Skip to main content
Plant Physiology logoLink to Plant Physiology
. 2021 May 20;187(4):1893–1914. doi: 10.1093/plphys/kiab228

Cellular export of sugars and amino acids: role in feeding other cells and organisms

Ji-Yun Kim 1, Eliza P -I Loo 1, Tin Yau Pang 2, Martin Lercher 2, Wolf B Frommer 1,3, Michael M Wudick 1,✉,
PMCID: PMC8644676  PMID: 34015139

Abstract

Sucrose, hexoses, and raffinose play key roles in the plant metabolism. Sucrose and raffinose, produced by photosynthesis, are translocated from leaves to flowers, developing seeds and roots. Translocation occurs in the sieve elements or sieve tubes of angiosperms. But how is sucrose loaded into and unloaded from the sieve elements? There seem to be two principal routes: one through plasmodesmata and one via the apoplasm. The best-studied transporters are the H+/SUCROSE TRANSPORTERs (SUTs) in the sieve element-companion cell complex. Sucrose is delivered to SUTs by SWEET sugar uniporters that release these key metabolites into the apoplasmic space. The H+/amino acid permeases and the UmamiT amino acid transporters are hypothesized to play analogous roles as the SUT-SWEET pair to transport amino acids. SWEETs and UmamiTs also act in many other important processes—for example, seed filling, nectar secretion, and pollen nutrition. We present information on cell type-specific enrichment of SWEET and UmamiT family members and propose several members to play redundant roles in the efflux of sucrose and amino acids across different cell types in the leaf. Pathogens hijack SWEETs and thus represent a major susceptibility of the plant. Here, we provide an update on the status of research on intercellular and long-distance translocation of key metabolites such as sucrose and amino acids, communication of the plants with the root microbiota via root exudates, discuss the existence of transporters for other important metabolites and provide potential perspectives that may direct future research activities.


An update on intercellular and long-distance translocation of sugars and amino acids, including plant-root microbiota communication, other metabolite transporters is provided, and perspectives are discussed.

Introduction

Cells can secrete specific compounds for various functions, for example, disposal, protection from osmotic damage, feeding of other cells—either a neighboring cell in the same organism or cells from other organisms—or solute distribution in multicellular organisms, and defense. A well-studied example is Corynebacterium glutamicum, which effectively secretes glutamate and is therefore used for the industrial production of glutamate (Nakayama et al., 2018). Corynebacterium glutamicum secretes glutamate via a mechanosensitive efflux transporter. Many bacteria secrete valine into their biofilms where it serves as an antibiotic (Valle et al., 2008). This review focuses on processes in which major metabolites, in particular sugars and amino acids, are secreted from plant cells. Key physiological aspects discussed here relate to the distribution of assimilates in plants, as well as to the exchange of metabolites with other organisms, in particular nectar secretion and feeding of beneficial and pathogenic microbes. This review highlights families of transporters for metabolites—sugars and amino acids—and their role in carbon and nitrogen allocation: SWEETs and UmamiTs, as well as additional transporters for other metabolites and their roles in physiology, pathogenesis, and symbiosis.

Advances

  • SWEET sugar transporters and UmamiT amino acid transporters are expressed in specific cell types that play roles in secretory functions.

  • Amino acid metabolism in two phloem cell types, phloem parenchyma and CCs, are distinct, indicating the metabolism in the two phloem cell types may shape the relative amino acid composition of the phloem sap.

  • Numerous SWEETs across different plant families are induced during arbuscular mycorrhizal fungus and rhizobial symbiosis, implicating SWEETs in symbiotic nutrition.

  • Metabolomic studies reveal region dependent root exudation under various growth conditions.

SWEET and UmamiT transporters: evolution and structure

Members of the SemiSWEET-SWEET sugar transporter superfamily had originally been described as homologs of Medicago truncatula NODULIN 3 (MtN3), based on the observation that its transcript levels increased during nodulation (Gamas et al., 1996). SWEETs belong to an ancient family with members present already in Archaea. Plant genomes typically contain approximately 20 SWEETs with two conserved PQ-loop repeats (Table 1; Supplemental Figure S1; Supplemental Table S1).

Table 1.

Transporters for metabolites with potential roles in cellular efflux discussed in this study

Transporter Family Super Family Conserved Domain(s) PFAM Interpro TCDB PDB
SWEET (MtN3-like) SWEET/Semi SWEET SWEET Sugar efflux transporter, PQ-loop repeat PF04193–PF03083 IPR004316 2.A.123 5CTH, 5XPD, 5CTG
UmamiT (MtN21-like) P-DME DMT EamA-like repeat PF00892 IPR000620 2.A.7.4 5I20a
MATE MATE MviN MatE PF01554 IPR002528 2.A.66 5Y50
ALMT ArAE UspB PF11744 IPR020966 9.A.85 N.D.

Abbreviations: ABC, ATP-binding cassette; ArAE, aromatic acid exporter; MtN3 or MtN21-like, Medicago truncatula nodulin 3 or 21- like; MviN, mouse virulence N; N.D., not determined; PDB, protein database; P-DME, plant drug/metabolite exporter; PFAM, protein family; TCDB, transporter classification database; UspB, universal stress protein-B.

a

Bacterial homolog.

UmamiT amino acid transporters had originally been described as MtN21-like nodulins (Dinkeloo et al., 2018). They belong to the plant-specific branch of the Drug/Metabolite Exporter (P-DME) family, which is part of the larger family of drug/metabolite transporters (DMT), and which shares sequence homology with members of the five transmembrane-domain Bacterial/Archaeal transporter (BAT) family. Prokaryotic DMT paralogs function as amino acid efflux transporters, for example, the Escherichia coli O-acetyl-serine/cysteine exporter EamA (Franke et al., 2003). UmamiTs contain two EamA-like domains (Table 1). Plants typically have approximately 50 UmamiT paralogs per haploid genome (Table 1; Supplemental Figure S2; Supplemental Table S2).

Structures of SWEETs, SemiSWEETs, their ancestral prokaryotic homolog, and a distant UmamiT homolog have been resolved by X-ray crystallography (Xu et al., 2014; Tao et al., 2015; Tsuchiya et al., 2016; Han et al., 2017; Latorraca et al., 2017; Figure 1, A–D). While eukaryotic SWEETs are composed of seven transmembrane helices, with an apparent parallel symmetry axis (3 + 1 + 3), prokaryotic SemiSWEETs are amongst the smallest known transporters with only three helices (Figure 1A). SemiSWEET pores are formed by parallel oriented dimers, while eukaryotic SWEETs contain a central fourth transmembrane helix that orients the second repeat in the same orientation as in dimeric SemiSWEETs (Xu et al., 2014; Tao et al., 2015; Han et al., 2017; Figure 1, A and C). SemiSWEETs have two gates, and the transport cycle alternates between outside open, occluded and open inside conformations (Latorraca et al., 2017).

Figure 1.

Figure 1

Topology of SemiSWEET, SWEET, SemiUmamiT and UmamiT. A, Bacterial SemiSWEET unit comprised of a triple helix bundle (green). B, Bacterial SemiUmamiT topology based on bioinformatic analyses (Aramemnon) and the structure of BAT1 (Jack et al., 2001). C, Topology of eukaryotic SWEETs comprised of two triple helix bundles (light blue) fused via an additional linker helix (gray). D, Predicted topology for UmamiT based on bioinformatic analyses (Aramemnon) and the structure of the amino acid exporter YddG (Tsuchiya et al., 2016). TM, transmembrane domain; THB, triple helix bundle. Numbers indicate transmembrane helices (represented as boxes).

UmamiTs are predicted to contain ten transmembrane helices, similar to their distant prokaryotic homolog YddG, which functions as an amino acid exporter (Tsuchiya et al., 2016; Figure 1D). Most DMTs consist of inverted structural repeats, which are basket-shaped and related by a two-fold pseudosymmetry, yielding a substrate binding cavity at the center (Tsuchiya et al., 2016). Similar to SemiSWEETs, the prokaryotic BATs may form “half” transporters with five transmembrane helixes (Figure 1B). UmamiTs may have arisen by intragenic duplication from such ancestral transporter domains (Jack et al., 2001). Interestingly, the genome of Arabidopsis (Arabidopsis thaliana) also seems to code for “half”- or SemiUmamiTs. For instance, UmamiT43 is predicted with five transmembrane segments (http://aramemnon.uni‐koeln.de), but has not yet been functionally characterized.

SWEET and UmamiT substrates

The activity of SWEETs was identified through a screen of polytopic membrane proteins with unknown functions coexpressed with genetically encoded glucose or sucrose sensors in human embryonic kidney (HEK293T) cells (Chen et al., 2010, 2012). It was hypothesized that human cells, which are cultured in media with a neutral pH, and which lack plasma membrane H+-ATPases, may provide favorable conditions for identifying sucrose efflux transporters that might function as uniporters or sucrose/H+ antiporters (Fieuw and Patrick, 1993). SWEETs, similar as SemiSWEETs, can transport hexoses and/or the disaccharide sucrose (Table 2, Supplemental Figure S1; Chen et al., 2010, 2012; Xu et al., 2014; Tao et al., 2015). Phylogenetically, Arabidopsis SWEET members fall into four clades in which clade 3 members preferably mediate sucrose transport (Supplemental Figure S1). In addition to sugar transport, several SWEETs are capable of transporting gibberellic acid (GA), which at first sight, neither resembles glucose nor sucrose (Kanno et al., 2016; Morii et al., 2020; Table 2). Notably, GA biosynthesis pathway genes are enriched in phloem cell types where AtSWEET11-13 and several other GA transporters, such as AtNPF4.6, are enriched (Figure 2; Kim et al., 2021). Although all characterized clade III SWEETs are plasma membrane-localized, other members (i.e. clades II, IV) were also detected in vacuolar and ER membranes (Table 2, Supplemental Figure S1).

Table 2.

SWEETs in Arabidopsis

Gene Name (Alternative Name) Substrate(s) Locali-zation Physiological Role Reference
SWEET1 Glucose PM N.D. Chen et al., 2010
SWEET2 2-DOG TP Resistance to Pythium spp Chen et al., 2015a; Veillet et al., 2017; Sellami et al., 2019; Desrut et al., 2020
SWEET3 2-DOG N.D. N.D. Chen et al., 2015b; Desrut et al., 2020; Liao et al., 2020
SWEET4 Glucose PM Sugar supply to the axial tissues, freezing and drought tolerance and nonhost resistance Chen et al., 2010; Liu et al., 2016; Desrut et al., 2020; Liao et al., 2020
SWEET5 (VEX1) Glucose ND Possibly transport of sugars in vegetative cell of pollen grains Engel et al., 2005; Borges et al., 2008; Chen et al., 2010; Borges et al., 2012; Liao et al., 2020
SWEET6 2-DOG ER N.D. Chen et al., 2010; Lee et al., 2011; Chen et al., 2015b
SWEET7 Glucose N.D. N.D. Chen et al., 2010; Liao et al., 2020
SWEET8 (RPG1) Glucose PM Microspore development, pollen mitosis, primexine deposition, tapetum efflux Drakakaki et al., 2006; Chen et al., 2010; Sun et al., 2013; Veillet et al., 2017; Liao et al., 2020
SWEET9 Sucrose, weak glucose, GA PM, TGN Nectar secretion Lin et al., 2014; Kanno et al., 2016; Durand et al., 2018
SWEET10 Sucrose, GA N.D. Floral transition Chen et al., 2012, 2015a; Kanno et al., 2016; Durand et al., 2018; Andrés et al., 2020; Desrut et al., 2020)
SWEET11 Sucrose, glucose, fructose, GA PM Efflux of sucrose from PP for phloem loading, embryo nutrition, vascular development, freezing tolerance, salicylic acid-mediated defense response Chen et al., 2012, 2015c; Eom et al., 2015; Le Hir et al., 2015; Durand et al., 2016, 2018; Kanno et al., 2016; Gebauer et al., 2017; dos Anjos et al., 2018; Walerowski et al., 2018; Dinant et al., 2019; Sellami et al., 2019; Desrut et al., 2020; Huang et al., 2020; Wei et al., 2020; Zhao et al., 2020; Fichtner et al., 2021
SWEET12 Sucrose, glucose, fructose, GA PM Efflux of sucrose from PP for phloem loading, embryo nutrition, vascular development, freezing tolerance, salicylic acid-mediated defense response Chen et al., 2012, 2015c; Duan et al., 2014; Eom et al., 2015; Le Hir et al., 2015; Durand et al., 2016, 2018; Kanno et al., 2016; Gebauer et al., 2017; dos Anjos et al., 2018; Walerowski et al., 2018; Dinant et al., 2019; Sellami et al., 2019; Desrut et al., 2020; Huang et al., 2020; Wei et al., 2020; Zhao et al., 2020; Fichtner et al., 2021
SWEET13 (RPG2) Sucrose, GA PM Anther dehiscence, germination, seed development, vegetative growth, microspore development, pollen mitosis, primexine deposition, tapetum efflux Durand et al., 2016, 2018; Kanno et al., 2016; Han et al., 2017; Sellami et al., 2019; Andrés et al., 2020; Zhao et al., 2020; Fichtner et al., 2021
SWEET14 Sucrose, GA PM Anther dehiscence, germination, seed development, vegetative growth Durand et al., 2016, 2018; Kanno et al., 2016; Sellami et al., 2019; Andrés et al., 2020
SWEET15 (SAG29) Sucrose PM Embryo nutrition, accelerated senescence in overexpression lines Chen et al., 2010, 2015c; Seo et al., 2011; Matallana-Ramirez et al., 2013; Qi et al., 2015; Durand et al., 2016, 2018; Gao et al., 2016; Rasheed et al., 2016; Zhao et al., 2016; Gebauer et al., 2017; Kihira et al., 2017; Sellami et al., 2019; Desrut et al., 2020; Huang et al., 2020; Zhang et al., 2020; Zhao et al., 2020
SWEET16 Glucose, fructose, sucrose TP Overexpression shows altered germination rate, growth phenotype, and stress tolerance Blommel et al., 2004; Klemens et al., 2013; Guo et al., 2014; Walerowski et al., 2018; Sellami et al., 2019; Aubry et al., 2021
SWEET17 fructose TP fructose homeostasis regulation Chardon et al., 2013; Guo et al., 2014; Veillet et al., 2017; Walerowski et al., 2018; Aubry et al., 2021

Substrates, subcellular localization, and physiological roles of Arabidopsis SWEET family members.

Abbreviations: 2-DOG, 2-deoxyglucose (glucose analog); PM, plasma membrane; RPG1 or 2, RUPTURED POLLEN GRAIN 1 or 2; SAG29, SENESCENCE ASSOCIATED GENE 29; TGN, trans-Golgi network; TP, tonoplast; VEX1, VEGETATIVE CELL EXPRESSED 1.

Figure 2.

Figure 2

Leaf cell type specificity of GA transporters identified in heterologous system (Xenopus oocytes and yeast) and in planta (*). Dot plot showing transcript enrichment of GA transporters across 19 clusters of the leaf scRNA-seq data (Kim et al., 2021). The diameter of the dot indicates the percentage of cells in the cluster in which transcripts for that gene were detected, while the color of each dot represents the average log-scaled expression of each gene across all cells within a given cluster (see legend at lower right side). Cell types assigned to each cluster are indicated in the upper right panel. BS1, bundle sheath; BS2, bundle sheath cells enriched with photosynthetic processes, XP1 and XP2, xylem cells related with the bundle sheath; XP3, xylem cells enriched with vascular parenchyma markers; PCXP, procambium cells related to XP1; PCPP, procambium cells related to PP cells with transfer cell identity (PP1); u.a., unassigned. Note that NPF proteins transport additional substrates as reviewed in Corratgé-Faillie and Lacombe, 2017. The mRNA counts of SWEET9, SWEET10, SWEET14, NPF5.3, NPF4.2, NPF4.1, NPF2.5, NPF2.4, NPF2.1 were not detected in the dataset (Kim et al., 2021). For detailed information about the dataset and description of the clusters/subclusters, refer to Kim et al. (2021).

Given the comparatively high number and diverse chemical properties of amino acids (i.e. charge, polarity, aromaticity), the substrate recognition and transport mechanism of UmamiTs is likely more complex compared to that of SWEETs. While some prokaryotic DMTs, like YddG, seem to specifically transport aromatic amino acids, YdeD exports cysteine, asparagine, and glutamine, RhtA exports threonine and homoserine, and Rickettsia prowazekii Sam imports S-adenosylmethionine (Franke et al., 2003; Livshits et al., 2003; Tucker et al., 2003; Doroshenko et al., 2007; Tsuchiya et al., 2016). A similar preference can currently not be attributed to any UmamiTs, based on the admittedly limited data available. Phenylalanine—the sole aromatic amino acid included in UmamiT transport assays so far—was shown to be a substrate for UmamiT14, 24, and 25 (Besnard et al., 2016, 2018). However, the same transporters were also able to transport up to 13 additional proteinogenic amino acids and structurally related metabolites (Table 3, Supplemental Figure S2). Interestingly, UmamiT5/WAT1 (WALLS ARE THIN 1) facilitates vacuolar influx of indole-3-acetic acid, which is structurally similar to tryptophan (Ranocha et al., 2013). Phylogenetic evidence indicates that AtUmamiT5 belongs to a distinct clade (clade V; Supplemental Figure S2) that contains UmamiT1, also in the tonoplast (Schmidt et al., 2007). It is tempting to speculate that these members all mediate auxin transport across the tonoplast. γ-aminobutyric acid was transported by UmamiT23-25 (Besnard et al., 2018), citrulline by UmamiT18/SIAR1 (SILIQUES ARE RED 1; Ladwig et al., 2012). Though initially thought to be plasma membrane transporters, some UmamiTs localized to the vacuole and/or the ER (Table 3; Supplemental Figure S2).

Table 3.

UmamiTs in Arabidopsis

Gene Name (Alternative Name) Substrate(s) Locali-zation Physiological Role References
UmamiT1 N.D. TP N.D. Schmidt et al., 2007
UmamiT5 (WAT1) auxin (IAA) TP Vacuolar auxin influx Ranocha et al., 2010, 2013
UmamiT11 Gln PM Likely cellular efflux to support embryo growth Müller et al., 2015
UmamiT14 Glu, Phe, Gln/Arg, Ala, Ser, Gly, Asn, Pro, Thr, Val, His, Ile, Leu, citrulline PM Likely cellular efflux to support embryo growth, phloem unloading in roots Müller et al., 2015; Besnard et al., 2016
UmamiT18 (SIAR1) Asp, Gln/Arg, Ala, Asn, Thr, Val, His, Ile, Leu PM Phloem unloading in roots, apoplasmic release of amino acids in seeds Ladwig et al., 2012; Besnard et al., 2016
UmamiT23 Gln/Arg, Glu, GABA, Asp, Thr N.D. N.D. Besnard et al., 2018
UmamiT24 Gln/Arg, Ala, Glu, GABA, Phe, Val, Gly, Asp, Thr, Ser, Ile TP Involved in transient storage of amino acids Besnard et al., 2018
UmamiT25 Gln/Ala, Glu, Leu, GABA, Phe, Val, Gly, Asp, Thr, Ser, Ile, Pro PM Amino acid export from the endosperm Besnard et al., 2018
UmamiT28 Gln PM Likely cellular efflux to support embryo growth Müller et al., 2015
UmamiT29 Gln PM Likely cellular efflux to support embryo growth Müller et al., 2015
UmamiT36 (RTP1) N.D. ER N.D. Pan et al., 2016

Substrates, subcellular localization, and physiological roles of Arabidopsis UmamiT family members.

Abbreviations: ER, endoplasmic reticulum; GABA, γ-aminobutyric acid; IAA: indole-3-acetic acid; RTP1, RESISTANCE TO PHYTOPHTHORA PARASITICA 1; SIAR1, SILIQUES ARE RED 1; TP, tonoplast; WAT1,WALLS ARE THIN 1.

Transport mechanisms of SWEETs and UmamiTs

The study of sugar transport mechanism in plants started more than 40 years ago (Giaquinta, 1976). Since then, various sugar transporters from different species were characterized in heterologous expression systems (e.g. Riesmeier et al., 1992; Gahrtz et al., 1994; Sauer and Stolz, 1994; Carpaneto et al., 2005; Chen et al., 2012). SUTs, the first identified sucrose transporters, share common features with amino acid permeases (AAPs). Both SUTs and AAPs cotransport protons and sucrose (SUT1) or amino acids (AAPs) into cells at a stoichiometry of 1:1 (Fischer et al., 1995; Boorer et al., 1996a, 1996b). As proton symporters, their activity is determined by the proton motive force.

SWEETs are characterized by their ability to mediate bidirectional transport, their low-affinity for sugars, and pH-independency (Chen et al., 2012). In the absence of direct evidence, these characteristics are consistent with SWEETs functioning as uniporter, meaning that the concentration gradient of sugar determines whether flux is inward or outward. This is consistent with physiological observations, which support import in a few cases and efflux in many (see below).

Amino acids can occur as positively/negatively charged or neutral molecules. Although heterologous expression of UmamiTs revealed their bona fide ability to bidirectionally transport certain amino acids (Ladwig et al., 2012; Müller et al., 2015), physiological evidence is consistent with a uniport mechanism for UmamiTs. Thereby positively charged amino acids would show a tendency to be taken up into the cell, while negatively charged amino acids would rather exit the cell—even against a concentration gradient. Other transport mechanisms, such as a proton symport-coupled uptake (as used by AAPs) or proton antiport-mediated export of amino acids would be needed to allow the transport of amino acids against their (electro)chemical gradient. Based on the YddG crystal structure, a unique alternating-access transport mechanism was proposed, characterized by bending motions of transmembrane segments 3, 4, and 9 (Tsuchiya et al., 2016).

Roles for SWEETs in phloem loading

The identification of SUT sucrose/H+ symporters and the demonstration that SUT1 homologs were essential for phloem loading in potato (Solanum tuberosum), tobacco (Nicotiana tabacum), Arabidopsis, and maize (Zea mays) implicated a yet unidentified mechanism for sucrose export for cells along the path from synthesis in the mesophyll to the sites of loading at the sieve element-companion cell complex (SECC; Riesmeier et al., 1992, 1993, 1994; Bürkle et al., 1998; Gottwald et al., 2000; Slewinski et al., 2009). Although ample evidence had been assembled for the existence of the sucrose/H+ symporters and their role in importing sucrose into the SECC, essentially nothing was known about proteins involved in the efflux of sucrose and their location in the leaf. Various quantitative studies on the distribution of plasmodesmata had implied the interface between the SECC and the adjacent phloem parenchyma (PP) inside the phloem as the apoplasmic transfer site.

Arabidopsis

SWEET11 and 12 were shown to be expressed in specific phloem cells, most likely the PP (Chen et al., 2012). Single-cell RNA-sequencing (scRNA-seq), in combination with confocal microscopy enabled unambiguous assignment of SWEET11 and SWEET12 to the PP (Kim et al., 2021; Figure 3A). The phenotype of T-DNA mutants was consistent with the role of these two SWEETs in sucrose efflux from PP (Chen et al., 2012). The resulting model for phloem loading assumes that sucrose produced by photosynthesis in mesophyll cells (MCs) is transported to PP through plasmodesmata. Sucrose is then exported into the apoplasm by SWEET11 and SWEET12. Sucrose is then taken up actively into the SECC by SUT1/SUC2, energized by H+-ATPases (Figure 4A). However, there are some caveats to this model—if apoplasmic transport functioned as the exclusive path, ablation of key members is expected to be lethal. However, suc2 and sweet11;12 double mutants are both viable and fertile (Srivastava et al., 2008; Chen et al., 2012). Therefore, it is likely that other transporters or other routes exist. Notably, mRNA levels of SWEET13 was increased in sweet11;12 mutants (Chen et al., 2012). The presence of SWEET13 transcripts in the same cells as SWEET11 and 12 indicates additive activities (Figure 3A; Kim et al., 2021). Likely, distinct routes coexist, possibly using plasmodesmata.

Figure 3.

Figure 3

Cell type-specific transcript enrichment of SWEET and UmamiT family genes. A, Dot plot showing transcript enrichment of SWEET and UmamiT family genes across 19 clusters of the leaf scRNA-seq data (Kim et al., 2021). The diameter of the dot indicates the percentage of cells in the cluster in which transcripts for that gene were detected, while the color of each dot represents the average log-scaled expression of each gene across all cells within a given cluster. Cell types assigned to each cluster are indicated in the upper panel. Note that only SWEET and UmamiT family transcripts detected in the leaf scRNA-seq dataset were included in the plot. For detailed information about the dataset and description of the clusters/subclusters, refer to Kim et al. (2021). B, Violin plots illustrating the transcript enrichment of clade VI UmamiT38, UmamiT41, UmamiT45, UmamiT44, UmamiT46, and UmamiT47 in the subclusters of cluster 4. C, UmamiT45-coexpressed genes related to amino acid transport. Coexpression data were obtained from the ATTED-II coexpression database (http://atted.jp). The logit score (MR, mutual rank; Obayashi et al., 2014) for the UmamiT45-coexpressed UmamiTs, AAP family protein PUT1 (At1g31820), AAP1 (At1g58360), cationic amino acid transporters CAT4 (At3g03720), CAT9 (At1g05940), and CAT2 (At1g58030) is indicated. The logit score for UmamiT45 was 14.2.

Figure 4.

Figure 4

The role of SWEETs and UmamiTs in Arabidopsis, rice, and maize. A, The role of SWEETs and UmamiTs during seed filling (upper panel), nectar secretion (middle panel), and phloem loading in Arabidopsis (lower panel). Tissues of the seed (marked with distinct colors) are shown in the top left illustration. The same colors refer to tissues depicted in the panel. Note that the spatial distribution of SWEETs and UmamiTs changes dynamically during seed development. The schematic presented here illustrates early developmental stages (heart stage). B, The role of SWEETs in seed filling (upper panel) and phloem loading in maize (lower panel). C, The role of rice SWEETs in seed filling (upper panel) and pathogen growth (lower panel). Arrows indicate the direction of sugar or amino acid flow. OI, outer integument; II, inner integument; MCE, micropylar endosperm; EN, endosperm; EM, embryo; SE, sieve element; NP, nectary parenchyma; GC, guard cell; BS, bundle sheath; BETL, basal endosperm transfer layer; abBS, abaxial bundle sheath; VP, vascular parenchyma; NE, nucellar epidermis; NP, nucellar projection. Figure was created with Biorender.

Maize

Although the leaf morphology of C4 monocotyledous species with Kranz anatomy is distinct, maize uses a homolog of the Arabidopsis SUT1/SUC2—named ZmSUT1 for phloem loading as well (Slewinski et al., 2009). Phylogenetically, ZmSUT1 does not belong to the same dicot branch but fulfills the same function of supplying the SECC with sucrose. In maize, the closely related SWEET13s (SWEET13a, b, and c) are among the genes with the highest transcript levels in leaves (Emms et al., 2016; Bezrutczyk et al., 2018a). Knock-out mutants show symptoms of severe phloem loading defects. A combination of scRNA-seq, in situ hybridization, and translational GUS fusions indicates that all three SWEET13s are preferentially expressed in the two abaxial bundle sheath cells of the rank-2 intermediate veins (cells responsible for sucrose export from leaves; Fritz et al., 1983, 1989; Bezrutczyk et al., 2021). As sut1 and sweet13a;b;c triple mutants are viable and fertile, maize also appears to use additional phloem loading pathways or compensate in yet unknown ways (Botha, 2013; Figure 4B).

Rice

It was tempting to hypothesize that rice (Oryza sativa) homologs of ZmSUT1 and ZmSWEET13 would be key players for phloem loading in rice. However, neither mutants in OsSUT1 nor in the closest OsSWEET13 showed symptoms of phloem loading defects (Eom et al., 2012, 2019). Thus, despite the phylogenetic relationship among grasses, distinct phloem loading mechanisms seem to exist. Interestingly, another member of the SUT family, namely OsSUT2, might provide the driving force for phloem loading via symplasmic mechanisms (Eom et al., 2012). OsSUT2, as a vacuolar sucrose/H+ symporter, could be responsible for extremely high levels of sucrose in the cytosol of MCs, which could enable diffusion down a concentration gradient to sieve elements via plasmodesmata. Since sucrose concentration is estimated to reach almost 600 mM in rice (Hayashi and Chino, 1990), very higher sucrose concentrations would be needed for efficient translocation.

Physiological roles of UmamiTs

Transcripts from most UmamiTs belonging to clade VI (Supplemental Figure S2) were broadly detected in cells from almost all cell types of the leaf, with an apparent preferential accumulation in cells from the bundle sheath/xylem cells, overlapping with transcripts from other amino acid transporters (Figure 3, A–C). More cell type-specific expression was observed for AtUmamit5/WAT1 in epidermal cells, whereas transcripts of AtUmamiT10, 27, and 31 were almost exclusive for guard cells (Figure 3A), pointing toward stomatal functions. As cell types in the leaf have distinct metabolic activities reflected by the differential transcript level of metabolic pathway genes (Kim et al., 2021), it will be interesting to assess the substrate specificity and the role of these UmamiTs in respect to the cell types where they are enriched (Box 1). More detailed analyses were performed using reporters for a number of UmamiT family members, yielding evidence for roles in phloem loading and seed filling.

Box 1.

single-cell RNA-seq as an approach to identify additional transporters

Single-cell transcriptomics is a rapidly evolving field that enables profiling transcriptomes of individual cells derived from complex organs. This revolutionary technology was made possible by capturing individual cells and sequencing low amounts of RNA. A major goal of single-cell transcriptomic studies is to obtain transcriptome signatures of individual cells and cluster distinct cell types (or states) within complex tissues and associate these transcriptomic cellular states with the functional state of each cell type. scRNA-seq has been applied to different tissues of diverse plant species (Denyer et al., 2019; Jean-Baptiste et al., 2019; Ryu et al., 2019; Shulse et al., 2019; Zhang et al., 2019; Liu et al., 2020, 2020; Wendrich et al., 2020; Bezrutczyk et al., 2021; Kim et al., 2021). These studies identified the transcriptomes of major cell types or subtypes (or states) that were not previously well defined, for example, vascular cell types in different developmental states.

Recently, computational pipelines were established to characterize the metabolic heterogeneity with cell type resolution based on single-cell data (Xiao et al., 2019; Box 1, Figures A and B). Using these approaches, the activity of metabolic pathways can be assessed at the cell type level (Xiao et al., 2019; Kim et al., 2021).

Panel A shows a Uniform Manifold Approximation and Projection (UMAP) dimensional reduction projection of transcriptome profiles from 5,230 Arabidopsis leaf cells grouped into distinct clusters (Kim et al., 2021). Each dot represents an individual cell colored according to cell type. Panel B shows a UMAP plot filtered for metabolic genes (same dataset), demonstrating that clustering patterns are retained but shifted, intimating distinct expression patterns for metabolic genes (Kim et al., 2021). This approach may help identify transporters, as, depending on the substrate, transporters likely correlate with the respective metabolic pathways of that cell.Inline graphic

Figure: Landscape of metabolic gene expression in single leaf cells. A, UMAP plot of metabolic gene expression profiles of leaf cells from the Arabidopsis leaf scRNA-seq dataset (Kim et al., 2021). UMAP was used for visualization by reducing data to two-dimensions. Each dot indicates one cell; colors indicate cell type described in the legend (BS1, bundle sheath; BS2, bundle sheath cells enriched with photosynthetic processes; XP1 and XP2, xylem cells related with the bundle sheath; XP3, xylem cells enriched with vascular parenchyma markers; PCXP, procambium cells related to XP1; PCPP, procambium cells related to PP cells with transfer cell identity (PP1); CC, companion cells, u.a., unassigned). B, UMAP plot of metabolic gene expression profiles of same dataset as in (A).

Roles of UmamiTs in phloem loading

Amino acids are the main transport forms of organic nitrogen in the phloem of most plants. We may therefore hypothesize to find similar pairs of transporters analogous to the SWEET-SUT pair for organic nitrogen. In accordance with a role in amino acid efflux from PP in Arabidopsis, transcripts of multiple UmamiTs were enriched in the same cells as transcripts from SWEET11, 12, and 13. Six out of seven PP-specific UmamiTs are coexpressed with each other and with SWEET11 and 12 (Kim et al., 2021; Figure 3A). Similar to SWEET11 and 12, UmamiT18/SIAR1 is expressed both in the PP of leaves and in seeds (Ladwig et al., 2012; Kim et al., 2021). Mutants show reduced amino acid accumulation in seeds, possibly implicating UmamiT18 in transport processes in phloem loading as well as seed filling (Ladwig et al., 2012; Kim et al., 2021). These similarities to sucrose transport are striking, since also in this case multiple SWEETs contribute to phloem loading, and SWEETs have dual roles in phloem loading and seed filling. The large number of UmamiTs that are coexpressed in PP, may indicate that they are needed to maximize flux and cover the diverse set of substrates. Transcripts of AAP2, AAP4, and AAP5 amino acid H+/symporters are enriched in companion cells (CCs), and may thus play analogous roles for importing amino acids as SUTs do for sucrose (Kim et al., 2021). Notably, the amino acid metabolism in PP and CC is very different, indicating that metabolic activities shape the amino acid composition of the phloem sap (Kim et al., 2021; Box 1). Further characterization of UmamiTs and cell specific metabolism will be useful to understand the regulation of amino acid allocation.

Roles for SWEETs and UmamiTs in phloem unloading

The partitioning of sucrose and amino acid is strongly dependent on phloem loading in the source regions and unloading in the sink regions of the plant. Several SWEETs and UmamiTs from various species are known to be expressed in different sink tissues (Figure 4; e.g. Kryvoruchko et al., 2016; Zhen et al., 2018; Jeena et al., 2019; Wang et al., 2019; Ren et al., 2021). Impaired activity of SWEETs or UmamiTs causes defects in basic physiological processes reflecting their broad role. Additional roles of SWEETs and UmamiTs in pathogen susceptibility have also been reported in numerous studies and are summarized in Box 2 (Buell et al., 2003; van Damme et al., 2009; Chen et al., 2010; Ranocha et al., 2010; Smeekens et al., 2010; Denancé et al., 2013; Zeier, 2013; Hahn et al., 2014; Schwelm et al., 2015; Struck, 2015; Bezrutczyk et al., 2018b; Besnard et al., 2021; Prior et al., 2021).

Box 2.

SWEETs and UmamiTs: roles in pathogen susceptibility

Upon identification of the SWEETs, their involvement in pathogen susceptibility was promptly reported, that is, upregulation of OsSWEETs during Xanthomonas oryzae pv oryzae (Xoo) infection. Later, the bacterial blight resistance-conferring locus Xa13 was shown to correspond to OsSWEET11 (originally called Os8N3), and xa25 and xa41 to OsSWEET13 and OsSWEET14, respectively. Noteworthy, only sucrose-transporting clade III SWEETs function as susceptibility genes. Induction of these SWEETs occurs by binding to their promoters of Xoo-secreted TAL (transcriptional activator-like proteins) effectors (Figure 4C). Interestingly, North American Xoo isolates lack TAL effectors and are weak virulence-inducing pathogens compared to African/Asian strains. Therefore, the ability to induce SWEETs was likely crucial for turning Xoo into a “super pathogen.” SWEETs from several plant species are induced by diverse pathogens, including biotrophic bacteria, oomycetes, and fungi (Chen et al., 2010; Smeekens et al., 2010; Bezrutczyk et al., 2018b). As the genomes of various pathogens lack TAL effector homologs (Buell et al., 2003; Hahn et al., 2014; Schwelm et al., 2015), these pathogens possibly use alternative SWEET induction mechanisms. In Arabidopsis, Pseudomonas syringae (Pst DC3000) presumably activates a bZIP transcription factor that eventually induces SWEETs and UmamiTs (Prior et al., 2021). Pathogens likely activate SWEETs to gain access to carbon skeletons, energy, and nutrients for efficient reproduction.

Changes in the free amino acid pool composition/homeostasis affect plant defense responses (van Damme et al., 2009; Zeier, 2013) and pathogen nutrition (Struck, 2015). In addition, the expression of genes coding for amino acid transporters was altered upon pathogen infection (Pratelli and Pilot, 2014).

Implication of UmamiTs in pathogen susceptibility was shown repeatedly and is mediated by increased salicylic acid (SA) levels. For instance, auxin-transporting AtUmamiT5/WAT1 is required for secondary cell-wall deposition (Ranocha et al., 2010), and wat1 mutants conferred broad-spectrum resistances against vascular pathogens, likely due to reduced root auxin levels (Denancé et al., 2013).

Likewise, AtUmamiT36/RTP1 (RESISTANCE TO PHYTOPHTHORA PARASITICA 1) mutants displayed increased Pst DC3000 and G. cichoracearum resistance (Pan et al., 2016). Enhanced pathogen resistance upon overexpression of a MtN21-like amino acid transporter was shown for UmamiT14.

Plants overexpressing UmamiT14 displayed increased SA-levels and enhanced resistance toward the biotrophic Hyaloperonospora arabidopsidis, likely due to a constitutive immune response (Besnard et al., 2021). Together, these results suggest that misregulation (i.e. up-/downregulation) of UmamiTs and subsequent altered amino acid accumulation/composition can trigger enhanced pathogen resistance, albeit without clear correlations to changes in specific amino acids.

Roles for SWEETs in seed filling

Growth and development of the embryo depends on adequate supply with photoassimilates from maternal tissues. The unfertilized ovule is symplasmically isolated from the maternal tissues before fertilization. Post fertilization, a drastic increase of plasmodesmata can be observed between the terminal sieve element and neighboring cells at the chalazal region (the seed nutrient unloading zone) forming a symplasmically connected unloading domain (ULD). Along the unloading path, the unloading zone and the integuments, the layers between outer and inner integuments, and endosperm and the embryo are symplasmically isolated (Thorne, 1985; Stadler et al., 2005; Werner et al., 2011). Consequently, sugars and amino acids must be exported from one cell into the apoplasm before then can be reimported in the adjacent cell.

One of the most elegant systems for studying metabolite efflux is the “empty seed technique” established in legumes (Wolswinkel and Ammerlaan, 1983; Thorne, 1985; Fieuw and Patrick, 1993; Walker et al., 2000). In this technique, the embryo is surgically removed and the seed coat is filled with solutions known to influence assimilate transport. By assaying the contents of the solution, it is possible to study transport processes involved in the release of assimilates to the developing seed. The studies revealed that sucrose efflux from the seed coat occurs, in part, by sucrose/H+ antiport (in part mediated by SUT; Baud et al., 2005; Zhang et al., 2007), as well as by proton gradient-independent mechanisms (Walker et al., 1995; De Jong et al., 1996). Consistent with their function as uniporters, SWEET members could contribute to the proton-independent efflux from the seed coat. The cell type specificity of SWEET4, 11, 12, 15 is developmentally regulated and each of them likely contributed to sucrose transfer across the different symplasmic barriers (Chen et al., 2015c; Lu et al., 2020). Triple sweet11;12;15 mutants accumulated starch in the seed coat and showed severe defects in seed development, implicating important roles in sucrose efflux at distinct steps in seed filling (Chen et al., 2015c; Figure 4A). However, as sweet11;12;15 mutants were viable, additional transport mechanisms are likely to exist. It is likely that a yet to be identified sucrose/H+ antiporters may be responsible for this function (Fieuw and Patrick, 1993; Walker et al., 1995).

The evidence for SWEETs in feeding tissues in the developing seed is not limited to Arabidopsis, but also exists in crops. In rice, OsSWEET11 and 15 are essential for transporting sugar through distinct apoplasmic pathways (Figure 4C; Ma et al., 2017; Yang et al., 2018). In maize, ZmSWEET4c, ZmSWEET11, and ZmSWEET15b were found to be localized at different stages of seed development (Sekhon et al., 2011; Li et al., 2014; Sosso et al., 2015). ZmSWEET4c is likely involved in translocating cell wall invertase-derived hexoses in and/or across the basal endosperm transfer layer (BETL), a cell layer of endosperm characterized by cell wall invaginations that amplify the plasma membrane surface area (Sosso et al., 2015; Figure 4B). Interestingly, ZmSWEET4c may be a target of domestication that was likely recruited by farmers and breeders who selected for large grains. Although substantial progress has been made, the full path of sucrose in none of the species has been unraveled.

Roles for UmamiTs in amino acid supply to seeds

The transport of amino acids and sucrose shares commonalities since both processes must undergo similar symplasmic and apoplasmic steps. Uptake of amino acids into the embryo has been shown to occur via H+/amino acid symporters such as the AAPs, while efflux processes are mediated by proton gradient-independent, transporter-mediated mechanisms (Lanfermeijer et al., 1990; de Jong et al., 1997; Tegeder et al., 2000; Sanders et al., 2009; Zhang et al., 2015; Karmann et al., 2018). Analogous to the roles of several SWEETs, multiple UmamiTs are implied in proton-independent efflux of amino acids to supply the developing seed (Karmann et al., 2018; Figure 4A). Interestingly, several UmamiTs localize to the ULD where symplasmic transport through plasmodesmata is considered as the dominant route (Stadler et al., 2005). In the plasma membrane of the ULD, UmamiT11, 14, 18, and 24 have been implicated in the export of amino acid to the developing embryo (Figure 4A;Table 3). In early seed development, UmamiT11 and 14 are present in cells at the end of the funicular vasculature, which are adjacent to the protoxylem, CCs, and sieve elements. In accordance with a disrupted export of amino acids from the chalazal zone, seeds of umamit11 and 14 single mutants were smaller (Müller et al., 2015). Only at later stages (torpedo stage), UmamiT28 was detected in the cellularizing endosperm and the endothelium layer of the inner integuments (Müller et al., 2015). Up to the late torpedo stage of embryo development, UmamiT29 was found in the middle layer of the inner integument, followed by a shift in localization to the inner layer of the outer integuments in later developmental stages (Müller et al., 2015). Mutants of either UmamiT28 or 29 produced smaller seeds with a trend to accumulate amino acids (Müller et al., 2015). Contrarily to the tonoplast-localized UmamiT24, which might be involved in temporary amino acid storage, the plasma membrane-localized UmamiT25 presumably mediates amino acid export from the endosperm (Besnard et al., 2018). Mutations in either gene resulted in reduced seed amino acid content, likely due to reduced uptake (Besnard et al., 2018). Taken together, the developmentally controlled differential expression of UmamiTs across several symplasmic seed tissues is suggestive of their roles in transfer of amino acid export from maternal to filial tissues. The relatively high number and overlapping expression patterns of UmamiTs involved in amino acid transport in seeds point toward redundant functions for at least some of the proteins (Müller et al., 2015). However, the full path of amino acid translocation remains to be unraveled.

Roles for SWEETs in pollen nutrition

Pollen germination and tube growth initially rely on nutrient storage in the pollen grain. Pollen grains, pollen tubes, and the anther tapetum are sink tissues that are symplasmically isolated, requiring an unloading pathway through the apoplasmic space. In Arabidopsis, AtSWEET8 and AtSWEET13, also known as RUPTURED POLLEN GRAIN (RPG) 1 and 2, respectively, were suggested to function in the efflux of sugar in the tapetum and microspores for pollen cell wall synthesis (Guan et al., 2008; Sun et al., 2013; Kanno et al., 2016). Mutations in SWEET8 and SWEET13 resulted in pollen cell wall defects and reduced male fertility. AtSWEET13 could partially rescue the defective pollen phenotype of atsweet8, suggesting functional redundancy (Sun et al., 2013). In coniferous Wilson’s spruce (Picea wilsonii), PwSWEET1 was implied in supplying glucose for proper pollen germination and pollen tube growth (Zhou et al., 2020). In rice, OsSWEET11 has been implicated in the export of sugars during pollen development as OsSWEET11-silenced plants showed low fertility and pollen viability (Yang et al., 2006; Eom et al., 2015).

Roles for SWEETs in nectar secretion

To attract and reward pollinators, plants secrete nectar—a sugar-rich solution, which contains volatile compounds produced in the nectary. The mechanism for nectar secretion was reported through a study using Arabidopsis, turnip (Brassica rapa), and coyote tobacco (Nicotiana attenuate; Lin et al., 2014). Arabidopsis SWEET9 is highly expressed in the nectary and similar as its homologs in N. attenuata and B. rapa show sucrose uniport activity. Loss of AtSWEET9 resulted impaired nectar secretion. The current model proposes that sucrose synthesized in the nectary parenchyma cells is secreted via AtSWEET9 into the apoplasm and hydrolyzed by cell wall invertases, which cause an osmotic gradient to sustain water secretion (Figure 4A). However, as secreted nectars require a fast and active secretion, it likely cannot solely be mediated by uniport. How AtSWEET9, as a uniporter, can secrete sugar to high levels is still a conundrum and whether other mechanisms are involved in the process remains to be elucidated. In petunia, NEC1, the homolog of SWEET9, is highly expressed in nectaries and likely plays a similar role (Ge et al., 2000).

SWEETs in vacuolar transport

The vacuole occupies more than 80% of the plant cell volume and is separated from the cytosol by a semi-permeable membrane, the tonoplast. The vacuole is the primary compartment for maintaining cellular homeostasis, turgor pressure, detoxification, and importantly, storage of sugars. After the identification of the first tonoplast-localized monosaccharide transporter (Wormit et al., 2006), multiple vacuolar transporters mediating transport of sugars by facilitated diffusion and active transport have been described (e.g. Aluri and Buttner, 2007; Eom et al., 2011; Payyavula et al., 2011; Poschet et al., 2011; Schulz et al., 2011; Pommerrenig et al., 2018). Clade IV SWEETs have been shown to be responsible for the efflux of fructose, glucose, and sucrose from the vacuole (Jeena et al., 2019). AtSWEET17, the first characterized vacuolar fructose transporter, was implied in determining leaf fructose content under normal and stress conditions (Chardon et al., 2013). AtSWEET17 and its close paralog AtSWEET16 were also shown to be highly expressed in the root vacuoles. Mutations and overexpression of SWEET16 and SWEET17 resulted in various growth phenotypes under normal and abiotic stress conditions, reflecting the vital role of vacuolar SWEET-mediated sugar efflux in the development and stress tolerance of plants (Chardon et al., 2013; Klemens et al., 2013; Guo et al., 2014). AtSWEET2, another tonoplast-enriched clade I SWEET, has been hypothesized to prevent sugar loss from roots (Chen et al., 2015a). AtSWEET2 was also shown to be induced during Pythium infection. As sweet2 mutants were susceptible to the oomycete, it has been predicted that AtSWEET2 modulates sugar secretion to limit carbon loss to the rhizosphere. A detailed description of the role of SWEETs is presented in Table 2.

Nutrition of symbiota and microbiota

The observation that SWEETs and UmamiTs were originally identified as nodulins may indicate possible roles in symbiosis (Box 3). Besides the highly evolved symbiotic system in legumes, many plants are colonized by mycorrhiza and rhizobials, and all plants interact closely with endo- and ectophytic microbiota. Notably, plants are thought to secrete 15%–40% of their photosynthate into the soil, presumably for feeding root-colonizing microbiota (Lynch and Whipps, 1990). Although the composition of exudates is influenced by various factors, primary metabolites including sugars, amino acids, and organic acids, are secreted in larger quantities than secondary metabolites (Badri et al., 2008). For instance, in maize, sugars constitute 64% of the root exudate, whereas amino acids and low molecular weight organic acids (LMWOAs) represent 22% and 14%, respectively (Hütsch et al., 2002).

Box 3.

Symbiosis as a basis for microbiota establishment

The “symbiosis cascade effect” hypothesis describes how plant symbiosis establishment influences and subsequently drives the assembly plant root microbiota (Uroz et al., 2019). The establishment of symbiosis affects intra- and intercellular communication, transcriptional reprogramming, rerouting of metabolite signaling pathways, hence, root exudate composition and root architecture. In turn, this influences the establishment and/or modifies the microbial community structure (Uroz et al., 2019).

Root AMF and nitrogen-fixing/nodulation symbiosis affects the root microbiome. The root microbial community structure of L. japonicus mutants impaired in nodulation or symbiosis is similar among diverse mutants but distinct from that of the wild-type. The altered community structure was retained even in nitrogen-supplemented soil where nodulation is prevented in wild-type (Zgadzaj et al., 2016). Similarly, M. truncatula mutants impaired in nodulation and/or AMF symbioses assemble more similar root bacterial communities compared to wild-type (Offre et al., 2007). AMF and nodulation symbiosis play important roles in structuring the root fungal and bacterial communities, and disruption of symbiosis causes major shifts in bacterial and fungal assembly (Thiergart et al., 2019; Wang et al., 2020).

Numerous SWEETs across different plant families were shown to be induced during AMF and rhizobial symbiosis (Fiorilli et al., 2015; Manck-Götzenberger and Requena, 2016), indicating an evolutionarily conserved role for some SWEETs in symbiosis. Although the exact role of SWEETs in microbiotal sugar feeding remains yet unclear, their roles in AMF and rhizobia symbiosis are a compelling basis to speculate about a role for sugar transporters in shaping the root microbiota. It should be noted that while evidence for symbioses as a basis for root microbiota establishment exists, there are differences between symbiosis and root microbiota formation (Sasse et al., 2018).

Sugars: SWEETs

Plants grown in a rhizospheric microbial culture showed significantly less sugar reuptake compared to cultures in sterile conditions (Kuzyakov and Jones, 2006). Considering the broad roles of SWEETs in sugar secretion, it is conceivable that they also play a role in sugar efflux from roots. Unsurprisingly, arbuscular mycorrhizal fungus (AMF) and nitrogen-fixing rhizobia induce SWEETs across different plant families (Fiorilli et al., 2015; Manck-Götzenberger and Requena, 2016). Studies in potato roots revealed that colonization by the AMF Rhizophagus irregularis affected steady state mRNA levels of 22 of 35 StSWEETs. StSWEET2c, StSWEET7a, and StSWEET12a showed the highest induction in arbuscule-containing cells (Manck-Götzenberger and Requena, 2016). Transcriptome profiles from AMF-colonized soybean (Glycine max) revealed induction of GmSWEET6 and GmSWEET15 in arbuscule-containing cells (Manck-Götzenberger and Requena, 2016). In Medicago, the glucose transporter MtSWEET1b/SWEET1.2 was induced in arbuscule-containing cells where it localized to peri-arbuscular membranes, presumably to facilitate nutrient exchange between host plant and AMF during symbiosis (An et al., 2019). SWEETs have also been implicated in legume-rhizobium symbiosis. In lotus (Lotus japonicus), SWEET3 is preferentially expressed in nodules infected by Mesorhizobium loti, indicating a role in sugar transfer toward rhizobia (Pini et al., 2017; SugiyaMa et al., 2017). LjSWEET3 was induced also by AMF, hinting at common regulatory mechanisms for AMF and rhizobial symbiosis, that could influence the assembly of root microbiota (Zgadzaj et al., 2016; SugiyaMa et al., 2017; Uroz et al., 2019). In Sinorhizobium meliloti-infected Medicago, MtSWEET11 translocated from plasma membranes to transcellular infection threads and symbiosomes (Table 4; Figure 5) presumably mediating efflux of sugar into the symbiosomes (Kryvoruchko et al., 2016). A summary of the role of SWEETs in symbiosis and microbiota-feeding is presented in Table 4 and Figure 5.

Table 4.

Roles of root exudates and their efflux transporters in plant root–microbiota interactions

Role in Plant–Microbe Interactions
Reference(s)
Root Exudate Attributed Transporter (Substrate)
Sugar AMF and rhizobia symbiosis mutants are incapable of assembling normal root and rhizosphere microbiota GmSWEET6, MtSWEET1b (glu), StSWEET2c, StSWEET7a, StSWEET12a Localized in arbuscules-containing cells, indicated to function in supplying sugar to AMFs Manck-Götzenberger and Requena, 2016; An et al., 2019; Zhao et al., 2019
LjSWEET3, MtSWEET11 (suc) Induced in nodules, transfection threads, and symbiosomes induced by nitrogen-fixing and non-nitrogen fixing rhizobia Offre et al., 2007; Zgadzaj et al., 2016; SugiyaMa et al., 2017; Thiergart et al., 2019; Wang et al., 2020

Amino acid Microbial-derived products increase amino acids efflux from roots of plants grown under hydroponic conditions AtUmamiT14, AtUmamiT18 (gln) umamit14 and umamit18 show decreased amino acid export from roots Besnard et al., 2016

LMWOAs Exogenous application of LMWOAs present in rhizospheric exudates result in selection for growth-promoting taxa, and stimulation of soil microbial activities AtALMT1 (malate) Overexpression of AtALMT1 and induction by MAMPs recruits beneficial rhizobacteria that induce plant immunity Rudrappa et al., 2008; Lakshmanan et al., 2012; Kobayashi et al., 2013; Macias-Benitez et al., 2020
LjALMT4 (malate) Specifically expressed in nodule vasculature bundle, and involved in bidirectional transport of malate in nodules Takanashi, 2016
TaALMT1 Taalmt1 assembles differentially enriched root bacterial OTUs compared to wild-type Mahoney, 2017
Citrate-supplemented soil causes decrease in the richness and diversity of the bacterial communities MtMATE67 (citrate) Induced by S. meliloti and localized to nodules and symbiosomes. Transports citrate to increase Fe(III) availability for rhizobia Kryvoruchko et al., 2018
LjMATE1 (citrate) Induced by rhizobia and AMF. Supports nodule function by providing citrate iron translocation to nodules Takanashi et al., 2013; Handa et al., 2015

Others AtPDR2 Increased phenolic compounds and reduced sugar content in root exudates of atpdr2 increases relative abundance of beneficial bacterial OTUs Badri et al., 2008;, 2009
N.D. PaPDR1 (strigolactones) Root exudates of mutants show reduced AMF hyphal branching and promotes parasitic seed germination. A. thaliana overexpressing PaPDR1 increased secretion of synthetic strigolactone analog Kretzschmar et al., 2012; Sasse et al., 2015; De Cuyper and Goormachtig, 2017
N.D. AtPDR8 (scopoletin), AtPDR9 (coumarin) Atpdr9 mutants assemble different root-associated microbiota compared to wild-type in Fe-limiting conditions. atpdr9 is incapable of microbiota-mediated plant growth rescue Fourcroy et al., 2016; Voges et al., 2019; Harbort et al., 2020
Diterpenes: Diterpenoid-deficient maize mutant shows altered rhizospheric microbiota, largely attributed to deficiency of diterpene in roots NtPDR1 NtPDR1 overexpressor indicated a role for NtPDR1 in efflux of diterpenes from roots in response to biotic triggers Crouzet et al., 2013; Murphy et al., 2021
Triterpenes: Triterpene biosynthesis mutants/analysis and exogenous triterpene treatment indicated roles in regulating bacterial growth N.D. N.D. Huang et al., 2019
Isoflavonoid: Essential for plant-rhizobium symbiosis. Simulated exudation of the most abundant rhizobia-inducing flavonoids shows interaction with diverse soil bacteria MtABCG10 (isoflavonoid) Fungal elicitor treatment on RNAi silenced-MtABCG10 shows increase in isoflavonoids in root exudates. Increased susceptibility to pathogenic fungi in mutants. Subramanian et al., 2007; Banasiak et al., 2013; Poole et al., 2018

Abbreviations: ABCG, ATP-binding cassette transporter G; gln, glutamine; glu, glucose; MAMP, microbe-associated molecular pattern; OTU, operational taxonomic unit; PDR, pleiotropic drug resistance; suc, sucrose.

Figure 5.

Figure 5

Transporters with potential roles in root metabolite efflux for feeding microbiota and symbiota. Microbes actively recruited to the proximities of the root surface (rhizosphere) or colonizing the internal root tissues (endosphere) constitute the root microbiota. Malate secretion by ALMT1 during pathogen challenge recruits B. subtilis for induced systemic resistance. ALMT1 also mediates the transport of malate into rhizobia-symbiosomes. Microbial-derived products trigger increased amino acid efflux in the roots, likely via UmamiTs. Sugar exported via SWEETs to arbuscule containing cells (AMF symbiosis) or nodules (rhizobia symbiosis) serves to maintain favorable growth conditions for symbiosis. MATE transporters are involved in the efflux of citrate, which can be metabolized by microbes. AMF and rhizobacteria symbioses trigger the symbiosis cascade effect that could be a basis for the establishment of mutualistic interactions in the root. Substrates for the corresponding transport proteins are indicated (gln, glutamine; glu, glucose, suc, sucrose). Asterisks indicate putative substrates based on cross-reference to homologs. Figure was created with Biorender.

Amino acids: UmamiTs and GDU1

Dyshomeostasis in free amino acid pool impacts the defense response of plants, which in turn, could affect symbiosis (van Damme et al., 2009; Zeier, 2013; Pratelli and Pilot, 2014; Struck, 2015; Pan et al., 2016; Besnard et al., 2021; Box 2). In sterile conditions, amino acid reuptake into roots was higher compared to exudation (Phillips et al., 2004). Addition of microbial products significantly increased net amino acid exudation, likely due to effects on the relative activity of amino acid secretion and passive uptake by root microbes (Phillips et al., 2004). Mutants of AtUmamiT14 and AtUmamiT18 transferred lower amino acid amounts from roots to media, implicating the two UmamiTs in amino acid secretion toward the rhizosphere (Besnard et al., 2016). GLUTAMINE DUMPER 1 (GDU1), which is mainly expressed in the root vasculature is also postulated to play a role in amino acid secretion from roots as overexpressors efflux elevated amino acid amounts into media (Pratelli et al., 2010). The molecular function of GDU1, a single membrane spanning protein, is yet to be revealed. The role of UmamiTs and GDUs in amino acid exudation and reuptake in nonsterile soil-grown plants will be an important research target.

LMWOAs (malate and citrate): ALMTs and multidrug and toxic compound extrusion transporters

Under certain stress conditions and in the presence of soil microbes increased exudation of LMWOAs (mainly citrate and malate) has been observed (Boldt-Burisch et al., 2019). Interestingly, structure and activity of soil bacterial communities changed more dramatically upon treatment with LMWOA, compared to sugar treatment (Shi et al., 2011; Macias-Benitez et al., 2020). Although we do not understand which transporters are involved in LMWOA efflux for nutrition of microbiota, candidate transporters for malate and citrate come from three transporters superfamilies (Table 1). Malate efflux at the plasma membrane and import into vacuoles is mediated by aluminum-activated malate transporters (ALMT) and tonoplast dicarboxylate transporters (TDT), respectively (Igamberdiev and Eprintsev, 2016; Frei et al., 2018). The first ALMT was isolated from aluminum-tolerant wheat (Triticum aestivum), followed by identification of a functional ortholog in Arabidopsis, AtALMT1 (Sasaki et al., 2004; Hoekenga et al., 2006). Characterization of various ALMTs unveiled roles in a wide range of physiological processes, including plant–microbe interactions (Piñeros et al., 2008; Meyer et al., 2011; De Angeli et al., 2013; Ramesh et al., 2015). Pseudomonas infection, and in particular, bacterial elicitors and jasmonic acid signaling induce AtALMT1 in the absence of a low pH or an aluminum-rich environment, leading to root secretion of malate, thereby promoting colonization by beneficial rhizobacteria and conferring systemic immunity against subsequent pathogen infection (Table 4; Figure 5; Rudrappa et al., 2008; Lakshmanan et al., 2012; Berendsen et al., 2018). In wheat, lines carrying either wild-type or mutant TaALMT1 alleles showed differences in host-associated bacterial taxa, reminiscent of that caused by symbiosis disruptions (Thiergart et al., 2019; Wang et al., 2020). Whether such differences are attributed to malate secretion have not yet been tested (Mahoney et al., 2017). The malate efflux transporter LjALMT4 is highly expressed in M. loti-infected nodule junctions of young nodules and in vascular bundles of mature nodules (Figure 5). Interestingly, LjALMT4-mediated malate efflux was independent of aluminum ions (Takanashi et al., 2016).

The efflux of citrate is mediated by the multidrug and toxic compound extrusion (MATE) transporters, which were originally identified as energy-dependent efflux transporters in bacteria that confer drug resistance (Morita et al., 1998). Plant genomes carry a large number of MATE orthologs, also known as DETOXIFICATION (DTX) proteins, for example, 58 MATE members in Arabidopsis (Hvorup et al., 2003). Citrate-transporting MATEs were identified in barley (Hordeum vulgare Al-ACTIVATED CITRATE TRANSPORTER 1, HvAACT1), sorghum (Sorghum vulgare, SbMATE), Arabidopsis (AtMATE), rice (FERRIC REDUCTASE DEFECTIVE LIKE 1 and 4, OsFRDL1 and 4), rice bean (Vigna umbellate, VuMATE1, VuMATE2), and maize (ZmMATE1; Furukawa et al., 2007; Magalhaes et al., 2007; Yokosho et al., 2009; Maron et al., 2010; Yokosho et al., 2011; Liu et al., 2018). Although their direct involvement in microbe feeding requires more research, the ability of MATE transporters to efflux citrate, which can be metabolized by microbes, might imply such roles (Figure 5;Furukawa et al., 2007; Baetz and Martinoia, 2014). The iron-activated plasma membrane citrate efflux transporter MtMATE67 is induced by S. meliloti inoculation. MtMATE67 localizes to plasma membranes of symbiosomes, also known as the peribacteriod membrane, where it primarily transports citrate into the symbiosome to increase the availability of Fe (III) for rhizobia (Kryvoruchko et al., 2018). Similarly, the nodule-specific citrate transporter LjMATE1 supports nodule function by providing citrate for iron translocation to the nodule infection zone (Takanashi et al., 2013). Interestingly, LjMATE1 is also upregulated during AMF R. irregularis infection (Handa et al., 2015). Table 4 and Figure 5 summarize the current state for ALMTs, MATE, and other organic transporters (not mentioned in text, but reviewed in Sasse et al., 2018; Stassen et al., 2021) in plant–microbe symbiosis.

Concluding remarks and future perspectives

The pressure-flow hypothesis (Münch, 1930) is still the most widely accepted mechanism for long-distance phloem transport. However, despite many contributions since then, our current understanding of assimilate allocation is still limited. The phloem sap contains many hundreds of metabolites (Fiehn, 2003), and for most we have no clue how they are transported (see “Outstanding questions”). For instance, malate is an important constituent of the phloem sap which affects nitrate uptake by roots (Touraine et al., 1992). However, a bona fide malate transporter that effluxes malate from the PP or imports malate by proton symport into the SECC remains elusive. It is conceivable that some members of the dicarboxylate transporter family are involved in malate secretion from the PP (Taniguchi et al., 2002). However, selecting candidates for transport assays and verifying their physiological roles is challenging. As substrates which enter and leave the cells are largely dependent on the distribution and function of plasmodesmata and transmembrane transporters, transcriptomic profiles at the single-cell level serves as a reliable map for selecting transporter candidates for a wide range of metabolites, hormones, and even ions (Box 1; e.g. Denyer et al., 2019; Jean-Baptiste et al., 2019; Ryu et al., 2019; Shulse et al., 2019; Xiao et al., 2019; Zhang et al., 2019; Liu et al., 2020; Wendrich et al., 2020; Bezrutczyk et al., 2021; Kim et al., 2021). Single-cell analysis may also serve as tool to dissect the integration between plants and the pathogenic (or commensal microorganisms). Several SWEETs in rice have been implicated in disease susceptibility to Xanthomonas oryzae pv. oryzae (Chen et al., 2010; Smeekens et al., 2010; Bezrutczyk et al., 2018a,b; Box 2), yet our current knowledge on other efflux transporters that may function as susceptibility factors is limited. Although our current technology is limited in capturing eukaryotic RNAs with polyA tails, development for simultaneously capturing eukaryotic and prokaryotic transcripts at once, and performing a combined single-cell and spatial transcriptomics, will provide insight on host cell–pathogen interaction at the cell type and spatial resolution. Importantly, the development of genetically encoded biosensors targeted to various cellular compartments will empower dissection of the mechanisms for distribution and fluxes of different nutrients. We expect that these technologies will allow us to better understand symbiosis establishment, plant–pathogen interaction, and enable us to systematically engineer nutrient flux in plants to increase crop yield in the future.

Supplemental data

The following materials are available in the online version of this article.

Supplemental Figure S1 . Phylogenetic analysis of SWEET family proteins of Arabidopsis (At), Oryza sativa (Os, rice), and Zea mays (Zm, maize).

Supplemental Figure S2 . The phylogeny of 46 members of the A. thaliana UmamiT family.

Supplemental Table S1 . Gene IDs of SWEETs used for phylogenetic trees.

Supplemental Table S2 . Gene IDs of UmamiTs used for phylogenetic trees.

Outstanding questions

  • When/how do sugars/amino acids cross plasmodesmata?

  • How are symplasmic and apoplasmic pathways coordinated?

  • How is the efflux from one cell coordinated with capacity for uptake by adjacent cells?

  • How is the demand-supply of sugars/amino acids in sink-source tissues coordinated?

  • How are SWEETs/UmamiTs dynamically regulated during development/upon environmental cues?

  • How are import and export processes coordinated during seed filling? By which transporters—especially the unloading of sucrose/amino acids in the unloading domain? What is the step-by-step path of sucrose/amino acids in model and crop plants?

  • Do SWEETs/efflux transporters play roles in pathogen susceptibility to other diseases?

  • Are there clade-specific UmamiT substrate and/or localization patterns?

  • How is malate loaded into the phloem?

  • Which transporters are involved in feeding microbiota? How are their activities controlled? How can pathogen feeding be prevented?

  • Do other efflux transporters serve as host susceptibility factors for pathogens?

Funding

This research was supported by Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy—EXC-2048/1—Project ID 390686111, Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) SFB 1208—Project-ID 267205415, the Alexander von Humboldt Professorship, the National Science Foundation (SECRETome Project: Systematic Evaluation of CellulaR ExporT from plant cells, IOS-1546879) and the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement no. 951292, Sympore) to W.B.F.

Conflict of interest statement. None declared.

Supplementary Material

kiab228_Supplementary_Data

J.Y.K., W.B.F., and M.M.W. conceptualized the review. All authors contributed major sections of this manuscript and prepared specific figures and tables (J.Y.K.: scRNA-seq analysis, SWEETs, other efflux transporters, M.M.W.: UmamiTs, phylogenetic analysis, E.L.: other efflux transporters, symbiosis and microbiota sections, T.Y.P. and M.L.: metabolic gene expression analysis).

The author responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (https://academic.oup.com/plphys/pages/general-instructions) is: Michael M. Wudick (wudick@hhu.de).

References

  1. Aluri S, Buttner M (2007) Identification and functional expression of the Arabidopsis thaliana vacuolar glucose transporter 1 and its role in seed germination and flowering. Proc Natl Acad Sci USA 104: 2537–2542 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. An J, Zeng T, Ji C, de Graaf S, Zheng Z, Xiao TT, Deng X, Xiao S, Bisseling T, Limpens E, et al. (2019) A Medicago truncatula SWEET transporter implicated in arbuscule maintenance during arbuscular mycorrhizal symbiosis. New Phytol 224: 396–408 [DOI] [PubMed] [Google Scholar]
  3. Andrés F, Kinoshita A, Kalluri N, Fernández V, Falavigna VS, Cruz TMD, Jang S, Chiba Y, Seo M, Mettler-Altmann T, et al. (2020) The sugar transporter SWEET10 acts downstream of FLOWERING LOCUS T during floral transition of Arabidopsis thaliana. BMC Plant Biol 20: 53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. dos Anjos L, Pandey PK, Moraes TA, Feil R, Lunn JE, Stitt M (2018) Feedback regulation by trehalose 6-phosphate slows down starch mobilization below the rate that would exhaust starch reserves at dawn in Arabidopsis leaves. Plant Direct 2: e00078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Aubry E, Hoffmann B, Vilaine F, Gilard F, Klemens PAW, Guérard F, Gakière B, Neuhaus HE, Bellini C, Dinant S, et al. (2021) A vacuolar hexose transport is required for xylem development in the inflorescence stem of Arabidopsis. bioRxiv 2020.12.09.417345 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Badri DV, Loyola-Vargas VM, Broeckling CD, De-la-Peña C, Jasinski M, Santelia D, Martinoia E, Sumner LW, Banta LM, Stermitz F, et al. (2008) Altered profile of secondary metabolites in the root exudates of Arabidopsis ATP-binding cassette transporter mutants. Plant Physiol 146: 762–771 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Badri DV, Quintana N, Kassis EGE, Kim HK, Choi YH, Sugiyama A, Verpoorte R, Martinoia E, Manter DK, Vivanco JM (2009) An ABC transporter mutation alters root exudation of phytochemicals that provoke an overhaul of natural soil microbiota. Plant Physiol 151: 2006–2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Baetz U, Martinoia E (2014) Root exudates: the hidden part of plant defense. Trends Plant Sci 19: 90–98 [DOI] [PubMed] [Google Scholar]
  9. Banasiak J, Biała W, Staszków A, Swarcewicz B, Kępczyńska E, Figlerowicz M, Jasiński M (2013) A Medicago truncatula ABC transporter belonging to subfamily G modulates the level of isoflavonoids. J Exp Bot 64: 1005–1015 [DOI] [PubMed] [Google Scholar]
  10. Baud S, Wuilleme S, Lemoine R, Kronenberger J, Caboche M, Lepiniec L, Rochat C (2005) The AtSUC5 sucrose transporter specifically expressed in the endosperm is involved in early seed development in Arabidopsis. Plant J 43: 824–836 [DOI] [PubMed] [Google Scholar]
  11. Berendsen RL, Vismans G, Yu K, Song Y, de Jonge R, Burgman WP, Burmølle M, Herschend J, Bakker PAHM, Pieterse CMJ (2018) Disease-induced assemblage of a plant-beneficial bacterial consortium. ISME J 12: 1496–1507 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Besnard J, Pratelli R, Zhao C, Sonawala U, Collakova E, Pilot G, Okumoto S (2016) UMAMIT14 is an amino acid exporter involved in phloem unloading in Arabidopsis roots. J Exp Bot 67: 6385–6397 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Besnard J, Sonawala U, Maharjan B, Collakova E, Finlayson SA, Pilot G, McDowell J, Okumoto S (2021) Increased expression of UMAMIT amino acid transporters results in activation of salicylic acid dependent stress response. Front Plant Sci 11: 606386. doi: 10.3389/fpls.2020.606386 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Besnard J, Zhao C, Avice J-C, Vitha S, Hyodo A, Pilot G, Okumoto S (2018) Arabidopsis UMAMIT24 and 25 are amino acid exporters involved in seed loading. J Exp Bot 69: 5221–5232 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bezrutczyk M, Hartwig T, Horschman M, Char SN, Yang J, Yang B, Frommer WB, Sosso D (2018a) Impaired phloem loading in zmsweet13a,b,c sucrose transporter triple knock-out mutants in Zea mays. New Phytol 218: 594–603 [DOI] [PubMed] [Google Scholar]
  16. Bezrutczyk M, Yang J, Eom J-S, Prior M, Sosso D, Hartwig T, Szurek B, Oliva R, Vera-Cruz C, White FF, et al. (2018b) Sugar flux and signaling in plant-microbe interactions. Plant J 93: 675–685 [DOI] [PubMed] [Google Scholar]
  17. Bezrutczyk M, Zöllner NR, Kruse CPS, Hartwig T, Lautwein T, Köhrer K, Frommer WB, Kim J-Y (2021) Evidence for phloem loading via the abaxial bundle sheath cells in maize leaves. Plant Cell 33: 531–547 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Blommel PG, Smith DW, Bingman CA, Dyer DH, Rayment I, Holden HM, Fox BG, Phillips GN Jr. (2004) Crystal structure of gene locus At3g16990 from Arabidopsis thaliana. Proteins 57: 221–222 [DOI] [PubMed] [Google Scholar]
  19. Boldt-Burisch K, Schneider BU, Naeth MA, Hüttl RF (2019) Root exudation of organic acids of herbaceous pioneer plants and their growth in sterile and non-sterile nutrient-poor, sandy soils from post-mining sites. Pedosphere 29: 34–44 [Google Scholar]
  20. Boorer KJ, Frommer WB, Bush DR, Kreman M, Loo DD, Wright EM (1996a) Kinetics and specificity of a H+/amino acid transporter from Arabidopsis thaliana. J Biol Chem 271: 2213–2220 [DOI] [PubMed] [Google Scholar]
  21. Boorer KJ, Loo DDF, Frommer WB, Wright EM (1996b) Transport mechanism of the cloned potato H+/sucrose cotransporter StSUT1. J Biol Chem 271: 25139–25144 [DOI] [PubMed] [Google Scholar]
  22. Borges F, Gardner R, Lopes T, Calarco JP, Boavida LC, Slotkin RK, Martienssen RA, Becker JD (2012) FACS-based purification of Arabidopsis microspores, sperm cells and vegetative nuclei. Plant Methods 8: 44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Borges F, Gomes G, Gardner R, Moreno N, McCormick S, Feijó JA, Becker JD (2008) Comparative transcriptomics of Arabidopsis sperm cells. Plant Physiol 148: 1168–1181 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Botha CEJ (2013) A tale of two neglected systems-structure and function of the thin- and thick-walled sieve tubes in monocotyledonous leaves. Front Plant Sci 4: 297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Buell CR, Joardar V, Lindeberg M, Selengut J, Paulsen IT, Gwinn ML, Dodson RJ, Deboy RT, Durkin AS, Kolonay JF, et al. (2003) The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc Natl Acad Sci USA 100: 10181–10186 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Bürkle L, Hibberd JM, Quick WP, Kühn C, Hirner B, Frommer WB (1998) The H+-sucrose cotransporter NtSUT1 is essential for sugar export from tobacco leaves. Plant Physiol 118: 59–68 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Carpaneto A, Geiger D, Bamberg E, Sauer N, Fromm J, Hedrich R (2005) Phloem-localized, proton-coupled sucrose carrier ZmSUT1 mediates sucrose efflux under the control of the sucrose gradient and the proton motive force. J Biol Chem 280: 21437–21443 [DOI] [PubMed] [Google Scholar]
  28. Chardon F, Bedu M, Calenge F, Klemens PA, Spinner L, Clement G, Chietera G, Leran S, Ferrand M, Lacombe B, et al. (2013) Leaf fructose content is controlled by the vacuolar transporter SWEET17 in Arabidopsis. Curr Biol 23: 697–702 [DOI] [PubMed] [Google Scholar]
  29. Chen HY, Huh JH, Yu YC, Ho LH, Chen L-Q, Tholl D, Frommer WB, Guo W-J (2015a) The Arabidopsis vacuolar sugar transporter SWEET2 limits carbon sequestration from roots and restricts Pythium infection. Plant J 83: 1046–1058 [DOI] [PubMed] [Google Scholar]
  30. Chen LQ, Cheung LS, Feng L, Tanner W, Frommer WB (2015b) Transport of sugars. Annu Rev Biochem 84: 865–894 [DOI] [PubMed] [Google Scholar]
  31. Chen LQ, Hou B-H, Lalonde S, Takanaga H, Hartung ML, Qu X-Q, Guo W-J, Kim J-G, Underwood W, Chaudhuri B, et al. (2010) Sugar transporters for intercellular exchange and nutrition of pathogens. Nature 468: 527–532 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Chen LQ, Lin IW, Qu X-Q, Sosso D, McFarlane HE, Londoño A, Samuels AL, Frommer WB (2015c) A cascade of sequentially expressed sucrose transporters in the seed coat and endosperm provides nutrition for the Arabidopsis embryo. Plant Cell 27: 607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Chen LQ, Qu X-Q, Hou B-H, Sosso D, Osorio S, Fernie AR, Frommer WB (2012) Sucrose efflux mediated by SWEET proteins as a key step for phloem transport. Science 335: 207. [DOI] [PubMed] [Google Scholar]
  34. Corratgé-Faillie C, Lacombe B (2017) Substrate (un)specificity of Arabidopsis NRT1/PTR FAMILY (NPF) proteins. J Exp Bot 68: 3107–3113 [DOI] [PubMed] [Google Scholar]
  35. Crouzet J, Roland J, Peeters E, Trombik T, Ducos E, Nader J, Boutry M (2013) NtPDR1, a plasma membrane ABC transporter from Nicotiana tabacum, is involved in diterpene transport. Plant Mol Biol 82: 181–192 [DOI] [PubMed] [Google Scholar]
  36. van Damme M, Zeilmaker T, Elberse J, Andel A, de Sain-van der Velden M, van den Ackerveken G (2009) Downy mildew resistance in Arabidopsis by mutation of HOMOSERINE KINASE. Plant Cell 21: 2179–2189 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. De Angeli A, Zhang J, Meyer S, Martinoia E (2013) AtALMT9 is a malate-activated vacuolar chloride channel required for stomatal opening in Arabidopsis. Nat Commun 4: 1804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. De Cuyper C, Goormachtig S (2017) Strigolactones in the rhizosphere: friend or foe? Mol Plant Microbe Interact 30: 683–690 [DOI] [PubMed] [Google Scholar]
  39. De Jong A, Koerselman-Kooij JW, Schuurmans JAMJ, Borstlap AC (1996) Characterization of the uptake of sucrose and glucose by isolated seed coat halves of developing pea seeds. Evidence that a sugar facilitator with diffusional kinetics is involved in seed coat unloading. Planta 199: 486–492 [Google Scholar]
  40. Denancé N, Ranocha P, Oria N, Barlet X, Rivière M-P, Yadeta KA, Hoffmann L, Perreau F, Clément G, Maia-Grondard A, et al. (2013) Arabidopsis wat1 (walls are thin1)-mediated resistance to the bacterial vascular pathogen, Ralstonia solanacearum, is accompanied by cross-regulation of salicylic acid and tryptophan metabolism. Plant J 73: 225–239 [DOI] [PubMed] [Google Scholar]
  41. Denyer T, Ma X, Klesen S, Scacchi E, Nieselt K, Timmermans MCP (2019) Spatiotemporal developmental trajectories in the Arabidopsis root revealed using high-throughput single-cell RNA sequencing. Dev Cell 48: 840–852.e5 [DOI] [PubMed] [Google Scholar]
  42. Desrut A, Moumen B, Thibault F, Le Hir R, Coutos-Thévenot P, Vriet C (2020) Beneficial rhizobacteria Pseudomonas simiae WCS417 induce major transcriptional changes in plant sugar transport. J Exp Bot 71: 7301–7315 [DOI] [PubMed] [Google Scholar]
  43. Dinant S, Wolff N, De Marco F, Vilaine F, Gissot L, Aubry E, Sandt C, Bellini C, Le Hir R (2019) Synchrotron FTIR and Raman spectroscopy provide unique spectral fingerprints for Arabidopsis floral stem vascular tissues. J Exp Bot 70: 871–884 [DOI] [PubMed] [Google Scholar]
  44. Dinkeloo K, Boyd S, Pilot G (2018) Update on amino acid transporter functions and on possible amino acid sensing mechanisms in plants. Sem Cell Dev Biol 74: 105–113 [DOI] [PubMed] [Google Scholar]
  45. Doroshenko V, Airich L, Vitushkina M, Kolokolova A, Livshits V, Mashko S (2007) YddG from Escherichia coli promotes export of aromatic amino acids. FEMS Microbiol Lett 275: 312–318 [DOI] [PubMed] [Google Scholar]
  46. Drakakaki G, Zabotina O, Delgado I, Robert S, Keegstra K, Raikhel N (2006) Arabidopsis reversibly glycosylated polypeptides 1 and 2 are essential for pollen development. Plant Physiol 142: 1480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Duan Z, Homma A, Kobayashi M, Nagata N, Kaneko Y, Fujiki Y, Nishida I (2014) Photoassimilation, assimilate translocation and plasmodesmal biogenesis in the source leaves of Arabidopsis thaliana grown under an increased atmospheric CO2 concentration. Plant Cell Physiol 55: 358–369 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Durand M, Mainson D, Porcheron B, Maurousset L, Lemoine R, Pourtau N (2018) Carbon source–sink relationship in Arabidopsis thaliana: the role of sucrose transporters. Planta 247: 587–611 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Durand M, Porcheron B, Hennion N, Maurousset L, Lemoine R, Pourtau N (2016) Water deficit enhances C export to the roots in Arabidopsis thaliana plants with contribution of sucrose transporters in both shoot and roots. Plant Physiol 170: 1460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Emms DM, Covshoff S, Hibberd JM, Kelly S (2016) Independent and parallel evolution of new genes by gene duplication in two origins of C4 photosynthesis provides new insight into the mechanism of phloem loading in C4 species. Mol Biol Evol 33: 1796–1806 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Engel ML, Holmes-Davis R, McCormick S (2005) Green sperm. Identification of male gamete promoters in Arabidopsis. Plant Physiol 138: 2124–2133 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Eom J-S, Chen L-Q, Sosso D, Julius BT, Lin IW, Qu X-Q, Braun DM, Frommer WB (2015) SWEETs, transporters for intracellular and intercellular sugar translocation. Curr Opin Plant Biol 25: 53–62 [DOI] [PubMed] [Google Scholar]
  53. Eom JS, Cho JI, Reinders A, Lee SW, Yoo Y, Tuan PQ, Choi SB, Bang G, Park YI, Cho MH, et al. (2011) Impaired function of the tonoplast-localized sucrose transporter in rice, OsSUT2, limits the transport of vacuolar reserve sucrose and affects plant growth. Plant Physiol 157: 109–119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Eom J-S, Choi S-B, Ward JM, Jeon J-S (2012) The mechanism of phloem loading in rice (Oryza sativa). Mol Cells 33: 431–438 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Eom J-S, Luo D, Atienza-Grande G, Yang J, Ji C, Luu VT, Huguet-Tapia JC, Liu B, Nguyen H, Schmidt SM, et al. (2019) Diagnostic kit for rice blight resistance. Nat Biotechnol 37: 1372–1379 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Fichtner F, Barbier FF, Annunziata MG, Feil R, Olas JJ, Mueller-Roeber B, Stitt M, Beveridge CA, Lunn JE (2021) Regulation of shoot branching in Arabidopsis by trehalose 6-phosphate. New Phytol 229: 2135–2151 [DOI] [PubMed] [Google Scholar]
  57. Fiehn O (2003) Metabolic networks of Cucurbita maxima phloem. Phytochem 62: 875–886 [DOI] [PubMed] [Google Scholar]
  58. Fieuw S, Patrick JW (1993) Mechanism of photosynthate efflux from Vicia faba L. seed coats. J Exp Bot 44: 65–74 [Google Scholar]
  59. Fiorilli V, Vallino M, Biselli C, Faccio A, Bagnaresi P, Bonfante P (2015) Host and non-host roots in rice: cellular and molecular approaches reveal differential responses to arbuscular mycorrhizal fungi. Front Plant Sci 6: 636. doi: 10.3389/fpls.2015.00636 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Fischer WN, Kwart M, Hummel S, Frommer WB (1995) Substrate specificity and expression profile of amino acid transporters (AAPs) in Arabidopsis. J Biol Chem 270: 16315–16320 [DOI] [PubMed] [Google Scholar]
  61. Fourcroy P, Tissot N, Gaymard F, Briat J-F, Dubos C (2016) Facilitated Fe nutrition by phenolic compounds excreted by the Arabidopsis ABCG37/PDR9 transporter requires the IRT1/FRO2 high-affinity root Fe2+ transport system. Mol Plant 9: 485–488 [DOI] [PubMed] [Google Scholar]
  62. Franke I, Resch A, Daßler T, Maier T, Böck A (2003) YfiK from Escherichia coli promotes export of O-acetylserine and cysteine. J Bacteriol 185: 1161–1166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Frei B, Eisenach C, Martinoia E, Hussein S, Chen X-Z, Arrivault S, Neuhaus HE (2018) Purification and functional characterization of the vacuolar malate transporter tDT from Arabidopsis. J Biol Chem 293: 4180–4190 [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Fritz E, Evert RF, Heyser W (1983) Microautoradiographic studies of phloem loading and transport in the leaf of Zea mays L. Planta 159: 193–206 [DOI] [PubMed] [Google Scholar]
  65. Fritz E, Evert RF, Nasse H (1989) Loading and transport of assimilates in different maize leaf bundles. Planta 178: 1–9 [DOI] [PubMed] [Google Scholar]
  66. Furukawa J, Yamaji N, Wang H, Mitani N, Murata Y, Sato K, Katsuhara M, Takeda K, Ma JF (2007) An aluminum-activated citrate transporter in barley. Plant Cell Physiol 48: 1081–1091 [DOI] [PubMed] [Google Scholar]
  67. Gahrtz M, Stolz J, Sauer N (1994) A phloem-specific sucrose-H+ symporter from Plantago major L. supports the model of apoplastic phloem loading. Plant J 6: 697–706 [DOI] [PubMed] [Google Scholar]
  68. Gamas P, Niebel Fde C, Lescure N, Cullimore J (1996) Use of a subtractive hybridization approach to identify new Medicago truncatula genes induced during root nodule development. Mol Plant Microbe Interact 9: 233–42 [DOI] [PubMed] [Google Scholar]
  69. Gao S, Gao J, Zhu X, Song Y, Li Z, Ren G, Zhou X, Kuai B (2016) ABF2, ABF3, and ABF4 promote ABA-mediated chlorophyll degradation and leaf senescence by transcriptional activation of chlorophyll catabolic genes and senescence-associated genes in Arabidopsis. Mol Plant 9: 1272–1285 [DOI] [PubMed] [Google Scholar]
  70. Ge Y-X, Angenent GC, Wittich PE, Peters J, Franken J, Busscher M, Zhang L-M, Dahlhaus E, Kater MM, Wullems GJ, et al. (2000) NEC1, a novel gene, highly expressed in nectary tissue of Petunia hybrida. Plant J 24: 725–734 [DOI] [PubMed] [Google Scholar]
  71. Gebauer P, Korn M, Engelsdorf T, Sonnewald U, Koch C, Voll LM (2017) Sugar accumulation in leaves of Arabidopsis sweet11/sweet12 double mutants enhances priming of the salicylic acid-mediated defense response. Front Plant Sci 8: 1378. doi: 10.3389/fpls.2017.01378 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Giaquinta RT (1976) Evidence for phloem loading from the apoplast. Chemical modification of membrane sulfhydryl groups. Plant Physiol 57: 872–875 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Gottwald JR, Krysan PJ, Young JC, Evert RF, Sussman MR (2000) Genetic evidence for the in planta role of phloem-specific plasma membrane sucrose transporters. Proc Natl Acad Sci USA 97: 13979–13984 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Guan Y-F, Huang X-Y, Zhu J, Gao J-F, Zhang H-X, Yang Z-N (2008) RUPTURED POLLEN GRAIN1, a member of the MtN3/saliva gene family, is crucial for exine pattern formation and cell integrity of microspores in Arabidopsis. Plant Physiol 147: 852. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Guo W-J, Nagy R, Chen H-Y, Pfrunder S, Yu Y-C, Santelia D, Frommer WB, Martinoia E (2014) SWEET17, a facilitative transporter, mediates fructose transport across the tonoplast of Arabidopsis roots and leaves. Plant Physiol 164: 777–789 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Hahn M, Viaud M, Kan J van (2014) The genome of Botrytis cinerea, a ubiquitous broad host range necrotroph. InDean RA, Lichens-Park A, Kole C, eds, Genomics of Plant-Associated Fungi and Oomycetes: Dicot Pathogens. Springer, Berlin, Heidelberg, pp 19–44 [Google Scholar]
  77. Han L, Zhu Y, Liu M, Zhou Y, Lu G, Lan L, Wang X, Zhao Y, Zhang XC (2017) Molecular mechanism of substrate recognition and transport by the AtSWEET13 sugar transporter. Proc Natl Acad Sci USA 114: 10089–10094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Handa Y, Nishide H, Takeda N, Suzuki Y, Kawaguchi M, Saito K (2015) RNA-seq transcriptional profiling of an arbuscular mycorrhiza provides insights into regulated and coordinated gene expression in Lotus japonicus and Rhizophagus irregularis. Plant Cell Physiol 56: 1490–1511 [DOI] [PubMed] [Google Scholar]
  79. Harbort CJ, Hashimoto M, Inoue H, Niu Y, Guan R, Rombolà AD, Kopriva S, Voges MJEEE, Sattely ES, Garrido-Oter R, et al. (2020) Root-secreted coumarins and the microbiota interact to improve Iron nutrition in Arabidopsis. Cell Host Microbe 28: 825–837.e6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Hayashi H, Chino M (1990) Chemical composition of phloem sap from the uppermost internode of the rice plant. Plant Cell Physiol 31: 247–251 [Google Scholar]
  81. Hoekenga OA, Maron LG, Piñeros MA, Cançado GMA, Shaff J, Kobayashi Y, Ryan PR, Dong B, Delhaize E, Sasaki T, et al. (2006) AtALMT1, which encodes a malate transporter, is identified as one of several genes critical for aluminum tolerance in Arabidopsis. Proc Natl Acad Sci USA 103: 9738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Huang AC, Jiang T, Liu Y-X, Bai Y-C, Reed J, Qu B, Goossens A, Nützmann H-W, Bai Y, Osbourn A (2019) A specialized metabolic network selectively modulates Arabidopsis root microbiota. Science 364: eaau6389. doi: 10.1126/science.aau6389 [DOI] [PubMed] [Google Scholar]
  83. Huang C, Yu J, Cai Q, Chen Y, Li Y, Ren Y, Miao Y (2020) Triple-localized WHIRLY2 influences leaf senescence and silique development via carbon allocation. Plant Physiol 184: 1348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Hütsch BW, Augustin J, Merbach W (2002) Plant rhizodeposition—an important source for carbon turnover in soils. J Plant Nutr Soil Sci 165: 397–407 [Google Scholar]
  85. Hvorup RN, Winnen B, Chang AB, Jiang Y, Zhou X-F, Saier Jr MH (2003) The multidrug/oligosaccharidyl-lipid/polysaccharide (MOP) exporter superfamily. Eur J Biochem 270: 799–813 [DOI] [PubMed] [Google Scholar]
  86. Igamberdiev AU, Eprintsev AT (2016) Organic acids: the pools of fixed carbon involved in redox regulation and energy balance in higher plants. Front Plant Sci 7: 1042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Jack DL, Yang NM, Saier MH (2001) The drug/metabolite transporter superfamily. Eur J Biochem 268: 3620–3639 [DOI] [PubMed] [Google Scholar]
  88. Jean-Baptiste K, McFaline-Figueroa JL, Alexandre CM, Dorrity MW, Saunders L, Bubb KL, Trapnell C, Fields S, Queitsch C, Cuperus JT (2019) Dynamics of gene expression in single root cells of Arabidopsis thaliana. Plant Cell 31: 993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Jeena GS, Kumar S, Shukla RK (2019) Structure, evolution and diverse physiological roles of SWEET sugar transporters in plants. Plant Mol Biol 100: 351–365 [DOI] [PubMed] [Google Scholar]
  90. de Jong A, Koerselman-Kooij JW, Schuurmans JAMJ, Borstlap AC (1997) The mechanism of amino acid efflux from seed coats of developing pea seeds as revealed by uptake experiments. Plant Physiol 114: 731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Kanno Y, Oikawa T, Chiba Y, Ishimaru Y, Shimizu T, Sano N, Koshiba T, Kamiya Y, Ueda M, Seo M (2016) AtSWEET13 and AtSWEET14 regulate gibberellin-mediated physiological processes. Nat Commun 7: 13245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Karmann J, Müller B, Hammes UZ (2018) The long and winding road: transport pathways for amino acids in Arabidopsis seeds. Plant Reprod 31: 253–261 [DOI] [PubMed] [Google Scholar]
  93. Kihira M, Taniguchi K, Kaneko C, Ishii Y, Aoki H, Koyanagi A, Kusano H, Suzui N, Yin Y-G, Kawachi N, et al. (2017) Arabidopsis thaliana FLO2 is Involved in efficiency of photoassimilate translocation, which is associated with leaf growth and aging, yield of seeds and seed quality. Plant Cell Physiol 58: 440–450 [DOI] [PubMed] [Google Scholar]
  94. Kim JY, Symeonidi E, Tin PY, Denyer T, Weidauer D, Miras M., Wudick M, Lercher M, Timmermans MCP, Frommer WB (2021) Distinct identities of leaf phloem cells revealed by single cell transcriptomics. Plant Cell 33: 511–530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Klemens PAW, Patzke K, Deitmer J, Spinner L, Le Hir R, Bellini C, Bedu M, Chardon F, Krapp A, Neuhaus HE (2013) Overexpression of the vacuolar sugar carrier AtSWEET16 modifies germination, growth, and stress tolerance in Arabidopsis. Plant Physiol 163: 1338–1352 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Kobayashi Y, Lakshmanan V, Kobayashi Y, Asai M, Iuchi S, Kobayashi M, Bais HP, Koyama H (2013) Overexpression of AtALMT1 in the Arabidopsis thaliana ecotype Columbia results in enhanced Al-activated malate excretion and beneficial bacterium recruitment. Plant Signal Behav 8: e25565.. doi: 10.4161/psb.25565 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Kretzschmar T, Kohlen W, Sasse J, Borghi L, Schlegel M, Bachelier JB, Reinhardt D, Bours R, Bouwmeester HJ, Martinoia E (2012) A petunia ABC protein controls strigolactone-dependent symbiotic signalling and branching. Nature 483: 341–344 [DOI] [PubMed] [Google Scholar]
  98. Kryvoruchko IS, Routray P, Senjuti S, Torres-Jerez I, Tejada-Jiménez M, Finney LA, Nakashima J, Pislariu CI, Benedito VA, Gonzalez-Guerrero M, et al. (2018) An Iron-activated citrate transporter, MtMATE67, is required for symbiotic nitrogen fixation. Plant Physiol 176: 2315–2329 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Kryvoruchko IS, Sinharoy S, Torres-Jerez I, Sosso D, Pislariu CI, Guan D, Murray J, Benedito VA, Frommer WB, Udvardi MK (2016) MtSWEET11, a nodule-specific sucrose transporter of Medicago truncatula root nodules. Plant Physiol 171: 554–565 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Kuzyakov Y, Jones DL (2006) Glucose uptake by maize roots and its transformation in the rhizosphere. Soil Biol Biochem 38: 851–860 [Google Scholar]
  101. Ladwig F, Stahl M, Ludewig U, Hirner AA, Hammes UZ, Stadler R, Harter K, Koch W (2012) Siliques are Red1 from Arabidopsis acts as a bidirectional amino acid transporter that is crucial for the amino acid homeostasis of siliques. Plant Physiol 158: 1643–1655 [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Lakshmanan V, Kitto SL, Caplan JL, Hsueh Y-H, Kearns DB, Wu Y-S, Bais HP (2012) Microbe-associated molecular patterns-triggered root responses mediate beneficial rhizobacterial recruitment in Arabidopsis. Plant Physiol 160: 1642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Lanfermeijer FC, Koerselman-Kooij JW, Borstlap AC (1990) Changing kinetics of l-valine uptake by immature pea cotyledons during development. Planta 181: 576–582 [DOI] [PubMed] [Google Scholar]
  104. Latorraca NR, Fastman NM, Venkatakrishnan AJ, Frommer WB, Dror RO, Feng L (2017) Mechanism of substrate translocation in an alternating access transporter. Cell 169: 96–107.e12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Lee J, Lee H, Kim J, Lee S, Kim DH, Kim S, Hwang I (2011) Both the hydrophobicity and a positively charged region flanking the C-terminal region of the transmembrane domain of signal-anchored proteins play critical roles in determining their targeting specificity to the endoplasmic reticulum or endosymbiotic organelles in Arabidopsis cells. Plant Cell 23: 1588–1607 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Le Hir R, Spinner L, Klemens PAW, Chakraborti D, de Marco F, Vilaine F, Wolff N, Lemoine R, Porcheron B, Géry C, et al. (2015) Disruption of the sugar transporters AtSWEET11 and AtSWEET12 affects vascular development and freezing tolerance in Arabidopsis. Mol Plant 8: 1687–1690 [DOI] [PubMed] [Google Scholar]
  107. Li G, Wang D, Yang R, Logan K, Chen H, Zhang S, Skaggs MI, Lloyd A, Burnett WJ, Laurie JD, et al. (2014) Temporal patterns of gene expression in developing maize endosperm identified through transcriptome sequencing. Proc Natl Acad Sci USA 111: 7582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Liao S, Wang L, Li J, Ruan Y-L (2020) Cell wall invertase is essential for ovule development through sugar signaling rather than provision of carbon nutrients. Plant Physiol 183: 1126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Lin IW, Sosso D, Chen L-Q, Gase K, Kim S-G, Kessler D, Klinkenberg PM, Gorder MK, Hou B-H, Qu X-Q, et al. (2014) Nectar secretion requires sucrose phosphate synthases and the sugar transporter SWEET9. Nature 508: 546–549 [DOI] [PubMed] [Google Scholar]
  110. Liu MY, Lou HQ, Chen WW, Piñeros MA, Xu JM, Fan W, Kochian LV, Zheng SJ, Yang JL (2018) Two citrate transporters coordinately regulate citrate secretion from rice bean root tip under aluminum stress. Plant Cell Environ 41: 809–822 [DOI] [PubMed] [Google Scholar]
  111. Liu X, Zhang Y, Yang C, Tian Z, Li J (2016) AtSWEET4, a hexose facilitator, mediates sugar transport to axial sinks and affects plant development. Sci Rep 6: 24563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Liu Z, Zhou Y, Guo J, Li J, Tian Z, Zhu Z, Wang J, Wu R, Zhang B, Hu Y, et al. (2020) Global dynamic molecular profiling of stomatal lineage cell development by single-cell RNA sequencing. Mol Plant 13: 1178–1193 [DOI] [PubMed] [Google Scholar]
  113. Livshits VA, Zakataeva NP, Aleshin VV, Vitushkina MV (2003) Identification and characterization of the new gene rhtA involved in threonine and homoserine efflux in Escherichia coli. Res Microbiol 154: 123–135 [DOI] [PubMed] [Google Scholar]
  114. Lu J, Le Hir R, Gomez-Paez D-M, Coen O, Péchoux C, Jasinski S, Magnani E (2020) The nucellus: between cell elimination and sugar transport. Plant Physiol 185: 478–490 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Lynch JM, Whipps JM (1990) Substrate flow in the rhizosphere. Plant and Soil 129: 1–10 [Google Scholar]
  116. Ma L, Zhang D, Miao Q, Yang J, Xuan Y, Hu Y (2017) Essential role of sugar transporter OsSWEET11 during the early stage of rice grain filling. Plant Cell Physiol 58: 863–873 [DOI] [PubMed] [Google Scholar]
  117. Macias-Benitez S, Garcia-Martinez AM, Caballero Jimenez P, Gonzalez JM, Tejada Moral M, Parrado Rubio J (2020) Rhizospheric organic acids as biostimulants: monitoring feedbacks on soil microorganisms and biochemical properties. Front Plant Sci 11: 633. doi: 10.3389/fpls.2020.00633 [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Magalhaes JV, Liu J, Guimarães CT, Lana UGP, Alves VMC, Wang Y-H, Schaffert RE, Hoekenga OA, Piñeros MA, Shaff JE, et al. (2007) A gene in the multidrug and toxic compound extrusion (MATE) family confers aluminum tolerance in sorghum. Nat Genet 39: 1156–1161 [DOI] [PubMed] [Google Scholar]
  119. Mahoney AK, Yin C, Hulbert SH (2017) Community structure, species variation, and potential functions of rhizosphere-associated bacteria of different winter wheat (Triticum aestivum) cultivars. Front Plant Sci 8: 132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Manck-Götzenberger J, Requena N (2016) Arbuscular mycorrhiza symbiosis induces a major transcriptional reprogramming of the potato SWEET sugar transporter family. Front Plant Sci 7: 487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Maron LG, Piñeros MA, Guimarães CT, Magalhaes JV, Pleiman JK, Mao C, Shaff J, Belicuas SNJ, Kochian LV (2010) Two functionally distinct members of the MATE (multi-drug and toxic compound extrusion) family of transporters potentially underlie two major aluminum tolerance QTLs in maize. Plant J 61: 728–740 [DOI] [PubMed] [Google Scholar]
  122. Matallana-Ramirez LP, Rauf M, Farage-Barhom S, Dortay H, Xue G-P, Dröge-Laser W, Lers A, Balazadeh S, Mueller-Roeber B (2013) NAC transcription factor ORE1 and senescence-induced BIFUNCTIONAL NUCLEASE1 (BFN1) constitute a regulatory cascade in Arabidopsis. Mol Plant 6: 1438–1452 [DOI] [PubMed] [Google Scholar]
  123. Meyer S, Scholz-Starke J, De Angeli A, Kovermann P, Burla B, Gambale F, Martinoia E (2011) Malate transport by the vacuolar AtALMT6 channel in guard cells is subject to multiple regulation. Plant J 67: 247–257 [DOI] [PubMed] [Google Scholar]
  124. Morii M, Sugihara A, Takehara S, Kanno Y, Kawai K, Hobo T, Hattori M, Yoshimura H, Seo M, Ueguchi-Tanaka M (2020) The dual function of OsSWEET3a as a gibberellin and glucose transporter Is important for young shoot development in rice. Plant Cell Physiol 61: 1935–1945 [DOI] [PubMed] [Google Scholar]
  125. Morita Y, Kodama K, Shiota S, Mine T, Kataoka A, Mizushima T, Tsuchiya T (1998) NorM, a putative multidrug efflux protein, of Vibrio parahaemolyticus and its homolog in Escherichia coli. Antimicrob Agents Chemother 42: 1778–1782 [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Müller B, Fastner A, Karmann J, Mansch V, Hoffmann T, Schwab W, Suter-Grotemeyer M, Rentsch D, Truernit E, Ladwig F, et al. (2015) Amino acid export in developing Arabidopsis seeds depends on UmamiT facilitators. Curr Biol 25: 3126–3131 [DOI] [PubMed] [Google Scholar]
  127. Münch E (1930) Die Stoffbewegungen in der Pflanze. Gustav Fischer Verlag, Jena [Google Scholar]
  128. Murphy KM, Edwards J, Louie KB, Bowen BP, Sundaresan V, Northen TR, Zerbe P (2021) Bioactive diterpenoids impact the composition of the root-associated microbiome in maize (Zea mays). Sci Rep 11: 333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Nakayama Y, Hashimoto K-I, Sawada Y, Sokabe M, Kawasaki H, Martinac B (2018) Corynebacterium glutamicum mechanosensitive channels: towards unpuzzling “glutamate efflux” for amino acid production. Biophys Rev 10: 1359–1369 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Obayashi T, Okamura Y, Ito S, Tadaka S, Aoki Y, Shirota M, Kinoshita K (2014) ATTED-II in 2014: evaluation of gene coexpression in agriculturally important plants. Plant Cell Physiol 55: e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Offre P, Pivato B, Siblot S, Gamalero E, Corberand T, Lemanceau P, Mougel C (2007) Identification of bacterial groups preferentially associated with mycorrhizal roots of Medicago truncatula. Appl Environ Microbiol 73: 913–921 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Pan Q, Cui B, Deng F, Quan J, Loake GJ, Shan W (2016) RTP1 encodes a novel endoplasmic reticulum (ER)-localized protein in Arabidopsis and negatively regulates resistance against biotrophic pathogens. New Phytol 209: 1641–1654 [DOI] [PubMed] [Google Scholar]
  133. Payyavula RS, Tay KH, Tsai CJ, Harding SA (2011) The sucrose transporter family in Populus: the importance of a tonoplast PtaSUT4 to biomass and carbon partitioning. Plant J 65: 757–70 [DOI] [PubMed] [Google Scholar]
  134. Phillips DA, Fox TC, King MD, Bhuvaneswari TV, Teuber LR (2004) Microbial products trigger amino acid exudation from plant roots. Plant Physiol 136: 2887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Piñeros MA, Cançado GMA, Maron LG, Lyi SM, Menossi M, Kochian LV (2008) Not all ALMT1-type transporters mediate aluminum-activated organic acid responses: the case of ZmALMT1 – an anion-selective transporter. Plant J 53: 352–367 [DOI] [PubMed] [Google Scholar]
  136. Pini F, East AK, Appia-Ayme C, Tomek J, Karunakaran R, Mendoza-Suárez M, Edwards A, Terpolilli JJ, Roworth J, Downie JA, et al. (2017) Bacterial biosensors for in Vivo spatiotemporal mapping of root secretion. Plant Physiol 174: 1289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Pommerrenig B, Ludewig F, Cvetkovic J, Trentmann O, Klemens PAW, Neuhaus HE (2018) In concert: orchestrated changes in carbohydrate homeostasis are critical for plant abiotic stress tolerance. Plant Cell Physiol 59: 1290–1299 [DOI] [PubMed] [Google Scholar]
  138. Poole P, Ramachandran V, Terpolilli J (2018) Rhizobia: from saprophytes to endosymbionts. Nat Rev Microbiol 16: 291–303 [DOI] [PubMed] [Google Scholar]
  139. Poschet G, Hannich B, Raab S, Jungkunz I, Klemens PAW, Krueger S, Wic S, Neuhaus HE, Büttner M (2011) A novel Arabidopsis vacuolar glucose exporter is involved in cellular sugar homeostasis and affects the composition of seed storage compounds. Plant Physiol 157: 1664–1676 [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Pratelli R, Pilot G (2014) Regulation of amino acid metabolic enzymes and transporters in plants. J Exp Bot 65: 5535–5556 [DOI] [PubMed] [Google Scholar]
  141. Pratelli R, Voll LM, Horst RJ, Frommer WB, Pilot G (2010) Stimulation of nonselective amino acid export by glutamine dumper proteins. Plant Physiol 152: 762–773 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Prior MJ, Selvanayagam J, Kim J-G, Tomar M, Jonikas M, Mudgett MB, Smeekens S, Hanson J, Frommer WB (2021) Arabidopsis bZIP11 is a susceptibility factor during Pseudomonas syringae infection. Mol Plant Microbe Interact 34: 439–447 [DOI] [PubMed] [Google Scholar]
  143. Qi T, Wang J, Huang H, Liu B, Gao H, Liu Y, Song S, Xie D (2015) Regulation of jasmonate-induced leaf senescence by antagonism between bHLH subgroup IIIe and IIId factors in Arabidopsis. Plant Cell 27: 1634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Ramesh SA, Tyerman SD, Xu B, Bose J, Kaur S, Conn V, Domingos P, Ullah S, Wege S, Shabala S, et al. (2015) GABA signalling modulates plant growth by directly regulating the activity of plant-specific anion transporters. Nat Commun 6: 7879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Ranocha P, Denancé N, Vanholme R, Freydier A, Martinez Y, Hoffmann L, Köhler L, Pouzet C, Renou J-P, Sundberg B, et al. (2010) Walls are thin 1 (WAT1), an Arabidopsis homolog of Medicago truncatula NODULIN21, is a tonoplast-localized protein required for secondary wall formation in fibers. Plant J 63: 469–483 [DOI] [PubMed] [Google Scholar]
  146. Ranocha P, Dima O, Nagy R, Felten J, Corratgé-Faillie C, Novák O, Morreel K, Lacombe B, Martinez Y, Pfrunder S, et al. (2013) Arabidopsis WAT1 is a vacuolar auxin transport facilitator required for auxin homoeostasis. Nat Commun 4: 2625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Rasheed S, Bashir K, Matsui A, Tanaka M, Seki M (2016) Transcriptomic analysis of soil-grown Arabidopsis thaliana roots and shoots in response to a drought stress. Front Plant Sci 7: 180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Ren Y, Li M, Guo S, Sun H, Zhao J, Zhang J, Liu G, He H, Tian S, Yu Y, et al. (2021) Evolutionary gain of oligosaccharide hydrolysis and sugar transport enhanced carbohydrate partitioning in sweet watermelon fruits. Plant Cell. doi: 10.1093/plcell/koab055 [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Riesmeier JW, Hirner B, Frommer WB (1993) Potato sucrose transporter expression in minor veins indicates a role in phloem loading. Plant Cell 5: 1591–1598 [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Riesmeier JW, Willmitzer L, Frommer WB (1992) Isolation and characterization of a sucrose carrier cDNA from spinach by functional expression in yeast. EMBO J 11: 4705–13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Riesmeier JW, Willmitzer L, Frommer WB (1994) Evidence for an essential role of the sucrose transporter in phloem loading and assimilate partitioning. EMBO J 13: 1–7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Rudrappa T, Czymmek KJ, Paré PW, Bais HP (2008) Root-secreted malic acid recruits beneficial soil bacteria. Plant Physiol 148: 1547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Ryu KH, Huang L, Kang HM, Schiefelbein J (2019) Single-cell RNA sequencing resolves molecular relationships among individual plant cells. Plant Physiol 179: 1444–1456 [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Sanders A, Collier R, Trethewy A, Gould G, Sieker R, Tegeder M (2009) AAP1 regulates import of amino acids into developing Arabidopsis embryos. Plant J 59: 540–552 [DOI] [PubMed] [Google Scholar]
  155. Sasaki T, Yamamoto Y, Ezaki B, Katsuhara M, Ahn SJ, Ryan PR, Delhaize E, Matsumoto H (2004) A wheat gene encoding an aluminum-activated malate transporter. Plant J 37: 645–653 [DOI] [PubMed] [Google Scholar]
  156. Sasse J, Martinoia E, Northen T (2018) Feed your friends: do plant exudates shape the root microbiome? Trends Plant Sci 23: 25–41 [DOI] [PubMed] [Google Scholar]
  157. Sasse J, Simon S, Gübeli C, Liu G-W, Cheng X, Friml J, Bouwmeester H, Martinoia E, Borghi L (2015) Asymmetric localizations of the ABC transporter PaPDR1 trace paths of directional strigolactone transport. Curr Biol 25: 647–655 [DOI] [PubMed] [Google Scholar]
  158. Sauer N, Stolz J (1994) SUC1 and SUC2: two sucrose transporters from Arabidopsis thaliana; expression and characterization in baker’s yeast and identification of the histidine-tagged protein. Plant J 6: 67–77 [DOI] [PubMed] [Google Scholar]
  159. Schmidt UG, Endler A, Schelbert S, Brunner A, Schnell M, Neuhaus HE, Marty-Mazars D, Marty F, Baginsky S, Martinoia E (2007) Novel tonoplast transporters identified using a proteomic approach with vacuoles isolated from cauliflower buds. Plant Physiol 145: 216–229 [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Schulz A, Beyhl D, Marten I, Wormit A, Neuhaus E, Poschet G, Buttner M, Schneider S, Sauer N, Hedrich R (2011) Proton-driven sucrose symport and antiport are provided by the vacuolar transporters SUC4 and TMT1/2. Plant J 68: 129–136 [DOI] [PubMed] [Google Scholar]
  161. Schwelm A, Fogelqvist J, Knaust A, Jülke S, Lilja T, Bonilla-Rosso G, Karlsson M, Shevchenko A, Dhandapani V, Choi SR, et al. (2015) The Plasmodiophora brassicae genome reveals insights in its life cycle and ancestry of chitin synthases. Sci Rep 5: 11153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Sekhon RS, Lin H, Childs KL, Hansey CN, Buell CR, de Leon N, Kaeppler SM (2011) Genome-wide atlas of transcription during maize development. Plant J 66: 553–563 [DOI] [PubMed] [Google Scholar]
  163. Sellami S, Le Hir R, Thorpe MR, Vilaine F, Wolff N, Brini F, Dinant S (2019) Salinity effects on sugar homeostasis and vascular anatomy in the stem of the Arabidopsis thaliana inflorescence. Int J Mol Sci 20: 3167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Seo PJ, Park JM, Kang SK, Kim SG, Park CM (2011) An Arabidopsis senescence-associated protein SAG29 regulates cell viability under high salinity. Planta 233: 189–200 [DOI] [PubMed] [Google Scholar]
  165. Shi S, Richardson AE, O’Callaghan M, DeAngelis KM, Jones EE, Stewart A, Firestone MK, Condron LM (2011) Effects of selected root exudate components on soil bacterial communities. FEMS Microbiol Ecol 77: 600–610 [DOI] [PubMed] [Google Scholar]
  166. Shulse CN, Cole BJ, Ciobanu D, Lin J, Yoshinaga Y, Gouran M, Turco GM, Zhu Y, O’Malley RC, Brady SM, et al. (2019) High-throughput single-cell transcriptome profiling of plant cell types. Cell Rep 27: 2241–2247.e4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Slewinski TL, Meeley R, Braun DM (2009) Sucrose transporter1 functions in phloem loading in maize leaves. J Exp Bot 60: 881–892 [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Smeekens S, Ma J, Hanson J, Rolland F (2010) Sugar signals and molecular networks controlling plant growth. Curr Opin Plant Biol 13: 274–279 [DOI] [PubMed] [Google Scholar]
  169. Sosso D, Luo D, Li Q-B, Sasse J, Yang J, Gendrot G, Suzuki M, Koch KE, McCarty DR, Chourey PS, et al. (2015) Seed filling in domesticated maize and rice depends on SWEET-mediated hexose transport. Nat Genet 47: 1489–1493 [DOI] [PubMed] [Google Scholar]
  170. Srivastava AC, Ganesan S, Ismail IO, Ayre BG (2008) Functional characterization of the Arabidopsis AtSUC2 Sucrose/H+ symporter by tissue-specific complementation reveals an essential role in phloem loading but not in long-distance transport. Plant Physiol 148: 200–211 [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Stadler R, Lauterbach C, Sauer N (2005) Cell-to-cell movement of green fluorescent protein reveals post-phloem transport in the outer integument and identifies symplastic domains in Arabidopsis seeds and embryos. Plant Physiol 139: 701–712 [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Stassen MJJ, , HsuS-H, , PieterseCMJ, , StringlisIA (. 2021) Coumarin Communication Along the Microbiome–Root–Shoot Axis. Trends in Plant Science 26: 169–183 [DOI] [PubMed] [Google Scholar]
  173. Struck C (2015) Amino acid uptake in rust fungi. Front Plant Sci 6: 40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Subramanian S, Stacey G, Yu O (2007) Distinct, crucial roles of flavonoids during legume nodulation. Trends Plant Sci 12: 282–285 [DOI] [PubMed] [Google Scholar]
  175. Sugiyama A, Saida Y, Yoshimizu M, Takanashi K, Sosso D, Frommer WB, Yazaki K (2017) Molecular characterization of LjSWEET3, a sugar transporter in nodules of Lotus japonicus. Plant Cell Physiol 58: 298–306 [DOI] [PubMed] [Google Scholar]
  176. Sun MX, Huang XY, Yang J, Guan YF, Yang ZN (2013) Arabidopsis RPG1 is important for primexine deposition and functions redundantly with RPG2 for plant fertility at the late reproductive stage. Plant Reprod 26: 83–91 [DOI] [PubMed] [Google Scholar]
  177. Takanashi K, Sasaki T, Kan T, Saida Y, Sugiyama A, Yamamoto Y, Yazaki K (2016) A dicarboxylate transporter, LjALMT4, mainly expressed in nodules of Lotus japonicus. Mol Plant Microbe Interact 29: 584–592 [DOI] [PubMed] [Google Scholar]
  178. Takanashi K, Yokosho K, Saeki K, Sugiyama A, Sato S, Tabata S, Ma JF, Yazaki K (2013) LjMATE1: A citrate transporter responsible for iron supply to the nodule infection zone of Lotus japonicus. Plant Cell Physiol 54: 585–594 [DOI] [PubMed] [Google Scholar]
  179. Taniguchi M, Taniguchi Y, Kawasaki M, Takeda S, Kato T, Sato S, Tabata S, Miyake H, Sugiyama T (2002) Identifying and characterizing plastidic 2-oxoglutarate/malate and dicarboxylate transporters in Arabidopsis thaliana. Plant Cell Physiol 43: 706–717 [DOI] [PubMed] [Google Scholar]
  180. Tao Y, Cheung LS, Li S, Eom J-S, Chen L-Q, Xu Y, Perry K, Frommer WB, Feng L (2015) Structure of a eukaryotic SWEET transporter in a homotrimeric complex. Nature 527: 259–263 [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Tegeder M, Offler CE, Frommer WB, Patrick JW (2000) Amino acid transporters are localized to transfer cells of developing Pea seeds. Plant Physiol 122: 319–325 [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Thiergart T, Zgadzaj R, Bozsóki Z, Garrido-Oter R, Radutoiu S, Schulze-Lefert P (2019) Lotus japonicus symbiosis genes impact microbial interactions between symbionts and multikingdom commensal communities. mBio 10: e01833–19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Thorne JH (1985) Phloem unloading of C and N assimilates in developing seeds. Annu Rev Plant Physiol 36: 317–343 [Google Scholar]
  184. Touraine B, Muller B, Grignon C (1992) Effect of phloem-translocated malate on NO3 uptake by roots of intact soybean plants. Plant Physiol 99: 1118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Tsuchiya H, Doki S, Takemoto M, Ikuta T, Higuchi T, Fukui K, Usuda Y, Tabuchi E, Nagatoishi S, Tsumoto K, et al. (2016) Structural basis for amino acid export by DMT superfamily transporter YddG. Nature 534: 417–420 [DOI] [PubMed] [Google Scholar]
  186. Tucker AM, Winkler HH, Driskell LO, Wood DO (2003) S-adenosylmethionine transport in Rickettsia prowazekii. J Bacteriol 185: 3031–3035 [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Uroz S, Courty PE, Oger P (2019) Plant symbionts are engineers of the plant-associated microbiome. Trends Plant Sci 24: 905–916 [DOI] [PubMed] [Google Scholar]
  188. Valle J, Re SD, Schmid S, Skurnik D, D’Ari R, Ghigo J-M (2008) The amino acid valine Is secreted in continuous-flow bacterial biofilms. J Bacteriol 190: 264–274 [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Veillet F, Gaillard C, Lemonnier P, Coutos-Thévenot P, La Camera S (2017) The molecular dialogue between Arabidopsis thaliana and the necrotrophic fungus Botrytis cinerea leads to major changes in host carbon metabolism. Sci Rep 7: 17121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Voges MJEEE, Bai Y, Schulze-Lefert P, Sattely ES (2019) Plant-derived coumarins shape the composition of an Arabidopsis synthetic root microbiome. Proc Natl Acad Sci USA 116: 12558–12565 [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Walerowski P, Gündel A, Yahaya N, Truman W, Sobczak M, Olszak M, Rolfe S, Borisjuk L, Malinowski R (2018) Clubroot disease stimulates early steps of phloem differentiation and recruits SWEET sucrose transporters within developing galls. Plant Cell 30: 3058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Walker NA, Patrick JW, Zhang W-H, Fieuw S (1995) Efflux of photosynthate and acid from developing seed coats of Phaseolus vulgaris L.: a chemiosmotic analysis of pump-driven efflux. J Exp Bot 46: 539–549 [Google Scholar]
  193. Walker NA, Zhang WH, Harrington G, Holdaway N, Patrick JW (2000) Effluxes of solutes from developing seed coats of Phaseolus vulgaris L. and Vicia faba l.: locating the effect of turgor in a coupled chemiosmotic system. J Exp Bot 51: 1047–1055 [DOI] [PubMed] [Google Scholar]
  194. Wang H, Yan S, Xin H, Huang W, Zhang H, Teng S, Yu Y-C, Fernie AR, Lu X, Li P, et al. (2019) A subsidiary cell-localized glucose transporter promotes stomatal conductance and photosynthesis. Plant Cell 31: 1328–1343 [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Wang X, Feng H, Wang Y, Wang M, Xie X, Chang H, Wang L, Qu J, Sun K, He W, et al. (2020) Mycorrhizal symbiosis modulates the rhizosphere microbiota to promote rhizobia-legume symbiosis. Mol Plant 14: 503–516 [DOI] [PubMed] [Google Scholar]
  196. Wei X, Nguyen STT, Collings DA, McCurdy DW (2020) Sucrose regulates wall ingrowth deposition in phloem parenchyma transfer cells in Arabidopsis via affecting phloem loading activity. J Exp Bot 71: 4690–4702 [DOI] [PubMed] [Google Scholar]
  197. Wendrich JR, Yang B, Vandamme N, Verstaen K, Smet W, Van de Velde C, Minne M, Wybouw B, Mor E, Arents HE, et al. (2020) Vascular transcription factors guide plant epidermal responses to limiting phosphate conditions. Science 370: eaay4970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Werner D, Gerlitz N, Stadler R (2011) A dual switch in phloem unloading during ovule development in Arabidopsis. Protoplasma 248: 225–235 [DOI] [PubMed] [Google Scholar]
  199. Wolswinkel P, Ammerlaan A (1983) Phloem unloading in developing seeds of Vicia faba L.: the effect of several inhibitors on the release of sucrose and amino acids by the seed coat. Planta 158: 205–215 [DOI] [PubMed] [Google Scholar]
  200. Wormit A, Trentmann O, Feifer I, Lohr C, Tjaden J, Meyer S, Schmidt U, Martinoia E, Neuhaus HE (2006) Molecular identification and physiological characterization of a novel monosaccharide transporter from Arabidopsis involved in vacuolar sugar transport. Plant Cell 18: 3476–3490 [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Xiao Z, Dai Z, Locasale JW (2019) Metabolic landscape of the tumor microenvironment at single cell resolution. Nat Commun 10: 3763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Xu Y, Tao Y, Cheung LS, Fan C, Chen L-Q, Xu S, Perry K, Frommer WB, Feng L (2014) Structures of bacterial homologues of SWEET transporters in two distinct conformations. Nature 515: 448–452 [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Yang B, Sugio A, White FF (2006) Os8N3 is a host disease-susceptibility gene for bacterial blight of rice. Proc Natl Acad Sci USA 103: 10503–10508 [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Yang J, Luo D, Yang B, Frommer WB, Eom J-S (2018) SWEET11 and 15 as key players in seed filling in rice. New Phytol 218: 604–615 [DOI] [PubMed] [Google Scholar]
  205. Yokosho K, Yamaji N, Ma JF (2011) An Al-inducible MATE gene is involved in external detoxification of Al in rice. Plant J 68: 1061–1069 [DOI] [PubMed] [Google Scholar]
  206. Yokosho K, Yamaji N, Ueno D, Mitani N, Ma JF (2009) OsFRDL1 is a citrate transporter required for efficient translocation of iron in rice. Plant Physiol 149: 297–305 [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Zeier J (2013) New insights into the regulation of plant immunity by amino acid metabolic pathways. Plant Cell Environ 36: 2085–2103 [DOI] [PubMed] [Google Scholar]
  208. Zgadzaj R, Garrido-Oter R, Jensen DB, Koprivova A, Schulze-Lefert P, Radutoiu S (2016) Root nodule symbiosis in Lotus japonicus drives the establishment of distinctive rhizosphere, root, and nodule bacterial communities. Proc Natl Acad Sci USA 113: E7996–E8005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Zhang H, Zhang L, Wu S, Chen Y, Yu D, Chen L (2020) AtWRKY75 positively regulates age-triggered leaf senescence through gibberellin pathway. Plant Divers. doi: 10.1016/j.pld.2020.10.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Zhang L, Garneau MG, Majumdar R, Grant J, Tegeder M (2015) Improvement of pea biomass and seed productivity by simultaneous increase of phloem and embryo loading with amino acids. Plant J 81: 134–146 [DOI] [PubMed] [Google Scholar]
  211. Zhang T-Q, Xu Z-G, Shang G-D, Wang J-W (2019) A single-cell RNA sequencing profiles the developmental landscape of Arabidopsis root. Mol Plant 12: 648–660 [DOI] [PubMed] [Google Scholar]
  212. Zhang W-H, Zhou Y, Dibley KE, Tyerman SD, Furbank RT, Patrick JW (2007) Nutrient loading of developing seeds. Funct Plant Biol 34: 314–331 [DOI] [PubMed] [Google Scholar]
  213. Zhao S, Chen A, Chen C, Li C, Xia R, Wang X (2019) Transcriptomic analysis reveals the possible roles of sugar metabolism and export for positive mycorrhizal growth responses in soybean. Physiol Plant 166: 712–728 [DOI] [PubMed] [Google Scholar]
  214. Zhao W, Faust F, Schubert S (2020) Potassium is a potential toxicant for Arabidopsis thaliana under saline conditions. J Plant Nutr Soil Sci 183: 455–467 [Google Scholar]
  215. Zhao Y, Chan Z, Gao J, Xing L, Cao M, Yu C, Hu Y, You J, Shi H, Zhu Y, et al. (2016) ABA receptor PYL9 promotes drought resistance and leaf senescence. Proc Natl Acad Sci USA 113: 1949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Zhen Q, Fang T, Peng Q, Liao L, Zhao L, Owiti A, Han Y (2018) Developing gene-tagged molecular markers for evaluation of genetic association of apple SWEET genes with fruit sugar accumulation. Horticulture Research 5: 14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Zhou Y, Cui X, Hu A, Miao Y, Zhang L (2020) Characterization and functional analysis of pollen-specific PwSWEET1 in Picea wilsonii. J Forest Res 31: 1913–1922 [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

kiab228_Supplementary_Data

Articles from Plant Physiology are provided here courtesy of Oxford University Press

RESOURCES