Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Departamento de Biotecnología-Biología Vegetal, Escuela Técnica Superior de Ingeniería Agronómica, Alimentaría y de Biosistemas, UPM, Madrid, Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Centro de Biotecnología y Genómica de Plantas, Universidad Politécnica de Madrid (UPM) - Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA/CSIC), Campus de Montegancedo UPM, Pozuelo de Alarcón (Madrid), Spain
Departamento de Biotecnología-Biología Vegetal, Escuela Técnica Superior de Ingeniería Agronómica, Alimentaría y de Biosistemas, UPM, Madrid, Spain
The two most abundant macronutrients in plant cells are carbon (C) and nitrogen (N). Coordination of their cellular metabolism is a fundamental factor in guaranteeing the optimal growth and development of plants. N availability and assimilation profoundly affect plant gene expression and modulate root and stem architecture, thus affecting whole plant growth and crop yield. N status also affects C fixation, as it is an important component of the photosynthetic machinery in leaves. Reciprocally, increasing C supply promotes N uptake and assimilation. There is extensive knowledge of the different mechanisms that plants use for sensing and signalling their nutritional status to regulate the assimilation, metabolism and transport of C and N. However, the crosstalk between C and N pathways has received much less attention. Plant growth and development are greatly affected by suboptimal C/N balance, which can arise from nutrient deficiencies or/and environmental cues. Mechanisms that integrate and respond to changes in this specific nutritional balance have started to arise. This review will examine the specific responses to C/N imbalance in plants by focusing on the main inorganic and organic metabolites involved, how they are sensed and transported, and the interconnection between the early signalling components and hormonal networks that underlies plants’ adaptive responses.
1 INTRODUCTION
Plant growth and yield rely on nutrient-signalling pathways efficiently coordinated with development and environmental conditions. The two most abundant macronutrients in plant cells are carbon (C) and nitrogen (N), and responses to either of these nutrients have been studied extensively. However, the balance between C- and N-compounds has received less attention despite the fact that the C/N ratio, in addition to the C status or the N availability alone, significantly affects plant growth and development (Martin et al., 2002). The control of C and N interaction is assisted by a complex network of signals emanating from nitrate (NO3-), ammonium (NH4+), and N-containing metabolites as well as from C metabolism (Coruzzi & Zhou, 2001; Miller et al., 2008; Stitt & Krapp, 1999). Subsequently, sensing and transport of these signals are coordinated with several gene regulatory networks to maintain global C/N homeostasis in the plant.
C is assimilated through photosynthesis from CO2 and the resulting sugars, in particular sucrose and glucose, are converted through glycolysis and the tricarboxylic acid cycle (TCA) to provide energy and C-skeletons, such as 2-oxoglutarate (2-OG; also known as α-ketoglutarate), for amino acid biosynthesis. N nutrients include inorganic (NO3- and NH4+) and organic compounds (i.e., amino acids), the latter being synthesized by incorporating NH4+ into the C-skeletons. NO3- is first reduced by the nitrate reductase to nitrite (NO2-) and then by nitrite reductase to NH4+. Next, NH4+ is incorporated into glutamate (Glu) to generate glutamine (Gln). Gln is converted back to Glu using 2-oxoglutarate (2-OG), an intermediate of the tricarboxylic acid cycle (TCA). Glu serves as a N-donor for the biosynthesis of amino acids and other N-containing compounds (Figure 1).
Schematic representation of sensors, transporters and major signalling players that underlie C/N balance regulation. Central panel depicts the coordination between C fixation and N assimilation. Left panel illustrates CO2 sensing, fixation and control of stomatal opening, as well as the downstream transport of sucrose from leaves to sink organs, influenced by N levels. Right panel shows NO3 uptake, transport and cytoplasmic and nuclear signalling in the root cells in response to N availability. Finally, the mechanism that coordinates glucose (Glc) signalling and N utilization in rice, involving HKX7-ARE4 TF complex and translocation to the nucleus, is also depicted.
N assimilation is highly energy demanding and, in leaves, the photosynthetic metabolism is responsible for providing reducing power and ATP, whereas it is the oxidative pentose phosphate pathway that provides NADPH for biosynthesis in roots. Moreover, large amounts of N are invested in the photosynthetic machinery, as it requires enzymes built of amino acids. In addition, in land plants, the sites where C and N are absorbed are spatially separated. NO3- and NH4+ are taken up by the root system and assimilated both in roots and leaves, while C assimilation occurs in the leaf. Therefore, keeping appropriate C/N ratios throughout the plant requires a spatial exchange of energy and C-skeletons with N-compounds supported by dedicated sensing, signalling and transport mechanisms. This intricate relationship explains the interrelated effects produced by changes in C/N ratios. For example, N-deficiency negatively affects the photosynthetic output, an effect reverted by increasing N levels. Concomitantly, illuminated and photosynthetically active leaves support higher levels of N assimilation compared to dark conditions (Matt et al., 2001; Nunes-Nesi et al., 2010; Reed et al., 1983; Zheng et al., 2009).
Here, we focus on the different aspects of C and N interactions and their connections to reveal the complexity of the mechanisms underlying the regulation of their balance in plants. Thereby, we have reviewed major findings on the role of key C- and N-compounds, their sensors and transporters, and their early signalling events to highlight their relationships and impact on the regulation of the C/N balance in plants. To facilitate reading, we have mostly used gene acronyms throughout the text and a list of complete gene names and their abbreviations have been included separately (Table S1).
2 SUGAR AND NITROGEN SIGNALS
The control of C and N interaction involves signals emanating from C metabolism and inorganic and organic N-compounds. Although a certain number of molecules signalling C and N are known, we will focus in this section on the direct and essential signalling molecules involved in the interaction of both macronutrients.
Concerning C, atmospheric CO2 is fixed in mature leaves to generate sugars (Smith & Stitt, 2007). Plants are able to sense changes in the CO2 concentration in the intercellular air spaces of leaves throughout the day and respond by opening and closing their stomata (Zhang et al., 2018). Many studies have investigated the effect of elevated CO2 on plant nutrient uptake, especially on N. The results of these studies are highly variable, but most of them show that high CO2 concentration significantly reduces the efficiency of N acquisition by decreasing the rate of N uptake (Gojon et al., 2023; Soussana et al., 2005; Taub & Wang, 2008). However, the effect of high CO2 on N uptake seems to be different depending on the N source, affecting more negatively NO3- uptake than NH4+ (Rubio-Asensio & Bloom, 2017). This effect has been described in several species like Arabidopsis thaliana, wheat (Bloom et al., 2010), Medicago truncatula (Guo et al., 2013), or rice (Shimono & Bunce, 2009). Regarding Arabidopsis, Li et al. (2008) reported that elevated CO2 enhanced the expression of transcripts of glycolytic and TCA cycle enzymes, downregulated plastid functions and modulated the expression of N metabolism genes, which resembled N-deficiency conditions. Moreover, elevated CO2 promoted the storage and consumption of C-compounds, including several organic acids of the TCA cycle. However, the increase in organic acids, substrates for transamination, did not result in increased amino acid accumulation except for aromatic amino acids. Also, high CO2 produced a dramatic increase of heat shock proteins and transcripts of glutathione transferases (Li et al., 2008), resembling an abiotic stress response. Authors argued that increased photosynthetic CO2 fixation alters the apparent C/N balance and this imbalance seems to constrain the possible growth benefits from elevated CO2. These effects are more pronounced under N deficiency, producing a decreased shoot/root ratio, reduced NO3- content and accumulation of anthocyanin in seedlings (Takatani et al., 2014). Additionally, in mature plants, high CO2 under N deficiency also promotes leaf chlorosis, anthocyanin accumulation and increased senescence-associated gene expression (Aoyama et al., 2014). Indirect and physiological processes may explain the negative effects of elevated CO2 on N uptake and assimilation. For example, high CO2 triggers the closure of the stomata, which produces a reduction in stomatal conductance and transpiration. This could result in lower translocation of nutrients via the xylem and reduced acquisition of nutrients via roots (Gojon et al., 2023; Tanner & Beevers, 2001). This could also be explained as a consequence of an excessive accumulation of C-compounds (Li et al., 2008; Krämer et al., 2022). Nevertheless, a CO2-sensor has recently been identified, thus emerging as a possible direct signalling pathway (Takahashi et al., 2022; see sensors and transporters section).
Plant trioses phosphate synthesized by photosynthetic activity are converted into transient starch in chloroplasts or converted to sucrose in the cytoplasm through a series of enzymatic reactions (McClain & Sharkey, 2019). Sucrose is then transported to sink organs and stored in vacuoles or cleaved into fructose and glucose or UDP glucose. Out of these compounds, glucose and sucrose are the main sugar signals and they can show hormone-like activities, being directly sensed and producing some direct effects on N signalling or metabolism (Li et al., 2021a; Sakr et al., 2018). For example, in Arabidopsis, the expression of key genes of NO3- uptake, reduction and assimilation is enhanced by exogenously supplied sugar, which is partly dependent on the glucose sensor HEXOKINASE1 (HXK1)-mediated signalling (Lejay et al., 2003; Price et al., 2004; Figure 1). In rice, upon glucose perception, a mechanism involving the glucose sensor HXK7 and the ARE4 transcription factor (TF) activates the expression of a subset of high-affinity NO3- transporter genes (Ma et al., 2023a; see sensors and transporters; Figure 1). On the other hand, excessive accumulation of glucose in response to high C availability leads to repression of photosynthetic genes and produces stomatal closure in an HXK1-dependent manner, which entails reduced nutrient acquisition (Kelly et al., 2012, 2013). All these mechanisms seem to contribute to the maintenance of C/N balance in the plant. Interestingly, the impact of high C availability on plant development seems to be dependent on N levels. It was observed that a high concentration of glucose (6%) on MS media (N sufficiency) produces seedling growth arrest involving abscisic acid (ABA) and ethylene signalling. However, in the absence of N, a lower concentration of 2% glucose resulted in a similar developmental arrest without involvement of ABA or ethylene signalling (Cho et al., 2010; Lu et al., 2015a; Pourtau et al., 2004). This situation (low N and 2% glucose supply) induced leaf yellowing and changes in gene expression with high similarity to those observed during developmental senescence (Pourtau et al., 2004, 2006; Wingler et al., 2004, 2009). During leaf senescence, N from photosynthetic proteins of old leaves is exported and reused for the photosynthetic apparatus of younger organs or stored in seed storage proteins. Although photosynthesis declines during this process, in the absence of strong sinks, old leaves can accumulate sugars and signal an excess of C relative to low N availability (Wingler, 2018), suggesting that this situation might provoke a high C/N stress response.
The other major plant monosaccharide, fructose, has also been identified as a specific sugar signal. It was first noticed from observations that psicose, a fructose analogue, inhibits lettuce root growth in an HXK1-independent manner (Kato-Noguchi et al., 2005). Thus, specific components of fructose signalling, independent of HXK1-mediated glucose signalling, have been identified. These components included the NAC089 TF (Li et al., 2011) and the cytosolic FINS1 protein (fructose-1,6-bisphosphatase; see sensors and transporters section). Like glucose, the addition of exogenous fructose to the medium produces seedling growth arrest in Arabidopsis (Cho & Yoo, 2011). But interestingly, and unlike glucose, decreasing the concentration of fructose in the absence of N did not trigger developmental arrest and had little effect on Arabidopsis seedling growth, suggesting that fructose can signal independently of glucose and N (Cho & Yoo, 2011). Additionally, it seems that fructose might also be a shoot-to-root signal for the start of the light period since it peaks in the early morning and is low during the rest of the diurnal cycle (Brauner et al., 2018).
Sucrose, like glucose, also produces seedling growth arrest depending on N availability (Gao et al., 2008). However, there are several sucrose-specific responses unique to plants and unrelated to those produced by its derivates or by HXK1 (Chiou & Bush, 1998; Loreti et al., 2000; Wind et al., 2010). For example, sucrose downregulation of Beta vulgaris sucrose transporter SUT1 is not elicited by hexoses and is not impeded by an HXK inhibitor (Chiou & Bush, 1998; Vaughn et al., 2002). Also, sucrose-mediated PAP1 induction, a TF involved in anthocyanin metabolism, is not produced by other sugars (Teng et al., 2005). In this case, the mechanism of sucrose perception remains unknown, although several candidates have been proposed. For instance, there is a pseudo-perception mechanism dependent of sucrose (and not hexose) concentration based on nascent peptide-mediated ribosome stalling at the stop codon of the Arabidopsis uORF2 upstream of the bZIP11 TF (Rahmani et al., 2009). This mechanism has a direct connection with N signalling since bZIP11 reprograms amino acid metabolism in response to sucrose (Hanson et al., 2008). A novel sucrose signalling pathway has been proposed linking plant response to constant changes in sucrose status with photosynthesis, and consequently control of plant growth and development (Nguyen et al., 2016).
Like sucrose, trehalose 6-phosphate (T6P) is a non-reducing disaccharide. During the evolution of higher plants, sucrose became the most common sugar and T6P (intermediate in the synthesis of trehalose) evolved a new role as ‘plant insulin’ by signalling how much sucrose is available for growth. T6P levels in plants increase along with sucrose levels and the first regulates the production of the latter (Figure 1). If the plant is producing too much sucrose during photosynthesis, T6P signals to invest some fixed C into organic acids and amino acids. At night, T6P also functions as an inhibitor of starch breakdown to control the kinetics of sucrose synthesis and avoid the complete depletion of starch reserves (Baena-González & Lunn 2020; Fichtner et al. 2021a; Figueroa & Lunn 2016). Regarding T6P interaction with N, there are three direct connexions. First, in high sugar/low N seedling growth arrest experiments, endogenous T6P concentration, and not glucose, correlates with C and N availability in the medium (Li et al., 2020b). Second, high T6P levels stimulate NO3- assimilation and synthesis of organic acids to generate C skeletons for amino acid synthesis (Figueroa et al., 2016). Third, flowering in short-day conditions relies on T6P signalling under low N (Olas et al., 2019).
In addition, to serve as C skeletons for amino acid synthesis, the organic acids generated in the TCA cycle are also precursors of hormones and other signalling molecules in plants. A recent work has suggested that these organic acids from the TCA cycle can act as hormone-like molecules and activate signalling pathways. Thus, exogenous addition of some of them (i.e., succinate and α-ketoglutarate) produces distinct root phenotypes in stem cell divisions and growth, and these phenotypes are not sufficiently explained by changes in ATP or ROS levels (Zhang et al., 2023). In any case, these effects and a possible connection with N or nutritional status remain unexplored.
Finally, NO3- is a positive signal for the induction of sugar unloading and assimilation, and several key genes for C metabolism, like pyruvate kinase, phosphoenolpyruvate carboxylase, 6-phosphogluconate dehydrogenase, and glucose-6-phosphate dehydrogenase, are induced by NO3- (Scheible et al., 2004; Wang et al., 2003). NO3-, as well as NH4+ and amino acids, activate the TOR kinase (Cao et al., 2019; Liu et al., 2021; see C/N signalling section), but the direct molecular links between different N sensors and N-signalling pathways responsible for TOR activation are still unknown in plants (Liu et al., 2021).
3 SUGAR AND NITROGEN SENSORS AND TRANSPORTERS
Maintaining an appropriate balance or ratio of C and N involves the sensing and signalling of C and N status throughout the plant organs. NO3- and/or NH4+ are usually taken up by the root system and NO3- is transported to the leaf. Meanwhile, C is assimilated in the form of sucrose, which is the main long-distance transported sugar, and exported from the source leaves to sink organs via phloem (Figure 1). Assimilated N is converted to amino acids, transiently stored in source leaves and roots and relocated to sink organs (Lalonde et al., 2004; Sakr et al., 2018). After assimilation, N is mainly transported to the mesophyll cells through the xylem in the form of Gln, Asp, Glu, and Asn for C assimilation. Thus, sink organs, such as flowers, fruits, seeds, and younger leaves, are the primary reserve for C and N assimilates, while source organs, such as older leaves and green plant parts, export carbohydrates produced by photosynthesis. The CO2 assimilation capacity is mainly affected by the N content of the leaves and is generally positively correlated with it. Under N deficiency conditions, leave's chlorophyll content and photosynthetic rate decrease. In consequence, plants have developed sensing mechanisms to control their photosynthetic output in response to the levels of C and the status of N in the soil and in their root system (Wingler et al., 2018; Xuan et al., 2017). The mechanisms underlying sugar and NO3- sensing are complex and involve different N and sugar transporters. These proteins must coordinate the N assimilation with C metabolism, nutrient availability and other environmental factors (Zheng et al., 2009).
At the front line, plants are able to sense changes in the CO2 concentration, and elevation of leaf CO2 concentration triggers a rapid stomatal closing, leading to reduced transpiration. Changes in transpiration have been associated with alteration in translocation of nutrients and their acquisition through the roots (Gojon et al., 2023). Sensing of CO2 involves negative regulators of high CO2-induced stomatal closure, the HT1 (Hashimoto et al., 2006), CBC1 and CBC2 kinases (Hiyama et al., 2017), and positive regulators of early CO2 signal transduction, the MPK4/MPK12 kinases (Tõldsepp et al., 2018). Very recently, it was discovered that MPK4/12 and HT1 constitute the long-sought primary stomatal CO2 sensor upstream of the CBC1 kinase. Thus, increased CO2 levels trigger the interaction between MPK4/MPK12 and HT1, which inhibits HT1 kinase activity and promotes stomatal opening. At low CO2, HT1 phosphorylates and activates the downstream CBC1 kinase (Takahashi et al., 2022; Figure 1).
Plants have evolved a complex mechanism to sense different sugars, including sucrose, glucose and fructose, causing a response that is sometimes specific to the type of sugar (Rolland et al., 2006; Smeekens & Hellmann, 2014). Sucrose is the main form of C; it is synthesized in the cytoplasm of mesophyll cells and initially transported out of mesophyll cells into apoplast via SWEET transporters (Chen et al., 2012). Then, sucrose is imported into companion cells from the apoplast through membrane-localized SUTs/SUCs symporters (Scofield et al., 2007; Figure 1). While SUT/SUC family transporters studied to date, including vacuolar SUTs, are sucrose/H+ symporters, SWEETs are able to independently transport sucrose and fructose across cell membranes (Gautam et al., 2022; Hu et al., 2021; Jeena et al., 2019). SWEETs also participate in abiotic stress responses by maintaining intracellular sugar concentrations (Lu et al., 2019; Mathan et al., 2021; Miao et al., 2017). Interestingly, some SWEET members have also been demonstrated to transport hormones in addition to sugars. Specifically, two Arabidopsis SWEET genes (AtSWEET13 and AtSWEET14) are involved in the transportation of various forms of gibberellins (GAs; Kanno et al., 2016). Sucrose transport and N levels need to be coordinated to maintain C/N homeostasis. It has been described that the transport from source to sink mediated by different sugar transporters is more efficient at sufficient than at deficient N levels. Accordingly, N levels affect the expression of SWEETs and SUTs in different plant species like Arabidopsis, maize or rice. For instance, AtSWEET16 expression decreases under low N levels and its overexpression enhances growth and N content at adequate N supply but not under N deficiency (Klemens et al., 2013). Interestingly, it has been showed that enhanced expression of OsSWEET11 and OsSWEET14 increased grain filling and grain yield even under low N conditions in rice (Fei et al., 2021). In barley, HvSWEET11 expression is induced more efficiently by NH4+ than by NO3-, suggesting a differential role of SWEET transporters depending on the source of N (Lopes et al., 2008). Finally, AtSWEET17 modulates fructose balance under different abiotic stresses, including low N (Chardon et al., 2013; Guo et al., 2014). In summary, N status seems to regulate the functionality of some of the sugar transporters to balance C/N levels by affecting their gene expression and protein levels.
The evolutionary conserved HXK1 hexokinase family participates in sugar metabolism and sensing (Moore et al., 2003). This family of proteins catalyse the phosphorylation of hexose sugars in the cytoplasm as the first step of the glycolytic pathway. However, Arabidopsis HXK1 can function as a glucose sensor, independently of its catalytic activity, by forming a nuclear complex with the vacuolar H+-ATPase B1 (VHA-B1), the 26S proteasome AAA-ATPase subunit (RPT5B), and putative TFs (Cho et al., 2006). Thus, HXK1 interacts with these proteins in the nucleus to control transcription and diverse glucose responses. Under high glucose levels, the HXK-mediated pathway inhibits photosynthesis and transpiration as well as seedling development but, under N sufficiency, HXK1 induces growth in vegetative and reproductive organs (Cho et al., 2007; Vanderwall et al., 2023). Recently, in rice, the glucose sensor OsHXK7 has been found to form a complex with the ARE4 TF in the cytoplasm. Upon sensing a glucose signal, ARE4 is released and translocated into the nucleus to activate the expression of a subset of high-affinity NO3- transporter genes to increase NO3- uptake and accumulation (Ma et al., 2023b; Figure 1). Thus, this mechanism coordinates glucose signalling and N utilization in rice. Interestingly, this mechanism displays a diurnal pattern in response to circadian changes of soluble sugars (Ma et al., 2023a).
Fructose is also one of the prevalent sugars in plants (Kato-Noguchi et al., 2005). Fructose signalling induces seedling developmental arrest and interacts with plant stress hormone signalling in a similar manner than glucose. Under N deficiency, fructose overaccumulates in leaves (Chardon et al., 2013) and exerts its regulation on seedling growth through FINS1, a cytosolic fructose 1,6-bisphosphatase, whose function in fructose signalling is independent of its catalytic activity in sugar metabolism (Cho & Yoo, 2011). FINS1 protein has been suggested to be a fructose sensor capable of locating to the nucleus, where it interacts with WUS/WOX TFs to bind WUS/WOX promoters and regulates their activity to control stem cell function in a fructose dose-dependent manner (Cho & Yoo 2011; Li et al., 2023).
NO3- is the main form of N absorbed by plants and, similarly to some sugars, it acts as a signal in plant physiology and development (Wang et al., 2018b). NO3- is absorbed from the soil into root cells, usually by the dual-affinity NRT1 and high-affinity NRT2 transporters (Chopin et al., 2007; Liu et al., 2022b; O'Brien et al., 2016; Tsay et al., 1993). NRT1.1/NPF6.3 (hereafter NRT1.1) is one of the most well-studied members of its family and was initially characterized as a transporter involved in NO3- uptake by roots as well as root-to-shoot NO3- translocation (Léran et al., 2013; Liu et al., 1999). Its affinity for NO3- can change in response to substrate availability. Thus, phosphorylation of a threonine residue, Thr101, switches NRT1.1 from low- to high-affinity state (Liu & Tsay, 2003). Later, it was shown that NRT1.1 functions as a NO3- sensor independently of its function as a transporter (Ho et al., 2009). It is involved in NO3--dependent gene expression and root development and its ability to sense NO3- concentrations depends on its phosphorylation state. Thus, the CIPK23 kinase phosphorylates NRT1.1 to maintain a low-level primary response under low NO3- concentration (Ho et al., 2009). On the contrary, ABI2 phosphatase can dephosphorylate CIPK23, thereby suppressing its kinase activity, thus enhancing NRT1.1 activity (Léran et al., 2015). Subsequently, the association of the NRT1.1 with the cyclic nucleotide-gated channel protein (CNGC15) promotes the Ca+2 influx through the cell membrane (Wang et al., 2021c). Then, several Ca+2-dependent kinases (CPK10/30/32) are able to induce the phosphorylation of the NLP7 TF to promote its shuttling from cytoplasm to the nucleus (Liu et al., 2017). NLP7 activates primary NO3- response genes (Konishi et al., 2013; Liu et al., 2022a) and it has been shown to be the primary NO3- sensor in plants. Upon binding to NO3-, NLP7 suffers a conformational change favoring its nuclear retention and NO3--responsive transcription (Liu et al., 2022a; Marchive et al., 2013; Figure 1). Interestingly, some members of the NRT1/NPF family have also been shown to transport additional substrates (Kanstrup et al., 2022). NRT1.1 transports auxin that mediates differential root growth depending on soil's NO3- supply (Krouk et al., 2010). Thus, NO3- influx competes with auxin import, the latter inhibiting root elongation and branching, to adjust root architecture to NO3- availability. Similarly, NPF5.12/TOB1 has been shown to transport NO3- and indole-3-butyric acid, an auxin precursor, to mediate cytokinin regulation of root architecture (Michniewicz et al., 2019). Recently, it has been discovered that sugars can also be substrates for NRT1/NPF activity. Thus, SUCROSE AND GLUCOSE CARRIER 1 (ZmSUGCAR1), a maize NRT1/NPF-type transporter, is able to transport sucrose and glucose. Loss-of-function mutation of ZmSUGCAR1 caused a significant decrease in sucrose and glucose content, provoking a reduction in kernel size. Moreover, it seems that this function is conserved in cereals since orthologues of ZmSUGCAR1 from wheat and sorghum displayed similar sugar transport activities (Yang et al., 2022).
While members of the NRT1 protein family serve mainly as low-affinity transporters, the NRT2 protein family functions as high-affinity transporters. The roots are the predominant site of expression for all AtNRT2 genes and the responsivity and induced expression of different NRT2 genes to changes in NO3- availability vary significantly. Thus, in Arabidopsis, NRT2.1, NRT2.2, NRT2.4, and NRT2.5 are mediators of high-affinity NO3- uptake, although the latter two proteins function only under N starvation (Ruffel et al., 2011). The high-affinity NO3- transporter NRT2.1 is the main transporter under low N conditions (Zhuo et al., 1999) and is transcriptionally and post-transcriptionally regulated by C- and N-derived metabolites, making it a central player in the integration of these signals. Thus, NRT2.1 is induced both by N starvation and by light and sugars, suggesting that it mediates the coordination between N signalling and photosynthesis (Cerezo et al., 2001; Gansel et al., 2001; Lejay et al., 1999, 2003). Lateral root initiation is strongly repressed in Arabidopsis by the combination of high external sucrose and low external NO3- and NRT2.1 seems to act as a sensor for this C/N balance response in lateral root formation (Asim et al., 2020; Little et al., 2005; Zhang et al., 2007).
Finally, the glutamate receptor GLR1.1 has also an important role in C/N balance response. The plant glutamate-like receptors (GLRs) are homologs of mammalian ionotropic glutamate receptors (iGluRs) and function in several processes related to plant growth, development, physiology and environmental stress responses (Weiland et al., 2015). Thus, the activities or transcript levels of several genes encoding enzymes related to C or N metabolism are reduced in plants defective in this gene. Interestingly, these plants also showed altered expression of several ABA biosynthetic and signalling genes together with enhanced drought tolerance, suggesting a role for this receptor in plant response to water stress (Kang et al., 2003, 2004). Further studies will elucidate how AtGLR1.1 integrates C/N ratios and stress response with plant development.
4 C/N SIGNALLING
TOR and SnRK1 are two kinase complexes conserved among eukaryotes that play antagonistic key roles in nutrient signalling (Artins et al., 2024; Artins & Caldana, 2022; Dobrenel et al., 2016; Fichtner et al., 2021b; McCready et al., 2020; Peixoto & Baena-Gonzalez, 2022; Ryabova et al., 2019; Wu et al., 2019). The TOR pathway responds to energy, nutrients, environmental cues and hormones and is activated under nutrient sufficiency to adjust growth according to available resources. Important enzymes involved in the metabolism of C and N compounds are putatively phosphorylated by TOR (Van Leene et al., 2019). It is worth mentioning that, while TOR expression in mammals is detected in all cells (Chiu et al., 1994), TOR expression in plants is limited to the zones where cell proliferation is coupled to cell growth, mostly cytosolic, such as embryo, endosperm and primary apical and root meristems (Menand et al., 2002). Particularly relevant for this growth is the glucose-TOR signalling that integrates light, sugar availability (photosynthetic activity and C-assimilation) and auxin levels to control meristem activation in roots and shoots (Chen et al., 2018; Li et al., 2017; Pfeiffer et al., 2016; Xiong et al., 2013; Figure 2). The absence of TOR expression in differentiated mature cells, whose growth process is mostly related to vacuole and cell-wall expansion instead of cytosolic, points to the existence of a different nutrient sensing mechanism (Menand et al., 2002). In addition to its activation by sugars (sucrose and glucose), TOR also couples changes in sugar levels to the metabolism of N-compounds. For instance, studies in Chlamydomonas showed that impaired CO2 fixation inhibits TOR activity, which correlates with a gradual decrease of most amino acids (Mallen-Ponce et al., 2022). Starch synthesis is essential to keep photosynthetic electron transport capacity and, when impaired, causes a drop in photosynthetic function and a redirection of photosynthetically fixed C (Saroussi et al., 2019). Remarkably, a mutant unable to synthesize starch had higher TOR activity, probably due to the concomitant elevation of Gln concentration (Mallen-Ponce et al., 2022), an amino acid known to activate TOR (Gonzalez & Hall, 2017). TOR activity can also be altered by inorganic and organic N-signals. In leaf primordium of Arabidopsis under N-deficiency, TOR is activated when NO3- and NH4 are supplied, independently of their assimilation or glucose-energy and hormone signalling pathways (Liu et al., 2021; Figure 2). In the same study, it was found that 15 out of the 20 proteinogenic amino acids were able to activate TOR, with several of them being derived from specific C-compounds produced in glycolysis (pyruvate) and TCA (oxalacetate). A previous study showed the ability of certain amino acids to upregulate TOR signalling and impact homeostasis and organization of actin cytoskeleton and endomembranes (Cao et al., 2019; Figure 2). In line with the above, TOR activates growth in response to C and N signals through the promotion of protein synthesis by phosphorylation of the E2Fa TF and the S6K ribosomal protein kinase (Li et al., 2017; Henriques et al., 2013; Schepetilnikov et al., 2013; Xiong et al., 2013; Figure 2). Genetic or chemical downregulation of components of TOR leads to a fast increase in the levels of amino acid levels in plants (Salem et al., 2017), which could be explained by the inability to use them efficiently for protein synthesis. TOR also promotes protein translation by coordinating ribosomal RNA synthesis with nucleotide availability (C- and N-containing compounds) and its activity is inhibited when nucleotide biosynthesis is disrupted (Busche et al., 2021; Figure 2).
Graphic representation of SnRK1 and TOR regulatory networks and their interaction with C/N signalling pathways. The left side of the figure depicts the relationship between SnRK1 and sugar-related transporters and signalling proteins, as well as NO3- sensing. The right side represents links between TOR, protein and nucleotide synthesis, and regulators of C/N balance. Both sides are interconnected by C and N-uptake/assimilation processes shown in the middle of the figure. The formation of complexes is indicated by dotted lines. Glc: Glucose. AAs: Amino Acids.
It is well-known that energy and C starvation activate SnRK1 to phosphorylate and inactivate TOR. In addition, SnRK1 also modulates the phosphorylation status of about 500 proteins (Cho et al., 2016; Li et al., 2020b; Nukarinen et al., 2016; Van Leene et al., 2022), affecting the activity of key enzymes for N, C or fatty acid metabolism (Harthill et al., 2006; Kulma et al., 2004). For example, SnRK1 also induces autophagy processes through differential phosphorylation of ATG (AuTophaGy) proteins, and acts upstream of the negative regulation exerted by TOR on the same processes (Chen et al., 2017; Huang et al., 2019; Liu & Bassham, 2010; Mugume et al., 2020; Pu et al., 2017; Soto-Burgos & Bassham, 2017; Figure 2). Activation of SnRK1 also triggers its translocation to the nucleus to promote transcription of over 1,000 genes destined to reduce energy consumption (Baena-González et al., 2007; Ramon et al., 2019). Part of this regulation is exerted by the phosphorylation of bZIP63 and putatively other C/S1 bZIP TFs (Baena-González et al., 2007; Baena-González & Sheen, 2008; Mair et al., 2015). For instance, SnRK1 promotes the catabolism of triacylglycerols (TAGs) and amino acids during seedling establishment to fuel gluconeogenesis through bZIP63 (Henninger et al., 2022; Figure 2). SnRK1 has also been proposed to couple energy metabolism with growth under diurnal cycles through the clock-regulated bZIP63 by preventing the acceleration of starch degradation at the end of the night (Viana et al., 2021). Interestingly, SnRK1 signalling is higher at the end of the night, coinciding with the lowest levels of T6P (Peixoto et al., 2021). As mentioned above, T6P is a proxy for sucrose levels and promotes starch biosynthesis, interplaying antagonistically with SnRK1 to control sucrose homeostasis (Figure 2). A clear connection between the role of SnRK1 on the control of sucrose levels and N metabolism comes from maize. Lower activity of SnRK1 under increased sucrose levels has been linked to the translocation of the O2 TF to the nucleus to promote zein synthesis, the main N-storage compound in maize seeds (Li et al., 2020a). However, when SnRK1 is artificially overexpressed in Arabidopsis plants grown under elevated glucose/sucrose levels, growth arrest is promoted due to enhanced phosphorylation and activity of SnRK1 (Jossier et al., 2009). Finally, SnRK1 activity is also under the control of N-signals given that NO3- promotes degradation of the catalytic subunit of the SnRK1 complex (KIN10), while NO3- depletion induces KIN10 accumulation (Wang et al., 2022; Figure 2). Whether this mechanism depends on C/N ratios remains to be elucidated.
Interestingly, KIN10 phosphorylates NLP7, the NO3- sensor, to promote its cytoplasmic localization and degradation (Wang et al., 2022). Thus, N-sufficiency favours NLP7 nuclear localization and signalling, both of which would be counteracted by SnRK1 (KIN10) under N-deficiency. NLP7 was also found to orchestrate a NO3--activated pathway that resulted in improved light energy utilization under high light and low N conditions. This pathway enhances the expression of a subunit of the chloroplast VAR2 protease by a cascade of direct transcriptional activations. These activations are carried out chronologically by NLP7 and the HB52/54 TFs, which facilitate the removal of photodamaged-thylakoid membrane proteins (Ariga et al., 2022; Figure 2). In this sense, the chloroplast is an important organelle for nutrient homeostasis and the location of the PII protein, a highly conserved signal transduction regulator of C/N interactions. In plants, the PII gene is encoded by the nuclear genome and targeted to the chloroplast, and its transcription is positively regulated by light and sucrose and negatively by amino acids (Hsieh et al., 1998; Nunes-Nesi et al., 2010). Although its role in C and N metabolism may not be as important as in bacterial systems, several results indicate that PII connects C/N pathways in plants. For instance, PII regulates the first committed step of Arg biosynthesis by interacting with a chloroplastic kinase (Chen et al., 2006; Ferrario-Méry et al., 2006) and PII mutants show reduced accumulation of Arg and Gln, and higher levels of starch and sugars when grown under NH4+ (Ferrario-Méry et al., 2005, 2008). In addition, PII seems to inhibit lipid biosynthesis since it interacts with and inhibits the first enzyme of fatty acid synthesis in plastids (ACCase), an inhibition reverted by 2-OG (Feria Bourrellier et al., 2010). Interestingly, PII expression is upregulated during early seed development by the WRI1 TF, a master positive regulator of glycolysis and seed-storage oil synthesis. This effect might enable plants to fine-tune fatty acid content (Baud et al., 2010) and to channel some of the C-skeletons towards N-compounds.
An aspect not very well studied in C/N responses is the possible existence of organ-type or/and developmental-specific signalling mechanisms. For instance, another relevant study on this matter revealed that C- and N-acquisition is coordinated between roots and shoots by the mobile HY5 TF (Chen et al., 2016). Upon illumination of the shoot, HY5 stimulates NO3- uptake by moving from shoot to root through the phloem, where it activates its own expression and NRT2.1 expression. This effect is enhanced by sucrose/glucose via an unknown mechanism, probably to keep balanced C/N ratios. In addition, HY5 also affects both sucrose metabolism and shoot-to-root transport by promoting the expression of T6P synthase (TPS1) and SWEET11/12 sucrose efflux transporters (Chen et al., 2016; Figures 1 and 2). These SWEET genes are also among those directly transactivated by several bZIP TFs that were found to promote apical dominance and restrict lateral organ formation by favouring redistribution of C- and N-compounds from source leaves to apical organs (Kreisz et al., 2024). Loss of S1-type bZIPs function produced reduced leaf sugar export to the apoplast for long-distance transport to distal sinks, which increased the symplastic sugar flux to proximal sinks (increased C/N ratios in rosette leaves). Simultaneously, the lack of glutaminase transactivation rendered increased levels of Gln and Asn in seedling roots and increased Gln in the shoot just before bolting (Kreisz et al., 2024). These and previous findings position the S1-type bZIPs as key players in C/N and energy responses with a wide regulatory spectrum. For instance, under low-energy conditions, they promote the degradation of proline and branched amino acids to feed C skeletons into alternative respiratory pathways for ATP production (Pedrotti et al., 2018; Figure 2). They also heterodimerize with the C-type bZIP63 to mediate key aspects of SnRK1 functions (Ehlert et al., 2006; Henninger et al., 2022; Muralidhara et al., 2021; Mair et al., 2015). The above results also point to the possible existence of regulatory and molecular interactions between HY5 and S1-type bZIPs, which may have been hidden so far due to the genetic redundancy of the bZIPs.
14-3-3 proteins (also and hereafter called General Regulatory Factors, GRFs) can be found in different subcellular locations and have extensive regulatory effects on C and N metabolism in plants. GRF proteins generally interact with phosphorylated target proteins to regulate their functions by different mechanisms (Jaspert et al., 2011; Zhao et al., 2021). Several of the GRF target proteins are enzymes of the C and N metabolism (Comparot et al., 2003) as relevant as NO3- reductase (Bachmann et al., 1996), sucrose-phosphate synthase (Moorhead et al., 1999; Toroser et al., 1998), starch synthase III family members (Sehnke et al., 2001), trehalose-6-phosphate synthase or cytosolic and plastidic forms of glutamine synthetase (Moorhead et al., 1999; Figure 2). GRF proteins also seem to be directly linked to energy status since 5´-AMP has been shown to reduce the interaction between GRF and their targets (Athwal et al., 1998). Unlike animals, plants use mostly protons and H+-ATPase pumps to transduce chemical and environmental signals and support a range of physiological functions (Sze et al., 1999). Interestingly, several GRFs have been shown to interact with and regulate chloroplast, mitochondrial and plasma membrane ATPases (Bunney et al., 2001; Jahn et al., 1997; Sze et al., 1999). Several lines of evidence point to the existence of functional redundancy in the GRF family and their differential regulation of metabolic enzymes. Thus, Arabidopsis GRF1, GRF3 and GRF8 proteins showed overlapping and specific interactions with 150-200 proteins. Furthermore, when the activities of enzymes targeted by GRF were compared by using gain- and loss-of-function mutants, results confirmed the existence of partial functional redundancy in this family of regulators (Shin et al., 2011). In a separate study, overexpression of Arabidopsis GRFs triggered decreased levels of sugars, malate, citrate and N-containing-compounds as well as a reduction in the activities of known GRF-interacting-enzymes such as the TCA cycle enzymes isocitrate dehydrogenase and malate dehydrogenase (Diaz et al., 2011). Additional evidence associating their negative regulation on the activity of several C- and N-biosynthetic enzymes was obtained from the downregulation of several GRFs in potato and Arabidopsis (Sehnke et al., 2001; Świȩdrych et al., 2002; Zuk et al., 2003, 2005). In addition to regulating the activity of their target enzymes, molecular evidence of proteolytic degradation of GRF target proteins has been obtained in Medicago truncatula (Lima et al., 2006) and under sugar starvation in Arabidopsis (Cotelle et al., 2000).
The stress caused by high C/N induces the accumulation of GRF proteins, and hypersensitivity to this stress has been observed when ectopically overexpressing GRF1 (Sato et al., 2011). High C/N also induces ATL31/CNI1 expression, an E3-ligase that interacts with GRF proteins at the membrane and mediates their degradation by the proteasome. This mechanism allows modulation and fine-tuning of GRF activity to control the response and reduce stress sensitivity to high C/N stress (Aoyama et al., 2014; Sato et al., 2011; Yasuda et al., 2014). ATL31 ligase, as well as it closest relative ATL6, are under the positive control of a group of kinases. CIPK7, 12 and 14 phosphorylate and stabilize ATL31 and both activities are required for the interaction and degradation of GRF target proteins (Yasuda et al., 2017). Likewise, the FER kinase mediates ATL6 phosphorylation to enhance the interaction between ATL6 and its target GRF proteins at the membrane (Xu et al., 2019). Not surprisingly, loss-of-function in these kinases (CIPKs or FER) causes hypersensitivity to high C/N ratios (Xu et al., 2019; Yasuda et al., 2017). It was also found that the direct phosphorylation of ATL31 by CIPK14 is preceded by the interaction of the kinase with the CBL8 Ca2+ sensor at the plasma membrane in a Ca2+-dependent manner (Batistič et al., 2010; Yasuda et al., 2017). This event highlights the existence and relevance of non-nuclear control mechanisms for C/N responses and growth. In support of this evidence, FER is also a membrane receptor kinase that mediates C/N responses and has additional roles in the control of growth and stress responses (Liu et al., 2023; Song et al., 2022; Tang et al., 2022; Yang et al., 2015). FER is bound and activated by RALF1, a secreted peptide that induces a rapid increase in cytoplasmic Ca2+ (Haruta et al., 2014). FER, in turn, interacts with and phosphorylates ATL6 at the membrane to reduce GRF levels and sensitivity to high C/N (Xu et al., 2019). Moreover, RALF1 increases TOR activation by enhancing FER-TOR interactions, promoting true leaf growth under N-deficient conditions in Arabidopsis (Song et al., 2022). All these results underscore the existence of different tissue-specific outcomes of the RALF1-FER interaction and the impact of several subcellular compartments and developmental stages to signal the C/N status (Figure 2).
5 PHYTOHORMONES IN C/N SIGNALLING
Efficient and rapid communication throughout the plant body is key for successful coordination of growth and function in multiple organs with different nutrient needs. This requires both local and long-distance signalling for the integration of environmental, nutritional and developmental signals (Yu et al., 2015; Wheeldon & Bennett, 2021). In recent years, phytohormones have been increasingly linked to energy and nutrient homeostasis, particularly C and N metabolism (Ahmad et al., 2023; Fabregas & Fernie, 2021, 2022; Fichtner et al., 2021b; Jamsheer et al., 2019; Sakr et al., 2018). In some cases, the impact of high C availability on plant development and hormone signalling seems to be dependent on N levels (Cho et al., 2010; Lu et al., 2015b; Pourtau et al., 2004). This section will highlight their emerging roles in C/N nutritional balance by focusing on the links between hormone pathways, the kinase complexes TOR and SnRK1, and C and N signalling.
5.1 Abscisic acid
ABA regulates growth at multiple developmental stages throughout the life of the plant and a wide range of physiological processes in response to stress signals and changing environments (Chen et al., 2020; Parwez et al., 2022). The core of ABA signalling relies on the hormone recognition by intracellular receptors (PYR1/PYL/RCAR), triggering a conformational change to allow binding to PP2Cs phosphatases, thus releasing their inhibition on SnRK2 protein kinases (Umezawa et al., 2010; Figure 3). SnRK2s activate multiple physiological responses by phosphorylating downstream substrates such ABF TFs (Umezawa et al., 2013; Wang et al., 2013).
Overview of phytohormones main signalling mechanisms linked to C/N balance. The different hormonal and signalling components are shown from left to right: Cytokinins (CK), abscisic acid (ABA), auxins, ethylene (ET), strigolactones (SL), gibberellins (GA) and brassinosteroids (BR). The elements of each hormone's network are shown in the same colour. The formation of complexes is indicated by dotted lines. Glc: Glucose.
Over the years, various links have been established between ABA and the key metabolic regulators SnRK1 and TOR. Genetic, molecular and physiological analyses have revealed the involvement of SnRK1 in both sugar and ABA signalling pathways (Jossier et al., 2009), including largely overlapping transcriptional responses (Carianopol et al., 2020). Thus, under optimal conditions (low ABA), PP2C-SnRK2 complexes directly bind and inhibit SnRK1 (Belda-Palazón et al., 2020, 2022; Rodrigues et al., 2013), while stress scenarios would increase ABA levels and allow the coordinated activation of both ABA and SnRK1 pathways (Figure 3). A direct link between SnRK1 and ABA perception has also been identified since the overexpression of an ABA receptor (RCAR6) leads to the activation of both SnRK1 and SnRK2 (Yoshida et al., 2024). Downstream of ABA reception, SnRK1 activity is known to regulate ABA-related TFs, such as FUS3 (Chan et al., 2017; Gazzarrini et al., 2004; Tsai & Gazzarrini, 2012) or ABF2 (Rodrigues et al., 2013). On the other hand, reciprocal regulation of TOR and SnRK2 has been described. Thus, TOR phosphorylates ABA receptors to prevent activation of SnRK2 under favourable environmental conditions, whereas under stress ABA-activated SnRK2s trigger the dissociation and inhibition of the TOR complex (Wang et al., 2018a; Figure 3).
C and N signalling networks also interact with ABA at different intersection points. The genetic screen of mutagenized Arabidopsis plants for altered sugar sensitivity has often led to the identification of ABA-deficient or insensitive mutants (Fabregas and Fernie 2021). On the other hand, several lines of evidence have linked root adaptation to NO3-, a response that couples lateral root growth to NO3- levels with ABA signalling (Contreras-López et al., 2022; Kim et al., 2004; Léran et al., 2015; Ondzighi-Assoume et al., 2016; Rodrigues et al., 2013; Signora et al., 2001). The ABI1 phosphatase provides a direct crosstalk between C/N balance and ABA signalling pathway. ABI1 overexpression produces a tolerance phenotype under extremely high C/N ratios, accompanied by altered expression of ABA and SnRK1s-related genes (Lu et al., 2015a). The positive and master regulators of ABA signalling, ABI4 and ABI5, have also been linked to glucose and NO3- metabolism (Signora et al., 2001; Fabregas and Fernie 2021). ABI5 mediates the acclimation of Arabidopsis plants to low CO2 availability downstream of the phosphoenolpyruvate carboxylase PPC2, a response that requires the regulation of C and N metabolism (You et al., 2020). ABI4 act upstream of the transcription factor CHO1, which acts as a node of hormonal and nutritional signalling and regulates seedling growth under excess supply of glucose and NO3- (Yamagishi et al., 2009). Finally, ABI4 and ABI5 act synergistically in the activation of the diacylglycerol acyltransferase DGAT1 expression, the rate-limiting enzyme in TAG biosynthesis, in Arabidopsis seedlings under abiotic stresses, N deficiency and C/N imbalance (Yang et al., 2011; Kong et al., 2013).
5.2 Auxins
Auxin is an essential hormone that integrates endogenous developmental signals and exogenous environmental cues to coordinate the growth and development of spatially distant organs (Casanova-Sáez et al., 2021; Zhao, 2018). Over the years, accumulating evidence has demonstrated that the auxin pathway interacts with and coordinates C and N metabolism to shape the plant architecture at different growth stages and under dynamic environmental conditions (Fichtner et al., 2020; Gutiérrez et al., 2007). It has been shown that different soluble sugars derived from photosynthesis regulate auxin biosynthesis via HXK-dependent and independent signalling, as well as via PIF TFs (Lilley et al., 2012; Sairanen et al., 2012; Figure 3). Auxins, in turn, fine-tune sugar metabolism and source-sink C partitioning (Fichtner et al., 2020; Mishra et al., 2022). At the molecular level, in both shoot and root apical meristems, systemic glucose and enhanced auxin accumulation activates the ROP2 GTPase, which physically interacts with and triggers TOR phosphorylation and activation, upregulating cell cycle genes and DNA synthesis (Li et al., 2017; Schepetilnikov et al., 2013, 2017; Xiong et al., 2013; Figure 3). However, in the root elongation zone, glucose-activated TOR phosphorylates and stabilizes PIN2, an auxin efflux facilitator, to maintain the low auxin levels required for cell expansion and root patterning (Yuan et al., 2020). As with glucose, ROP2 has been described to integrate the presence of auxin and primary N signals (NO3-, NH4+ and Gln) to promote cell division and true leaf development through TOR activation (Li et al., 2017; Liu et al., 2021; Schepetilnikov et al., 2017; Figure 3). In addition, TOR has been recently described to promote the translation of the auxin-induced ARF7, ARF19 and LBD16 TFs, essential players in root branching, in a sugar-dependent manner (Stitz et al., 2023). How this C-mediated regulation by TOR is affected by N-availability remains to be determined. Consequently, disruption of the TOR pathway leads to repression of auxin biosynthesis, transport and signalling, resulting in impaired auxin-mediated plant growth and stress responses (Deng et al., 2016, 2017; Dong et al., 2015; Li et al., 2017; Schepetilnikov et al., 2013). On the other side, some links between NO3--supply and auxin translocation and accumulation have been reported. At low N availability, NRT1.1 prevents auxin accumulation by lowering its biosynthesis and transport to the tip of the lateral root primordia (LRP), thereby contributing to slowing down LRP development and emergence in NO3--poorer areas (Krouk et al., 2010; Maghiaoui et al. 2020; Mounier et al., 2014). The integration of N signals into auxin signalling has also been described in combination with cytokinins (CKs). NLP7 coordinates shoot growth in response to NO3- supply via stimulating CKs biosynthesis and transport from root to shoot. There, the enhanced expression of CRFs directly regulates PIN transporters, promoting the accumulation and, thereby, stimulating shoot growth (Asim et al., 2020; Abualia et al., 2022; Kumar et al., 2023; Figure 3). Collectively, these findings reinforce that auxin accumulation, distribution and function in the plant is vastly controlled by sugar metabolism and N availability.
5.3 Brassinosteroids
It is well established that brassinosteroids (BRs) play critical roles in plant growth and developmental processes but also responses to environmental stimuli (Nolan et al., 2020). At the first steps of their signalling cascade, BRs binding induces the phosphorylation of the receptor kinase BRI1, which triggers a phosphorylation cascade that results in the inactivation of the BIN2 kinase and subsequent activation of the BES1/BZR1 TFs, key positive regulators of downstream BR-responsive genes (Nolan et al., 2020; Han et al., 2023). In recent years, BR signalling research has led to the characterization of crosstalk between these core components, C metabolism, N availability and TOR signalling (Han et al., 2023). BRs are required for normal plastid development and CO2 assimilation (Schröder et al., 2014) as well as for TCA cycle activity, starch accumulation and biomass production (Siddiqui et al., 2018). Conversely, C availability modulates BR signalling in a dose-dependent manner (Han et al., 2023; Zhang et al., 2015). Thus, although proper concentration of sugar activates BR signalling, the pathway is inhibited under sugar starvation or surplus, mainly by reducing BZR1/BES1 activity (Han et al., 2023). Sugar starvation causes BZR1/BES1 degradation via selective autophagy (Nolan et al., 2017; Wang et al., 2021b; Zhang et al., 2016), a pathway enhanced by BIN2 and SnRK1, and negatively regulated by TOR (Liao et al., 2023; Soto-Burgos & Bassham, 2017; Figure 3). On the other hand, high concentrations of sucrose also inhibit BR-signalling by promoting the phosphorylation and cytosolic retention of BZR1 (Hao et al., 2016; Zhang et al., 2021). Moreover, BR signalling is also highly dependent on the availability of N in the environment (Han et al., 2023). Mild N deficiency induces BR biosynthesis (Jia et al., 2020) and also increases BR signalling through the accumulation and activation of BES1. Activated BES1 suppresses the expression and DNA binding ability of the LBD37/38/39 TFs, thus enhancing the expression of genes that stimulate foraging response under low N conditions (Chai et al., 2022; Figure 3).
5.4 Cytokinins
CKs are hormones that participate in the regulation of plant growth, nutritional signalling, and responses to abiotic and biotic stresses (Li et al., 2021b). CKs signalling participates in local and long-distance shoot–root communication, and the adjustment of plant architecture in response to C metabolism and external and internal N status. Both signals have been shown to induce CK synthesis and accumulation via the upregulation of IPT genes (Wang et al., 2021a; Abualia et al., 2023). On one hand, photosynthetically generated sugars induce the biosynthesis and transportation of CK precursors from the root to the shoot to activate CK signalling and stimulate plant growth (Kiba et al., 2019; Figure 3). In return, CKs can also influence C metabolism and many CK response mutants display altered sensitivity to sugars (Radchuk et al., 2023; Wang et al., 2021a). Also, CKs are implicated in sugar-induced anthocyanin accumulation in Arabidopsis leaves (Sakr et al., 2018). As mentioned, CKs function as a root-to-shoot long-distance N status signal that regulates both N-related metabolism and transport to adjust plant growth to N availability. In roots, NO3- supply induces CKs biosynthesis and translocation through the xylem via NLP7. Then, CK accumulation promotes shoot growth through increasing expression of auxin efflux carriers and induces a systemic downstream signal that regulates NO3- transporter expression in roots (Abualia et al., 2022; Landrein et al., 2018; Poitout et al., 2018; Figure 3). Consistently, numerous components of the CK signal transduction pathway, including type-A ARR as well as CRF TFs, have been identified as NO3--responsive genes (Abualia et al., 2023). Although CK-dependent coordinated regulation of C and N metabolism has already been observed (Reguera et al., 2013), the role of this phytohormone in the combined regulation of C/N balance in the plant needs to be elucidated.
5.5 Gibberellins
GAs and C metabolism interplay in the control of photomorphogenesis processes like cotyledon opening and the regulation of light-responsive gene expression (Feng et al., 2008). Modulation of GA levels in response to light and C metabolism extends to environmental CO2 (Ribeiro et al., 2012), circadian rhythms, as they track daily fluctuations in sugar levels (Paparelli et al., 2013), and the TCA cycle, with the intermediate 2-OG (2-oxoglutarate) controlling the rate of biosynthesis of GAs and amino acids (Araújo et al., 2014). GAs are also involved in the induction of the anthocyanin biosynthesis pathway, which is dependent on the sucrose stabilisation of DELLA proteins (Li et al., 2014). DELLAs are nuclear proteins lacking a DNA-binding domain that negatively regulate GA responses through interaction with other TFs and are targeted for degradation in the presence of GA and the GA receptor GID1 (Daviere & Achard, 2013). Regarding N metabolism, the only DELLA protein in rice (SLR1) functions as a pivotal factor in the integration of GA signalling pathways and N deficiency-induced adaptations on plant architecture. On one hand, OsGRF4 TF activity is inhibited by DELLAs, thus balancing N assimilation, C fixation and plant growth. Increasing the abundance of OsGRF4 tips the GRF-DELLA balance to favour GRF4, which increases growth and N-use efficiency (NUE; Li et al., 2018; Figure 3) On the other hand, DELLA accumulation increases the stability of OsNGR5, a protein that promotes rice tillering in the presence of N, by competitively binding to the gibberellin receptor GID1 (Wu et al., 2020; Figure 3). In addition, Camut et al. (2021) revealed that NO3- enhances the accumulation of bioactive GAs in both Arabidopsis and wheat, thus reducing the abundance of DELLAs and promoting plant growth. Finally, a few reports suggest the interaction of GA signalling with TOR and SnRK1. Zhang et al. (2018) revealed that disruption of the major RAPTOR isoform, RAPTOR1B, leads to a decrease in transcript levels of the GA receptor GID1. On the other hand, SnRK1 can repress GA biosynthesis through FUS3 activation (Curaba et al., 2004; Gazzarrini et al., 2004; Figure 3).
5.6 Other phytohormones
Several other phytohormones have been related to N and C pathways regulation but their direct contribution to the control of plant response to C/N balance has not been clearly established yet. For instance, high levels of strigolactones (SLs) in the root modify its architecture to increase their coverage in the surrounding soil and the root exudates that promote plant–microbe symbiotic interactions, which together improve mineral nutrients and water uptake (Barbier et al., 2023; Dun et al., 2023; Kelly et al., 2023; Sun et al., 2022). SLs are also transported through the xylem from roots to shoots, where they can act as a signal to coordinate shoot branching with the nutrient status of the roots (Dun et al., 2023; Kelly et al., 2023). Moreover, Sun et al. (2023) revealed a crosstalk between the SL and GA pathways in response to changes in N availability in rice. Under high N supply, the D53 and SLR1-DELLA proteins, both key repressors of the SL and GA signalling pathways, respectively, bind to the GRF4 TF, reducing the activation of N metabolism-associated genes. This regulation is reverted under N limitation by SL-mediated promoted degradation of the D53 and SLR1 proteins (Figure 3). SLs also mediate C metabolism and sugar signalling in early seedling development (Li et al., 2016) and shoot branching. In this respect, SL negative effect over tillering in rice involves the formation of D3-D14-D53 complex and subsequent degradation of D14 and D53. This mechanism is dependent on sucrose levels, which can alleviate SL-mediated inhibition of bud outgrowth through D3 repression and the triggering of a D14-dependent induction of D53 accumulation (Patil et al., 2022; Figure 3). This crosstalk and the role of MAX2 (D3 orthologue) have also been described in Arabidopsis (Bennett et al., 2006; Bertheloot et al., 2020).
Finally, genetic and phenotypic approaches have uncovered an antagonistic interplay between sugars and ethylene pathways (Sakr et al., 2018), where glucose sensors HXK1 and HKL1 have been described as direct regulators of glucose-dependent transcriptional repression of some ethylene biosynthesis genes. Meanwhile, HKL1 can also act as a positive effector of ethylene responses (Karve et al., 2012). EIN3 TF, whose stability is enhanced by ethylene and reduced by glucose, is also a key regulator of this interaction (Yanagisawa et al., 2003; Figure 3). The relationship between ethylene and C signalling has been described in developmental processes like the dark-to-light transition of seedlings (Brenya et al., 2023; Zhong et al., 2014), adaptation to high atmospheric CO2 (Smet et al., 2020) or the long-distance transport of sucrose from source leaves to sink organs (Tong et al., 2022). Interestingly, ethylene biosynthesis and signalling are negatively regulated by the master regulators TOR (Dong et al., 2015; Fu et al., 2021; Zhuo et al., 2020; Figure 3) and SnRK1 (Kim et al., 2017; Lumba et al., 2012). As with sugars, ethylene also plays an essential role in plant responses to mineral nutrient availability, including N. In summary, the exposition to sub or supra-optimal N status stimulates ethylene biosynthesis and signalling in the plant (Figure 3). Then, induced ethylene mediates a range of adaptive responses that will re-model root architecture and modulate the overall assimilation, translocation and accumulation of N (Khan et al., 2015; Ma et al., 2023b).
6 CONCLUSION
In this review, we have focused on aspects of the C and N metabolism and signalling that may be underlying mechanisms responsible for maintaining C/N balance in plants and provided a comprehensive, as well as reasonably summarized picture of their features. We still have numerous gaps in our understanding of the molecular and hormonal control of this important process, as well as their connection with plant development and stress responses. Results from these studies will open new avenues to improve crop yield by enhancing their ability to cope with nutrient imbalance and stress scenarios.
AUTHOR CONTRIBUTIONS
All authors participated in literature search, figure design and manuscript writing. IF-P drew the figures. M.P. and L.O-S planned and edited the manuscript. All the authors have read and approved the final version of the manuscript.
ACKNOWLEDGEMENTS
This work was supported with grants PID2019-109154RB-I00 and PID2022-142059OB-I00 funded by spanish MCIN/AEI/10.13039/501100011033 and MCIN/AEI/10.13039/501100011033/ FEDER, UE, respectively. IF-P is supported by a PhD contract (FPI) funded by grant SEV-2016-0672 to the CBGP (Centre of Excellence Severo Ochoa program of the Agencia Estatal de Investigación, Spain).
Table S1. List of complete gene names in alphabetical order according to their acronyms.
Please note: The publisher is not responsible for the content or functionality of any supporting information supplied by the authors. Any queries (other than missing content) should be directed to the corresponding author for the article.
BIBLIOGRAPHY
Abualia, R., Ötvös, K., Novák, O., Bouguyon, E., Domanegg, K., Krapp, A., Nacry, P., Gojon, A., Lacombe, B. & Benková, E. (2022) Molecular framework integrating nitrate sensing in root and auxin-guided shoot adaptive responses. Proceedings of the National Academy of Sciences, 119(31), e2122460119. https://doi.org/10.1073/pnas.2122460119.
Ahmad, N., Jiang, Z., Zhang, L., Hussain, I. & Yang, X. (2023) Insights on Phytohormonal Crosstalk in Plant Response to Nitrogen Stress: A Focus on Plant Root Growth and Development. International Journal of Molecular Sciences, 24(4).
Aoyama, S., Huarancca Reyes, T., Guglielminetti, L., Lu, Y., Morita, Y., Sato, T. & Yamaguchi, J. (2014) Ubiquitin ligase ATL31 functions in leaf senescence in response to the balance between atmospheric CO2 and nitrogen availability in Arabidopsis. Plant & cell physiology, 55(2), 293-305. https://doi.org/10.1093/pcp/pcu002.
Araújo, W. L., Martins, A. O., Fernie, A. R. & Tohge, T. (2014) 2-Oxoglutarate: linking TCA cycle function with amino acid, glucosinolate, flavonoid, alkaloid, and gibberellin biosynthesis. Frontiers in Plant Science, 5. https://doi.org/10.3389/fpls.2014.00552.
Ariga, T., Sakuraba, Y., Zhuo, M., Yang, M. & Yanagisawa, S. (2022) The Arabidopsis NLP7-HB52/54-VAR2 pathway modulates energy utilization in diverse light and nitrogen conditions. Current Biology, 32(24), 5344-5353.e6. https://doi.org/10.1016/j.cub.2022.10.024.
Artins, A. & Caldana, C. (2022) The metabolic homeostaTOR: The balance of holding on or letting grow. Current Opinion in Plant Biology, 66, 102196. https://doi.org/10.1016/j.pbi.2022.102196.
Artins, A., Martins, M. C. M., Meyer, C., Fernie, A. R. & Caldana, C. (2024) Sensing and regulation of C and N metabolism – novel features and mechanisms of the TOR and SnRK1 signaling pathways. The Plant Journal, n/a(n/a). https://doi.org/10.1111/tpj.16684.
Asim, M., Ullah, Z., Xu, F., An, L., Aluko, O. O., Wang, Q. & Liu, H. (2020) Nitrate Signaling, Functions, and Regulation of Root System Architecture: Insights from Arabidopsis thaliana. Genes, 11(6).
Athwal, G. S., Huber, J. L. & Huber, S. C. (1998) Phosphorylated Nitrate Reductase and 14-3-3 Proteins1: Site of Interaction, Effects of Ions, and Evidence for an AMP-Binding Site on 14-3-3 Proteins. Plant Physiology, 118(3), 1041-1048. https://doi.org/10.1104/pp.118.3.1041.
Bachmann, M., Huber, J. L., Liao, P.-C., Gage, D. A. & Huber, S. C. (1996) The inhibitor protein of phosphorylated nitrate reductase from spinach (Spinacia oleracea) leaves is a 14-3-3 protein. FEBS Letters, 387(2-3), 127-131. https://doi.org/10.1016/0014-5793(96)00478-4.
Baena-González, E. & Lunn, J. E. (2020) SnRK1 and trehalose 6-phosphate – two ancient pathways converge to regulate plant metabolism and growth. Current Opinion in Plant Biology, 55, 52-59. https://doi.org/10.1016/j.pbi.2020.01.010.
Baena-González, E., Rolland, F., Thevelein, J. M. & Sheen, J. (2007) A central integrator of transcription networks in plant stress and energy signalling. Nature, 448(7156), 938-942. https://doi.org/10.1038/nature06069.
Barbier, F., Fichtner, F. & Beveridge, C. (2023) The strigolactone pathway plays a crucial role in integrating metabolic and nutritional signals in plants. Nature Plants, 9(8), 1191-1200. https://doi.org/10.1038/s41477-023-01453-6.
Batistič, O., Waadt, R., Steinhorst, L., Held, K. & Kudla, J. (2010) CBL-mediated targeting of CIPKs facilitates the decoding of calcium signals emanating from distinct cellular stores. The Plant Journal, 61(2), 211-222. https://doi.org/10.1111/j.1365-313X.2009.04045.x.
Baud, S., Feria Bourrellier, A. B., Azzopardi, M., Berger, A., Dechorgnat, J., Daniel-Vedele, F., Lepiniec, L., Miquel, M., Rochat, C., Hodges, M. & Ferrario-Méry, S. (2010) PII is induced by WRINKLED1 and fine-tunes fatty acid composition in seeds of Arabidopsis thaliana. The Plant Journal, 64(2), 291-303. https://doi.org/10.1111/j.1365-313X.2010.04332.x.
Belda-Palazón, B., Adamo, M., Valerio, C., Ferreira, L. J., Confraria, A., Reis-Barata, D., Rodrigues, A., Meyer, C., Rodriguez, P. L. & Baena-González, E. (2020) A dual function of SnRK2 kinases in the regulation of SnRK1 and plant growth. Nature Plants, 6(11), 1345-1353. https://doi.org/10.1038/s41477-020-00778-w.
Belda-Palazón, B., Costa, M., Beeckman, T., Rolland, F. & Baena-González, E. (2022) ABA represses TOR and root meristem activity through nuclear exit of the SnRK1 kinase. Proceedings of the National Academy of Sciences, 119(28), e2204862119. https://doi.org/10.1073/pnas.2204862119.
Bennett, T., Sieberer, T., Willett, B., Booker, J., Luschnig, C. & Leyser, O. (2006) The Arabidopsis MAX Pathway Controls Shoot Branching by Regulating Auxin Transport. Current Biology, 16(6), 553-563. https://doi.org/10.1016/j.cub.2006.01.058.
Bertheloot, J., Barbier, F., Boudon, F., Perez-Garcia, M. D., Péron, T., Citerne, S., Dun, E., Beveridge, C., Godin, C. & Sakr, S. (2020) Sugar availability suppresses the auxin-induced strigolactone pathway to promote bud outgrowth. New Phytologist, 225(2), 866-879. https://doi.org/10.1111/nph.16201.
Bloom, A. J., Burger, M., Asensio, J. S. R. & Cousins, A. B. (2010) Carbon Dioxide Enrichment Inhibits Nitrate Assimilation in Wheat and Arabidopsis. Science, 328(5980), 899-903. https://doi.org/10.1126/science.1186440.
Brauner, K., Birami, B., Brauner, H. A. & Heyer, A. G. (2018) Diurnal periodicity of assimilate transport shapes resource allocation and whole-plant carbon balance. The Plant Journal, 94(5), 776-789. https://doi.org/10.1111/tpj.13898.
Brenya, E., Dutta, E., Herron, B., Walden, L. H., Roberts, D. M. & Binder, B. M. (2023) Ethylene-mediated metabolic priming increases photosynthesis and metabolism to enhance plant growth and stress tolerance. PNAS Nexus, 2(7), pgad216. https://doi.org/10.1093/pnasnexus/pgad216.
Bunney, T. D., van Walraven, H. S. & de Boer, A. H. (2001) 14-3-3 protein is a regulator of the mitochondrial and chloroplast ATP synthase. Proceedings of the National Academy of Sciences, 98(7), 4249-4254. https://doi.org/10.1073/pnas.061437498.
Busche, M., Scarpin, M. R., Hnasko, R. & Brunkard, J. O. (2021) TOR coordinates nucleotide availability with ribosome biogenesis in plants. The Plant Cell, 33(5), 1615-1632. https://doi.org/10.1093/plcell/koab043.
Camut, L., Gallova, B., Jilli, L., Sirlin-Josserand, M., Carrera, E., Sakvarelidze-Achard, L., Ruffel, S., Krouk, G., Thomas, S. G., Hedden, P., Phillips, A. L., Daviere, J. M. & Achard, P. (2021) Nitrate signaling promotes plant growth by upregulating gibberellin biosynthesis and destabilization of DELLA proteins. Curr Biol, 31(22), 4971-4982 e4. https://doi.org/10.1016/j.cub.2021.09.024.
Cao, P., Kim, S.-J., Xing, A., Schenck, C. A., Liu, L., Jiang, N., Wang, J., Last, R. L. & Brandizzi, F. (2019) Homeostasis of branched-chain amino acids is critical for the activity of TOR signaling in Arabidopsis. eLife, 8, e50747. https://doi.org/10.7554/eLife.50747.
Carianopol, C. S., Chan, A. L., Dong, S., Provart, N. J., Lumba, S. & Gazzarrini, S. (2020) An abscisic acid-responsive protein interaction network for sucrose non-fermenting related kinase1 in abiotic stress response. Communications Biology, 3(1), 145. https://doi.org/10.1038/s42003-020-0866-8.
Casanova-Sáez, R., Mateo-Bonmatí, E. & Ljung, K. (2021) Auxin Metabolism in Plants. Cold Spring Harbor Perspectives in Biology, 13(3). https://doi.org/10.1101/cshperspect.a039867.
Cerezo, M., Tillard, P., Filleur, S., Muños, S. p., Daniel-Vedele, F. o. & Gojon, A. (2001) Major Alterations of the Regulation of Root NO3 − Uptake Are Associated with the Mutation of Nrt2.1 and Nrt2.2 Genes in Arabidopsis. Plant Physiology, 127(1), 262-271. https://doi.org/10.1104/pp.127.1.262.
Chai, S., Chen, J., Yue, X., Li, C., Zhang, Q., de Dios, V. R., Yao, Y. & Tan, W. (2022) Interaction of BES1 and LBD37 transcription factors modulates brassinosteroid-regulated root forging response under low nitrogen in arabidopsis. Frontiers in Plant Science, 13. https://doi.org/10.3389/fpls.2022.998961.
Chan, A., Carianopol, C., Tsai, A. Y.-L., Varatharajah, K., Chiu, R. S. & Gazzarrini, S. (2017) SnRK1 phosphorylation of FUSCA3 positively regulates embryogenesis, seed yield, and plant growth at high temperature in Arabidopsis. Journal of Experimental Botany, 68(15), 4219-4231. https://doi.org/10.1093/jxb/erx233.
Chardon, F., Bedu, M., Calenge, F., Klemens, Patrick A. W., Spinner, L., Clement, G., Chietera, G., Léran, S., Ferrand, M., Lacombe, B., Loudet, O., Dinant, S., Bellini, C., Neuhaus, H. E., Daniel-Vedele, F. & Krapp, A. (2013) Leaf Fructose Content Is Controlled by the Vacuolar Transporter SWEET17 in Arabidopsis. Current Biology, 23(8), 697-702. https://doi.org/10.1016/j.cub.2013.03.021.
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. Proceedings of the National Academy of Sciences, 115(50), 12823-12828. https://doi.org/10.1073/pnas.1809526115.
Chen, K., Li, G.-J., Bressan, R. A., Song, C.-P., Zhu, J.-K. & Zhao, Y. (2020) Abscisic acid dynamics, signaling, and functions in plants. Journal of Integrative Plant Biology, 62(1), 25-54. https://doi.org/10.1111/jipb.12899.
Chen, L.-Q., Qu, X.-Q., Hou, B.-H., Sosso, D., Osorio, S., Fernie, A. R. & Frommer, W. B. (2012) Sucrose Efflux Mediated by SWEET Proteins as a Key Step for Phloem Transport. Science, 335(6065), 207-211. https://doi.org/10.1126/science.1213351.
Chen, L., Su, Z.-Z., Huang, L., Xia, F.-N., Qi, H., Xie, L.-J., Xiao, S. & Chen, Q.-F. (2017) The AMP-Activated Protein Kinase KIN10 Is Involved in the Regulation of Autophagy in Arabidopsis. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.01201.
Chen, Y. M., Ferrar, T. S., Lohmeir-Vogel, E., Morrice, N., Mizuno, Y., Berenger, B., Ng, K. K. S., Muench, D. G. & Moorhead, G. B. G. (2006) The PII Signal Transduction Protein of Arabidopsis thaliana Forms an Arginine-regulated Complex with Plastid N-Acetyl Glutamate Kinase. Journal of Biological Chemistry, 281(9), 5726-5733. https://doi.org/10.1074/jbc.M510945200.
Chiou, T.-J. & Bush, D. R. (1998) Sucrose is a signal molecule in assimilate partitioning. Proceedings of the National Academy of Sciences, 95(8), 4784-4788. https://doi.org/10.1073/pnas.95.8.4784.
Chiu, M. I., Katz, H. & Berlin, V. (1994) RAPT1, a mammalian homolog of yeast Tor, interacts with the FKBP12/rapamycin complex. Proceedings of the National Academy of Sciences, 91(26), 12574-12578. https://doi.org/10.1073/pnas.91.26.12574.
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. Journal of Experimental Botany, 67(9), 2745-2760. https://doi.org/10.1093/jxb/erw107.
Chopin, F., Orsel, M., Dorbe, M.-F., Chardon, F., Truong, H.-N., Miller, A. J., Krapp, A. & Daniel-Vedele, F. o. (2007) The Arabidopsis ATNRT2.7 Nitrate Transporter Controls Nitrate Content in Seeds. The Plant Cell, 19(5), 1590-1602. https://doi.org/10.1105/tpc.107.050542.
Comparot, S., Lingiah, G. & Martin, T. (2003) Function and specificity of 14-3-3 proteins in the regulation of carbohydrate and nitrogen metabolism. Journal of Experimental Botany, 54(382), 595-604. https://doi.org/10.1093/jxb/erg057.
Contreras-López, O., Vidal, E. A., Riveras, E., Alvarez, J. M., Moyano, T. C., Sparks, E. E., Medina, J., Pasquino, A., Benfey, P. N., Coruzzi, G. M. & Gutiérrez, R. A. (2022) Spatiotemporal analysis identifies ABF2 and ABF3 as key hubs of endodermal response to nitrate. Proceedings of the National Academy of Sciences, 119(4), e2107879119. https://doi.org/10.1073/pnas.2107879119.
Coruzzi, G. M. & Zhou, L. (2001) Carbon and nitrogen sensing and signaling in plants: emerging ‘matrix effects’. Current Opinion in Plant Biology, 4(3), 247-253. https://doi.org/10.1016/S1369-5266(00)00168-0.
Cotelle, V., Meek, S. E. M., Provan, F., Milne, F. C., Morrice, N. & MacKintosh, C. (2000) 14-3-3s regulate global cleavage of their diverse binding partners in sugar-starved Arabidopsis cells. The EMBO Journal, 19(12), 2869-2876-2876. https://doi.org/10.1093/emboj/19.12.2869.
Curaba, J., Moritz, T., Blervaque, R., Parcy, F., Raz, V., Herzog, M. & Vachon, G. (2004) AtGA3ox2, a key gene responsible for bioactive gibberellin biosynthesis, is regulated during embryogenesis by LEAFY COTYLEDON2 and FUSCA3 in Arabidopsis. Plant Physiol, 136(3), 3660-9.
Deng, K., Dong, P., Wang, W., Feng, L., Xiong, F., Wang, K., Zhang, S., Feng, S., Wang, B., Zhang, J. & Ren, M. (2017) The TOR Pathway Is Involved in Adventitious Root Formation in Arabidopsis and Potato. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.00784.
Deng, K., Yu, L., Zheng, X., Zhang, K., Wang, W., Dong, P., Zhang, J. & Ren, M. (2016) Target of Rapamycin Is a Key Player for Auxin Signaling Transduction in Arabidopsis. Frontiers in Plant Science, 7. https://doi.org/10.3389/fpls.2016.00291.
Diaz, C., Kusano, M., Sulpice, R., Araki, M., Redestig, H., Saito, K., Stitt, M. & Shin, R. (2011) Determining novel functions of Arabidopsis14-3-3 proteins in central metabolic processes. BMC Systems Biology, 5(1), 192. https://doi.org/10.1186/1752-0509-5-192.
Dong, P., Xiong, F., Que, Y., Wang, K., Yu, L., Li, Z. & Maozhi, R. (2015) Expression profiling and functional analysis reveals that TOR is a key player in regulating photosynthesis and phytohormone signaling pathways in Arabidopsis. Frontiers in Plant Science, 6. https://doi.org/10.3389/fpls.2015.00677.
Dun, E. A., Brewer, P. B., Gillam, E. M. J. & Beveridge, C. A. (2023) Strigolactones and Shoot Branching: What Is the Real Hormone and How Does It Work?Plant and Cell Physiology, 64(9), 967-983. https://doi.org/10.1093/pcp/pcad088.
Ehlert, A., Weltmeier, F., Wang, X., Mayer, C. S., Smeekens, S., Vicente-Carbajosa, J. & Dröge-Laser, W. (2006) Two-hybrid protein–protein interaction analysis in Arabidopsis protoplasts: establishment of a heterodimerization map of group C and group S bZIP transcription factors. The Plant Journal, 46(5), 890-900. https://doi.org/10.1111/j.1365-313X.2006.02731.x.
Fàbregas, N. & Fernie, A. R. (2021) The interface of central metabolism with hormone signaling in plants. Current Biology, 31(23), R1535-R1548. https://doi.org/10.1016/j.cub.2021.09.070.
Fàbregas, N. & Fernie, A. R. (2022) The reliance of phytohormone biosynthesis on primary metabolite precursors. Journal of Plant Physiology, 268, 153589. https://doi.org/10.1016/j.jplph.2021.153589.
Feng, S., Martinez, C., Gusmaroli, G., Wang, Y., Zhou, J., Wang, F., Chen, L., Yu, L., Iglesias-Pedraz, J. M., Kircher, S., Schafer, E., Fu, X., Fan, L. M. & Deng, X. W. (2008) Coordinated regulation of Arabidopsis thaliana development by light and gibberellins. Nature, 451(7177), 475-9.
Feria Bourrellier, A. B., Valot, B., Guillot, A., Ambard-Bretteville, F., Vidal, J. & Hodges, M. (2010) Chloroplast acetyl-CoA carboxylase activity is 2-oxoglutarate–regulated by interaction of PII with the biotin carboxyl carrier subunit. Proceedings of the National Academy of Sciences, 107(1), 502-507. https://doi.org/10.1073/pnas.0910097107.
Ferrario-Méry, S., Besin, E., Pichon, O., Meyer, C. & Hodges, M. (2006) The regulatory PII protein controls arginine biosynthesis in Arabidopsis. FEBS Letters, 580(8), 2015-2020. https://doi.org/10.1016/j.febslet.2006.02.075.
Ferrario-Méry, S., Bouvet, M., Leleu, O., Savino, G., Hodges, M. & Meyer, C. (2005) Physiological characterisation of Arabidopsis mutants affected in the expression of the putative regulatory protein PII. Planta, 223(1), 28-39. https://doi.org/10.1007/s00425-005-0063-5.
Fichtner, F. & Lunn, J. E. (2021a) The Role of Trehalose 6-Phosphate (Tre6P) in Plant Metabolism and Development. Annual Review of Plant Biology, 72(Volume 72, 2021), 737-760. https://doi.org/10.1146/annurev-arplant-050718-095929.
Fichtner, F., Dissanayake, I. M., Lacombe, B. & Barbier, F. (2021b) Sugar and Nitrate Sensing: A Multi-Billion-Year Story. Trends in Plant Science, 26(4), 352-374. https://doi.org/10.1016/j.tplants.2020.11.006.
Fichtner, F., Olas, J. J., Feil, R., Watanabe, M., Krause, U., Hoefgen, R., Stitt, M. & Lunn, J. E. (2020) Functional Features of TREHALOSE-6-PHOSPHATE SYNTHASE1, an Essential Enzyme in Arabidopsis[OPEN]. The Plant Cell, 32(6), 1949-1972. https://doi.org/10.1105/tpc.19.00837.
Figueroa, C. M., Feil, R., Ishihara, H., Watanabe, M., Kölling, K., Krause, U., Höhne, M., Encke, B., Plaxton, W. C., Zeeman, S. C., Li, Z., Schulze, W. X., Hoefgen, R., Stitt, M. & Lunn, J. E. (2016) Trehalose 6–phosphate coordinates organic and amino acid metabolism with carbon availability. The Plant Journal, 85(3), 410-423. https://doi.org/10.1111/tpj.13114.
Figueroa, C. M. & Lunn, J. E. (2016) A Tale of Two Sugars: Trehalose 6-Phosphate and Sucrose. Plant Physiology, 172(1), 7-27. https://doi.org/10.1104/pp.16.00417.
Gansel, X., Muños, S., Tillard, P. & Gojon, A. (2001) Differential regulation of the NO3– and NH4+ transporter genes AtNrt2.1 and AtAmt1.1 in Arabidopsis: relation with long-distance and local controls by N status of the plant. The Plant Journal, 26(2), 143-155. https://doi.org/10.1046/j.1365-313x.2001.01016.x.
Gao, P., Xin, Z. & Zheng, Z.-L. (2008) The OSU1/QUA2/TSD2-Encoded Putative Methyltransferase Is a Critical Modulator of Carbon and Nitrogen Nutrient Balance Response in Arabidopsis. PLOS ONE, 3(1), e1387. https://doi.org/10.1371/journal.pone.0001387.
Gautam, T., Dutta, M., Jaiswal, V., Zinta, G., Gahlaut, V. & Kumar, S. (2022) Emerging Roles of SWEET Sugar Transporters in Plant Development and Abiotic Stress Responses. Cells, 11(8).
Gazzarrini, S., Tsuchiya, Y., Lumba, S., Okamoto, M. & McCourt, P. (2004) The Transcription Factor FUSCA3 Controls Developmental Timing in Arabidopsis through the Hormones Gibberellin and Abscisic Acid. Developmental Cell, 7(3), 373-385. https://doi.org/10.1016/j.devcel.2004.06.017.
Gojon, A., Cassan, O., Bach, L., Lejay, L. & Martin, A. (2023) The decline of plant mineral nutrition under rising CO2: physiological and molecular aspects of a bad deal. Trends in Plant Science, 28(2), 185-198. https://doi.org/10.1016/j.tplants.2022.09.002.
González, A. & Hall, M. N. (2017) Nutrient sensing and TOR signaling in yeast and mammals. The EMBO Journal, 36(4), 397-408-408. https://doi.org/10.15252/embj.201696010.
Guo, H., Sun, Y., Li, Y., Liu, X., Ren, Q., Zhu-Salzman, K. & Ge, F. (2013) Elevated CO2 Modifies N Acquisition of Medicago truncatula by Enhancing N Fixation and Reducing Nitrate Uptake from Soil. PLOS ONE, 8(12), e81373. https://doi.org/10.1371/journal.pone.0081373.
Guo, W.-J., Nagy, R., Chen, H.-Y., Pfrunder, S., Yu, Y.-C., Santelia, D., Frommer, W. B. & Martinoia, E. (2014) SWEET17, a Facilitative Transporter, Mediates Fructose Transport across the Tonoplast of Arabidopsis Roots and LeavesPlant Physiology, 164(2), 777-789. https://doi.org/10.1104/pp.113.232751.
Gutiérrez, R. A., Lejay, L. V., Dean, A., Chiaromonte, F., Shasha, D. E. & Coruzzi, G. M. (2007) Qualitative network models and genome-wide expression data define carbon/nitrogen-responsive molecular machines in Arabidopsis. Genome Biology, 8(1), R7. https://doi.org/10.1186/gb-2007-8-1-r7.
Han, C., Wang, L., Lyu, J., Shi, W., Yao, L., Fan, M. & Bai, M.-Y. (2023) Brassinosteroid signaling and molecular crosstalk with nutrients in plants. Journal of Genetics and Genomics, 50(8), 541-553. https://doi.org/10.1016/j.jgg.2023.03.004.
Hanson, J., Hanssen, M., Wiese, A., Hendriks, M. M. W. B. & Smeekens, S. (2008) The sucrose regulated transcription factor bZIP11 affects amino acid metabolism by regulating the expression of ASPARAGINE SYNTHETASE1 and PROLINE DEHYDROGENASE2. The Plant Journal, 53(6), 935-949. https://doi.org/10.1111/j.1365-313X.2007.03385.x.
Hao, Y., Wang, H., Qiao, S., Leng, L. & Wang, X. (2016) Histone deacetylase HDA6 enhances brassinosteroid signaling by inhibiting the BIN2 kinase. Proceedings of the National Academy of Sciences, 113(37), 10418-10423. https://doi.org/10.1073/pnas.1521363113.
Harthill, J. E., Meek, S. E. M., Morrice, N., Peggie, M. W., Borch, J., Wong, B. H. C. & MacKintosh, C. (2006) Phosphorylation and 14-3-3 binding of Arabidopsis trehalose-phosphate synthase 5 in response to 2-deoxyglucose. The Plant Journal, 47(2), 211-223. https://doi.org/10.1111/j.1365-313X.2006.02780.x.
Haruta, M., Sabat, G., Stecker, K., Minkoff, B. B. & Sussman, M. R. (2014) A Peptide Hormone and Its Receptor Protein Kinase Regulate Plant Cell Expansion. Science, 343(6169), 408-411. https://doi.org/10.1126/science.1244454.
Hashimoto, M., Negi, J., Young, J., Israelsson, M., Schroeder, J. I. & Iba, K. (2006) Arabidopsis HT1 kinase controls stomatal movements in response to CO2. Nature Cell Biology, 8(4), 391-397. https://doi.org/10.1038/ncb1387.
Henninger, M., Pedrotti, L., Krischke, M., Draken, J., Wildenhain, T., Fekete, A., Rolland, F., Müller, M. J., Fröschel, C., Weiste, C. & Dröge-Laser, W. (2022) The evolutionarily conserved kinase SnRK1 orchestrates resource mobilization during Arabidopsis seedling establishment. The Plant Cell, 34(1), 616-632. https://doi.org/10.1093/plcell/koab270.
Henriques, R., Magyar, Z. & Bögre, L. (2013) S6K1 and E2FB are in mutually antagonistic regulatory links controlling cell growth and proliferation in Arabidopsis. Plant Signaling & Behavior, 8(6), e24367. https://doi.org/10.4161/psb.24367.
Hiyama, A., Takemiya, A., Munemasa, S., Okuma, E., Sugiyama, N., Tada, Y., Murata, Y. & Shimazaki, K.-i. (2017) Blue light and CO2 signals converge to regulate light-induced stomatal opening. Nature Communications, 8(1), 1284. https://doi.org/10.1038/s41467-017-01237-5.
Hsieh, M.-H., Lam, H.-M., van de Loo, F. J. & Coruzzi, G. (1998) A PII-like protein in Arabidopsis: Putative role in nitrogen sensing. Proceedings of the National Academy of Sciences, 95(23), 13965-13970. https://doi.org/10.1073/pnas.95.23.13965.
Hu, Z., Tang, Z., Zhang, Y., Niu, L., Yang, F., Zhang, D. & Hu, Y. (2021) Rice SUT and SWEET Transporters. International Journal of Molecular Sciences, 22(20).
Jahn, T., Fuglsang, A. T., Olsson, A., Brüntrup, I. M., Collinge, D. B., Volkmann, D., Sommarin, M., Palmgren, M. G. & Larsson, C. (1997) The 14-3-3 protein interacts directly with the C-terminal region of the plant plasma membrane H+-ATPase. The Plant Cell, 9(10), 1805-1814. https://doi.org/10.1105/tpc.9.10.1805.
Jamsheer K. M., Jindal, S. & Laxmi, A. (2019) Evolution of TOR–SnRK dynamics in green plants and its integration with phytohormone signaling networks. Journal of Experimental Botany, 70(8), 2239-2259. https://doi.org/10.1093/jxb/erz107.
Jaspert, N., Throm, C. & Oecking, C. (2011) Arabidopsis 14-3-3 Proteins: Fascinating and Less Fascinating Aspects. Frontiers in Plant Science, 2(96). https://doi.org/10.3389/fpls.2011.00096.
Jeena, G. S., Kumar, S. & Shukla, R. K. (2019) Structure, evolution and diverse physiological roles of SWEET sugar transporters in plants. Plant Molecular Biology, 100(4), 351-365. https://doi.org/10.1007/s11103-019-00872-4.
Jia, Z., Giehl, R. F. H. & von Wirén, N. (2020) The Root Foraging Response under Low Nitrogen Depends on DWARF1-Mediated Brassinosteroid Biosynthesis1. Plant Physiology, 183(3), 998-1010. https://doi.org/10.1104/pp.20.00440.
Jossier, M., Bouly, J.-P., Meimoun, P., Arjmand, A., Lessard, P., Hawley, S., Grahame Hardie, D. & Thomas, M. (2009) SnRK1 (SNF1-related kinase 1) has a central role in sugar and ABA signalling in Arabidopsis thaliana. The Plant Journal, 59(2), 316-328. https://doi.org/10.1111/j.1365-313X.2009.03871.x.
Kang, J., Mehta, S. & Turano, F. J. (2004) The putative glutamate receptor 1.1 (AtGLR1.1) in Arabidopsis thaliana Regulates Abscisic Acid Biosynthesis and Signaling to Control Development and Water Loss. Plant and Cell Physiology, 45(10), 1380-1389. https://doi.org/10.1093/pcp/pch159.
Kang, J. & Turano, F. J. (2003) The putative glutamate receptor 1.1 (AtGLR1.1) functions as a regulator of carbon and nitrogen metabolism in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, 100(11), 6872. https://doi.org/10.1073/pnas.1030961100.
Kanstrup, C. & Nour-Eldin, H. H. (2022) The emerging role of the nitrate and peptide transporter family: NPF in plant specialized metabolism. Current Opinion in Plant Biology, 68, 102243. https://doi.org/10.1016/j.pbi.2022.102243.
Karve, A., Xia, X. & Moore, B. d. (2012) Arabidopsis Hexokinase-Like1 and Hexokinase1 Form a Critical Node in Mediating Plant Glucose and Ethylene ResponsesPlant Physiology, 158(4), 1965-1975. https://doi.org/10.1104/pp.112.195636.
Kelly, G., David-Schwartz, R., Sade, N., Moshelion, M., Levi, A., Alchanatis, V. & Granot, D. (2012) The Pitfalls of Transgenic Selection and New Roles of AtHXK1: A High Level of AtHXK1 Expression Uncouples Hexokinase1-Dependent Sugar Signaling from Exogenous SugarPlant Physiology, 159(1), 47-51. https://doi.org/10.1104/pp.112.196105.
Kelly, G., Moshelion, M., David-Schwartz, R., Halperin, O., Wallach, R., Attia, Z., Belausov, E. & Granot, D. (2013) Hexokinase mediates stomatal closure. The Plant Journal, 75(6), 977-988. https://doi.org/10.1111/tpj.12258.
Kelly, J. H., Tucker, M. R. & Brewer, P. B. (2023) The Strigolactone Pathway Is a Target for Modifying Crop Shoot Architecture and Yield. Biology, 12(1).
Khan, M. I. R., Trivellini, A., Fatma, M., Masood, A., Francini, A., Iqbal, N., Ferrante, A. & Khan, N. A. (2015) Role of ethylene in responses of plants to nitrogen availability. Frontiers in Plant Science, 6. https://doi.org/10.3389/fpls.2015.00927.
Kiba, T., Takebayashi, Y., Kojima, M. & Sakakibara, H. (2019) Sugar-induced de novo cytokinin biosynthesis contributes to Arabidopsis growth under elevated CO2. Scientific Reports, 9(1), 7765. https://doi.org/10.1038/s41598-019-44185-4.
Kim, S., Kang, J.-y., Cho, D.-I., Park, J. H. & Kim, S. Y. (2004) ABF2, an ABRE-binding bZIP factor, is an essential component of glucose signaling and its overexpression affects multiple stress tolerance. The Plant Journal, 40(1), 75-87. https://doi.org/10.1111/j.1365-313X.2004.02192.x.
Klemens, P. A. W., Patzke, K., Deitmer, J., Spinner, L., Le Hir, R., Bellini, C., Bedu, M., Chardon, F., Krapp, A. & Neuhaus, H. E. (2013) Overexpression of the Vacuolar Sugar Carrier AtSWEET16 Modifies Germination, Growth, and Stress Tolerance in ArabidopsisPlant Physiology, 163(3), 1338-1352. https://doi.org/10.1104/pp.113.224972.
Kong, Y., Chen, S., Yang, Y. & An, C. (2013) ABA-insensitive (ABI) 4 and ABI5 synergistically regulate DGAT1 expression in Arabidopsis seedlings under stress. FEBS Letters, 587(18), 3076-3082. https://doi.org/10.1016/j.febslet.2013.07.045.
Konishi, M. & Yanagisawa, S. (2013) Arabidopsis NIN-like transcription factors have a central role in nitrate signalling. Nature Communications, 4(1), 1617. https://doi.org/10.1038/ncomms2621.
Krämer, K., Brock, J. & Heyer, A. G. (2022) Interaction of Nitrate Assimilation and Photorespiration at Elevated CO2. Frontiers in Plant Science, 13. https://doi.org/10.3389/fpls.2022.897924.
Kreisz, P., Hellens, A. M., Fröschel, C., Krischke, M., Maag, D., Feil, R., Wildenhain, T., Draken, J., Braune, G., Erdelitsch, L., Cecchino, L., Wagner, T. C., Ache, P., Mueller, M. J., Becker, D., Lunn, J. E., Hanson, J., Beveridge, C. A., Fichtner, F., Barbier, F. F. & Weiste, C. (2024) S1 basic leucine zipper transcription factors shape plant architecture by controlling C/N partitioning to apical and lateral organs. Proceedings of the National Academy of Sciences, 121(7), e2313343121. https://doi.org/10.1073/pnas.2313343121.
Krouk, G., Lacombe, B., Bielach, A., Perrine-Walker, F., Malinska, K., Mounier, E., Hoyerova, K., Tillard, P., Leon, S., Ljung, K., Zazimalova, E., Benkova, E., Nacry, P. & Gojon, A. (2010) Nitrate-Regulated Auxin Transport by NRT1.1 Defines a Mechanism for Nutrient Sensing in Plants. Developmental Cell, 18(6), 927-937. https://doi.org/10.1016/j.devcel.2010.05.008.
Kulma, A., Villadsen, D., Campbell, D. G., Meek, S. E. M., E. Harthill, J., Nielsen, T. H. & MacKintosh, C. (2004) Phosphorylation and 14-3-3 binding of Arabidopsis 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase. The Plant Journal, 37(5), 654-667. https://doi.org/10.1111/j.1365-313X.2003.01992.x.
Kumar, N., Caldwell, C. & Iyer-Pascuzzi, A. S. (2023) The NIN-LIKE PROTEIN 7 transcription factor modulates auxin pathways to regulate root cap development in Arabidopsis. Journal of Experimental Botany, 74(10), 3047-3059. https://doi.org/10.1093/jxb/erad058.
Lalonde, S., Wipf, D. & Frommer, W. B. (2004) TRANSPORT MECHANISMS FOR ORGANIC FORMS OF CARBON AND NITROGEN BETWEEN SOURCE AND SINK. Annual Review of Plant Biology, 55(Volume 55, 2004), 341-372. https://doi.org/10.1146/annurev.arplant.55.031903.141758.
Landrein, B., Formosa-Jordan, P., Malivert, A., Schuster, C., Melnyk, C. W., Yang, W., Turnbull, C., Meyerowitz, E. M., Locke, J. C. W. & Jönsson, H. (2018) Nitrate modulates stem cell dynamics in Arabidopsis shoot meristems through cytokinins. Proceedings of the National Academy of Sciences, 115(6), 1382-1387. https://doi.org/10.1073/pnas.1718670115.
Lejay, L., Gansel, X., Cerezo, M., Tillard, P., Müller, C., Krapp, A., von Wirén, N., Daniel-Vedele, F. o. & Gojon, A. (2003) Regulation of Root Ion Transporters by Photosynthesis: Functional Importance and Relation with Hexokinase. The Plant Cell, 15(9), 2218-2232. https://doi.org/10.1105/tpc.013516.
Lejay, L., Tillard, P., Lepetit, M., Olive, F. D., Filleur, S., Daniel-Vedele, F. & Gojon, A. (1999) Molecular and functional regulation of two NO3– uptake systems by N- and C-status of Arabidopsis plants. The Plant Journal, 18(5), 509-519. https://doi.org/10.1046/j.1365-313X.1999.00480.x.
Léran, S., Edel, K. H., Pervent, M., Hashimoto, K., Corratgé-Faillie, C., Offenborn, J. N., Tillard, P., Gojon, A., Kudla, J. & Lacombe, B. (2015) Nitrate sensing and uptake in Arabidopsis are enhanced by ABI2, a phosphatase inactivated by the stress hormone abscisic acid. Science Signaling, 8(375), ra43-ra43. https://doi.org/10.1126/scisignal.aaa4829.
Li, G. D., Pan, L. N., Jiang, K., Takahashi, I., Nakamura, H., Xu, Y. W., Asami, T. & Shen, R. F. (2016) Strigolactones are involved in sugar signaling to modulate early seedling development in Arabidopsis. Plant Biotechnology, 33(2), 87-97. https://doi.org/10.5511/plantbiotechnology.16.0326a.
Li, P., Ainsworth, E. A., Leakey, A. D. B., Ulanov, A., Lozovaya, V., Ort, D. R. & Bohnert, H. J. (2008) Arabidopsis transcript and metabolite profiles: ecotype-specific responses to open-air elevated [CO2]. Plant, Cell & Environment, 31(11), 1673-1687. https://doi.org/10.1111/j.1365-3040.2008.01874.x.
Li, P., Wind, J. J., Shi, X., Zhang, H., Hanson, J., Smeekens, S. C. & Teng, S. (2011) Fructose sensitivity is suppressed in Arabidopsis by the transcription factor ANAC089 lacking the membrane-bound domain. Proceedings of the National Academy of Sciences, 108(8), 3436-3441. https://doi.org/10.1073/pnas.1018665108.
Li, X., Cai, W., Liu, Y., Li, H., Fu, L., Liu, Z., Xu, L., Liu, H., Xu, T. & Xiong, Y. (2017) Differential TOR activation and cell proliferation in Arabidopsis root and shoot apexes. Proceedings of the National Academy of Sciences, 114(10), 2765-2770. https://doi.org/10.1073/pnas.1618782114.
Li, X., Sanagi, M., Lu, Y., Nomura, Y., Stolze, S. C., Yasuda, S., Saijo, Y., Schulze, W. X., Feil, R., Stitt, M., Lunn, J. E., Nakagami, H., Sato, T. & Yamaguchi, J. (2020b) Protein Phosphorylation Dynamics Under Carbon/Nitrogen-Nutrient Stress and Identification of a Cell Death-Related Receptor-Like Kinase in Arabidopsis. Frontiers in Plant Science, 11. https://doi.org/10.3389/fpls.2020.00377.
Li, Y., Van den Ende, W. & Rolland, F. (2014) Sucrose Induction of Anthocyanin Biosynthesis Is Mediated by DELLA. Molecular Plant, 7(3), 570-572. https://doi.org/10.1093/mp/sst161.
Liao, C.-Y., Pu, Y., Nolan, T. M., Montes, C., Guo, H., Walley, J. W., Yin, Y. & Bassham, D. C. (2023) Brassinosteroids modulate autophagy through phosphorylation of RAPTOR1B by the GSK3-like kinase BIN2 in Arabidopsis. Autophagy, 19(4), 1293-1310. https://doi.org/10.1080/15548627.2022.2124501.
Lilley, J. L. S., Gee, C. W., Sairanen, I., Ljung, K. & Nemhauser, J. L. (2012) An Endogenous Carbon-Sensing Pathway Triggers Increased Auxin Flux and Hypocotyl ElongationPlant Physiology, 160(4), 2261-2270. https://doi.org/10.1104/pp.112.205575.
Lima, L., Seabra, A., Melo, P., Cullimore, J. & Carvalho, H. (2006) Phosphorylation and subsequent interaction with 14-3-3 proteins regulate plastid glutamine synthetase in Medicago truncatula. Planta, 223(3), 558-567. https://doi.org/10.1007/s00425-005-0097-8.
Little Daniel, Y., Rao, H., Oliva, S., Daniel-Vedele, F., Krapp, A. & Malamy Jocelyn, E. (2005) The putative high-affinity nitrate transporter NRT2.1 represses lateral root initiation in response to nutritional cues. Proceedings of the National Academy of Sciences, 102(38), 13693-13698. https://doi.org/10.1073/pnas.0504219102.
Liu, K.-H., Huang, C.-Y. & Tsay, Y.-F. (1999) CHL1 Is a Dual-Affinity Nitrate Transporter of Arabidopsis Involved in Multiple Phases of Nitrate Uptake. The Plant Cell, 11(5), 865-874. https://doi.org/10.1105/tpc.11.5.865.
Liu, K.-H., Liu, M., Lin, Z., Wang, Z.-F., Chen, B., Liu, C., Guo, A., Konishi, M., Yanagisawa, S., Wagner, G. & Sheen, J. (2022a) NIN-like protein 7 transcription factor is a plant nitrate sensor. Science, 377(6613), 1419-1425. https://doi.org/10.1126/science.add1104.
Liu, K.-h., Niu, Y., Konishi, M., Wu, Y., Du, H., Sun Chung, H., Li, L., Boudsocq, M., McCormack, M., Maekawa, S., Ishida, T., Zhang, C., Shokat, K., Yanagisawa, S. & Sheen, J. (2017) Discovery of nitrate–CPK–NLP signalling in central nutrient–growth networks. Nature, 545(7654), 311-316. https://doi.org/10.1038/nature22077.
Liu, K. H. & Tsay, Y. F. (2003) Switching between the two action modes of the dual-affinity nitrate transporter CHL1 by phosphorylation. The EMBO Journal, 22(5), 1005-1013-1013. https://doi.org/10.1093/emboj/cdg118.
Liu, X., Hu, B. & Chu, C. (2022b) Nitrogen assimilation in plants: current status and future prospects. Journal of Genetics and Genomics, 49(5), 394-404. https://doi.org/10.1016/j.jgg.2021.12.006.
Lopes, M. S. & Araus, J. L. (2008) Comparative genomic and physiological analysis of nutrient response to NH4+, NH4+/NO3- and NO3- in barley seedlings. Physiologia Plantarum, 134(1), 134-150. https://doi.org/10.1111/j.1399-3054.2008.01114.x.
Loreti, E., Alpi, A. & Perata, P. (2000) Glucose and Disaccharide-Sensing Mechanisms Modulate the Expression of α-amylase in Barley Embryos1. Plant Physiology, 123(3), 939-948. https://doi.org/10.1104/pp.123.3.939.
Lu, Y., Sasaki, Y., Li, X., Mori, I. C., Matsuura, T., Hirayama, T., Sato, T. & Yamaguchi, J. (2015a) ABI1 regulates carbon/nitrogen-nutrient signal transduction independent of ABA biosynthesis and canonical ABA signalling pathways in Arabidopsis. Journal of Experimental Botany, 66(9), 2763-2771. https://doi.org/10.1093/jxb/erv086.
Lu, Y., Yamaguchi, J. & Sato, T. (2015b) Integration of C/N-nutrient and multiple environmental signals into the ABA signaling cascade. Plant Signaling & Behavior, 10(12), e1048940. https://doi.org/10.1080/15592324.2015.1048940.
Ma, B., Ma, T., Xian, W., Hu, B. & Chu, C. (2023b) Interplay between ethylene and nitrogen nutrition: How ethylene orchestrates nitrogen responses in plants. Journal of Integrative Plant Biology, 65(2), 399-407. https://doi.org/10.1111/jipb.13355.
Maghiaoui, A., Bouguyon, E., Cuesta, C., Perrine-Walker, F., Alcon, C., Krouk, G., Benková, E., Nacry, P., Gojon, A. & Bach, L. (2020) The Arabidopsis NRT1.1 transceptor coordinately controls auxin biosynthesis and transport to regulate root branching in response to nitrate. Journal of Experimental Botany, 71(15), 4480-4494. https://doi.org/10.1093/jxb/eraa242.
Mair, A., Pedrotti, L., Wurzinger, B., Anrather, D., Simeunovic, A., Weiste, C., Valerio, C., Dietrich, K., Kirchler, T., Nägele, T., Vicente Carbajosa, J., Hanson, J., Baena-González, E., Chaban, C., Weckwerth, W., Dröge-Laser, W. & Teige, M. (2015) SnRK1-triggered switch of bZIP63 dimerization mediates the low-energy response in plants. eLife, 4, e05828. https://doi.org/10.7554/eLife.05828.
Mallén-Ponce, M. J., Pérez-Pérez, M. E. & Crespo, J. L. (2022) Photosynthetic assimilation of CO2 regulates TOR activity. Proceedings of the National Academy of Sciences, 119(2), e2115261119. https://doi.org/10.1073/pnas.2115261119.
Marchive, C., Roudier, F., Castaings, L., Bréhaut, V., Blondet, E., Colot, V., Meyer, C. & Krapp, A. (2013) Nuclear retention of the transcription factor NLP7 orchestrates the early response to nitrate in plants. Nature Communications, 4(1), 1713. https://doi.org/10.1038/ncomms2650.
Martin, T., Oswald, O. & Graham, I. A. (2002) Arabidopsis Seedling Growth, Storage Lipid Mobilization, and Photosynthetic Gene Expression Are Regulated by Carbon:Nitrogen Availability. Plant Physiology, 128(2), 472. https://doi.org/10.1104/pp.010475.
Mathan, J., Singh, A. & Ranjan, A. (2021) Sucrose transport in response to drought and salt stress involves ABA-mediated induction of OsSWEET13 and OsSWEET15 in rice. Physiologia Plantarum, 171(4), 620-637. https://doi.org/10.1111/ppl.13210.
Matt, P., Geiger, M., Walch-Liu, P., Engels, C., Krapp, A. & Stitt, M. (2001) The immediate cause of the diurnal changes of nitrogen metabolism in leaves of nitrate-replete tobacco: a major imbalance between the rate of nitrate reduction and the rates of nitrate uptake and ammonium metabolism during the first part of the light period. Plant, Cell & Environment, 24(2), 177-190. https://doi.org/10.1111/j.1365-3040.2001.00676.x.
McClain, A. M. & Sharkey, T. D. (2019) Triose phosphate utilization and beyond: from photosynthesis to end product synthesis. Journal of Experimental Botany, 70(6), 1755-1766. https://doi.org/10.1093/jxb/erz058.
McCready, K., Spencer, V. & Kim, M. (2020) The Importance of TOR Kinase in Plant Development. Frontiers in Plant Science, 11. https://doi.org/10.3389/fpls.2020.00016.
Menand, B., Desnos, T., Nussaume, L., Berger, F., Bouchez, D., Meyer, C. & Robaglia, C. (2002) Expression and disruption of the Arabidopsis TOR (target of rapamycin) gene. Proceedings of the National Academy of Sciences, 99(9), 6422-6427. https://doi.org/10.1073/pnas.092141899.
Miao, H., Sun, P., Liu, Q., Miao, Y., Liu, J., Zhang, K., Hu, W., Zhang, J., Wang, J., Wang, Z., Jia, C., Xu, B. & Jin, Z. (2017) Genome-wide analyses of SWEET family proteins reveal involvement in fruit development and abiotic/biotic stress responses in banana. Scientific Reports, 7(1), 3536. https://doi.org/10.1038/s41598-017-03872-w.
Michniewicz, M., Ho, C.-H., Enders, T. A., Floro, E., Damodaran, S., Gunther, L. K., Powers, S. K., Frick, E. M., Topp, C. N., Frommer, W. B. & Strader, L. C. (2019) TRANSPORTER OF IBA1 Links Auxin and Cytokinin to Influence Root Architecture. Developmental Cell, 50(5), 599-609.e4. https://doi.org/10.1016/j.devcel.2019.06.010.
Miller, A. J., Fan, X., Shen, Q. & Smith, S. J. (2008) Amino acids and nitrate as signals for the regulation of nitrogen acquisition. Journal of Experimental Botany, 59(1), 111-119. https://doi.org/10.1093/jxb/erm208.
Mishra, B. S., Sharma, M. & Laxmi, A. (2022) Role of sugar and auxin crosstalk in plant growth and development. Physiologia Plantarum, 174(1), e13546. https://doi.org/10.1111/ppl.13546.
Moore, B., Zhou, L., Rolland, F., Hall, Q., Cheng, W.-H., Liu, Y.-X., Hwang, I., Jones, T. & Sheen, J. (2003) Role of the Arabidopsis Glucose Sensor HXK1 in Nutrient, Light, and Hormonal Signaling. Science, 300(5617), 332-336. https://doi.org/10.1126/science.1080585.
Moorhead, G., Douglas, P., Cotelle, V., Harthill, J., Morrice, N., Meek, S., Deiting, U., Stitt, M., Scarabel, M., Aitken, A. & MacKintosh, C. (1999) Phosphorylation-dependent interactions between enzymes of plant metabolism and 14-3-3 proteins. The Plant Journal, 18(1), 1-12. https://doi.org/10.1046/j.1365-313X.1999.00417.x.
Mounier, E., Pervent, M., Ljung, K., Gojon, A. & Nacry, P. (2014) Auxin-mediated nitrate signalling by NRT1.1 participates in the adaptive response of rabidopsis root architecture to the spatial heterogeneity of nitrate availability. Plant, Cell & Environment, 37(1), 162-174. https://doi.org/10.1111/pce.12143.
Mugume, Y., Kazibwe, Z. & Bassham, D. C. (2020) Target of Rapamycin in Control of Autophagy: Puppet Master and Signal Integrator. International Journal of Molecular Sciences, 21(21).
Muralidhara, P., Weiste, C., Collani, S., Krischke, M., Kreisz, P., Draken, J., Feil, R., Mair, A., Teige, M., Müller, M. J., Schmid, M., Becker, D., Lunn, J. E., Rolland, F., Hanson, J. & Dröge-Laser, W. (2021) Perturbations in plant energy homeostasis prime lateral root initiation via SnRK1-bZIP63-ARF19 signaling. Proceedings of the National Academy of Sciences, 118(37), e2106961118. https://doi.org/10.1073/pnas.2106961118.
Nolan, T. M., Brennan, B., Yang, M., Chen, J., Zhang, M., Li, Z., Wang, X., Bassham, D. C., Walley, J. & Yin, Y. (2017) Selective Autophagy of BES1 Mediated by DSK2 Balances Plant Growth and Survival. Developmental Cell, 41(1), 33-46.e7. https://doi.org/10.1016/j.devcel.2017.03.013.
Nolan, T. M., Vukašinović, N., Liu, D., Russinova, E. & Yin, Y. (2020) Brassinosteroids: Multidimensional Regulators of Plant Growth, Development, and Stress Responses. The Plant Cell, 32(2), 295. https://doi.org/10.1105/tpc.19.00335.
Nukarinen, E., Nägele, T., Pedrotti, L., Wurzinger, B., Mair, A., Landgraf, R., Börnke, F., Hanson, J., Teige, M., Baena-Gonzalez, E., Dröge-Laser, W. & Weckwerth, W. (2016) Quantitative phosphoproteomics reveals the role of the AMPK plant ortholog SnRK1 as a metabolic master regulator under energy deprivation. Scientific Reports, 6(1), 31697. https://doi.org/10.1038/srep31697.
Nunes-Nesi, A., Fernie, A. R. & Stitt, M. (2010) Metabolic and Signaling Aspects Underpinning the Regulation of Plant Carbon Nitrogen Interactions. Molecular Plant, 3(6), 973-996. https://doi.org/10.1093/mp/ssq049.
O'Brien, José A., Vega, A., Bouguyon, E., Krouk, G., Gojon, A., Coruzzi, G. & Gutiérrez, Rodrigo A. (2016) Nitrate Transport, Sensing, and Responses in Plants. Molecular Plant, 9(6), 837-856. https://doi.org/10.1016/j.molp.2016.05.004.
Olas, J. J., Van Dingenen, J., Abel, C., Działo, M. A., Feil, R., Krapp, A., Schlereth, A. & Wahl, V. (2019) Nitrate acts at the Arabidopsis thaliana shoot apical meristem to regulate flowering time. New Phytologist, 223(2), 814-827. https://doi.org/10.1111/nph.15812.
Ondzighi-Assoume, C. A., Chakraborty, S. & Harris, J. M. (2016) Environmental Nitrate Stimulates Abscisic Acid Accumulation in Arabidopsis Root Tips by Releasing It from Inactive Stores. The Plant Cell, 28(3), 729-745. https://doi.org/10.1105/tpc.15.00946.
Paparelli, E., Parlanti, S., Gonzali, S., Novi, G., Mariotti, L., Ceccarelli, N., van Dongen, J. T., Kölling, K., Zeeman, S. C. & Perata, P. (2013) Nighttime Sugar Starvation Orchestrates Gibberellin Biosynthesis and Plant Growth in ArabidopsisThe Plant Cell, 25(10), 3760-3769. https://doi.org/10.1105/tpc.113.115519.
Parwez, R., Aftab, T., Gill, S. S. & Naeem, M. (2022) Abscisic acid signaling and crosstalk with phytohormones in regulation of environmental stress responses. Environmental and Experimental Botany, 199, 104885. https://doi.org/10.1016/j.envexpbot.2022.104885.
Patil, S. B., Barbier, F. F., Zhao, J., Zafar, S. A., Uzair, M., Sun, Y., Fang, J., Perez-Garcia, M.-D., Bertheloot, J., Sakr, S., Fichtner, F., Chabikwa, T. G., Yuan, S., Beveridge, C. A. & Li, X. (2022) Sucrose promotes D53 accumulation and tillering in rice. New Phytologist, 234(1), 122-136. https://doi.org/10.1111/nph.17834.
Pedrotti, L., Weiste, C., Nägele, T., Wolf, E., Lorenzin, F., Dietrich, K., Mair, A., Weckwerth, W., Teige, M., Baena-González, E. & Dröge-Laser, W. (2018) Snf1-RELATED KINASE1-Controlled C/S1-bZIP Signaling Activates Alternative Mitochondrial Metabolic Pathways to Ensure Plant Survival in Extended Darkness. The Plant Cell, 30(2), 495-509. https://doi.org/10.1105/tpc.17.00414.
Peixoto, B. & Baena-González, E. (2022) Management of plant central metabolism by SnRK1 protein kinases. Journal of Experimental Botany, 73(20), 7068-7082. https://doi.org/10.1093/jxb/erac261.
Peixoto, B., Moraes, T. A., Mengin, V., Margalha, L., Vicente, R., Feil, R., Höhne, M., Sousa, A. G. G., Lilue, J., Stitt, M., Lunn, J. E. & Baena-González, E. (2021) Impact of the SnRK1 protein kinase on sucrose homeostasis and the transcriptome during the diel cycle. Plant Physiology, 187(3), 1357-1373. https://doi.org/10.1093/plphys/kiab350.
Pfeiffer, A., Janocha, D., Dong, Y., Medzihradszky, A., Schöne, S., Daum, G., Suzaki, T., Forner, J., Langenecker, T., Rempel, E., Schmid, M., Wirtz, M., Hell, R. & Lohmann, J. U. (2016) Integration of light and metabolic signals for stem cell activation at the shoot apical meristem. eLife, 5, e17023. https://doi.org/10.7554/eLife.17023.
Poitout, A., Crabos, A., Petřík, I., Novák, O., Krouk, G., Lacombe, B. & Ruffel, S. (2018) Responses to Systemic Nitrogen Signaling in Arabidopsis Roots Involve trans-Zeatin in Shoots. The Plant Cell, 30(6), 1243-1257. https://doi.org/10.1105/tpc.18.00011.
Pourtau, N., Jennings, R., Pelzer, E., Pallas, J. & Wingler, A. (2006) Effect of sugar-induced senescence on gene expression and implications for the regulation of senescence in Arabidopsis. Planta, 224(3), 556-568. https://doi.org/10.1007/s00425-006-0243-y.
Pourtau, N., Marès, M., Purdy, S., Quentin, N., Ruël, A. & Wingler, A. (2004) Interactions of abscisic acid and sugar signalling in the regulation of leaf senescence. Planta, 219(5), 765-772. https://doi.org/10.1007/s00425-004-1279-5.
Price, J., Laxmi, A., St. Martin, S. K. & Jang, J.-C. (2004) Global Transcription Profiling Reveals Multiple Sugar Signal Transduction Mechanisms in Arabidopsis[W]. The Plant Cell, 16(8), 2128-2150. https://doi.org/10.1105/tpc.104.022616.
Pu, Y., Luo, X. & Bassham, D. C. (2017) TOR-Dependent and -Independent Pathways Regulate Autophagy in Arabidopsis thaliana. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.01204.
Radchuk, V., Belew, Z. M., Gündel, A., Mayer, S., Hilo, A., Hensel, G., Sharma, R., Neumann, K., Ortleb, S., Wagner, S., Muszynska, A., Crocoll, C., Xu, D., Hoffie, I., Kumlehn, J., Fuchs, J., Peleke, F. F., Szymanski, J. J., Rolletschek, H., Nour-Eldin, H. H. & Borisjuk, L. (2023) SWEET11b transports both sugar and cytokinin in developing barley grains. The Plant Cell, 35(6), 2186-2207. https://doi.org/10.1093/plcell/koad055.
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 PeptidePlant Physiology, 150(3), 1356-1367. https://doi.org/10.1104/pp.109.136036.
Ramon, M., Dang, T. V. T., Broeckx, T., Hulsmans, S., Crepin, N., Sheen, J. & Rolland, F. (2019) Default Activation and Nuclear Translocation of the Plant Cellular Energy Sensor SnRK1 Regulate Metabolic Stress Responses and Development. The Plant Cell, 31(7), 1614-1632. https://doi.org/10.1105/tpc.18.00500.
Reed, A. J., Canvin, D. T., Sherrard, J. H. & Hageman, R. H. (1983) Assimilation of [15N]Nitrate and [15N]Nitrite in Leaves of Five Plant Species under Light and Dark Conditions 1. Plant Physiology, 71(2), 291-294. https://doi.org/10.1104/pp.71.2.291.
Reguera, M., Peleg, Z., Abdel-Tawab, Y. M., Tumimbang, E. B., Delatorre, C. A. & Blumwald, E. (2013) Stress-Induced Cytokinin Synthesis Increases Drought Tolerance through the Coordinated Regulation of Carbon and Nitrogen Assimilation in RicePlant Physiology, 163(4), 1609-1622. https://doi.org/10.1104/pp.113.227702.
Ribeiro, D. M., Araújo, W. L., Fernie, A. R., Schippers, J. H. M. & Mueller-Roeber, B. (2012) Action of Gibberellins on Growth and Metabolism of Arabidopsis Plants Associated with High Concentration of Carbon DioxidePlant Physiology, 160(4), 1781-1794. https://doi.org/10.1104/pp.112.204842.
Rodrigues, A., Adamo, M., Crozet, P., Margalha, L., Confraria, A., Martinho, C., Elias, A., Rabissi, A., Lumbreras, V., González-Guzmán, M., Antoni, R., Rodriguez, P. L. & Baena-González, E. (2013) ABI1 and PP2CA Phosphatases Are Negative Regulators of Snf1-Related Protein Kinase1 Signaling in ArabidopsisThe Plant Cell, 25(10), 3871-3884. https://doi.org/10.1105/tpc.113.114066.
Rolland, F., Baena-Gonzalez, E. & Sheen, J. (2006) SUGAR SENSING AND SIGNALING IN PLANTS: Conserved and Novel Mechanisms. Annual Review of Plant Biology, 57(1), 675-709. https://doi.org/10.1146/annurev.arplant.57.032905.105441.
Rubio-Asensio, J. S. & Bloom, A. J. (2017) Inorganic nitrogen form: a major player in wheat and Arabidopsis responses to elevated CO2. Journal of Experimental Botany, 68(10), 2611-2625. https://doi.org/10.1093/jxb/erw465.
Ruffel, S., Krouk, G., Ristova, D., Shasha, D., Birnbaum, K. D. & Coruzzi, G. M. (2011) Nitrogen economics of root foraging: Transitive closure of the nitrate–cytokinin relay and distinct systemic signaling for N supply vs. demand. Proceedings of the National Academy of Sciences, 108(45), 18524-18529. https://doi.org/10.1073/pnas.1108684108.
Ryabova, L. A., Robaglia, C. & Meyer, C. (2019) Target of Rapamycin kinase: central regulatory hub for plant growth and metabolism. Journal of Experimental Botany, 70(8), 2211-2216. https://doi.org/10.1093/jxb/erz108.
Sakr, S., Wang, M., Dédaldéchamp, F., Perez-Garcia, M.-D., Ogé, L., Hamama, L. & Atanassova, R. (2018) The Sugar-Signaling Hub: Overview of Regulators and Interaction with the Hormonal and Metabolic Network. International Journal of Molecular Sciences, 19(9).
Salem, M. A., Li, Y., Wiszniewski, A. & Giavalisco, P. (2017) Regulatory-associated protein of TOR (RAPTOR) alters the hormonal and metabolic composition of Arabidopsis seeds, controlling seed morphology, viability and germination potential. The Plant Journal, 92(4), 525-545. https://doi.org/10.1111/tpj.13667.
Saroussi, S., Karns, D. A. J., Thomas, D. C., Bloszies, C., Fiehn, O., Posewitz, M. C. & Grossman, A. R. (2019) Alternative outlets for sustaining photosynthetic electron transport during dark-to-light transitions. Proceedings of the National Academy of Sciences, 116(23), 11518-11527. https://doi.org/10.1073/pnas.1903185116.
Sato, T., Maekawa, S., Yasuda, S., Domeki, Y., Sueyoshi, K., Fujiwara, M., Fukao, Y., Goto, D. B. & Yamaguchi, J. (2011) Identification of 14-3-3 proteins as a target of ATL31 ubiquitin ligase, a regulator of the C/N response in Arabidopsis. The Plant Journal, 68(1), 137-146. https://doi.org/10.1111/j.1365-313X.2011.04673.x.
Scheible, W. R., Morcuende, R., Czechowski, T., Fritz, C., Osuna, D., Palacios-Rojas, N., Schindelasch, D., Thimm, O., Udvardi, M. K. & Stitt, M. (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(1), 2483-99.
Schepetilnikov, M., Dimitrova, M., Mancera-Martínez, E., Geldreich, A., Keller, M. & Ryabova, L. A. (2013) TOR and S6K1 promote translation reinitiation of uORF-containing mRNAs via phosphorylation of eIF3h. The EMBO Journal, 32(8), 1087-1102-1102. https://doi.org/10.1038/emboj.2013.61.
Schepetilnikov, M., Makarian, J., Srour, O., Geldreich, A., Yang, Z., Chicher, J., Hammann, P. & Ryabova, L. A. (2017) GTPase ROP2 binds and promotes activation of target of rapamycin, TOR, in response to auxin. The EMBO Journal, 36(7), 886-903-903. https://doi.org/10.15252/embj.201694816.
Schröder, F., Lisso, J., Obata, T., Erban, A., Maximova, E., Giavalisco, P., Kopka, J., Fernie, A. R., Willmitzer, L. & Müssig, C. (2014) Consequences of induced brassinosteroid deficiency in Arabidopsis leaves. BMC Plant Biology, 14(1), 309. https://doi.org/10.1186/s12870-014-0309-0.
Scofield, G. N., Hirose, T., Aoki, N. & Furbank, R. T. (2007) Involvement of the sucrose transporter, OsSUT1, in the long-distance pathway for assimilate transport in rice. Journal of Experimental Botany, 58(12), 3155-3169. https://doi.org/10.1093/jxb/erm153.
Sehnke, P. C., Chung, H.-J., Wu, K. & Ferl, R. J. (2001) Regulation of starch accumulation by granule-associated plant 14-3-3 proteins. Proceedings of the National Academy of Sciences, 98(2), 765-770. https://doi.org/10.1073/pnas.98.2.765.
Shimono, H. & Bunce, J. A. (2009) Acclimation of nitrogen uptake capacity of rice to elevated atmospheric CO2 concentration. Annals of Botany, 103(1), 87-94. https://doi.org/10.1093/aob/mcn209.
Shin, R., Jez, J. M., Basra, A., Zhang, B. & Schachtman, D. P. (2011) 14-3-3 proteins fine-tune plant nutrient metabolism. FEBS Lett, 585(1), 143-7. https://doi.org/10.1016/j.febslet.2010.11.025.
Siddiqui, H., Hayat, S. & Bajguz, A. (2018) Regulation of photosynthesis by brassinosteroids in plants. Acta Physiologiae Plantarum, 40(3), 59. https://doi.org/10.1007/s11738-018-2639-2.
Signora, L., De Smet, I., Foyer, C. H. & Zhang, H. (2001) ABA plays a central role in mediating the regulatory effects of nitrate on root branching in Arabidopsis. The Plant Journal, 28(6), 655-662. https://doi.org/10.1046/j.1365-313x.2001.01185.x.
Smeekens, S. & Hellmann, H. A. (2014) Sugar sensing and signaling in plants. Frontiers in plant science, 5, 113-113. https://doi.org/10.3389/fpls.2014.00113.
Smet, D., Depaepe, T., Vandenbussche, F., Callebert, P., Nijs, I., Ceulemans, R. & Van Der Straeten, D. (2020) The involvement of the phytohormone ethylene in the adaptation of Arabidopsis rosettes to enhanced atmospheric carbon dioxide concentrations. Environmental and Experimental Botany, 177, 104128. https://doi.org/10.1016/j.envexpbot.2020.104128.
Song, L., Xu, G., Li, T., Zhou, H., Lin, Q., Chen, J., Wang, L., Wu, D., Li, X., Wang, L., Zhu, S. & Yu, F. (2022) The RALF1-FERONIA complex interacts with and activates TOR signaling in response to low nutrients. Molecular Plant, 15(7), 1120-1136. https://doi.org/10.1016/j.molp.2022.05.004.
Soto-Burgos, J. & Bassham, D. C. (2017) SnRK1 activates autophagy via the TOR signaling pathway in Arabidopsis thaliana. PLOS ONE, 12(8), e0182591. https://doi.org/10.1371/journal.pone.0182591.
Soussana, J.-F., Teyssonneyre, F., Picon-Cochard, C. & Dawson, L. (2005) A trade-off between nitrogen uptake and use increases responsiveness to elevated CO2 in infrequently cut mixed C3 grasses. New Phytologist, 166(1), 217-230. https://doi.org/10.1111/j.1469-8137.2005.01332.x.
Stitt, M. & Krapp, A. (1999) The interaction between elevated carbon dioxide and nitrogen nutrition: the physiological and molecular background. Plant, Cell & Environment, 22(6), 583-621. https://doi.org/10.1046/j.1365-3040.1999.00386.x.
Stitz, M., Kuster, D., Reinert, M., Schepetilnikov, M., Berthet, B., Reyes-Hernández, J., Janocha, D., Artins, A., Boix, M., Henriques, R., Pfeiffer, A., Lohmann, J., Gaquerel, E. & Maizel, A. (2023) TOR acts as a metabolic gatekeeper for auxin-dependent lateral root initiation in Arabidopsis thaliana. The EMBO Journal, 42(10), e111273. https://doi.org/10.15252/embj.2022111273.
Sun, H., Li, W., Burritt, D. J., Tian, H., Zhang, H., Liang, X., Miao, Y., Mostofa, M. G. & Tran, L.-S. P. (2022) Strigolactones interact with other phytohormones to modulate plant root growth and development. The Crop Journal, 10(6), 1517-1527. https://doi.org/10.1016/j.cj.2022.07.014.
Świȩdrych, A., Prescha, A., Matysiak-Kata, I., Biernat, J. & Szopa, J. (2002) Repression of the 14-3-3 Gene Affects the Amino Acid and Mineral Composition of Potato Tubers. Journal of Agricultural and Food Chemistry, 50(7), 2137-2141. https://doi.org/10.1021/jf0112825.
Sze, H., Li, X. & Palmgren, M. G. (1999) Energization of Plant Cell Membranes by H+-Pumping ATPases: Regulation and Biosynthesis. The Plant Cell, 11(4), 677-689. https://doi.org/10.1105/tpc.11.4.677.
Takahashi, Y., Bosmans, K. C., Hsu, P.-K., Paul, K., Seitz, C., Yeh, C.-Y., Wang, Y.-S., Yarmolinsky, D., Sierla, M., Vahisalu, T., Waszczak, C., McCammon, J. A., Kangasjärvi, J., Zhang, L., Kollist, H., Trac, T. & Schroeder, J. I. (2022) Stomatal CO2/bicarbonate sensor consists of two interacting protein kinases, Raf-like HT1 and non-kinase-activity requiring MPK12/MPK4. Science Advances, 8(49), eabq6161. https://doi.org/10.1126/sciadv.abq6161.
Takatani, N., Ito, T., Kiba, T., Mori, M., Miyamoto, T., Maeda, S.-I. & Omata, T. (2014) Effects of high CO2 on growth and metabolism of Arabidopsis seedlings during growth with a constantly limited supply of nitrogen. Plant & cell physiology, 55(2), 281-292. https://doi.org/10.1093/pcp/pct186.
Tanner, W. & Beevers, H. (2001) Transpiration, a prerequisite for long-distance transport of minerals in plants?Proceedings of the National Academy of Sciences, 98(16), 9443-9447. https://doi.org/10.1073/pnas.161279898.
Taub, D. R. & Wang, X. (2008) Why are Nitrogen Concentrations in Plant Tissues Lower under Elevated CO2? A Critical Examination of the Hypotheses. Journal of Integrative Plant Biology, 50(11), 1365-1374. https://doi.org/10.1111/j.1744-7909.2008.00754.x.
Teng, S., Keurentjes, J., Bentsink, L. n., Koornneef, M. & Smeekens, S. (2005) Sucrose-Specific Induction of Anthocyanin Biosynthesis in Arabidopsis Requires the MYB75/PAP1 Gene. Plant Physiology, 139(4), 1840-1852. https://doi.org/10.1104/pp.105.066688.
Tõldsepp, K., Zhang, J., Takahashi, Y., Sindarovska, Y., Hõrak, H., Ceciliato, P. H. O., Koolmeister, K., Wang, Y.-S., Vaahtera, L., Jakobson, L., Yeh, C.-Y., Park, J., Brosche, M., Kollist, H. & Schroeder, J. I. (2018) Mitogen-activated protein kinases MPK4 and MPK12 are key components mediating CO2-induced stomatal movements. The Plant Journal, 96(5), 1018-1035. https://doi.org/10.1111/tpj.14087.
Toroser, D., Athwal, G. S. & Huber, S. C. (1998) Site-specific regulatory interaction between spinach leaf sucrose-phosphate synthase and 14-3-3 proteins. FEBS Letters, 435(1), 110-114. https://doi.org/10.1016/S0014-5793(98)01048-5.
Tsai, A. Y.-L. & Gazzarrini, S. (2012) AKIN10 and FUSCA3 interact to control lateral organ development and phase transitions in Arabidopsis. The Plant Journal, 69(5), 809-821. https://doi.org/10.1111/j.1365-313X.2011.04832.x.
Tsay, Y. F., Schroeder, J. I., Feldmann, K. A. & Crawford, N. M. (1993) The herbicide sensitivity gene CHL1 of Arabidopsis encodes a nitrate-inducible nitrate transporter. Cell, 72(5), 705-13. https://doi.org/10.1016/0092-8674(93)90399-b.
Umezawa, T., Nakashima, K., Miyakawa, T., Kuromori, T., Tanokura, M., Shinozaki, K. & Yamaguchi-Shinozaki, K. (2010) Molecular Basis of the Core Regulatory Network in ABA Responses: Sensing, Signaling and Transport. Plant and Cell Physiology, 51(11), 1821-1839. https://doi.org/10.1093/pcp/pcq156.
Umezawa, T., Sugiyama, N., Takahashi, F., Anderson, J. C., Ishihama, Y., Peck, S. C. & Shinozaki, K. (2013) Genetics and Phosphoproteomics Reveal a Protein Phosphorylation Network in the Abscisic Acid Signaling Pathway in Arabidopsis thaliana. Science Signaling, 6(270), rs8-rs8. https://doi.org/10.1126/scisignal.2003509.
Vanderwall, M. & Gendron, J. M. (2023) HEXOKINASE1 and glucose-6-phosphate fuel plant growth and development. Development, 150(20), dev202346. https://doi.org/10.1242/dev.202346.
Van Leene, J., Eeckhout, D., Gadeyne, A., Matthijs, C., Han, C., De Winne, N., Persiau, G., Van De Slijke, E., Persyn, F., Mertens, T., Smagghe, W., Crepin, N., Broucke, E., Van Damme, D., Pleskot, R., Rolland, F. & De Jaeger, G. (2022) Mapping of the plant SnRK1 kinase signalling network reveals a key regulatory role for the class II T6P synthase-like proteins. Nature Plants, 8(11), 1245-1261. https://doi.org/10.1038/s41477-022-01269-w.
Van Leene, J., Han, C., Gadeyne, A., Eeckhout, D., Matthijs, C., Cannoot, B., De Winne, N., Persiau, G., Van De Slijke, E., Van de Cotte, B., Stes, E., Van Bel, M., Storme, V., Impens, F., Gevaert, K., Vandepoele, K., De Smet, I. & De Jaeger, G. (2019) Capturing the phosphorylation and protein interaction landscape of the plant TOR kinase. Nature Plants, 5(3), 316-327. https://doi.org/10.1038/s41477-019-0378-z.
Vaughn, M. W., Harrington, G. N. & Bush, D. R. (2002) Sucrose-mediated transcriptional regulation of sucrose symporter activity in the phloem. Proceedings of the National Academy of Sciences, 99(16), 10876-10880. https://doi.org/10.1073/pnas.172198599.
Viana, A. J. C., Matiolli, C. C., Newman, D. W., Vieira, J. G. P., Duarte, G. T., Martins, M. C. M., Gilbault, E., Hotta, C. T., Caldana, C. & Vincentz, M. (2021) The sugar-responsive circadian clock regulator bZIP63 modulates plant growth. New Phytologist, 231(5), 1875-1889. https://doi.org/10.1111/nph.17518.
Wang, M., Le Gourrierec, J., Jiao, F., Demotes-Mainard, S., Perez-Garcia, M.-D., Ogé, L., Hamama, L., Crespel, L., Bertheloot, J., Chen, J., Grappin, P. & Sakr, S. (2021a) Convergence and Divergence of Sugar and Cytokinin Signaling in Plant Development. International Journal of Molecular Sciences, 22(3).
Wang, P., Nolan, T. M., Clark, N. M., Jiang, H., Montes-Serey, C., Guo, H., Bassham, D. C., Walley, J. W. & Yin, Y. (2021b) The F-box E3 ubiquitin ligase BAF1 mediates the degradation of the brassinosteroid-activated transcription factor BES1 through selective autophagy in Arabidopsis. The Plant Cell, 33(11), 3532-3554. https://doi.org/10.1093/plcell/koab210.
Wang, P., Xue, L., Batelli, G., Lee, S., Hou, Y.-J., Van Oosten, M. J., Zhang, H., Tao, W. A. & Zhu, J.-K. (2013) Quantitative phosphoproteomics identifies SnRK2 protein kinase substrates and reveals the effectors of abscisic acid action. Proceedings of the National Academy of Sciences, 110(27), 11205-11210. https://doi.org/10.1073/pnas.1308974110.
Wang, R., Okamoto, M., Xing, X. & Crawford, N. M. (2003) Microarray analysis of the nitrate response in Arabidopsis roots and shoots reveals over 1,000 rapidly responding genes and new linkages to glucose, trehalose-6-phosphate, iron, and sulfate metabolism. Plant physiology, 132(2), 556-567. https://doi.org/10.1104/pp.103.021253.
Wheeldon, C. D. & Bennett, T. (2021) There and back again: An evolutionary perspective on long-distance coordination of plant growth and development. Seminars in Cell & Developmental Biology, 109, 55-67. https://doi.org/10.1016/j.semcdb.2020.06.011.
Wingler, A. (2018) Transitioning to the Next Phase: The Role of Sugar Signaling throughout the Plant Life Cycle. Plant Physiology, 176(2), 1075-1084. https://doi.org/10.1104/pp.17.01229.
Wingler, A., Marès, M. & Pourtau, N. (2004) Spatial patterns and metabolic regulation of photosynthetic parameters during leaf senescence. New Phytologist, 161(3), 781-789. https://doi.org/10.1111/j.1469-8137.2004.00996.x.
Wingler, A., Masclaux-Daubresse, C. & Fischer, A. M. (2009) Sugars, senescence, and ageing in plants and heterotrophic organisms. Journal of Experimental Botany, 60(4), 1063-1066. https://doi.org/10.1093/jxb/erp067.
Wu, Y., Shi, L., Li, L., Fu, L., Liu, Y., Xiong, Y. & Sheen, J. (2019) Integration of nutrient, energy, light, and hormone signalling via TOR in plants. Journal of Experimental Botany, 70(8), 2227-2238. https://doi.org/10.1093/jxb/erz028.
Xu, G., Chen, W., Song, L., Chen, Q., Zhang, H., Liao, H., Zhao, G., Lin, F., Zhou, H. & Yu, F. (2019) FERONIA phosphorylates E3 ubiquitin ligase ATL6 to modulate the stability of 14-3-3 proteins in response to the carbon/nitrogen ratio. Journal of Experimental Botany, 70(21), 6375-6388. https://doi.org/10.1093/jxb/erz378.
Xuan, W., Beeckman, T. & Xu, G. (2017) Plant nitrogen nutrition: sensing and signaling. Current Opinion in Plant Biology, 39, 57-65. https://doi.org/10.1016/j.pbi.2017.05.010.
Yamagishi, K., Tatematsu, K., Yano, R., Preston, J., Kitamura, S., Takahashi, H., McCourt, P., Kamiya, Y. & Nambara, E. (2009) CHOTTO1, a double AP2 domain protein of Arabidopsis thaliana, regulates germination and seedling growth under excess supply of glucose and nitrate. Plant Cell Physiol, 50(2), 330-40.
Yanagisawa, S., Yoo, S.-D. & Sheen, J. (2003) Differential regulation of EIN3 stability by glucose and ethylene signalling in plants. Nature, 425(6957), 521-525. https://doi.org/10.1038/nature01984.
Yang, B., Wang, J., Yu, M., Zhang, M., Zhong, Y., Wang, T., Liu, P., Song, W., Zhao, H., Fastner, A., Suter, M., Rentsch, D., Ludewig, U., Jin, W., Geiger, D., Hedrich, R., Braun, D. M., Koch, K. E., McCarty, D. R., Wu, W.-H., Li, X., Wang, Y. & Lai, J. (2022) The sugar transporter ZmSUGCAR1 of the nitrate transporter 1/peptide transporter family is critical for maize grain filling. The Plant Cell, 34(11), 4232-4254. https://doi.org/10.1093/plcell/koac256.
Yang, Y., Yu, X., Song, L. & An, C. (2011) ABI4 Activates DGAT1 Expression in Arabidopsis Seedlings during Nitrogen DeficiencyPlant Physiology, 156(2), 873-883. https://doi.org/10.1104/pp.111.175950.
Yasuda, S., Sato, T., Maekawa, S., Aoyama, S., Fukao, Y. & Yamaguchi, J. (2014) Phosphorylation of Arabidopsis Ubiquitin Ligase ATL31 Is Critical for Plant Carbon/Nitrogen Nutrient Balance Response and Controls the Stability of 14-3-3 Proteins*. Journal of Biological Chemistry, 289(22), 15179-15193. https://doi.org/10.1074/jbc.M113.533133.
Yoshida, T., Mergner, J., Yang, Z., Liu, J., Kuster, B., Fernie, A. R. & Grill, E. (2024) Integrating multi-omics data reveals energy and stress signaling activated by abscisic acid in Arabidopsis. The Plant Journal, 119(2), 1112-1133. https://doi.org/10.1111/tpj.16765.
Yuan, X., Xu, P., Yu, Y. & Xiong, Y. (2020) Glucose-TOR signaling regulates PIN2 stability to orchestrate auxin gradient and cell expansion in Arabidopsis root. Proceedings of the National Academy of Sciences, 117(51), 32223-32225. https://doi.org/10.1073/pnas.2015400117.
Zhang, H., Rong, H. & Pilbeam, D. (2007) Signalling mechanisms underlying the morphological responses of the root system to nitrogen in Arabidopsis thaliana. Journal of Experimental Botany, 58(9), 2329-2338. https://doi.org/10.1093/jxb/erm114.
Zhang, T., Noll, S. E., Peng, J. T., Klair, A., Tripka, A., Stutzman, N., Cheng, C., Zare, R. N. & Dickinson, A. J. (2023) Chemical imaging reveals diverse functions of tricarboxylic acid metabolites in root growth and development. Nature Communications, 14(1), 2567. https://doi.org/10.1038/s41467-023-38150-z.
Zhang, Y., Liu, Z., Wang, J., Chen, Y., Bi, Y. & He, J. (2015) Brassinosteroid is required for sugar promotion of hypocotyl elongation in Arabidopsis in darkness. Planta, 242(4), 881-893. https://doi.org/10.1007/s00425-015-2328-y.
Zhang, Z., Zhu, J.-Y., Roh, J., Marchive, C., Kim, S.-K., Meyer, C., Sun, Y., Wang, W. & Wang, Z.-Y. (2016) TOR Signaling Promotes Accumulation of BZR1 to Balance Growth with Carbon Availability in Arabidopsis. Current Biology, 26(14), 1854-1860. https://doi.org/10.1016/j.cub.2016.05.005.
Zhao, X., Li, F. & Li, K. (2021) The 14-3-3 proteins: regulators of plant metabolism and stress responses. Plant Biology, 23(4), 531-539. https://doi.org/10.1111/plb.13268.
Zhao, Y. (2018) Essential Roles of Local Auxin Biosynthesis in Plant Development and in Adaptation to Environmental Changes. Annual Review of Plant Biology, 69(Volume 69, 2018), 417-435. https://doi.org/10.1146/annurev-arplant-042817-040226.
Zhong, S., Shi, H., Xue, C., Wei, N., Guo, H. & Deng, X. W. (2014) Ethylene-orchestrated circuitry coordinates a seedling's response to soil cover and etiolated growth. Proceedings of the National Academy of Sciences, 111(11), 3913-3920. https://doi.org/10.1073/pnas.1402491111.
Zhuo, D., Okamoto, M., Vidmar, J. J. & Glass, A. D. M. (1999) Regulation of a putative high-affinity nitrate transporter ( ) in roots of. The Plant Journal, 17(5), 563-568. https://doi.org/10.1046/j.1365-313X.1999.00396.x.
Zhuo, F., Xiong, F., Deng, K., Li, Z. & Ren, M. (2020) Target of Rapamycin (TOR) Negatively Regulates Ethylene Signals in Arabidopsis. International Journal of Molecular Sciences, 21(8).
Żuk, M., Skała, J., Biernat, J. & Szopa, J. (2003) Repression of six 14-3-3 protein isoforms resulting in the activation of nitrate and carbon fixation key enzymes from transgenic potato plants. Plant Science, 165(4), 731-741. https://doi.org/10.1016/S0168-9452(03)00231-0.
Zuk, M., Weber, R. & Szopa, J. (2005) 14-3-3 Protein Down-regulates Key Enzyme Activities of Nitrate and Carbohydrate Metabolism in Potato Plants. Journal of Agricultural and Food Chemistry, 53(9), 3454-3460. https://doi.org/10.1021/jf0485584.
Please check your email for instructions on resetting your password.
If you do not receive an email within 10 minutes, your email address may not be registered,
and you may need to create a new Wiley Online Library account.
Request Username
Can't sign in? Forgot your username?
Enter your email address below and we will send you your username
If the address matches an existing account you will receive an email with instructions to retrieve your username
The full text of this article hosted at iucr.org is unavailable due to technical difficulties.