Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Oct 6.
Published in final edited form as: Annu Rev Cell Dev Biol. 2021 Aug 5;37:341–367. doi: 10.1146/annurev-cellbio-010521-015047

Dynamic Nutrient Signaling Networks in Plants

Lei Li 1, Kun-hsiang Liu 1,2, Jen Sheen 1
PMCID: PMC8497281  NIHMSID: NIHMS1691452  PMID: 34351784

Abstract

Nutrients are vital to life through intertwined sensing, signaling, and metabolic processes. Emerging research focuses on how distinct nutrient signaling networks integrate and coordinate gene expression, metabolism, growth, and survival. We review the multifaceted roles of sugars, nitrate, and phosphate as essential plant nutrients in controlling complex molecular and cellular mechanisms of dynamic signaling networks. Key advances in central sugar and energy signaling mechanisms mediated by the evolutionarily conserved master regulators HEXOKINASE1 (HXK1), TARGET OF RAPAMYCIN (TOR), and SNF1-RELATED PROTEIN KINASE1 (SNRK1) are discussed. Significant progress in primary nitrate sensing, calcium signaling, transcriptome analysis, and root–shoot communication to shape plant biomass and architecture are elaborated. Discoveries on intracellular and extracellular phosphate signaling and the intimate connections with nitrate and sugar signaling are examined. This review highlights the dynamic nutrient, energy, growth, and stress signaling networks that orchestrate systemwide transcriptional, translational, and metabolic reprogramming, modulate growth and developmental programs, and respond to environmental cues.

Keywords: glucose signaling, HEXOKINASE1, HXK1, TARGET OF RAPAMYCIN, TOR, SNF1-RELATED PROTEIN KINASE1, SNRK1, nitrate signaling, phosphate signaling

INTRODUCTION

Nutrient signaling is emerging as the most ancient and fundamental mechanism to modulate cellular activities and organismal development through integration with other intrinsic regulators and environmental inputs in multicellular plants and animals. In contrast to the previously prevailing notion that nutrients automatically feed into cellular metabolism and growth, nutrient signaling mechanisms are complex due to being part of the tailored regulatory networks in diverse cells, tissues, and organs with specialized physiology, metabolism, and functions. There is increasing recognition of the essential roles of nutrient signaling networks in plant development and agriculture as well as in human health, disease, and aging.

Free from the organic nutrient constraint in animals, plants play a central role in bridging the conversion of inorganic element resources on Earth to the production of organic molecules and nutrients. For example, carbon (C) in the atmosphere as CO2 is fixed through photosynthesis to generate sugars in the shoot. Inorganic nitrogen (N) in nitrate (NO 3) and phosphorus (P) in phosphate (PO43-/Pi) are acquired and assimilated from the soil by roots to synthesize nucleic acids, amino acids, lipids, vitamins, and adenosine triphosphate (ATP). Photosynthetic plants possess remarkable versatility in acquiring and synthesizing nutrients by sensing and responding to wide-ranging inorganic nutrients in soil or inside cells and ultimately fuel, build, and sustain all lives (Figure 1). In this review, we examine key molecular and cellular mechanisms intimately linked to sensing and dynamic primary responses to sugars (Baena-González & Lunn 2020, Broeckx et al. 2016, Dobrenel et al. 2016, Li & Sheen 2016, Ryabova et al. 2019, Sheen 2014, Shi et al. 2018, Wu et al. 2019), nitrate (Liu et al. 2020, Vidal et al. 2020), and phosphate (Chien et al. 2018, Ham et al. 2018, Puga et al. 2017) as major nutrient signals. We highlight and discuss emerging new knowledge and universal concepts generated from the rapidly growing repertoire of interdisciplinary experimental approaches and tools largely in Arabidopsis thaliana that uncover expanded complexity in plant nutrient sensing and signaling.

Figure 1.

Figure 1

Carbon, nitrogen, and phosphorus are essential nutrients that control integrated signaling networks in plants. (a) Sugars are generated by photosynthesis from source leaves and translocated to sink organs. Nitrate (NO 3) and phosphate (PO 3-4) are acquired by roots from the soil and transported to shoots. (b) Glucose-sensor HXK1 and energy sensors TOR and SNRK1 respond to different concentrations of sugars. NRT1.1 is a nitrate transceptor and mediates nitrate-activated CPK-NLP signaling to induced PNRs. Phosphate is taken up by PHT1 and converted to InsP8, which is sensed by SPX to repress PHR1 in PSRs. Other TFs also regulate PNR and PSR. (c) ❶ (Top) HXK1-mediated growth promotion in adult Arabidopsis plants and (bottom) growth arrest in seedlings on high glucose in WT and hxk1 plants. ❷ WT, snrk1, and SNRK1-OX plants. ❸ WT and conditional tor seedlings. ❹ Seedlings supplied with KCl (control) and different nitrogen sources, NH4+, Gln, and NO3. ❺ Seedlings grown in the medium with (+) or without (−) Phosphate (Pi). Abbreviations: CPK, CALCIUM DEPENDENT PROTEIN KINASE; Glc, glucose; Gln, glutamine; HXK1, HEXOKINASE1; InsP8, inositol phyrophosphate; NLP, NIN-LIKE PROTEIN; OX, overexpression; P, phosphorylation; PHR1, PHOSPHATE STARVATION RESPONSE1; PHT1, PHOSPHATE TRANSPORTER1; PNR, primary nitrate response; PSR, phosphate-starvation response; SNRK1, SNF1-RELATED PROTEIN KINASE1; SPX, SYG1/Pho81/XPR1; T6P, trehalose 6-phosphate; TF, transcription factor; TOR, TARGET OF RAPAMYCIN; WT, wild type.

SUGAR SENSING AND SIGNALING

In photosynthetic plants, glucose and sucrose show hormonelike activities and modulate many essential processes, including embryogenesis, germination, seedling development, vegetative growth, reproduction, and senescence. How glucose or other sugars and carbon metabolites function as direct and essential signaling molecules remains an exciting research challenge, as key mutants or severe nutrient deficiency are often lethal or deleterious with pleiotropic phenotypes. There are two leading mechanisms to perceive and transduce sugar signals. Direct sensing is triggered by sugar-binding sensors with diverse affinity and specificity, such as glucose-binding sensor HEXOKINASE1 (HXK1), as well as other potential sugar sensors, including REGULATOR OF G-PROTEIN SIGNALING1 (RGS1) and TREHALOSE 6-PHOSPHATE (T6P) SYNTHASE1 (TPS1). Sugar-derived bioenergetic molecules and metabolites modulate signaling proteins, including glucose-activated TARGET OF RAPAMYCIN (TOR) and sugar-repressed SNF1-RELATED PROTEIN KINASE1 (SNRK1), which relay sugar signals indirectly (Figure 1). In this review, we discuss the sugar signals, the potential sugar sensors, and the sugar signaling networks discovered in the plant system (Figure 2).

Figure 2.

Figure 2

Sugar sensing and signaling in plants. Glucose sensor HXK1 controls multiple biological processes at different glucose concentrations uncoupled from metabolism. Sucrose and glucose metabolism generate diverse sugars and metabolites that could potentially be perceived by distinct sensors. Abbreviations: Acetyl CoA, acetyl coenzyme A; FBP, fructose-1,6-bisphosphatase; G1P, glucose 1-phosphate; G6P, glucose 6-phosphate; Gln, glutamine; GlcNAc, N-acetylglucosamine; HXK, HEXOKINASE; INV, invertase; OGT, O-LINKED N-ACETYLGLUCOSAMINE (O-GLcNAc) TRANSFERASE; R5P, ribose 5-phosphate; RGS1, REGULATOR OF G-PROTEIN SIGNALING1; SNRK1, SNF-RELATED PROTEIN KINASE1; SUS, SUCROSE SYNTHASE; T6P, trehalose 6-phosphate; TOR, TARGET OF RAPAMYCIN; TPP, T6P PHOSPHATASE; TPS, T6P SYNTHASE; UDP, uridine diphosphate.

Glucose Sensors and Signaling

Using Arabidopsis. thaliana as a reference plant, genetic screens have identified glucose-insensitive (gin) and glucose-oversensitive mutants based on the developmental arrest phenotype mediated by high glucose during seedling germination. Extensive characterization of sugar mutants revealed unexpected positive or negative interactions between glucose and plant hormones, including positive abscisic acid (ABA) and auxin relays and antagonistic cytokinin and ethylene signaling (Moore et al. 2003, Rolland et al. 2006, Sheen 2014). Importantly, gin2 mutants define HXK1 as a prominent glucose sensor (Jang et al. 1997, Moore et al. 2003) despite the presence of five other HXK and HXK-LIKE (HKL) genes that may provide overlapping or distinct functions in sugar sensing and metabolism (Karve et al. 2008). HXK1 mutants lacking catalytic activity but retaining strong glucose binding still support various signaling functions in transcription, cell proliferation, root and inflorescence growth, leaf expansion, fertility, and senescence (Figure 2), thus demonstrating the uncoupling of glucose signaling from glucose metabolism (Feng et al. 2015, Jang et al. 1997, Moore et al. 2003). The catalytic activity of HXK1—but not the glucose-sensor function—is required for lesion formation and cell death in myo-inositol-1-phosphate synthase mutants, illustrating the distinct roles of HXK1 in glucose signaling and metabolic pathways (Figure 2) (Bruggeman et al. 2015).

Although the majority of HXK1 protein is located at the outer membrane of mitochondria (Karve et al. 2008), HXK1 is detected in the nucleus and is associated with target genes based on proteomic and chromatin immunoprecipitation (ChIP) analyses (Cho et al. 2006). The potential actions of the plant HXK glucose sensor in the nucleus are supported by the loss of the glucose-dependent degradation of the nuclear transcription factor (TF) ETHYLENE INSENSITIVE3 (EIN3) through proteasome and reporter gene expression in gin2, as well as the nuclear localization of OsHXK5/6 in rice and specific HXKs from other plant species (Aguilera-Alvarado & Sánchez-Nieto 2017, Cho et al. 2009, Yanagisawa et al. 2003). Apple MdHXK1 appears to phosphorylate and stabilize TF MdbHLH3 in the nucleus to promote anthocyanin synthesis (Hu et al. 2016). To precisely and fully dissect the direct and indirect glucose signaling networks, it will be informative to develop more specific and quantitative glucose-induced phenotypes and molecular readouts in different cell types, organs, and developmental stages (Figure 1) (Baena-González et al. 2007, Kelly et al. 2012, Li & Sheen 2016, Moore et al. 2003, Xiong et al. 2013). Systematic and comparative studies of the large numbers of HXKs and HKLs with different subcellular localization in diverse plant species from unicellular green algae to flowering plants may lead to fresh findings in sugar signaling (Aguilera-Alvarado & Sánchez-Nieto 2017, Cho et al. 2009, Karve et al. 2008, Roth et al. 2019, Ulfstedt et al. 2018).

Besides intracellular and nuclear glucose sensing, the seven–transmembrane domain protein RGS1 has been proposed to play a critical role as a glucose sensor on the plasma membrane in plants. High levels of glucose stimulate RGS1 phosphorylation by the protein kinase (PK) WNK8-Gβγ complex, which leads to RGS1 endocytosis and G-protein signaling, cell division, and marker gene activation (Urano et al. 2012). A deep screen for Arabidopsis G-protein-interacting proteins has identified that RGS1-HXK1-INTERACTING PROTEIN1 (RHIP1) serves as the physical scaffold for these two sensors. The three mutants hxk1, rgs1, and rhip1 exhibit complex relationships in different sugar assays (Huang et al. 2015). It will be crucial to determine whether and how glucose binds directly to RGS1 as a potential sensor and how RHIP1 mediates the distinct regulation of marker genes in different glucose signaling pathways.

Novel Sugar Signaling

Sucrose is transported from photosynthetic leaf sources to support the growth and development of sink organs, including the root apical meristem (RAM) and shoot apical meristem (SAM), where new cells are supplied from the stem cell niche, growing primary and lateral roots as well as leaf primordia, young leaves, flowers, fruit, and seeds (Figure 1). There are several sucrose-specific responses unique to plants. The best studied example is the sucrose-induced repression of translation (Li & Sheen 2016, Rahmani et al. 2009). The 5′ leader of the transcript encoding Arabidopsis TF BASIC LEUCINE ZIPPER (bZIP)11 harbors several upstream open reading frames (uORFs). The second uORF encoding a 28-residue peptide can act as an attenuator peptide by stalling ribosomes on the mRNA in the presence of high levels of sucrose but not glucose (Rahmani et al. 2009). For specific sucrose sensing, various sucrose transporters and synthases with different subcellular localizations and potential regulatory functions are candidates to be sucrose sensors (Chen et al. 2015, Stein & Granot 2019).

Intensive research on plant trehalose biosynthesis by T6P SYNTHASES (TPSs) and T6P PHOSPHATASES (TPPs) has suggested unique sugar sensing and signaling mechanisms because of the intriguing regulatory effects of T6P on plant growth, development, and stress resistance (Figure 2) (Broeckx et al. 2016, Figueroa & Lunn 2016). The essential role of T6P in plant development was first shown by the recessive embryo lethality of tps1 mutants (Eastmond et al. 2002). T6P levels were modulated by expressing Escherichia coli genes encoding TPS to increase T6P or TPP to lower T6P in transgenic Arabidopsis plants or to complement tps1. These experiments demonstrated that T6P performs critical roles in regulating carbohydrate use for growth and in modulating shoot growth, branching, and flowering (Schluepmann et al. 2003). By controlling TPS1 expression in both leaves and the SAM, T6P was shown to activate key flowering regulators (Wahl et al. 2013). Direct T6P manipulation using sunlight-triggered release and plant-permeable analogs of T6P during the grain-filling period increases wheat yield. T6P spraying also leads to drought tolerance in wheat plants (Griffiths et al. 2016). However, the molecular basis of T6P sensing and signaling awaits further elucidation. Research findings postulate that T6P may exert its functions by direct or indirect binding to SNRK1, thereby reducing SNRK1’s activity (Baena-González & Lunn 2020, Figueroa & Lunn 2016, Zhai et al. 2018, Zhang et al. 2009). Significantly, the characterization of various forms of TPS1 revealed its nuclear localization and the regulatory functions of its noncatalytic domains in the N- and C-termini, suggesting potential mechanisms of regulation by sucrose and T6P (Fichtner et al. 2020). In maize, meristem fate is regulated by TPPs, but the inflorescence branching regulated by TPPs (RAMOSA3 and TPP4) is uncoupled from enzymatic activity, implicating another possible sugar-sensor function (Claeys et al. 2019).

Besides glucose, sucrose, and T6P, other sugar signals and sensors have also been implicated in regulating plant transcription, metabolism, and development (Figure 2) (Li & Sheen 2016). An integrated study using a cell-based functional screen and genetic mutations has identified the nuclear-localized FRUCTOSE1–6-BIPHOSPHATASE (FBP) as a putative fructose sensor uncoupled from its catalytic activity (Cho & Yoo 2011). It will be interesting to determine how FBP may be connected to the fructose-specific signaling suppressor TF ANAC089 in the nucleus that shares downstream interactions with ABA and ethylene signaling pathways (Li et al. 2011). Another novel finding in plant sugar signaling is the recent demonstration that Arabidopsis SEC encodes a specific O-LINKED N-ACETYLGLUCOSAMINE (O-GlcNAc) TRANSFERASE (OGT). O-GlcNAcylation suppresses DELLA, which is a repressor of multiple TFs, for example BRASSINAZOLE RESISTANT1 (BZR1), PHYTOCHROME INTERACTING FACTOR4 (PIF4) and PIF5, and JASMONATE-ZIM-DOMAIN PROTEIN1 (JAZ1), which are involved in brassinosteroid, light, and jasmonate signaling, respectively (Zentella et al. 2016). Unexpectedly, DELLA is also mono-O-fucosylated by the O-fucosyltransferase encoded by SPY to activate DELLA repression (Zentella et al. 2017). Large-scale proteomic analyses by mass spectrometry have identified many regulatory proteins with either O-GlcNAcylation or phosphorylation, opening the way for future exploration of OGT functions in nutrient signaling (Xu et al. 2017).

TOR IN NUTRIENT AND ENERGY SIGNALING

TOR is an atypical but evolutionarily conserved Ser/Thr PK acting as a master signaling integrator to dynamically orchestrate cell metabolism, biogenesis, organ growth, and developmental transitions in response to nutrient, energy, hormone, and environmental cues (Figure 3) (Dobrenel et al. 2016, Ryabova et al. 2019, Shi et al. 2018, Wu et al. 2019). The plant TOR COMPLEX1 (TORC1) has two conserved regulatory subunits, LETHAL-WITH-SEC13-PROTEIN8 (LST8) and REGULATORY-ASSOCIATED-PROTEIN-OF-MTOR1 (RAPTOR). The generation of conditional tor mutants, plants with hyper- or hypoactive TOR expression, regulatory subunit mutants, and specific TOR chemical inhibitors has enabled significant progress in dissecting the plant TOR signaling network by circumventing the embryonic lethality of the null tor mutants (Caldana et al. 2013, Deprost et al. 2007, Dobrenel et al. 2016, Menand et al. 2002, Moreau et al. 2012, Ren et al. 2011, Ren et al. 2012, Ryabova et al. 2019, Shi et al. 2018, Xiong & Sheen 2012, Xiong et al. 2013). The conservation of the TOR-S6K (RPS6 KINASE)-RPS6 (RIBOSOMAL PROTEIN S6) phosphorylation cascade provides quantitative assays to monitor in vivo TOR activity based on phosphor-motif-specific antibodies (Dobrenel et al. 2016, Enganti et al. 2018, Schepetilnikov et al. 2013, Xiong & Sheen 2012).

Figure 3.

Figure 3

TOR acts as a central hub to integrate nutrient, energy, and environmental cues to orchestrate growth and development. The TOR complex is composed of TOR, RAPTOR, and LST8 and is promoted by the ATPase cochaperone complex TTT-WAC-RUVBL for dimerization and activation. ATP generated from the mitochondrial ETC, glucose, light, nutrients (N, P, or S), and amino acids promotes TOR activation. TOR phosphorylation of substrates (with orange circled Ps) activates (orange text) or represses (blue text) their functions in various biological processes. RPS6 KINASE (S6K) can also phosphorylate targets (with grey circled Ps). TOR is inhibited by SNRK1 and ABA. TOR phosphorylates and inhibits the ABA receptor PYL. Reciprocally, the ABA-PYL-SNRK2 cascade inhibits TOR activity through RAPTOR phosphorylation. Abbreviations: ABA, abscisic acid; ATG, AUTOPHAGY-RELATED; BIN2, BRASSINOSTEROID INSENSITIVE2; BR, brassinosteroid; COP1, CONSTITUTIVELY PHOTOMORPHOGENIC1; E2F, E2 TRANSCRIPTION FACTOR; EIF3h, EUKARYOTIC INITIATION FACTOR 3h; ETC, electron transport chain; G6P, glucose 6-phosphate; HXK, HEXOKINASE; LARP1, LA-RELATED PROTEIN1; LST8, LETHAL-WITH-SEC13-PROTEIN8; MRF1, MA3-DOMAIN-CONTAINING-TRANSLATION-REGULATORY-FACTOR1; PP2C, PROTEIN PHOSPHATASE 2C; PYL, PYRABACTIN RESISTANCE-LIKE; RAPTOR, REGULATORY-ASSOCIATED-PROTEIN-OF-MTOR1; ROP2, RHO-OF-PLANTS2; RPS6, RIBOSOMAL PROTEIN S6; RUVBL, RUVB-LIKE AAA ATPASE1; SNRK, SNF1-RELATED PROTEIN KINASE; TAP46, 2A PHOSPHATASE ASSOCIATED PROTEIN OF 46 KD; TOR, TARGET OF RAPAMYCIN; TTT, TEL2-TTI1-TTI2; WAC, WWDOMAIN CONTAINING ADAPTOR WITH COILED-COIL; YAK1, YET ANOTHER KINASE1.

Upstream Regulators of TOR Activity

At the critical checkpoint of photoautotrophic transition from seed to self-sufficient seedling in Arabidopsis, photosynthesis-derived glucose drives TOR signaling relays through glycolysis and mitochondrial bioenergetics to control root meristem and leaf primordia activation (Li et al. 2017, Xiong et al. 2013). Sugar and energy supplies from the mitochondria and chloroplasts are critical for the active TORC1 as a dimer promoted by the ATPase cochaperone complex TTT (TEL2-TTI1-TTI2)-WAC-RUVBL (RUVB-LIKE AAA ATPASE1) (Brunkard et al. 2020, Shi et al. 2018, Wu et al. 2019, Van Leene et al. 2019). Distinct from the activation of TOR in mammals, plant glucose-TOR signaling cannot be replaced by hormones and dictates very rapid (2-h) global transcriptional reprogramming of remarkable gene sets involved in central and secondary metabolism, the cell cycle, transcription, translation, signaling, transport, protein folding, and nutrient assimilation as well as the biosynthesis of nucleotides, lipids, amino acids, cell walls, lignins, and glucosinolates (Scarpin et al. 2020, Wu et al. 2019, Xiong et al. 2013).

Extensive studies in Arabidopsis plants have revealed that glucose metabolism and energy status play pivotal roles in activating TOR by interacting with light, hormone, clock, and other nutrient signals (Figure 3) (Mohammed et al. 2018, Wu et al. 2019). Although glucose alone is sufficient to activate root meristem proliferation via TOR signaling, glucose and light are synergistically required to activate cell proliferation by TOR signaling based on the expression of pCYCB1;1::GUS as a mitotic marker in leaf primordia (Li et al. 2017, Xiong et al. 2013). Light-enhanced translation is stimulated by white, far-red, and blue light perception via the phyA and CRY1/2 photoreceptors through the convergent CONSTITUTIVELY PHOTOMORPHOGENIC1 (COP1)-auxin-TOR-RPS6 signaling relay (Chen et al. 2018). The requirement for auxin signaling in TOR activation is mediated through Rho-like small GTPase RHO-OF-PLANTS2 (ROP2). Plants expressing constitutively active ROP2 contain high levels of active TOR. ROP2 physically interacts with and, when GTP bound, activates TOR in vitro. TOR activation in response to auxin is abolished in ROP-deficient rop2,4,6 plants, and green fluorescent protein (GFP) tagged TOR can associate with endosome-like structures in ROP2-overexpressing plants (Li et al. 2017, Schepetilnikov et al. 2017). Sugar-promoted hypocotyl elongation requires both TOR and brassinosteroid signaling mediated by the TF BZR1, which is degraded via autophagy suppressed by TOR (Zhang et al. 2016b).

In the SAM, sucrose and photoreceptors additively promote the expression of WUSCHEL (WUS) encoding a homeodomain master TF by suppressing COP1 for stem cell regulation in the organizing center, which is partially prevented by the specific TOR inhibitor AZD8055 (Pfeiffer et al. 2016). However, the expression of WOX5 related to WUS in the root quiescent center is not diminished in the tor-es mutant (Xiong et al. 2013). TOR activation stimulated by sugar, energy, and light signaling may exert differential regulations and functions in the RAM and SAM. It will be interesting to identify the molecular regulators mediating the TOR activation of WUS in the SAM. TOR is also an essential energy sensor to coordinate the circadian clock and plant growth. Nicotinamide treatment modulates the cytosolic ATP concentration monitored by an ATP biosensor and alters glucose-TOR signaling on period length, meristem activation, and root growth (N. Zhang et al. 2019).

Despite the lack of gene homologs encoding established sensors and transducers mediating TOR activation by amino acids in animals, sulfur (S), P, and N nutrients all influence plant TOR signaling (Figure 3) (Wu et al. 2019). For instance, Leu, Ile, and Val have been linked to TOR activation from the characterization of an Arabidopsis mutant that overaccumulated branched-chain amino acids with a reorganized actin cytoskeleton and endomembranes (Cao et al. 2019) or the inhibitory amino acids for the nighttime oxygen consumption rate stimulated by Pro and Ala (O’Leary et al. 2020). Importantly, a systematic study showed that TOR can be activated by diverse nitrogen signals, including nitrate, ammonium, and 15 ammino acids, via convergent ROP2 signaling to promote cell proliferation in the leaf primordium (Liu 2021 et al, in press).

S limitation is transduced to shoot TOR signaling and downregulation of glucose metabolism causing decreased rates of translation, lowered meristematic activity, and elevated rates of autophagy (Y. Dong et al. 2017). In Chlamydomonas, phosphate starvation sharply decreases LST8 abundance in TORC1, which is mediated by the TF PSR1 controlling phosphate-starvation responses (PSRs) (Couso et al. 2020). These new findings represent a fertile ground to investigate the diverse nutrient connections and downstream response mechanisms in TOR signaling unique to plants.

Downstream Effectors of the TOR Signaling Network

Direct and indirect TOR effectors have been identified in mediating the downstream TOR signaling network in plants. Systems, cellular, biochemical, and genetic analyses continue to uncover TOR phosphorylation of the TFs E2FA and E2FB, which are critical to S-phase gene activation. The finding reveals a previously unrecognized checkpoint in controlling cell cycle entry based on nutrient-energy status in meristems and primordia. The successful identification of low-abundance TFs offers an innovative approach for discovering novel TOR substrates with complex and combinatorial phosphorylation sites (Li et al. 2017, Xiong et al. 2013). Remarkably, plant TOR promotes protein translation by multiple mechanisms. Over 100 Arabidopsis genes encoding ribosomal proteins and protein synthesis machineries are activated by glucose-TOR signaling (Scarpin et al. 2020, Xiong et al. 2013). ChIP studies have revealed that nuclear TOR regulates rRNA synthesis by directly binding to the 45S rRNA promoter and the 5′ transcribed spacer DNA motifs (Ren et al. 2011). TOR directly phosphorylates TAP46 to inhibit PROTEIN PHOSPHATASE 2A (PP2A) and promote translation and N assimilation (Ahn et al. 2011). Moreover, S6K activated by TOR phosphorylates RPS6 and MA3-DOMAIN-CONTAINING-TRANSLATION-REGULATORY-FACTOR1 (MRF1) to enhance translation (Chen et al. 2018, Dobrenel et al. 2016, Enganti et al. 2018, Lee et al. 2017, Ren et al. 2012). Comprehensive phenotypic analyses of rps6a and rps6b mutants and genetic complementation support downstream roles for RPS6 in light- and nutrient-dependent TOR functions in root, leaf, and flowering regulation (Ren et al. 2012). S6K phosphorylation of MRF1 promotes its association with EUKARYOTIC INITIATION FACTOR4A-1 (EIF4A-1) and light polysomal fractions, which may reboot translation during the dark-to-light transition (Lee et al. 2017). Unexpectedly, TOR-S6K signaling promotes translation reinitiation of uORF-containing mRNAs via phosphorylation of EUKARYOTIC INITIATION FACTOR 3h (EIF3h) (Ryabova et al. 2019, Schepetilnikov et al. 2013, Schepetilnikov et al. 2017). Phosphorylation of plant LA-RELATED-PROTEIN1 (LARP1) is inhibited by Torin2. Rigorous bioinformatics analyses suggest that LARP1 controls translation of conserved and plant-specific mRNAs with 5′-TERMINAL OLIGOPYRIMIDINE (Scarpin et al. 2020, Van Leene et al. 2019).

Recent discoveries have revealed that TOR or S6K phosphorylation can mediate the repression of substrates in surprisingly diverse biological processes in plants (Figure 3). The S6K-BIN2 (BRASSINOSTEROID INSENSITIVE2) relay promotes brassinosteroid signaling (Xiong et al. 2017). Direct TOR phosphorylation blocks the PK YAK1 as a negative regulator in cell cycle and flowering (Barrada et al. 2019, Forzani et al. 2019), prevents ATG1 or ATG13 activation in autophagy (Soto-Burgos et al. 2018, Van Leene et al. 2019), and inhibits the ABA receptor PYLs involved in stress responses (P. Wang et al. 2018). ABA-activated SNRK2 phosphorylates RAPTOR to inactivate TORC1, indicating reciprocal negative regulatory loops to balance TOR-growth and ABA-stress signaling (P. Wang et al. 2018). Interestingly, ETHYLENE-INSENSITIVE PROTEIN2 (EIN2), a previously defined key regulator in ethylene signaling, is phosphorylated by TOR to prevent its nuclear translocation and repressor function in glucose-TOR signaling, which is uncoupled from the ethylene responses mediated by the CONSTITUTIVE-TRIPLE-RESPONSE1-EIN2 link through specific phosphorylation sites (Fu et al. 2021). These findings suggest that central signaling hubs can be shared but differentially modulated by diverse hormone and nutrient signaling pathways using distinct phosphorylation codes that can be specified by upstream PKs. Unexpectedly, TOR dynamically represses cell–cell transport and sugar allocation via plasmodesmata during the sink-to-source leaf transition (Brunkard et al. 2020).

Important advances in our understanding of the glucose-TOR signaling network have been greatly facilitated by comprehensive and comparative RNA sequencing (RNA-seq), ribosome sequencing (Ribo-seq), and phosphoproteomics and interactomics analyses using Arabidopsis seedlings or cultured cells. These large-scale studies provide rich data resources for future exploration of both upstream and downstream network components and TOR substrate candidates controlling conserved or unique plant processes (Scarpin et al. 2020, Van Leene et al. 2019; Xiong et al., 2013).

SNRK1 IN STRESS AND ENERGY SIGNALING

To adapt to the daily light–dark cycle or unpredictable environmental stresses that compromise photosynthesis and respiration, plants need to control energy homeostasis for sustained growth and survival during sugar and energy deprivation. Studies on the regulation of universal DARK-INDUCIBLE (DIN) genes have provided crucial evidence to molecularly connect the glucose-repressible DIN transcription to the Arabidopsis energy sensor PKs KIN10 and KIN11 under diverse stress conditions (Baena-González et al. 2007, Sheen 2014). The Arabidopsis SNRK1 heterotrimeric complexes, consisting of the catalytic α subunit (KIN10/11) and the regulatory subunits (KINβ1/2/3 and the plant-specific KINβγ), share evolutionarily conserved structure in yeast SUCROSE-NONFERMENTING1 (SNF1) and human AMP-ACTIVATED PROTEIN KINASE (AMPK) (Figure 4). The Arabidopsis kin10,11 and kinβγ mutants are embryonic lethal, hindering research progress in the past (Baena-González et al. 2007, Broeckx et al. 2016, Emanuelle et al. 2015, Ramon et al. 2013, Ramon et al. 2019).

Figure 4.

Figure 4

SNRK1 plays a central role in sensing nutrient deprivation and stress to promote catabolism but inhibit anabolism and growth. The SNRK1 complex is composed of catalytic a (KIN10/11) and regulatory β and βγ subunits. SNRK1 is activated by darkness, stress, starvation, hypoxia, and SNAKs but inhibited by sugars, metabolites, and FLZ repressors. SNRK1 phosphorylation positively (orange text) or negatively (blue text) regulates targets in broad biological processes. The catalytic a subunits (KIN10/11) can translocate to the nucleus to phosphorylate various TFs. KIN10 represses EIN3, MYC2, WRI1, IDD8, and PIF4 but activates FUS3, bZIP63, and SPCH. SNRK1 also phosphorylates and inhibits cytosolic enzymes, including SPS, NR, TPS7/8/9, and HMGR; translation regulators ISO4E, ISO4G, and ELF4E; and other signaling proteins in various biological responses. SNRK1 antagonizes TOR complex signaling by phosphorylating RAPTOR. Abbreviations: ABA, abscisic acid; ABI1, ABA INSENSITIVE1; ATG1, AUTOPHAGY RELATED1; bZIP, BASIC LEUCINE ZIPPER; EIN3, ETHYLENE INSENSITIVE3; ELF4E, EUKARYOTIC TRANSLATION INITIATION FACTOR 4E; F2KP, fructose-2,6-bisphosphatase; FLZ, FCS-LIKE ZINC FINGER; FUS3, FUSCA3; G1P, glucose 1-phosphate; G6P, glucose 6-phosphate; HMGR, 3-HYDOXY-3-METHYLGLUTARYL–COA REDUCTASE; IDD8, INDETERMINATE DOMAIN8; KIN10/11, KINASE10/11; MPK, MITOGEN-ACTIVATED PROTEIN KINASE; NR, NITRATE REDUCTASE; P, phosphorylation; PIF4, PHYTOCHROME INTERACTING FACTOR4; PP2CA, PROTEIN PHOSPHATASE 2CA; PTP, PROTEIN TYROSINE PHOSPHATASE; RAPTOR, REGULATORY-ASSOCIATED-PROTEIN-OF-MTOR1; SNAK, SNRK1-ACTIVATING KINASE; SNRK1, SNF1-RELATED KINASE1; SPCH, SPEECHLESS; SPS, SUCROSE PHOSPHATE SYNTHASE; T6P, trehalose 6-phosphate; TF, transcription factor; TOR, TARGET OF RAPAMYCIN; TPS, T6P SYNTHASE; UDP, uridine diphosphate; WRI1, WRINKLED1.

The SNRK1 Transcription Network

Remarkably, the transient expression of KIN10 alone controls the activation and repression of thousands of genes involved in transcription, signaling, catabolism, and anabolism in a cell-based assay. These KIN10 target genes are either positively correlated with stress and starvation genes or negatively correlated with the genes activated by glucose, sucrose, and elevated CO2 levels identified in cultured cells, seedlings, and adult leaves from genome-wide analyses. The high correlation between KIN10 target genes and the transcriptome reprogramming in multiple physiological responses provides compelling evidence to support a critical and convergent role of KIN10 in mediating sugar- or energy-deprivation responses triggered by darkness, starvation, or diverse stress conditions. Transgenic KIN10 overexpression confers enhanced starvation tolerance and life span extension and alters plant architecture and developmental transitions. Conditional kin10,11 deficiency abrogates the transcriptional switch in darkness or stress signaling and impairs starch mobilization at night but also arrests growth by unclear mechanisms (Baena-González et al. 2007, Broeckx et al. 2016, Pedrotti et al. 2018, Sheen 2014).

Significant research advances have revealed how direct phosphorylation by KIN10/11 promotes or represses specific TFs involved in various biological functions, primarily plant-specific and not previously recognized (Figure 4). The best studied example demonstrated that KIN10 phosphorylates bZIP63 to enhance heterodimerization with TF bZIP-S and activate target genes in alternative mitochondrial metabolic pathways to ensure plant survival in extended darkness (Mair et al. 2015, Pedrotti et al. 2018). The TPS1-T6P-KIN10-bZIP63-PRR7 (PSEUDO-RESPONSE REGULATOR7) signaling cascade regulates the sucrose-dependent circadian phase (Frank et al. 2018). Interestingly, FUSCA3 (FUS3), a key regulator in embryogenesis, is phosphorylated by KIN10 to modulate KIN10-dependent postembryonic development and flowering (Tsai & Gazzarrini 2012). Phosphorylation of SPCHLESS by KIN10 promotes sucrose-dependent stabilization and stomata development (Han et al. 2020). Many studies have shown that KIN10 can also directly phosphorylate and repress TFs controlling diverse biological processes, including MYC2 in the salt-stress response (Im et al. 2014), INDETERMINATE DOMAIN8 (IDD8) in flowering (Jeong et al. 2015), EIN3 in senescence (Kim et al. 2017), WRINKLED1 (WRI1) in lipid biosynthesis (Zhai et al. 2017), and PHYTOCHROME INTERACTING FACTOR4 (PIF4) in thermoresponsive hypocotyl elongation requiring TPS1-T6P signaling (Hwang et al. 2019). KIN10 phosphorylation mediates diverse regulatory mechanisms, reducing MYC2 and EIN3 protein stability or IDD8 transcriptional activity.

The SNRK1 Signaling Network

Besides plant-specific transcriptional controls, similar to human AMPK, KIN10 phosphorylates ATG1 to activate autophagy (Chen et al. 2017) and RAPTOR to repress TOR signaling (Nukarinen et al. 2016). Quantitative phosphoproteomics, proteomics, and metabolomics analyses have been conducted with Arabidopsis plants in extended darkness using a conditional kin10,11 mutant (Nukarinen et al. 2016, Pedrotti et al. 2018) or submergence stress (hypoxia) based on dominant-negative KIN10 plants (Cho et al. 2016). These complementary approaches confirmed the SNRK1-dependent phosphorylation of multiple metabolic enzymes in the cytosol in vivo, supporting an SNRK1-mediated redirection of carbon flux toward glycolysis and mitochondrial respiration. Novel PROTEIN TYROSINE PHOSPHATASE (PTP)-MITOGEN-ACTIVATED PROTEIN KINASE (MPK) signaling (Cho et al. 2016) and phosphorylation of photosynthesis-related proteins in the chloroplast were also detected (Nukarinen et al. 2016). It remains to be determined whether SNRK1 enters the chloroplast and phosphorylates organelle-specific targets (Ruiz-Gayosso et al. 2018). Surprisingly, KIN10 phosphorylation represses general translation via EIF4E and EIFISO4E inhibition (Bruns et al. 2019) but targets plant-specific EIFISO4G1 to enhance the translation of mRNAs for hypoxia and submergence tolerance (Cho et al. 2016, Cho et al. 2019) (Figure 4).

Novel SNRK1 Regulation

Despite structural conservation, the regulation of SNRK1 exhibits many features not found in human AMPK. SNRK1 is insensitive to AMP/ADP activation and resistant to T-loop dephosphorylation by protein phosphatases (PPs). KIN10 is constitutively active by T-loop autophosphorylation. Nuclear KIN10 activates a DIN6-LUC reporter but is repressed by myristoylated KINβ1/2 in cell-based assays. Importantly, transgenic plants with altered KIN10 localization modify metabolic stress responses and display novel phenotypes in shoots and roots (Broeckx et al. 2016, Ramon et al. 2019). Significantly, T6P appears to inhibit KIN10 by directly binding and decreasing its interaction, phosphorylation, and activation by upstream SNRK1-ACTIVATING KINASE (SNAK)1/2 in vitro (Shen et al. 2009, Zhai et al. 2018). However, the binding site remains unknown and the Kd is higher than the endogenous T6P level estimated at low micromolar concentrations as a signaling molecule in plants (Baena-González & Lunn 2020, Figueroa & Lunn 2016). Other sugar metabolites can also repress SNRK1 by unknown mechanisms (Nunes et al. 2013, Zhang et al. 2009). KINβ3 and KINβγ are detected in the chloroplast and SNRK1 activity is modulated by maltose (Ruiz-Gayosso et al. 2018). KIN10 activation and ABA signaling share overlapping target genes, and specific PP2Cs appear to serve as the converged hub for the coordinated activation of ABA and energy/stress signaling pathways (Jossier et al. 2009, Rodrigues et al. 2013). FSC-LIKE ZINC FINGER (FLZ)6/10 interacting with SNRK1 in the cytosol represents plant-specific repressors (Jamsheer K et al. 2018) (Figure 4). Future research may uncover new mechanisms of SNRK1 regulation to clarify additional layers of molecular controls in the dynamic sugar/energy/stress signaling networks.

NITRATE SENSING AND SIGNALING

Plants have evolved sophisticated sensing and signaling networks to respond and adapt to fluctuating nitrate availability from starvation to broad concentrations in the soil (Figure 5). Using nitrate reductase–deficient plants, nitrate was demonstrated to act as a signaling molecule uncoupled from its metabolism, and nitrate-induced TF ANR1 regulates root architecture (Scheible et al. 1997, Wang et al. 2004, Zhang & Forde 1998). Extensive transcriptome analyses and network modeling have identified thousands of nitrate-responsive genes and key TFs for the dynamic and spatiotemporal control of nitrogen uptake and assimilation, carbon metabolism, sulfate assimilation, transcription regulation, and hormone metabolism (Canales et al. 2014, Scheible et al. 2004, Vidal et al. 2013, Vidal et al. 2020, Walker et al. 2017, Wang et al. 2004, Y.-Y. Wang et al. 2018). Primary nitrate responses (PNRs) occur in minutes and activate the synthesis and signaling of the plant hormones cytokinin and auxin but degrade ABA. Nitrate-induced cytokinin promotes root meristem proliferation and acts as a systemic signal to promote shoot growth (Liu et al. 2017, Naulin et al. 2020, Sakakibara 2020, Vidal et al. 2013, Yan et al. 2016) (Figure 5).

Figure 5.

Figure 5

The nitrate sensing and signaling network. Nitrate triggers the nuclear Ca2+-CPK-NLP6/7 cascade and induces genes encoding activators (orange text) and repressors (blue text) of the PNRs. Nitrate-activated genes encoding TFs, including NIGT1s and LBD37/38/39, repress N-starvation genes. Nitrate activates the synthesis and signaling of the plant hormones cytokinin and auxin but degrades ABA. The activity of the nitrate transceptor NRT1.1 is regulated by CBL1/9-CIPK23, which is inhibited by ABI2/PP2C. Nitrogen starvation activates the CLE1/3/4/7 peptides to repress lateral root development and triggers CEPs to move to the shoot to activate CEPRs. Activated CEPD1/2 proteins in the shoot then translocate to the root to enhance nitrate uptake. Abbreviations: ABA, abscisic acid; ABI2, ABA INSENSITIVE2; AFB, AUXIN-SIGNALING F-BOX; ANR, ARABIDOPSIS NITRATE REGULATED1; CBL, CALCINEURIN B-LIKE; CEP, C-TERMINALLY ENCODED PEPTIDE; CEPD, CEP DOWNSTREAM; CEPR; CEP RECEPTOR; CIPK, CBL-INTERACTING PROTEIN KINASE; CLE, CLAVATA3/EMBRYO SURROUNDING REGION-RELATED; CPK, CALCIUM DEPENDENT PROTEIN KINASE; CRF, CYTOKININ-RESPONSE FACTOR; CYP, cytochrome P450; IPT3, ISOPENTENYLTRANSFERASE3; LBD, LATERAL ORGAN BOUNDARIES DOMAIN; NAC, NAM, ATAF, CUC; NIGT1, NITRATE-INDUCIBLE GARP-TYPE TRANSCRIPTIONAL REPRESSOR1; NLP, NIN-LIKE PROTEIN; NRT1.1, NITRATE TRANSPORTER 1.1; PNR, primary nitrate response; PP2C, PROTEIN PHOSPHATASE2C; TAR, TRYPTOPHAN AMINOTRANSFERASE RELATED; TCP, TEOSINTE BRANCHED 1/CINCINNATA/PROLIFERATING CELL FACTOR; TF, transcription factor; TGA, TGACG SEQUENCE-SPECIFIC BINDING PROTEIN1.

The Nitrate Transcription Network

The characterization of nin-like-protein7 (nlp7) or dominant-negative nlp6 mutants and NIN-LIKE PROTEIN (NLP)6/7 binding to nitrate-responsive elements has defined master TFs in PNRs (Castaings et al. 2009, Konishi & Yanagisawa 2013, Marchive et al. 2013). Time series ChIP-sequence (ChIP-seq) and DNA-adenine-methyltransferase-identification experiments efficiently captured genomewide transient and stable NLP7 target genes (Alvarez et al. 2020). NLP8 plays a unique role in nitrate-stimulated seed germination by activating CYP707A2 for ABA degradation (Yan et al. 2016). Nitrate-induced TF genes encoding activators (ANR1, TGA1/4, and NAC4) and repressors [LATERAL ORGAN BOUNDARIES DOMAIN (LBD)37/38/39, NITRATE-INDUCIBLE GARP-TYPE TRANSCRIPTIONAL REPRESSOR (NIGT)1/2, and CYTOKININ-RESPONSE FACTOR (CRF)4] of nitrate responses have been extensively characterized for unique or overlapping functions in nitrate signaling and root development (Canales et al. 2014, Vidal et al. 2013, Vidal et al. 2020, Y.-Y. Wang et al. 2018). LBD37/38/39 and NIGT1/2 dynamically fine-tune nitrate-regulated transcription (Kiba et al. 2018, Maeda et al. 2018, Medici et al. 2015). The rapid, dynamic, dose-responsive, and hierarchical features of the nitrate regulatory network are demonstrated by the integration of TF-target gene data sets from time series genomewide RNA-seq, ChIP-seq, and DNase-seq experiments with computational and network analyses. TGA1 and CRF4 are proposed TF hubs, and TGA1 overexpression strongly promotes the root system (Alvarez et al. 2019, Brooks et al. 2019, Swift et al. 2020, Varala et al. 2018). Mutant analyses based on nitrogen utilization efficiency in rice have identified key TFs, including GROWTH-REGULATING FACTOR4 promoting N uptake and N/C metabolism (Li et al. 2018) and NITROGEN-MEDIATED TILLER GROWTH RESPONSE5 promoting tilling/branching (Wu et al. 2020). Large-scale yeast one-hybrid (Y1H) screens between TFs and promoters in N metabolism have proposed a comprehensive N-related transcriptional regulatory network, providing valuable resources for future research (Gaudinier et al. 2018). Research advances are crucial to connecting the expansive transcription network to upstream nitrate sensing and signaling mechanisms (Figure 5).

Nitrate Sensing and Signaling

NRT1.1/CHL1/NITRATE TRANSPORTER1/PEPTIDE TRANSPORTER (NPF)6.3 was the first identified dual-affinity nitrate transporter and transceptor, as the P492L mutant shows compromised nitrate transport but retains nitrate signaling (Ho et al. 2009). T101 in NRT1.1 is phosphorylated by calcium-regulated CALCINEURIN B-LIKE (CBL)1/9-CBL-INTERACTING PROTEIN KINASE (CIPK)23 and modulates nitrate transport, sensing, and signaling (Figure 5) (Bouguyon et al. 2015, Ho et al. 2009, Y-Y. Wang et al. 2018, X. Zhang et al. 2019). CBL1/9-CIPK23 is inactivated by a specific ABA INSENSITIVE (ABI)2/PP2C, which integrates nitrate and ABA-stress signaling (Léran et al. 2015). In rice, nitrate promotes OsNRT1.1B-mediated activation of the E3 ubiquitin ligase OsNBIP1 to degrade a repressor OsSPX4 and activate OsNLP3 (Hu et al. 2019). New insights are necessary to distinguish molecular mechanisms for different NRT1.1 states and subcellular localizations in extracellular or intracellular nitrate sensing and signaling. Recent studies also suggest that NRT1.13 may be a potential nitrate transceptor in regulating flowering and branching (Chen et al. 2021).

Calcium signaling in PNR was first indicated by applying inhibitors of calcium channels and effectors to diminish nitrate-induced transcription (Liu et al. 2017, Liu et al. 2020, Riveras et al. 2015, Sakakibara et al. 1997). The use of an ultrasensitive calcium biosensor was critical to demonstrating that nitrate triggers calcium signaling in the cytosol and nucleus. A targeted functional screen of CALCIUM-DEPENDENT PKs (CPKs) identified group III CPK10/30/32 as redundant key regulators in PNR, which are impaired in a conditional icpk triple mutant. S205 phosphorylation by CPK10/30/32 is essential for nuclear NLP7 retention to activate the nitrate-responsive transcriptome and optimal root–shoot development (Figure 5). Intriguingly, a calcium channel CYCLIC NUCLEOTIDE-GATED ION CHANNEL15 (CNGC15) has been reported to interact with NRT1.1 and facilitate nitrate-induced calcium influx from the plasma membrane at the root tip (Wang et al. 2021) Future exploration of dynamic protein phosphorylation events and nitrate-regulated genes encoding PKs and PPs may reveal novel mechanisms in PNR (Liu et al. 2017, Liu et al. 2020).

Nitrogen-Starvation Responses

Nitrogen deficiency in plants triggers changes in root foraging, transcriptome, metabolism, and nitrogen uptake (Krapp et al. 2011). Nuclear TEOSINTE BRANCHED1/ CINCINNATA/PROLIFERATING CELL FACTOR (TCP)20 and NLP6/7 form heterodimers and regulate nitrate-assimilation genes and primary root meristems under N starvation (Guan et al. 2017). Genome-wide Y1H screens identified four related NIGT1s as TF repressors of a N-starvation-responsive promoter. NLP7 activates NIGT1 transcription, but NIGT1s suppress NLP7 and the N-starvation-responsive transcriptome to optimize dynamic N acquisition, remobilization, and signaling (Kiba et al. 2018, Maeda et al. 2018). Nitrogen starvation also triggers peptide genes for local and systemic N-demand signaling. CLE1/3/4/7 peptides are induced in the root vasculature and repress lateral root development (Araya et al. 2014). Split-root and grafting assays indicate that secreted C-TERMINALLY ENCODED PEPTIDES (CEPs) are induced and move from the N-starved roots through the xylem stream to shoots where the CEPs are perceived by receptor kinases (Tabata et al. 2014). Phloem-specific C-TERMINALLY ENCODED DOWNSTREAM PEPTIDE (CEPD)1/2 proteins are then induced in shoots and translocated to roots through phloem to enhance NRT2.1 for high-affinity nitrate uptake (Figure 5) (Ohkubo et al. 2017). Understanding the sensing mechanisms in N-deficiency responses and their integration with the complex nitrate signaling network represents future research challenges.

PHOSPHATE SENSING AND SIGNALING

The acquisition and assimilation of inorganic phosphate is essential to support plant growth. As phosphate availability is often limited in the soil, plants have evolved PSRs mediated by complex sensing and signaling networks. Universal and organ-specific PSRs occur in shoots and roots as well as through root-to-shoot or shoot-to-root systemic communication and regulation. In shoots, PSRs reduce photosynthesis and leaf expansion but promote anthocyanin, sugar, and starch accumulation and early flowering. In roots, PSRs lead to primary root arrest and meristem exhaustion but promote extensive root exudates (organic acids and phosphatases) to solubilize phosphate as well as lateral root and root-hair development to increase the surface area for phosphate forage. Universal PSRs stimulate uptake, translocation, recycling, and remobilization of phosphate and membrane phospholipid remodeling (Figure 6a) (Chien et al. 2018, Ham et al. 2018, Puga et al. 2017, Raghothama 2000).

Figure 6.

Figure 6

(a) PSRs in shoots and roots. Pi is taken up by PHT1 transporters and converted to the signaling molecule InsP8, which is perceived by the SPX sensor to inhibit PHR1. Under Pi starvation, the key TFs PHR1/PHL induce transcriptome reprogramming to promote PSRs in both shoots and roots as well as shoot-to-root mobile mRNAs and miRNA. Shoot miR399 inhibits root PHO2, encoding a ubiquitin E2 conjugase for the degradation of PHT1 and PHO1 transporters. In roots, low external Pi induces iron-dependent ROS by derepressing ferroxidase LPR1/2 and promoting nuclear TF STOP1 to enhance malate transporter ALMT1. Activated CLE14 peptide inhibits the key TFs SHR/SCR and WOX5 as well as auxin transporter PINs and arrests the primary root meristem. Root PSRs also induce the hormones auxin and cytokinin to promote lateral root and root hair growth. (b) High N levels promote PSRs by multiple mechanisms. Nitrate enhances TF PHR1 by stabilizing the protein and degrading the SPX repressor. The nitrate-NRT1.1-CPK-NLP cascade or the nitrate-NRT1.1-NBIP-SPX-NLP relay activates NIGT1s to repress SPXs and PHO2 encoding repressors of the PSR but activates PHT1s to enhance Pi uptake. Nitrate also enhances the PSR by a PHO2-dependent mechanism independent of PHR1. Abbreviations: ALMT1, ALUMINUM-ACTIVATED MALATE TRANSPORTER1; ALS3, ALUMINUM SENSITIVE3; AUX1, AUXIN TRANSPORTER PROTEIN1; CLE14, CLAVATA3/ESR-RELATED14; CPK, CALCIUM-DEPENDENT PROTEIN KINASE; ER, endoplasmic reticulum; InsP8, inositol pyrophosphate8; LHW, LONESOME HIGHWAY; LPR, LOW PHOSPHATE ROOT; miRNA, microRNA; mRNA, messenger RNA; NBIP, NRT1.1B INTERACTING PROTEIN; NIGT1, NITRATE-INDUCIBLE GARP-TYPE TRANSCRIPTIONAL REPRESSOR1; NLA, NITROGEN LIMITATION ADAPTATION; NLP, NIN-LIKE PROTEIN; NRT1.1, NITRATE TRANSPORTER 1.1; P, phosphorylation; PDR2, PHOSPHATE DEFICIENCY RESPONSE2; PHF1, PHOSPHATE TRANSPORTER TRAFFIC FACILITATOR1; PHL, PHR1-LIKE; PHO, PHOSPHATE; PHR, PHOSPHATE STARVATION RESPONSE1; PHT1, PHOSPHATE TRANSPORTER1; Pi, phosphate; PIN, PIN-FORMED; PSR, phosphate-starvation response; ROS, reactive oxygen species; SCR, SCARECROW; SHR, SHORT-ROOT; SPX, SYG1/Pho81/XPR1; STAR1, SENSITIVE TO ALUMINUM RHIZOTOXICITY1; STOP1, SENSITIVE TO PROTON TOXICITY1; TF, transcription factor; TIR1, TRANSPORT INHIBITOR RESPONSE1; TMO5, TARGET OF MONOPTEROS5; Ub, ubiquitination; VIH, VIP-HOMOLOG; WOX5, WUSCHEL RELATED HOMEOBOX5.

Internal Phosphate Sensing and Signaling

Comprehensive characterization of the ARABIDOPSIS PHOSPHATE-STARVATION-RESPONSE1 (phr1) and PHR-LIKE (phl) mutants has established that PHOSPHATE STARVATION RESPONSE (PHR)1 and related PHLs are central TFs controlling transcriptome reprogramming and a broad range of PSRs in shoots and roots (Bustos et al. 2010, Rubio et al. 2001, Sun et al. 2016, Thibaud et al. 2010). In Arabidopsis and rice, the evolutionarily conserved SPX-domain repressors interact with and inhibit PHR1 in a phosphate-dependent manner. Consistently, spx1,2 mutants enhance PSRs, whereas SPX1-overexpression plants reduce PSRs (Puga et al. 2014, Wang et al. 2014). Comprehensive structural, biochemical, genetic, and phenotypic analyses provide compelling evidence that SPXs are sensors that bind to InsP8 as a metabolite signal indicator of intracellular phosphate levels. Inositol pyrophosphate kinases such as VIP-HOMOLOG1/2 (VIH1/2) play a key role in the phosphate-to-InsP8 signal relay as vih1,2 plants exhibit constitutive PSR despite abundant phosphate (Dong et al. 2019, Wild et al. 2016, Zhu et al. 2019) (Figure 6a).

The PSR involves extensive systemic signaling between shoots and roots. Older source leaves appear to be the first to suffer phosphate deprivation, and PSRs are subsequently initiated in both shoots and roots. Protein-encoding mRNAs and microRNAs (miRNAs) can play regulatory roles through long-distance communications (Figure 6a). PSR-induced miRNA399 expression occurs earlier in shoots during the PSR and moves through phloem to roots to suppress mRNAs encoding PHO2, a ubiquitin E2 conjugase, which targets the PHO1 exporter and PHT1 transporters for ubiquitination and degradation (Bari et al. 2006, Chiou et al. 2006, Huang et al. 2013, Lin et al. 2008, Liu et al. 2012). The mRNA encoding an E3 ligase NITROGEN LIMITATION ADAPTATION (NLA) targeting plasma membrane PHT1s is suppressed by miR827 to elevate cytosolic phosphate levels during PSR (Kant et al. 2011, Lin et al. 2013). Unexpectedly, the EXS domain of PHO1 is sufficient to provide an enigmatic root-to-shoot signal to attenuate phosphate signaling despite low internal phosphate levels in shoots (Wege et al. 2016). Phloem fluid transcriptome analyses discovered abundant mobile mRNAs in Arabidopsis and cucumber, which require thorough investigation of their functions in the PSR (Thieme et al. 2015, Zhang et al. 2016a).

External Phosphate Signaling

Transcriptome analyses from split-root experiments support root-specific responses to extracellular phosphate status independent of intracellular phosphate concentrations, which are modulated systemically (Bustos et al. 2010, Thibaud et al. 2010). Primary root arrest is the consequence of iron-dependent reactive oxygen species (ROS) signaling triggered by low external phosphate levels (Figure 6a). Genetic analyses identified LOW-PHOSPHATE-ROOT1/2 (LPR1/2) encoding cell-wall targeted ferroxidases in promoting ROS. LPR1/2 are modulated by three upstream negative regulators, P5-type ATPase PHOSPHATE DEFICIENCY RESPONSE (PDR)2 and ABC transporter complex ALS3/STAR1. Upon Pi limitation, apoplastic iron at low pH stimulates the rapid nuclear accumulation of the TF STOP1 and activates a malate transporter gene ALMT1 to promote local malate-Fe3+ and ROS signaling (Balzergue et al. 2017, Godon et al. 2019, Mora-Macías et al. 2017). Derepressed LPR1/2 levels trigger cell-specific Fe3+, ROS, and callose accumulation, which impairs TF SHR movement and primary root growth in the meristem and elongation zones (Balzergue et al. 2017, J. Dong et al. 2017, Mora-Macías et al. 2017, Müller et al. 2015). Finally, activated CLE14 peptide signaling switches cell differentiation in the stem cell niche via CLAVATA2 (CLV2)/PERCEPTION OF ARABIDOPSIS DANGER SIGNAL2 (PEPR2) receptors, which eventually exhausts the primary root meristem and diminishes the key TFs WOX5, SCARECROW (SCR), and SHORT ROOT (SHR) as well as PIN1 auxin transporters (Gutiérrez-Alanís et al. 2017).

Integrative Phosphate Signaling

PSRs are intimately integrated with plant hormones and other nutrients (Figure 6a) (Chien et al. 2018, Ham et al. 2018, Puga et al. 2017). To forage phosphate resources in the topsoil, the PSR promotes auxin synthesis, transport, and signaling to enhance lateral root initiation and elongation by activating an auxin receptor gene TIR1 in root pericycle cells (Pérez-Torres et al. 2008). PSR-stimulated root-hair elongation requires auxin from the root tip and is significantly reduced in the phr1/phl (Bustos et al. 2010), auxin synthesis taa1, and auxin transport aux1 mutants (Bhosale et al. 2018, Bustos et al. 2010). However, the enhanced root-hair density requires the auxin-activated TFs TARGET OF MONOPTEROS 5 (TMO5)/LONESOME HIGHWAY (LHW) and cytokinin synthesis in the xylem but signaling in the epidermis (Wendrich et al. 2020).

Exciting findings have uncovered the critical role of nitrate signaling in the PSR by multiple mechanisms (Figure 6b). Nitrate controls PHR1 accumulation and is indispensable in transcriptome reprogramming during PSR. The nitrate-NRT1.1-CPK-NLP7-NIGT1 cascade suppresses PHO2 expression, but the nitrate requirement can be bypassed in pho2 independent of PHR1 and NRT1.1 (Medici et al. 2019). Nitrate- and PHR1-activated NIGT1s also suppress SPXs encoding a pivotal repressor but activate PHT1s in the PSR (Kiba et al. 2018, Maeda et al. 2018, Medici et al. 2015, Medici et al. 2019, Ueda et al. 2020). In rice, nitrate stimulates OsNRT1.1B-OsNBIP1interaction and OsSPX4 degradation, which leads to nuclear translocation and the activation of OsPHR2 and OsNLP3, a pathway that is likely conserved in other plants (Hu et al. 2019). As phosphate uptake requires high energy, sugars are essential to promote PSR-mediated transcriptome changes by unknown signaling mechanisms. These findings reveal the priority for nutrient acquisition, as PSRs are not activated until both sugars and nitrate are available (Lei et al. 2011, Medici et al. 2019, Müller et al. 2007).

CHALLENGES AND OUTLOOK

Plant nutrient signaling networks form the foundation of the global food chain and support interconnected ecosystems. The fixation of CO2 and the release of O2 through photosynthesis have shaped the evolution of all lives in the history of Earth. Advances in our knowledge of the global cycles of C, N, and P and their intertwined nutrient regulatory networks in plants may offer solutions for agricultural innovations as well as mitigating climate changes and balancing ecosystems.

Despite significant research progress and discoveries, many questions remain regarding how the key nutrient, energy, and metabolite sensors and signaling regulators function and integrate into the interdependent nutrient signaling networks as well as interconnected sensor coordination. Diverse sensors perceiving physiological nutrient and metabolite signals with distinct specificity and affinity at broad concentration ranges are only beginning to emerge. The main challenges are to uncouple precise sensing and signaling functions from nutrient transport and metabolic processes, which are tightly intertwined. Specific chemical inhibitors or activators as well as live-imaging and time series analyses of sensitive and specific reporters and biosensors may circumvent mutant limitations due to lethality, redundancy, and long-term pleiotropic effects. Thoughtfully designed global profiling experiments and genome-wide high-throughput screens should enable the discovery of more nutrient sensors, direct downstream targets, subcellular actions, and novel molecular wiring and mechanisms.

Investigation of the complex nutrient, hormone, peptide, and metabolite interactions in organ-, tissue-, and cell-specific responses mediated by local or systemic signaling will lead to new surprises. Both nitrate and phosphate signaling rely on sugars by largely unknown molecular mechanisms. Future discoveries should encompass universal sensors, signaling components, and regulatory mechanisms in multicellular organisms from plants to humans. Unique and specialized regulators in diverse plant species will also be discovered. Understanding and exploration of the emerging relationships of C, N, and P and their integration with intrinsic regulators and diverse environmental cues may inspire new and tailored designs for beneficial or enhanced traits in nutrient uptake, assimilation, mobilization, and utilization and storage while uncoupling stress responses compromising growth and reproduction. Sophisticated plans and practices in agricultural applications for more efficient fertilizer acquisition and usage may help realize the ultimate goals of sustainable agriculture and ecosystems.

ACKNOWLEDGMENTS

We apologize to the many authors whose important contributions could not be cited because of the expanded scope and space limitation. We thank Hsing-Yi Cho for her contribution to Figure 1. This work was supported by National Institutes of Health grants GM060493 and GM129093, National Science Foundation grants, and Special Project (PJ009106) Rural Development Administration-Korea to J.S. K.H.L. is supported by Northwest Agricultural & Forestry University start-up funds, and National Natural Science Foundation of China grants NSFC-31670246 and NSFC-31870227.

Footnotes

DISCLOSURE STATEMENT

The authors are not aware of any affiliations, memberships, funding, or financial holdings that might be perceived as affecting the objectivity of this review.

LITERATURE CITED

  1. Aguilera-Alvarado GP, Sánchez-Nieto S. 2017. Plant hexokinases are multifaceted proteins. Plant Cell Physiol. 58:1151–60 [DOI] [PubMed] [Google Scholar]
  2. Ahn CS, Han J-A, Lee H-S, Lee S, Pai H-S. 2011. The PP2A regulatory subunit Tap46, a component of the TOR signaling pathway, modulates growth and metabolism in plants. Plant Cell 23:185–209 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Alvarez JM, Moyano TC, Zhang T, Gras DE, Herrera FJ, et al. 2019. Local changes in chromatin accessibility and transcriptional networks underlying the nitrate response in Arabidopsis roots. Mol. Plant 12:1545–60 [DOI] [PubMed] [Google Scholar]
  4. Alvarez JM, Schinke A-L, Brooks MD, Pasquino A, Leonelli L, et al. 2020. Transient genome-wide interactions of the master transcription factor NLP7 initiate a rapid nitrogen-response cascade. Nat. Commun 11:1157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Araya T, Miyamoto M, Wibowo J, Suzuki A, Kojima S, et al. 2014. CLE-CLAVATA1 peptide-receptor signaling module regulates the expansion of plant root systems in a nitrogen-dependent manner. PNAS 111:2029–34 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Baena-González E, Lunn JE. 2020. SnRK1 and trehalose 6-phosphate—two ancient pathways converge to regulate plant metabolism and growth. Curr. Opin. Plant Biol 55:52–59 [DOI] [PubMed] [Google Scholar]
  7. Baena-González E, Rolland F, Thevelein JM, Sheen J. 2007. A central integrator of transcription networks in plant stress and energy signalling. Nature 448:938–42 [DOI] [PubMed] [Google Scholar]
  8. Balzergue C, Dartevelle T, Godon C, Laugier E, Meisrimler C, et al. 2017. Low phosphate activates STOP1-ALMT1 to rapidly inhibit root cell elongation. Nat. Commun 8:15300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bari R, Pant BD, Stitt M, Scheible W-R. 2006. PHO2, microRNA399, and PHR1 define a phosphate-signaling pathway in plants. Plant Physiol. 141:988–99 [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Barrada A, Djendli M, Desnos T, Mercier R, Robaglia C, et al. 2019. A TOR-YAK1 signaling axis controls cell cycle, meristem activity and plant growth in Arabidopsis. Development 146:dev171298. [DOI] [PubMed] [Google Scholar]
  11. Bhosale R, Giri J, Pandey BK, Giehl RF, Hartmann A, et al. 2018. A mechanistic framework for auxin dependent Arabidopsis root hair elongation to low external phosphate. Nat. Commun 9:1409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bouguyon E, Brun F, Meynard D, Kubeš M, Pervent M, et al. 2015. Multiple mechanisms of nitrate sensing by Arabidopsis nitrate transceptor NRT1.1. Nat. Plants 1:15015. [DOI] [PubMed] [Google Scholar]
  13. Broeckx T, Hulsmans S, Rolland F. 2016. The plant energy sensor: evolutionary conservation and divergence of SnRK1 structure, regulation, and function. J. Exp. Bot 67:6215–52 [DOI] [PubMed] [Google Scholar]
  14. Brooks MD, Cirrone J, Pasquino AV, Alvarez JM, Swift J, et al. 2019. Network Walking charts transcriptional dynamics of nitrogen signaling by integrating validated and predicted genome-wide interactions. Nat. Commun 10:1569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bruggeman Q, Prunier F, Mazubert C, de Bont L, Garmier M, et al. 2015. Involvement of Arabidopsis hexokinase1 in cell death mediated by myo-inositol accumulation. Plant Cell 27:1801–14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Brunkard JO, Xu M, Scarpin MR, Chatterjee S, Shemyakina EA, et al. 2020. TOR dynamically regulates plant cell–cell transport. PNAS 117:5049–58 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bruns AN, Li S, Mohannath G, Bisaro DM. 2019. Phosphorylation of Arabidopsis eIF4E and eIFiso4E by SnRK1 inhibits translation. FEBS J. 286:3778–96 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bustos R, Castrillo G, Linhares F, Puga MI, Rubio V, et al. 2010. A central regulatory system largely controls transcriptional activation and repression responses to phosphate starvation in Arabidopsis. PLOS Genet. 6:e1001102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Caldana C, Li Y, Leisse A, Zhang Y, Bartholomaeus L, et al. 2013. Systemic analysis of inducible target of rapamycin mutants reveal a general metabolic switch controlling growth in Arabidopsis thaliana. Plant J. 73:897–909 [DOI] [PubMed] [Google Scholar]
  20. Canales J, Moyano TC, Villarroel E, Gutiérrez RA. 2014. Systems analysis of transcriptome data provides new hypotheses about Arabidopsis root response to nitrate treatments. Front. Plant Sci 5:22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Cao P, Kim S-J, Xing A, Schenck CA, Liu L, et al. 2019. Homeostasis of branched-chain amino acids is critical for the activity of TOR signaling in Arabidopsis. eLife 8:e50747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Castaings L, Camargo A, Pocholle D, Gaudon V, Texier Y, et al. 2009. The nodule inception-like protein 7 modulates nitrate sensing and metabolism in Arabidopsis. Plant J. 57:426–35 [DOI] [PubMed] [Google Scholar]
  23. Chen G-H, Liu M-J, Xiong Y, Sheen J, Wu S-H. 2018. TOR and RPS6 transmit light signals to enhance protein translation in deetiolating Arabidopsis seedlings. PNAS 115:12823–28 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chen HY, Lin SH, Cheng LH, Wu JJ, Lin YC, et al. 2021. Potential transceptor AtNRT1.13 modulates shoot architecture and flowering time in a nitrate-dependent manner. Plant Cell. doi: 10.1093/plcell/koab051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Chen L, Su Z-Z, Huang L, Xia F-N, Qi H, et al. 2017. The AMP-activated protein kinase KIN10 is involved in the regulation of autophagy in Arabidopsis. Front. Plant Sci 8:1201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Chen L-Q, Cheung LS, Feng L, Tanner W, Frommer WB. 2015. Transport of sugars. Annu. Rev. Biochem 84:865–94 [DOI] [PubMed] [Google Scholar]
  27. Chien P-S, Chiang C-P, Leong SJ, Chiou T-J. 2018. Sensing and signaling of phosphate starvation: from local to long distance. Plant Cell Physiol. 59:1714–22 [DOI] [PubMed] [Google Scholar]
  28. Chiou T-J, Aung K, Lin S-I, Wu C-C, Chiang S-F, Su C-I. 2006. Regulation of phosphate homeostasis by microRNA in Arabidopsis. Plant Cell 18:412–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Cho H-Y, Lu M-YJ, Shih M-C. 2019. The SnRK1-eIFiso4G1 signaling relay regulates the translation of specific mRNAs in Arabidopsis under submergence. New Phytol. 222:366–81 [DOI] [PubMed] [Google Scholar]
  30. Cho H-Y, Wen T-N, Wang Y-T, Shih M-C. 2016. Quantitative phosphoproteomics of protein kinase SnRK1 regulated protein phosphorylation in Arabidopsis under submergence. J. Exp. Bot 67:2745–60 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Cho J-I, Ryoo N, Eom J-S, Lee D-W, Kim H-B, et al. 2009. Role of the rice hexokinases OsHXK5 and OsHXK6 as glucose sensors. Plant Physiol. 149:745–59 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Cho Y-H, Yoo S-D. 2011. Signaling role of fructose mediated by FINS1/FBP in Arabidopsis thaliana. PLOS Genet. 7:e1001263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Cho Y-H, Yoo S-D, Sheen J. 2006. Regulatory functions of nuclear hexokinase1 complex in glucose signaling. Cell 127:579–89 [DOI] [PubMed] [Google Scholar]
  34. Claeys H, Vi SL, Xu X, Satoh-Nagasawa N, Eveland AL, et al. 2019. Control of meristem determinacy by trehalose 6-phosphate phosphatases is uncoupled from enzymatic activity. Nat. Plants 5:352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Couso I, Pérez-Pérez ME, Ford MM, Martínez-Force E, Hicks LM, et al. 2020. Phosphorus availability regulates TORC1 signaling via LST8 in Chlamydomonas. Plant Cell 32:69–80 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Deprost D, Yao L, Sormani R, Moreau M, Leterreux G, et al. 2007. The Arabidopsis TOR kinase links plant growth, yield, stress resistance and mRNA translation. EMBO Rep. 8:864–70 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Dobrenel T, Caldana C, Hanson J, Robaglia C, Vincentz M, et al. 2016. TOR signaling and nutrient sensing. Annu. Rev. Plant Biol 67:261–85 [DOI] [PubMed] [Google Scholar]
  38. Dong J, Ma G, Sui L, Wei M, Satheesh V, et al. 2019. Inositol pyrophosphate InsP8 acts as an intracellular phosphate signal in Arabidopsis. Mol. Plant 12:1463–73 [DOI] [PubMed] [Google Scholar]
  39. Dong J, Piñeros MA, Li X, Yang H, Liu Y, et al. 2017. An Arabidopsis ABC transporter mediates phosphate deficiency-induced remodeling of root architecture by modulating iron homeostasis in roots. Mol. Plant 10:244–59 [DOI] [PubMed] [Google Scholar]
  40. Dong Y, Silbermann M, Speiser A, Forieri I, Linster E, et al. 2017. Sulfur availability regulates plant growth via glucose-TOR signaling. Nat. Commun 8:1174. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  41. Eastmond PJ, Van Dijken AJ, Spielman M, Kerr A, Tissier AF, et al. 2002. Trehalose-6-phosphate synthase 1, which catalyses the first step in trehalose synthesis, is essential for Arabidopsis embryo maturation. Plant J. 29:225–35 [DOI] [PubMed] [Google Scholar]
  42. Emanuelle S, Hossain MI, Moller IE, Pedersen HL, Meene AM, et al. 2015. SnRK1 from Arabidopsis thaliana is an atypical AMPK. Plant J. 82:183–92 [DOI] [PubMed] [Google Scholar]
  43. Enganti R, Cho SK, Toperzer JD, Urquidi-Camacho RA, Cakir OS, et al. 2018. Phosphorylation of ribosomal protein RPS6 integrates light signals and circadian clock signals. Front. Plant Sci 8:2210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Feng J, Zhao S, Chen X, Wang W, Dong W, et al. 2015. Biochemical and structural study of Arabidopsis hexokinase 1. Acta Crystallogr. Sect. D 71:367–75 [DOI] [PubMed] [Google Scholar]
  45. Fichtner F, Olas JJ, Feil R, Watanabe M, Krause U, et al. 2020. Functional features of TREHALOSE-6-PHOSPHATE SYNTHASE1, an essential enzyme in Arabidopsis. Plant Cell 32:1949–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Figueroa CM, Lunn JE. 2016. A tale of two sugars: trehalose 6-phosphate and sucrose. Plant Physiol. 172:7–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Forzani C, Duarte GT, Van Leene J, Clément G, Huguet S, et al. 2019. Mutations of the AtYAK1 kinase suppress TOR deficiency in Arabidopsis. Cell Rep. 27:3696–708.e5 [DOI] [PubMed] [Google Scholar]
  48. Frank A, Matiolli CC, Viana AJ, Hearn TJ, Kusakina J, et al. 2018. Circadian entrainment in Arabidopsis by the sugar-responsive transcription factor bZIP63. Curr. Biol 28:2597–606.e6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Fu L, Liu Y, Qin G, Wu P, Zi H, et al. 2021. The TOR–EIN2 axis mediates nuclear signalling to modulate plant growth. Nature. doi: 10.1038/s41586-021-03310-y [DOI] [PubMed] [Google Scholar]
  50. Gaudinier A, Rodriguez-Medina J, Zhang L, Olson A, Liseron-Monfils C, et al. 2018. Transcriptional regulation of nitrogen-associated metabolism and growth. Nature 563:259–64 [DOI] [PubMed] [Google Scholar]
  51. Godon C, Mercier C, Wang X, David P, Richaud P, et al. 2019. Under phosphate starvation conditions, Fe and Al trigger accumulation of the transcription factor STOP1 in the nucleus of Arabidopsis root cells. Plant J. 99:937–49 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Griffiths CA, Sagar R, Geng Y, Primavesi LF, Patel MK, et al. 2016. Chemical intervention in plant sugar signalling increases yield and resilience. Nature 540:574–78 [DOI] [PubMed] [Google Scholar]
  53. Guan P, Ripoll J-J, Wang R, Vuong L, Bailey-Steinitz LJ, et al. 2017. Interacting TCP and NLP transcription factors control plant responses to nitrate availability. PNAS 114:2419–24 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Gutiérrez-Alanís D, Yong-Villalobos L, Jiménez-Sandoval P, Alatorre-Cobos F, Oropeza-Aburto A, et al. 2017. Phosphate starvation-dependent iron mobilization induces CLE14 expression to trigger root meristem differentiation through CLV2/PEPR2 signaling. Dev. Cell 41:555–70.e3 [DOI] [PubMed] [Google Scholar]
  55. Ham B-K, Chen J, Yan Y, Lucas WJ. 2018. Insights into plant phosphate sensing and signaling. Curr. Opin. Biotechnol 49:1–9 [DOI] [PubMed] [Google Scholar]
  56. Han C, Liu Y, Shi W, Qiao Y, Wang L, et al. 2020. KIN10 promotes stomatal development through stabilization of the SPEECHLESS transcription factor. Nat. Commun 11:4214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Ho C-H, Lin S-H, Hu H-C, Tsay Y-F. 2009. CHL1 functions as a nitrate sensor in plants. Cell 138:1184–94 [DOI] [PubMed] [Google Scholar]
  58. Hu B, Jiang Z, Wang W, Qiu Y, Zhang Z, et al. 2019. Nitrate–NRT1.1B–SPX4 cascade integrates nitrogen and phosphorus signalling networks in plants. Nat. Plants 5:401. [DOI] [PubMed] [Google Scholar]
  59. Hu D-G, Sun C-H, Zhang Q-Y, An J-P, You C-X, Hao Y-J. 2016. Glucose sensor MdHXK1 phosphorylates and stabilizes MdbHLH3 to promote anthocyanin biosynthesis in apple. PLOS Genet. 12:e1006273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Huang J-P, Tunc-Ozdemir M, Chang Y, Jones AM. 2015. Cooperative control between AtRGS1 and AtHXK1 in a WD40-repeat protein pathway in Arabidopsis thaliana. Front. Plant Sci 6:851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Huang T-K, Han C-L, Lin S-I, Chen Y-J, Tsai Y-C, et al. 2013. Identification of downstream components of ubiquitin-conjugating enzyme PHOSPHATE2 by quantitative membrane proteomics in Arabidopsis roots. Plant Cell 25:4044–60 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Hwang G, Kim S, Cho J-Y, Paik I, Kim J-I, Oh E. 2019. Trehalose-6-phosphate signaling regulates thermoresponsive hypocotyl growth in Arabidopsis thaliana. EMBO Rep. 20:e47828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Im JH, Cho YH, Kim GD, Kang GH, Hong JW, Yoo SD. 2014. Inverse modulation of the energy sensor Snf1-related protein kinase 1 on hypoxia adaptation and salt stress tolerance in Arabidopsis thaliana. Plant Cell Environ. 37:2303–12 [DOI] [PubMed] [Google Scholar]
  64. Jamsheer KM, Sharma M, Singh D, Mannully CT, Jindal S, et al. 2018. FCS-like zinc finger 6 and 10 repress SnRK 1 signalling in Arabidopsis. Plant J. 94:232–45 [DOI] [PubMed] [Google Scholar]
  65. Jang J-C, León P, Zhou L, Sheen J. 1997. Hexokinase as a sugar sensor in higher plants. Plant Cell 9:5–19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Jeong E-Y, Seo PJ, Woo JC, Park C-M. 2015. AKIN10 delays flowering by inactivating IDD8 transcription factor through protein phosphorylation in Arabidopsis. BMC Plant Biol. 15:110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Jossier M, Bouly JP, Meimoun P, Arjmand A, Lessard P, et al. 2009. SnRK1 (SNF1-related kinase 1) has a central role in sugar and ABA signalling in Arabidopsis thaliana. Plant J. 59:316–28 [DOI] [PubMed] [Google Scholar]
  68. Kant S, Peng M, Rothstein SJ. 2011. Genetic regulation by NLA and microRNA827 for maintaining nitrate-dependent phosphate homeostasis in Arabidopsis. PLOS Genet. 7:e1002021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Karve A, Rauh BL, Xia X, Kandasamy M, Meagher RB, et al. 2008. Expression and evolutionary features of the hexokinase gene family in Arabidopsis. Planta 228:411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Kelly G, David-Schwartz R, Sade N, Moshelion M, Levi A, et al. 2012. The pitfalls of transgenic selection and new roles of AtHXK1: A high level of AtHXK1 expression uncouples hexokinase1-dependent sugar signaling from exogenous sugar. Plant Physiol. 159:47–51 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Kiba T, Inaba J, Kudo T, Ueda N, Konishi M, et al. 2018. Repression of nitrogen starvation responses by members of the Arabidopsis GARP-type transcription factor NIGT1/HRS1 subfamily. Plant Cell 30:925–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Kim G-D, Cho Y-H, Yoo S-D. 2017. Regulatory functions of cellular energy sensor SNF1-Related Kinase1 for leaf senescence delay through ETHYLENE-INSENSITIVE3 repression. Sci. Rep 7:3193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Konishi M, Yanagisawa S. 2013. Arabidopsis NIN-like transcription factors have a central role in nitrate signalling. Nat. Commun 4:1617. [DOI] [PubMed] [Google Scholar]
  74. Krapp A, Berthomé R, Orsel M, Mercey-Boutet S, Yu A, et al. 2011. Arabidopsis roots and shoots show distinct temporal adaptation patterns toward nitrogen starvation. Plant Physiol. 157:1255–82 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Lee D-H, Park SJ, Ahn CS, Pai H-S. 2017. MRF family genes are involved in translation control, especially under energy-deficient conditions, and their expression and functions are modulated by the TOR signaling pathway. Plant Cell 29:2895–920 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Lei M, Liu Y, Zhang B, Zhao Y, Wang X, et al. 2011. Genetic and genomic evidence that sucrose is a global regulator of plant responses to phosphate starvation in Arabidopsis. Plant Physiol. 156:1116–30 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Léran S, Edel KH, Pervent M, Hashimoto K, Corratgé-Faillie C, et al. 2015. Nitrate sensing and uptake in Arabidopsis are enhanced by ABI2, a phosphatase inactivated by the stress hormone abscisic acid. Sci. Signal 8:ra43. [DOI] [PubMed] [Google Scholar]
  78. Li L, Sheen J. 2016. Dynamic and diverse sugar signaling. Curr. Opin. Plant Biol 33:116–25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Li P, Wind JJ, Shi X, Zhang H, Hanson J, et al. 2011. Fructose sensitivity is suppressed in Arabidopsis by the transcription factor ANAC089 lacking the membrane-bound domain. PNAS 108:3436–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Li S, Tian Y, Wu K, Ye Y, Yu J, et al. 2018. Modulating plant growth-metabolism coordination for sustainable agriculture. Nature. 560:595–600 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Li X, Cai W, Liu Y, Li H, Fu L, et al. 2017. Differential TOR activation and cell proliferation in Arabidopsis root and shoot apexes. PNAS 114:2765–70 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Lin S-I, Chiang S-F, Lin W-Y, Chen J-W, Tseng C-Y, et al. 2008. Regulatory network of microRNA399 and PHO2 by systemic signaling. Plant Physiol. 147:732–46 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Lin W-Y, Huang T-K, Chiou T-J. 2013. NITROGEN LIMITATION ADAPTATION, a target of microRNA827, mediates degradation of plasma membrane–localized phosphate transporters to maintain phosphate homeostasis in Arabidopsis. Plant Cell 25:4061–74 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Liu K-H, Diener A, Lin Z, Liu C, Sheen J. 2020. Primary nitrate responses mediated by calcium signalling and diverse protein phosphorylation. J. Exp. Bot 71:4428–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Liu K-H, Niu Y, Konishi M, Wu Y, Du H, et al. 2017. Discovery of nitrate–CPK–NLP signalling in central nutrient–growth networks. Nature 545:311–16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Liu T-Y, Huang T-K, Tseng C-Y, Lai Y-S, Lin S-I, et al. 2012. PHO2-dependent degradation of PHO1 modulates phosphate homeostasis in Arabidopsis. Plant Cell 24:2168–83 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Liu Y, Duan X, Zhao X, Ding W, Wang Y, et al. 2021. Diverse Nitrogen Signals Activate Convergent ROP2-TOR Signaling in Arabidopsis. Development Cell. In press [DOI] [PubMed] [Google Scholar]
  88. Maeda Y, Konishi M, Kiba T, Sakuraba Y, Sawaki N, et al. 2018. A NIGT1-centred transcriptional cascade regulates nitrate signalling and incorporates phosphorus starvation signals in Arabidopsis. Nat. Commun 9:1376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Mair A, Pedrotti L, Wurzinger B, Anrather D, Simeunovic A, et al. 2015. SnRK1-triggered switch of bZIP63 dimerization mediates the low-energy response in plants. eLife 4:e05828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Marchive C, Roudier F, Castaings L, Bréhaut V, Blondet E, et al. 2013. Nuclear retention of the transcription factor NLP7 orchestrates the early response to nitrate in plants. Nat. Commun 4:1713. [DOI] [PubMed] [Google Scholar]
  91. Medici A, Marshall-Colon A, Ronzier E, Szponarski W, Wang R, et al. 2015. AtNIGT1/HRS1 integrates nitrate and phosphate signals at the Arabidopsis root tip. Nat. Commun 6:6274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Medici A, Szponarski W, Dangeville P, Safi A, Dissanayake IM, et al. 2019. Identification of molecular integrators shows that nitrogen actively controls the phosphate starvation response in plants. Plant Cell 31:1171–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Menand B, Desnos T, Nussaume L, Berger F, Bouchez D, et al. 2002. Expression and disruption of the Arabidopsis TOR (target of rapamycin) gene. PNAS 99:6422–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Mohammed B, Bilooei SF, Dóczi R, Grove E, Railo S, et al. 2018. Converging light, energy and hormonal signaling control meristem activity, leaf initiation, and growth. Plant Physiol. 176:1365–81 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Moore B, Zhou L, Rolland F, Hall Q, Cheng W-H, et al. 2003. Role of the Arabidopsis glucose sensor HXK1 in nutrient, light, and hormonal signaling. Science 300:332–36 [DOI] [PubMed] [Google Scholar]
  96. Mora-Macías J, Ojeda-Rivera JO, Gutiérrez-Alanís D, Yong-Villalobos L, Oropeza-Aburto A, et al. 2017. Malate-dependent Fe accumulation is a critical checkpoint in the root developmental response to low phosphate. PNAS 114:E3563–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Moreau M, Azzopardi M, Clément G, Dobrenel T, Marchive C, et al. 2012. Mutations in the Arabidopsis homolog of LST8/GβL, a partner of the target of rapamycin kinase, impair plant growth, flowering, and metabolic adaptation to long days. Plant Cell 24:463–81 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Müller J, Toev T, Heisters M, Teller J, Moore KL, et al. 2015. Iron-dependent callose deposition adjusts root meristem maintenance to phosphate availability. Dev. Cell 33:216–30 [DOI] [PubMed] [Google Scholar]
  99. Müller R, Morant M, Jarmer H, Nilsson L, Nielsen TH. 2007. Genome-wide analysis of the Arabidopsis leaf transcriptome reveals interaction of phosphate and sugar metabolism. Plant Physiol. 143:156–71 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Naulin PA, Armijo GI, Vega AS, Tamayo KP, Gras DE, et al. 2020. Nitrate induction of primary root growth requires cytokinin signaling in Arabidopsis thaliana. Plant Cell Physiol. 61:342–52 [DOI] [PubMed] [Google Scholar]
  101. Nukarinen E, Nägele T, Pedrotti L, Wurzinger B, Mair A, et al. 2016. Quantitative phosphoproteomics reveals the role of the AMPK plant ortholog SnRK1 as a metabolic master regulator under energy deprivation. Sci. Rep 6:31697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Nunes C, Primavesi LF, Patel MK, Martinez-Barajas E, Powers SJ, et al. 2013. Inhibition of SnRK1 by metabolites: tissue-dependent effects and cooperative inhibition by glucose 1-phosphate in combination with trehalose 6-phosphate. Plant Physiol. Biochem 63:89–98 [DOI] [PubMed] [Google Scholar]
  103. O’Leary BM, Oh GGK, Lee CP, Millar AH. 2020. Metabolite regulatory interactions control plant respiratory metabolism via Target of Rapamycin (TOR) kinase activation. Plant Cell 32:666–82 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Ohkubo Y, Tanaka M, Tabata R, Ogawa-Ohnishi M, Matsubayashi Y. 2017. Shoot-to-root mobile polypeptides involved in systemic regulation of nitrogen acquisition. Nat. Plants 3:17029. [DOI] [PubMed] [Google Scholar]
  105. Pedrotti L, Weiste C, Nägele T, Wolf E, Lorenzin F, et al. 2018. Snf1-RELATED KINASE1-controlled C/S1-bZIP signaling activates alternative mitochondrial metabolic pathways to ensure plant survival in extended darkness. Plant Cell 30:495–509 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Pérez-Torres C-A, López-Bucio J, Cruz-Ramírez A, Ibarra-Laclette E, Dharmasiri S, et al. 2008. Phosphate availability alters lateral root development in Arabidopsis by modulating auxin sensitivity via a mechanism involving the TIR1 auxin receptor. Plant Cell 20:3258–72 [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Pfeiffer A, Janocha D, Dong Y, Medzihradszky A, Schöne S, et al. 2016. Integration of light and metabolic signals for stem cell activation at the shoot apical meristem. eLife 5:e17023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Puga MI, Mateos I, Charukesi R, Wang Z, Franco-Zorrilla JM, et al. 2014. SPX1 is a phosphate-dependent inhibitor of Phosphate Starvation Response 1 in Arabidopsis. PNAS 111:14947–52 [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Puga MI, Rojas-Triana M, de Lorenzo L, Leyva A, Rubio V, Paz-Ares J. 2017. Novel signals in the regulation of Pi starvation responses in plants: facts and promises. Curr. Opin. Plant Biol 39:40–49 [DOI] [PubMed] [Google Scholar]
  110. Raghothama K 2000. Phosphate transport and signaling. Curr. Opin. Plant Biol 3:182–87 [PubMed] [Google Scholar]
  111. Rahmani F, Hummel M, Schuurmans J, Wiese-Klinkenberg A, Smeekens S, Hanson J. 2009. Sucrose control of translation mediated by an upstream open reading frame-encoded peptide. Plant Physiol. 150:1356–67 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Ramon M, Dang TVT, Broeckx T, Hulsmans S, Crepin N, et al. 2019. Default activation and nuclear translocation of the plant cellular energy sensor SnRK1 regulate metabolic stress responses and development. Plant Cell 31:1614–32 [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Ramon M, Ruelens P, Li Y, Sheen J, Geuten K, Rolland F. 2013. The hybrid Four-CBS-Domain KINβγ subunit functions as the canonical γ subunit of the plant energy sensor SnRK1. Plant J. 75:11–25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Ren M, Qiu S, Venglat P, Xiang D, Feng L, et al. 2011. Target of rapamycin regulates development and ribosomal RNA expression through kinase domain in Arabidopsis. Plant Physiol. 155:1367–82 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Ren M, Venglat P, Qiu S, Feng L, Cao Y, et al. 2012. Target of rapamycin signaling regulates metabolism, growth, and life span in Arabidopsis. Plant Cell 24:4850–74 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Riveras E, Alvarez JM, Vidal EA, Oses C, Vega A, Gutiérrez RA. 2015. The calcium ion is a second messenger in the nitrate signaling pathway of Arabidopsis. Plant Physiol. 169:1397–404 [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Rodrigues A, Adamo M, Crozet P, Margalha L, Confraria A, et al. 2013. ABI1 and PP2CA phosphatases are negative regulators of Snf1-related protein kinase1 signaling in Arabidopsis. Plant Cell 25:3871–84 [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Rolland F, Baena-Gonzalez E, Sheen J. 2006. Sugar sensing and signaling in plants: conserved and novel mechanisms. Annu. Rev. Plant Biol 57:675–709 [DOI] [PubMed] [Google Scholar]
  119. Roth MS, Westcott DJ, Iwai M, Niyogi KK. 2019. Hexokinase is necessary for glucose-mediated photosynthesis repression and lipid accumulation in a green alga. Commun. Biol 2:347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Rubio V, Linhares F, Solano R, Martín AC, Iglesias J, et al. 2001. A conserved MYB transcription factor involved in phosphate starvation signaling both in vascular plants and in unicellular algae. Genes Dev. 15:2122–33 [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Ruiz-Gayosso A, Rodríguez-Sotres R, Martínez-Barajas E, Coello P. 2018. A role for the carbohydrate-binding module (CBM) in regulatory SnRK1 subunits: the effect of maltose on SnRK1 activity. Plant J. 96:163–75 [DOI] [PubMed] [Google Scholar]
  122. Ryabova LA, Robaglia C, Meyer C. 2019. Target of Rapamycin kinase: central regulatory hub for plant growth and metabolism. J. Exp. Bot 70:2211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Sakakibara H 2020. Cytokinin biosynthesis and transport for systemic nitrogen signaling. Plant J. [DOI] [PubMed] [Google Scholar]
  124. Sakakibara H, Kobayashi K, Deji A, Sugiyama T. 1997. Partial characterization of the signaling pathway for the nitrate-dependent expression of genes for nitrogen-assimilatory enzymes using detached maize leaves. Plant Cell Physiol. 38:837–43 [Google Scholar]
  125. Scarpin MR, Leiboff S, Brunkard JO. 2020. Parallel global profiling of plant TOR dynamics reveals a conserved role for LARP1 in translation. eLife 9:e58795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Scheible W-R, Gonzalez-Fontes A, Lauerer M, Muller-Rober B, Caboche M, Stitt M. 1997. Nitrate acts as a signal to induce organic acid metabolism and repress starch metabolism in tobacco. Plant Cell 9:783–98 [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Scheible W-R, Morcuende R, Czechowski T, Fritz C, Osuna D, et al. 2004. Genome-wide reprogramming of primary and secondary metabolism, protein synthesis, cellular growth processes, and the regulatory infrastructure of Arabidopsis in response to nitrogen. Plant Physiol. 136:2483–99 [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Schepetilnikov M, Dimitrova M, Mancera-Martínez E, Geldreich A, Keller M, Ryabova LA. 2013. TOR and S6K1 promote translation reinitiation of uORF-containing mRNAs via phosphorylation of eIF3h. EMBO J. 32:1087–102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Schepetilnikov M, Makarian J, Srour O, Geldreich A, Yang Z, et al. 2017. GTPase ROP2 binds and promotes activation of target of rapamycin, TOR, in response to auxin. EMBO J. 36:886–903 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Schluepmann H, Pellny T, van Dijken A, Smeekens S, Paul M. 2003. Trehalose 6-phosphate is indispensable for carbohydrate utilization and growth in Arabidopsis thaliana. PNAS 100:6849–54 [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Sheen J 2014. Master regulators in plant glucose signaling networks. J. Plant Biol 57:67–79 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Shen W, Reyes MI, Hanley-Bowdoin L. 2009. Arabidopsis protein kinases GRIK1 and GRIK2 specifically activate SnRK1 by phosphorylating its activation loop. Plant Physiol. 150:996–1005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Shi L, Wu Y, Sheen J. 2018. TOR signaling in plants: conservation and innovation. Development 145:dev160887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Soto-Burgos J, Zhuang X, Jiang L, Bassham DC. 2018. Dynamics of autophagosome formation. Plant Physiol. 176:219–29 [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Stein O, Granot D. 2019. An overview of sucrose synthases in plants. Front. Plant Sci 10:95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Sun L, Song L, Zhang Y, Zheng Z, Liu D. 2016. Arabidopsis PHL2 and PHR1 act redundantly as the key components of the central regulatory system controlling transcriptional responses to phosphate starvation. Plant Physiol. 170:499–514 [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Swift J, Alvarez JM, Araus V, Gutiérrez RA, Coruzzi GM. 2020. Nutrient dose-responsive transcriptome changes driven by Michaelis–Menten kinetics underlie plant growth rates. PNAS 117:12531–40 [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Tabata R, Sumida K, Yoshii T, Ohyama K, Shinohara H, Matsubayashi Y. 2014. Perception of root-derived peptides by shoot LRR-RKs mediates systemic N-demand signaling. Science 346:343–46 [DOI] [PubMed] [Google Scholar]
  139. Thibaud MC, Arrighi JF, Bayle V, Chiarenza S, Creff A, et al. 2010. Dissection of local and systemic transcriptional responses to phosphate starvation in Arabidopsis. Plant J. 64:775–89 [DOI] [PubMed] [Google Scholar]
  140. Thieme CJ, Rojas-Triana M, Stecyk E, Schudoma C, Zhang W, et al. 2015. Endogenous Arabidopsis messenger RNAs transported to distant tissues. Nat. Plants 1:15025. [DOI] [PubMed] [Google Scholar]
  141. Tsai AYL, Gazzarrini S. 2012. AKIN10 and FUSCA3 interact to control lateral organ development and phase transitions in Arabidopsis. Plant J. 69:809–21 [DOI] [PubMed] [Google Scholar]
  142. Ueda Y, Kiba T, Yanagisawa S. 2020. Nitrate-inducible NIGT1 proteins modulate phosphate uptake and starvation signalling via transcriptional regulation of SPX genes. Plant J. 102:448–66 [DOI] [PubMed] [Google Scholar]
  143. Ulfstedt M, Hu G-Z, Eklund DM, Ronne H. 2018. The ability of a charophyte alga hexokinase to restore glucose signaling and glucose repression of gene expression in a glucose-insensitive Arabidopsis hexokinase mutant depends on its catalytic activity. Front. Plant Sci 9:1887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Urano D, Phan N, Jones JC, Yang J, Huang J, et al. 2012. Endocytosis of the seven-transmembrane RGS1 protein activates G-protein-coupled signalling in Arabidopsis. Nat. Cell Biol 14:1079–88 [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Van Leene J, Han C, Gadeyne A, Eeckhout D, Matthijs C, et al. 2019. Capturing the phosphorylation and protein interaction landscape of the plant TOR kinase. Nat. Plants 5:316–27 [DOI] [PubMed] [Google Scholar]
  146. Varala K, Marshall-Colón A, Cirrone J, Brooks MD, Pasquino AV, et al. 2018. Temporal transcriptional logic of dynamic regulatory networks underlying nitrogen signaling and use in plants. PNAS 115:6494–99 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Vidal EA, Alvarez JM, Araus V, Riveras E, Brooks M, et al. 2020. Nitrate in 2020: thirty years from transport to signaling networks. Plant Cell 32:2094–119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Vidal EA, Moyano TC, Riveras E, Contreras-López O, Gutiérrez RA. 2013. Systems approaches map regulatory networks downstream of the auxin receptor AFB3 in the nitrate response of Arabidopsis thaliana roots. PNAS 110:12840–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Wahl V, Ponnu J, Schlereth A, Arrivault S, Langenecker T, et al. 2013. Regulation of flowering by trehalose-6-phosphate signaling in Arabidopsis thaliana. Science 339:704–7 [DOI] [PubMed] [Google Scholar]
  150. Walker L, Boddington C, Jenkins D, Wang Y, Grønlund JT, et al. 2017. Changes in gene expression in space and time orchestrate environmentally mediated shaping of root architecture. Plant Cell 29:2393–412 [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Wang P, Zhao Y, Li Z, Hsu C-C, Liu X, et al. 2018. Reciprocal regulation of the TOR kinase and ABA receptor balances plant growth and stress response. Mol. Cell 69:100–12.e6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Wang R, Tischner R, Gutiérrez RA, Hoffman M, Xing X, et al. 2004. Genomic analysis of the nitrate response using a nitrate reductase-null mutant of Arabidopsis. Plant Physiol. 136:2512–22 [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Wang X, Feng C, Tian L, Hou C, Tian W, et al. 2021. A transceptor-channel complex couples nitrate sensing to calcium signaling in Arabidopsis. Mol. Plant doi: 10.1016/j.molp.2021.02.005. [DOI] [PubMed] [Google Scholar]
  154. Wang Y-Y, Cheng Y-H, Chen K-E, Tsay Y-F. 2018. Nitrate transport, signaling, and use efficiency. Annu. Rev. Plant Biol 69:85–122 [DOI] [PubMed] [Google Scholar]
  155. Wang Z, Ruan W, Shi J, Zhang L, Xiang D, et al. 2014. Rice SPX1 and SPX2 inhibit phosphate starvation responses through interacting with PHR2 in a phosphate-dependent manner. PNAS 111:14953–58 [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Wege S, Khan GA, Jung J-Y, Vogiatzaki E, Pradervand S, et al. 2016. The EXS domain of PHO1 participates in the response of shoots to phosphate deficiency via a root-to-shoot signal. Plant Physiol. 170:385–400 [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Wendrich JR, Yang B, Vandamme N, Verstaen K, Smet W, et al. 2020. Vascular transcription factors guide plant epidermal responses to limiting phosphate conditions. Science 370:eaay4970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Wild R, Gerasimaite R, Jung J-Y, Truffault V, Pavlovic I, et al. 2016. Control of eukaryotic phosphate homeostasis by inositol polyphosphate sensor domains. Science 352:986–90 [DOI] [PubMed] [Google Scholar]
  159. Wu K, Wang S, Song W, Zhang J, Wang Y, et al. 2020. Enhanced sustainable green revolution yield via nitrogen-responsive chromatin modulation in rice. Science 367: eaaz2046. [DOI] [PubMed] [Google Scholar]
  160. Wu Y, Shi L, Li L, Fu L, Liu Y, et al. 2019. Integration of nutrient, energy, light, and hormone signalling via TOR in plants. J. Exp. Bot 70:2227–38 [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Xiong F, Zhang R, Meng Z, Deng K, Que Y, et al. 2017. Brassinosteriod Insensitive 2 (BIN2) acts as a downstream effector of the Target of Rapamycin (TOR) signaling pathway to regulate photoautotrophic growth in Arabidopsis. New Phytol. 213:233–49 [DOI] [PubMed] [Google Scholar]
  162. Xiong Y, McCormack M, Li L, Hall Q, Xiang C, Sheen J. 2013. Glucose-TOR signalling reprograms the transcriptome and activates meristems. Nature 496:181–86 [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Xiong Y, Sheen J. 2012. Rapamycin and glucose-target of rapamycin (TOR) protein signaling in plants. J. Biol. Chem 287:2836–42 [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Xu S-L, Chalkley RJ, Maynard JC, Wang W, Ni W, et al. 2017. Proteomic analysis reveals O-GlcNAc modification on proteins with key regulatory functions in Arabidopsis. PNAS 114:E1536–43 [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Yan D, Easwaran V, Chau V, Okamoto M, Ierullo M, et al. 2016. NIN-like protein 8 is a master regulator of nitrate-promoted seed germination in Arabidopsis. Nat. Commun 7:13179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Yanagisawa S, Yoo S-D, Sheen J. 2003. Differential regulation of EIN3 stability by glucose and ethylene signalling in plants. Nature 425:521–25 [DOI] [PubMed] [Google Scholar]
  167. Zentella R, Hu J, Hsieh W-P, Matsumoto PA, Dawdy A, et al. 2016. O-GlcNAcylation of master growth repressor DELLA by SECRET AGENT modulates multiple signaling pathways in Arabidopsis. Genes Dev. 30:164–76 [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Zentella R, Sui N, Barnhill B, Hsieh W-P, Hu J, et al. 2017. The Arabidopsis O-fucosyltransferase SPINDLY activates nuclear growth repressor DELLA. Nat. Chem. Biol 13:479–85 [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Zhai Z, Keereetaweep J, Liu H, Feil R, Lunn JE, Shanklin J. 2018. Trehalose 6-phosphate positively regulates fatty acid synthesis by stabilizing WRINKLED1. Plant Cell 30:2616–27 [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Zhai Z, Liu H, Shanklin J. 2017. Phosphorylation of WRINKLED1 by KIN10 results in its proteasomal degradation, providing a link between energy homeostasis and lipid biosynthesis. Plant Cell 29:871–89 [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Zhang H, Forde BG. 1998. An Arabidopsis MADS box gene that controls nutrient-induced changes in root architecture. Science 279:407–9 [DOI] [PubMed] [Google Scholar]
  172. Zhang N, Meng Y, Li X, Zhou Y, Ma L, et al. 2019. Metabolite-mediated TOR signaling regulates the circadian clock in Arabidopsis. PNAS 116:25395–97 [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Zhang X, Cui Y, Yu M, Su B, Gong W, et al. 2019. Phosphorylation-mediated dynamics of nitrate transceptor NRT1.1 regulate auxin flux and nitrate signaling in lateral root growth. Plant Physiol. 181:480–98 [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Zhang Y, Primavesi LF, Jhurreea D, Andralojc PJ, Mitchell RA, et al. 2009. Inhibition of SNF1-related protein kinase1 activity and regulation of metabolic pathways by trehalose-6-phosphate. Plant Physiol. 149:1860–71 [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Zhang Z, Zheng Y, Ham B-K, Chen J, Yoshida A, et al. 2016a. Vascular-mediated signalling involved in early phosphate stress response in plants. Nat. Plants 2:16033. [DOI] [PubMed] [Google Scholar]
  176. Zhang Z, Zhu J-Y, Roh J, Marchive C, Kim S-K, et al. 2016b. TOR signaling promotes accumulation of BZR1 to balance growth with carbon availability in Arabidopsis. Curr. Biol 26:1854–60 [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Zhu J, Lau K, Puschmann R, Harmel RK, Zhang Y, et al. 2019. Two bifunctional inositol pyrophosphate kinases/phosphatases control plant phosphate homeostasis. eLife 8:e43582. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES