Abstract

Small-molecule phytohormones exert control over plant growth, development, and stress responses by coordinating the patterns of gene expression within and between cells. Increasing evidence indicates that currently recognized plant hormones are part of a larger group of regulatory metabolites that have acquired signaling properties during the evolution of land plants. This rich assortment of chemical signals reflects the tremendous diversity of plant secondary metabolism, which offers evolutionary solutions to the daunting challenges of sessility and other unique aspects of plant biology. A major gap in our current understanding of plant regulatory metabolites is the lack of insight into the direct targets of these compounds. Here, we illustrate the blurred distinction between classical phytohormones and other bioactive metabolites by highlighting the major scientific advances that transformed the view of jasmonate from an interesting floral scent to a potent transcriptional regulator. Lessons from jasmonate research generally apply to other phytohormones and thus may help provide a broad understanding of regulatory metabolite–protein interactions. In providing a framework that links small-molecule diversity to transcriptional plasticity, we hope to stimulate future research to explore the evolution, functions, and mechanisms of perception of a broad range of plant regulatory metabolites.

ADVANCES
  • The past century of plant hormone research has helped reveal how low-molecular-weight metabolites modulate transcription to promote plasticity in growth, development, and environmental responses.

  • Recent knowledge of the evolutionary origins of phytohormone biosynthesis, perception, and signaling is helping redefine the hormone concept to reflect the unique biology of plants.

  • Currently recognized phytohormones are part of a larger group of metabolites that alter gene expression with different biological effects and degrees of conservation within land plants.

  • Classical genetic approaches to study jasmonates and other phytohormones provide a roadmap to exploring the activities of other regulatory metabolites.

  • Advances in mass spectrometry, molecular modeling, and artificial intelligence-based structural prediction provide new tools to identify regulatory metabolite–protein interactions in plants.

Introduction

Plants use a diverse repertoire of chemical messengers to coordinate their growth and development with informational cues from the environment. Among organic signaling compounds, small-molecule plant hormones (i.e. phytohormones) have been the focus of much research during the past century. Currently recognized phytohormones influence virtually all aspects of growth, development, and environmental acclimation and tend to be broadly conserved among land plants. Evolutionary reconstruction of plant hormone response pathways using phylogenomic approaches has provided important insights into the emergence of small molecules as drivers of phenotypic plasticity and innovations in life history traits (Bowman et al. 2017; Blázquez et al. 2020). Moreover, the molecular mechanisms governing phytohormone homeostasis and action have emerged as key targets for improving the productivity and stress resilience of crops (Hedden 2003; Bailey-Serres et al. 2019).

What is the meaning of “hormone” as it applies to the unique biology of plants? Beyond the limited number of compounds elevated to phytohormone status in textbooks, what other plant metabolites have signaling properties? What are the mechanisms by which these molecules are perceived and transduced to downstream effectors? How do such metabolites modulate transcription, and how are these functions acquired during evolution? Finally, how can this knowledge inform the development of new crop varieties that are more productive, nutritious, and better adapted to changing climate conditions? These are the main questions that motivated this topical review.

Phytohormones as regulatory metabolites

Here, we develop the hypothesis that currently recognized plant hormones are part of a larger and more diverse group of metabolites that acquired regulatory attributes over the course of embryophyte evolution. This supposition is based on the increasing recognition that many secondary (also referred to as “specialized”) metabolites exhibit hormone-like effects on growth, development, and responses to the environment. The mechanisms by which these bioactive compounds act remain largely unknown but are often associated with changes in gene expression. The past century of plant hormone research, combined with new approaches to systematically map protein–metabolite interactions, provides a general roadmap for future progress in this area. We hope that this conceptual framework helps guide future research aimed at understanding the evolution, functions, and mechanisms of action of plant regulatory metabolites beyond currently recognized phytohormones.

Overview of phytohormone actions and origins

The identification of genes required for plant hormone biosynthesis and action resulted largely from the use of genetic approaches applied to the model eudicot species Arabidopsis (Arabidopsis thaliana). This branch of research greatly advanced our understanding of how “linear” phytohormone-response pathways function at the molecular level, including the discovery of ethylene as the first gaseous signaling molecule (Kende 1998), the elucidation of mechanisms to perceive steroids and auxins at the cell surface (He et al. 2000; Yu et al. 2023), and the paradigm that some hormones act as molecular glue to promote the reorganization of transcription complexes (Dharmasiri et al. 2003; Tan et al. 2007; Sheard et al. 2010). A deep mechanistic understanding of phytohormone-response pathways has also provided a foundation on which to explore more complex questions, such as the evolutionary origins of chemical communication in plants, cross-regulation between phytohormone-response pathways, the ecological relevance of phytohormones, and strategies to engineer economically important plant traits.

Currently recognized plant hormones exert their effects by triggering signal transduction pathways that modulate the activities of specific cellular effectors, most often transcription factors (TFs). It is now evident that the phenotypic outputs of hormone signaling cascades are strongly influenced by synergistic and antagonistic interactions with many other signaling pathways (Bari and Jones 2009; Santner et al. 2009; Altmann et al. 2020). Due to this interconnectivity, phytohormone-response networks regulate gene expression and other physiological processes in a context-specific manner that is shaped by multiple developmental and environmental (e.g. light) cues (Fig. 1). This complex web of interactions likely explains the pleiotropic effects of hormones and why they appear to “do” so many things. A better understanding of how phytohormones and other bioactive metabolites interact to influence gene expression networks has the potential to improve crop performance and thus represents a major goal of modern plant science (Berens et al. 2017).

Common themes in plant hormone biosynthesis and signaling. A) Phytohormones can be viewed as secondary (i.e. specialized) metabolites that are derived from several major classes of primary metabolites. Some hormones, including JA-Ile and cytokinin (CK), are synthesized from more than one primary metabolite class. Peptide hormones are omitted for simplicity. B) Phytohormone-response pathways consist of several highly integrated processes (bold text) that collectively modulate the activity of one or more effectors. Hormone biosynthetic pathways utilize precursors from primary metabolism and give rise to bioactive compounds whose intracellular concentrations are several orders of magnitude below those of primary metabolic precursors. Bioactive hormones (depicted as a generic chemical structure) are subject to oxidation, conjugation, and other catabolic reactions that inactivate the signal (dashed arrow). Hormone perception is mediated by specific proteinaceous receptors. Many receptors are associated with a ligand-induced enzymatic activity (small curved arrow). Ligand binding transmits the signal to a downstream effector such as a TF. Effector TFs bind to specific cis-regulatory elements (CRE) in response genes and work together with various coactivators and chromatin-modifying enzymes to control the rate of transcription via RNA polymerase II (Pol II). TFs and other effectors generate physiological responses. These responses are subject to multiple levels of feedback regulation, for example, as a means of preventing the overactivation of a response. Hormone-responsive TFs are subject to many additional layers of regulation, including inputs from multiple developmental and environmental signals (green arrows). As a consequence, the expression of any given target gene is context dependent; effector TFs constitute the apex of the regulatory hierarchy, with hormone-receptor inputs serving a condition-specific role. ABA, abscisic acid; BR, brassinosteroid; ET, ethylene; GA, gibberellin; IAA, indole acetic acid (auxin); SA, salicylic acid; SL, strigolactone.
Figure 1.

Common themes in plant hormone biosynthesis and signaling. A) Phytohormones can be viewed as secondary (i.e. specialized) metabolites that are derived from several major classes of primary metabolites. Some hormones, including JA-Ile and cytokinin (CK), are synthesized from more than one primary metabolite class. Peptide hormones are omitted for simplicity. B) Phytohormone-response pathways consist of several highly integrated processes (bold text) that collectively modulate the activity of one or more effectors. Hormone biosynthetic pathways utilize precursors from primary metabolism and give rise to bioactive compounds whose intracellular concentrations are several orders of magnitude below those of primary metabolic precursors. Bioactive hormones (depicted as a generic chemical structure) are subject to oxidation, conjugation, and other catabolic reactions that inactivate the signal (dashed arrow). Hormone perception is mediated by specific proteinaceous receptors. Many receptors are associated with a ligand-induced enzymatic activity (small curved arrow). Ligand binding transmits the signal to a downstream effector such as a TF. Effector TFs bind to specific cis-regulatory elements (CRE) in response genes and work together with various coactivators and chromatin-modifying enzymes to control the rate of transcription via RNA polymerase II (Pol II). TFs and other effectors generate physiological responses. These responses are subject to multiple levels of feedback regulation, for example, as a means of preventing the overactivation of a response. Hormone-responsive TFs are subject to many additional layers of regulation, including inputs from multiple developmental and environmental signals (green arrows). As a consequence, the expression of any given target gene is context dependent; effector TFs constitute the apex of the regulatory hierarchy, with hormone-receptor inputs serving a condition-specific role. ABA, abscisic acid; BR, brassinosteroid; ET, ethylene; GA, gibberellin; IAA, indole acetic acid (auxin); SA, salicylic acid; SL, strigolactone.

Plant hormones are derived from nearly all classes of primary metabolites, including isoprenoids, amino acids, fatty acids, and nucleotides (Fig. 1 and Box 1). Given their potent and wide-ranging effects on gene expression, endogenous hormone concentrations are subject to exquisite feedback control at the level of biosynthesis, transport, and catabolism. Bioactive end products of hormone biosynthetic pathways are perceived by specific proteinaceous receptors that, upon ligand binding, transduce the signal to a major TF or other effector (Fig. 1). High-affinity binding of hormones to their cognate receptor accounts for the ability of these metabolites to exert physiological effects at very low concentrations. Elucidation of the 3-dimensional structures of hormone–receptor complexes has yielded unprecedented insight into how ligand binding is linked to rapid changes in effector activity (Takeuchi et al. 2021). Some hormones induce a conformational change (i.e. allosteric regulation) in their receptor, whereas others act as molecular glues to promote protein–protein interactions that transduce the signal to an effector. Ligand–receptor interactions typically activate target effectors through changes in the activity of a receptor-linked enzyme, such as a kinase, phosphatase, or ubiquitin ligase. The integration of hormone perception with enzymatic function is further illustrated by the recent discovery of plant hormone receptors having α/β-hydrolase or nucleotide cyclase activity (Takeuchi et al. 2021; Qi et al. 2022).

Box 1.
Plant hormones are found in unexpected places

The tendency of phytohormones to exert broad-spectrum effects (Fig. 2) may reflect their discovery through the use of bioassays designed to identify substances that modulate plant growth and cell division. In this context, it is interesting to note that activity-driven purification of phytohormones employed starting materials not only from plant sources but also complex chemical mixtures from non-plant sources, including coal gas (ethylene), herring sperm DNA (cytokinin), human urine (auxin), and fungal exudates (gibberellins and jasmonates); only later was it determined that these metabolites are endogenous plant compounds. The powerful biological effects of classical hormones may explain why these low-hanging fruits on the larger tree of regulatory metabolites were the first to be identified using bioassay-based approaches biased toward easily measurable phenotypes (e.g. growth) that involve large-scale reprogramming of gene expression. Endogenous regulatory metabolites with more subtle effects on a narrow range of effectors surely exist (Fig. 2), but alternative experimental approaches may be needed to identify them (see Concluding remarks and future perspectives). Many phytohormones are produced by nonplant taxa, including bacteria, fungi, and algae, which lack the canonical hormone response pathways found in land plants. The production of phytohormones in these lineages has fascinating implications for understanding how processes such as horizontal gene transfer and endosymbiosis have shaped the relationships of plants with other groups of organisms (Bari and Jones 2009; Morffy and Strader 2020).

The advent of functional phylogenomic approaches has spurred remarkable progress in understanding the evolution of phytohormone-response pathways and, more broadly, the understudied interface between metabolism and signaling (Bowman et al. 2017; Blázquez et al. 2020; Rieseberg et al. 2023). Small-molecule plant hormones have features ascribed to both primary and secondary metabolites (Erb and Kliebenstein 2020; Fàbregas and Fernie 2021). Primary metabolites, such as amino acids, fatty acids, and sugars, are relatively abundant compounds (e.g. micromolar range) whose critical roles in the plant lifecycle stem from their use as building blocks that fuel tissue growth and increases in biomass. Although many hormones are also essential for the plant lifecycle, this dependency reflects the signaling properties of these metabolites, which typically accumulate to minuscule levels (e.g. subnanomolar) relative to the primary compounds from which they are derived. Whereas primary metabolites are produced by all organisms and cell types, the distribution of phytohormones in the tree of life is more sporadic and includes recent innovations within land plants (Box 1) (Bowman et al. 2017; Weng et al. 2021; Chini et al. 2023; Rieseberg et al. 2023). These collective observations support the view that small-molecule phytohormones trace their evolutionary origins to specialized metabolites that were co-opted for signaling roles in either the ancestral land plant or within land plants (Bowman et al. 2017; Rieseberg et al. 2023). Accordingly, plants use small-molecule signals in ways that are often unique to the plant kingdom (Box 2).

Box 2.
Phytohormone signaling pathways are unique

The term “hormone” was coined by animal physiologists to convey the notion of chemical messengers that coordinate the growth and metabolic activity in different parts of the body (Tata 2005). This view was founded on observations that mammalian hormones are produced in specific locations, transported through the bloodstream, and generate physiological effects in target tissues. Plant scientists have generally, but not universally (Trewavas 1981; Bennett and Leyser 2014), embraced the hormone paradigm because it accounts for the potent biological effects of endogenous plant growth and immune regulators (Davies 2010). Beyond a general role in mediating intercellular communication, however, fundamental differences between plant and animal biology suggest that extrapolating hormone signaling mechanisms from animals to plants is naive at best, and at worst hinders a full appreciation of how plants use small molecules as informational cues. Indeed, the independent evolution of multicellularity in plants endowed them with signal communication networks that were likely shaped by unique features of the embryophyte lineage: a thick cell wall that restricts cell mobility; an autotrophic, sessile lifestyle enabled by photosynthetic plastids; increased developmental flexibility (i.e. totipotency) of interdependent above- and belowground organs; and an extraordinary diversity of metabolites to help meet the profound challenges of sessility. While both plants and animals rely on transcriptional programming to specify patterns of cell identity and growth, the signaling components responsible for executing these developmental programs in plants and animals are largely nonhomologous (Meyerowitz 2002). Moreover, many of the storied signaling components that helped to establish the animal hormone paradigm, including G protein-coupled receptors and associated second messengers such as cyclic AMP, either appear to be missing in plants (Urano and Jones 2014) or function in plants in ways that deviate from the animal paradigm (Wong et al. 2023). Plant cells also appear to be more autonomous than their animal counterparts, such that most plant cell types likely have the capacity to both synthesize and perceive a given chemical signal. Accordingly, the classic hormone concept of “action at a distance” does not necessarily apply to plants, even though many plant signaling compounds are transported long distances. If we have learned anything from the past century of phytohormone research, it is that plants use low-molecular-weight metabolites as informational cues in ways that are often unique to the plant kingdom.

Phytohormones versus regulatory metabolites: what's the difference?

There is general consensus about the properties that define a phytohormone (Davies 2010). These characteristics include the ability to pleiotropically modulate growth, developmental, and immune-related processes, activity at very low concentrations, rapid biological responses to fluctuating concentrations of the hormone signal, and a high degree of conservation in the plant kingdom. Plant hormones are also perceived by dedicated receptors that transduce the signal to major effectors of physiological change.

Beyond these general features, how should we think about phytohormones conceptually? As is the case for other complex multicellular organisms, the spatial and temporal action of transcriptional regulators (e.g. TFs) plays a dominant role in executing a genetic blueprint that specifies rudimentary plant phenotypes. Phytohormone networks appear to confer plasticity to these hard-wired processes by providing real-time, environmentally dependent inputs that coordinate gene expression patterns between cell types, tissue types, and organs. This conceptual view of phytohormones as conditional modulators of gene expression has implications for understanding the stepwise assembly of hormone-response pathways during evolution (Bowman et al. 2017; Mutte et al. 2018; Blázquez et al. 2020). Increasing evidence indicates that protein components of various phytohormone signaling cascades predated the recruitment of the hormone into the pathway and thus have the capacity to control output responses independently of the hormone (Mutte et al. 2018; Sun et al. 2019; Bunsick et al. 2021). For example, auxin appears to have been recruited as a signal based on its ability to promote protein–protein interactions between preexisting components of the pathway, thereby enabling efficient control of TF activity (Mutte et al. 2018; Cao et al. 2022). In this scenario, the major drivers of plant phenotype are ancient transcriptional regulators such as TFs, which over evolutionary time acquired the capacity to receive and integrate information from internal (e.g. metabolic) and external (e.g. light) cues (Fig. 1).

A unified phytohormone concept also needs to address the lack of a clear distinction between known hormones and an increasing number of plant metabolites that exhibit regulatory properties (Erb and Kliebenstein 2020). Evolutionary theory predicts that plants possess a diverse array of regulatory metabolites that differ not only in their biosynthetic origin and phylogenetic distribution but also in their range of biological effects (Fig. 2). We propose that hormones should be viewed not as an exclusive group of endogenous mediators but rather as part of a spectrum of compounds that acquired signaling attributes during or after the colonization of land by plants. This framework is consistent with current knowledge of how specialized metabolites in general, and phytohormones in particular, evolved (Bowman et al. 2017; Blázquez et al. 2020; Rieseberg et al. 2023). Phylogenomic analyses suggested that most hormone-response pathways emerged early in the evolution of land plants and, because of their essential roles in the plant lifecycle, were retained in present-day lineages (Bowman et al. 2017). Regulatory metabolites whose distribution is restricted to specific plant taxa were either recruited later in evolution or perhaps discarded from other lineages due to their lack of utility in a particular environment. This evolutionary perspective accounts for the existence of lineage-specific regulatory metabolites (e.g. certain glucosinolate derivatives) that confer subtle—and thus difficult-to-detect—fitness advantages in a given environment (Fig. 2) (Erb and Kliebenstein 2020).

Diversity of plant regulatory metabolites. A) Chemical structures of currently recognized phytohormones (top group) and representative regulatory metabolites (bottom group). B) Plant regulatory metabolites (gray circles) differ not only in their chemical structures but also in their phylogenetic distribution (y axis) and range of biological effects (x axis). For example, phytohormone action is typically linked to the differential expression of hundreds (or thousands) of genes, whereas other regulatory metabolites modulate the expression of a smaller set of genes. Owing to their strong biological effects and high degree of conservation in land plants, currently recognized phytohormones tend to occupy the upper right quadrant (dashed circle) of the plot. 3OHPGSL, 3-hydroxypropylglucosinolate; ABA, abscisic acid; BL, brassinolide; CK, cytokinin (trans-zeatin); DMNT, (E)-3,8-dimethyl-1,4,7-nonatriene; ET, ethylene; GA, gibberellic acid (GA3); I3C, indole-3-carbinol; IAA, indole-3-acetic acid (auxin); Jasmonic acid (JA)-Ile, jasmonoyl-L-isoleucine; NHP, N-hydroxy-pipecolic acid; SA, salicylic acid; SL, strigolactone (strigol); βCC, β-cyclocitral.
Figure 2.

Diversity of plant regulatory metabolites. A) Chemical structures of currently recognized phytohormones (top group) and representative regulatory metabolites (bottom group). B) Plant regulatory metabolites (gray circles) differ not only in their chemical structures but also in their phylogenetic distribution (y axis) and range of biological effects (x axis). For example, phytohormone action is typically linked to the differential expression of hundreds (or thousands) of genes, whereas other regulatory metabolites modulate the expression of a smaller set of genes. Owing to their strong biological effects and high degree of conservation in land plants, currently recognized phytohormones tend to occupy the upper right quadrant (dashed circle) of the plot. 3OHPGSL, 3-hydroxypropylglucosinolate; ABA, abscisic acid; BL, brassinolide; CK, cytokinin (trans-zeatin); DMNT, (E)-3,8-dimethyl-1,4,7-nonatriene; ET, ethylene; GA, gibberellic acid (GA3); I3C, indole-3-carbinol; IAA, indole-3-acetic acid (auxin); Jasmonic acid (JA)-Ile, jasmonoyl-L-isoleucine; NHP, N-hydroxy-pipecolic acid; SA, salicylic acid; SL, strigolactone (strigol); βCC, β-cyclocitral.

While it is obvious that all small-molecule hormones are regulatory metabolites, it is less clear which of these compounds should be categorized as phytohormones. Dwelling on this ambiguous distinction has limited practical value because the evolutionary tinkering that recruits metabolites into their myriad roles is blind to such classification schemes. One solution to this quandary is to phase out the use of the term “hormone” in favor of a more general term such as regulatory metabolite (Bennett and Leyser 2014). A quick survey of the literature, however, suggests that the plant hormone/phytohormone labels are here to stay. Suffice it to say that currently recognized phytohormone pathways mediate potent and pleiotropic biological effects in a wide range of plant species. Debating whether or not a particular metabolite should be classified as a hormone is less meaningful than determining how that compound functions at a mechanistic level; such insights are needed to obtain a satisfactory understanding of the physiological, ecological, and evolutionary relevance of bioactive metabolites.

The expanding universe of plant regulatory metabolites

Twenty-five years ago, Kende and Zeevaart (1997) predicted that the list of 5 “classical” plant hormones, namely abscisic acid, auxin, cytokinin, gibberellin, and ethylene, would expand as the functions of other bioactive compounds became recognized. This expectation has materialized with the general acceptance of brassinosteroid, jasmonate, salicylic acid, and strigolactone as the newest members of the group (Santner et al. 2009; Takeuchi et al. 2021). Beyond these known phytohormones, to what extent do plant metabolites have signaling attributes and how were these properties acquired during evolution? One way to address this question is through the study of presumptive regulatory metabolites whose chemical identity is unknown. The KARRIKIN INSENSITIVE 2 receptor, which perceives smoke-derived compounds (i.e. karrikins) that promote seed germination, is thought to perceive an endogenous ligand, but the identity of this compound(s) remains elusive (Waters and Nelson 2023). Another example is the cytochrome P450 KLUH/KLU, which in Arabidopsis is involved in the production of a mobile growth-promoting factor that is likely distinct from known phytohormones (Anastasiou et al. 2007).

A second approach to expand our knowledge of plant regulatory metabolites is to identify the molecular targets of endogenous compounds that exhibit signaling properties. The list of such compounds is lengthy and includes peptides, sugars (e.g. trehalose-6-phosphate), nucleotides (e.g. cyclic nucleotides), and amino acids (e.g. Glu). The direct targets of several of these compounds have been identified and studied in detail (Sandoval and Santiago 2020). Below, we highlight a few examples of secondary metabolites whose regulatory properties and modes of action are unknown but are under active investigation. Readers are referred to recent reviews for a more in-depth discussion of bioactive plant compounds originating from secondary metabolism (Erb and Kliebenstein 2020; Durán-Medina et al. 2021; Weng et al. 2021; Cai and Aharoni 2022).

Plant secondary metabolites account for much of the chemical diversity in nature. This large assortment of compounds exerts direct effects on plant-interacting organisms and also promotes adaptation to abiotic stress conditions. Interestingly, an increasing number of specialized metabolites are reported to possess signaling attributes reminiscent of phytohormones (Fig. 2) (Erb and Kliebenstein 2020). Flavonoids and polyamines have long been known for their ability to influence gene expression and other physiological processes (Peer and Murphy 2007; Davies 2010; Pourcel et al. 2013). Glucosinolates comprise a phylogenetically restricted group of defense compounds, some of which exert effects on growth and stress resilience (Kim et al. 2015; Malinovsky et al. 2017; Salehin et al. 2019; Katz et al. 2020) (Fig. 2). N-hydroxy-pipecolic acid, β-aminobutyric acid, and indole have received increasing attention for their immunomodulatory properties (Erb et al. 2015; Cai and Aharoni 2022; Shields et al. 2022). Chemically diverse isoprenoids, including volatile terpenoids, have been intensively studied for their signaling roles in growth, defense, and interplant communication (Zuo et al. 2019; Bai et al. 2021; Ninkovic et al. 2021; Hu 2022; Wang and Erb 2022). Oxidative cleavage of carotenoids is also a rich source of metabolites with regulatory functions, beyond the well-characterized abscisic acid and strigolactone hormones (Wang et al. 2021). Examples include the promotion of leaf stress responses by β-cyclocitral (Dinneny et al. 2019) and β-ionone (Felemban et al. 2023), the mediation of plant rhizosphere interactions by zaxinone (Wang, Haider, et al. 2019), and the control of root architecture by anchorene (Jia et al. 2019). Oxygenated fatty acids (i.e. oxylipins) and their derivatives, including green leaf volatiles (Matsui 2006), azelaic acid (Jung et al. 2009), and 12-oxo-phytodienoic acid (OPDA) (Wang et al. 2020), are also recognized as signaling compounds. These examples highlight the regulatory properties of plant secondary metabolites and raise the central question: what is their mode of action?

Consistent with their biological activity, most of the above-mentioned bioactive compounds have been shown to modulate nuclear gene expression. Low-molecular-weight metabolites functionally interact with target proteins in well-defined ways, including as enzyme substrates and products, cofactors, allosteric modulators, and molecular glues (Fig. 3). In theory, each of the protein–metabolite interactions has the potential to alter physiological responses through changes in gene expression. Specific examples of how small molecules may regulate transcription include the binding of β-caryophyllene to a TOPLESS-like transcriptional corepressor (Nagashima et al. 2019), the activation of a NIN-like protein TF by nitrate (Liu et al. 2022), the increase in cytosolic Ca2+ levels by Z-3-hexenal (Aratani et al. 2023), the perception of β-aminobutyric acid by an aspartyl tRNA synthetase (Luna et al. 2014), and the modulation of reactive oxygen species homeostasis by flavonols (Pourcel et al. 2013; Sugimoto et al. 2022; Daryanavard et al. 2023) (Fig. 3). Regulatory metabolites may also alter gene expression networks by affecting the metabolism or action of known phytohormones. For example, auxin signaling and transport are influenced by certain flavonols (Brown et al. 2001; Peer and Murphy 2007), apocarotenoids (Jia et al. 2019), and glucosinolates (Katz et al. 2015, 2020). Priming of plant defense responses by various volatile signals is also mediated by phytohormones, most notably jasmonate and salicylic acid (Fig. 3) (Engelberth et al. 2004; Dani and Loreto 2022; Hu 2022; Li et al. 2022).

Low-molecular-weight metabolites interact with target proteins to modulate cellular processes that affect gene expression. A) Schematic diagram of an enzyme (E) interacting with its substrate (S) to generate a product (P). Enzymes often contain cofactors (c) to assist in catalysis and may also bind allosteric modulators (a) that positively or negatively regulate catalysis through conformational changes in the active site. B) Small molecules bind to proteinaceous receptors (R) and transmit information via 2 general mechanisms. First, and analogous to enzyme–metabolite interactions, allosteric-type ligands induce a conformational change in the receptor, which in turn propagates the signal (e.g. ABA, GA, and SL perception). Second, binding of molecular glue-type ligands creates a unique surface that interacts with a coreceptor (CoR) protein (e.g. IAA, JA, and BR perception). Many currently recognized phytohormone receptors possess (or are associated with) an enzymatic activity involved in signal transmission. Some plant receptors, including TIR1 and COI1, are associated with a cofactor (c) that is required for receptor structure and function. C) Examples of how gene expression may be modulated by the interactions of specific proteins or cellular processes with low-molecular-weight metabolites other than known phytohormones. See main text for details.
Figure 3.

Low-molecular-weight metabolites interact with target proteins to modulate cellular processes that affect gene expression. A) Schematic diagram of an enzyme (E) interacting with its substrate (S) to generate a product (P). Enzymes often contain cofactors (c) to assist in catalysis and may also bind allosteric modulators (a) that positively or negatively regulate catalysis through conformational changes in the active site. B) Small molecules bind to proteinaceous receptors (R) and transmit information via 2 general mechanisms. First, and analogous to enzyme–metabolite interactions, allosteric-type ligands induce a conformational change in the receptor, which in turn propagates the signal (e.g. ABA, GA, and SL perception). Second, binding of molecular glue-type ligands creates a unique surface that interacts with a coreceptor (CoR) protein (e.g. IAA, JA, and BR perception). Many currently recognized phytohormone receptors possess (or are associated with) an enzymatic activity involved in signal transmission. Some plant receptors, including TIR1 and COI1, are associated with a cofactor (c) that is required for receptor structure and function. C) Examples of how gene expression may be modulated by the interactions of specific proteins or cellular processes with low-molecular-weight metabolites other than known phytohormones. See main text for details.

It is also likely that the signaling properties of some secondary metabolites reflect dedicated perception systems that remain to be discovered. In their review of secondary metabolites that exhibit signaling properties, Erb and Kliebenstein (2020) concluded that these compounds affect growth and development by mechanisms that “are barely distinguishable from mechanisms normally assigned to plant hormones.” Although this may be true at a conceptual level, studies claiming to identify the direct target of a bioactive plant secondary metabolite are sparse. Additional work is clearly needed to define the physiological relevance and precise modes of action of these and other protein–metabolite interactions.

Lessons from jasmonate

The past century of plant hormone research, together with emerging technologies to elucidate broader networks of plant metabolite–protein interactions, provides a general road map for future efforts to identify and characterize regulatory metabolites. As a case study, we highlight the major experimental approaches and discoveries that led to the current understanding of jasmonate as a lipid-derived signal for promoting transcriptional flexibility. Detailed mechanistic knowledge of the jasmonate response pathway and its evolutionary origins offer valuable insights into how metabolites are recruited into roles that promote transcriptional pliability in dynamic environments. Although we draw mainly on studies performed with Arabidopsis as a model system, increasing attention to other model and nonmodel species has provided key advances into our understanding of jasmonate biology. Readers are referred to several recent reviews for a comprehensive discussion of jasmonate biosynthesis, signaling, and evolution (Chini et al. 2016; Wasternack and Song 2017; Howe et al. 2018; Wasternack and Feussner 2018; Wang, Wu, et al. 2019; Monte 2023).

Exogenous jasmonates promote diverse responses

The discovery of jasmonates (collectively referred to as JAs) dates back more than half a century to the identification of methyl-JA (MeJA) as a component of floral scent (Demole et al. 1962). This work laid the foundation for the chemical synthesis of JAs, which are widely used as additives in the food and fragrance industries (Wasternack 2015). Jasmonic acid (JA) was first identified as a plant growth inhibitory substance from cultures of the fungus Lasiodiplodia theobromae (Aldridge et al. 1971). This finding foreshadowed the current paradigm that pathogens often produce phytohormones or structurally related compounds to co-opt host plant immunity (Box 1) (Bari and Jones 2009). The availability of synthetic JAs facilitated “spray and pray” approaches to investigate the effects of exogenous compounds in a wide range of plant species. These studies provided evidence that JAs inhibit growth (Aldridge et al. 1971; Staswick et al. 1992), promote senescence (Ueda and Kato 1980), and attract insect pollinators (Baker et al. 1981). Subsequent studies showed that exogenous JAs also induce the accumulation of various stress-related compounds, including proteinase inhibitors (PIs), defensive secondary metabolites, pathogenesis-related proteins, and vegetative storage proteins (Farmer and Ryan 1990; Parthier 1990; Staswick 1990; Gundlach et al. 1992). Newly emerging molecular approaches further demonstrated that low concentrations of exogenous JAs provoke global changes in plant gene expression (Farmer et al. 1992; Sembdner and Parthier 1993; Creelman and Mullet 1997). This early era of jasmonate research established that exogenous JAs elicit growth- and stress-related responses by a mechanism that likely depends on changes in gene expression.

Quantification of endogenous jasmonates

The ability of any exogenous substance to elicit a biological response is insufficient to establish a role for that compound as a regulatory factor. Rather, it is necessary to demonstrate that specific physiological effects coincide with the spatial and temporal accumulation of the metabolite in plant tissues. The discovery that JAs have signaling properties prompted the development of methods to measure these compounds in plant tissues (Creelman and Mullet 1997). Analytical platforms for quantifying JAs range from high-performance liquid chromatography (Yamane et al. 1981) and qualitative radio immunolabeling (Knöfel et al. 1984) to quantitative GC-MS analyses (Müller et al. 2002) and current state-of-the-art UPLC-MS/MS approaches capable of detecting target metabolites in the pmol range (e.g. Balcke et al. 2012). The low abundance of JAs in plant tissues is consistent with the notion that phytohormones act at concentrations several orders of magnitude below that of most primary metabolites. Analysis using increasingly sensitive instruments revealed the chemical diversity of JAs and related oxylipins (i.e. oxygenated fatty acids) in a broad range of plant species (Creelman and Mullet 1997; Wasternack and Feussner 2018).

Elucidation of the jasmonate biosynthetic pathway

A complete understanding of regulatory metabolites requires knowledge of how and where the compound is produced in plant cells. A combination of biochemical and genetic approaches was particularly useful for elucidating the enzymatic reactions that convert polyunsaturated fatty acids to JA and the major receptor-active form of the phytohormone, jasmonoyl-L-isoleucine (JA-Ile) (Browse 2009) (Fig. 4). Classical biochemical fractionation and labeling experiments demonstrated that JA is produced from chloroplastic pools of α-linolenic acid (18:3) via lipoxygenase (LOX) enzymes (Vick and Zimmerman 1984). The elucidation of jasmonate biosynthesis has been extensively reviewed (Wasternack 2015), and the pathway is summarized in Fig. 4. The intracellular location of jasmonate biosynthetic enzymes was validated using immunocytochemical and cell biological approaches (Schaller and Weiler 1997; Wasternack and Feussner 2018). Activity-based purification of several jasmonate biosynthetic enzymes from diverse plant species facilitated the molecular cloning of the corresponding genes using information deduced from the amino acid sequences of the purified proteins (e.g. Song and Brash 1991). These studies confirmed the presence of plastidic or peroxisomal targeting sequences in the relevant enzymes. The development of plant genomic resources, including sequencing of the Arabidopsis genome (Arabidopsis Genome Initiative 2000), greatly facilitated efforts to identify genes involved in jasmonate biosynthesis and other aspects of jasmonate biology.

Major advances in elucidating the jasmonate pathway. A) Substantial milestones in jasmonate research are listed together with the number of jasmonate-related publications, based on the results of a 2023 PubMed search for “jasmonate.” Due to space constraints, not all milestones are included. See Wasternack (2015) for a comprehensive history of jasmonate research. B) Scheme of JA-Ile biosynthesis and signaling. See main text for details. 10,11-EOT, (11S)-10,11-epoxy-octadecatrienoic acid; 11-HPHT, 11(S)-hydroperoxy-hexadecatrienoic acid; 12,13-EOT, (13S)-12,13-epoxy-octadecatrienoic acid; 13-HPOT, 13(S)-hydroperoxy-octadecatrienoic acid; 13-LOX, 13-LIPOXYGENASE; 4,5-ddh-JA, 4,5-didehydro JA; AOC, ALLENE OXIDE CYCLASE; AOS, ALLENE OXIDE SYNTHASE; COI1, CORONATINE INSENSITIVE 1; dnOPDA, dinor-oxo-phytodienoic acid; JA, (+)-7-iso-jasmonic acid; JA-Ile, (+)-7-iso-jasmonoyl-isoleucine; JAZ, JAZMONATE ZIM DOMAIN; MED, MEDIATOR complex; NINJA, NOVEL INTERACTOR OF JAZ; OPC-4, 3-oxo-2-(2-(Z)-pentenyl cyclopentane-1-butanoic acid; OPC-6, 3-oxo-2-(2-(Z)-pentenyl)-cyclopentane-1-hexanoic acid); OPC-8, 3-oxo-2-(2-(Z)-pentenyl)-cyclopentane-1-octanoic acid; OPDA, 12-oxo-phytodienoic acid; OPR2, OPDA REDUCTASE 2; OPR3, OXO-PHYTODIENOIC ACID REDUCTASE 3; Pol II, RNA polymerase II; SCF, SKP1-CUL1-F-box; tnOPDA, tetranor-OPDA; TPL, TOPLESS; Ub, ubiquitin.
Figure 4.

Major advances in elucidating the jasmonate pathway. A) Substantial milestones in jasmonate research are listed together with the number of jasmonate-related publications, based on the results of a 2023 PubMed search for “jasmonate.” Due to space constraints, not all milestones are included. See Wasternack (2015) for a comprehensive history of jasmonate research. B) Scheme of JA-Ile biosynthesis and signaling. See main text for details. 10,11-EOT, (11S)-10,11-epoxy-octadecatrienoic acid; 11-HPHT, 11(S)-hydroperoxy-hexadecatrienoic acid; 12,13-EOT, (13S)-12,13-epoxy-octadecatrienoic acid; 13-HPOT, 13(S)-hydroperoxy-octadecatrienoic acid; 13-LOX, 13-LIPOXYGENASE; 4,5-ddh-JA, 4,5-didehydro JA; AOC, ALLENE OXIDE CYCLASE; AOS, ALLENE OXIDE SYNTHASE; COI1, CORONATINE INSENSITIVE 1; dnOPDA, dinor-oxo-phytodienoic acid; JA, (+)-7-iso-jasmonic acid; JA-Ile, (+)-7-iso-jasmonoyl-isoleucine; JAZ, JAZMONATE ZIM DOMAIN; MED, MEDIATOR complex; NINJA, NOVEL INTERACTOR OF JAZ; OPC-4, 3-oxo-2-(2-(Z)-pentenyl cyclopentane-1-butanoic acid; OPC-6, 3-oxo-2-(2-(Z)-pentenyl)-cyclopentane-1-hexanoic acid); OPC-8, 3-oxo-2-(2-(Z)-pentenyl)-cyclopentane-1-octanoic acid; OPDA, 12-oxo-phytodienoic acid; OPR2, OPDA REDUCTASE 2; OPR3, OXO-PHYTODIENOIC ACID REDUCTASE 3; Pol II, RNA polymerase II; SCF, SKP1-CUL1-F-box; tnOPDA, tetranor-OPDA; TPL, TOPLESS; Ub, ubiquitin.

Mutants galore

The importance of well-defined mutants (e.g. lines harboring null alleles) for elucidating the jasmonate pathway and understanding the functions of JAs cannot be overstated. Identification of jasmonate-deficient Arabidopsis mutants in forward genetic screens proved instrumental not only for validating early models of jasmonate biosynthesis but also for uncovering new steps in the pathway (Schaller and Weiler 1997; Browse 2009; Wasternack and Feussner 2018). This approach was facilitated by the easily recognizable phenotypes of JA deficiency in Arabidopsis, including male sterility and increased susceptibility to pathogens and insects. The first jasmonate biosynthetic mutant to be identified was jasmonic acid resistant 1 (jar1), which displays insensitivity to JA- and MeJA-inhibited root growth (Staswick et al. 1992). Because these JAs strongly elicit responses in wild-type plants, the most straightforward interpretation at the time was that jar1 is a jasmonate perception mutant. However, map-based cloning of JAR1 showed that it encodes a JA-conjugating enzyme responsible for the production of JA-Ile, indicating that jar1 is in fact a biosynthetic mutant that fails to produce a positive regulator (i.e. JA-Ile) of jasmonate responses (Staswick et al. 2002; Staswick and Tiryaki 2004). JAR1 is a member of the GH3 family of acyl acid amido synthetases, which conjugate amino acids to many phytohormones and likely other regulatory metabolites as well (Staswick and Tiryaki 2004; Westfall et al. 2012). Interestingly, a natural sequence polymorphism abrogating the expression of JAR1 accounts for much of the variation in induced defense responses within wild populations of Nicotiana attenuata (Ray et al. 2023).

Another key advance in understanding the biosynthesis and function of JAs was the discovery that severe JA deficiency in Arabidopsis results in male sterility due to defects in pollen development and anther dehiscence (Feys et al. 1994; Browse 2009). The reversibility of this phenotype by exogenous JA was first documented in a fatty acid desaturase triple mutant (fad3 fad7 fad8) that fails to produce trienoic fatty acid precursors of JA (McConn and Browse 1996). The application of JA to flower buds of the fad3 fad7 fad8 mutant recovered its fertility, demonstrating that the conversion of 18:3 to JA is essential for normal flower development. Subsequent screens for male-sterile mutants that could be chemically complemented by JA treatment led to the isolation of additional biosynthetic mutants, including those impaired in ALLENE OXIDE SYNTHASE, 12-OXO-PHYTODIENOIC ACID REDUCTASE 3 (OPR3), and DEFECTIVE IN ANTHER DEHISCENCE1 (DAD1) (reviewed by Browse 2009) (Fig. 4). The latter enzyme is a phospholipase A1 that acts in floral tissues to produce fatty acid precursors of JA from plastidial galactolipids (Ishiguro et al. 2001). In parallel with studies on the role of JAs in male fertility, the availability of JA biosynthetic mutants was instrumental in establishing a causal link between JAs and resistance to a broad spectrum of biotic agents, including arthropod herbivores and necrotrophic pathogens (Glazebrook 2005; Howe and Jander 2008; Bari and Jones 2009; Browse 2009).

Functional analysis of genes encoding biosynthetic enzymes and other components of the jasmonate pathway was advanced through the use of reverse genetic approaches that leveraged newly developed molecular tools, most notably the sequenced Arabidopsis genome (Arabidopsis Genome Initiative 2000) and sequence-indexed insertional mutants (Alonso et al. 2003). These resources were particularly important for the characterization of jasmonate biosynthetic enzymes encoded by multigene families, as functional redundancy among paralogs often precludes the identification of mutants in forward genetic screens (Schilmiller et al. 2007; Delfin et al. 2022). The 13-LOX gene family in Arabidopsis exemplifies the use of reverse genetics in helping to decipher the subfunctionalization of specific family members. All 4 13-LOX enzymes (LOX2, LOX3, LOX4, and LOX6) in Arabidopsis oxygenate 18:3 and contribute to JA production, likely explaining why Arabidopsis 13-lox mutants have not been identified in forward genetic screens. Detailed analysis of single and multiorder lox mutants showed that each paralog preferentially acts in different tissues and cell types or is involved in other aspects of oxylipin biosynthesis (Bannenberg et al. 2009; Caldelari et al. 2011; Chauvin et al. 2013; Grebner et al. 2013; Mochizuki et al. 2016). Reverse genetic approaches were also useful for uncovering the existence of alternative pathways that operate in the absence of the major JA biosynthetic route. For example, in mutants lacking OPR3, OPDA can be converted to 4,5-di-dehydro-JA in the peroxisome and then reduced to JA in the cytosol by OPR2 (Fig. 4) (Chini et al. 2018; Howe 2018). Although OPR2 contributes weakly to JA production in Arabidopsis, recent evidence suggests that this alternative biosynthetic route predated the OPR3-dependent pathway in the evolution of jasmonate biosynthesis (Chini et al. 2023).

A satisfactory understanding of phytohormones and other regulatory metabolites depends on knowledge obtained from diverse plant species. In general, the roles of JAs in promoting senescence, growth inhibition, and broad-spectrum resistance to herbivores and pathogens are broadly conserved across land plants (Monte et al. 2019). However, whereas JAs appear to be essential for normal reproductive development in all plants examined to date, defects in the JA pathway often give rise to reproductive phenotypes that are specific to a particular species or lineage. The rice extra glume1 (eg1) mutant, which is defective in a DAD1 homolog, is impaired in early stages of spikelet and floret development but show normal stamen and pollen maturation (Cai et al. 2014). Disruption of JA biosynthesis in the maize (Zea mays) sex determination mutant tasselseed 1 (ts1) results in feminization of the male inflorescence (tassel) but does not affect early spikelet and floret development (Acosta et al. 2009). The lack of jasmonate responsiveness in the tomato (Solanum lycopersicum) jasmonic acid-insensitive 1 (jai1) mutant impairs the maternal control of seed development while decreasing but not eliminating pollen viability (Li et al. 2004; Goetz et al. 2012). These examples illustrate how the study of mutants in diverse plants adds important perspectives to our understanding of jasmonate biology.

Regulation of jasmonate biosynthesis and catabolism

A substantial advance toward understanding the regulation of jasmonate biosynthesis was the discovery that various stress-related treatments result in the rapid accumulation of endogenous JAs. Examples of eliciting stimuli include the treatment of cultured plant cells with fungal cell wall fragments (Gundlach et al. 1992) and mechanical wounding of leaves (Creelman et al. 1992; Farmer and Ryan 1992). The latter observation linked groundbreaking work on wound-inducible defensive PIs in tomato to the discovery that exogenous JAs rapidly activate the expression of PI-encoding genes (Green and Ryan 1972; Farmer and Ryan 1990; Farmer et al. 1992). A model in which the perception of wounding and other stress-related signals at the cell surface triggers the production of JA and concomitant defense responses has garnered strong support over the past 3 decades (Farmer and Ryan 1992; Howe 2019). These early insights into the roles of JAs in inducing the expression of genes encoding defensive compounds and their accumulation opened up new avenues of investigation into how this lipid-based signaling pathway controls plant immunity (Glazebrook 2005; Karban and Baldwin 2007; Campos et al. 2014; Wang, Wu, et al. 2019).

Owing to its simplicity and effectiveness in inducing jasmonate biosynthesis, mechanical wounding has been used extensively to study the dynamics of jasmonate responses within and between different plant tissues. In general, JA and JA-Ile levels increase within minutes of tissue damage and peak within approximately 1 h (Table 1) (Baldwin et al. 1997; McConn and Browse 1996; Glauser et al. 2009; Koo et al. 2009). Basal levels of OPDA are approximately 500–600 pmol/g fresh weight (FW) in undamaged Arabidopsis leaves and 30–40 pmol/g FW in roots and increase 40-fold in leaves and 6-fold in roots after leaf wounding (Table 1). Although basal JA levels are lower than basal OPDA levels and are often below the limit of quantification in undamaged tissues, JA levels can increase up to 500-fold after mechanical wounding. Peak levels of wound-induced JA-Ile are typically much less (e.g. 10-fold) than peak levels of JA in wounded tissues (Table 1). This observation is consistent with the “prohormone” concept in which JA availability is a limiting factor for the JAR1-catalyzed formation of JA-Ile (Staswick and Tiryaki 2004; Acosta and Farmer 2010; Koo 2018). The large differences in the relative abundances of JA biosynthetic precursors and intermediates (18:3 > OPDA > JA > JA-Ile) also illustrate how the tightly regulated biosynthesis of these regulatory metabolites depends on an appropriate supply of precursors along the entire biosynthetic route (Table 1).

Table 1.

Distribution of jasmonates in Arabidopsis

TissueConditionOPDAaJAaJA-IleaReferences
LeavesResting600<10<2Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
Leaves1 h post leaf wounding25,0007,000400Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
RootsResting35<10<2Schulze et al. 2019
Roots1 h Post leaf wounding1006060Schulze et al. 2019
TissueConditionOPDAaJAaJA-IleaReferences
LeavesResting600<10<2Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
Leaves1 h post leaf wounding25,0007,000400Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
RootsResting35<10<2Schulze et al. 2019
Roots1 h Post leaf wounding1006060Schulze et al. 2019

aMetabolite levels are in pmol/g FW tissue.

Table 1.

Distribution of jasmonates in Arabidopsis

TissueConditionOPDAaJAaJA-IleaReferences
LeavesResting600<10<2Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
Leaves1 h post leaf wounding25,0007,000400Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
RootsResting35<10<2Schulze et al. 2019
Roots1 h Post leaf wounding1006060Schulze et al. 2019
TissueConditionOPDAaJAaJA-IleaReferences
LeavesResting600<10<2Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
Leaves1 h post leaf wounding25,0007,000400Schulze et al. (2019), Reymond et al. (2000), Koo et al. (2009)
RootsResting35<10<2Schulze et al. 2019
Roots1 h Post leaf wounding1006060Schulze et al. 2019

aMetabolite levels are in pmol/g FW tissue.

Precise control over the spatial and temporal accumulation of phytohormones depends not only on biosynthetic enzymes but also on factors involved in metabolite transport and catabolism (Li et al. 2021). In the case of JAs, several metabolic transformations are crucial for preventing the overactivation of costly jasmonate-mediated stress responses (Koo and Howe 2009; Heitz et al. 2016; Wasternack and Feussner 2018). For example, cytosolic pools of JA are depleted via the combined action of highly specific JA hydroxylases and sulfotransferases (Caarls et al. 2017; Smirnova et al. 2017; Fernández-Milmanda et al. 2020). JA-Ile itself is inactivated both by hydrolysis of the amide bond (Woldemariam et al. 2012; Widemann et al. 2013; Koo 2018) and by cytochrome P450 enzymes that oxidize the pentenyl side chain to generate the corresponding hydroxy and carboxy derivatives (Koo et al. 2011, 2014; Heitz et al. 2012; Aubert et al. 2015). The finding that JA induces the expression of genes encoding these catabolic enzymes is consistent with the notion of negative feedback regulation of JA-Ile levels (Heitz et al. 2016; Koo 2018). Notably, the rapid inducibility of these and other jasmonate-related genes has been exploited through “guilt-by-associated” coexpression approaches to identify components of the jasmonate response pathway (e.g. Koo et al. 2011).

Homeostasis of plant regulatory metabolites also involves processes that control the systemic production of the chemical signal. The jasmonate pathway provides an excellent model to address this question (Gao et al. 2023; Morin et al. 2023). JA and JA-Ile levels increase rapidly not only locally at the site of injury but also in distal unwounded leaves (Glauser et al. 2009; Koo et al. 2009) and roots (Gasperini et al. 2015). Leaf injury provoked by mechanical wounding or insect herbivory stimulates the long-distance translocation of OPDA and JA to mount defense responses in distal organs (Schulze et al. 2019; Li et al. 2020). Concomitant with this burst of JAs is the rapid propagation of slow wave electrical potentials governed by GLUTAMATE RECEPTOR-LIKE (GLR) proteins (Mousavi et al. 2013; Nguyen et al. 2018). Wounding also stimulates the release of extracellular elicitors such as Glu, which binds to specific GLRs to regulate long-distance Ca2+ waves (Toyota et al. 2018; Alfieri et al. 2020; Bellandi et al. 2022). Recent studies in Arabidopsis further showed that tissue injury results in the generation of aglucone elicitors of the GLR-dependent propagation of slow wave potentials via glucohydrolase enzymes (Gao et al. 2023). Consistent with these findings, mutants defective in specific GLRs (e.g. GLR3.3 and GLR3.6) are impaired in wound-induced slow wave potentials, long-distance increases in cytosolic Ca2+ and jasmonate accumulation in distal leaves (Mousavi et al. 2013; Nguyen et al. 2018; Toyota et al. 2018). Specific DAD1-like lipases appear to link wounding-responsive electrical signaling to the activation of JA biosynthesis (Morin et al. 2023).

Elucidation of the core jasmonate signaling pathway

By the mid-1990s, it was becoming increasingly evident that the potent effects of JAs on gene expression involve a specific mechanism that links the perception of these compounds to transcriptional reprogramming (Parthier 1990; Farmer and Ryan 1992; Wasternack and Parthier 1997). However, in the absence of any clues into how JAs function, which experimental approaches were best suited to tackle this issue? Once again, mutants identified in genetic screens provided the way forward. The first major advance in understanding the mode of action of JAs came from the identification of the Arabidopsis coronatine insensitive 1 (coi1) mutant. This mutant fails to respond to coronatine, which is a toxin produced by some pathogenic strains of Pseudomonas syringae (Feys et al. 1994). As a structural mimic of JA-Ile, exogenous coronatine elicits physiological responses that are very similar to those induced by exogenous JAs. The finding that coi1 mutants were insensitive to both coronatine and JA indicated that these compounds signal via a similar mechanism (Feys et al. 1994). Importantly, coi1 mutants are also male sterile and highly susceptible to attack by necrotrophic pathogens and herbivores (Feys et al. 1994; Thomma et al. 1998; Glazebrook 2005). The phenotypic similarity between JA biosynthesis and coi1 perception mutants provided compelling genetic evidence that endogenous JAs activate a COI1-dependent signaling pathway leading to the expression of jasmonate-responsive genes.

Map-based cloning studies revealed that COI1 encodes a member of the large (>500 in Arabidopsis) F-box family of proteins (Xie et al. 1998). F-box proteins act as specificity determinants in SCF (Skp1/Cullin/F-box)-type E3 ubiquitin ligase complexes that target specific substrates for ubiquitylation and proteolytic destruction (Shabek and Zheng 2014). Accordingly, it was suggested that COI1 recruits a negative regulator of jasmonate responses for targeted degradation, which in turn promotes fertility and defense responses (Xie et al. 1998). This model is supported by the sequence similarity between COI1 and the closely related F-box protein TRANSPORT INHIBITOR RESPONSE 1 (TIR1) and by emerging evidence that TIR1 is a central component of the auxin perception apparatus (Ruegger et al. 1998). An intensive search for COI1 substrates included genetic screens to identify additional regulators of the pathway (Ellis and Turner 2001; Hilpert et al. 2001; Xu et al. 2001; Jensen et al. 2002; Lorenzo et al. 2004), identification of COI1-interacting proteins (Devoto et al. 2002) and cataloging of JA-responsive genes that are expressed in a COI1-dependent manner (Devoto et al. 2005). Despite these efforts, viable candidate COI1 substrates were not identified, and progress toward understanding the molecular mechanism of jasmonate signaling was at an impasse for nearly a decade (Browse 2009). One notable discovery during this period was the identification of jasmonate-insensitive mutants that are defective in a basic helix–loop–helix-type TF called MYC2 (Berger et al. 1996; Boter et al. 2004; Lorenzo et al. 2004). Although MYC2 and its closely related proteins (e.g. MYC3 and MYC4 in Arabidopsis) are regarded as master positive regulators of jasmonate responses (Fernández-Calvo et al. 2011; Kazan and Manners 2013), a satisfactory model to explain how MYC activity is linked to jasmonate and COI1 failed to emerge at this time.

In 2007, the discovery of JASMONATE ZIM DOMAIN (JAZ) proteins as the elusive COI1 substrates was achieved through the analysis of gene coexpression modules (Thines et al. 2007; Yan et al. 2007) and positional cloning of a mutated gene that confers dominant insensitivity to JA (Chini et al. 2007). Thines and coworkers found that 8 related genes encoding uncharacterized proteins containing a ZINC-FINGER PROTEIN EXPRESSED IN INFLORESCENCE MERISTEM (ZIM) motif were rapidly induced by exogenous jasmonate (Thines et al. 2007). Similarly, genome-wide expression profiling showed that several wound-induced transcripts encoding these ZIM domain-containing proteins accumulated in a JA-dependent manner (Yan et al. 2007). Finally, the causative mutation underlying the JA-insensitive jai3-1 mutant phenotype was shown to reside in a chromosomal region that contains a member (JAZ3) of the same gene family (Chini et al. 2007). This plant-specific gene family was named JAZ, with 13 recognized members in Arabidopsis (Chini et al. 2007; Thines et al. 2007; Yan et al. 2007; Thireault et al. 2015). The size of the rapidly evolving JAZ family ranges from one in the common liverwort (Marchantia polymorpha) to more than 20 members in many flowering plants (Howe et al. 2018; Monte et al. 2019).

Three key findings associated with the discovery of JAZ helped to fill in the remaining major gaps in understanding the core mechanism of jasmonate signaling. First, mutated JAZ proteins that fail to interact with COI1 strongly repress the expression of jasmonate-responsive genes (Chini et al. 2007; Thines et al. 2007; Yan et al. 2007). Second, JAZ proteins exert repression by binding directly to MYC TFs (Chini et al. 2007) and many other TFs as well (Fernández-Calvo et al. 2011; Chini et al. 2016; Howe et al. 2018). Finally, the physical interaction between COI1 and JAZ is promoted by JA-Ile but not its metabolic precursors or derivatives, including OPDA, JA, and MeJA; it was therefore concluded that a COI1–JAZ complex constitutes the site of JA-Ile perception (Thines et al. 2007). These collective findings provided a basis for a model (Fig. 4) in which MYC-dependent responses are repressed by JAZ in the absence of JA-Ile. In response to stress-related or developmental cues that stimulate JA-Ile accumulation, JA-Ile promotes the SCFCOI1-dependent turnover of JAZ and subsequent expression of jasmonate-responsive genes.

Follow-up studies showed that coronatine also promotes the formation of COI1–JAZ complexes and that the C-terminal Jas motif of JAZ harbors a “degron” sequence that mediates the hormone-dependent COI1–JAZ interaction (Katsir et al. 2008; Melotto et al. 2008). Mutation of the degron motif stabilizes JAZ against hormone-dependent degradation, resulting in dominant repression of jasmonate responses (Chini et al. 2007; Thines et al. 2007; Yan et al. 2007; Melotto et al. 2008). Biochemical and structural analyses further showed that the COI1–JAZ interaction is mediated specifically by (+)-7-iso-JA-Ile, which is the major JA-Ile stereoisomer in plant tissues (Fonseca et al. 2009; Sheard et al. 2010). In validating and extending the pioneering work on JAR1 (Staswick and Tiryaki 2004), it is now evident that jasmonate responses elicited by exogenous JA and MeJA require metabolic transformation of these compounds to (+)-7-iso-JA-Ile in plant tissues.

Structural insights into jasmonate perception and signaling

The discovery that COI1–JAZ complexes are a high-affinity receptor for JA-Ile (Thines et al. 2007) was aided by the molecular glue paradigm, which first emerged from the study of auxin perception (Dharmasiri et al. 2003; Kepinski and Leyser 2005; Tan et al. 2007). The 3-dimensional structure of the ligand-bound COI1–JAZ ternary complex shows that an inositol pyrophosphate (likely InsP8) serves as a COI1 structural cofactor and that the JAZ degron simultaneously interacts with both the bound hormone and COI1 (Fonseca et al. 2009; Sheard et al. 2010; Laha et al. 2015). This model of COI1–JAZ as a high-affinity coreceptor does not exclude the possibility that JA-Ile first binds to COI1 with low affinity, followed by the formation of a high-affinity COI1–JAZ complex. This 2-step model of JA-Ile perception has received experimental support by showing that coronatine weakly interacts (i.e. low micromolar affinity) with COI1 in the absence of JAZ (Yan et al. 2018). Similar results were recently reported for TIR1, which can bind auxin at low micromolar affinity in the absence of an Aux/IAA substrate (Cao et al. 2022). These collective findings support a scenario in which molecular glue-type hormones were recruited as signals based on their ability to promote protein–protein interactions (Fig. 3).

The structural elucidation of MYC3 binding to the Jas domain of JAZ9 further clarified the molecular mechanism by which JAZ proteins repress MYC TFs (Zhang et al. 2015). In the absence of the hormone, the Jas motif assumes an α-helical structure that intercalates within the MYC N-terminal fold, thereby inhibiting the interaction of MYC with the MED25 subunit of the mediator complex (Çevik et al. 2012; Chen et al. 2012; Zhang et al. 2015). JAZ proteins recruit transcriptional corepressors of the TOPLESS family via the adaptor protein NINJA (Pauwels et al. 2010) and, in some cases, directly interact with TOPLESS through an EAR motif (Kagale et al. 2010; Shyu et al. 2012). JA-Ile-induced JAZ degradation releases MYCs from repression while simultaneously facilitating the recruitment of coactivators such as MED25 (Zhang et al. 2015; An et al. 2017) (Fig. 4). Thus, JAZ proteins perform dual functions as transcriptional repressors in the absence of JA-Ile and as coreceptors in the presence of the hormone. Understanding the structural and dynamic behavior of JAZ-associated transcriptional complexes represents an important future challenge (Zhang et al. 2017; Zhou et al. 2023). The prolific ability of JAZs to interact with proteins in other signaling pathways (Robert-Seilaniantz et al. 2011; Song et al. 2014; Berens et al. 2017; Howe et al. 2018; Zhou et al. 2023) provides a window into how regulatory metabolites connect pathways to reconfigure transcriptional programs in a flexible and dynamic manner.

Multiple layers of repression

A recurring theme in hormone biology is the existence of mechanisms to attenuate the amplitude and duration of responses through negative feedback control (Fig. 1). In the case of jasmonate, overactivation of the pathway results in stunted growth, decreased fecundity, loss of photosynthetic capacity, and, in its extreme, cell death (Baldwin 1998; Yan et al. 2007; Zhang and Turner 2008; Guo et al. 2018, 2022). Among the major mechanisms to attenuate these costly responses are the catabolism of bioactive JAs (see above) and de novo synthesis of JAZ repressors. The latter mechanism is based on the observation that JAZ genes are rapidly and strongly expressed by MYC2 in response to jasmonate (Chini et al. 2007; Thines et al. 2007; Chung et al. 2008; Moreno et al. 2013; Zander et al. 2020). Effective JAZ-mediated repression relies on the recruitment of various corepressor and adaptor proteins, including NINJA, TOPLESS, and epigenetic modifying enzymes (Pauwels et al. 2010; Howe et al. 2018; Zander 2021). JAZ stabilization resulting from alternative splicing events that eliminate or weaken COI1 interactions, as well as jasmonate-induced expression of factors that compete with MYCs for DNA binding, represent additional layers of negative feedback control (Chung et al. 2010; Nakata et al. 2013; Fonseca et al. 2014; Qi et al. 2015; Zhang et al. 2017; Wu et al. 2020). These many layers of signal attenuation highlight the negative fitness consequences of overactivating jasmonate responses.

Evolution of the jasmonate pathway

Functional phylogenomic approaches based on recently sequenced streptophyte (charophyte algae and land plants) genomes have provided important new insights into processes that shaped the emergence of metabolic and signaling pathways during the colonization of land by plants (Maeda and Fernie 2021; Bowman 2022; Rieseberg et al. 2023). Recent studies on the evolution of JAs have contributed to these efforts by revealing lineage-specific innovations in both JA biosynthetic and signaling pathways. Genome analysis revealed that bryophyte lineages (hornworts, liverworts, and mosses) contain orthologs for all core signaling JA components, including COI1, JAZ, NINJA, TPL, and MYC (Bowman et al. 2017; Monte et al. 2018). Functional analyses of the COI1-JAZ-MYC signaling module showed that the jasmonate pathway likely emerged in the common ancestor of land plants more than 450 million years ago (Bowman et al. 2017; Peñuelas et al. 2019; Chini et al. 2023). Interestingly, however, bryophytes neither synthesize nor respond to JA-Ile, which is the receptor-active form of the hormone in vascular plants. Rather, studies with M. polymorpha (Mp) and the lycophyte Huperzia selago identified several isomers of dnOPDA (dn-cis-OPDA and dn-iso-OPDA) as active ligands in these species (Monte et al. 2018; Kneeshaw et al. 2022). The selectivity of the MpCOI1–JAZ coreceptor for dnOPDA ligands was attributed to a single amino acid change in MpCOI1 that alters the size of the ligand-binding pocket, as well as specific residues in the Jas motif of MpJAZ. A broad survey of the green lineage recently confirmed and extended these insights into the evolutionary trajectory of JAs (Chini et al. 2023).

Functional phylogenomic studies have also provided insight into the extent to which the physiological functions of JAs are conserved in land plants. Transcriptional changes in M. polymorpha induced by treatment with dnOPDA are similar to transcriptional changes elicited by JA-Ile in flowering plants, including the induction of genes related to secondary metabolism and wounding responses (Devoto et al. 2005; Chini et al. 2018). However, whereas JA-Ile promotes defense and inhibits growth and fertility in Arabidopsis, dnOPDA appears to regulate defense and growth responses but not fertility in M. polymorpha (Monte et al. 2018). Given the reduced fertility of OPDA-deficient mutants of the moss Physcomitrium patens (Stumpe et al. 2010), additional work is needed to better understand the evolutionary basis of the roles of JAs in plant reproductive processes.

It is noteworthy that cyclopentenone JAs such as OPDA and dnOPDA are reactive electrophilic compounds that exhibit signaling properties independently of COI1. A recent study provided evidence that these compounds signal through a COI1-independent pathway to promote thermotolerance in diverse species, including Arabidopsis, M. polymorpha, and the charophyte alga Klebsormidium nitens (Monte et al. 2020). These findings suggest an ancestral role for JAs in regulating protection against heat stress, which was likely important for plant colonization of terrestrial habitats. Elucidation of the COI1-independent signaling pathway by which cyclopentanone JAs promote thermotolerance promises to provide new insight into how the larger family of lipid-derived JAs were recruited into distinct signaling pathways during plant evolution.

Concluding remarks and future perspectives

The extraordinary diversity of plant metabolism arises from adaptive processes such as the neofunctionalization of duplicated enzyme-encoding genes and increased enzyme promiscuity (Chae et al. 2014; Weng et al. 2021). As a consequence, hundreds of thousands of secondary metabolites constitute an evolving chemical “library” from which to recruit small molecules into roles that promote plant fitness in dynamic and often harsh environments. This process is exemplified by small-molecule plant hormones, which have acquired the ability to modulate cellular processes needed to help the plant adapt its growth and development to changing conditions. Currently recognized phytohormones appear to be part of a larger and more diverse group of chemical mediators whose action is linked in part to transcriptional control. Lessons from the study of the jasmonate pathway and other phytohormones provide a guide to drive future progress in unveiling the physiological roles of regulatory metabolites, their modes of perception and signaling, and their evolutionary history. Current knowledge gaps (see Outstanding Questions) pose exciting challenges for the future of plant science, which may help to inform efforts to modify signaling pathways for crop improvement.

A major bottleneck in understanding how plant metabolites influence gene expression and associated traits is the lack of insight into the direct targets of these compounds. The vast majority of small-molecule receptors in plants were identified through the use of mutants that are unresponsive to a defined eliciting cue, including (pro)hormones and hormone agonists. Following the identification of such mutants, molecular genetic techniques are employed to track down the causal mutation. The strength of this approach lies in linking a specific gene to a robust biological response, as well as its application to any bioactive compound whose chemical identity is known. Moreover, advances in genome-sequencing technology have greatly accelerated the process of identifying the mutated gene underlying a phenotype of interest (e.g. Schneeberger et al. 2009). Limitations of the classical genetic approach include the labor intensiveness of genetic screens and the possibility that perception mutants are difficult to identify due to functional redundancy within receptor-encoding gene families.

These considerations justify future research to systemically investigate regulatory metabolite–protein interactions in plants. Such efforts have historically been hampered by technical challenges related to the transient, low-affinity nature of these bimolecular associations. Nevertheless, recent advances in mass spectrometry have facilitated the development of powerful approaches to identify biologically relevant protein–metabolite interactions in various model organisms (Piazza et al. 2018; Luzarowski and Skirycz 2019). This technology can be applied in either a non-targeted or targeted manner (i.e. identification of proteins that bind to a specific metabolite) and is adaptable for use in plants (Schlossarek et al. 2023). Improvements in molecular modeling and other computational approaches for predicting protein–metabolite interactions, which are widely used for drug discovery (Lavecchia 2015), are also expected to be employed to identify candidate protein targets of plant regulatory metabolites. In the coming years, there is much to look forward to as these new technologies are combined with robust genetic approaches to delve deeper into the expanding universe of plant regulatory metabolites.

OUTSTANDING QUESTIONS
  • What are the direct targets of the myriad volatile and nonvolatile plant secondary metabolites that exhibit signaling attributes?

  • To what extent can untargeted mass spectrometry-based and predictive approaches be used to systematically identify protein–metabolite interactions with a wide range of affinities?

  • How do evolutionary forces recruit secondary metabolites into signaling pathways?

  • Beyond the small number of plant hormones known to act in the nucleus, to what extent do small molecules interact with and modulate the activities of transcriptional regulators?

  • How can knowledge of regulatory metabolites be used to develop crop varieties that are more productive, nutritious, and better adapted to changing climate conditions?

Acknowledgments

We thank Erich Grotewold, Danve Castroverde, and anonymous reviewers for helpful comments on the manuscript. We gratefully acknowledge current and past laboratory members for their many contributions to jasmonate research over the years. We also acknowledge the influence of the stimulating research environments in our respective institutions (Leibniz Institute of Plant Biochemistry and MSU-DOE Plant Research Laboratory), where there is a strong legacy of phytohormone research. G.A.H. thanks Hans Kende and Jan Zeevaart for inspiration in the early stage of his career, and the Fulbright U.S. Scholar program for the opportunity to discuss aspects of this work with colleagues in Japan. The graphical abstract and Figs. 1 and 3 were created with BioRender.com.

Author contributions

G.A.H. conceived the structure of the review; D.G. and G.A.H. wrote the paper.

Funding

This work was supported by the Chemical Sciences, Geosciences, and Biosciences Division, Basic Energy Sciences, Office of Science at the U.S. Department of Energy (through grant DE-FG02–91ER20021) and the Michigan AgBioResearch Project MICL02278 to G.A.H. and by the IPB core funding from the Leibniz Association, the German Research Foundation (grants GRK2498/400681449 and GA 2419/3-1) and the European Union (ERC, MECHANOJAS, 101088876) to D.G.

Dive Curated Terms

The following phenotypic, genotypic, and functional terms are of significance to the work described in this paper:

References

Acosta
IF
,
Farmer
EE
.
Jasmonates
.
Arabidopsis Book
.
2010
:
8
:
e0129
. https://doi.org/10.1199/tab.0129

Acosta
IF
,
Laparra
H
,
Romero
SP
,
Schmelz
E
,
Hamberg
M
,
Mottinger
JP
,
Moreno
MA
,
Dellaporta
SL
.
Tasselseed1 is a lipoxygenase affecting jasmonic acid signaling in sex determination of maize
.
Science
.
2009
:
323
(
5911
):
262
265
. https://doi.org/10.1126/science.1164645

Aldridge
DC
,
Galt
S
,
Giles
D
,
Turner
WB
.
Metabolites of Lasiodiplodia theobromae
.
J Chem Soc
.
1971
:
1971
:
1623
1627
. https://doi.org/10.1039/j39710001623

Alfieri
A
,
Doccula
FG
,
Pederzoli
R
,
Grenzi
M
,
Bonza
MC
,
Luoni
L
,
Candeo
A
,
Romano Armada
N
,
Barbiroli
A
,
Valentini
G
, et al.
The structural bases for agonist diversity in an Arabidopsis thaliana glutamate receptor-like channel
.
Proc Natl Acad Sci U S A
.
2020
:
117
(
1
):
752
760
. https://doi.org/10.1073/pnas.1905142117

Alonso
JM
,
Stepanova
AN
,
Leisse
TJ
,
Kim
CJ
,
Chen
H
,
Shinn
P
,
Stevenson
DK
,
Zimmerman
J
,
Barajas
P
,
Cheuk
R
, et al.
Genome-wide insertional mutagenesis of Arabidopsis thaliana
.
Science
.
2003
:
301
(
5633
):
653
657
. https://doi.org/10.1126/science.1086391

Altmann
M
,
Altmann
S
,
Rodriguez
PA
,
Weller
B
,
Elorduy Vergara
L
,
Palme
J
,
Marín-de la Rosa
N
,
Sauer
M
,
Wenig
M
,
Villaécija-Aguilar
JA
, et al.
Extensive signal integration by the phytohormone protein network
.
Nature
.
2020
:
583
(
7815
):
271
276
. https://doi.org/10.1038/s41586-020-2460-0

An
C
,
Li
L
,
Zhai
Q
,
You
Y
,
Deng
L
,
Wu
F
,
Chen
R
,
Jiang
H
,
Wang
H
,
Chen
Q
, et al.
Mediator subunit MED25 links the jasmonate receptor to transcriptionally active chromatin
.
Proc Natl Acad Sci U S A
.
2017
:
114
(
42
):
E8930
E8939
. https://doi.org/10.1073/pnas.1710885114

Anastasiou
E
,
Kenz
S
,
Gerstung
M
,
MacLean
D
,
Timmer
J
,
Fleck
C
,
Lenhard
M
.
Control of plant organ size by KLUH/CYP78A5-dependent intercellular signaling
.
Dev Cell
.
2007
:
13
(
6
):
843
856
. https://doi.org/10.1016/j.devcel.2007.10.001

Arabidopsis Genome Initiative
.
Analysis of the genome sequence of the flowering plant Arabidopsis thaliana
.
Nature
.
2000
:
408
(
6814
):
796
815
. https://doi.org/10.1038/35048692

Aratani
Y
,
Uemura
T
,
Hagihara
T
,
Matsui
K
,
Toyota
M
.
Green leaf volatile sensory calcium transduction in Arabidopsis
.
Nat Commun
.
2023
:
14
(
1
):
6236
. https://doi.org/10.1038/s41467-023-41589-9

Aubert
Y
,
Widemann
E
,
Miesch
L
,
Pinot
F
,
Heitz
T
.
CYP94-mediated jasmonoyl-isoleucine hormone oxidation shapes jasmonate profiles and attenuates defence responses to Botrytis cinerea infection
.
J Exp Bot
.
2015
:
66
(
13
):
3879
3892
. https://doi.org/10.1093/jxb/erv190

Bai
Y
,
Fernández-Calvo
P
,
Ritter
A
,
Huang
AC
,
Morales-Herrera
S
,
Bicalho
KU
,
Karady
M
,
Pauwels
L
,
Buyst
D
,
Njo
M
, et al.
Modulation of Arabidopsis root growth by specialized triterpenes
.
New Phytol
.
2021
:
230
(
1
):
228
243
. https://doi.org/10.1111/nph.17144

Bailey-Serres
J
,
Parker
JE
,
Ainsworth
EA
,
Oldroyd
GED
,
Schroeder
JI
.
Genetic strategies for improving crop yields
.
Nature
.
2019
:
575
(
7781
):
109
118
. https://doi.org/10.1038/s41586-019-1679-0

Baker
TC
,
Nishida
R
,
Roelofs
WL
.
Close-range attraction of female oriental fruit moths to herbal scent of male hairpencils
.
Science
.
1981
:
214
(
4527
):
1359
1361
. https://doi.org/10.1126/science.214.4527.1359

Balcke
GU
,
Handrick
V
,
Bergau
N
,
Fichtner
M
,
Henning
A
,
Stellmach
H
,
Tissier
A
,
Hause
B
,
Frolov
A
.
An UPLC-MS/MS method for highly sensitive high-throughput analysis of phytohormones in plant tissues
.
Plant Methods
.
2012
:
8
(
1
):
47
. https://doi.org/10.1186/1746-4811-8-47

Baldwin
IT
.
Jasmonate-induced responses are costly but benefit plants under attack in native populations
.
Proc Natl Acad Sci U S A
.
1998
:
95
(
14
):
8113
8118
. https://doi.org/10.1073/pnas.95.14.8113

Baldwin
IT
,
Zhang
ZP
,
Diab
N
,
Ohnmeiss
TE
.
Quantification, correlations and manipulations of wound-induced changes in jasmonic acid and nicotine in Nicotiana sylvestris
.
Planta
.
1997
:
201
(
4
):
397
404
. https://doi.org/10.1007/s004250050082

Bannenberg
G
,
Martínez
M
,
Hamberg
M
,
Castresana
C
.
Diversity of the enzymatic activity in the lipoxygenase gene family of Arabidopsis thaliana
.
Lipids
.
2009
:
44
(
2
):
85
95
. https://doi.org/10.1007/s11745-008-3245-7

Bari
R
,
Jones
JDG
.
Role of plant hormones in plant defence responses
.
Plant Mol Biol
.
2009
:
69
(
4
):
473
488
. https://doi.org/10.1007/s11103-008-9435-0

Bellandi
A
,
Papp
D
,
Breakspear
A
,
Joyce
J
,
Johnston
MG
,
de Keijzer
J
,
Raven
EC
,
Ohtsu
M
,
Vincent
TR
,
Miller
AJ
, et al.
Diffusion and bulk flow of amino acids mediate calcium waves in plants
.
Sci Adv
.
2022
:
8
(
42
):
eabo6693
. https://doi.org/10.1126/sciadv.abo6693

Bennett
T
,
Leyser
O
. The auxin question: a philosophical overview. In:
Zažímalová
E
,
Petrášek
J
,
Benková
E
, editors.
Auxin and its role in plant development
.
Vienna
:
Springer
;
2014
. p.
3
19
.

Berens
ML
,
Berry
HM
,
Mine
A
,
Argueso
CT
,
Tsuda
K
.
Evolution of hormone signaling networks in plant defense
.
Annu Rev Phytopathol
.
2017
:
55
(
1
):
401
425
. https://doi.org/10.1146/annurev-phyto-080516-035544

Berger
S
,
Bell
E
,
Mullet
JE
.
Two methyl jasmonate-insensitive mutants show altered expression of AtVsp in response to methyl jasmonate and wounding
.
Plant Physiol
.
1996
:
111
:
525
531
.

Blázquez
MA
,
Nelson
DC
,
Weijers
D
.
Evolution of plant hormone response pathways
.
Annu Rev Plant Biol
.
2020
:
71
(
1
):
327
353
. https://doi.org/10.1146/annurev-arplant-050718-100309

Boter
M
,
Ruíz-Rivero
O
,
Abdeen
A
,
Prat
S
.
Conserved MYC transcription factors play a key role in jasmonate signaling both in tomato and Arabidopsis
.
Genes Dev
.
2004
:
18
(
13
):
1577
1591
. https://doi.org/10.1101/gad.297704

Bowman
JL
.
The origin of a land flora
.
Nat Plants
.
2022
:
8
(
12
):
1352
1369
. https://doi.org/10.1038/s41477-022-01283-y

Bowman
JL
,
Kohchi
T
,
Yamato
KT
,
Jenkins
J
,
Shu
S
,
Ishizaki
K
,
Yamaoka
S
,
Nishihama
R
,
Nakamura
Y
,
Berger
F
, et al.
Insights into land plant evolution garnered from the Marchantia polymorpha genome
.
Cell
.
2017
:
171
(
2
):
287
304.e15
. https://doi.org/10.1016/j.cell.2017.09.030

Brown
DE
,
Rashotte
AM
,
Murphy
AS
,
Normanly
J
,
Tague
BW
,
Peer
WA
,
Taiz
L
,
Muday
GK
.
Flavonoids act as negative regulators of auxin transport in vivo in arabidopsis
.
Plant Physiol
.
2001
:
126
(
2
):
524
535
. https://doi.org/10.1104/pp.126.2.524

Browse
J
.
The power of mutants for investigating jasmonate biosynthesis and signaling
.
Phytochemistry
.
2009
:
70
(
13–14
):
1539
1546
. https://doi.org/10.1016/j.phytochem.2009.08.004

Bunsick
M
,
McCullough
R
,
McCourt
P
,
Lumba
S
.
Plant hormone signaling: is upside down right side up?
Curr Opin Plant Biol
.
2021
:
63
:
102070
. https://doi.org/10.1016/j.pbi.2021.102070

Caarls
L
,
Elberse
J
,
Awwanah
M
,
Ludwig
NR
,
de Vries
M
,
Zeilmaker
T
,
Van Wees
SCM
,
Schuurink
RC
,
Van den Ackerveken
G
.
Arabidopsis JASMONATE-INDUCED OXYGENASES down-regulate plant immunity by hydroxylation and inactivation of the hormone jasmonic acid
.
Proc Natl Acad Sci U S A
.
2017
:
114
(
24
):
6388
6393
. https://doi.org/10.1073/pnas.1701101114

Cai
J
,
Aharoni
A
.
Amino acids and their derivatives mediating defense priming and growth tradeoff
.
Curr Opin Plant Biol
.
2022
:
69
:
102288
. https://doi.org/10.1016/j.pbi.2022.102288

Cai
Q
,
Yuan
Z
,
Chen
M
,
Yin
C
,
Luo
Z
,
Zhao
X
,
Liang
W
,
Hu
J
,
Zhang
D
.
Jasmonic acid regulates spikelet development in rice
.
Nat Commun
.
2014
:
5
(
1
):
3476
. https://doi.org/10.1038/ncomms4476

Caldelari
D
,
Wang
G
,
Farmer
EE
,
Dong
X
.
Arabidopsis lox3 lox4 double mutants are male sterile and defective in global proliferative arrest
.
Plant Mol Biol
.
2011
:
75
(
1–2
):
25
33
. https://doi.org/10.1007/s11103-010-9701-9

Campos
ML
,
Kang
J-H
,
Howe
GA
.
Jasmonate-triggered plant immunity
.
J Chem Ecol
.
2014
:
40
(
7
):
657
675
. https://doi.org/10.1007/s10886-014-0468-3

Cao
S
,
Kang
S
,
Mao
H
,
Yao
J
,
Gu
L
,
Zheng
N
.
Defining molecular glues with a dual-nanobody cannabidiol sensor
.
Nat Commun
.
2022
:
13
(
1
):
815
. https://doi.org/10.1038/s41467-022-28507-1

Çevik
V
,
Kidd
BN
,
Zhang
P
,
Hill
C
,
Kiddle
S
,
Denby
KJ
,
Holub
EB
,
Cahill
DM
,
Manners
JM
,
Schenk
PM
, et al.
MEDIATOR25 acts as an integrative hub for the regulation of jasmonate-responsive gene expression in Arabidopsis
.
Plant Physiol
.
2012
:
160
(
1
):
541
555
. https://doi.org/10.1104/pp.112.202697

Chae
L
,
Kim
T
,
Nilo-Poyanco
R
,
Rhee
SY
.
Genomic signatures of specialized metabolism in plants
.
Science
.
2014
:
344
(
6183
):
510
513
. https://doi.org/10.1126/science.1252076

Chauvin
A
,
Caldelari
D
,
Wolfender
J-L
,
Farmer
EE
.
Four 13-lipoxygenases contribute to rapid jasmonate synthesis in wounded Arabidopsis thaliana leaves: a role for lipoxygenase 6 in responses to long-distance wound signals
.
New Phytol
.
2013
:
197
(
2
):
566
575
. https://doi.org/10.1111/nph.12029

Chen
R
,
Jiang
H
,
Li
L
,
Zhai
Q
,
Qi
L
,
Zhou
W
,
Liu
X
,
Li
H
,
Zheng
W
,
Sun
J
, et al.
The Arabidopsis mediator subunit MED25 differentially regulates jasmonate and abscisic acid signaling through interacting with the MYC2 and ABI5 transcription factors
.
Plant Cell
.
2012
:
24
(
7
):
2898
2916
. https://doi.org/10.1105/tpc.112.098277

Chini
A
,
Fonseca
S
,
Fernández
G
,
Adie
B
,
Chico
JM
,
Lorenzo
O
,
García-Casado
G
,
López-Vidriero
I
,
Lozano
FM
,
Ponce
MR
, et al.
The JAZ family of repressors is the missing link in jasmonate signalling
.
Nature
.
2007
:
448
(
7154
):
666
671
. https://doi.org/10.1038/nature06006

Chini
A
,
Gimenez-Ibanez
S
,
Goossens
A
,
Solano
R
.
Redundancy and specificity in jasmonate signalling
.
Curr Opin Plant Biol
.
2016
:
33
:
147
156
. https://doi.org/10.1016/j.pbi.2016.07.005

Chini
A
,
Monte
I
,
Zamarreño
AM
,
García-Mina
JM
,
Solano
R
.
Evolution of the jasmonate ligands and their biosynthetic pathways
.
New Phytol
.
2023
:
238
(
5
):
2236
2246
. https://doi.org/10.1111/nph.18891

Chini
A
,
Monte
I
,
Zamarreño
AM
,
Hamberg
M
,
Lassueur
S
,
Reymond
P
,
Weiss
S
,
Stintzi
A
,
Schaller
A
,
Porzel
A
, et al.
An OPR3-independent pathway uses 4,5-didehydrojasmonate for jasmonate synthesis
.
Nat Chem Biol
.
2018
:
14
(
2
):
171
178
. https://doi.org/10.1038/nchembio.2540

Chung
HS
,
Cooke
TF
,
Depew
CL
,
Patel
LC
,
Ogawa
N
,
Kobayashi
Y
,
Howe
GA
.
Alternative splicing expands the repertoire of dominant JAZ repressors of jasmonate signaling
.
Plant J
.
2010
:
63
(
4
):
613
622
. https://doi.org/10.1111/j.1365-313X.2010.04265.x

Chung
HS
,
Koo
AJK
,
Gao
X
,
Jayany
S
,
Thines
B
,
Jones
AD
,
Howe
GA
.
Regulation and function of Arabidopsis JASMONATE-ZIM-domain genes in response to wounding and herbivory
.
Plant Physiol
.
2008
:
146
(
3
):
952
964
. https://doi.org/10.1104/pp.107.115691

Creelman
RA
,
Mullet
JE
.
Biosynthesis and action of jasmonates in plants
.
Annu Rev Plant Physiol Plant Mol Biol
.
1997
:
48
(
1
):
355
381
. https://doi.org/10.1146/annurev.arplant.48.1.355

Creelman
RA
,
Tierney
ML
,
Mullet
JE
.
Jasmonic acid/methyl jasmonate accumulate in wounded soybean hypocotyls and modulate wound gene expression
.
Proc Natl Acad Sci U S A
.
1992
:
89
(
11
):
4938
4941
. https://doi.org/10.1073/pnas.89.11.4938

Dani
KGS
,
Loreto
F
.
Plant volatiles as regulators of hormone homeostasis
.
New Phytol
.
2022
:
234
(
3
):
804
812
. https://doi.org/10.1111/nph.18035

Daryanavard
H
,
Postiglione
AE
,
Mühlemann
JK
,
Muday
GK
.
Flavonols modulate plant development, signaling, and stress responses
.
Curr Opin Plant Biol
.
2023
:
72
:
102350
. https://doi.org/10.1016/j.pbi.2023.102350

Davies
PJ
. The plant hormones: their nature, occurrence, and functions. In:
Davies
PJ
, editor.
Plant hormones: biosynthesis, signal transduction, action!
.
Dordrecht, Netherlands
:
Springer
;
2010
. p.
1
15
.

Delfin
JC
,
Kanno
Y
,
Seo
M
,
Kitaoka
N
,
Matsuura
H
,
Tohge
T
,
Shimizu
T
.
AtGH3.10 is another jasmonic acid-amido synthetase in Arabidopsis thaliana
.
Plant J
.
2022
:
110
(
4
):
1082
1096
. https://doi.org/10.1111/tpj.15724

Demole
E
,
Lederer
E
,
Mercier
D
.
Isolement et détermination de la structure du jasmonate de méthyle, constituant odorant caractéristique de l’essence de jasmin
.
Helv Chim Acta
.
1962
:
45
(
2
):
675
685
. https://doi.org/10.1002/hlca.19620450233

Devoto
A
,
Ellis
C
,
Magusin
A
,
Chang
H-S
,
Chilcott
C
,
Zhu
T
,
Turner
JG
.
Expression profiling reveals COI1 to be a key regulator of genes involved in wound- and methyl jasmonate-induced secondary metabolism, defence, and hormone interactions
.
Plant Mol Biol
.
2005
:
58
(
4
):
497
513
. https://doi.org/10.1007/s11103-005-7306-5

Devoto
A
,
Nieto-Rostro
M
,
Xie
D
,
Ellis
C
,
Harmston
R
,
Patrick
E
,
Davis
J
,
Sherratt
L
,
Coleman
M
,
Turner
JG
.
COI1 links jasmonate signalling and fertility to the SCF ubiquitin-ligase complex in Arabidopsis
.
Plant J
.
2002
:
32
(
4
):
457
466
. https://doi.org/10.1046/j.1365-313X.2002.01432.x

Dharmasiri
N
,
Dharmasiri
S
,
Jones
AM
,
Estelle
M
.
Auxin action in a cell-free system
.
Curr Biol
.
2003
:
13
(
16
):
1418
1422
. https://doi.org/10.1016/S0960-9822(03)00536-0

Dinneny
J
,
Al-Babili
S
,
Benfey
PN
.
β-Cyclocitral is a conserved root growth regulator
.
Proc Natl Acad Sci U S A
.
2019
:
116
(
21
):
10563
10567
. https://doi.org/10.1073/pnas.1821445116

Durán-Medina
Y
,
Ruiz-Cortés
BE
,
Guerrero-Largo
H
,
Marsch-Martínez
N
.
Specialized metabolism and development: an unexpected friendship
.
Curr Opin Plant Biol
.
2021
:
64
:
102142
. https://doi.org/10.1016/j.pbi.2021.102142

Ellis
C
,
Turner
JG
.
The Arabidopsis mutant cev1 has constitutively active jasmonate and ethylene signal pathways and enhanced resistance to pathogens
.
Plant Cell
.
2001
:
13
(
5
):
1025
1033
. https://doi.org/10.1105/tpc.13.5.1025

Engelberth
J
,
Alborn
HT
,
Schmelz
EA
,
Tumlinson
JH
.
Airborne signals prime plants against insect herbivore attack
.
Proc Natl Acad Sci U S A.
2004
:
101
(
6
):
1781
1785
. https://doi.org/10.1073/pnas.0308037100

Erb
M
,
Kliebenstein
DJ
.
Plant secondary metabolites as defenses, regulators, and primary metabolites: the blurred functional trichotomy
.
Plant Physiol
.
2020
:
184
(
1
):
39
52
. https://doi.org/10.1104/pp.20.00433

Erb
M
,
Veyrat
N
,
Robert
CAM
,
Xu
H
,
Frey
M
,
Ton
J
,
Turlings
TCJ
.
Indole is an essential herbivore-induced volatile priming signal in maize
.
Nat Commun
.
2015
:
6
(
1
):
6273
. https://doi.org/10.1038/ncomms7273

Fàbregas
N
,
Fernie
AR
.
The reliance of phytohormone biosynthesis on primary metabolite precursors
.
J Plant Physiol
.
2021
:
268
:
153589
. https://doi.org/10.1016/j.jplph.2021.153589

Farmer
EE
,
Johnson
RR
,
Ryan
CA
.
Regulation of expression of proteinase inhibitor genes by methyl jasmonate and jasmonic acid
.
Plant Physiol
.
1992
:
98
(
3
):
995
1002
. https://doi.org/10.1104/pp.98.3.995

Farmer
EE
,
Ryan
CA
.
Interplant communication: airborne methyl jasmonate induces synthesis of proteinase inhibitors in plant leaves
.
Proc Natl Acad Sci U S A
.
1990
:
87
(
19
):
7713
7716
. https://doi.org/10.1073/pnas.87.19.7713

Farmer
EE
,
Ryan
CA
.
Octadecanoid precursors of jasmonic acid activate the synthesis of wound-inducible proteinase inhibitors
.
Plant Cell
.
1992
:
4
(
2
):
129
134
. https://doi.org/10.2307/3869566

Felemban
A
,
Moreno
JC
,
Mi
J
,
Ali
S
,
Sham
A
,
AbuQamar
SF
,
Al-Babili
S
.
The apocarotenoid β-ionone regulates the transcriptome of Arabidopsis thaliana and increases its resistance against Botrytis cinerea
.
Plant J
.
2023
:
117
(
2
):
541
560
.

Fernández-Calvo
P
,
Chini
A
,
Fernández-Barbero
G
,
Chico
J-M
,
Gimenez-Ibanez
S
,
Geerinck
J
,
Eeckhout
D
,
Schweizer
F
,
Godoy
M
,
Franco-Zorrilla
JM
, et al.
The Arabidopsis bHLH transcription factors MYC3 and MYC4 are targets of JAZ repressors and act additively with MYC2 in the activation of jasmonate responses
.
Plant Cell
.
2011
:
23
(
2
):
701
715
. https://doi.org/10.1105/tpc.110.080788

Fernández-Milmanda
GL
,
Crocco
CD
,
Reichelt
M
,
Mazza
CA
,
Köllner
TG
,
Zhang
T
,
Cargnel
MD
,
Lichy
MZ
,
Fiorucci
A-S
,
Fankhauser
C
, et al.
A light-dependent molecular link between competition cues and defence responses in plants
.
Nat Plants
.
2020
:
6
(
3
):
223
230
. https://doi.org/10.1038/s41477-020-0604-8

Feys
B
,
Benedetti
CE
,
Penfold
CN
,
Turner
JG
.
Arabidopsis mutants selected for resistance to the phytotoxin coronatine are male sterile, insensitive to methyl jasmonate, and resistant to a bacterial pathogen
.
Plant Cell
.
1994
:
6
(
5
):
751
759
. https://doi.org/10.2307/3869877

Fonseca
S
,
Chini
A
,
Hamberg
M
,
Adie
B
,
Porzel
A
,
Kramell
R
,
Miersch
O
,
Wasternack
C
,
Solano
R
.
(+)-7-iso-Jasmonoyl-L-isoleucine is the endogenous bioactive jasmonate
.
Nat Chem Biol
.
2009
:
5
(
5
):
344
350
. https://doi.org/10.1038/nchembio.161

Fonseca
S
,
Fernández-Calvo
P
,
Fernández
GM
,
Díez-Díaz
M
,
Gimenez-Ibanez
S
,
López-Vidriero
I
,
Godoy
M
,
Fernández-Barbero
G
,
Van Leene
J
,
De Jaeger
G
, et al.
bHLH003, bHLH013 and bHLH017 are new targets of JAZ repressors negatively regulating JA responses
.
PLoS One
.
2014
:
9
(
1
):
e86182
. https://doi.org/10.1371/journal.pone.0086182

Gao
Y-Q
,
Jimenez-Sandoval
P
,
Tiwari
S
,
Stolz
S
,
Wang
J
,
Glauser
G
,
Santiago
J
,
Farmer
EE
.
Ricca's factors as mobile proteinaceous effectors of electrical signaling
.
Cell
.
2023
:
186
(
7
):
1337
1351.e20
. https://doi.org/10.1016/j.cell.2023.02.006

Gasperini
D
,
Chauvin
A
,
Acosta
IF
,
Kurenda
A
,
Stolz
S
,
Chételat
A
,
Wolfender
J-L
,
Farmer
EE
.
Axial and radial oxylipin transport
.
Plant Physiol
.
2015
:
169
(
3
):
2244
2254
. https://doi.org/10.1104/pp.15.01104

Glauser
G
,
Dubugnon
L
,
Mousavi
SAR
,
Rudaz
S
,
Wolfender
J-L
,
Farmer
EE
.
Velocity estimates for signal propagation leading to systemic jasmonic acid accumulation in wounded Arabidopsis
.
J Biol Chem
.
2009
:
284
(
50
):
34506
34513
. https://doi.org/10.1074/jbc.M109.061432

Glazebrook
J
.
Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens
.
Annu Rev Phytopathol
.
2005
:
43
(
1
):
205
227
. https://doi.org/10.1146/annurev.phyto.43.040204.135923

Goetz
S
,
Hellwege
A
,
Stenzel
I
,
Kutter
C
,
Hauptmann
V
,
Forner
S
,
McCaig
B
,
Hause
G
,
Miersch
O
,
Wasternack
C
, et al.
Role of cis-12-oxo-phytodienoic acid in tomato embryo development
.
Plant Physiol
.
2012
:
158
(
4
):
1715
1727
. https://doi.org/10.1104/pp.111.192658

Grebner
W
,
Stingl
NE
,
Oenel
A
,
Mueller
MJ
,
Berger
S
.
Lipoxygenase6-dependent oxylipin synthesis in roots is required for abiotic and biotic stress resistance of Arabidopsis
.
Plant Physiol
.
2013
:
161
(
4
):
2159
2170
. https://doi.org/10.1104/pp.113.214544

Green
TR
,
Ryan
CA
.
Wound-induced proteinase inhibitor in plant leaves: a possible defense mechanism against insects
.
Science
.
1972
:
175
(
4023
):
776
777
. https://doi.org/10.1126/science.175.4023.776

Gundlach
H
,
Müller
MJ
,
Kutchan
TM
,
Zenk
MH
.
Jasmonic acid is a signal transducer in elicitor-induced plant cell cultures
.
Proc Natl Acad Sci U S A
.
1992
:
89
(
6
):
2389
2393
. https://doi.org/10.1073/pnas.89.6.2389

Guo
Q
,
Major
IT
,
Kapali
G
,
Howe
GA
.
MYC transcription factors coordinate tryptophan-dependent defence responses and compromise seed yield in Arabidopsis
.
New Phytol
.
2022
:
236
(
1
):
132
145
. https://doi.org/10.1111/nph.18293

Guo
Q
,
Yoshida
Y
,
Major
IT
,
Wang
K
,
Sugimoto
K
,
Kapali
G
,
Havko
NE
,
Benning
C
,
Howe
GA
.
JAZ repressors of metabolic defense promote growth and reproductive fitness in Arabidopsis
.
Proc Natl Acad Sci U S A
.
2018
:
115
(
45
):
E10768
E10777
. https://doi.org/10.1073/pnas.1811828115

He
Z
,
Wang
ZY
,
Li
J
,
Zhu
Q
,
Lamb
C
,
Ronald
P
,
Chory
J
.
Perception of brassinosteroids by the extracellular domain of the receptor kinase BRI1
.
Science
.
2000
:
288
(
5475
):
2360
2363
. https://doi.org/10.1126/science.288.5475.2360

Hedden
P
.
The genes of the green revolution
.
Trends Genet
.
2003
:
19
(
1
):
5
9
. https://doi.org/10.1016/S0168-9525(02)00009-4

Heitz
T
,
Smirnova
E
,
Widemann
E
,
Aubert
Y
,
Pinot
F
,
Ménard
R
.
The rise and fall of jasmonate biological activities
.
Subcell Biochem
.
2016
:
86
:
405
426
. https://doi.org/10.1007/978-3-319-25979-6_16

Heitz
T
,
Widemann
E
,
Lugan
R
,
Miesch
L
,
Ullmann
P
,
Désaubry
L
,
Holder
E
,
Grausem
B
,
Kandel
S
,
Miesch
M
, et al.
Cytochromes P450 CYP94C1 and CYP94B3 catalyze two successive oxidation steps of plant hormone jasmonoyl-isoleucine for catabolic turnover
.
J Biol Chem
.
2012
:
287
(
9
):
6296
6306
. https://doi.org/10.1074/jbc.M111.316364

Hilpert
B
,
Bohlmann
H
,
op den Camp
RO
,
Przybyla
D
,
Miersch
O
,
Buchala
A
,
Apel
K
.
Isolation and characterization of signal transduction mutants of Arabidopsis thaliana that constitutively activate the octadecanoid pathway and form necrotic microlesions
.
Plant J
.
2001
:
26
(
4
):
435
446
. https://doi.org/10.1046/j.1365-313X.2001.2641036.x

Howe
GA
.
Plant hormones: metabolic end run to jasmonate
.
Nat Chem Biol
.
2018
:
14
(
2
):
109
110
. https://doi.org/10.1038/nchembio.2553

Howe
GA
.
Auspicious beginnings for the defense hormone jasmonate
.
Plant Cell
.
2019
:
31
(
6
):
1198
1199
. https://doi.org/10.1105/tpc.19.00333

Howe
GA
,
Jander
G
.
Plant immunity to insect herbivores
.
Annu Rev Plant Biol
.
2008
:
59
(
1
):
41
66
. https://doi.org/10.1146/annurev.arplant.59.032607.092825

Howe
GA
,
Major
IT
,
Koo
AJ
.
Modularity in jasmonate signaling for multistress resilience
.
Annu Rev Plant Biol
.
2018
:
69
(
1
):
387
415
. https://doi.org/10.1146/annurev-arplant-042817-040047

Hu
L
.
Integration of multiple volatile cues into plant defense responses
.
New Phytol
.
2022
:
233
(
2
):
618
623
. https://doi.org/10.1111/nph.17724

Ishiguro
S
,
Kawai-Oda
A
,
Ueda
J
,
Nishida
I
,
Okada
K
.
The DEFECTIVE IN ANTHER DEHISCIENCE gene encodes a novel phospholipase A1 catalyzing the initial step of jasmonic acid biosynthesis, which synchronizes pollen maturation, anther dehiscence, and flower opening in Arabidopsis
.
Plant Cell
.
2001
:
13
(
10
):
2191
2209
. https://doi.org/10.1105/tpc.010192

Jensen
AB
,
Raventos
D
,
Mundy
J
.
Fusion genetic analysis of jasmonate-signalling mutants in Arabidopsis
.
Plant J
.
2002
:
29
(
5
):
595
606
. https://doi.org/10.1046/j.0960-7412.2001.01241.x

Jia
K-P
,
Dickinson
AJ
,
Mi
J
,
Cui
G
,
Xiao
TT
,
Kharbatia
NM
,
Guo
X
,
Sugiono
E
,
Aranda
M
,
Blilou
I
, et al.
Anchorene is a carotenoid-derived regulatory metabolite required for anchor root formation in Arabidopsis
.
Sci Adv
.
2019
:
5
(
11
):
eaaw6787
. https://doi.org/10.1126/sciadv.aaw6787

Jung
HW
,
Tschaplinski
TJ
,
Wang
L
,
Glazebrook
J
,
Greenberg
JT
.
Priming in systemic plant immunity
.
Science
.
2009
:
324
(
5923
):
89
91
. https://doi.org/10.1126/science.1170025

Kagale
S
,
Links
MG
,
Rozwadowski
K
.
Genome-wide analysis of ethylene-responsive element binding factor-associated amphiphilic repression motif-containing transcriptional regulators in Arabidopsis
.
Plant Physiol
.
2010
:
152
(
3
):
1109
1134
. https://doi.org/10.1104/pp.109.151704

Karban
R
,
Baldwin
IT
.
Induced responses to herbivory
.
Chicago
:
University of Chicago Press
;
2007
.

Katsir
L
,
Schilmiller
AL
,
Staswick
PE
,
He
SY
,
Howe
GA
.
COI1 is a critical component of a receptor for jasmonate and the bacterial virulence factor coronatine
.
Proc Natl Acad Sci U S A
.
2008
:
105
(
19
):
7100
7105
. https://doi.org/10.1073/pnas.0802332105

Katz
E
,
Bagchi
R
,
Jeschke
V
,
Rasmussen
ARM
,
Hopper
A
,
Burow
M
,
Estelle
M
,
Kliebenstein
DJ
.
Diverse allyl glucosinolate catabolites independently influence root growth and development
.
Plant Physiol
.
2020
:
183
(
3
):
1376
1390
. https://doi.org/10.1104/pp.20.00170

Katz
E
,
Nisani
S
,
Yadav
BS
,
Woldemariam
MG
,
Shai
B
,
Obolski
U
,
Ehrlich
M
,
Shani
E
,
Jander
G
,
Chamovitz
DA
.
The glucosinolate breakdown product indole-3-carbinol acts as an auxin antagonist in roots of Arabidopsis thaliana
.
Plant J
.
2015
:
82
(
4
):
547
555
. https://doi.org/10.1111/tpj.12824

Kazan
K
,
Manners
JM
.
MYC2: the master in action
.
Mol Plant
.
2013
:
6
(
3
):
686
703
. https://doi.org/10.1093/mp/sss128

Kende
H
.
Plant biology and the Nobel Prize
.
Science
.
1998
:
282
(
5389
):
627
. https://doi.org/10.1126/science.282.5389.627b

Kende
H
,
Zeevaart
J
.
The five “classical” plant hormones
.
Plant Cell
.
1997
:
9
:
1197
1210
. https://doi.org/10.1105/tpc.9.7.1197

Kepinski
S
,
Leyser
O
.
The Arabidopsis F-box protein TIR1 is an auxin receptor
.
Nature
.
2005
:
435
(
7041
):
446
451
. https://doi.org/10.1038/nature03542

Kim
JI
,
Dolan
WL
,
Anderson
NA
,
Chapple
C
.
Indole glucosinolate biosynthesis limits phenylpropanoid accumulation in Arabidopsis thaliana
.
Plant Cell
.
2015
:
27
(
5
):
1529
1546
. https://doi.org/10.1105/tpc.15.00127

Kneeshaw
S
,
Soriano
G
,
Monte
I
,
Hamberg
M
,
Zamarreño
ÁM
,
García-Mina
JM
,
Franco-Zorrilla
JM
,
Kato
N
,
Ueda
M
,
Rey-Stolle
MF
, et al.
Ligand diversity contributes to the full activation of the jasmonate pathway in Marchantia polymorpha
.
Proc Natl Acad Sci U S A
.
2022
:
119
(
36
):e2202930119. https://doi.org/10.1073/pnas.2202930119

Knöfel
H-D
,
Brückner
C
,
Kramell
R
,
Sembdner
G
,
Schreiber
K
.
A radioimmunoassay for jasmonic acid
.
Biochem Physiol Pflanz
.
1984
:
179
(
4
):
317
325
. https://doi.org/10.1016/S0015-3796(84)80048-7

Koo
AJ
.
Metabolism of the plant hormone jasmonate: a sentinel for tissue damage and master regulator of stress response
.
Phytochem Rev
.
2018
:
17
(
1
):
51
80
. https://doi.org/10.1007/s11101-017-9510-8

Koo
AJK
,
Cooke
TF
,
Howe
GA
.
Cytochrome P450 CYP94B3 mediates catabolism and inactivation of the plant hormone jasmonoyl-L-isoleucine
.
Proc Natl Acad Sci U S A
.
2011
:
108
(
22
):
9298
9303
. https://doi.org/10.1073/pnas.1103542108

Koo
AJK
,
Gao
X
,
Jones
AD
,
Howe
GA
.
A rapid wound signal activates the systemic synthesis of bioactive jasmonates in Arabidopsis
.
Plant J
.
2009
:
59
(
6
):
974
986
. https://doi.org/10.1111/j.1365-313X.2009.03924.x

Koo
AJK
,
Howe
GA
.
The wound hormone jasmonate
.
Phytochemistry
.
2009
:
70
(
13–14
):
1571
1580
. https://doi.org/10.1016/j.phytochem.2009.07.018

Koo
AJ
,
Thireault
C
,
Zemelis
S
,
Poudel
AN
,
Zhang
T
,
Kitaoka
N
,
Brandizzi
F
,
Matsuura
H
,
Howe
GA
.
Endoplasmic reticulum-associated inactivation of the hormone jasmonoyl-L-isoleucine by multiple members of the cytochrome P450 94 family in Arabidopsis
.
J Biol Chem
.
2014
:
289
(
43
):
29728
29738
. https://doi.org/10.1074/jbc.M114.603084

Laha
D
,
Johnen
P
,
Azevedo
C
,
Dynowski
M
,
Weiß
M
,
Capolicchio
S
,
Mao
H
,
Iven
T
,
Steenbergen
M
,
Freyer
M
, et al.
VIH2 regulates the synthesis of inositol pyrophosphate InsP8 and jasmonate-dependent defenses in Arabidopsis
.
Plant Cell
.
2015
:
27
(
4
):
1082
1097
. https://doi.org/10.1105/tpc.114.135160

Lavecchia
A
.
Machine-learning approaches in drug discovery: methods and applications
.
Drug Discov Today
.
2015
:
20
(
3
):
318
331
. https://doi.org/10.1016/j.drudis.2014.10.012

Li
L-L
,
Li
Z
,
Lou
Y
,
Meiners
SJ
,
Kong
C-H
.
(−)-Loliolide is a general signal of plant stress that activates jasmonate-related responses
.
New Phytol
.
2022
:
238
(
5
):
2099
2112
. https://doi.org/10.1111/nph.18644

Li
M
,
Wang
F
,
Li
S
,
Yu
G
,
Wang
L
,
Li
Q
,
Zhu
X
,
Li
Z
,
Yuan
L
,
Liu
P
.
Importers drive leaf-to-leaf jasmonic acid transmission in wound-induced systemic immunity
.
Mol Plant
.
2020
:
13
(
10
):
1485
1498
. https://doi.org/10.1016/j.molp.2020.08.017

Li
M
,
Yu
G
,
Cao
C
,
Liu
P
.
Metabolism, signaling, and transport of jasmonates
.
Plant Commun
.
2021
:
2
(
5
):
100231
. https://doi.org/10.1016/j.xplc.2021.100231

Li
L
,
Zhao
Y
,
McCaig
BC
,
Wingerd
BA
,
Wang
J
,
Whalon
ME
,
Pichersky
E
,
Howe
GA
.
The tomato homolog of CORONATINE-INSENSITIVE1 is required for the maternal control of seed maturation, jasmonate-signaled defense responses, and glandular trichome development
.
Plant Cell
.
2004
:
16
(
1
):
126
143
. https://doi.org/10.1105/tpc.017954

Liu
K-H
,
Liu
M
,
Lin
Z
,
Wang
Z-F
,
Chen
B
,
Liu
C
,
Guo
A
,
Konishi
M
,
Yanagisawa
S
,
Wagner
G
, et al.
NIN-like protein 7 transcription factor is a plant nitrate sensor
.
Science
.
2022
:
377
(
6613
):
1419
1425
. https://doi.org/10.1126/science.add1104

Lorenzo
O
,
Chico
JM
,
Sánchez-Serrano
JJ
,
Solano
R
.
JASMONATE-INSENSITIVE1 encodes a MYC transcription factor essential to discriminate between different jasmonate-regulated defense responses in Arabidopsis
.
Plant Cell
.
2004
:
16
(
7
):
1938
1950
. https://doi.org/10.1105/tpc.022319

Luna
E
,
van Hulten
M
,
Zhang
Y
,
Berkowitz
O
,
López
A
,
Pétriacq
P
,
Sellwood
MA
,
Chen
B
,
Burrell
M
,
van de Meene
A
, et al.
Plant perception of β-aminobutyric acid is mediated by an aspartyl-tRNA synthetase
.
Nat Chem Biol
.
2014
:
10
(
6
):
450
456
. https://doi.org/10.1038/nchembio.1520

Luzarowski
M
,
Skirycz
A
.
Emerging strategies for the identification of protein–metabolite interactions
.
J Exp Bot
.
2019
:
70
(
18
):
4605
4618
. https://doi.org/10.1093/jxb/erz228

Maeda
HA
,
Fernie
AR
.
Evolutionary history of plant metabolism
.
Annu Rev Plant Biol
.
2021
:
72
(
1
):
185
216
. https://doi.org/10.1146/annurev-arplant-080620-031054

Malinovsky
FG
,
Thomsen
M-LF
,
Nintemann
SJ
,
Jagd
LM
,
Bourgine
B
,
Burow
M
,
Kliebenstein
DJ
.
An evolutionarily young defense metabolite influences the root growth of plants via the ancient TOR signaling pathway
.
eLife
.
2017
:
6
:
e29353
. https://doi.org/10.7554/eLife.29353

Matsui
K
.
Green leaf volatiles: hydroperoxide lyase pathway of oxylipin metabolism
.
Curr Opin Plant Biol
.
2006
:
9
(
3
):
274
280
. https://doi.org/10.1016/j.pbi.2006.03.002

McConn
M
,
Browse
J
.
The critical requirement for linolenic acid is pollen development, not photosynthesis, in an Arabidopsis mutant
.
Plant Cell
.
1996
:
8
(
3
):
403
416
. https://doi.org/10.2307/3870321

Melotto
M
,
Mecey
C
,
Niu
Y
,
Chung
HS
,
Katsir
L
,
Yao
J
,
Zeng
W
,
Thines
B
,
Staswick
P
,
Browse
J
, et al.
A critical role of two positively charged amino acids in the jas motif of Arabidopsis JAZ proteins in mediating coronatine- and jasmonoyl isoleucine-dependent interactions with the COI1 F-box protein
.
Plant J
.
2008
:
55
(
6
):
979
988
. https://doi.org/10.1111/j.1365-313X.2008.03566.x

Meyerowitz
EM
.
Plants compared to animals: the broadest comparative study of development
.
Science
.
2002
:
295
(
5559
):
1482
1485
. https://doi.org/10.1126/science.1066609

Mochizuki
S
,
Sugimoto
K
,
Koeduka
T
,
Matsui
K
.
Arabidopsis lipoxygenase 2 is essential for formation of green leaf volatiles and five-carbon volatiles
.
FEBS Lett
.
2016
:
590
(
7
):
1017
1027
. https://doi.org/10.1002/1873-3468.12133

Monte
I
.
Jasmonates and salicylic acid: evolution of defense hormones in land plants
.
Curr Opin Plant Biol
.
2023
:
76
:
102470
. https://doi.org/10.1016/j.pbi.2023.102470

Monte
I
,
Franco-Zorrilla
JM
,
García-Casado
G
,
Zamarreño
AM
,
García-Mina
JM
,
Nishihama
R
,
Kohchi
T
,
Solano
R
.
A single JAZ repressor controls the jasmonate pathway in Marchantia polymorpha
.
Mol Plant
.
2019
:
12
(
2
):
185
198
. https://doi.org/10.1016/j.molp.2018.12.017

Monte
I
,
Ishida
S
,
Zamarreño
AM
,
Hamberg
M
,
Franco-Zorrilla
JM
,
García-Casado
G
,
Gouhier-Darimont
C
,
Reymond
P
,
Takahashi
K
,
García-Mina
JM
, et al.
Ligand–receptor co-evolution shaped the jasmonate pathway in land plants
.
Nat Chem Biol
.
2018
:
14
(
5
):
480
488
. https://doi.org/10.1038/s41589-018-0033-4

Monte
I
,
Kneeshaw
S
,
Franco-Zorrilla
JM
,
Chini
A
,
Zamarreño
AM
,
García-Mina
JM
,
Solano
R
.
An ancient COI1-independent function for reactive electrophilic oxylipins in thermotolerance
.
Curr Biol
.
2020
:
30
(
6
):
962
971.e3
. https://doi.org/10.1016/j.cub.2020.01.023

Moreno
JE
,
Shyu
C
,
Campos
ML
,
Patel
LC
,
Chung
HS
,
Yao
J
,
He
SY
,
Howe
GA
.
Negative feedback control of jasmonate signaling by an alternative splice variant of JAZ10
.
Plant Physiol
.
2013
:
162
(
2
):
1006
1017
. https://doi.org/10.1104/pp.113.218164

Morffy
N
,
Strader
LC
.
Old town roads: routes of auxin biosynthesis across kingdoms
.
Curr Opin Plant Biol
.
2020
:
55
:
21
27
. https://doi.org/10.1016/j.pbi.2020.02.002

Morin
H
,
Chételat
A
,
Stolz
S
,
Marcourt
L
,
Glauser
G
,
Wolfender
J-L
,
Farmer
EE
.
Wound-response jasmonate dynamics in the primary vasculature
.
New Phytol
.
2023
:
240
(
4
):
1484
1496
. https://doi.org/10.1111/nph.19207

Mousavi
SAR
,
Chauvin
A
,
Pascaud
F
,
Kellenberger
S
,
Farmer
EE
.
GLUTAMATE RECEPTOR-LIKE genes mediate leaf-to-leaf wound signalling
.
Nature
.
2013
:
500
(
7463
):
422
426
. https://doi.org/10.1038/nature12478

Müller
A
,
Düchting
P
,
Weiler
EW
.
A multiplex GC-MS/MS technique for the sensitive and quantitative single-run analysis of acidic phytohormones and related compounds, and its application to Arabidopsis thaliana
.
Planta
.
2002
:
216
(
1
):
44
56
. https://doi.org/10.1007/s00425-002-0866-6

Mutte
SK
,
Kato
H
,
Rothfels
C
,
Melkonian
M
,
Wong
GK-S
,
Weijers
D
.
Origin and evolution of the nuclear auxin response system
.
eLife
.
2018
:
7
:
e33399
. https://doi.org/10.7554/eLife.33399

Nagashima
A
,
Higaki
T
,
Koeduka
T
,
Ishigami
K
,
Hosokawa
S
,
Watanabe
H
,
Matsui
K
,
Hasezawa
S
,
Touhara
K
.
Transcriptional regulators involved in responses to volatile organic compounds in plants
.
J Biol Chem
.
2019
:
294
(
7
):
2256
2266
. https://doi.org/10.1074/jbc.RA118.005843

Nakata
M
,
Mitsuda
N
,
Herde
M
,
Koo
AJK
,
Moreno
JE
,
Suzuki
K
,
Howe
GA
,
Ohme-Takagi
M
.
A bHLH-type transcription factor, ABA-INDUCIBLE BHLH-TYPE TRANSCRIPTION FACTOR/JA-ASSOCIATED MYC2-LIKE1, acts as a repressor to negatively regulate jasmonate signaling in Arabidopsis
.
Plant Cell
.
2013
:
25
(
5
):
1641
1656
. https://doi.org/10.1105/tpc.113.111112

Nguyen
CT
,
Kurenda
A
,
Stolz
S
,
Chételat
A
,
Farmer
EE
.
Identification of cell populations necessary for leaf-to-leaf electrical signaling in a wounded plant
.
Proc Natl Acad Sci U S A
.
2018
:
115
(
40
):
10178
10183
. https://doi.org/10.1073/pnas.1807049115

Ninkovic
V
,
Markovic
D
,
Rensing
M
.
Plant volatiles as cues and signals in plant communication
.
Plant Cell Environ
.
2021
:
44
(
4
):
1030
1043
. https://doi.org/10.1111/pce.13910

Parthier
B
.
Jasmonates: hormonal regulators or stress factors in leaf senescence?
J Plant Growth Regul
.
1990
:
9
(
1–4
):
57
. https://doi.org/10.1007/BF02041942

Pauwels
L
,
Barbero
GF
,
Geerinck
J
,
Tilleman
S
,
Grunewald
W
,
Pérez
AC
,
Chico
JM
,
Bossche
RV
,
Sewell
J
,
Gil
E
, et al.
NINJA connects the co-repressor TOPLESS to jasmonate signalling
.
Nature
.
2010
:
464
(
7289
):
788
791
. https://doi.org/10.1038/nature08854

Peer
WA
,
Murphy
AS
.
Flavonoids and auxin transport: modulators or regulators?
Trends Plant Sci
.
2007
:
12
(
12
):
556
563
. https://doi.org/10.1016/j.tplants.2007.10.003

Peñuelas
M
,
Monte
I
,
Schweizer
F
,
Vallat
A
,
Reymond
P
,
García-Casado
G
,
Franco-Zorrilla
JM
,
Solano
R
.
Jasmonate-related MYC transcription factors are functionally conserved in Marchantia polymorpha
.
Plant Cell
.
2019
:
31
(
10
):
2491
2509
. https://doi.org/10.1105/tpc.18.00974

Piazza
I
,
Kochanowski
K
,
Cappelletti
V
,
Fuhrer
T
,
Noor
E
,
Sauer
U
,
Picotti
P
.
A map of protein-metabolite interactions reveals principles of chemical communication
.
Cell
.
2018
:
172
(
1–2
):
358
372.e23
. https://doi.org/10.1016/j.cell.2017.12.006

Pourcel
L
,
Irani
NG
,
Koo
AJK
,
Bohorquez-Restrepo
A
,
Howe
GA
,
Grotewold
E
.
A chemical complementation approach reveals genes and interactions of flavonoids with other pathways
.
Plant J
.
2013
:
74
(
3
):
383
397
. https://doi.org/10.1111/tpj.12129

Qi
L
,
Kwiatkowski
M
,
Chen
H
,
Hoermayer
L
,
Sinclair
S
,
Zou
M
,
Del Genio
CI
,
Kubeš
MF
,
Napier
R
,
Jaworski
K
, et al.
Adenylate cyclase activity of TIR1/AFB auxin receptors in plants
.
Nature
.
2022
:
611
(
7934
):
133
138
. https://doi.org/10.1038/s41586-022-05369-7

Qi
T
,
Wang
J
,
Huang
H
,
Liu
B
,
Gao
H
,
Liu
Y
,
Song
S
,
Xie
D
.
Regulation of jasmonate-induced leaf senescence by antagonism between bHLH subgroup IIIe and IIId factors in Arabidopsis
.
Plant Cell
.
2015
:
27
(
6
):
1634
1649
. https://doi.org/10.1105/tpc.15.00110

Ray
R
,
Halitschke
R
,
Gase
K
,
Leddy
SM
,
Schuman
MC
,
Rodde
N
,
Baldwin
IT
.
A persistent major mutation in canonical jasmonate signaling is embedded in an herbivory-elicited gene network
.
Proc Natl Acad Sci U S A
.
2023
:
120
(
35
):e2308500120. https://doi.org/10.1073/pnas.2308500120

Reymond
P
,
Weber
H
,
Damond
M
,
Farmer
EE
.
Differential gene expression in response to mechanical wounding and insect feeding in Arabidopsis
.
Plant Cell
.
2000
:
12
:
707
720
.

Rieseberg
TP
,
Dadras
A
,
Fürst-Jansen
JMR
,
Dhabalia Ashok
A
,
Darienko
T
,
de Vries
S
,
Irisarri
I
,
de Vries
J
.
Crossroads in the evolution of plant specialized metabolism
.
Semin Cell Dev Biol
.
2023
:
134
:
37
58
. https://doi.org/10.1016/j.semcdb.2022.03.004

Robert-Seilaniantz
A
,
Grant
M
,
Jones
JDG
.
Hormone crosstalk in plant disease and defense: more than just jasmonate-salicylate antagonism
.
Annu Rev Phytopathol
.
2011
:
49
(
1
):
317
343
. https://doi.org/10.1146/annurev-phyto-073009-114447

Ruegger
M
,
Dewey
E
,
Gray
WM
,
Hobbie
L
,
Turner
J
,
Estelle
M
.
The TIR1 protein of Arabidopsis functions in auxin response and is related to human SKP2 and yeast grr1p
.
Genes Dev
.
1998
:
12
(
2
):
198
207
. https://doi.org/10.1101/gad.12.2.198

Salehin
M
,
Li
B
,
Tang
M
,
Katz
E
,
Song
L
,
Ecker
JR
,
Kliebenstein
DJ
,
Estelle
M
.
Auxin-sensitive aux/IAA proteins mediate drought tolerance in Arabidopsis by regulating glucosinolate levels
.
Nat Commun
.
2019
:
10
(
1
):
4021
. https://doi.org/10.1038/s41467-019-12002-1

Sandoval
PJ
,
Santiago
J
.
In vitro analytical approaches to study plant ligand-receptor interactions
.
Plant Physiol
.
2020
:
182
(
4
):
1697
1712
. https://doi.org/10.1104/pp.19.01396

Santner
A
,
Calderon-Villalobos
LIA
,
Estelle
M
.
Plant hormones are versatile chemical regulators of plant growth
.
Nat Chem Biol
.
2009
:
5
(
5
):
301
307
. https://doi.org/10.1038/nchembio.165

Schaller
F
,
Weiler
EW
.
Enzymes of octadecanoid biosynthesis in plants–12-oxo-phytodienoate 10,11-reductase
.
Eur J Biochem
.
1997
:
245
(
2
):
294
299
. https://doi.org/10.1111/j.1432-1033.1997.t01-1-00294.x

Schilmiller
AL
,
Koo
AJK
,
Howe
GA
.
Functional diversification of acyl-coenzyme A oxidases in jasmonic acid biosynthesis and action
.
Plant Physiol
.
2007
:
143
(
2
):
812
824
. https://doi.org/10.1104/pp.106.092916

Schlossarek
D
,
Zhang
Y
,
Sokolowska
EM
,
Fernie
AR
,
Luzarowski
M
,
Skirycz
A
.
Don’t let go: co-fractionation mass spectrometry for untargeted mapping of protein-metabolite interactomes
.
Plant J
.
2023
:
113
(
5
):
904
914
. https://doi.org/10.1111/tpj.16084

Schneeberger
K
,
Ossowski
S
,
Lanz
C
,
Juul
T
,
Petersen
AH
,
Nielsen
KL
,
Jørgensen
J-E
,
Weigel
D
,
Andersen
SU
.
SHOREmap: simultaneous mapping and mutation identification by deep sequencing
.
Nat Methods
.
2009
:
6
(
8
):
550
551
. https://doi.org/10.1038/nmeth0809-550

Schulze
A
,
Zimmer
M
,
Mielke
S
,
Stellmach
H
,
Melnyk
CW
,
Hause
B
,
Gasperini
D
.
Wound-Induced shoot-to-root relocation of JA-Ile precursors coordinates Arabidopsis growth
.
Mol Plant
.
2019
:
12
(
10
):
1383
1394
. https://doi.org/10.1016/j.molp.2019.05.013

Sembdner
G
,
Parthier
B
.
The biochemistry and the physiological and molecular actions of jasmonates
.
Annu Rev Plant Physiol Plant Mol Biol
.
1993
:
44
(
1
):
569
589
. https://doi.org/10.1146/annurev.pp.44.060193.003033

Shabek
N
,
Zheng
N
.
Plant ubiquitin ligases as signaling hubs
.
Nat Struct Mol Biol
.
2014
:
21
(
4
):
293
296
. https://doi.org/10.1038/nsmb.2804

Sheard
LB
,
Tan
X
,
Mao
H
,
Withers
J
,
Ben-Nissan
G
.
Jasmonate perception by inositol-phosphate-potentiated COI1–JAZ co-receptor
.
Nature
.
2010
:
468
(
7322
):
400
405
. https://doi.org/10.1038/nature09430

Shields
A
,
Shivnauth
V
,
Castroverde
CDM
.
Salicylic acid and N-hydroxypipecolic acid at the fulcrum of the plant immunity-growth equilibrium
.
Front Plant Sci
.
2022
:
13
:
841688
. https://doi.org/10.3389/fpls.2022.841688

Shyu
C
,
Figueroa
P
,
Depew
CL
,
Cooke
TF
,
Sheard
LB
,
Moreno
JE
,
Katsir
L
,
Zheng
N
,
Browse
J
,
Howe
GA
.
JAZ8 lacks a canonical degron and has an EAR motif that mediates transcriptional repression of jasmonate responses in Arabidopsis
.
Plant Cell
.
2012
:
24
(
2
):
536
550
. https://doi.org/10.1105/tpc.111.093005

Smirnova
E
,
Marquis
V
,
Poirier
L
,
Aubert
Y
,
Zumsteg
J
,
Ménard
R
,
Miesch
L
,
Heitz
T
.
Jasmonic acid oxidase 2 hydroxylates jasmonic acid and represses basal defense and resistance responses against Botrytis cinerea infection
.
Mol Plant
.
2017
:
10
(
9
):
1159
1173
. https://doi.org/10.1016/j.molp.2017.07.010

Song
WC
,
Brash
AR
.
Purification of an allene oxide synthase and identification of the enzyme as a cytochrome P-450
.
Science
.
1991
:
253
(
5021
):
781
784
. https://doi.org/10.1126/science.1876834

Song
S
,
Qi
T
,
Wasternack
C
,
Xie
D
.
Jasmonate signaling and crosstalk with gibberellin and ethylene
.
Curr Opin Plant Biol
.
2014
:
21
:
112
119
. https://doi.org/10.1016/j.pbi.2014.07.005

Staswick
PE
.
Novel regulation of vegetative storage protein genes
.
Plant Cell
.
1990
:
2
(
1
):
1
6
. https://doi.org/10.2307/3869045

Staswick
PE
,
Su
W
,
Howell
SH
.
Methyl jasmonate inhibition of root growth and induction of a leaf protein are decreased in an Arabidopsis thaliana mutant
.
Proc Natl Acad Sci U S A
.
1992
:
89
(
15
):
6837
6840
. https://doi.org/10.1073/pnas.89.15.6837

Staswick
PE
,
Tiryaki
I
.
The oxylipin signal jasmonic acid is activated by an enzyme that conjugates it to isoleucine in Arabidopsis
.
Plant Cell
.
2004
:
16
(
8
):
2117
2127
. https://doi.org/10.1105/tpc.104.023549

Staswick
PE
,
Tiryaki
I
,
Rowe
ML
.
Jasmonate response locus JAR1 and several related Arabidopsis genes encode enzymes of the firefly luciferase superfamily that show activity on jasmonic, salicylic, and indole-3-acetic acids in an assay for adenylation
.
Plant Cell
.
2002
:
14
(
6
):
1405
1415
. https://doi.org/10.1105/tpc.000885

Stumpe
M
,
Göbel
C
,
Faltin
B
,
Beike
AK
,
Hause
B
,
Himmelsbach
K
,
Bode
J
,
Kramell
R
,
Wasternack
C
,
Frank
W
, et al.
The moss Physcomitrella patens contains cyclopentenones but no jasmonates: mutations in allene oxide cyclase lead to reduced fertility and altered sporophyte morphology
.
New Phytol
.
2010
:
188
(
3
):
740
749
. https://doi.org/10.1111/j.1469-8137.2010.03406.x

Sugimoto
K
,
Zager
JJ
,
Aubin
BS
,
Lange
BM
,
Howe
GA
.
Flavonoid deficiency disrupts redox homeostasis and terpenoid biosynthesis in glandular trichomes of tomato
.
Plant Physiol
.
2022
:
188
(
3
):
1450
1468
. https://doi.org/10.1093/plphys/kiab488

Sun
Y
,
Harpazi
B
,
Wijerathna-Yapa
A
,
Merilo
E
,
de Vries
J
,
Michaeli
D
,
Gal
M
,
Cuming
AC
,
Kollist
H
,
Mosquna
A
.
A ligand-independent origin of abscisic acid perception
.
Proc Natl Acad Sci U S A
.
2019
:
116
(
49
):
24892
24899
. https://doi.org/10.1073/pnas.1914480116

Takeuchi
J
,
Fukui
K
,
Seto
Y
,
Takaoka
Y
,
Okamoto
M
.
Ligand-receptor interactions in plant hormone signaling
.
Plant J
.
2021
:
105
(
2
):
290
306
. https://doi.org/10.1111/tpj.15115

Tan
X
,
Calderon-Villalobos
LIA
,
Sharon
M
,
Zheng
C
,
Robinson
CV
,
Estelle
M
,
Zheng
N
.
Mechanism of auxin perception by the TIR1 ubiquitin ligase
.
Nature
.
2007
:
446
(
7136
):
640
645
. https://doi.org/10.1038/nature05731

Tata
JR
.
One hundred years of hormones
.
EMBO Rep
.
2005
:
6
(
6
):
490
496
. https://doi.org/10.1038/sj.embor.7400444

Thines
B
,
Katsir
L
,
Melotto
M
,
Niu
Y
,
Mandaokar
A
,
Liu
G
,
Nomura
K
,
He
SY
,
Howe
GA
,
Browse
J
.
JAZ repressor proteins are targets of the SCF(COI1) complex during jasmonate signalling
.
Nature
.
2007
:
448
(
7154
):
661
665
. https://doi.org/10.1038/nature05960

Thireault
C
,
Shyu
C
,
Yoshida
Y
,
St Aubin
B
,
Campos
ML
,
Howe
GA
.
Repression of jasmonate signaling by a non-TIFY JAZ protein in Arabidopsis
.
Plant J
.
2015
:
82
(
4
):
669
679
. https://doi.org/10.1111/tpj.12841

Thomma
BP
,
Eggermont
K
,
Penninckx
IA
,
Mauch-Mani
B
,
Vogelsang
R
,
Cammue
BP
,
Broekaert
WF
.
Separate jasmonate-dependent and salicylate-dependent defense-response pathways in Arabidopsis are essential for resistance to distinct microbial pathogens
.
Proc Natl Acad Sci U S A
.
1998
:
95
(
25
):
15107
15111
. https://doi.org/10.1073/pnas.95.25.15107

Toyota
M
,
Spencer
D
,
Sawai-Toyota
S
,
Jiaqi
W
,
Zhang
T
,
Koo
AJ
,
Howe
GA
,
Gilroy
S
.
Glutamate triggers long-distance, calcium-based plant defense signaling
.
Science
.
2018
:
361
(
6407
):
1112
1115
. https://doi.org/10.1126/science.aat7744

Trewavas
A
.
How do plant growth substances work?
Plant Cell Environ
.
1981
:
4
(
3
):
203
228
. https://doi.org/10.1111/j.1365-3040.1981.tb01048.x

Ueda
J
,
Kato
J
.
Isolation and identification of a senescence-promoting substance from wormwood (Artemisia absinthium L)
.
Plant Physiol
.
1980
:
66
(
2
):
246
249
. https://doi.org/10.1104/pp.66.2.246

Urano
D
,
Jones
AM
.
Heterotrimeric G protein-coupled signaling in plants
.
Annu Rev Plant Biol
.
2014
:
65
(
1
):
365
384
. https://doi.org/10.1146/annurev-arplant-050213-040133

Vick
BA
,
Zimmerman
DC
.
Biosynthesis of jasmonic acid by several plant species
.
Plant Physiol
.
1984
:
75
(
2
):
458
461
. https://doi.org/10.1104/pp.75.2.458

Wang
K-D
,
Borrego
EJ
,
Kenerley
CM
,
Kolomeits
MV
.
Oxylipins other than jasmonic acid are xylem-resident signals regulating systemic resistance induced by Trichoderma virens in maize
.
Plant Cell
.
2020
:
32
(
1
):
16
185
. https://doi.org/10.1105/tpc.19.00487

Wang
L
,
Erb
M
.
Volatile uptake, transport, perception, and signaling shape a plant's nose
.
Essays Biochem
.
2022
:
66
(
5
):
695
702
. https://doi.org/10.1042/EBC20210092

Wang
JY
,
Haider
I
,
Jamil
M
,
Fiorilli
V
,
Saito
Y
,
Mi
J
,
Baz
L
,
Kountche
BA
,
Jia
K-P
,
Guo
X
, et al.
The apocarotenoid metabolite zaxinone regulates growth and strigolactone biosynthesis in rice
.
Nat Commun
.
2019
:
10
(
1
):
810
. https://doi.org/10.1038/s41467-019-08461-1

Wang
JY
,
Lin
P-Y
,
Al-Babili
S
.
On the biosynthesis and evolution of apocarotenoid plant growth regulators
.
Semin Cell Dev Biol
.
2021
:
109
:
3
11
. https://doi.org/10.1016/j.semcdb.2020.07.007

Wang
J
,
Wu
D
,
Wang
Y
,
Xie
D
.
Jasmonate action in plant defense against insects
.
J Exp Bot
.
2019
:
70
(
13
):
3391
3400
. https://doi.org/10.1093/jxb/erz174

Wasternack
C
.
How jasmonates earned their laurels: past and present
.
J Plant Growth Regul
.
2015
:
34
(
4
):
761
794
. https://doi.org/10.1007/s00344-015-9526-5

Wasternack
C
,
Feussner
I
.
The oxylipin pathways: biochemistry and function
.
Annu Rev Plant Biol
.
2018
:
69
(
1
):
363
386
. https://doi.org/10.1146/annurev-arplant-042817-040440

Wasternack
C
,
Parthier
B
.
Jasmonate-signalled plant gene expression
.
Trends Plant Sci
.
1997
:
2
(
8
):
302
307
. https://doi.org/10.1016/S1360-1385(97)89952-9

Wasternack
C
,
Song
S
.
Jasmonates: biosynthesis, metabolism, and signaling by proteins activating and repressing transciption
.
J Exp Bot
.
2017
:
68
(
6
)
1303
1321
. https://doi.org/10.1093/jxb/erw443

Waters
MT
,
Nelson
DC
.
Karrikin perception and signalling
.
New Phytol
.
2023
:
237
(
3
):
1525
1541
. https://doi.org/10.1111/nph.18598

Weng
J-K
,
Lynch
JH
,
Matos
JO
,
Dudareva
N
.
Adaptive mechanisms of plant specialized metabolism connecting chemistry to function
.
Nat Chem Biol
.
2021
:
17
(
10
):
1037
1045
. https://doi.org/10.1038/s41589-021-00822-6

Westfall
CS
,
Zubieta
C
,
Herrmann
J
,
Kapp
U
,
Nanao
MH
,
Jez
JM
.
Structural basis for prereceptor modulation of plant hormones by GH3 proteins
.
Science
.
2012
:
336
(
6089
):
1708
1711
. https://doi.org/10.1126/science.1221863

Widemann
E
,
Miesch
L
,
Lugan
R
,
Holder
E
,
Heinrich
C
,
Aubert
Y
,
Miesch
M
,
Pinot
F
,
Heitz
T
.
The amidohydrolases IAR3 and ILL6 contribute to jasmonoyl-isoleucine hormone turnover and generate 12-hydroxyjasmonic acid upon wounding in Arabidopsis leaves
.
J Biol Chem
.
2013
:
288
(
44
):
31701
31714
. https://doi.org/10.1074/jbc.M113.499228

Woldemariam
MG
,
Onkokesung
N
,
Baldwin
IT
,
Galis
I
.
Jasmonoyl-L-isoleucine hydrolase 1 (JIH1) regulates jasmonoyl-L-isoleucine levels and attenuates plant defenses against herbivores
.
Plant J
.
2012
:
72
(
5
):
758
767
. https://doi.org/10.1111/j.1365-313X.2012.05117.x

Wong
A
,
Chi
W
,
Yu
J
,
Bi
C
,
Tian
X
,
Yang
Y
,
Gehring
C
.
Plant adenylate cyclases have come full circle
.
Nat Plants
.
2023
:
9
(
9
):
1389
1397
. https://doi.org/10.1038/s41477-023-01486-x

Wu
F
,
Deng
L
,
Zhai
Q
,
Zhao
J
,
Chen
Q
,
Li
C
.
Mediator subunit MED25 couples alternative splicing of JAZ genes with fine-tuning of jasmonate signaling
.
Plant Cell
.
2020
:
32
(
2
):
429
448
. https://doi.org/10.1105/tpc.19.00583

Xie
DX
,
Feys
BF
,
James
S
,
Nieto-Rostro
M
,
Turner
JG
.
COI1: an Arabidopsis gene required for jasmonate-regulated defense and fertility
.
Science
.
1998
:
280
(
5366
):
1091
1094
. https://doi.org/10.1126/science.280.5366.1091

Xu
L
,
Liu
F
,
Wang
Z
,
Peng
W
,
Huang
R
,
Huang
D
,
Xie
D
.
An Arabidopsis mutant cex1 exhibits constant accumulation of jasmonate-regulated AtVSP, Thi2.1 and PDF1.2
.
FEBS Lett
.
2001
:
494
(
3
):
161
164
. https://doi.org/10.1016/S0014-5793(01)02331-6

Yamane
H
,
Takahashi
N
,
Ueda
J-I
,
Kato
J
.
Resolution of (±)-methyl jasmonate by high performance liquid chromatography and the inhibitory effect of (+)-enantiomer on the growth of rice seedlings
.
Agric Biol Chem
.
1981
:
45
(
7
):
1709
1711
. https://doi.org/10.1080/00021369.1981.10864755

Yan
Y
,
Stolz
S
,
Chételat
A
,
Reymond
P
,
Pagni
M
,
Dubugnon
L
,
Farmer
EE
.
A downstream mediator in the growth repression limb of the jasmonate pathway
.
Plant Cell
.
2007
:
19
(
8
):
2470
2483
. https://doi.org/10.1105/tpc.107.050708

Yan
J
,
Yao
R
,
Chen
L
,
Li
S
,
Gu
M
,
Nan
F
,
Xie
D
.
Dynamic perception of jasmonates by the F-box protein COI1
.
Mol Plant
.
2018
:
11
:
1237
1247
. https://doi.org/10.1016/j.molp.2018.07.007

Yu
Y
,
Tang
W
,
Lin
W
,
Li
W
,
Zhou
X
,
Li
Y
,
Chen
R
,
Zheng
R
,
Qin
G
,
Cao
W
, et al.
ABLs and TMKs are co-receptors for extracellular auxin
.
Cell
.
2023
:
186
(
25
):
5457
5471.e17
. https://doi.org/10.1016/j.cell.2023.10.017

Zander
M
.
Many ways to repress! JAZ's agony of choices
.
Mol Plant
.
2021
:
14
(
5
):
714
716
. https://doi.org/10.1016/j.molp.2021.04.010

Zander
M
,
Lewsey
MG
,
Clark
NM
,
Yin
L
,
Bartlett
A
,
Saldierna Guzmán
JP
,
Hann
E
,
Langford
AE
,
Jow
B
,
Wise
A
, et al.
Integrated multi-omics framework of the plant response to jasmonic acid
.
Nat Plants
.
2020
:
6
(
3
):
290
302
. https://doi.org/10.1038/s41477-020-0605-7

Zhang
F
,
Ke
J
,
Zhang
L
,
Chen
R
,
Sugimoto
K
,
Howe
GA
,
Xu
HE
,
Zhou
M
,
He
SY
,
Melcher
K
.
Structural insights into alternative splicing-mediated desensitization of jasmonate signaling
.
Proc Natl Acad Sci U S A
.
2017
:
114
(
7
):
1720
1725
. https://doi.org/10.1073/pnas.1616938114

Zhang
Y
,
Turner
JG
.
Wound-induced endogenous jasmonates stunt plant growth by inhibiting mitosis
.
PLoS One
.
2008
:
3
(
11
):
e3699
. https://doi.org/10.1371/journal.pone.0003699

Zhang
F
,
Yao
J
,
Ke
J
,
Zhang
L
,
Lam
VQ
,
Xin
X-F
,
Zhou
XE
,
Chen
J
,
Brunzelle
J
,
Griffin
PR
, et al.
Structural basis of JAZ repression of MYC transcription factors in jasmonate signalling
.
Nature
.
2015
:
525
(
7568
):
269
273
. https://doi.org/10.1038/nature14661

Zhou
XE
,
Zhang
Y
,
Yao
J
,
Zheng
J
,
Zhou
Y
,
He
Q
,
Moreno
J
,
Lam
VQ
,
Cao
X
,
Sugimoto
K
, et al.
Assembly of JAZ-JAZ and JAZ-NINJA complexes in jasmonate signaling
.
Plant Commun
.
2023
:
4
(
6
):
100639
–. https://doi.org/10.1016/j.xplc.2023.100639

Zuo
Z
,
Weraduwage
SM
,
Lantz
AT
,
Sanchez
LM
,
Weise
SE
,
Wang
J
,
Childs
KL
,
Sharkey
TD
.
Isoprene acts as a signaling molecule in gene networks important for stress responses and plant growth
.
Plant Physiol
.
2019
:
180
(
1
):
124
152
. https://doi.org/10.1104/pp.18.01391

Author notes

The author responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (https://dbpia.nl.go.kr/plphys/pages/General-Instructions) is: Gregg A. Howe.

Conflict of interest statement. None declared.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.