6 – Bar graph illustrates the olfactory choices of S. exigua regarding filtered airflow and filtered airflow. Each bar represents the count of olfactory choices made by S. exigua. ..................................................................................................... 31 Figure 3.1 – Thin layer chromatography of acetone extracts of S2 (left panel) and R4 (right panel) plants. ..................................................................................................... 65 Figure 3.2 - Thin layer chromatography of acetone extracts of two R4 plants grown in different conditions.. ................................................................................................ 65 Figure 3.3 – Spectrophotometer Full scan of MHR and HS A. fatua lines ranging from 400- 750 nm. ..................................................................................................... 66 Figure 3.4 – Spectrophotometer, Acetone soluble pigment concentrations of A. fatua lines detected at 480 nm. ............................................................................................... 67 Figure 3.5 – Spectrophotometer, Acetone soluble pigment concentrations of A. fatua lines detected at 652 nm. ............................................................................................... 68 Figure 3.6 – Field derived lines measuring a. Y(NO) – Non-regulated energy dissipation, b. Y(NPQ) - Regulated energy dissipation, c. Y(II) – Effective Photochemical yield, and d. ETR - Electron transfer rate. ..................................................... 71 viii LIST OF FIGURES CONTINUED Figure 3.7 – 93 RIL groups and progenitor lines measuring a. Y(NO) – Non- regulated energy dissipation, b. Y(NPQ) - Regulated energy dissipation, c. Y(II) – Effective Photochemical yield, and d. ETR - Electron transfer rate........................... 72 Figure 3.7 – 06 RIL groups and progenitor lines measuring a. Y(NO) – Non- regulated energy dissipation, b. Y(NPQ) - Regulated energy dissipation, c. Y(II) – Effective Photochemical yield, and d. ETR - Electron transfer rate........................... 73 Figure 4.1 – Linkage Groups for the 06 RIL Population. Black Bands Represent a SNP. ..................................................................................................................... 86 ix ABSTRACT Intense herbicide usage has led to the evolution of herbicide resistant weeds, which threaten food production and security. The multiple herbicide resistant (MHR) Avena fatua (wild oat) lines investigated here are resistant to all members of selective herbicide families available for A. fatua control. The research in this thesis is designed to help understand some of the ecological, biochemical, and genetic aspects of MHR. First, MHR lines with elevated volatile organic compound (VOC) levels and herbicide susceptible lines were used to compare the feeding behavior of Spodoptera exigua (beet armyworm), and potential role of VOCs to mitigate herbicide injury. Results for feeding behavior were mixed, possibly being influenced by environmental and genetic changes more than VOCs. Exposing VOCs to A. fatua lines found that linalool reduced flucarbazone injury of HS plants, while a combined VOC treatment generally increased herbicide injury. MHR responded differently than HS plants to some treatments, suggesting that MHR has fundamental VOC perception alterations. Other studies compared plant pigments and energy management capabilities and showed that MHR lines had higher beta-carotene and chlorophyll b concentrations, as well as enhanced photosynthetic and excess energy management capabilities in MHR lines. Finally, two populations of recombinant inbred lines (RILs) were phenotyped for herbicide resistance and used to discover several quantitative trait loci (QTL) associated with resistance. Overall, this work contributes to our understanding of MHR and will lay the groundwork for future studies. 1 CHAPTER ONE INTRODUCTION Multiple Herbicide Resistant Avena fatua Weeds are the primary biotic cause of crop yield loss, reducing yield by an annual worldwide average of 28% (Vila et al., 2021). Before the creation of synthetic herbicides, enormous efforts were required to control weed populations by manual weeding or tillage. The adoption of herbicides has since revolutionized weed control by substantially reducing costs and manual labor, as well as increasing the overall efficacy of weed control (Heap 2014). The economic cost of weed management in the United States is estimated to be $33 billion annually (Chauhan 2020). In most modern crop production systems, producers use selective herbicides, which do not injure crops while killing weeds. However, the very properties that facilitated their continuous use have also led to overreliance and overuse, imposing intense selection pressure on target weed species. As a result, herbicide-resistant populations of many weeds are now widespread across worldwide cropping systems. The first recorded observation of herbicide resistance was to 2,4-dichorophenoxyacetic (2,4-D), an auxin mimic herbicide that was discovered in 1950 (Reddy and Nandula 2012). Since 1950, approximately 521 unique cases of herbicide-resistant weeds (species x site of action) have been discovered, or a rate of 16 new cases reported annually (Heap 2014). Large amounts of herbicides are used each year. For reference, glyphosate, one of the most widely used herbicides, is applied at an annual rate of 804,000 tons (Rivas-Garcia et al., 2022). 2 Herbicide resistant weed populations can be resistant to one herbicide (HR) or multiple herbicides (MHR). Two types of biological mechanisms confer herbicide resistance: target site resistance (TSR) and non-target site resistance (NTSR). TSR, the most common mechanism, confers resistance due to a mutation in the gene encoding the herbicide target enzyme. This change inhibits the herbicide from binding to the target site or leads to overexpression of the target enzyme, allowing the enzyme to still function in the presence of herbicide (Powles and Yu 2010; Fang et al., 2019). NSTR includes other mechanisms that confer herbicide resistance outside of the target enzyme, such as a reduction in herbicide uptake or translocation, enhanced herbicide metabolism, decreased rate of herbicide activation and/or sequestration, or a point mutation that confers herbicide resistance (Délye et al., 2013, Dyer 2018, Jugulam and Shyam 2019). The focus of this literature review is Avena fatua L. (wild oat). As an annual, allohexaploid, and predominantly self-pollinating monocotyledonous weed, A. fatua thrives across temperate agricultural regions globally (Dahiya et al., 2019). Its adaptability extends to cereal systems, root crops, legumes, vegetables, and even ornamental plants, demonstrating its competitive nature (Beckie et al., 2012). Ranked as the second worst grass weed in the world (Beckie et al., 2012), A. fatua presents severe challenges for agricultural management. It exhibits various physiological traits that enhance its competitive advantage, including extended seed dormancy and early seed shattering and dispersal before harvest (Mahajan and Chauhan 2021). According to estimates, A. fatua has infested over 11 million hectares of agricultural land in the U.S. Great Plains and the Pacific Northwest and have caused an annual crop loss valued at approximately $1 billion (Beckie et al., 2012). 3 Avena fatua’s persistence surpasses borders, affecting cereal crops in Europe, North America, and Australia (Wrzesińska et al., 2016; Jäck et al., 2017; Dyer 2018; Mahajan and Chauhan 2021; Ņečajeva et al., 2021). Herbicide resistance biotypes of A. fatua is present among all these locations. For example, random survey in the Western Australian grain belt looked at 677 cropping fields and found 107 A. fatua populations that had range of resistant capabilities (Owen et al., 2009). In North America, and several European countries, A. fatua populations are being discovered to have multiple herbicide resistant capabilities (Travlos et al., 2011; Beckie et al., 2012; Adamczewski et al., 2019). Indicating that herbicide resistant A. fatua is becoming a greater threat to food production, and food security globally. Overall, research on Avena fatua resistant biotypes is divided into multiple sections; 1.finding the mechanisms conferring herbicide resistance like enhanced metabolism or target site mutations (Ahmad-Hamdani et al., 2013; Yu et al., 2013), 2. managing these populations to avoid more herbicide resistance emergence (Nietschke et al., 1996; Cavan et al., 2001), and 3. exploring the differences in resistant of A. fatua as compared to other resistant weeds, or a sensitive biotype (De Prado and Franco, 2004; Alizade et al., 2020). This thesis is investigating the later and the final chapter relates to finding mechanism conferring herbicide resistance. In Teton County, Montana, USA, producers controlled A. fatua populations exclusively with the preemergence herbicide triallate (HRAC Group 15) for over three decades after its commercialization in 1962. Herbicide Resistance Action Committee (HRAC) grouping is a classification system used to categorize herbicides based on their mode of action (Beffa et al., 2019). Producer complaints about triallate lack of control of A. fatua in 1993 led the Dyer lab to collect seeds and test them for resistance in the controlled condition of a greenhouse. These 4 populations were confirmed to be triallate resistant and were subjected to recurrent selection with triallate for five generations to create the inbred line 93R (Kern et al., 2002). This line was also found to be cross-resistant to the unrelated herbicide, difenzoquat (HRAC Group Ø), which has an unknown mode of action. An additional sensitive line HS2 was derived from a nondormant biotype used for seed dormancy research (Naylor and Jana 1976, Johnson et al., 1995). In 2006, after a complaint from producers that pinoxaden (HRAC Group 1) failed to control A. fatua during its first year of commercial use, seeds collected from fields near the 93R source were subjected to recurrent selection using pinoxaden for two generations to create the R3 and R4 inbred lines (Keith et al., 2015). They were confirmed to be resistant to members of the acetyl- CoA carboxylase (HRAC Group 1), acetolactate synthase (HRAC Group 2), and photosystem I (HRAC Group 22) inhibitors, as well as triallate and difenzoquat (Lehnhoff et al., 2013). An herbicide susceptible A. fatua line (HS1) was created from seeds collected nearby and was later confirmed to be 100% HS (Lehnhoff et al., 2013). Other resistant lines 18R, 20R2, and 20R1, were derived from seeds collected in 2018 (18R) and 2020 (20R1 and 20R2) from nearby fields and were suspected to be resistant to glyphosate as well as all R3 and R4 herbicides (Keith et al in preparation). In 2002, the Dyer lab made reciprocal crosses between 93R and HS1 to determine the genetic basis for triallate resistance. Triallate resistance segregated in a 15:1 HS to HR ratio in the F2 generation, indicating that two recessive genes controlled the resistance phenotype (Kern et al., 2002). No target site mutations known to confer resistance to Group 1 or Group 2 inhibitors were found in R3 and R4 A. fatua lines, indicating that NTSR mechanisms were responsible for the MHR phenotype (Keith et al., 2015). Additionally, RNA - Seq analysis 5 showed that R3 and R4 lines have constitutively elevated gene transcript levels relative to HS1, including genes with functions in xenobiotic catabolism, stress response, redox maintenance, and transcriptional regulation (Keith et al., 2017). R3 and R4 resistance to flucarbazone, imazmethabenz, and pinoxaden is controlled by three closely linked nuclear genes (Burns et al., 2018). Finally, the R3 and R4 lines' response to herbicide stress is similar to other abiotic stressors, as determined by proteomic biochemical and immunological assays (Burns et al., 2018). The findings from these studies reveal that the resistant lines possess NTSR mechanisms to multiple herbicides, making them a distinctive model for research. Volatile Organic Compounds Plants are stationary organisms: they have evolved numerous defense mechanisms to combat abiotic and biotic stresses. One response to stress is the release of biogenic volatile organic compounds (VOCs) that not only function to attract pollinators and in plant–plant communication, but also induce plant defense mechanisms against insects and pathogens (Ueda et al., 2012; Kalske et al., 2019). When a plant perceives a stress, it emits a unique VOC profile which can contain a mix of hundreds of VOCs, and these profiles differ among plant species (Landi et al., 2020; Tiwari et al., 2020; Fitzky et al., 2021). Plant VOCs include members of the very large terpene family, including hemiterpenes (C5), monoterpenes (C10), sesquiterpenes (C15), homoterpenes (C11 and C16), diterpenes (C20), and triterpenes (C30) (Vivaldo et al., 2017, Zhou and Jander 2021). Mechanically injured plants release a mix of green leaf volatiles (GLVs) consisting of aldehydes, esters, and alcohols of six-carbon compounds (Ameye et al., 2018). 6 Notably, all plants constitutively emit VOCs; however, when they encounter stress, they release different VOC profiles dependent on response. Emitted VOCs from a stressed plant can be detected by surrounding receiver plants alerting them to prepare for incipient stresses by priming defensive mechanisms, including enhanced production of defensive chemicals, induction of defensive proteins and enzymes, and activation or reinforcement of physical barriers (Llusia et al., 2005; Brilli et al., 2009; Yuan et al., 2009; Kalske et al., 2019; Sako et al., 2020; Zhou and Jander 2021). For example, drought stress causes a plant to warn other plants around it and induce a defensive mode to counteract the stress, like closing the stomata to reserve water (Copolovici et al., 2014). In contrast, plants experiencing a strong temperature change emit a different profile of VOCs that elicit different physiological responses (Holopainen and Gershenzon 2010). The previously described Dyer lab MHR A. fatua line (R4) was found to constitutively emit a different VOC profile than the HS1 line (Keith et al., in preparation). The MHR and HS lines were placed individually in glass collection chambers, and VOCs were collected on a super- Q trap (Alltech Associates Inc, Deerfield, IL) for 6-hour periods, eluted, and quantified using gas chromatography/mass spectrometry. Of more than 75 VOCs detected, nine were emitted at significantly different levels between untreated HS and MHR plants. Of these nine, five were identified as monoterpenoids, three as GLVs, and one as methyl salicylate. The A. fatua monoterpenoids emitted at significantly higher levels in MHR plants were β-myrcene, limonene, linalool, α-terpinene, and α-pinene. GLVs identified as decanal was also emitted at higher levels in MHR lines. 7 Of the nine VOCs listed above, linalool, limonene, and methyl salicylate were chosen for further study. In addition, methyl jasmonate was included because of its well-documented role in defense signaling (Tamogami et al., 2008). The jasmonate signaling pathway regulates the plant response to abiotic stressors like drought, salt, and cold (Wang et al., 2021). This pathway also regulates responses to biotic stress imposed by insect herbivores, pathogens, or other fungal diseases. (Turner et al., 2002). Transcriptome studies show that jasmonate signaling is also involved in changes in defense gene expression in response to mechanical and herbivory plant wounding, host plant resistance to phloem-feeding insects, enhanced chemical and VOC based defenses, priming direct and indirect defenses, pathogen resistance and systemic transmission of defense signaling (Howe and Jander 2008). The monoterpenoid linalool enhances direct and indirect defenses mostly against biotic stressors like insects, pathogens, and fungal infections (Zhang et al., 2023). Exposure to linalool can also induce similar responses. For example, exposure to exogenous linalool has been shown to activate oxidative bursts and inhibit infection by Colletotichum acutatum, a pathogen which causes citrus post-bloom fruit drop in Citrus × aurantium (Sweet Orange) (Sun et al., 2014). Similarly, exogenous exposure to limonene has been shown to activate disease resistances to multiple organisms and confer limited thermotolerance in Quercus ilex (Holm Oak) (Llusia et al., 2005; Fujioka et al., 2015). Therefore, the inherent properties of linalool and limonene protect plants from biotic stressors and likely have broader regulatory roles in plant stress response. Methyl salicylate, another methyl ester compound, has major signaling roles in systemic acquired resistance (SAR), an induced defense response that confers long-lasting protection against a broad spectrum of microorganisms (Park et al., 2007). Methyl salicylate is primarily 8 associated with biotic stressors and is released by plants undergoing stress such as pathogen attack. (Chen et al., 2019). In addition, methyl salicylate itself has direct antimicrobial properties and can act as a local defense compound within the emitting plant as Aphidoidea (aphid) and other herbivores have been documented to be deterred by methyl salicylate (Verma et al., 2023). In receiver plants, methyl salicylate can be demethylated to salicylic acid, which further amplifies the defense signaling network by upregulating the expression of pathogenesis-related genes and priming the plant's defenses (Chen et al., 2019). Changes in VOC emission profiles have been investigated after herbicide treatment of susceptible plants and in transgenic herbicide-tolerant plants. Exposure to sublethal glyphosate doses altered VOC/GLV emission patterns but did not alter parasitoid behavior on herbicide susceptible Zea mays (maize) plants (D’Alessandro et al., 2006). Transgenic glyphosate-resistant Glycine max (soybean) plants with an insensitive 5-enolpyruvoylshikimate-3-phosphate synthase did not show constitutive VOC changes. However, emissions of several monoterpenoids were preferentially elevated in response to herbivory by Chrysodeixis includens (soybean looper) in the resistant line, as compared to their glyphosate sensitive counterparts (Strapasson et al., 2016). In Oryza sativa (rice), no VOC differences were detected between transgenic glufosinate- resistant and non-transgenic cultivars (Choi and Kim 2013). Besides these studies on transgenic plants, I am not aware of any VOC analyses of field-selected, herbicide resistant populations of any plant species. Plant Pigments and Non-Photochemical Quenching Chapter Four of this thesis examines plant pigment concentrations and non- photochemical quenching (NPQ) capacity of MHR and HS A. fatua lines. Quantifying plant 9 pigments and NPQ can give insight into linkages with abiotic stress defense mechanisms and possibly with NTSR mechanisms. As mentioned above, NTSR mechanisms are mechanisms that confer herbicide resistance by reducing herbicide uptake or translocation, enhancing rates of herbicide metabolism, decreasing rates of herbicide activation and/or sequestration, possibly by a point mutation in a gene other than one encoding the herbicide’s target enzyme, or other unidentified mechanisms (Délye et al., 2013; Dyer 2018; Jugulam and Shyam 2019). Since many herbicides cause increased reactive oxygen species (ROS) content (Traxler et al., 2023), an enhanced ability to quench ROS could also confer NTSR. Herbicides function by specifically binding to target enzymes that disrupt essential metabolic pathways within plants. This interaction leads to a cascade of events that ultimately leads to the plant’s mortality. Some herbicides like HRAC groups 5, 6, 10, 12, 13, 14, 22, 23, 32, 33, 34, and others bind enzymes in plant photosystems, (Beffa et al., 2019) and generally, shut down electron transport in the photosystem. As a result, ROS generation can increase rapidly and dramatically (Matzrafi et al., 2017; Caverzan et al., 2019; Traxler et al., 2023). ROS molecules are integral in plant signaling, growth, development, and response to biotic and abiotic factors (Tripathy and Oelmüller 2012, Das and Roychoudhury 2014). There are four primary types of ROS: free radicals O2•− (superoxide radical), •OH (hydroxyl radical), non- radicals H2O2 (hydrogen peroxide), and 1O2 (singlet oxygen). While these ROS molecules are beneficial in small amounts, large amounts can damage DNA, lipids, and proteins and ultimately cause apoptosis (Das and Roychoudhury 2014). Plants ameliorate excessive ROS levels using enzymatic and non-enzymatic mechanisms, including plant pigments, which play a significant role in ROS homeostasis (Choudhary et al., 2017). 10 Plants contain molecules called biochromes, or more commonly, pigments, that absorb or reflect light. Many biological processes depend on visible light, using wavelengths ranging from ultraviolet light (<380 nm) to far-red light (>700 nm). Plant pigments absorb energy specific wavelengths to power and influence biological processes (Carvalho et al., 2011), like photosynthesis, germination, fruiting, growth direction, and pigment formation. Light quality, quantity, duration, direction, and intensity are all involved in photomorphogenesis, the growth and development of plants in response to light (Arsovski et al., 2012). Photomorphogenesis is influenced by protein photoreceptors that interpret incoming light and activate processes to regulate signaling pathways (Paik and Huq 2019). A critical part of photomorphogenesis is regulating plant pigment levels in response to environmental variables. For example, abiotic stress can cause an accumulation of carotenoids and anthocyanins as part of an overall defense response (Carvalho et al., 2011; Demmig-Adams et al., 2020; Stewart et al., 2021). Plant pigments are classified into four categories: flavonoids, carotenoids, chlorophyll, and betalains, all of which have roles in photosynthesis, light harvesting, and photoprotection (Chen 2015). Since betalains are only found in the order Caryophyllales, my research focused on the other three main plant pigment classes. Flavonoids are essential secondary metabolites that play critical roles in plant development and defense. They are classified into 12 subclasses based on the oxidation degree of their central heterocycle (Shen et al., 2022). These pigments reflect wavelengths that confer purple, blue, and red pigmentation in fruits, vegetables, and flowers (Carvalho et al., 2011). Anthocyanidins, an unstable subclass of flavonoids, generally exist in their glycosylated form as anthocyanins. This large subclass contains over 650 compounds which act as free radical 11 acceptors that can quench and neutralize ROS before they cause damage to plant cells (Han et al., 2012; Shen et al., 2022). The second class of plant pigments consisting of carotenoids, lipophilic secondary metabolites crucial for photoprotection and light harvesting, are grouped into two sub-classes: carotenes and xanthophylls. Carotenes are non-oxygenated molecules, while xanthophylls contain oxygen (Amorim-Carrilho et al., 2014), and both are responsible for the orange, yellow, and red colors in fruits, vegetables, and flowers (Carvalho et al., 2011). Carotenes and xanthophyll both play photoprotective roles by dissipating photosynthetic excitation energy as heat scavenges ROS and suppresses lipid peroxidation (Tiwari et al., 2017). Additionally, carotenes and xanthophylls are antenna molecules capable of harvesting light and transmitting photon energy to the photosystems. Importantly, they significantly detoxify ROS (Han et al., 2012; Maoka 2020). Carotenes exist in either a ground state or in one of two excited states after photon absorption. In its lower excited state, transferred energy is used in photosynthesis, while the higher energy state (triplet–triplet) releases transferred energy as heat through polyene vibration. However, the triplet–triplet state can also potentially create ROS, particularly singlet oxygen. Excess ROS is primarily quenched by antioxidants, but more stressful conditions can cause cellular damage, such as lipid peroxidation and protein oxidation, ultimately leading to DNA damage (Choudhary et al., 2017; Hashimoto et al., 2018; Maoka 2020). Xanthophylls are also involved in light harvesting and photoprotection. They dissipate excess energy by a cyclic enzymatic process that removes epoxy groups, forming de-epoxidized xanthophylls. These de-epoxidized xanthophylls, particularly zeaxanthin, play a vital role in 12 enhancing the structural integrity of chloroplast membranes, enhancing their resistance to oxidative damage by ROS. This cycle promotes the dissipation of excess energy, effectively preventing it from overexciting the photosynthetic reaction centers. During periods of intense light exposure, violaxanthin, which contains epoxy groups, is enzymatically converted into zeaxanthin, the de-epoxidized form, passing through the intermediary antheraxanthin. This safely releases energy before interacting with oxygen molecules (Maoka 2020). Genetic and transgenic efforts to alter xanthophyll biosynthesis and overexpress beta- carotene hydroxylase genes have improved the abiotic stress tolerance of Arabidopsis thaliana, Nicotiana tabacum, and Oryza sativa. (Davison et al., 2002; Götz et al., 2002; Römer and Fraser 2005; Du et al., 2010). Beta-carotene hydroxylase genes are responsible for the biosynthesis of zeaxanthin and carotenoid precursors, which can confer drought and oxidative stress resistance (Du et al., 2010). To my knowledge, enhanced herbicide tolerance based on elevated non- enzymatic antioxidants has not been documented. However, the detection of plant pigment concentrations in MHR and HS Avena fatua lines warrants further analysis since it is a potential NTSR mechanism. The third plant pigment class consists of chlorophylls, chlorophyll a and b, which are located in chloroplast light-harvesting complexes and capture solar energy used to drive photosynthesis and other biological processes (Murchie and Lawson 2013). If light energy is not used for photochemistry, then it is emitted as heat or light that could cause ROS generation. These outcomes do not exist in isolation but rather in competition, and are dependent upon light intensity, light quality, and genetics (Genesio et al., 2021). If light intensity exceeds the photosynthetic maximum, and if the generated excess energy is not quenching by the plant, it 13 could cause ROS generation and induces photooxidative damage (Roach and Krieger 2014). Antioxidant mechanisms can be activated to dissipate energy from excess light; however, photoreceptors such as far-red, red, blue, and UV are primarily utilized first. As far-red photoreceptors, indicated by the name are influenced by far-red light (Batschauer 1998). Plant photoreceptors can initiate light avoidance by causing chloroplasts to relocate inside plant cell (Carvalho et al., 2011; Tripathy and Oelmüller 2012). My research focused on photochemical yield and excess light quenching through regulated and non-regulated mechanisms, factors that are linked to abiotic stress tolerance (Chaerle et al., 2007; Hussain et al., 2023), and perhaps to NTSR to herbicides. Since ROS generation commonly occurs with herbicide injury, any mechanism that has an enhanced role to dissipate, quench ROS, would likely enhance NTSR herbicide resistance. Quantitative Trait Loci Analysis of MHR Avena fatua The issue of identifying herbicide resistant mechanisms is a complex and troublesome problem worldwide. As the global reliance on herbicides continues, so does the selection pressure, leading to a rapid increase in cases of herbicide resistance. Improved genetic analysis of NTSR mechanisms in A. fatua became more accessible with the availability of a reference genome for the closely related species Avena sativa (common oat) in 2021 (Yao et al., 2022). Avena sativa and Avena fatua are in the same genus; both are hexaploid (42 chromosomes), can be crossed without barriers, and a high degree of synteny is expected (Koroluk et al., 2022). Since previous studies in the Dyer lab have shown that several genes are potentially responsible for the MHR phenotype (Burns et al., 2018; Keith et al., 2017; Keith et al., 2015), genomic 14 approaches using quantitative statistical methodology is warranted to search for casual genes (Leon et al., 2021). QTL analysis is a statistical method that links phenotypic and genotypic data to specific regions of chromosomes. This tool attempts to explain the genetic basis of variation in complex traits (Farquhar, 1998). High quality phenotypic data results in higher predictive power to detect potential QTL markers (Würschum, 2012). Genotypic data requirements are stringent and often necessitate the use of recombinant inbred lines (RILs) for accurate analysis. This involves a cross between two parental lines, typically one possessing the trait of interest and the other lacking it. The progeny from this initial cross is then subjected to sibling mating or selfing for multiple generations to achieve isogeneity. The resulting RIL thus allows precise genetic mapping at specific loci (Broman, 2005). The Dyer lab has three RIL populations that were utilized for this research. The 93R inbred line was crossed with the HS S2 line to create a population of 124 RILs (Kern, Myers et al. 2002) and the inbred MHR line R4, which were also crossed with S2 to create the 06 (n = 120) RIL populations, respectively (Burns, Keith et al. 2018). To determine genomic differences between resistant and sensitive A. fatua lines that may contribute to herbicide resistance, genotyping-by-sequencing (GBS) was performed on all parental and RILs populations. GBS fragments the genome into small sequences and is aligned to a reference genome to identify single nucleotide polymorphisms (SNPs) (Deschamps et al., 2012). SNPs represent parental variations at individual nucleotides in the DNA sequence. To create linkage maps, SNPs from the genomic data are placed on a map according to their position on the chromosome and how often they are inherited together. The distance between adjacent SNPs on the map is measured in centimorgans (cM), units representing the distance between 15 genetic markers on a chromosome (Oliver et al., 2001). One cM represents a 1% chance that a marker at one genetic locus will be separated from a marker at a second locus due to recombination during meiosis. The closer the SNPs are to each other on a chromosome, the more likely they will be passed on together to the next generation (Ganal et al., 2009). SNPs are valuable for this purpose due to their abundance throughout the genome and the relative ease with which their frequency and location can be determined in the mapped individuals (Sudan, Sharma et al. 2023). While some SNPs can be directly associated with phenotypic differences, their primary utility in this context is to serve as markers for genetic mapping (Myles & Wayne, 2008). QTL analyses use the linkage map created from the SNPs and phenotypic data to calculate QTL positions using a statistical estimate known as the logarithm of the odds ratio (LOD). LOD scores thus estimate the relative strength of evidence for the presence of a QTL at a particular location (Van Ooijen, 1999). QTL analysis measures intervals between adjacent pairs of markers along each chromosome and gives a corresponding LOD score. Using these methods, one can find accurate and reliable QTL markers associated with potential candidate genes that confer NTSR mechanisms in A. fauta. Research Objectives 1) Evaluate the olfactory and herbivory preferences of a generalist insect species between MHR and HS lines of A. fatua under choice and no-choice conditions. 2) Examine how exposure to exogenous VOC applications affects the physiological responses of HS and MHR A. fatua lines to subsequent herbicide treatments. 16 3) Characterize and compare plant pigment concentrations and non-photochemical quenching capacities of MHR and HS A. fatua lines to determine if they are associated with herbicide resistance phenotypes. 4) Perform quantitative trait loci analysis using recombinant inbred lines of three A. fatua populations to identify genomic regions potentially associated with MHR phenotypes. Project Significance This project aims to understand the mechanisms of MHR in Avena fatua using physiological, biochemical, and molecular genetic tools. Insect herbivory studies will help to understand the associations between elevated VOC emission profiles and insect orientation and feeding choices and may have ecological implications for MHR plants in agroecosystems. Examining the effects of VOC exposure on subsequent herbicide efficacy will add to our understanding of the maintenance and evolution of MHR as affected by biochemical plant signaling. Comparative analysis of plant pigments and non-photochemical quenching capacities will add to our knowledge of previously undescribed NTSR mechanisms related to photooxidative stress tolerance. Discovery and analysis of quantitative trait loci for MHR will reveal specific genetic regions associated with this trait and represents a novel molecular genetic approach in weedy species to unravel the underlying mechanisms that confer resistance, potentially aiding the development of more effective, targeted control strategies. 17 CHAPTER TWO INSECT GENERALIST PREFERENCE TO MHR AVENA FATUA Summary This study aimed to explore the feeding preferences of the generalist herbivore, Spodoptera exigua (beet armyworm), between multiple herbicide resistant (MHR) and herbicide sensitive (HS) lines of Avena fatua (Wild oat) in both greenhouse and field settings. Beet armyworm Spodoptera exigua preferred the R4 (MHR) line over the S1 (HS) line and preferred S1 over 93R (MHR) in a controlled greenhouse environment. However, no consistent preferences were observed in other choice and no-choice trials or in field experiments. Olfactory tests produced inconclusive results between the R4 and S1 lines. This study suggests that environmental factors and genetic variation influence herbivory more than the released VOCs for the Avena fatua lines used in this study. The inconsistent findings across experiments call for further investigation into the intricate factors affecting herbivore preferences, including but not limited to VOC profiles and genetic variations in A. fatua lines. This study provides a foundation for understanding the potential ecological implications of herbicide resistance on herbivore behavior, which may have broader impacts on both weed and crop management strategies. 18 Introduction Plant volatile organic compounds (VOCs) and green leaf volatiles (GLVs) are primary determinants of insect attraction, orientation, and feeding preferences (Ameye et al., 2018). Most published research on attracting insects to plants focuses on indirect defenses utilizing VOCs to attract beneficial insects. This attraction of predatory/parasitoid species that colonize the insect herbivores is a form of biological control (Baldwin et al., 2002; Kessler & Baldwin, 2001; Rodriguez-Saona et al., 2011). However, for weedy or undesirable plants, VOCs could also play a role in attracting herbivores to aid in pest management. VOC sensing is detected by an organism’s capacity of olfaction, the sense of smell, to detect and extract information from said VOCs. What sets insects apart is their heavy reliance on olfaction to ensure survival and reproduction. For insects, this primarily involves sensing food, hosts, mates, and prey and performing group communication for aggregation and avoidance (Ache & Young, 2005; Fan et al., 2011) Without damage, plants constantly emit VOCs that influence their ecology. VOC differences between the MHR and HS A. fatua lines may influence the feeding preferences of an insect generalist herbivore. MHR and HS A. fatua line’s (VOC) emissions differ. Of the 75 unique VOCs detected from the A. fatua lines, nine were produced significantly more in the MHR R4 line. When the HS S2 plant was mechanically wounded, its emission profile was similar to the untreated MHR R4 line. This, along with several other biochemical, genetic, and proteomic changes, indicates that the MHR A. fatua lines being examined are constitutively primed to resist stressors (Burns et al., 2018; Dyer, 2018; Keith et al., 2017). It begs the question 19 of whether these VOC profiles also act as a subtle deterrent or attractant, possibly influencing the feeding preferences of an insect generalist herbivore. VOC attraction of insect herbivores has been studied for many plant and insect species (Anderson & Alborn, 1999; Rojas, 1999; Sun et al., 2014). For specific VOCs, limonene, and methyl salicylate have been shown to attract Helicoverpa armigera (cotton bollworm) larvae in petri dish and two-choice olfactometer assays (Di et al., 2017; Gregg et al., 2010). The constitutively elevated emissions of limonene and methyl salicylate in MHR R4 A. fatua led me to hypothesize that they might similarly attract S. exigua and consume more of the MHR lines. Research in many species has illuminated the molecular details of plant deterrence of insect pests, including the defensive mechanisms that the plant incorporates when being fed upon. Several defenses are typically induced following an herbivore attack. Herbivory damage increases cytosolic calcium levels, increases protein phosphorylation after activation of mitogen- activated protein kinases (MAPK), NADPH oxidase is activated, and the generation of reactive oxygen species (Mostafa et al., 2022). Additionally, the plant releases VOCs to attract beneficial insects or to repel the insects that are consuming it (Howe & Jander, 2008). Avena fatua is fed upon by many insects including Spodoptera exigua (beet armyworm) and Acridomorpha (grasshopper species). Spodoptera exigua (beet armyworm), a polyphagous insect is a significant threat to agriculture and one of the most destructive insects in the Spodoptera genus (Hafeez et al., 2021). Beet armyworm larvae feed on economically important crops like cotton, vegetables, and flowers (Hafeez et al., 2021; Moulton et al., 2000). Since this species is not native to Montana, I also tested the Montana-endemic Acridomorpha (grasshopper species) for a relevant comparison using a natural herbivore in field conditions. Grasshoppers are 20 monophagous to grass species but can be polyphagous depending on the scarcity of sustenance, making wild oat an applicable experimental plant species (Franzke et al., 2010; Joern, 1979). Methods Plant Material The Avena fatua MHR lines utilized in these experiments consisted of 93R, R4, R3, 18R, 20R2, 20R1, and the S1 and S2 sensitive lines. These plants were cultivated under controlled conditions in a greenhouse, following a 16-hour photoperiod, which included natural sunlight supplemented with mercury vapor lamps delivering an intensity of 165 μmol m−2 sec−1. The temperature was maintained at 25 ± 4°C. They were planted in a standard greenhouse soil mix comprising Bozeman silt, loam, and Sunshine mix #12 (Sun Gro Horticulture, Inc., Bellvue, WA) in a 1:1:1 ratio enriched with perlite. Weekly fertilization was carried out using Jack’s water-soluble 20 N–20 P–20 K fertilizer (JR Peters Inc., Allentown, PA). All plants for each experiment were uniformly located on the same greenhouse bench. Spodoptera exigua Choice and No-choice Herbivory Experiments Greenhouse Experiments Fourth instar larvae of Spodoptera exigua (beet armyworm) were purchased from Benzon Research Inc. (Carlisle, PA). On arrival, the larvae were maintained at 4°C and provided with an artificial diet supplied by Benzon Research Inc. Before the start of each experiment, the larvae were subjected to a 24-hour starvation period at room temperature. Whole plant choice and no-choice herbivory experiments were conducted using A. fatua plants at Zadok's stage 13 under the greenhouse conditions described above. For choice 21 experiments two plants were transplanted into a 20.32 cm diameter pot and for no choice one plant was transplanted into a 20.32 cm diameter pot 7 days before experiment. Each trial involved placing a single S. exigua larva on the soil surface, positioned 4 cm from the base of either a single MHR or HS plant (no-choice) or equidistantly between MHR and HS plants growing together in the same pot (choice), in 20.32 cm diameter pots. To prevent larvae escape, pots were covered with 70 cm x 30 cm wide, 300 µ mesh nylon insect rearing sleeves (MegaView Science Co., Ltd. Taiwan). The larvae were allowed to feed for 6 hours, after which they were removed, and all plant shoots were harvested at the soil surface. Individual leaves were separated, taped to graph paper, labeled, and photographed for quantification of herbivory. The extent of leaf area consumed was determined using the open- source Java image processing and analysis program ImageJ (Schneider et al., 2012). Field Experiments Whole plant choice and no-choice herbivory experiments were also conducted under field conditions using A. fatua plants and S. exigua. Plants were grown at the Arthur H. Post Research Farm near Bozeman, MT, located at latitude 35.67 N, longitude 111.15 W, and an elevation of 1,455 m. The soil type at this location is an Amsterdam silt loam. Each plant was grown in a buried 20.32 cm diameter pot with 3.81 cm extending above the soil surface to allow insect-rearing sleeves to be fastened around the pot base to prevent insect escape. Spodoptera exigua larvae used in field trials were stored, starved, and allowed to feed as described above, and leaf areas consumed were recorded as above. Acridomorpha (grasshopper species) were gathered directly at the Arthur H. Post Farm site. Grasshopper collection involved using a bug net to sweep native grasses in the vicinity of field plots. The captured grasshoppers were subsequently placed in individual 50 mL tubes, each fitted with an airhole on the cap. This 22 collection took place one day prior to the experiment, ensuring a 24-hour period of starvation similar to what S. exigua experienced. All experiments were initiated at dawn or early morning. This early morning timing of experiment initiated was determined upon environmental factors, mainly temperature, as S. exigua was observed to seek heat avoidance with high temperatures. Measurements In the analysis of herbivory results, two distinct metrics were employed: mean leaf area consumed by S. exigua larvae, expressed in square millimeters (mm²), and mean average leaf consumption, expressed as a percentage of the total plant leaf area. Data presented will employ the leaf area consumed in (mm2). S1 or S2 leaf consumption was compared with each MHR line in pairwise comparison. Spodoptera exigua Olfaction Choice Assay Olfaction preference experiments were performed by monitoring S. exigua orientation to individual air flows from two different, intact plants growing in soil (Fig. 1.1). Avena fatua lines R4 and S1 and Lactuca sativa L. (as an unrelated, broadleaf control species) were grown separately in 10.16 cm diameter pots enclosed in scentless, clear, closed polyethylene bags (12.70 x 10.16 x 60.96 cm). Air flow was controlled at 6 liters min-1 from a commercial grade air pump (EcoPlus ECOair 1) attached to 3/16” clear vinyl tubes (Eastrans Inc. Bayboro, NC). Air streams were filtered through C290 activated charcoal filters (Etsy, Fort Jones, CA), and then passed through LZQ-7 gas flow meters (Hilitand, China), and into the polyethylene bags via slits at the bottom. Exhaust tubes from plant bags were secured into slits at the top of each bag, leading to a cardboard platform with 2.54 cm diameter holes spaced 15.24 cm apart and covered with stainless steel mesh screen (1 mm), so that air from two bagged plants was passing through the two holes. At the initiation of each experiment, one S. exigua larva was placed on the 23 platform between the two mesh holes, and the number of orientations towards one or the other air hole, or no response, was recorded during a one-minute trial. Direct comparisons were made between the R4 and S1 lines, filtered airflow and L. sativa, and a control study where no plants were assessed. Figure 1.1. Experimental set up for S. exigua olfaction choice assay. The “x” in the middle of the mesh screen represents where the S. exigua was placed for each repetition. Statistics Herbivory choice and no-choice data from the greenhouse were analyzed using linear mixed-effects models in R Studio (RStudio TEAM, 2023). Differences in total leaf area consumed between different lines were assessed. Differences in average leaf area consumed were also recorded, but the patterns were the same as total leaf area consumed so those results are not presented. Choice data was subset to only compare changes in leaf area consumed between directly compared lines. An interaction term between target and competing lines was included to make this assessment. The same analyses were performed on data collected from the field but an additional factor representing insect (S. exigua and Acridomorpha) was included. For the olfactory experiments we used linear mixed-effects models to assess the choices made by S. 24 exigua (choice A, choice B, or none). Variation within and between trials was accounted for with random intercept adjustments for all models. Analysis of variance tests (ANOVA) were performed to assess treatment group differences. Estimated marginal means (EMMs) with Tukey adjustments were used for post-hoc comparisons between treatment groups (Searle et al., 1980). All treatments were pseudo replicated 2 to 20 times within an experiment, and experiments were run 1 to 7 times. Data from multiple trials were combined after confirming homoscedasticity using diagnostic plots. If only one experiment was performed, a linear model was used with no mixed effects. Results Greenhouse Choice Experiments In the greenhouse S. exigua preference varied across genotypes. Choice experiments showed that R4 leaves were consumed more than S1, as expressed in areas consumed and average areas consumed (P ≤ 0.001 and P ≤ 0.05, respectively). Conversely, S1 was consumed more than 93R in mean area and mean average consumed (both with significance levels of P ≤ 0.05) (Figure 1.2). No significant differences were detected for other combinations of S1 and other MHR lines or in the S2 choice trials (Table 1.1). There were also no significant differences found between S2 and S1 choice trails (data not shown). 25 Figure 1.2: Mean leaf area consumed in mm2 by S. exigua between A. fatua HS line S1 and MHR lines within the choice experiments. No * indicates no significant difference. * Indicates P ≤ 0.05. *** indicates P ≤ 0.001. a) 18R v S1, b) 20R1 v S1, c) 20R2 v S1, d) 93R v S1, e) R3 v S1, f) R4 v S1 Table 1.1: Mean A. fatua leaf areas consumed (in mm²) in greenhouse S. exigua choice experiments. Two-way comparisons were between line S2 and one of six MHR lines within the same pot. ‘n’ replicates; p > 0.05 indicates no significant difference between treatments. Leaf Area Consumed (mm2) S2 vs MHR lines Greenhouse, Choice HS HS Area Consumed MHR MHR Area Consumed 2 2 n P-Value Line (mm ) Line (mm ) S2 19.95 (± 8.17) 18R 45.12 (± 13.68) 11 P > 0.05 S2 25.33 (± 8.04) 20R1 46.99 (± 9.88) 16 P > 0.05 S2 39.92 (± 15.94) 20R2 66.55 (± 23.81) 17 P > 0.05 S2 40.16 (± 12.51) 93R 15.71 (± 6.78) 16 P > 0.05 S2 43.67 (± 10.24) R3 52.7 (± 10.66) 13 P > 0.05 S2 61.36 (± 16.23) R4 66.61 (± 10.90) 16 P > 0.05 26 Greenhouse No-choice Experiments In no-choice greenhouse experiments, S. exigua demonstrated significantly higher consumption of R4 and S2 compared to 20R2 in terms of mean leaf areas consumed (mm2) (Figure 1.3). Figure 1.3: Mean leaf area consumed in mm2 comparing all A. fatua lines in the greenhouse in the no choice experiments. Compact letter display is employed to illustrate significant differences. Field Choice and No-choice Experiments Under field conditions, there were no significant differences in leaf area consumption among any plant line pairwise comparisons, for both S. exigua and grasshopper insect herbivores (Tables 1.2, 1.3, 1.4, 1.5, 1.6) 27 Table 1.2: Mean A. fatua leaf areas consumed (in mm²) in field S. exigua choice experiments. Two-way comparisons were between line S1 and one of four MHR lines in the same pot. ‘n’ replicates; p > 0.05 indicates no significant difference between treatments. Leaf Area Consumed (mm2) S1 vs MHR Lines Field, Choice HS Area Consumed MHR Area Consumed Line (mm2) Line (mm2) n P-Value S1 20.11 (± 7.11) 20R1 28.90 (± 8.21) 8 P > 0.05 S1 28.43 (± 7.78) 93R 10.28 (± 3.31) 6 P > 0.05 S1 11.74 (± 7.49) R3 36.23 (± 13.70) 6 P > 0.05 S1 14.38 (± 4.43) R4 25.43 (± 6.83) 10 P > 0.05 Table 1.3: Mean A. fatua leaf areas consumed (in mm²) in field S. exigua choice experiments. Two-way comparisons were between line S2 and one of four MHR lines in the same pot. ‘n’ replicates; p > 0.05 indicates no significant difference between treatments. Leaf Area Consumed (mm2) S2 vs MHR Lines Field, Choice HS Area Consumed MHR Area Consumed Line (mm2 n P-Value ) Line (mm2) S2 14.53 (± 6.41) 20R1 22.46 (± 5.70) 7 P > 0.05 S2 32.65 (± 11.90) 20R2 34.92 (± 12.79) 4 P > 0.05 S2 26.70 (± 9.31) 93R 30.84 (± 13.44) 8 P > 0.05 S2 30.07 (± 8.19) R4 49.08 (± 7.78) 9 P > 0.05 Table 1.4: Mean A. fatua leaf areas consumed (in mm²) in field S. exigua no-choice experiments. Multi-way comparisons were between six A. fatua lines. ‘n’ replicates; compact letter design employed to see statistical differences. Leaf Area Consumed (mm2) All Lines, Field, No-Choice A. fatua Line Area Consumed (mm2) n Compact Letter Design S2 41.42 (± 10.27) 12 a S1 42.11 (± 7.43) 14 a 93R 31.00 (± 7.87) 9 a R4 66.02 (± 16.53) 12 a 20R1 26.35 (± 6.06) 11 a 20R2 29.57 (± 11.18) 7 a 28 Table 1.5: Mean A. fatua leaf areas consumed (in mm²) in field grasshopper choice experiments. Two-way comparisons were between line S1 by 20R1 and 20R2, and S2 by 20R1 and 20R2 in the same pot. ‘n’ replicates; p > 0.05 indicates no significant difference between treatments. Leaf Area Consumed (mm2) HS vs MHR Lines, Field, Choice, Grasshopper HS Area Consumed MHR Area Consumed 2 2 n P-Value Line (mm ) Line (mm ) S1 51.27 (± 10.36) 20R1 18.39 (± 7.65) 2 P > 0.05 S1 92.69 (± 46.06) 20R2 55.63 (± 39.01) 3 P > 0.05 S2 100.56 (± 59.14) 20R1 37.29 (± 23.52) 4 P > 0.05 S2 22.62 (± 22.62) 20R2 37.29 (± 28.17) 2 P > 0.05 Table 1.6: Mean A. fatua leaf areas consumed (in mm²) in field grasshopper no-choice experiments. Multi-way comparisons were between four A. fatua lines. ‘n’ replicates; compact letter design employed to see statistical differences. Leaf Area Consumed (mm2) Four Lines, Field, No-Choice, Grasshopper A. fatua Line Area Consumed (mm2) n Compact Letter Design S2 72.16 ± (26.5) 7 a S1 102.09 ± (37.5) 9 a 93R 77.61 ± (54.3) 4 a 18R 31.99 ± (9.3) 7 a Olfactory Preferences of S. exigua Greenhouse S. exigua olfactory preference experiments did not detect differences in orientation between MHR and HS A. fatua plants (Figure 1.4). However, orientations towards Lactuca plants occurred significantly more often than to filtered air alone (Figure 1.5). This can be seen as the number of times a S. exigua motioned to a hole emitting headspace VOCs. Control experiments comparing two streams of filtered air showed a no significant preference to filtered air flow (Figure 1.6). This indicates that the headspace VOCs being pushed through the tubes are being sensed, and S. exigua is indicating towards a VOC profile. 29 Figure 1.4 - Olfactory choices of S. exigua regarding A. fatua lines R4 and S1. Each bar represents the count of olfactory choices made by S. exigua. Compact letter design is employed to indicate significant differences among the groups. 30 Figure 1.5 –Olfactory choices of S. exigua regarding Lactuca sativa L. and filtered airflow. Each bar represents the count of olfactory choices made by S. exigua. Compact letter design is employed to indicate significant differences among the groups. 31 Figure 1.6 – Bar graph illustrates the olfactory choices of S. exigua regarding filtered airflow and filtered airflow. Each bar represents the count of olfactory choices made by S. exigua. Compact letter design is employed to indicate significant differences among the groups. Discussion The main objective of this study was to investigate the attraction and feeding choices of the insect generalist herbivore S. exigua between MHR and HS A. fatua seedling plants in greenhouse and field settings. Field experiments were replicated using native grasshoppers. In greenhouse choice trials, larvae of the generalist herbivore S. exigua consumed R4 significantly more than S1. However, S2 choice trials did not yield statistically significant results. This behavior aligns with S. exigua's documented polyphagous nature, indicating its willingness to consume plant species it prefers (Hafeez et al., 2021; Moulton et al., 2000; Vickerman & Trumble, 1999). However, when S2 and R4 were used in a choice experiment, their consumption was relatively equal. This suggests that S2 is preferable over S1, but only in the presence of R4, indicating that damage by an herbivore does not consistently differ among 32 resistance and sensitive lines. When S2 and S1 were present in a choice experiment there was no significant difference between them. However, a reversal occurred in the presence of 93R, where S1 was consumed significantly more. Indicating that S. exigua does prefer S1 over some other lines, showing that they are making a conscious choice towards any line. Also, S2 and 93R did not show any significant difference. So, these results may show that R4 is heavily preferred over S1 but not compared to other lines. This trend was consistent across both metrics of measurement: area consumed (mm2) and the average consumed area by to the total shoot area. In no-choice experiments, S2 and R4 were found to be the most consumed lines, but data was only significantly larger than 20R2 in total area consumed. This data highlights the preference for R4 and S2 and indicates that S. exigua does not consume one more than the other during choice experiments. Notably, the comparison between S2 and 20R2 yields a noteworthy contrast, with the pattern changing when a competing choice was offered. While these lines exhibited a significant difference in the no-choice experiment, the presence of both in a choice experiment resulted in a consumption pattern where 20R2 was consumed more, albeit the differences were not significant. Observations in the field did not reveal any significant differences in total leaf area consumed in all lines, and insect generalist. This implies that environmental factors can have a mediating effect on herbivore choice and total consumption. For example, wind, sun, temperature can all influence the feeding behavior as well as VOC production from plants (Tauber et al., 1976; Holopainen and Gershenzon, 2010). The olfactory preference experiments yielded contradictory results compared to the choice experiments between R4 and S1. In the choice experiments, no significant differences were observed between R4 and S1. These differences suggest that S. exigua may exhibit a 33 preference for VOCs from R4 and S1, although consumption was greater for R4 plants, indicating that while VOCs may guide initial orientation or attraction, other factors such as nutritional value, physical plant characteristics, or secondary metabolites could be influencing the ultimate feeding choice of an insect herbivore (Moreira et al., 2016). A lack of consistency in olfactory data led me to conduct additional olfactory preference experiments to troubleshoot the experimental design for potential flaws. These experiments were done to extend and complement S. exigua studies previously performed using only S1 and R4 A. fatua lines (Keith et al, in preparation). The pilot study that was conducted found that S. exigua consumed MHR R4 and R3 lines 2.2 to 3.0-fold more than HS S1 lines. It is important to note that the data collection methods and experimental design were nearly identical to these experiments documented here. The R4 and S1 comparison holds true for this experiment; however, the findings for R3 and S1 were not. Considering the relatively low number of significant findings across all experiments, perhaps the previous studies our experiments were motivated by was not representative of the true relationship between plant stress and insect generalist. Research in this area remains relatively scarce, and the comparison between herbicide resistance and herbivore behavior can be convoluted. However, some previous findings offer valuable insights, particularly in the context of R4 and S1 preferences discussed earlier. Two studies by the same research team found that herbicide resistant Amaranthus hybridus were more susceptible to damage from herbivores (Gassmann & Futuyma, 2005; Gassmann, 2005). These papers support the findings for the R4 and S1 choice experiments but the 93R and S1 choice experiments present a contrast. 34 As the majority of these herbivory experiments have yielded insignificant results, and several of these experiments, particularly those conducted in controlled greenhouse environments, have been thoroughly tested. Therefore, alternative factors must be explored that can influence the preference of an insect generalist. One key factor that could potentially contribute to these preferences is genetic differences among these A. fatua lines, which may control mechanisms that render one line more preferable to another. This could be changing the expression of a gene that unintentionally makes a plant more attractive to an herbivore. For example, transgenerational effects and genome polyploidy could modify the generation of VOCs’ profiles of offspring (Picazo-Aragonés et al., 2020). VOCs have also been documented to play a significant role in herbivory defenses (Brilli et al., 2019; Zhou & Jander, 2021). As R4 and S1 headspace VOC profiles have been collected and analyzed, the rest of the lines have not been. R4 was consumed significantly more than S1 in choice trials possibly indicating that the 9 elevated VOCs from R4 could be attracting S. exigua. However, the VOC emissions from the other lines are unknown, and would require further investigation. It should be noted that several factors could have influenced the outcomes of this study. These may include unidentified variables altering plant attractiveness, genetic variations among A. fatua lines, or even the possibility of human errors during data collection. Consequently, further research is required to completely explore and understand these factors. The potential application of this research, which investigates the use of an insect generalist species to control invasive weeds, presents certain challenges. Notably, our experiments lack a comparison with crop species. It would be interesting to compare different A. fatua lines with various crop species and assess their attractiveness to insect herbivores. Further, conducting an in-depth analysis of all 35 A. fatua lines VOC profiles would be extremely valuable. Such an analysis could carry significant ecological implications if herbicide-resistant weeds are indeed becoming more attractive to insect generalists. This scenario might lead to an increase in insect damage among both crop and weed species, potentially posing a greater challenge for producers than the innate issues of an herbicide-resistant weed. Conclusion Avena fatua R4 lines were significantly preferred over S1 lines and S1 was preferred compared to 93R for choice experiments in a controlled greenhouse setting. Unfortunately, the remaining experiments, whether in choice or no-choice conditions, did not provide clear evidence of a preference by the insect generalist. Our field experiments also yielded inconclusive results, and in the greenhouse, only two choice experiments produced significant findings. One of these indicated that an MHR line was not preferred. This means that there are other differences being observed other than herbicide resistant capabilities between these sensitive and multiple resistant A. fatua lines. However, this research has highlighted certain aspects that could be responsible for this outcome. For example, both environmental changes and genetic variations are suspected of strong influence on the behavior of herbivorous insects. This observation, drawn from the outcomes of both greenhouse and field experiments, contributes to the ongoing discussion about the complexities of herbicide resistance and ecological implications. 36 CHAPTER THREE EXOGENOUS VOLATILE ORGANIC COMPOUNDS EFFECTS ON HERBICIDE RESISTANCE IN TWO AVENA FATUA LINES Summary This research is aimed to determine the effects of VOCs on the efficacy of herbicides in multiple herbicide resistant (MHR) R4 and herbicide sensitive (HS) S2 Avena fatua lines. Prior evidence has shown that MHR A. fatua lines are constitutively primed to resist stress, and their elevated VOC emissions are maybe a part of that defense. This study explored the potential of VOCs in plant-plant communication and their influence on herbicide resistance. R4 and S2 plants were exposed to VOCs, including linalool or limonene, or a combination VOC treatment including linalool, limonene, methyl salicylate, and methyl jasmonate, followed by a herbicide treatment of flucarbazone, imazamethabenz, or pinoxaden at sublethal doses. Linalool demonstrated a protective effect against flucarbazone in S2 lines, but this was not observed with other VOCs or herbicides. The combined VOC treatment enhanced herbicide injury in both S2 and R4 lines, particularly against imazamethabenz. This study contributes to understanding VOCs’ priming roles in herbicide injury responses. Future research should focus on isolating VOC effects in controlled environments and exploring the mechanisms underlying VOC- mediated stress tolerance, thereby enhancing our understanding of VOC roles in plant defenses. 37 Introduction Plants exhibit many complex defense mechanisms, including the ability to emit volatile organic compounds (VOCs) in response to stress. In addition to inducing internal defense responses, these compounds serve as chemical messengers, alerting neighboring plants to upcoming environmental stress and reinforcing defensive mechanisms (Brilli et al., 2019). These defenses are critical in response to environmental stressors like temperature extremes and drought (Bezerra et al., 2021; Bolsoni et al., 2018). The influence of VOCs on plant resilience to abiotic and biotic stressors has been well-documented in the literature (Brilli et al., 2019; Dudareva et al., 2013). Methyl jasmonate and methyl salicylate are methyl ester VOCs involved in several regulatory roles in stress response pathways. Methyl salicylate is critical for activating defense responses mainly associated with biotic stressors like pathogen infections (Li et al., 2019). Methyl jasmonate has known roles in priming plant defense mechanisms (Santino et al., 2013) and is mainly associated with defense responses to abiotic stressors like drought, soil salinity, and temperature change (Wang et al., 2021). The exogenous application of VOCs such as methyl salicylate and methyl jasmonate has been documented to alter the plant’s herbicide response. Herbicides can generate reactive oxygen species (ROS). Methyl salicylate pre-treatment generally attenuated ROS content and increased ROS-scavenging enzyme activity after applications of several herbicides such as halosulfuron-methyl and mesosulfuron-methyl (Khatami et al., 2022; Li et al., 2020). In contrast, methyl salicylate pre-treatment exacerbated atrazine injury in Nicotiana tabacum (Tobacco) and Arabidopsis thaliana (Thale cress) (Silverman et al., 2005). Pre-treatment with methyl jasmonate has shown a reduction in the 38 efficacy of paraquat (PSI inhibitor, Group 22) in Hordeum vulgare (barley) and Nicotiana tabacum (Tobacco), lessening the herbicide’s impact on photosynthetic efficiency (Hristova & Popova, 2002; Silverman et al., 2005). One example of an exogenous VOC effect on herbicide resistant plants, methyl jasmonate was found to decrease the efficacy of the growth-inhibitory herbicide quinclorac in resistant biotypes of Echinochloa crus-galli (barnyardgrass), while it did not affect herbicide response in quinclorac-sensitive strains (Jingjing, 2020). The VOC profile emitted can vary between herbicide resistant and sensitive biotypes and may contribute to differential herbicide responses. For example, in Avena fatua, monoterpenoids such as linalool and limonene, which were found to be released in higher quantities from R4 lines (Keith et al in preparation). Linalool, when applied in high concentrations, has been demonstrated to negatively affect growth parameters in Lolium rigidum (rigid ryegrass), suggesting that it may prime plants to be more susceptible to herbicide treatments (Vasilakoglou et al., 2013). Contrastingly, limonene did not exhibit a similar inhibitory impact in this study. However, limonene has been recognized for its role in conferring resistance against abiotic stresses, and linalool is primarily linked to the defense against biotic stresses (Fujioka et al., 2015; Llusia et al., 2005; Sun et al., 2014; Zhang et al., 2023). These findings illustrate the intricate roles that such compounds play in plant physiology and stress response mechanisms. MHR A. fatua plants constitutively emit elevated levels of six monoterpenoid, decanal, and methyl salicylate VOCs, and preferentially induce their emission in response to insect herbivory and mechanical wounding (Keith et al in preparation). These results raise the interesting possibility that VOCs are directly involved in MHR maintenance and evolution. Specifically, elevated VOCs from MHR plants could influence neighboring HS A. fatua plants in 39 their tolerance to an herbicide. A plant’s use of VOCs for stress communication is a well- established defense strategy, suggesting that MHR A. fatua may be “warning” neighboring plants against stressors, specifically herbicide stress. My goal was to determine if exogenous VOCs change HS or MHR plant responses to subsequent herbicide treatments. This work is based on preliminary studies that demonstrated that limonene and linalool improve resistance in S2 plants to the herbicides flucarbazone and pinoxaden, respectively. The preliminary findings contradict existing studies which indicate that exogenous linalool and limonene exacerbate the effects of herbicides, suggesting instead that the VOCs may confer protective benefits (Vasilakoglou et al., 2013). The goal is to investigate the influence of these VOCs individually and combined effect with all four on the efficacy of herbicides in MHR R4 and HS S2 A. fatua lines. It is well-established that VOCs can induce an enhanced stress response in plants facing abiotic challenges (Holopainen & Gershenzon, 2010). Including reactions to high temperatures, drought, and salinity. These responses underscore the complex role of VOCs in enhancing plant resistance (Possell & Loreto, 2013). Methods Plant Material Avena fatua lines S2 and R4 were grown in individual 10.16 cm2 pots to Zadok’s stage 13 size. All experiments were initiated mid-morning to minimize any potential circadian effects. Each individual experiment was run with A. fatua lines, S2 or R4. Growing conditions as described in Chapter 2: Insect Generalist Preference to MHR Avena fatua 40 Exogenous VOC Treatments Plants were placed in clear 61 cm x 61 cm (38L) polyethylene bags held upright with wire tomato cages, and the top of the bag was left open, for open bag systems. Closed bag was held up with a bamboo stake without wire tomato cages. Linalool (Sigma Aldrich, L2602) and limonene (Sigma Aldrich, 62118) were diluted in 100% ethanol to final concentrations of 0.56 M and 0.62 M, respectively. Control treatments received 100% ethanol only. 10 µL of each solution was dispensed onto 7.1 cm2 of 3mm filter paper (Whatman, United Kingdom), which was taped to the top of a 46 cm tall bamboo stake and secured upright in the center of plant pots. CO2 concentrations in closed and open bags were monitored (SCD30 CO2 Module, Sensirion, Switzerland). In both open and closed systems, plants were exposed to VOC treatments or ethanol for 6 hours under ambient greenhouse conditions. VOC mixtures were used to assess a synergistic treatment comprising linalool, limonene, methyl jasmonate (Sigma Aldrich, 392707), and methyl salicylate (Sigma Aldrich, 76631). Methyl jasmonate and methyl salicylate were gradually diluted in glass-distilled water to achieve a final concentration of 3 mM for each compound. To this solution, we added 0.125% (v/v) of a nonionic surfactant (X-77, Loveland Products Inc., Loveland, CO). Plants were relocated to a darkened room where they were uniformly misted with either the 3 mM methyl jasmonate and methyl salicylate solution or a control solution (consisting solely of 0.125% surfactant in water) until the point of runoff. After the application, the plants remained in the dark for two hours to avoid photodegradation of methyl salicylate and methyl jasmonate. Following this, the plants were transferred to the greenhouse where 0.56 M linalool and 0.62 M limonene was applied as described above. Linalool and limonene were administered onto the same piece of filter paper, 41 which was then removed after a 6-hour exposure, culminating in an 8-hour period for the combined VOC treatment. Herbicide Treatments After VOC treatment, plants were treated with a low concentration of herbicide. S2 plants were treated with flucarbazone (EverestTM), imazamethabenz (ASSERTTM), and pinoxaden (AXIALTM). Plants were treated with either 73.9, 83.7, or 98.8 g a.i. ha-1 (R4 plants), or 3, 4.9, 6.9, or 9.9 g a.i. ha-1 (S2 plants) flucarbazone, or 896.9, 1,029.4, or 1,175.2 g a.i. ha-1 (R4 plants) or 66.3, 110.5, 154.6, or 221 g a.i. ha-1 (S2 plants) imazamethabenz, or 13.1, 17.5, or 21.9 g a.i. ha-1 (R4 plants), or 2.6, 4.4, 6.1, or 8.8 g a.i. ha-1 (S2 plants) pinoxaden. Herbicides were applied using a moving nozzle sprayer in a total volume of 94L ha-1 as described (Keith et al., 2015). After herbicide sprays had dried, plants were returned to the greenhouse for 21 days, after which shoots were harvested, dried at 40°C for one week, and weighed. Statistics Mean dry biomass data from exogenous VOC and supplemental herbicide treatment were analyzed using linear mixed-effects models in R Studio (RStudio TEAM, 2023). To determine weight differences among treatments. Mean dry biomass data was subset by VOC and herbicide treatment to compare changes in the same herbicide, A. fatua line, and VOC treatments. Variation within and between trials was determined using random intercept adjustments for all models. Analysis of variance (ANOVA) was performed to assess treatment group differences. Estimated marginal means (EMMs) with Tukey adjustments were used for post-hoc comparisons among treatment groups (Searle et al., 1980). The S2 open bag exogenous linalool treatments included 16-31 technical replicates, which each experiment repeated 3-7 times. The rest of the 42 treatment combinations in open and closed bags were included 5-33 technical replicates with 1-3 experimental repeats. Data from multiple trials using the same VOC and herbicide, and all control treatments, were combined after confirming homoscedasticity using diagnostic plots. The R4 closed bag testing exogenous limonene with flucarbazone and pinoxaden, and exogenous linalool with all three herbicide treatments were performed only once, and so a linear model was used with no mixed effects for data analysis. Results Open Bag Experiments Two control treatments were used for open bag experiments: a 'no VOC' treatment and an 'ethanol' treatment. Ethanol, while not classified as a VOC, was utilized as a solvent in treatments involving linalool and limonene. Comparative analysis revealed no significant difference in the outcomes between the 'no VOC' and 'ethanol' treatments. Given this finding and ethanol's role in other treatments, we have designated ethanol as a standard control for all experimental comparisons and will be documented as “None”. This approach enables a consistent standard across treatments, focusing on VOC effects. Table 2.1 shows the mean dry biomass of S2 lines with exogenous linalool effects in open bag experiments. Linalool significantly reduced herbicide injury of plants treated with 3 g a.i. ha-1 flucarbazone, and increased herbicide injury with 4.4 g a.i. ha-1 pinoxaden. 43 Table 2.1: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Linalool Across Herbicide Treatments in Open Bags with P-Values Mean Dry Herbicide VOC Biomass n P Herbicide Dose (g ± SE) None 1.55 (± 0.04) 63 None None > 0.05 Linalool 1.53 (± 0.04) 67 None 0.39 (± 0.03) 31 flucarbazone 4.9 g a.i. ha-1 ≤ 0.05 Linalool 0.57 (± 0.08) 27 None 0.33 (± 0.02) 31 flucarbazone 9.9 g a.i. ha-1 > 0.05 Linalool 0.30 (± 0.01) 27 None 0.58 (± 0.06) 20 imazamethabenz 110.5 g a.i. ha-1 > 0.05 Linalool 0.70 (± 0.06) 20 None 0.45 (± 0.04) 21 imazamethabenz 221 g a.i. ha-1 > 0.05 Linalool 0.45 (± 0.03) 20 None 1.48 (± 0.08) 16 pinoxaden 4.4 g a.i. ha-1 ≤ 0.001 Linalool 0.64 (± 0.10) 16 None 0.34 (± 0.02) 16 pinoxaden 8.8 g a.i. ha-1 ≤ 0.01 Linalool 0.64 (± 0.10) 16 Table 2.2 shows the mean dry biomass of S2 lines with exogenous limonene effects in open bag experiments. There was no significance detected among any of the treatments. 44 Table 2.2: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Limonene Across Herbicide Treatments in Open Bags with P-Values Mean Dry Herbicide Herbicide Dose VOC Biomass n p (g ± SE) None 1.72 (± 0.06) 8 None None > 0.05 Limonene 1.64 (± 0.06) 8 None 0.48 (± 0.08) 8 flucarbazone 4.9 g a.i. ha-1 > 0.05 Limonene 0.59 (± 0.13) 8 None 0.41 (± 0.05) 8 flucarbazone 9.9 g a.i. ha-1 > 0.05 Limonene 0.34 (± 0.04) 8 None 0.62 (± 0.09) 8 imazamethabenz 110.5 g a.i. ha-1 > 0.05 Limonene 0.59 (± 0.05) 8 None 0.40 (± 0.05) 8 imazamethabenz 221 g a.i. ha-1 > 0.05 Limonene 0.39 (± 0.04) 8 None 1.01 (± 0.12) 12 pinoxaden 4.4 g a.i. ha-1 > 0.05 Limonene 1.12 (± 0.12) 12 None 0.43 (± 0.05) 12 pinoxaden 8.8 g a.i. ha-1 > 0.05 Limonene 0.34 (± 0.05) 12 CO2 Concentrations The rest of the experiments were performed using a closed bag system, as cross- contamination or evaporation could be occurring in open bag experiments. CO2 concentrations were monitored in a closed system and found that the CO2 levels did decrease to 150 ppm and stayed steady at 150 ppm for the remainder of the experiment, indicating the plants were experiencing stress. In contrast, the open bag systems had CO2 levels above 350 ppm. CO2 above 350 ppm is considered non-stress (Dippery et al., 1995; Gerhart & Ward, 2010). 45 The control treatments used for closed bag experiments were “no VOC in closed bag”, and “ethanol in a closed bag”. Table 2.3 shows the mean dry biomass of S2 lines with exogenous linalool effects in a closed bag experiment. For flucarbazone treatments at 3 and 6.9 g a.i. ha-1, the linalool treated plants had significantly higher biomass compared to the control. The imazamethabenz treatments yielded no significant differences. For the pinoxaden-treated group, a significant difference was recorded at the 2.6 g a.i. ha-1 dose as the linalool treatment was significantly lower than the control. There were no other significant differences among the other pinoxaden treatments. 46 Table 2.3: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Linalool Across Herbicide Treatments in Closed Bags with P-Values Mean Dry Herbicide Herbicide Dose VOC n p Biomass (g ± SE) None 1.48 (± 0.04) 29 None None > 0.05 Linalool 1.32 (± 0.07) 33 None 0.51 (± 0.04) 18 flucarbazone 3 g a.i. ha-1 ≤ 0.01 Linalool 0.63 (± 0.05) 18 None 0.52 (± 0.03) 18 flucarbazone 4.9 g a.i. ha-1 > 0.05 Linalool 0.55 (± 0.05) 18 None 0.39 (± 0.02) 18 flucarbazone 6.9 g a.i. ha-1 ≤ 0.01 Linalool 0.50 (± 0.03) 18 None 0.47 (± 0.07) 5 imazamethabenz 66.3 g a.i. ha-1 > 0.05 Linalool 0.45 (± 0.04) 5 None 0.37 (± 0.04) 11 imazamethabenz 110.5 g a.i. ha-1 > 0.05 Linalool 0.30 (± 0.02) 11 None 0.37 (± 0.04) 5 imazamethabenz 154.6 g a.i. ha-1 > 0.05 Linalool 0.43 (± 0.02) 5 None 1.19 (± 0.05) 15 pinoxaden 2.6 g a.i. ha-1 ≤ 0.01 Linalool 1.00 (± 0.08) 15 None 0.45 (± 0.06) 15 pinoxaden 4.4 g a.i. ha-1 > 0.05 Linalool 0.47 (± 0.06) 15 None 0.34 (± 0.04) 15 pinoxaden 6.1 g a.i. ha-1 > 0.05 Linalool 0.32 (± 0.03) 15 Table 2.4 shows the mean dry biomass of S2 lines with exogenous limonene effects in a closed bag experiment. Limonene reduced the herbicide injury significantly for the 4.9 g a.i. ha-1 dose of flucarbazone. No other significant differences were found. 47 Table 2.4: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Limonene Across Herbicide Treatments in Closed Bags with P-Values Mean Dry Herbicide Herbicide Dose VOC n P Biomass (g ± SE) None 1.28 (± 0.05) 27 > None None Limonene 1.14 (± 0.06) 31 0.05 None 0.24 (± 0.02) 8 flucarbazone 4.9 g a.i. ha-1 ≤ Limonene 0.30 (± 0.02) 8 0.001 None 0.74 (± 0.10) 10 > imazamethabenz 66.3 g a.i. ha-1 Limonene 0.56 (± 0.04) 10 0.05 None 0.41 (± 0.04) 10 > imazamethabenz 110.5 g a.i. ha-1 Limonene 0.39 (± 0.01) 10 0.05 None 0.36 (± 0.02) 10 > imazamethabenz 154.6 g a.i. ha-1 Limonene 0.35 (± 0.02) 10 0.05 None 0.92 (± 0.14) 9 > pinoxaden 2.6 g a.i. ha-1 Limonene 0.90 (± 0.09) 9 0.05 None 0.83 (± 0.07) 9 > pinoxaden 4.4 g a.i. ha-1 Limonene 0.80 (± 0.07) 9 0.05 None 0.35 (± 0.04) 9 > pinoxaden 6.1 g a.i. ha-1 Limonene 0.29 (± 0.02) 9 0.05 Table 2.5 shows the mean dry biomass of R4 lines with exogenous linalool effects in a closed bag experiment. No significant differences were observed. 48 Table 2.5: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for R4 Lines with Exogenous Linalool Across Herbicide Treatments in Closed Bags with P-Values A. Mean Dry fatua Herbicide Herbicide Dose VOC Biomass n P Line (g ± SE) None 1.36 (± 0.07) 26 R4 None None > 0.05 Linalool 1.36 (± 0.07) 26 None 0.73 (± 0.05) 6 R4 flucarbazone 73.9 g a.i. ha-1 > 0.05 Linalool 0.66 (± 0.05) 6 None 0.66 (± 0.05) 6 R4 flucarbazone 78.8 g a.i. ha-1 > 0.05 Linalool 0.71 (± 0.06) 6 None 0.69 (± 0.05) 6 R4 flucarbazone 83.7 g a.i. ha-1 > 0.05 Linalool 0.82 (± 0.08) 6 None 1.03 (± 0.13) 11 R4 imazamethabenz 896.9 g a.i. ha-1 > 0.05 Linalool 0.97 (± 0.09) 11 None 0.82 (± 0.06) 11 R4 imazamethabenz 1,029.4 g a.i. ha-1 > 0.05 Linalool 0.96 (± 0.09) 11 None 0.79 (± 0.08) 11 R4 imazamethabenz 1,175.2 g a.i. ha-1 > 0.05 Linalool 0.80 (± 0.04) 11 None 0.74 (± 0.06) 9 R4 pinoxaden 17.5 g a.i. ha-1 > 0.05 Linalool 0.69 (± 0.05) 9 Table 2.6 shows the mean dry biomass of R4 lines with exogenous limonene effects in a closed bag experiment. No significant differences were observed. 49 Table 2.6: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for R4 Lines with Exogenous Limonene Across Herbicide Treatments in Closed Bags with P-Values A. Mean Dry fatua Herbicide Herbicide Dose VOC Biomass n P Line (g ± SE) None 0.85 (± 0.03) 26 R4 None None > 0.05 Limonene 0.85 (± 0.03) 26 None 0.28 (± 0.04) 8 R4 flucarbazone 78.8 g a.i. ha-1 > 0.05 Limonene 0.39 (± 0.01) 8 None 0.39 (± 0.05) 12 R4 imazamethabenz 1,029.4 g a.i. ha-1 > 0.05 Limonene 0.40 (± 0.07) 12 None 0.74 (± 0.06) 8 R4 pinoxaden 17.5 g a.i. ha-1 > 0.05 Limonene 0.77 (± 0.08) 8 Table 2.7 shows the mean dry biomass of S2 lines with the combined exogenous VOCs and its effects on herbicide potency. The combined VOCs had a significantly negative effect for the imazamethabenz and pinoxaden treatment. 50 Table 2.7: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for S2 Lines with Exogenous Combined VOCs Across Herbicide Treatments in Closed Bags with P-Values Mean Dry Biomass Herbicide Herbicide Dose VOC n P (g ± SE) None 1.17 (± 0.06) 42 > None None Combined 1.16 (± 0.06) 42 0.05 VOCs None 0.33 (± 0.02) 18 -1 > flucarbazone 4.9 g a.i. ha Combined 0.33 (± 0.02) 18 0.05 VOCs None 0.43 (± 0.04) 33 -1 ≤ imazamethabenz 110.45 g a.i. ha Combined 0.27 (± 0.02) 32 0.001 VOCs None 0.36 (± 0.04) 16 ≤ pinoxaden 4.4 g a.i. ha-1 Combined 0.26 (± 0.02) 16 0.05 VOCs Table 2.8 shows the mean dry biomass of R4 lines with the combined exogenous VOCs and its effects on herbicide potency. Combined VOCs had a significantly negative effect on dry biomass for imazamethabenz treatment. No significant differences were found for the other treatments. 51 Table 2.8: Pairwise Comparisons of Mean Dry Biomass (Grams ± SE) for R4 Lines with Exogenous Combined VOCs Across Herbicide Treatments in Closed Bags with P-Values Mean Dry A. fatua Herbicide Herbicide Dose VOC Biomass n P Line (g ± SE) None 1.18 (± 0.10) 26 R4 None None Combined > 0.05 1.14 (± 0.11) 28 VOCs None 0.81 (± 0.08) 18 R4 flucarbazone 78.8 g a.i. ha-1 Combined > 0.05 0.69 (± 0.06) 18 VOCs None 1.06 (± 0.11) 18 R4 imazamethabenz 1,029.4 g a.i. ha-1 Combined ≤ 0.05 0.77 (± 0.10) 18 VOCs None 0.39 (± 0.01) 25 R4 pinoxaden 17.5 g a.i. ha-1 Combined > 0.05 0.35 (± 0.04) 25 VOCs Discussion The objective of these studies was to investigate whether VOCs influence the efficacy of herbicides in MHR R4 and HS S2 A. fatua lines. Prior research has demonstrated that MHR A. fatua lines have an inherent readiness to withstand various stressors (Burns et al., 2018; Dyer, 2018; Keith et al., 2017), and that they exhibit higher VOC emissions than S2 counterparts, indicating a possible role in plant-plant communication. To test these ideas, MHR and HS plants were exposed to exogenous VOCs, followed by treatments with flucarbazone, imazamethabenz, and pinoxaden herbicides at several sublethal doses. Preliminary experiments, in a closed bag system, demonstrated that limonene and linalool exhibited a protective effect against flucarbazone and pinoxaden applications, respectively. It was later suggested by a colleague that these enclosed environments might impose CO2 52 limitations on the treated plants. Therefore, subsequent experiments I conducted were in an open bag setup to mitigate CO2 constraints. Despite the open-bag approach, concerns regarding potential cross-contamination and evaporation led to a reversion to the closed-bag method in later experiments. CO2 levels within these closed environments were monitored, revealing a decline in concentration after the first hour, eventually stabilizing at 150 ppm. At 150 ppm a plant is considered CO2 stressed (Dippery et al., 1995; Gerhart & Ward, 2010). Consequently, all experiments conducted in closed bags were classified as experiencing CO2-stressed conditions. Notably, control experiments within these CO2-stressed environments displayed no significant differences in plant growth compared to those conducted without bag confinement. This outcome suggests that the CO2 limitation inherent to the closed-bag setup did not notably impact plant growth under the conditions of our study. Linalool had a protective effect in closed and open bags to flucarbazone, however it increased injury to pinoxaden. Pretreatment of S2 plants with linalool reduced injury from 4.9 g a.i. ha-1 flucarbazone in the open bag system (Table 2.1). In the closed bag system, 3, and 6.9 g a.i. ha-1 flucarbazone treatments found that linalool treated plants had higher dry biomass weight (Table 2.3). Indicating that linalool pre-treatments have a protective role to sublethal flucarbazone doses. However, the pinoxaden dose of 4.4 g. a.i. ha-1 shows a potentiating effect of pinoxaden, as it was significantly lower weight than the control in the open bag systems. For the closed bag systems, linalool had the same negative effect for 2.6 g a.i. ha-1 dose of pinoxaden. This indicates that linalool has a protective role against flucarbazone, but a non-protective role for pinoxaden. There were no significant differences found for any of the imazamethabenz 53 sublethal herbicide doses. R4 plant exposed to linalool in the closed system did not show any significance either (Table 2.5). Limonene had a slight protective effect against flucarbazone but did not show any protective roles for the other herbicides used. The pretreatment of exogenous limonene did not confer protection to S2 plants against any herbicide treatments when conducted in an open bag system (Table 2.2). However, in a closed system environment, limonene pretreatment was observed to mitigate the damage caused by a 4.9 g a.i. ha-1 dose of flucarbazone, suggesting a specific protective effect of limonene against this herbicide dose (Table 2.4). R4 plants treated with limonene did not exhibit any significant response to various herbicide treatments (Table 2.6). The absence of a significant protective effect in several cases implies that limonene may not provide resistance to herbicide induced stress. These findings imply that the role of limonene might be more aligned with conferring resistance to different abiotic stresses rather than mitigating the effects of herbicide induced abiotic stress (Fujioka et al., 2015; Llusia et al., 2005). The combined effects of the four VOCs generally increased the injury to all the herbicides showing a lack of protection across all experiments. The application of linalool, limonene, methyl salicylate, and methyl jasmonate in a combined treatment did not grant any enhanced protection against imazamethabenz or pinoxaden with S2 plants and instead enhanced the effects of those herbicides significantly (Table 2.7). Priming those plants to be more prone to herbicide injury. R4 plants exposed to the combined VOC treatment had a similar effect for the imazamethabenz treatment (Table 2.8). However, there was no effect with flucarbazone and pinoxaden for the R4 plants. 54 The application of linalool to S2 plants demonstrated an attenuation of the effects of flucarbazone across various sublethal doses, suggesting a protective role of linalool against this specific herbicide. However, this protective effect was not observed when linalool was combined with other VOCs such as limonene, methyl salicylate, and methyl jasmonate. The combined treatment of these VOCs with linalool appeared to exacerbate herbicide induced injury, indicating a potential synergistic effect that intensifies herbicide injury. Intriguingly, although imazamethabenz shares the same target enzyme (acetolactate synthase) as flucarbazone, they belong to different chemical classes: imazamethabenz to the imidazolinones and flucarbazone to the sulfonylureas. This difference could imply that linalool's protective role may be more specific to the sulfonylurea class of herbicides, as opposed to the imidazolinones. The protective role of limonene in our study presents an area of uncertainty. Notably, limonene was observed to confer protection to S2 plant lines against a specific dose of flucarbazone 4.9 g a.i. ha-1. This finding was from experiments conducted within a closed bag system. This isolated instance might suggest a potential protective role of limonene. However, such protective effects were not found in the open bag system. Furthermore, similar experiments involving R4 plants also failed to demonstrate any protective effects of limonene against the tested herbicides. These contrasting results imply that limonene's protective role is either non- existent or highly conditional, possibly influenced by the environmental setup of the experiment. The absence of a consistent protective effect across different experimental conditions and plant lines suggests that limonene may not play a significant role in herbicide resistance for the herbicides tested. 55 In the combined VOC treatments experiments, the combined application of linalool, limonene, methyl salicylate, and methyl jasmonate demonstrated an exacerbation of herbicide induced injury in both S2 and R4 plant lines. For the S2 lines, this combined VOC treatment intensified the effects of imazamethabenz and pinoxaden. Similarly, in R4 lines, the combined VOC application was observed to potentiate the herbicide injury of imazamethabenz. These findings suggest that the simultaneous presence of these specific VOCs does not confer a protective effect against the herbicides tested. Instead, it appears to enhance the herbicidal injury. This outcome indicates that the protective roles, if any, of individual VOCs like linalool or limonene do not translate into a combined treatment context. This research, while yielding significant results, should be viewed in the context of certain methodological issues. One notable limitation was the small-time frame of the combined VOC treatments, which restricted the experiments to a single herbicide dose. I also encountered variability in plant responses, particularly in terms of dry biomass weight following herbicide application. This was most evident in the S2 lines, where even low herbicide concentrations led to high herbicide injury, indicating a higher sensitivity than anticipated. Additionally, changes in dry weight responses within the same herbicide and VOC treatment groups could be credited to variabilities in experimental application and environmental conditions, such as differential light exposure (cloudy or sunny days) and watering regimes. Despite these challenges, the validity of the results is supported by the control treatments within each experiment. These controls were subjected to the same conditions as the treated plants, thereby providing a reliable baseline for comparison. 56 These results add to the discussion of VOCs and priming stress response mechanisms. It has been documented numerous times that VOCs through plant-plant communication can warn other plants of impending stress events, and receiver plants effectively mitigating or lessening the impact of the stressor (Brilli et al., 2019; Ueda et al., 2012). These indirect defenses can alter the way plants handle abiotic and biotic effects. As there are some significant differences among the S2 lines response to herbicide after being exposed to a VOC, but there is not a consistent outcome with specific herbicides and VOCs. Additionally, some results contrast certain herbicide treatments, therefore adding to the complexity of VOC priming defenses for herbicide applications. Future studies should research precise mechanisms of volatile organic compounds that are known to enhance stress tolerance such as methyl jasmonate and methyl salicylate. It would also be beneficial to employ a more precise experimental setup, such as the use of a sealed growth chamber. This would allow VOC treatments to be primarily focused without the confounding factors of cross-contamination or evaporation, while also facilitating the critical regulation of CO2 levels. In addition to a controlled environment, a detailed methodical approach would incorporate testing ranges of VOCs concentrations, and incorporating other VOCs that are emitted from the MHR R4 lines, thereby fulling understanding the complex and nuanced mechanisms that go into the VOC mediated herbicide resistance for A. fatua lines. 57 Conclusion In conclusion, this study has provided valuable insights into the interactions between VOCs and herbicides in MHR R4 and HS S2 A. fatua lines. The research revealed responses of these lines to various combinations of VOCs and herbicides, to understand the complexity of plant stress responses. While linalool demonstrated a protective effect against flucarbazone in S2 plants under certain conditions, this effect was not seen in R4 plants. Limonene did not have a strong effect to prime plants against herbicide stress. The combined application of multiple VOCs tended to exacerbate herbicide induced injury, suggesting that these compounds may interact synergistically to strengthen herbicide injury rather than lessen them. More research is required to see if methyl jasmonate and methyl salicylate have significant effects towards herbicide application. These findings contribute to the understanding of plant-plant communication and VOCs' role in priming stress response mechanisms, particularly in the context of herbicide resistance. 58 CHAPTER FOUR PLANT PIGMENTS DETECTION, PHOTOSYNTHETIC AND ENERGY MANAGEMENT PARAMETERS OF MULTIPLE HERBICIDE RESISTANT AVENA FATUA Summary Due to the observations of color differences of R4 and S2 A. fatua lines grown in a growth chamber, plant pigment and photosynthetic fluorometry of multiple herbicide sensitive and resistant Avena fatua lines, and two derived recombinant inbred lines (RILs) populations were examined. Thin layer chromatography (TLC) to physically separate and observe pigments, revealed some pigment differences among lines, warranting further investigation. Spectrophotometric analyses were then used to compare pigment absorbance values of acetone extracts, revealing significantly higher levels of beta-carotene in and chlorophyll b in MHR plants. Through pulse amplitude modulated (PAM) fluorometry, changes in distinct light energy management capabilities, such as lower non-regulated energy quenching and higher effective photochemical yield, were identified in R4 and R3 plants. This indicates that these lines are better equipped to utilize light energy and quench excess energy through non-destructive mechanisms. This research provides foundational insights into pigment levels, photosynthetic behavior, and light-energy management of HS and MHR A. fatua lines. Future work will be required to determine if there is a direct connection to non-target site resistance (NTSR) mechanisms. 59 Introduction The multiple herbicide resistance capabilities conferred to Avena fatua lines arise through NTSR mechanisms (Burns et al., 2018; Dyer, 2018; Keith et al., 2017). NTSR mechanisms include a reduction in herbicide uptake or translocation, enhanced rates of herbicide metabolism, a decreased rate of herbicide activation and sequestration, a mutation in a gene other than one encoding the herbicide’s target enzyme, or other unidentified mechanisms (Délye et al., 2013; Dyer, 2018; Jugulam & Shyam, 2019). Since many herbicides cause increased reactive oxygen species (ROS) content (Traxler et al., 2023), an enhanced ability to quench ROS could also confer NTSR capabilities. Plants have innate abilities to quench ROS through antioxidants, specifically enzymatic and non-enzymatic antioxidants (Ahmad et al., 2010; Sharma et al., 2022). Non-enzymatic antioxidants such as carotenoids and flavonoids play crucial roles in quenching ROS. The upregulation of carotenoids and flavonoids has been documented to alleviate abiotic stress in crops (Davison et al., 2002; Du et al., 2010; Götz et al., 2002; Römer & Fraser, 2005). If these protective molecules are found to be in higher concentration in MHR lines than HS, this might be an accomplice or causal to the NTSR mechanism. A plants’ non-photochemical quenching (NPQ) capacity is also linked to abiotic stress tolerance (Müller et al., 2001). Chlorophyll fluorometry, a non-invasive technique for assessing PSII activity, offers a tool to measure and compare photosynthetic mechanisms and physiological characteristics across plants. This method provides detailed insights into the efficiency and health of the photosynthetic process, allowing for a deeper understanding of plant responses under various environmental conditions (Murchie & Lawson, 2013). These assays show several fluorescence parameters that can show maximum/minimum chlorophyll fluorescence, PSII 60 operating efficiency, and quantum yield of photochemistry. These parameters commonly quantify photosynthetic performance, such as electron transfer rate and total photochemical yield of photons (Murchie & Lawson, 2013). In addition, chlorophyll fluorometry can also measure non- photochemical quenching (NPQ) mechanisms, which are photoprotective processes that remove excess energy through excited states. Non-stressed plants rely on these processes to allow the quenching of excitation energy (Ruban et al., 2012). Simplified, chlorophyll fluorometry can quantify the amount of light being quenched and the amount being used in photochemistry (Baker, 2008). The approach pulse amplitude fluorometry (PAM) estimates the number of photons being used in photochemistry versus NPQ. For this procedure, a plant is exposed to an intense light beam and the number of photons emitted is determined. PAM can also calculate the electron transport rate (ETR) of electrons passing through PSII, giving a reliable estimate of the plant's photosynthetic capacity (Raymond & Bunthawin, 2010). As such, ETR can be used to estimate the plant's overall photosynthetic electron transport efficiency. This efficiency is a crucial determinant of how effectively the plant converts light energy into chemical energy during photosynthesis, directly influencing its growth and productivity (Murchie & Lawson, 2013). Thus, plant pigment detection/quantification and fluorometry can give a reliable idea of how plants utilize light energy for photochemistry, manage excess energy and ROS. Several parameters were measured to identify these metrics and their potential role in NTSR mechanisms in MHR Avena fatua lines. 61 Methods Plant Material All Avena fatua lines were grown to Zadok’s stage 13 under the same greenhouse conditions described in Chapter 2: Insect Generalist Preference to MHR Avena fatua. RILs were used for pulse amplitude modulated fluorometry readings. These included the 93R inbred line that was crossed with the S2 HS line to create 124 RILs (Kern et al., 1996). Similarly, seed from the population found in 2006 were used to create the inbred MHR lines R4, which were crossed with S2 to create the 06 RIL (n = 120) population. Progeny from both crosses was advanced using single seed descent (Jinks & Pooni, 1976) for six generations to achieve a homozygosity rate of 98.44 %. Out of the 120 06 RILs, 35 lines were used, that all had a range of herbicide resistance capabilities. Out of the 124 93 RILs, 23 lines were used, that all had a range of herbicide resistant capabilities as well. Thin Layer Chromatography For thin layer chromatography, plants were grown in either controlled environment chambers (500 µmol m-2 sec1 PAR; high light) or under the greenhouse conditions (~300 µmol m-2 sec1 PAR; low light). Shoots from HS2 and MHR plants were harvested at the soil surface, and 0.6 g (FW) were immediately frozen in liquid nitrogen and stored at -80 °C. All subsequent steps were conducted under very low light conditions (~50 µmol m-2 sec1) to protect light- sensitive pigments. Shoot material was ground three times in a liquid nitrogen pre-cooled pestle and mortar, transferred into 2 mL Eppendorf tubes, acetone (400 µL) was added and vortexed for 2 minutes. Tubes were centrifuged 14,500 RPM at 2 °C for 5 minutes, and 150 µL of the supernatant was removed and spotted onto silica gel plates (20 x 20 cm TLC plates, Sigma 62 Aldrich, Z122785, St. Louis, MO) at 2 cm above the bottom edge with 20 µL microcapillary pipettes (Drummond Microcaps, 1-000-9000, Broomall, PA). TLC plates were placed into a 20 cm rectangular glass chromatography tank (Biostep, BS120.173A, Sun Valley, CA) containing 40:60 petroleum ether: acetone (by vol.) to a depth of 1.5 cm. After 20 minutes, plates were removed, dried, and photographed under room light and UV light. Spectrophotometer Assay Shoot material was harvested from HS and MHR plants was immediately frozen in liquid nitrogen and stored at -80 °C. Shoot material (0.3 g FW) was placed in pre-chilled 2 mL Eppendorf tubes containing three 2 mm glass beads. The frozen tubes were placed in a bead beater (Mini-Beadbeater-96 Biospec, Bartlesville, OK) and ran for one minute. Acetone (200 µL) was added to each tube, followed by vortexing (2 minutes) and centrifugation at 14,500 RPM at 2 °C for 5 minutes. Supernatant (100 µL) was removed to a new tube and kept on ice. A subsample (15 µL) of supernatant was mixed into 985 µL acetone in a 2 mL quartz cuvette (Sigma Aldrich, St. Louis, MO), and samples were loaded into a microplate reader (SpectraMax M2, Molecular Devices, San Jose, CA) and scanned using two modes. Mode 1 was a full scan of 300 nm – 750 nm at 2 nm intervals, and Mode 2 was a kinetic scan at 451, 480, 530, 538, 652, and 665 nm, which are absorbance maxima for Beta-carotene, betaxanthin, anthocyanin, betacyanin, chlorophyll b, and chlorophyll a, respectively (Peters & Noble, 2014). Pulse Amplitude Modulated Fluorometry Avena fatua field derived lines and 06R and 93R RIL populations were grown as described above. Individual RILs were grouped according to their resistance phenotypes. 06 RILs were grouped according to their resistance towards flucarbazone and triallate. 93 RILS 63 were grouped based on their resistance levels to triallate. These resistance groups are defined as Low (< 24% of resistant parent’s effective dose), Medium (25% to 75% of resistant parent’s effective dose), and High (>76% of resistant parent’s effective dose). Within the 06R RIL population, there were 14 lines classified as 'Low' resistance, 8 as 'Medium', and 13 as 'High'. For the 93R RILs, there were 7 lines in the 'Low' resistance group, 9 in the 'Medium', and 7 in the 'High' resistance category. Plants were transferred to a darkened room (~10 µmol m2 sec) and kept for 30 minutes before the experiment continued. These low light conditions were maintained throughout experiments. A two-saturation pulse analysis was performed using a PAM MINI II instrument (Heinz WALZ, Effeltrich, Germany). Intact plants were first exposed to a one second initial actinic light saturation pulse, followed by a secondary saturation pulse of one second after 30 seconds. Actinic light from a quasi-constant light source in the instrument was utilized to stimulate photosynthesis, while saturation pulses were used to saturate primary photosynthetic molecules. The parameters measured were the fluorescence minimum, maximum, effective photochemical yield, regulated and non-regulated quenching, and electron transport rate. Statistics Plant pigment and PAM data were analyzed using linear mixed-effects models in R Studio (RStudio TEAM, 2023). Variation within and between trials was assessed with random intercept adjustments for all models. Analysis of variance tests (ANOVA) were performed to assess treatment group differences. Estimated marginal means (EMMs) with Tukey adjustments were used for post-hoc comparisons among treatment groups (Searle et al., 1980). The Plant pigment detection using spectrophotometry experiments were replicated 4 times, and treatments 64 were technically replicated 7-10 times. Plant fluorimetry experiments were replicated 3 times and treatments were technically replicated 3-12 times. Field derived A. fatua lines were utilized more for both and are responsible for the high number of replicates. Data from experimental repeats were combined after confirming homoscedasticity using diagnostic plots. Results Thin Layer Chromatography TLC assays were used to semi-quantitatively detect pigment levels in S2 and R4 plants grown under growth chamber (high light) and greenhouse (low light) conditions. Pigments were tentatively identified based on Rf values from similar published assays (Quach et al., 2004). An observation was made about the diminished presence of the pheophytin band in R4 plants grown in the growth chamber, compared to their greenhouse counterparts (Figure 3.1). However, the opposite was observed in the R4 plants when the experiment was repeated; the pheophytin band was much stronger in R4 plants grown in a growth chamber compared to greenhouse conditions (Figure 3.2). Pheophytin is a chlorophyll derivative that functions as primary electron acceptor and has partial roles in preventing photodamage under high light conditions (Hou, 2014). No consistent differences were observed across plant samples for the other plant pigments. However due to the inherit limitations of TLC a more robust approach was required. These observations highlighted potential differences between the plant lines. Motivating the quantification of plant pigment concentrations through spectrophotometric analysis of plants grown under consistent light conditions. 65 Figure 3.1 – Thin layer chromatography of acetone extracts of S2 (left panel) and R4 (right panel) plants. The left and right hand extracts in both panels are from plants grown in the greenhouse and growth chamber, respectively. Figure 3.2 - Thin layer chromatography of acetone extracts of two R4 plants grown in different conditions. The left extract is the R4 A. fatua line grown in the greenhouse, and the right extract is the R4 line grown in the growth chamber. Spectrophotometer Assay Spectrophotometer assays showed absorbance values of acetone extracts from MHR and HS plants in 400 to 750 nm scans and at specific wavelengths. Direct quantification of pigment concentrations from absorbance values is possible but relies on pure samples without interfering substances (Lichtenthaler, 1987; Wellburn, 1994). Therefore, I reported only absorbance data 66 from my pigment mixed samples here, and then used these data to determine estimated pigment concentrations subject to some interpretation. A full scan of A. fatua field derived lines (Table 3.3) reveals a notable distinction in the absorption concentration of acetone soluble pigments. Specifically, the lines R3, R4, and 20R2 exhibit higher absorption levels and possess a greater concentration of acetone soluble pigments when compared to the S1, S2, and 93R lines. Figure 3.3 – Spectrophotometer Full scan of MHR and HS A. fatua lines ranging from 400- 750 nm. Pigment absorbance values from two HS and four MHR A. fatua lines showed no significant differences at 451 nm, 530 nm, 538 nm, or 665 nm. The concentration of acetone soluble pigments detected at 480 nm showed that 20R2’s concentration was significantly higher than the concentration of 93R (Figure 3.4). The other field derived lines were not significantly 67 different from one another. Extraction of shoot material in the polar solvent acetone favored solvation of chlorophyll and carotenoids, while flavonoids like anthocyanins would be present in low abundance (Silva et al., 2017). The most common plant pigment that has an absorption maximum at 480 is beta-carotene. Figure 3.4 – Spectrophotometer, Acetone soluble pigment concentrations of A. fatua lines detected at 480 nm. Compact letter design is employed to illustrate significant differences. The line center in box is median, the box represents the interquartile range of data observed. Figure 3.5 displays the acetone soluble pigment concentrations detected at 652 nm. R4 was significantly higher in concentration compared to 93R. The plant pigment commonly found here with acetone extractions is chlorophyll b. 68 Figure 3.5 – Spectrophotometer, Acetone soluble pigment concentrations of A. fatua lines detected at 652 nm. Compact letter design is employed to illustrate significant differences. The line center in the box is the median, the box represents the interquartile range of data observed. Pulse Amplitude Modulated Fluorometry Assay PAM assays revealed several informative photosynthetic and energy management parameters from HS and MHR plants. These parameters include: 1) Y(NPQ), representing effective non-photochemical energy dissipation. This shows a change in regulated quenching capacity, where these plants are releasing excess energy as heat, that prevents damage to the photosynthetic mechanism. This utilizes the xanthophyll cycle to dissipate excess energy; 2) Y(NO) represents quantum yield of non-regulated energy dissipation, or the fraction of absorbed light that dissipated as heat through non-regulated processes with the triplet-triplet energy release; 3) Y(II) which represents quantum yield of PSII photochemistry, or the fraction of absorbed light that is used in the photochemical process, and 4) (ETR) represents electron 69 transport rate and is an estimate of the rate at which electrons are moving through the photosynthetic electron transport chain (Hashimoto et al., 2018; Raymond & Bunthawin, 2010). The sum of the first three measurements is equal to 1 for a given absorbed photon flux: Y(II)+Y(NPQ)+Y(NO) = 1. Note that mean values can create a total that deviates slightly from 1.0. Equations for these metrics are as follows: Y(II) = Fm′−F / Fm′; Y(NO) = F / Fm’; Y(NPQ) = F/Fm’ – F/Fm. Where Fm′ is the fluorescence level at the saturation pulse, F is the fluorescence level before the saturation pulse and Fm is the maximum fluorescence level. Table 3.1 shows the values of Fm’, F, and Fm for all the lines used for the PAM assays. Chlorophyll fluorescence measurements represent the intensity of fluorescence and are typically dimensionless and presented in arbitrary units (Baker, 2008; Murchie & Lawson, 2013; Raymond & Bunthawin, 2010). 70 Table 3.1: Fluorescence Measurements Used for Photosynthetic and Energy Management Parameters of All A. fatua Lines Tested (measurements are arbitrary units) A. fatua Lines and RIL resistant groups Fm' F Fm S2 2034.9 1637.1 2890.8 S1 2019.6 1578.1 2918.4 R3 1901.5 1424.7 3284.6 R4 1771.1 1279.6 3143.8 20R2 2344.4 2012.9 3198.1 93R 2101.6 1674.7 3014.6 Low resistant group for 06 RILs 2001.2 1497.8 3085.1 Medium resistant group for 06 RILS 2005.5 1512.1 3112.9 High resistant group for 06 RILs 1961.5 1501.1 3055.6 Low resistant group for 93 RILs 2446.2 2033.0 3339.4 Medium resistant group for 93 RILs 2217.6 1827.4 3312.7 High resistant group for 93 RILs 2271.5 1900.1 3245.0 Figure 3.6 shows Y(NPQ) regulated quenching capacity, Y(NO) non-regulated quenching capacity, Y(II) effective photochemical yield values, and the electron transfer rate for HS and MHR plants. Y(NPQ) values were not significantly different among A. fatua lines, while Y(NO) values were lower in both R4 and R3 plants compared to the other lines. Y(II) values were significantly higher in R3 and R4 plants than in other lines. Y(II) values in 20R2 were lower than in S1, R3, and R4 plants. ETR values of R3 and R4 were significantly higher than the other lines and that 20R2 ETR value is significantly less than S1, R3, and R4. In total, these PAM values indicate that R3 and R4 utilize light energy more for photochemical reactions, and release excess energy through non-regulated ways significantly less. 71 Figure 3.6 – Field derived lines measuring a. Y(NO) – Non-regulated energy dissipation, b. Y(NPQ) - Regulated energy dissipation, c. Y(II) – Effective Photochemical yield, and d. ETR - Electron transfer rate. Compact letter design is employed to display significant differences. A single significant difference existed between the 93 RILs and their progenitor lines (Figure 3.7). The medium resistant group (25 - 75%) has a higher Y(NPQ) value compared to S2 and the low (<24%) resistant group. This indicates that the medium resistant group has an ability to quench excess energy through regulated mechanisms. 72 Figure 3.7 – 93 RIL groups and progenitor lines measuring a. Y(NO) – Non-regulated energy dissipation, b. Y(NPQ) - Regulated energy dissipation, c. Y(II) – Effective photochemical yield, and d. ETR - Electron transfer rate. Compact letter design is employed to display significant differences. Several significant differences were found for the 06 RIL groups and their progenitor lines (Figure 3.8). Y(NO) values showed that S2 was significantly higher than R4 and the RIL groups. S2 was also found to be significantly lower for Y(II) and ETR R4 and RIL groups. Y(NPQ) values showed that R4 was significantly higher than S2 and RIL groups. S2 was significantly lower than R4 and the high RIL resistant group for Y(NPQ). This indicates that R4 and the high RIL resistant group utilize more energy for photochemical processes and releases energy through regulated methods. This also shows that S2 consistently utilizes energy at a lower rate and releases energy through non-regulated mechanisms. 73 Figure 3.7 – 06 RIL groups and progenitor lines measuring a. Y(NO) – Non-regulated energy dissipation, b. Y(NPQ) - Regulated energy dissipation, c. Y(II) – Effective Photochemical yield, and d. ETR - Electron transfer rate. Compact letter design is employed to display significant differences. Discussion This study aimed to explore plant pigment and fluorometry differences between HS and MHR, A. fatua lines. Thin layer chromatography and spectrophotometer assays was used to compare pigment levels. Avena fatua MHR and HS field derived lines were used along with RIL groups based on their resistant phenotype and were measurements to compare photosynthetic parameters. Thin layer chromatography showed preliminary evidence that pheophytin, a chlorophyll pigment that functions as a primary electron acceptor, was present in the two S2 plants grown under high and low light (Hou, 2014). However, the pheophytin band was not as pronounced in the R4 plant grown under high light. This experiment was replicated and found the opposite, 74 showing that R4 grown in greenhouse conditions had a less pronounced pheophytin band. While this difference is subtle, it may indicate that slight environmental conditions may influence the pheophytin band. Cloudy days with less light energy being absorbed by the plant could lead to a reduction of this pigment, for plants grown in the greenhouse. Or it could be minor energy fluctuations in the growth chamber that altered the pigments bands. It has been documented that plants can control the synthase of plant pigments based on the light conditions they receive (Carvalho et al., 2011; Stewart et al., 2021). For example, Lemna gibba (duckweed) showed rapid changes in chlorophyll after exposure to either high or low light. However, because of these initial findings, further investigation was conducted on the detection of plant pigments using spectrophotometry. Spectrophotometry assays showed significant differences for acetone soluble pigments at 480 and 652 nm. 20R2 had a significant increase in concentration from 93R. Beta-carotene has an absorption maximum near 480 nm, indicating a higher concentration of beta-carotene in 20R2. Increase in beta-carotene can lead to enhanced photoprotection, enhanced photosynthetic activity, and higher stress tolerance to abiotic factors (Chen, 2015; Maoka, 2020). Interestingly, this pigment is not elevated in 93R plants, possibly indicating a lower abiotic stress tolerance. The 93R line is resistant to two herbicides, and 20R2 is resistant to nine herbicides. This possibly indicates that 20R2 could have evolved with high synthesis capabilities of beta-carotene to mitigate herbicide stress. Researchers have found that by overexpressing beta-carotene hydroxylase in Arabidopsis led to enhanced abiotic stress tolerance (Davison et al., 2002). Hydroxylase is the gene responsible for beta-carotene production. However, more research must be conducted to see if this gene is overexpressed in 20R2 lines. 75 Additionally, R4 was in greater concentration at 652 nm than 93R. Chlorophyll b has an absorption peak at this wavelength (Lichtenthaler, 1987; Wellburn, 1994). This indicates that R4 lines have higher chlorophyll b concentrations than 93R or HS lines. This difference could result in better light harvesting and more effective energy transfer under suboptimal conditions (Chen, 2015; Voitsekhovskaja & Tyutereva, 2015). Additionally, the PAM assays revealed that 93R underperformed for effective photochemical yield (YII) and electron transport rate and R4 performed higher in the same metrics. These results suggest that 93R has less chlorophyll b and is underutilizing light energy for photochemical processes, and that R4 is utilizing more light energy to drive photochemical processes. PAM fluorometry revealed 3 significant differences between MHR and HS lines in Y(NPQ), Y(NO), Y(II), and ETR parameters. The Y(NO) measurement was significantly lower in the R4 and R3 lines, showing that these two lines don’t utilize non-regulated quenching mechanism as much as the other lines. Y(II) and ETR measurements showed that R4 and R3 were significantly higher than the other field derived lines. This indicates that R4 and R3 also have a greater extent of utilizing captured energy and that R4 and R3 are better suited to capitalize on absorbed light and quenching excess energy through a regulated process. The other MHR lines 93R and 20R2 performed equivalently to the HS A. fatua lines. These non-regulated quenching mechanisms refer to the triplet-triplet energy release. As mentioned above, this energy release is usually non-destructive and is controlled by antioxidants present in cells. However, if the light stress becomes too high, or additional stresses arise, excess ROS molecules can be generated and cause damage to the cells (Das & Roychoudhury, 2014; Hashimoto et al., 2018). This shows that 93R, 20R2, S1, and S2 rely more on non-regulated quenching mechanisms and 76 could experience higher photooxidative stress in strong light conditions. This could also mean that these lines rely more heavily on other antioxidants to suppress ROS generation during high stress situations. The 93 RIL population and its progenitor lines showed minimal significant differences. The regulated quenching proportion, Y(NPQ), for S2 and the low resistance group was significantly lower than the medium resistance group. However, no significant differences were observed in non-regulated quenching, effective photochemical yield proportions, or electron transport rate. Given that the parental lines 93R and S2 did not differ significantly in any of these measures, the consistency observed among the 93 RIL and its progenitor lines was expected. These results indicate that S2 and the medium resistant group don’t utilize or do not have enough xanthophyll molecules to mitigate this energy release, instead relying more heavily on triplet- triplet energy transfer. The 06 RIL population had significant differences in that all showed S2 being different than the R4 and RIL resistant groups, which was observed from the PAM results from the field derived lines. The S2 line showed a notably higher non-regulated quenching proportion, Y(NO), than the other lines and was significantly less for Y(NPQ), YII, and ETR. For regulated quenching proportion Y(NPQ), R4 was significantly increased than the other progenitor line and resistance groups. The high resistance group was significantly different from S2, and R4 was an intermediate between the two lines, which means that R4 consistently has the highest Y(NPQ) value compared to the other RIL groups and S2. However, it is important to note that this fluorometry data is not co-segregating with resistance. If it was, the data would have shown the 77 low resistant groups and the high resistant groups to have similar data spreads to its parental lines. The results indicate that these resistant groups are an intermediary from both parental lines. Using the 06 and 93 RIL populations, the goal was to see if these traits co-segregated with herbicide resistance. (Broman, 2005). However, since most RIL resistance group measurements were insignificant and most RIL resistance groups did not differ amongst themselves, this outcome is unlikely. Future work should incorporate two procedures: plant pigment detection using high-performance liquid chromatography coupled with absorbance detection and mass spectrometry to identify pigments definitively (Naharwal et al., 2023). And another PAM assay utilizing the photosynthesis inhibitor paraquat to see if MHR lines have an enhanced capability to quench excess energy through regulated or non-regulated mechanism, as well as examining if the paraquat stress significantly impacts their photochemical yield and electron transport rate. Conclusion This study involved the examination of plant pigment and fluorometry measurements among different A. fatua lines under controlled conditions. Initial observations from growth chamber experiments led to a deeper investigation using thin layer chromatography, revealing a pheophytin band difference between the lines. This initiated spectrophotometric assays to identify these differences, finding differences for 93R, 20R2 in beta-carotene concentrations and differences between 93R and R4 chlorophyll b concentrations. Through pulse amplitude modulated fluorometry, R3 and R4 lines showed increased capabilities in managing absorbed light energy, potentially hinting at their enhanced resistance against herbicide stress. The 06 RIL population demonstrated an intermediator of the progenitor lines, while the 93 RIL population 78 mirrored its progenitor lines. Although this study provides valuable insights into the photosynthetic and light-management capabilities of A. fatua lines, the direct linkage between these traits and non-target site resistance mechanisms remains to be definitively established. Future research, potentially leveraging advanced techniques like high-performance liquid chromatography, and fluorometry measurements with photosystem inhibitors could offer clearer insights for potential NTSR mechanism. 79 CHAPTER FIVE QUANTITATIVE TRAIT LOCI ANALYSIS OF MULTIPLE HERBICIDE RESISTANT AVENA FATUA Summary Quantitative trait Loci (QTL) discovery and mapping were employed to delve into the genetic basis of multiple herbicide resistance in Avena fatua. Two recombinant inbred line (RIL) populations, 93 RIL(93R x S2), and 06 RIL (R4 x S2) were used developed for this study. Three QTLs were associated with triallate resistance in the 93 RIL population, and five QTLs were found to associate with flucarbazone and triallate in the 06 RIL population. Present in the 06 RIL population, a major QTL 3D_427541473 (LOD of 41.21) was identified as associated with triallate resistance, while QTL 4A_358747067 (LOD of 4.79) emerged as a significant candidate for flucarbazone resistance. Significant QTLs were not identified for difenzoquat resistance in either population. Identifying QTLs associated with triallate and flucarbazone resistance provides valuable preliminary genetic markers for future fine mapping and functional validation studies. In related research, differentially expressed genes (DEGs) from previous transcriptome assays were mapped with the QTLs reported here, providing potential candidate genes for MHR. This research advances the genetic characterization of A. fatua MHR lines, with the long-term goal of identifying specific herbicide resistance mechanisms. 80 Introduction The issue of identifying herbicide resistant mechanisms is a complex and troublesome problem worldwide. As the global reliance on herbicides continues, so does the selection pressure, leading to a rapid increase in cases of herbicide resistance. Since 1950, 521 unique cases of herbicide-resistant weeds (species x site of action) have been discovered, or a rate of 16 unique cases reported annually (Heap, 2014). Given this alarming rate and the widespread implications of MHR for global agricultural production and food security, understanding the genetic mechanisms and causal genes that confer MHR is of utmost importance. Our species of study, Avena fatua (wild oat), has developed resistance to multiple herbicides through nontarget site resistance (NTSR) mechanisms, leading to numerous studies to identify specific mechanisms responsible for the MHR phenotype. In 2002, inheritance studies of reciprocal crosses of A. fatua lines 93R with S2 found that triallate resistance segregated in a 15:1 S:R ratio, indicating that two recessive genes confer triallate resistance (Kern et al., 2002). Additionally, 93R’s resistance to triallate is conferred by reduced rates of sulfoxidase activity, a pre-herbicide activation step required for toxicity (Kern et al., 1996). In 2015, DNA sequencing of 06R line R4 found no known target site mutations that confer resistance to acetolactate synthase (ALS; HRAC Group 2) or acetyl coA carboxylase (ACCase; Group 1) inhibitors, confirming the involvement of NTSR mechanisms (Keith et al., 2015). In 2017, R4 was shown to exhibit constitutively elevated transcript levels of differentially expressed genes that had roles in xenobiotic catabolism, stress response, redox maintenance, and transcriptional regulators, as compared to S1 levels (Keith et al., 2017). In 2018, additional inheritance studies showed that 81 three closely linked nuclear genes control the resistance of R3 and R4 lines to flucarbazone, imazmethabenz, and pinoxaden (Burns et al., 2018). QTL analysis is a statistical method that links phenotypic and genotypic data to specific regions of chromosomes. This tool attempts to explain the genetic basis of variation in complex traits (Farquhar, 1998). QTL positions can be found using a linkage map of the RIL populations and phenotypic data for each line in those populations. Accompanying this position is the verification estimate known as the logarithm of the odds ratio (LOD). LOD scores estimate the relative strength of evidence for the presence of a QTL at a particular location (Van Ooijen, 1999). QTL analysis measures intervals between adjacent pairs of markers along each chromosome and gives a corresponding LOD score. Using these methods, accurate and reliable QTL markers associated with potential candidate genes that confer NTSR mechanisms in A. fauta can be found. QTL analyses have helped researchers investigate NTSR mechanisms for single herbicides very recently in Eleucine indica (goosegrass), Sorghum bicolor (grain sorghum), Alopercurus myosuroides (blackgrass) and Phaseolus vulgaris (snap bean) (Cai et al., 2023; Cha et al., 2014; Dakota et al., 2022; Pandian et al., 2022; Pandian et al., 2020). In this work, two RIL populations segregating for MHR were genotyped and used along with associated phenotypic data to find genetic markers by a QTL analysis. With the end goal of discovering the causal genes for the herbicide resistance traits in A. fatua lines. 82 Methods Plant Material and Experimental Design RILs were used described in Chapter 4: Plant Pigment Detection, Photosynthetic and Energy Management Parameters of Multiple Herbicide Resistant Avena fatua. All plants were grown in the greenhouse conditions described in Chapter 2: Insect Generalist Preference to MHR Avena fatua. Phenotypic Data Seeds of the 93R and S2 parental lines and derived 93 RILs were treated with pre- emergence applications of 513 g a.i. ha-1, 1,026 g a.i. ha-1 or 2,052 g a.i. ha-1 doses of triallate. Triallate was applied to 1,33.5 cm3 of moistened soil with a moving nozzle sprayer as described in Chapter Three: Exogenous Volatile Organic Compounds Effects on Herbicide Resistance in Two Avena fatua Lines, mixed thoroughly, and treated soil spread evenly over seeds placed on top of 5 cm of untreated greenhouse soil in 15.24 cm pots. Pots were lightly watered and fertilized as above, and after 21 days, percent germination (of 8 seeds planted), shoot height, and a visual herbicide injury rating (0 = no injury to 5 = complete death) were recorded. These three parameters were combined by averaging the scores to create an overall resistance score, which was then used to rank RILs. This resulted in 12 individual phenotypic measurements to use for QTL mapping. 06R and S2 parental and derived 06 RIL lines were grown in the greenhouse conditions described above and treated with individual herbicides when they reached Zadok’s stage 13. The 06R lines were treated with flucarbazone (Everest™) at 19.7 g a.i. ha-1 or 39.4 g a.i. ha-1 doses or triallate (FAR-GO™) at 513 g a.i. ha-1, 1,026 g a.i. ha-1 or 2,052 g a.i. ha-1 doses, difenzoquat 83 (Avenge™) at 712.05 g a.i. ha-1, or pinoxaden (Axial™) at 17.5 g a.i. ha-1. Three weeks after treatment, visual injury scores were recorded, plant shoots were harvested at the soil surface, tissues were dried at 40°C for two weeks, and dry weights (DW) were recorded. Phenotypic data from the 06 lines were expressed as the percent of either an untreated progenitor line or untreated RIL control grown simultaneously with the treated RIL line. All treatments were applied to 3 or 4 biological replicates. For 06 lines, a basic injury rating system ranging from 1 to 5 was employed (0 = no injury to 5 = complete death). This resulted in 17 phenotypic measurements to use for QTL mapping. Additionally, 06R and S2 parental and derived 06 RIL lines were phenotyped for triallate as described above. Genotyping and Linkage Map Construction Reference genome alignment and linkage map construction were completed following the procedures described in (Fiedler et al., 2015). Briefly, genomic DNA was extracted from lyophilized young leaves of parental and RIL population plants by Novogene Corp. Inc. (Sacramento, CA). Genotyping was conducted through Genotype-by-Sequencing (GBS) techniques (Deschamps et al., 2012) on restriction enzyme fragmented DNA, and resulting sequences were aligned to the Avena sativa reference genome OT3098v2 (Peng et al., 2022). This process allows single nucleotide polymorphisms to be identified, which are single nucleotide changes at a specific position in the genome. SNP identification and calling used the criteria of 1) a minimum of four reads for each SNP call, 2) SNPs must be present in 50% of the relevant RIL population, and 3) have a minor allele frequency of at least 0.01. The filtered SNPs were then used for linkage mapping using Joinmap (Ooijen, 2018). Due to the limitations of 84 GBS techniques and the aggressive filtering of non-ideal and redundant SNPs, a relatively small number of SNPs were mapped for each population. QTL Analysis Standard and multiple interval mapping (SIM and MIM, respectively) were conducted using the maximum likelihood method as described by (Broman, 2005), using the R-studio packages qtl and ASMap (Taylor & Butler, 2017). Standard interval mapping was first used to focus on a single QTL at each locus and calculate LOD scores for each marker location. Permutation tests for QTL significance were simultaneously conducted with 1,000 replications, providing a statistical foundation for LOD score to determine QTL significance (Doerge & Rebaï, 1996; Kao et al., 1999). For this phase, Haley-Knott (HK) regression (Haley & Knott, 1992) was specifically applied as the regression-based method for QTL mapping. Briefly, HK linearly regresses phenotypic values of individuals against marker probabilities, which are calculated based on known marker genotypes and their respective recombination fractions (Haley & Knott, 1992). After completion of standard interval mapping, multiple interval mapping (Kao et al., 1999) was used to fit multiple QTLs into a model simultaneously. Then a stepwise QTL function (Kao et al., 1999) was employed, allowing for the addition and removal of QTLs based on their statistical significance. For 06 RILs, a LOD score > 3.0 and a significance level set at P < 0.05 were utilized for QTL detection. For 93 RILs, weaker phenotypic data required that LOD scores > 2.5 were considered significant. RNA-seq Analysis In this study, an RNA-seq analysis on the Avena fatua transcriptome (Keith et al., 2017) was reanalyzed utilizing the Avena sativa reference genome OT3098v2 (Peng et al., 2022)to 85 align sequence reads. Quality control of the sequencing reads was conducted using BBmap version 38.98 (Bushnell, 2014). Reads were aligned using Spliced Transcripts Alignment to a Reference (STAR) software version 2.7.10a (Dobin et al., 2013). This approach yielded an average mapping rate of approximately 75%, a significant improvement over the original 30% mapping rate achieved with the Oryza sativa reference genome. Post-alignment filtering using FeatureCounts ver. 2.0.3 (Liao et al., 2014) further refined the dataset, resulting in approximately 60% of the reads remaining aligned. Differential gene expression analysis was carried out using DESeq2 version 1.36.0 (Love et al., 2014), and differentially expressed genes (DEGs) with a log 2-fold change > 0.85 were mapped to identified QTLs. Gene annotations were based on E-values ≤ 10−4 Results Genotyping and Linkage Map Construction Genotyping resulted in 1,504 mapped SNPs in 24 linkage groups in the 93 RIL population and the 06 RILs had 1,834 mapped SNPs in 30 linkage groups. Linkage groups for the 06 RIL population are shown in Figure 4.1. Once all SNP markers were mapped, the allele frequency within each RIL population was assessed. Given that these RIL populations originated from a single cross between two parental lines and were advanced to the sixth generation through single seed descent, it would be expected that the proportion of each parental allele would approximate 50% (Hill et al., 2008). However, the frequency at individual loci could deviate from this expectation due to factors such as random segregation and recombination. Consistent with this, the 06 and 93 RILs displayed an allele frequency close to 50% for each parent. 86 Figure 4.1 – Linkage groups for the 06 RIL population. Black bands represent a SNP. QTL Analysis Three QTLs for triallate resistance were identified in the 93 RIL population (Table 4.1). First, 3D_427541473 was detected on chromosome 3D at 19.6 cM and was detected for seven different phenotypic measurements with LOD scores ranging from 3.27 to 5.56. Second, QTL 3D_347553285 was detected on chromosome 3D at 13.4 cM for one phenotypic measurement with a LOD score of 8.01. Third, QTL 1A_9291409 was detected at chromosome 1A at 1.2 cM for one phenotypic measurement with a LOD score of 2.66. 87 Table 4.1. Complete list of QTLs found in the 93 RIL population using MIM. Tri suffix represents Triallate (FAR-GO). All QTLs were detected using an alpha of 0.05, and LOD score above 2.5 for significant threshold. Position QTL ID Trait Chromosome LOD Flanking Markers (cM) Visual rating 513 g 4.71 a.i ha-1 Tri Size rating 1,026 g -1 3.27 a.i ha Tri Size scale 2,052 g a.i -1 4.50 ha Tri Score 2,052 g a.i ha-1 5.56 3D_427541473 - 3D_427541473 Tri 3D 18.6 3D_433601457 Total Score 2,052 g -1 3.54 a.i ha Tri Rank Overall 5.38 Phenotype of visual rating 1,026 g a.i ha-1 5.14 Tri Germination scale 3D_347553285 - 3D_347553285 3D 13.4 8.01 1,026 g a.i ha-1 Tri 3D_346231075 Ranking based on 1A_9291409 - 1A_9291409 injury 1,026 g a.i ha- 1A 1.2 2.66 1A_251522473 1 Tri Five distinct QTLs associated with resistance to triallate or flucarbazone were identified in the 06 RIL population (Table 4.2). These QTLs were localized to five different chromosomes: 4A, 1D, 2D2, 3D2, and 7A. The most prominent QTL for flucarbazone resistance, designated as 4A_358747067, was located on chromosome 4A at 15.8 cM. This QTL was identified across seven flucarbazone phenotype measures, displaying LOD scores that ranged from 3.15 to 4.76. Two additional QTLs implicated in flucarbazone resistance were identified as 1D_16113776, located on chromosome 1D at 0 cM, was found in three different phenotypic measures with LOD 88 scores between 3.20 and 3.3; and 2D_410716933, located on chromosome 2D2 at 4.2 cM, was found in four phenotypic measurements with LOD scores ranging from 3.51 to 3.85. For triallate resistance, a major QTL labeled 3D_1907891 was located on chromosome 3D at 22.0 cM. It was identified in four phenotypic assays for triallate resistance, exhibiting LOD scores that ranged from 26.67 to 41.21, the highest among all identified QTLs. Another QTL, 7A_66986043, was found on chromosome 7A at 0 cM and was identified in five triallate phenotypic assays with LOD scores ranging from 3.40 to 4.91. Notably, no QTLs were identified for resistance to either pinoxaden or difenzoquat. Moreover, no shared QTLs were found between flucarbazone and triallate resistance measures. 89 Table 4.2. Complete list of QTLs found in the 06 RIL population using MIM. The Flu suffix represents flucarbazone (Everest) Tri suffix represents Triallate (FAR-GO). All QTLs are above the alpha = 0.05, and LOD score above 3 for significant threshold. Peak QTL Flanking Trait Chromosome Position LOD Identification Markers (cM) Percent of Control 19.7 3.95 g a.i ha-1 Flu Percent of R4 19.7 g -1 4.49 a.i ha Flu Percent of S2 19.7 g a.i 4.76 ha-1 Flu 4A_358747067 Visual Score 19.7 g a.i 4A_358747067 -1 4A 15.8 4.38 - ha Flu 4A_358747067 Percent of R4 39.4 g -1 3.41 a.i. ha Flu Percent of S2 39.4 g a.i. ha-1 3.15 Flu Visual Score 39.4 g a.i. -1 4.52 ha Flu Percent of R4 19.7 g 3.20 a.i. ha-1 Flu Percent of R4 39.4 g 1D_16113776 - 1D_16113776 1D 0 3.28 a.i. ha-1 Flu 1D_16113776 Visual score 39.4 g a.i. ha-1 3.3 Flu Percent of S2 19.7 g 3.85 a.i. ha-1 Flu Percent of R4 39.4 g 3.51 a.i. ha-1 Flu 2D_234378723 2D_410716933 2D2 4.2 - Percent of S2 39.4 g 2D_410716933 3.66 a.i. ha-1 Flu Visual Score of 39.4 g 3.66 a.i. ha-1 Flu 90 Table 4.2 Continued. Peak QTL Flanking Trait Chromosome Position LOD Identification markers (cM) Percent of Control 41.21 1,026 g a.i ha-1 Tri Percent of R4 1,026 g 41.21 a.i ha-1 Tri 3D_1907899 - 3D_1907891 3D2 22.0 Percent of Control 3D_6229596 26.67 2,052 g a.i ha-1 Tri Percent of R4 2,052 g 26.67 a.i. ha-1 Tri Percent of Control 3.40 1,026 g a.i. ha-1 Tri Percent of R4 1,026 g 3.40 a.i. ha-1 Tri 7A_6698604 Percent of Control 7A_66986043 - 7A 0 4.91 3 2,052 g a.i. ha-1 Tri 7A_66986043 Percent of R4 2,052 g 4.91 a.i. ha-1 Tri Visual Score of 2,052 g 4.47 a.i. ha-1 Tri RNA-Seq Analysis RNA-seq analysis identified 13 DEGs located within QTL peaks that met predefined selection criteria (Figure 4.3). Four of these were significantly overexpressed in the R4 lines, while the remaining nine were attenuated in R4 compared to S2. Notably, all four genes showing elevated expression in R4 were localized to the 3D_1907891 major QTL for triallate resistance. Specifically, the genes overexpressed in R4 include 73C10 and U73C6, both of which are annotated as UDP-glycosyl-transferases, and LEUNIG, a transcriptional co-repressor, and NPL4, characterized as an NPL4-like protein. In contrast, within the same major-effect QTL (3D_1907891), three other genes—BOR4, FB345, and TBL1—were attenuated in R4. BOR4 is 91 identified as a boron transporter, FB345 as an F-box protein, and TBL1 as a trichome birefringence-like 1 protein. Under the 2D_410716933QTL for flucarbazone resistance, six genes displayed lower expression levels in R4 than in S2 lines. These genes were annotated as U71K2, a UDP- glycosyl-transferase; CAR8, possessing a protein C2 domain; RA51D, a DNA repair protein; FBT8, a likely folate-biopterin transporter; PIP21, an aquaporin protein; and HFA2B, a heat stress transcription factor. Table 4.3. DEGs underlying QTL peaks in R4 lines. Alias is the identifier for genes. Log 2-fold change values > 0.8 and P < 0.5 represent the magnitude and direction of gene expression, where positive and negative numbers represent under- and over-expression in R4 compared to S2. Log 2- QTL DEG fold Identification Alias change P-Value Annotation 73C10 -1.6974 0.0001456 UDP-glycosyl-transferase U73C6 -1.7595 0.000000024 UDP-glycosyl-transferase LEUNIG -2.0964 0.00000025 Transcriptional co-repressor 3D_1907891 NPL4 -1.0317 0.0000042 NPL4-Like protein BOR4 3.1637 0.00000000000019 Boron transporter 4 FB345 5.6962 0.00000000000023 F-box protein Protein trichome TBL1 8.4688 0.000000000004 birefringence U71K2 0.9975 0.00044746 UDP-glycosyl-transferase Protein C2-Domain ABA- CAR8 0.8970 0.00000123 related 2D_41071693 RA51D 1.8205 0.0000000001 DNA repair Protein 3 FBT8 0.8333 0.00050535 Folate-biopterin transporter PIP21 2.0713 0.0000127 Aquaporin protein Heat stress transcription HFA2B 0.8537 0.00021586 factor 92 Discussion The primary aim of this study was to employ QTL mapping techniques to identify genomic regions associated with herbicide resistance across three RIL populations. QTL mapping is a widely recognized approach for dissecting complex traits and pinpointing genomic regions that influence quantitative attributes in plant species (Farquhar, 1998; Myles & Wayne, 2008; Würschum, 2012). In the 93 RIL population, three QTLs were identified. 3D_427541473, with LOD scores ranging from 3.27 to 5.56. 3D_347553285 and 1A_9291409 with LOD scores of 8.01 and 2.66 respectively. 3D_427541473 was detected multiple times using different phenotypic parameters in the QTL analysis, strongly indicating that this region houses a gene conferring triallate resistance. The other two QTLs were only detected once, using the scoring system phenotypic parameter which suggests that these may not be strong genomic markers for herbicide resistance. Crucially, the resistant trait for triallate in 93R segregates to a 15:1 (sensitive to resistant) ratio. This lead the RIL population to have 7 out of 124 fully resistant lines, with phenotypic values being concentrated at low end of each phenotypic metric. Without a range of phenotypic values, QTL analysis becomes extremely difficult (Consortium, 2003). For the 06 RIL population, QTL analysis identified three regions associated with flucarbazone resistance: 4A_358747067, 1D_16113776, and 2D_410716933. Among these, 4A_358747067 was the most recurrent, appearing in seven different phenotypic measurements and displaying a higher LOD score than the other flucarbazone-associated QTLs. For the other flucarbazone resistance trait in the 06 RIL population, two additional QTLs, 1D_16113776 and 2D_410716933, exhibited lower LOD scores compared to 4A_358747067. Specifically, 1D_16113776 had LOD scores ranging from 3.20 to 3.30, while 2D_410716933 registered 93 scores between 3.51 and 3.85. Intriguingly, the latter QTL had flanking markers within which several DEGs were identified. As two recessive genes control triallate resistance in 93R (Kern et al., 2002), and the mechanism responsible for triallate resistant in R4 normally segregates. This means that although QTL are being mapped onto 3D chromosome, they are not indicating the same mechanism are responsible for this phenotype. The lack of strong LOD scores could be caused by the two recessive genes required for triallate resistance. All the DEGs meeting our criteria were downregulated in the R4 population and included U71K2, a UDP-glucosyl- transferase implicated in xenobiotic resistance, and CAR8, a C2-Domain Protein involved in ABA receptor recruitment. C2-domain proteins have significant roles in signal transduction and membrane transport (Edel & Kudla, 2016). RA51D is a DNA repair protein, and PIP21, which was identified as an aquaporin, both being under expressed would could have a role for herbicide resistance. FBT8 is a folate-biopterin transporter that could grant some herbicide resistant capabilities, but only to asulam (DHPS inhibitor, Group 8), an herbicide that targets folate biosynthesis, which cannot control monocots. Notably, HFA2B, a heat stress transcription factor, was detected to be under-expressed in R4, which is similar to the findings presented in (Keith et al., 2017). In relation to triallate resistance in the 06 RIL population, two QTLs were detected: 3D_1907891 and 7A_66986043. The latter appeared five times across various measurements with LOD scores between 3.40 and 4.91, while 3D_1907891 was found to have a major effect QTL, evident by its high LOD scores ranging from 26.67 to 41.21. Subsequent DEG analysis under this major QTL peak revealed that four genes, 73C10, U73C6, LEUNIG, and NPL4, were 94 expressed at higher levels in R4 than in S2 lines. Conversely, three genes, namely BOR4, FB245, and TBL1, were upregulated in S2 lines. Of the DEGs overexpressed in R4, 73C10 and U73C6 were categorized as UDP- glycosyl-transferases, enzymes previously noted for their role in xenobiotic resistance (Keith et al., 2017). These enzymes are part of a broader superfamily, the uridine diphosphate sugar- utilizing glycosyl-transferases, involved in various metabolic pathways, including xenobiotic resistance (Dimunová et al., 2022; Meech et al., 2012). Their presence under a major-effect QTL highlights their possible role in herbicide resistance mechanisms. LEUNIG, a transcriptional co- repressor, and its homologs have been reported to regulate both abiotic and biotic stress responses in plants (Gonzalez et al., 2007; Sitaraman et al., 2008). Nuclear Protein Localization 4 (NPL4) operates as a heterodimeric cofactor within the ubiquitin-proteasome pathway, a critical catabolic mechanism that targets unwanted or damaged proteins for degradation (Bays & Hampton, 2002). This pathway has been implicated in herbicide treatment responses; for instance, glyphosate treated Pisum sativa (Common pea) exhibited elevated levels of ubiquitinated proteins alongside heightened proteasome activity (Zulet et al., 2013). The ubiquitin-proteasome system has also modulated the effects of various abiotic stresses on plant growth and development (Mackinnon & Stone, 2022). In the R4 lines, DEGs with lower expression included BOR4, a Boron transporter; FB345, an F-box protein; and TBL1, a protein involved in trichome birefringence. FB345 is part of the SCF (Skip1, Cullin, F-box) complex, which facilitates the ubiquitination of target proteins, marking them for degradation by the 26S proteasome (Jackson & Eldridge, 2002). TBL1 is a 95 binding protein that interacts with pectin and other cell wall polysaccharides, potentially affecting defense mechanisms (Bischoff et al., 2010). The attenuated expression of FB345 in R4 is intriguing, especially when considering its functional similarities with NPL4, a gene expressed at higher levels in R4 lines. Both are involved in proteasome-mediated degradation; however, the SCF complex and the ubiquitin- proteasome pathway in which NPL4 is complex, housing many different genes (Roos‐Mattjus & Sistonen, 2004). The upregulation of genes involved in this pathway but found in different lines suggests that proteasome activity may be modulated context-dependent. It is also plausible that these pathways intersect or are involved in different parts of herbicide resistance or stress response. While this raises questions about the specific roles of these proteins and their interaction in herbicide resistance, it also opens avenues for future research. For instance, a thorough functional validation could find whether the down regulation of FB345 in R4 lines indeed affects proteasome activity in a manner like or different from NPL4. The genes that have been found in previous work for A. fatua should also be examined as their occurrence, highlighted by this QTL analysis, further suggests a possible role in herbicide resistance (Keith et al., 2017). Given that the GBS data for this study was aligned to the Avena sativa reference genome, it would be warranted to re-align the GBS data once the Avena fatua genome becomes available. Subsequent QTL analyses could then be conducted to compare QTL regions of interest. Additionally, advanced methodologies such as fine mapping could be employed to refine our understanding of the identified regions. Fine mapping is a targeted genetic technique designed to narrow down the precise location of a SNP as a marker associated with a specific trait. It is 96 especially useful when studying heterozygous individuals who possess different alleles at the locus of interest (Noroozi & Sattari, 2015). The fine mapping approach can pinpoint the gene or QTL more accurately than broader mapping methods. Conclusion In conclusion, this study employed QTL mapping to investigate the genomic regions associated with herbicide resistance across different RIL populations. Highlighted outcomes were the identification of specific QTLs, such as 3D_427541473 for triallate resistance and 4A_358747067 for flucarbazone resistance for 06 RILs. These QTLs stand as genomic regions likely to harbor genes conferring herbicide resistance, thereby serving as significant markers for future genetic studies. Intriguingly, the DEG analysis within these QTL regions provided insights into the complex biochemical pathways potentially involved in herbicide resistance. For instance, UDP-glycosyl-transferases were found to be overexpressed in R4 lines and were situated under a major-effect QTL, highlighting their likely importance in herbicide resistance mechanisms. The study also noted upregulation of genes related to proteasome-mediated degradation pathways in both R4 and S2 lines. The following steps should involve fine mapping techniques to narrow these identified QTL regions. This would be a logical extension to provide a more accurate pinpointing of the genes responsible for herbicide resistance. Cross-referencing these findings with the soon-to-be-released Avena fatua genome could yield even more precise results. 97 REFERENCES CITED Ache, B. W., & Young, J. M. (2005). Olfaction: Diverse Species, Conserved Principles. Neuron, 48(3), 417-430. https://doi.org/10.1016/j.neuron.2005.10.022 Adamczewski, K., Kaczmarek, S., Kierzek, R., & Matysiak, K. (2019). Significant increase of weed resistance to herbicides in Poland. Journal of Plant Protection Research, 59(2). Ahmad, P., Jaleel, C. A., Salem, M. A., Nabi, G., & Sharma, S. (2010). Roles of enzymatic and nonenzymatic antioxidants in plants during abiotic stress. Critical reviews in biotechnology, 30(3), 161-175. Ameye, M., Allmann, S., Verwaeren, J., Smagghe, G., Haesaert, G., Schuurink, R. C., & Audenaert, K. (2018). Green leaf volatile production by plants: a meta-analysis. New Phytologist, 220(3), 666-683. https://doi.org/10.1111/nph.14671 Amorim-Carrilho, K., Cepeda, A., Fente, C., & Regal, P. (2014). Review of methods for analysis of carotenoids. TrAC Trends in Analytical Chemistry, 56, 49-73. Anderson, P., & Alborn, H. (1999). Effects on oviposition behaviour and larval development of Spodoptera littoralis by herbivore-induced changes in cotton plants. Entomologia Experimentalis et Applicata, 92(1), 45-51. https://doi.org/10.1046/j.1570- 7458.1999.00523.x Arsovski, A. A., Galstyan, A., Guseman, J. M., & Nemhauser, J. L. (2012). Photomorphogenesis. The Arabidopsis Book, 10, e0147. https://doi.org/10.1199/tab.0147 Baker, N. R. (2008). Chlorophyll fluorescence: a probe of photosynthesis in vivo. Annu. Rev. Plant Biol., 59, 89-113. Baldwin, I. T., Kessler, A., & Halitschke, R. (2002). Volatile signaling in plant–plant–herbivore interactions: what is real? Current opinion in plant biology, 5(4), 351-354. Batschauer, A. (1998). Photoreceptors of higher plants. Planta, 206(4), 479-492. Bays, N. W., & Hampton, R. Y. (2002). Cdc48–Ufd1–Npl4: stuck in the middle with Ub. Current Biology, 12(10), R366-R371. Beckie, H. J., Francis, A., & Hall, L. M. (2012). The Biology of Canadian Weeds. 27. Avena fatua L. (updated). Canadian Journal of Plant Science, 92(7), 1329-1357. https://doi.org/10.4141/cjps2012-005 Beckie, H. J., Warwick, S. I., & Sauder, C. A. (2012). Basis for herbicide resistance in Canadian populations of wild oat (Avena fatua). Weed Science, 60(1), 10-18. 98 Beffa, R., Menne, H., & Köcher, H. (2019). Herbicide resistance action committee (HRAC): herbicide classification, resistance evolution, survey, and resistance mitigation activities. Modern crop protection compounds, 3, 5-32. Bezerra, R. H. S., Sousa-Souto, L., Santana, A. E. G., & Ambrogi, B. G. (2021). Indirect plant defenses: volatile organic compounds and extrafloral nectar. Arthropod-Plant Interactions, 15(4), 467-489. https://doi.org/10.1007/s11829-021-09837-1 Bischoff, V., Nita, S., Neumetzler, L., Schindelasch, D., Urbain, A., Eshed, R., Persson, S., Delmer, D., & Scheible, W.-R. (2010). TRICHOME BIREFRINGENCE and its homolog AT5G01360 encode plant-specific DUF231 proteins required for cellulose biosynthesis in Arabidopsis. Plant Physiology, 153(2), 590-602. Bittles, A. H., & Neel, J. V. (1994). The costs of human inbreeding and their implications for variations at the DNA level. Nature Genetics, 8(2), 117-121. Bolsoni, V. P., Oliveira, D. P. D., Pedrosa, G. D. S., & Souza, S. R. D. (2018). Volatile organic compounds (VOC) variation in Croton floribundus (L.) Spreng. related to environmental conditions and ozone concentration in an urban forest of the city of São Paulo, São Paulo State, Brazil. Hoehnea, 45(2), 184-191. https://doi.org/10.1590/2236-8906-60/2017 Brilli, F., Ciccioli, P., Frattoni, M., Prestininzi, M., Spanedda, A. F., & Loreto, F. (2009). Constitutive and herbivore-induced monoterpenes emitted byPopulus × euroamericanaleaves are key volatiles that orientChrysomela populibeetles. Plant, Cell & Environment, 32(5), 542-552. https://doi.org/10.1111/j.1365- 3040.2009.01948.x Brilli, F., Loreto, F., & Baccelli, I. (2019). Exploiting plant volatile organic compounds (VOCs) in agriculture to improve sustainable defense strategies and productivity of crops. Frontiers in Plant Science, 10, 436279. Broman, K. W. (2005). The Genomes of Recombinant Inbred Lines. Genetics, 169(2), 1133- 1146. https://doi.org/10.1534/genetics.104.035212 Burns, Keith, B. K., Refai, M. Y., Bothner, B., & Dyer, W. E. (2018). Constitutive redox and phosphoproteome changes in multiple herbicide resistant Avena fatua L. are similar to those of systemic acquired resistance and systemic acquired acclimation. Journal of plant physiology, 220, 105-114. Burns, E. E., Keith, B. K., Talbert, L. E., & Dyer, W. E. (2018). Non‐target site resistance to flucarbazone, imazamethabenz and pinoxaden is controlled by three linked genes in Avena fatua. Weed Research, 58(1), 8-16. https://doi.org/10.1111/wre.12279 Bushnell, B. (2014). BBMap: a fast, accurate, splice-aware aligner. 99 Cai, L., Comont, D., MacGregor, D., Lowe, C., Beffa, R., Neve, P., & Saski, C. (2023). The blackgrass genome reveals patterns of non‐parallel evolution of polygenic herbicide resistance. New Phytologist, 237(5), 1891-1907. Carvalho, R. F., Takaki, M., & Azevedo, R. A. (2011). Plant pigments: the many faces of light perception. Acta physiologiae plantarum, 33(2), 241-248. https://doi.org/10.1007/s11738- 010-0533-7 Caverzan, A., Piasecki, C., Chavarria, G., Stewart, C., & Vargas, L. (2019). Defenses Against ROS in Crops and Weeds: The Effects of Interference and Herbicides. International Journal of Molecular Sciences, 20(5), 1086. https://doi.org/10.3390/ijms20051086 Cha, T. S., Anne-Marie, K., & Chuah, T. S. (2014). Identification and characterization of RAPD– SCAR markers linked to glyphosate-susceptible and-resistant biotypes of Eleusine indica (L.) Gaertn. Molecular biology reports, 41, 823-831. Chaerle, L., Leinonen, I., Jones, H. G., & Van Der Straeten, D. (2007). Monitoring and screening plant populations with combined thermal and chlorophyll fluorescence imaging. Journal of Experimental Botany, 58(4), 773-784. Chauhan, B. S. (2020). Grand Challenges in Weed Management [Specialty Grand Challenge]. Frontiers in Agronomy, 1. https://doi.org/10.3389/fagro.2019.00003 Chen, C. (2015). Overview of Plant Pigments. In (pp. 1-7). Springer New York. https://doi.org/10.1007/978-1-4939-2356-4_1 Chen, L., Wang, W.-S., Wang, T., Meng, X.-F., Chen, T.-t., Huang, X.-X., Li, Y.-j., & Hou, B.-K. (2019). Methyl Salicylate Glucosylation Regulates Plant Defense Signaling and Systemic Acquired Resistance. Plant Physiology, 180(4), 2167-2181. https://doi.org/10.1104/pp.19.00091 Choi, H., & Kim, J.-H. (2013). Volatile constituents of herbicide glufosinate-tolerant and non- transgenic rice by solvent-assisted flavor evaporation and gas chromatography-mass spectrometry. Journal of the Korean Society for Applied Biological Chemistry, 56(3), 357-360. Choudhary, K. K., Chaudhary, N., Agrawal, S., & Agrawal, M. (2017). Reactive oxygen species: generation, damage, and quenching in plants during stress. Reactive Oxygen Species in Plants: Boon Or Bane‐Revisiting the Role of ROS, 89-115. Consortium, C. T. (2003). The nature and identification of quantitative trait loci: a community's view. Nature Reviews Genetics, 4(11), 911-916. 100 Copolovici, L., Kännaste, A., Remmel, T., & Niinemets, Ü. (2014). Volatile organic compound emissions from Alnus glutinosa under interacting drought and herbivory stresses. Environmental and Experimental Botany, 100, 55-63. https://doi.org/10.1016/j.envexpbot.2013.12.011 D’Alessandro, M., Held, M., Triponez, Y., & Turlings, T. C. (2006). The role of indole and other shikimic acid derived maize volatiles in the attraction of two parasitic wasps. Journal of Chemical Ecology, 32(12), 2733-2748. Dahiya, A., Sharma, R., Sindhu, S., & Sindhu, S. S. (2019). Resource partitioning in the rhizosphere by inoculated Bacillus spp. towards growth stimulation of wheat and suppression of wild oat (Avena fatua L.) weed. Physiology and Molecular Biology of Plants, 25, 1483-1495. Dakota, B. M., Saballos, A., Soler-Garzón, A., Brooks, M., Hart, J. P., Lipka, A. E., Miklas, P., Peachey, R. E., Tranel, P. J., & Williams, M. M. (2022). Multiple Genomic Regions Govern Tolerance to Sulfentrazone in Snap Bean (Phaseolus Vulgaris L.). Das, K., & Roychoudhury, A. (2014). Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Frontiers in environmental science, 2, 53. Davison, P. A., Hunter, C. N., & Horton, P. (2002). Overexpression of β-carotene hydroxylase enhances stress tolerance in Arabidopsis. Nature, 418(6894), 203-206. https://doi.org/10.1038/nature00861 Délye, C., Jasieniuk, M., & Le Corre, V. (2013). Deciphering the evolution of herbicide resistance in weeds. Trends in Genetics, 29(11), 649-658. Demmig-Adams, B., Stewart, J. J., López-Pozo, M., Polutchko, S. K., & Adams, W. W. (2020). Zeaxanthin, a Molecule for Photoprotection in Many Different Environments. Molecules, 25(24), 5825. https://doi.org/10.3390/molecules25245825 Deschamps, S., Llaca, V., & May, G. D. (2012). Genotyping-by-Sequencing in Plants. Biology, 1(3), 460-483. https://doi.org/10.3390/biology1030460 Di, C., Ning, C., Huang, L.-Q., & Wang, C.-Z. (2017). Design of larval chemical attractants based on odorant response spectra of odorant receptors in the cotton bollworm. Insect biochemistry and molecular biology, 84, 48-62. Dimunová, D., Matoušková, P., Podlipná, R., Boušová, I., & Skálová, L. (2022). The role of UDP-glycosyltransferases in xenobioticresistance. Drug Metabolism Reviews, 54(3), 282-298. 101 Dippery, J. K., Tissue, D. T., Thomas, R. B., & Strain, B. R. (1995). Effects of low and elevated CO2 on C3 and C4 annuals. Oecologia, 101(1), 13-20. https://doi.org/10.1007/bf00328894 Dobin, A., Davis, C. A., Schlesinger, F., Drenkow, J., Zaleski, C., Jha, S., Batut, P., Chaisson, M., & Gingeras, T. R. (2013). STAR: ultrafast universal RNA-seq aligner. Bioinformatics, 29(1), 15-21. Doerge, R., & Rebaï, A. (1996). Significance thresholds for QTL interval mapping tests. Heredity, 76(5), 459-464. Du, H., Wang, N., Cui, F., Li, X., Xiao, J., & Xiong, L. (2010). Characterization of the β- carotene hydroxylase gene DSM2 conferring drought and oxidative stress resistance by increasing xanthophylls and abscisic acid synthesis in rice. Plant Physiology, 154(3), 1304-1318. Dudareva, N., Klempien, A., Muhlemann, J. K., & Kaplan, I. (2013). Biosynthesis, function and metabolic engineering of plant volatile organic compounds. New Phytologist, 198(1), 16- 32. https://doi.org/10.1111/nph.12145 Dyer, W. E. (2018). Stress-induced evolution of herbicide resistance and related pleiotropic effects. Pest Management Science, 74(8), 1759-1768. https://doi.org/10.1002/ps.5043 Edel, K. H., & Kudla, J. (2016). Integration of calcium and ABA signaling. Current opinion in plant biology, 33, 83-91. Fan, J., Francis, F., Liu, Y., Chen, J., & Cheng, D. (2011). An overview of odorant-binding protein functions in insect peripheral olfactory reception. Genet. Mol. Res, 10(4), 3056- 3069. Fang, J., Zhang, Y., Liu, T., Yan, B., Li, J., & Dong, L. (2019). Target-Site and Metabolic Resistance Mechanisms to Penoxsulam in Barnyardgrass (Echinochloa crus-galli (L.) P. Beauv). Journal of Agricultural and Food Chemistry, 67(29), 8085-8095. https://doi.org/10.1021/acs.jafc.9b01641 Farquhar. (1998). QTL analysis in plants; where are we now? Heredity, 80(2), 137-142. Fiedler, J. D., Lanzatella, C., Okada, M., Jenkins, J., Schmutz, J., & Tobias, C. M. (2015). High‐ Density Single Nucleotide Polymorphism Linkage Maps of Lowland Switchgrass using Genotyping‐by‐Sequencing. The plant genome, 8(2), plantgenome2014.2010.0065. Fitzky, A. C., Peron, A., Kaser, L., Karl, T., Graus, M., Tholen, D., Pesendorfer, M., Mahmoud, M., Sandén, H., & Rewald, B. (2021). Diversity and interrelations among the constitutive VOC emission blends of four broad-leaved tree species at seedling stage. Frontiers in Plant Science, 12, 708711. 102 Franzke, A., Unsicker, S. B., Specht, J., Koehler, G., & Weisser, W. W. (2010). Being a generalist herbivore in a diverse world: how do diets from different grasslands influence food plant selection and fitness of the grasshopper Chorthippus parallelus? Ecological Entomology, 35(2), 126-138. Fujioka, K., Gotoh, H., Noumi, T., Yoshida, A., Noutoshi, Y., Inagaki, Y., Yamamoto, M., Ichinose, Y., Shiraishi, T., & Toyoda, K. (2015). Protection induced by volatile limonene against anthracnose disease in Arabidopsis thaliana. Journal of General Plant Pathology, 81(6), 415-419. https://doi.org/10.1007/s10327-015-0621-z Gassmann, A., & Futuyma, D. (2005). Consequence of herbivory for the fitness cost of herbicide resistance: photosynthetic variation in the context of plant–herbivore interactions. Journal of Evolutionary Biology, 18(2), 447-454. Gassmann, A. J. (2005). Resistance to herbicide and susceptibility to herbivores: environmental variation in the magnitude of an ecological trade-off. Oecologia, 145(4), 575-585. Genesio, L., Bassi, R., & Miglietta, F. (2021). Plants with less chlorophyll: A global change perspective. Global Change Biology, 27(5), 959-967. https://doi.org/10.1111/gcb.15470 Gerhart, L. M., & Ward, J. K. (2010). Plant responses to low [CO2] of the past. New Phytologist, 188(3), 674-695. Gonzalez, D., Bowen, A. J., Carroll, T. S., & Conlan, R. S. (2007). The Transcription Corepressor LEUNIG Interacts with the Histone Deacetylase HDA19 and Mediator Components MED14 (SWP) and CDK8 (HEN3) To Repress Transcription. Molecular and Cellular Biology, 27(15), 5306-5315. https://doi.org/10.1128/mcb.01912-06 Götz, T., Sandmann, G., & Römer, S. (2002). Expression of a bacterial carotene hydroxylase gene (crtZ) enhances UV tolerance in tobacco. Plant Molecular Biology, 50, 127-140. Gregg, P. C., Del Socorro, A. P., & Henderson, G. S. (2010). Development of a synthetic plant volatile-based attracticide for female noctuid moths. II. Bioassays of synthetic plant volatiles as attractants for the adults of the cotton bollworm,Helicoverpa armigera(Hübner) (Lepidoptera: Noctuidae). Australian Journal of Entomology, 49(1), 21-30. https://doi.org/10.1111/j.1440-6055.2009.00734.x Hafeez, M., Ullah, F., Khan, M. M., Li, X., Zhang, Z., Shah, S., Imran, M., Assiri, M. A., Fernández-Grandon, G. M., & Desneux, N. (2021). Metabolic-based insecticide resistance mechanism and ecofriendly approaches for controlling of beet armyworm Spodoptera exigua: a review. Environmental Science and Pollution Research, 1-17. Haley, C. S., & Knott, S. A. (1992). A simple regression method for mapping quantitative trait loci in line crosses using flanking markers. Heredity, 69(4), 315-324. https://doi.org/10.1038/hdy.1992.131 103 Han, R.-M., Zhang, J.-P., & Skibsted, L. H. (2012). Reaction Dynamics of Flavonoids and Carotenoids as Antioxidants. Molecules, 17(2), 2140-2160. https://doi.org/10.3390/molecules17022140 Hashimoto, H., Uragami, C., Yukihira, N., Gardiner, A. T., & Cogdell, R. J. (2018). Understanding/unravelling carotenoid excited singlet states. Journal of The Royal Society Interface, 15(141), 20180026. https://doi.org/10.1098/rsif.2018.0026 Heap, I. (2023). The International Herbicide-Resistant Weed Database https://dx.doi.org/10.1002/ps.3696 Hill, W. G., Goddard, M. E., & Visscher, P. M. (2008). Data and theory point to mainly additive genetic variance for complex traits. PLoS genetics, 4(2), e1000008. Holopainen, J. K., & Gershenzon, J. (2010). Multiple stress factors and the emission of plant VOCs. Trends in Plant Science, 15(3), 176-184. [Record #197 is using a reference type undefined in this output style.] Howe, G. A., & Jander, G. (2008). Plant Immunity to Insect Herbivores. Annual review of plant biology, 59(1), 41-66. https://doi.org/10.1146/annurev.arplant.59.032607.092825 Hristova, V., & Popova, L. (2002). Treatment with methyl jasmonate alleviates the effects of paraquat on photosynthesis in barley plants. Photosynthetica, 40(4), 567-574. Hussain, S., Wang, J., Asad Naseer, M., Saqib, M., Siddiqui, M. H., Ihsan, F., Xiaoli, C., Xiaolong, R., Hussain, S., & Ramzan, H. N. (2023). Water stress memory in wheat/maize intercropping regulated photosynthetic and antioxidative responses under rainfed conditions. Scientific Reports, 13(1). https://doi.org/10.1038/s41598-023-40644-1 Jäck, O., Menegat, A., & Gerhards, R. (2017). Winter wheat yield loss in response to Avena fatua competition and effect of reduced herbicide dose rates on seed production of this species. Journal of Plant Diseases and Protection, 124, 371-382. Jackson, P. K., & Eldridge, A. G. (2002). The SCF ubiquitin ligase: an extended look. Molecular cell, 9(5), 923-925. Jingjing, C. (2020). Differences and mechanisms of resistance to quinclorac induced by exogenous methyl jasmonate and sensitive barnyard grass. Yingyong Shengtai Xuebao, 31(7). Jinks, J., & Pooni, H. (1976). Predicting the properties of recombinant inbred lines derived by single seed descent. Heredity, 36(2), 253-266. Joern, A. (1979). Feeding patterns in grasshoppers (Orthoptera: Acrididae): factors influencing diet specialization. Oecologia, 38, 325-347. 104 Johnson, R. R., Cranston, H. J., Chaverra, M. E., & Dyer, W. E. (1995). Characterization of cDNA clones for differentially expressed genes in embryos of dormant and nondormant Avena fatua L. caryopses. Plant Molecular Biology, 28(1), 113-122. https://doi.org/10.1007/bf00042043 Jugulam, & Shyam. (2019). Non-Target-Site Resistance to Herbicides: Recent Developments. Plants, 8(10), 417. https://doi.org/10.3390/plants8100417 Kalske, A., Shiojiri, K., Uesugi, A., Sakata, Y., Morrell, K., & Kessler, A. (2019). Insect Herbivory Selects for Volatile-Mediated Plant-Plant Communication. Current Biology, 29(18), 3128-3133.e3123. https://doi.org/10.1016/j.cub.2019.08.011 Kao, C.-H., Zeng, Z.-B., & Teasdale, R. D. (1999). Multiple interval mapping for quantitative trait loci. Genetics, 152(3), 1203-1216. Keith, B. K., Burns, E. E., Bothner, B., Carey, C. C., Mazurie, A. J., Hilmer, J. K., Biyiklioglu, S., Budak, H., & Dyer, W. E. (2017). Intensive herbicide use has selected for constitutively elevated levels of stress-responsive mRNAs and proteins in multiple herbicide-resistantAvena fatuaL. Pest Management Science, 73(11), 2267-2281. https://doi.org/10.1002/ps.4605 Keith, B. K., Lehnhoff, E. A., Burns, E. E., Menalled, F. D., & Dyer, W. E. (2015). Characterisation ofAvena fatuapopulations with resistance to multiple herbicides. Weed Research, 55(6), 621-630. https://doi.org/10.1111/wre.12172 Kern, A. J., Myers, T. M., Jasieniuk, M., Murray, B. G., Maxwell, B. D., & Dyer, W. E. (2002). Two Recessive Gene Inheritance for Triallate Resistance in Avena fatua L. Journal of Heredity, 93(1), 48-50. https://doi.org/10.1093/jhered/93.1.48 Kern, A. J., Peterson, D. M., Miller, E. K., Colliver, C. C., & Dyer, W. E. (1996). Triallate Resistance inAvena fatuaL. Is Due to Reduced Herbicide Activation. Pesticide Biochemistry and Physiology, 56(3), 163-173. Kessler, A., & Baldwin, I. T. (2001). Defensive function of herbivore-induced plant volatile emissions in nature. Science, 291(5511), 2141-2144. Khatami, S. A., Barmaki, M., Alebrahim, M. T., & Bajwa, A. A. (2022). Salicylic Acid Pre- Treatment Reduces the Physiological Damage Caused by the Herbicide Mesosulfuron- methyl+ Iodosulfuron-methyl in Wheat (Triticum aestivum). Agronomy, 12(12), 3053. Kirstine, W., Galbally, I., Ye, Y., & Hooper, M. (1998). Emissions of volatile organic compounds (primarily oxygenated species) from pasture. Journal of Geophysical Research: Atmospheres, 103(D9), 10605-10619. https://doi.org/10.1029/97jd03753 105 Landi, M., Araniti, F., Flamini, G., Piccolo, E. L., Trivellini, A., Abenavoli, M. R., & Guidi, L. (2020). “Help is in the air”: volatiles from salt-stressed plants increase the reproductive success of receivers under salinity. Planta, 251, 1-15. Lehnhoff, E. A., Keith, B. K., Dyer, W. E., Peterson, R. K., & Menalled, F. (2013). Multiple herbicide resistance in wild oat and impacts on physiology, germinability, and seed production. Agronomy Journal, 105(3), 854-862. Li, N., Han, X., Feng, D., Yuan, D., & Huang, L.-J. (2019). Signaling Crosstalk between Salicylic Acid and Ethylene/Jasmonate in Plant Defense: Do We Understand What They Are Whispering? International Journal of Molecular Sciences, 20(3), 671. https://doi.org/10.3390/ijms20030671 Li, Y.-F., Huang, L.-L., Liu, X.-L., Li, X.-S., & Tan, H.-H. (2020). Exogenous salicylic acid alleviates halosulfuron-methyl toxicity by coordinating the antioxidant system and improving photosynthesis in soybean (Glycine max Merr.). Acta physiologiae plantarum, 42, 1-10. Liao, Y., Smyth, G. K., & Shi, W. (2014). featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics, 30(7), 923-930. Lichtenthaler, H. K. (1987). [34] Chlorophylls and carotenoids: pigments of photosynthetic biomembranes. In Methods in enzymology (Vol. 148, pp. 350-382). Elsevier. Llusia, J., Penuelas, J., Asensio, D., & Munne-Bosch, S. (2005). Airborne limonene confers limited thermotolerance to Quercus ilex. Physiologia Plantarum, 123(1), 40-48. https://doi.org/10.1111/j.1399-3054.2004.00426.x Love, M. I., Huber, W., & Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome biology, 15(12), 1-21. Mackinnon, E., & Stone, S. L. (2022). The Ubiquitin Proteasome System and Nutrient Stress Response [Mini Review]. Frontiers in Plant Science, 13. https://doi.org/10.3389/fpls.2022.867419 Mahajan, G., & Chauhan, B. S. (2021). Seed longevity and seedling emergence behavior of wild oat (Avena fatua) and sterile oat (Avena sterilis ssp. ludoviciana) in response to burial depth in eastern Australia. Weed Science, 69(3), 362-371. Maoka, T. (2020). Carotenoids as natural functional pigments. Journal of Natural Medicines, 74(1), 1-16. https://doi.org/10.1007/s11418-019-01364-x Matzrafi, M., Shaar-Moshe, L., Rubin, B., & Peleg, Z. (2017). Unraveling the transcriptional basis of temperature-dependent pinoxaden resistance in Brachypodium hybridum. Frontiers in Plant Science, 8, 1064. 106 Meech, R., Miners, J. O., Lewis, B. C., & Mackenzie, P. I. (2012). The glycosidation of xenobiotics and endogenous compounds: versatility and redundancy in the UDP glycosyltransferase superfamily. Pharmacology & therapeutics, 134(2), 200-218. Moreira, X., Abdala-Roberts, L., Rasmann, S., Castagneyrol, B., & Mooney, K. A. (2016). Plant diversity effects on insect herbivores and their natural enemies: current thinking, recent findings, and future directions. Current opinion in insect science, 14, 1-7. Mostafa, S., Wang, Y., Zeng, W., & Jin, B. (2022). Plant responses to herbivory, wounding, and infection. International Journal of Molecular Sciences, 23(13), 7031. Moulton, J. K., Pepper, D. A., & Dennehy, T. J. (2000). Beet armyworm (Spodoptera exigua) resistance to spinosad. Pest Management Science: formerly Pesticide Science, 56(10), 842-848. Müller, P., Li, X.-P., & Niyogi, K. K. (2001). Non-Photochemical Quenching. A Response to Excess Light Energy1. Plant Physiology, 125(4), 1558-1566. https://doi.org/10.1104/pp.125.4.1558 Murchie, E. H., & Lawson, T. (2013). Chlorophyll fluorescence analysis: a guide to good practice and understanding some new applications. Journal of Experimental Botany, 64(13), 3983-3998. https://doi.org/10.1093/jxb/ert208 Myles, C., & Wayne, M. (2008). Quantitative trait locus (QTL) analysis. Nature Education 1 (1), 208. Naylor, J., & Jana, S. (1976). Genetic adaptation for seed dormancy in Avena fatua. Canadian Journal of Botany, 54(3-4), 306-312. Ņečajeva, J., Bleidere, M., Jansone, Z., Gailīte, A., & Ruņģis, D. (2021). Variability of seed germination and dormancy characteristics and genetic analysis of Latvian Avena fatua populations. Plants, 10(2), 235. Nietschke, B., Medd, R., Matthews, J., Reeves, T., & Powles, S. (1996). Managing herbicide resistant wild oats–options and adoption. Proceedings of 11th Australian Weeds Conference, ed. RCH Shepherd, Noroozi, M., & Sattari, A. (2015). Fine mapping of quantitative trait loci for grain weight in rice (Oryzasativa L.): a review. Intl J Farm Alli Sci, 4(4), 318-321. Ooijen, V. (2018). JoinMap® 5, Software for the calculation of genetic linkage maps in experimental populations of diploid species. Kyazma BV, Wageningen, Netherlands. Owen, M. J., & Powles, S. B. (2009). Distribution and frequency of herbicide-resistant wild oat (Avena spp.) across the Western Australian grain belt. Crop and Pasture Science, 60(1), 25-31. 107 Paik, I., & Huq, E. (2019). Plant photoreceptors: Multi-functional sensory proteins and their signaling networks. Seminars in Cell & Developmental Biology, Pandian, B. A., Sexton‐Bowser, S., Prasad, P. V., & Jugulam, M. (2022). Current status and prospects of herbicide‐resistant grain sorghum (Sorghum bicolor). Pest Management Science, 78(2), 409-415. Pandian, B. A., Varanasi, A., Vennapusa, A. R., Sathishraj, R., Lin, G., Zhao, M., Tunnell, M., Tesso, T., Liu, S., & Prasad, P. V. (2020). Characterization, genetic analyses, and identification of QTLs conferring metabolic resistance to a 4-hydroxyphenylpyruvate dioxygenase inhibitor in sorghum (Sorghum bicolor). Frontiers in Plant Science, 11, 596581. Park, S.-W., Kaimoyo, E., Kumar, D., Mosher, S., & Klessig, D. F. (2007). Methyl Salicylate Is a Critical Mobile Signal for Plant Systemic Acquired Resistance. Science, 318(5847), 113- 116. https://doi.org/doi:10.1126/science.1147113 Peng, Y., Yan, H., Guo, L., Deng, C., Wang, C., Wang, Y., Kang, L., Zhou, P., Yu, K., & Dong, X. (2022). Reference genome assemblies reveal the origin and evolution of allohexaploid oat. Nature Genetics, 54(8), 1248-1258. Peters, R. D., & Noble, S. D. (2014). Spectrographic measurement of plant pigments from 300 to 800 nm. Remote sensing of environment, 148, 119-123. Picazo-Aragonés, J., Terrab, A., & Balao, F. (2020). Plant volatile organic compounds evolution: Transcriptional regulation, epigenetics and polyploidy. International Journal of Molecular Sciences, 21(23), 8956. Possell, M., & Loreto, F. (2013). The role of volatile organic compounds in plant resistance to abiotic stresses: responses and mechanisms. In Biology, controls and models of tree volatile organic compound emissions (pp. 209-235). Springer. Powles, S. B., & Yu, Q. (2010). Evolution in action: plants resistant to herbicides. Annual review of plant biology, 61, 317-347. Quach, H. T., Steeper, R. L., & Griffin, G. W. (2004). An Improved Method for the Extraction and Thin-Layer Chromatography of Chlorophyll a and b from Spinach. Journal of Chemical Education, 81(3), 385. https://doi.org/10.1021/ed081p385 Raymond, & Bunthawin, S. (2010). The Use of Pulse Amplitude Modulation (PAM) Fluorometry to Measure Photosynthesis in a CAM Orchid, Dendrobium spp. (D. cv. Viravuth Pink). International Journal of Plant Sciences, 171(6), 575-585. https://doi.org/10.1086/653131 Reddy, K., & Nandula, V. (2012). Herbicide resistant crops: history, development and current technologies. Indian Journal of Agronomy, 57(1), 1-7. 108 Rivas-Garcia, T., Espinosa-Calderón, A., Hernández-Vázquez, B., & Schwentesius-Rindermann, R. (2022). Overview of Environmental and Health Effects Related to Glyphosate Usage. Sustainability, 14(11), 6868. https://www.mdpi.com/2071-1050/14/11/6868 Roach, T., & Krieger-Liszkay, A. (2014). Regulation of photosynthetic electron transport and photoinhibition. Current Protein and Peptide Science, 15(4), 351-362. Rodriguez-Saona, C., Kaplan, I., Braasch, J., Chinnasamy, D., & Williams, L. (2011). Field responses of predaceous arthropods to methyl salicylate: a meta-analysis and case study in cranberries. Biological control, 59(2), 294-303. Rojas, J. C. (1999). Influence of Host Plant Damage on the Host-Finding Behavior of Mamestra brassicae (Lepidoptera: Noctuidae). Environmental Entomology, 28(4), 588-593. https://doi.org/10.1093/ee/28.4.588 Römer, S., & Fraser, P. D. (2005). Recent advances in carotenoid biosynthesis, regulation and manipulation. Planta, 221(3), 305-308. https://doi.org/10.1007/s00425-005-1533-5 Roos‐Mattjus, P., & Sistonen, L. (2004). The ubiquitin‐proteasome pathway. Annals of Medicine, 36(4), 285-295. https://doi.org/10.1080/07853890310016324 Ruban, A. V., Johnson, M. P., & Duffy, C. D. (2012). The photoprotective molecular switch in the photosystem II antenna. Biochimica et Biophysica Acta (BBA)-Bioenergetics, 1817(1), 167-181. Sako, K., Nguyen, H. M., & Seki, M. (2020). Advances in Chemical Priming to Enhance Abiotic Stress Tolerance in Plants. Plant and Cell Physiology, 61(12), 1995-2003. https://doi.org/10.1093/pcp/pcaa119 Santino, A., Taurino, M., De Domenico, S., Bonsegna, S., Poltronieri, P., Pastor, V., & Flors, V. (2013). Jasmonate signaling in plant development and defense response to multiple (a) biotic stresses. Plant cell reports, 32, 1085-1098. Schneider, C. A., Rasband, W. S., & Eliceiri, K. W. (2012). NIH Image to ImageJ: 25 years of image analysis. Nature Methods, 9(7), 671-675. https://doi.org/10.1038/nmeth.2089 Searle, S. R., Speed, F. M., & Milliken, G. A. (1980). Population Marginal Means in the Linear Model: An Alternative to Least Squares Means. The American Statistician, 34(4), 216- 221. https://doi.org/10.1080/00031305.1980.10483031 Sharma, S. K., Singh, D., Pandey, H., Jatav, R. B., Singh, V., & Pandey, D. (2022). An overview of roles of enzymatic and nonenzymatic antioxidants in plant. Antioxidant Defense in Plants: Molecular Basis of Regulation, 1-13. Shen, N., Wang, T., Gan, Q., Liu, S., Wang, L., & Jin, B. (2022). Plant flavonoids: Classification, distribution, biosynthesis, and antioxidant activity. Food chemistry, 383, 132531. 109 Silva, S., Costa, E., Calhau, C., Morais, R., & Pintado, M. (2017). Anthocyanin extraction from plant tissues: A review. Critical reviews in food science and nutrition, 57(14), 3072-3083. Silverman, Petracek, P. D., Heiman, D. F., Ju, Z., Fledderman, C. M., & Warrior, P. (2005). Salicylate Activity. 2. Potentiation of Atrazine. Journal of Agricultural and Food Chemistry, 53(25), 9769-9774. https://doi.org/10.1021/jf0513821 Silverman, F. P., Petracek, P. D., Fledderman, C. M., Ju, Z., Heiman, D. F., & Warrior, P. (2005). Salicylate activity. 1. Protection of plants from paraquat injury. Journal of Agricultural and Food Chemistry, 53(25), 9764-9768. Sitaraman, J., Bui, M., & Liu, Z. (2008). LEUNIG_HOMOLOG and LEUNIG perform partially redundant functions during Arabidopsis embryo and floral development. Plant Physiology, 147(2), 672-681. Stewart, J. J., Adams, W. W., López-Pozo, M., Doherty Garcia, N., McNamara, M., Escobar, C. M., & Demmig-Adams, B. (2021). Features of the Duckweed Lemna That Support Rapid Growth under Extremes of Light Intensity. Cells, 10(6), 1481. https://doi.org/10.3390/cells10061481 Strapasson, P., Pinto-Zevallos, D. M., Gomes, S. M. D. S., & Zarbin, P. H. (2016). Volatile organic compounds induced by herbivory of the soybean looper Chrysodeixis includens in transgenic glyphosate-resistant soybean and the behavioral effect on the parasitoid, Meteorus rubens. Journal of chemical ecology, 42(8), 806-813. Sudan, J., Sharma, S., Salgotra, R. K., Pandey, R. K., Neelam, D., & Singh, R. (2023). Elucidating the process of SNPs identification in non-reference genome crops. Journal of Biomolecular Structure and Dynamics, 1-9. Sun, X.-L., Wang, G.-C., Gao, Y., Zhang, X.-Z., Xin, Z.-J., & Chen, Z.-M. (2014). Volatiles Emitted from Tea Plants Infested by Ectropis obliqua Larvae Are Attractive to Conspecific Moths. Journal of Chemical Ecology, 40(10), 1080-1089. https://doi.org/10.1007/s10886-014-0502-5 Tamogami, S., Rakwal, R., & Agrawal, G. K. (2008). Interplant communication: airborne methyl jasmonate is essentially converted into JA and JA-Ile activating jasmonate signaling pathway and VOCs emission. Biochemical and Biophysical Research Communications, 376(4), 723-727. Tauber, M. J., & Tauber, C. A. (1976). Insect seasonality: diapause maintenance, termination, and postdiapause development. Annual review of entomology, 21(1), 81-107. Taylor, J., & Butler, D. (2017). R package ASMap: efficient genetic linkage map construction and diagnosis. arXiv preprint arXiv:1705.06916. 110 Tiwari, S., Kate, A., Mohapatra, D., Tripathi, M. K., Ray, H., Akuli, A., Ghosh, A., & Modhera, B. (2020). Volatile organic compounds (VOCs): Biomarkers for quality management of horticultural commodities during storage through e-sensing. Trends in Food Science & Technology, 106, 417-433. Tiwari, S., Tiwari, S., Singh, M., Singh, A., & Prasad, S. M. (2017). Generation mechanisms of reactive oxygen species in the plant cell: an overview. Reactive Oxygen Species in Plants: Boon Or Bane‐Revisiting the Role of ROS, 1-22. Travlos, I. S., Giannopolitis, C. N., & Economou, G. (2011). Diclofop resistance in sterile wild oat (Avena sterilis L.) in wheat fields in Greece and its management by other post- emergence herbicides. Crop Protection, 30(11), 1449-1454. Traxler, C., Gaines, T. A., Küpper, A., Luemmen, P., & Dayan, F. E. (2023). The nexus between reactive oxygen species and the mechanism of action of herbicides. Journal of Biological Chemistry, 105267. Tripathy, B. C., & Oelmüller, R. (2012). Reactive oxygen species generation and signaling in plants. Plant Signaling & Behavior, 7(12), 1621-1633. https://doi.org/10.4161/psb.22455 Turner, J. G., Ellis, C., & Devoto, A. (2002). The jasmonate signal pathway. The Plant Cell, 14(suppl_1), S153-S164. Ueda, H., Kikuta, Y., & Matsuda, K. (2012). Plant communication. Plant Signaling & Behavior, 7(2), 222-226. https://doi.org/10.4161/psb.18765 Vasilakoglou, I., Dhima, K., Paschalidis, K., & Ritzoulis, C. (2013). Herbicidal potential onLolium rigidumof nineteen major essential oil components and their synergy. Journal of Essential Oil Research, 25(1), 1-10. https://doi.org/10.1080/10412905.2012.751054 Verma, B., Kumar, N., & Pratap, G. (2023). SYSTEMIC ACQUIRED RESISTANCE. Dr. Ajay B. Jadhao, 87. Vickerman, D. B., & Trumble, J. T. (1999). Feeding preferences of Spodoptera exigua in response to form and concentration of selenium. Archives of Insect Biochemistry and Physiology: Published in Collaboration with the Entomological Society of America, 42(1), 64-73. Vila, M., Beaury, E. M., Blumenthal, D. M., Bradley, B. A., Early, R., Laginhas, B. B., Trillo, A., Dukes, J. S., Sorte, C. J., & Ibanez, I. (2021). Understanding the combined impacts of weeds and climate change on crops. Environmental Research Letters, 16(3), 034043. Vivaldo, G., Masi, E., Taiti, C., Caldarelli, G., & Mancuso, S. (2017). The network of plants volatile organic compounds. Scientific Reports, 7(1). https://doi.org/10.1038/s41598-017- 10975-x 111 Voitsekhovskaja, O., & Tyutereva, E. (2015). Chlorophyll b in angiosperms: functions in photosynthesis, signaling and ontogenetic regulation. Journal of plant physiology, 189, 51-64. Wang, Y., Mostafa, S., Zeng, W., & Jin, B. (2021). Function and Mechanism of Jasmonic Acid in Plant Responses to Abiotic and Biotic Stresses. International Journal of Molecular Sciences, 22(16), 8568. https://doi.org/10.3390/ijms22168568 Wellburn, A. R. (1994). The spectral determination of chlorophylls a and b, as well as total carotenoids, using various solvents with spectrophotometers of different resolution. Journal of plant physiology, 144(3), 307-313. Wrzesińska, B., Kierzek, R., & Obrępalska‐Stęplowska, A. (2016). Evaluation of six commonly used reference genes for gene expression studies in herbicide‐resistant Avena fatua biotypes. Weed Research, 56(4), 284-292. Würschum, T. (2012). Mapping QTL for agronomic traits in breeding populations. Theoretical and Applied Genetics, 125(2), 201-210. https://doi.org/10.1007/s00122-012-1887-6 Yu, Q., Ahmad-Hamdani, M., Han, H., Christoffers, M., & Powles, S. (2013). Herbicide resistance-endowing ACCase gene mutations in hexaploid wild oat (Avena fatua): insights into resistance evolution in a hexaploid species. Heredity, 110(3), 220-231. Yuan, J. S., Himanen, S. J., Holopainen, J. K., Chen, F., & Stewart, C. N. (2009). Smelling global climate change: mitigation of function for plant volatile organic compounds. Trends in Ecology & Evolution, 24(6), 323-331. https://doi.org/10.1016/j.tree.2009.01.012 Zhang, L., Su, Q. F., Wang, L. S., Lv, M. W., Hou, Y. X., & Li, S. S. (2023). Linalool: A ubiquitous floral volatile mediating the communication between plants and insects. Journal of Systematics and Evolution, 61(3), 538-549. https://doi.org/10.1111/jse.12930 Zhou , S., & Jander, G. (2021). Molecular ecology of plant volatiles in interactions with insect herbivores. Journal of Experimental Botany, 73(2), 449-462. https://doi.org/10.1093/jxb/erab413 Zulet, A., Gil-Monreal, M., Villamor, J. G., Zabalza, A., Van Der Hoorn, R. A. L., & Royuela, M. (2013). Proteolytic Pathways Induced by Herbicides That Inhibit Amino Acid Biosynthesis. PLoS ONE, 8(9), e73847. https://doi.org/10.1371/journal.pone.0073847