Abstract
Sulfur (S) is an essential element for life on Earth. Plants are able to take up and utilize sulfate (SO42–), the most oxidized inorganic form of S compounds on Earth, through the reductive S assimilatory pathway that couples with photosynthetic energy conversion. Organic S compounds are subsequently synthesized in plants and made accessible to animals, primarily as the amino acid methionine. Thus, plant S metabolism clearly has nutritional importance in the global food chain. S metabolites may be part of redox regulation and drivers of essential metabolic pathways as cofactors and prosthetic groups, such as Fe–S centers, CoA, thiamine, and lipoic acid. The evolution of the S metabolic pathways and enzymes reflects the critical importance of functional innovation and diversifications. Here we review the major evolutionary alterations that took place in S metabolism across different scales and outline research directions that may take advantage of understanding the evolutionary adaptations.
Keywords: Cysteine, evolution, glucosinolates, glutathione, metabolism, methionine, natural variation, sulfur
This review delves into major evolutionary alterations that shaped and diversified sulfur metabolism across the domains of life, highlighting the speciation that occurred in different plant taxa.
Introduction
Sulfate (SO42–) is the main source of sulfur (S) utilized by plants. It is taken up from soil into the root system and subsequently distributed to various plant organs and cellular/subcellular compartments that may play specialized roles in plant development and protection. Root uptake of SO42– from soil and internal SO42– distribution largely depend on functions of the membrane-bound sulfate transporters (SULTRs) (Takahashi et al., 2011; Maruyama-Nakashita, 2017). While SO42– can be temporarily stored in the vacuoles, depending on cellular demands for S, it may be mobilized to serve as the substrate for synthesis of S-containing metabolites in the cytosol, plastids, and mitochondria (Fig. 1).
Fig. 1.
Schematic representation of S assimilation in plants. APS, adenosine 5'-phosphosulfate; PAPS, 3'-phosphoadenosine 5'-phosphosulfate; PAP, 3'-phosphoadenosine 5'-phosphate; OAS, O-acetylserine; SULTR, sulfate transporter; ATPS, ATP sulfurylase; APR, APS reductase; APK, APS kinase; SiR, sulfite reductase; SERAT, serine acetyltransferase; OASTL, OAS (thiol)lyase; PAPST, PAPS/PAP transporter; SOT, sulfotransferase; SAL1, 3ʹ,5ʹ-bisphosphate nucleotidase.
Immediately after SO42– import across the plasma membranes or the plastid envelopes, adenylation of SO42– to adenosine 5'-phosphosulfate (APS) takes place in both the plastids and the cytosol through an enzymatic reaction catalyzed by the ATP sulfurylase (ATPS). In higher plants, APS—the activated form of SO42–—resides at the branching point of S metabolism in the plastids where the flux of S can further be partitioned into the reductive and oxidized S assimilation pathways, also described as the primary and the secondary pathways (Kopriva et al., 2012) (Fig. 1). In the reductive S assimilation pathway, the enzymatic reduction of APS to sulfide (S2–) takes place in a stepwise manner; APS reductase (APR) catalyzes the first step generating sulfite (SO32–) by transferring two electrons to the activated SO42– group of APS, and subsequently in the second step the ferredoxin-dependent sulfite reductase (SiR) provides six electrons to reduce SO32– to S2– (Saito, 2004; Takahashi et al., 2011). S2– then becomes the substrate for the O-acetylserine(thiol)lyase (OAS-TL) catalyzing the S2– incorporation into the amino acid skeleton of O-acetylserine (OAS) to form the S amino acid cysteine (Cys) (Hell and Wirtz, 2011). OAS-TL acts together with serine acetyltransferase (SERAT), the enzyme activating the amino acid Ser to OAS, in the Cys synthase complex (Droux et al., 1998; Wirtz et al., 2004, 2012). While the pathway for SO42– reduction to S2– localizes exclusively in the plastids, Cys biosynthesis takes place in the plastids as well as in the mitochondria and the cytosol (Fig. 1) (Takahashi et al., 2011; Maruyama-Nakashita, 2017).
Methionine (Met) and glutathione (GSH) are S compounds of vital importance. Met is an essential amino acid for animals and humans as they do not assimilate SO42– into Cys. Translation initiation of every protein requires Met. Apart from its essentiality for protein synthesis, Met serves as an intermediate for synthesis of S-adenosylmethionine, the major donor of methyl groups for numerous biological processes such as DNA and protein methylation and synthesis of various natural products, including polyamines, as well as 1-aminocyclopropane-1-carboxylate (ACC), the substrate for biosynthesis of the plant hormone ethylene (Hesse and Hoefgen, 2003; Hesse et al., 2004; Sauter et al., 2013). GSH is a tripeptide with its thiol group derived from Cys, playing important roles in redox regulation as well as detoxification, conjugation, and transport of various toxic metabolites and heavy metals (Noctor et al., 2012). Subsequent enzymatic oligomerization of the γ-Glu–Cys units of GSH into phytochelatins provides a mitigation mechanism notably against toxic heavy metals (Cobbett and Goldsbrough, 2002). GSH is essential for root development, as the reduction in its content causes a concentration-dependent reduction in root length (Vernoux et al., 2000; Bangash et al., 2019). Cys serves as the reduced S donor for synthesis of both Met and GSH. The Met biosynthesis from Cys occurs in the trans-sulfuration pathway with three distinct enzymatic reactions: the condensation of Cys and O-phosphohomoserine to form cystathionine; the hydrolysis of cystathionine to form homocysteine; and the methylation of homocysteine to produce Met, catalyzed by cystathionine γ-synthase, cystathionine β-lyase, and methionine synthase, respectively (Breitinger et al., 2001; Hesse et al., 2004). On the other hand, as indicated by the chemical formula γ-glutamylcysteinylglycine, GSH is synthesized from its constituting amino acids Glu, Cys, and Gly in a two-step ATP-dependent enzymatic reaction catalyzed by glutamate-cysteine ligase (also known as γ-glutamylcysteine synthetase) and glutathione synthetase (Noctor and Foyer, 1998; Takahashi et al., 2011). Loss of GSH synthesis is, therefore, lethal (Cairns et al., 2006).
In the oxidized S assimilation pathway, phosphorylation of APS to 3'-phosphoadenosine 5'-phosphosulfate (PAPS) via phospho-transfer from ATP takes place in both the plastids and the cytosol through the function of the APS kinase (APK) (Fig. 1) (Mugford et al., 2009). PAPS is the active form of sulfate used for sulfation of peptides and various specialized metabolites through condensation reactions catalyzed by the sulfotransferases (SOTs) in the cytosol and the Golgi apparatus (Klein and Papenbrock, 2004; Takahashi et al., 2011; Chan et al., 2019). Sulfation is an important modification step for animal proteins and hormones (Gunal et al., 2019), while in plants the best studied sulfated chemical compounds are specialized metabolites, for example glucosinolates, sulfoflavonoids, and peptide hormones such as phytosulfokine (Halkier and Gershenzon, 2006; Bednarek et al., 2009; Chan et al., 2019; Kaufmann and Sauter, 2019). While these modifications of specialized metabolites through sulfation yield a biproduct 3'-phosphoadenosine 5'-phosphate (PAP), subsequent metabolic conversion of this phosphonucleotide completes the pathway and prevents its signaling action (Chan et al., 2019). The chloroplast/mitochondrion dual-localized 3ʹ,5ʹ-bisphosphate nucleotidase SAL1 (Estavillo et al., 2011; Chan et al., 2016) and the PAPS/PAP transporters facilitating PAP import to these organelles (Gigolashvili et al., 2012; Ashykhmina et al., 2019) are key players in this metabolic cycle coupled with the function of the APK enzyme located at the branching point of S metabolism in plants (Fig. 1).
S metabolism is essential for all living organisms, but evolved differently across the evolutionary tree. Not only the arrangement of pathways and enzymatic reactions but also the enzymes themselves catalyzing the same reaction are functionally diversified and can be assembled in entirely different configurations. For example, the trans-sulfuration pathway between the S-amino acids Cys and Met operates in opposite directions in plants and animals (see later) (Giovanelli et al., 1985; Hesse and Hoefgen, 2003; Hesse et al., 2004; Brosnan and Brosnan, 2006). Diverse structural organization of SO42– activation and PAPS biosynthetic enzymes—the ATPS and the APK enzymes found in strikingly different forms in bacteria and eukaryotes (see below)—represents another prominent example of S metabolic diversification which has emerged across kingdoms (Patron et al., 2008). S metabolism impacts life on Earth at different scales given its prevalence and indispensability. The diversity in plant S metabolism extends from a global geological scale to variations among plant taxa and ecotypes. In this review we will discuss plant S metabolism from the evolutionary perspective, highlighting its diversity, and outline open questions in research areas that would benefit from a broader consideration of plant evolution.
Adaptation of S metabolism in early evolution of life
S played a crucial role in shaping the development of life on Earth. Indeed, most of the hypotheses about the origin of life place S at the core of the initial energy sources (Olson, 2021). This is in agreement with evolutionary phylogeny that places organisms conducting dissimilatory S reduction and anoxygenic photosynthesis at the root of the tree of life (Burini et al., 2018). The S reduction probably produced hydrogen sulfide (H2S) that in turn was utilized as an electron donor in anoxygenic photosynthesis and led to the production of oxidized S compounds, in particular SO42–. Despite the occurrence of the first dissimilatory SO42– reducers originating possibly already 3.5 billion years ago (Canfield and Raiswell, 1999), SO42– concentration in the marine environment increased, reaching ~1 mM around 2.5 billion years ago. This seems to be the onset of oxygenic photosynthesis and further radiation of SO42– reduction (Canfield and Raiswell, 1999). A further increase in SO42– concentration led to changes in Fe deposits and more widespread use of SO42– as a source of S for biochemical reactions (Burini et al., 2018). In addition, the increase in SO42– concentration in seawater has been implicated to have driven changes in abundance of even diverse groups of primary producers in the oceans (Ratti et al., 2011). While in the Paleozoic ocean green algae and cyanobacteria were particularly abundant, with an increase in SO42– in the seawater, the composition of phytoplankton shifted towards Eukaryotic Chl a+c-containing microalgae such as diatoms, coccolithophorids, and dinoflagellates (Ratti et al., 2011).
A remnant of the evolutionary process that has led to oxygenic photosynthesis and SO42– assimilation is the ability of cyanobacteria to switch to anoxygenic photosynthesis in environments where S2– is present (Fig. 2) (Cohen et al., 1986). Despite the fact that oxygenic photosynthesis evolved in S2–-rich water (Canfield, 1998), S2– is toxic to modern cyanobacteria because it inhibits cytochrome c oxidase and inactivates PSII (Cuevasanta et al., 2017). Nevertheless, cyanobacteria are able to grow in the presence of S2– using several different strategies. Firstly, some cyanobacterial species restrict oxygenic photosynthesis in the presence of S2–. Secondly, there are some species that are capable of oxidizing S2– to elemental S by sulfide:quinone reductase (SQR), thus providing electrons to PSI (de Beer et al., 2017). SQR catalyzes the oxidation of S2– to sulfane S and plays a key role in energy transduction and S2– detoxification. Thirdly, certain cyanobacteria exhibit various types of D1 proteins, including an S2–-resistant variant capable of sustaining both oxygenic and anoxygenic photosynthesis concurrently (Mulo et al., 2009). Fourthly, such habitats are advantageous for purely anoxygenic cyanobacteria lacking oxygenic photosynthesis (de Beer et al., 2017), which directly oxidize S2– to S by SQR and therefore channel the flow of electrons from S2– to PSI through the quinone pool. SQR is thus directly involved in the anoxygenic photosynthetic activity of cyanobacteria, allowing them to use S2– as the electron donor (Rahimzadeh Karvansara et al., 2023). Although, in most cyanobacteria, photosynthesis is completely inhibited by sulfide, 45% of the sequenced cyanobacterial genomes in fact contain SQR, indicating the importance of the anoxygenic photosynthetic pathway in fitness under extremely dynamic habitats (Fig. 2) (Xia et al., 2017). Diversity in the photosynthetic metabolic pathways transforming S compounds thus reflects mechanisms that could have played an important role in the shaping of the extant photosynthetic organisms.
Fig. 2.
Effects of H2S on cyanobacterial photosynthesis. In oxygenic environments, H2S inhibits PSII and prevents electron transfer and water oxidation (1). Many cyanobacteria possess sulfide:quinone reductase (SQR) able to oxidize H2S to elemental sulfur (S0) and transfer electrons to PSI (2). b6f, cytochrome b6f complex.
Diversification of S metabolism at the kingdom scale
Dissimilatory and assimilatory S metabolism
Organisms with dissimilatory metabolism use inorganic S compounds as electron donors in the redox process to obtain energy, while assimilatory metabolism serves to integrate S into bioorganic compounds. Dissimilatory metabolism can be divided into oxidative and reductive processes. Dissimilatory S-oxidizing organisms are found in the Archaea and the Bacteria domains, and they are distinguished by phototropic and chemolithotrophic aerobic as well as anaerobic life styles (Dahl et al., 2008). The most common substrates are S2–, elemental S, thiosulfate, and tetrathionate. Further diversification of these dissimilatory S oxidizers are the enzymatic systems involved in these oxidation reactions. The sulfur oxidoreductase (Sox) enzyme system is the main mechanism for oxidation of thiosulfate, but it can also use other substrates and transfers electrons to cytochrome c (Friedrich et al., 2001). Another common enzymatic mechanism is the dissimilatory sulfite reductase (Dsr) system, using various substrates to produce SO32– to be further oxidized to SO42–, usually through the activity of the dissimilatory APR and ATPS, generating ATP (Dahl et al., 2005). The operon structure and compositions of the Dsr system vary across different taxa (Anantharaman et al., 2018). In addition, S2– can also be oxidized by SQR, as in anoxygenic photosynthesis (Fig. 3). None of these enzymes for oxidative dissimilatory S metabolism, however, has found its way to being integrated into plants. However, sulfide is also oxidized in plants, by the mitochondrial sulfur dioxygenase Ethylmalonic Encephalopathy Protein1 (ETHE1), which is important for detoxification of sulfide in the mitochondria to prevent the inhibition of cytochrome c oxidase (Birke et al., 2012; Krüßel et al., 2014).
Fig. 3.
Dissimilatory sulfur metabolism. In dissimilatory sulfate reduction (in black arrows), SO42– is activated by ATP sulfurylase (ATPS) to APS, reduced to SO32– by dissimilatory APS reductase (APR), and further to S2– by dissimilatory sulfite reductase (DSR). Dissimilatory sulfite oxidation (in red arrows) uses various sulfur sources (sulfide, thiosulfate, elemental S, polysulfides, etc.) interconverted by different enzymes and complexes, such as sulfur oxidoredutase (SOX), sulfide:quinone oxidoreductase (SQR), and polysulfide reductase (PSR). S2– is then oxidized to SO32– and SO42– by DSR, APR, and ATPS.
The core dissimilatory SO42– reducing metabolism requires three enzymatic steps identical to the assimilatory SO42– reduction, namely activation of SO42– to APS, reduction of APS to SO32–, and reduction of SO32– to S2– (Fig. 3). The structure of these enzymes and the reaction mechanisms in dissimilatory SO42– reducers, however, are more similar to those identified in S-oxidizing organisms. This is particularly true for the dissimilatory APR and Dsr. The dissimilatory APR is formed by two subunits, AprA containing FAD cofactors and AprB binding two [4Fe–4S] clusters (Fritz et al., 2000). In contrast, the assimilatory APR or the PAPS reductases are structurally unrelated to AprA and AprB, even though some of them possess an Fe–S center (Patron et al., 2008) (the sulfonucleotide specificities of these assimilatory enzymes are discussed in the section ‘APS/PAPS-dependent SO42– reduction’ below).
The core of the SiR from dissimilatory organisms is also composed of two subunits DsrA and DsrB, while it requires a DsrC protein to bind the S atom transitioning the redox reactions. DsrAB binds four sirohemes and eight [4Fe–4S] clusters as prosthetic groups, although two of the sirohemes might be substituted by sirohydrochlorins, the metal-free form of siroheme (Oliveira et al., 2011). The assimilatory SiRs, on the other hand, exist in at least three different forms. Siroheme is an essential cofactor for SiR and is present in all three assimilatory forms. The monomeric plant SiR is dependent on the function of ferredoxins providing electrons and contains a siroheme and an [4Fe–4S] cluster. In contrast, the bacterial SiR obtains its electrons from NADPH, and it is an oligomer of flavoprotein CysJ catalytic subunits and CysI proteins binding a siroheme and an [4Fe–4S] cluster. SiRs in yeasts and many other fungi are again different, composed of two distinct subunits binding a siroheme, FAD, and FMN (Patron et al., 2008).
ATPS for SO42– activation is also found in different forms, while the core enzyme is less varied than the other two enzymes, except for the bacterial assimilatory ATPS. In addition, unlike the other two enzymes APR and SiR, ATPS is prone to different fusions, mainly with APK. In bacteria, ATPS is formed by two subunits, the CysD ATP pyrophosphatase and the CysN GTPase. Plants, green algae and cyanobacteria, and yeast and fungi possess different forms of a single protein ATPS (Patron et al., 2008). The phylogenetic relationships of these single protein ATPSs are, however, rather unexpected, because plant ATPSs are more similar to the animal counterparts than the green algal ATPS that probably originated from cyanobacteria (Patron et al., 2008).
S metabolism in plants and animals
Major distinctions and interkingdom variations in S metabolism can also be found between the assimilatory SO42– reducers, such as algae and plants, and animals that are unable to reduce SO42– and are dependent on organic S compounds in their nutrition. While plants metabolize SO42– in both the reductive S assimilation pathway to Cys and the oxidized S assimilation pathway to PAPS transferring the SO42– group to bioorganic compounds, animals possess only the latter pathway (Gunal et al., 2019). Thus, the entry of SO42– into metabolism, namely activation of SO42– by adenylation, is conserved between plants and animals as are the functions of the two enzymes, ATPS and APK. However, unlike plants, animals do not possess the enzymes APR and SiR and are thus unable to reduce SO42–. Interestingly, despite the evolutionary distance and different arrangement of SO42– assimilation, ATPS and APK from plants and animals are structurally conserved, as demonstrated by a greater extent of similarities of the plant-derived ATPSs found with the homologous enzymes from animals than with those of green algal and cyanobacterial origins (Patron et al., 2008).
Despite the structural similarities, the redox regulation of the APK enzyme appears to have evolved specifically in land plants, with conserved Cys residues playing pivotal roles in switching the enzyme activity on and off as they change the redox states and hence the protein conformation (Ravilious et al., 2012). In plants, the mechanism to control the APK enzyme activity may well be paralleled with redox regulations of the chloroplast-localizing APR isoenzyme (Bick et al., 2000) and the SAL1 3ʹ,5ʹ-bisphosphate nucleotidase (Chan et al., 2016). Under oxidative stress, the inhibition of the redox-sensitive APK enzyme may increase the flux of APS reduction and subsequent reductive S assimilation that leads to Cys biosynthesis by taking advantage of the oxidatively activated form of the APR isoenzyme (Bick et al., 2000), but limit the S metabolic flux toward PAPS synthesis that couples with the activity of the PAPS/PAP exchanger, sulfation of specialized metabolites in the cytosol and the Golgi, and dephosphorylation of PAP through the oxidation-sensitive SAL1 enzyme in the chloroplast (Gigolashvili et al., 2012; Chan et al., 2016, 2019). The proposed regulatory mechanism thus facilitates switching the flux of S at the branching point toward the reductive S assimilation pathway to meet the demand for synthesis of reduced GSH when plants are exposed to oxidative stress. The conserved Cys residues for the redox control of the APK enzyme appear to have been established after the emergence of land plant species, in conjunction with the bifurcating S assimilation pathway (Fig. 1) (Ravilious et al., 2012). It remains to be investigated how these redox control mechanisms in the oxidized and reductive S assimilation pathways are coordinated with the postulated function of PAP as a retrograde signal of oxidative stress from the plastids to the nucleus (Estavillo et al., 2011; Chan et al., 2016, 2019).
The retention of the oxidized S assimilation pathway—SO42– activation to APS and subsequent synthesis of PAPS—indicates an expansion of biological sulfation in animals. Animals produce a wide variety of sulfated products, including hormones, carbohydrates, and proteins, through the aid of the gene family of SOT with broader functional diversities than in plants (Coughtrie, 2016; Gunal et al., 2019). In fact, the common occurrence of sulfated extracellular carbohydrates in humans, such as chondroitin sulfate or heparan sulfate, is not observed in plants, where no sulfated carbohydrates from cell walls have been isolated (Gunal et al., 2019). Both plants and animals modify Tyr in the polypeptide chain by sulfation; however, in plants, this modification is present only in small peptides, and the sulfation of proteins as a mechanism for post-translational regulation is most probably confined to animals as it has not yet been described in plants (Kaufmann and Sauter, 2019; Stewart and Ronald, 2022). The sulfated small peptides play various roles in control of plant development and growth, such as those represented by function of the phytosulfokines, the Casparian strip integrity factors CIF1 and CIF2, root meristem growth factor RGF, or TWISTED SEED1 (Matsuzaki et al., 2010; Doblas et al., 2017; Nakayama et al., 2017; Doll et al., 2020).
Another major difference between plants and animals is found in pathways that serve for synthesis of S-containing amino acids (Fig. 4). Plants synthesize Cys directly from S2– and OAS, and use Cys for synthesis of homocysteine and Met through the trans-sulfuration and methylation pathways (Giovanelli et al., 1985; Hesse and Hoefgen, 2003; Hesse et al., 2004). The thiol group of Cys reacts with O-phosphohomoserine, the activated form of homoserine, to form cystathionine in the γ-replacement reaction catalyzed by cystathionine γ-synthase. Cystathionine then splits into homocysteine, pyruvate, and ammonium by cystathionine β-lyase. The thiol group of homocysteine is then methylated by methionine synthase to form Met. In contrast, the Cys synthase complex is not present in animals. They instead form Cys from Met through the transmethylation and reverse trans-sulfuration pathway (Brosnan and Brosnan, 2006). The pathway involves activation of Met to S-adenosylmethionine, transfer of the methyl group to methyl acceptor substrates, and hydrolysis of the resultant S-adenosylhomocysteine to homocysteine. The thiol group of homocysteine then joins with serine to form cystathionine in the β-replacement reaction catalyzed by cystathionine β-synthase. The cystathionine subsequently becomes the substrate for cystathionine γ-lyase breaking it down into Cys, α-ketobutyrate, and ammonium (Brosnan and Brosnan, 2006). Thus, plants and animals direct the flow of metabolism between the two S-containing amino acids in an opposite manner (Fig. 4).
Fig. 4.
Methionine–cysteine interconversions in plants and animals. Met synthesis in plants (upper panel) and trans-sulfuration pathway in animals (lower panel). CGS, cystathionine γ-synthase, CBL, cystathionine β-lyase; MS, methionine synthase; MAT, methionine adenosyltransferase; MT, methyltransferase; SAHH, S-adenosylhomocysteine hydrolase; CBS, cystathionine β-synthase; CSE, cystathionine γ-lyase; Acc, acceptor; CH3-Acc, methylated acceptor; THF, tetrahydrofolate.
Despite these distinctions of the S metabolic pathways, protein persulfidation on exposure to H2S appears consequential to regulation of enzymes in both plants and animals, and persulfidation is common to all kingdoms of life (Ogata et al., 2023). The discovery of the signaling function of S2– and its prevalent roles in human and animal diseases led to establishing similar concepts in plants. Recent studies in fact provide implications that S2– can act as a stress protectant and control a number of important physiological processes. Sulfide triggers stomatal closure and inhibits autophagy induction by ATG4 inactivation (Laureano-Marín et al., 2020; Scuffi et al., 2014; Aroca et al., 2020). However, the greatest difference from the animal system is that, in plants, S2– is an intermediate of the S assimilation pathway determining the flux of Cys biosynthesis (Calderwood and Kopriva, 2014). The full understanding of the proposed multifaceted roles of S2– in plants is still lacking.
APS/PAPS-dependent SO42– reduction
The reduction of activated SO42– is another step of the S assimilation pathway that underwent diversification in different organisms. The major difference is the nature of the substrate sulfonucleotide, APS or PAPS, as the activated form of SO42– being reduced, and the corresponding enzyme APS reductase or PAPS reductase, respectively, catalyzing the reactions (Kopriva and Koprivova, 2004; Carroll et al., 2005) (Fig. 5). The mechanism of SO42– reduction and the identity of the substrate and the enzyme have been controversially discussed especially in the plant community (Kopriva and Koprivova, 2004).
Fig. 5.
Diversity of enzymes catalyzing APS reduction in S assimilation. The schematic representation of enzymes reducing activated SO42– in bacteria and cyanobacteria, yeast and fungi, basal plants, flowering plants, and eukaryotic microalgae, namely the diatom Thalassiosira pseudonana (top) and the dinoflagellate Heterocapsa triquetra (bottom). Orange-colored APR enzymes mark the presence of the Fe–S clusters. Trx, thioredoxin or thioredoxin-like domain.
Since the reductive S assimilation pathway was first resolved in unicellular model organisms, Escherichai coli and yeast, both of which use PAPS and PAPS reductase, and because of the fact that the PAPS reductase activity was also present in photosynthesizing organisms, such as the cyanobacterium Synechococcus PCC 6301 (Peck, 1961; Wagner et al., 1978), the PAPS-dependent pathway was long considered universally responsible for the reduction of SO42– and believed to have also been adopted by plants. However, a later discovery of the APS-dependent activity in green algae and plants with SO32– bound to a carrier as a reaction product led to the identification of the corresponding enzyme named APS sulfotransferase (Schmidt, 1972). The bacterial and fungal PAPS reductase utilizes thioredoxin as an electron donor. In contrast, the plant APS sulfotransferase was shown to be independent of thioredoxin. In addition, APS sulfotransferase was regulated by a number of environmental conditions and metabolites, and was shown to play an important role in control of S assimilation in plants and algae (Brunold, 1990). When the corresponding gene for this enzyme was first identified in plants, the APS sulfotransferase was found to be a protein with two domains, one similar to bacterial PAPS reductases and the other a thioredoxin-like domain (Gutierrez-Marcos et al., 1996; Setya et al., 1996). The latter domain has a function of glutaredoxin, reflected in the ability of GSH to serve as the reductant for the enzyme (Bick et al., 1998). Subsequently, the enzyme was shown to carry an Fe–S center as a cofactor, and therefore, it was renamed as APS reductase (APR) (Kopriva et al., 2001) to be an enzyme specific to eukaryotes. However, a number of bacterial taxa demonstrated activities to utilize APS in SO42– reduction as well, but in this case using thioredoxin as an electron donor (Abola et al., 1999; Bick et al., 2000). Consequently, the presence of the Fe–S center was postulated to be the key factor determining the APS or PAPS specificity of the enzyme (Kopriva et al., 2002). However, this view changed when an APS-dependent form without the thioredoxin domain and the Fe–S cluster, APR-B, was found in the moss Physcomitrium patens (Kopriva et al., 2007; Stevenson et al., 2013). The APS and PAPS reductases have a common reaction mechanism, transferring the activated SO42– group from the sulfonucleotide to a conserved Cys residue of the protein, forming a stable intermediate, which is then reductively released as free SO32– by thioredoxin, glutaredoxin, or the C-terminal domain of the plant enzyme (Weber et al., 2000; Carroll et al., 2005).
The diversity of the sulfonucleotide reductases (Fig. 5) has been further extended by identification of genes in eukaryotic microalgae with APR-B fused to thioredoxin, or with APR being fused to ATPS (Patron et al., 2008). However, it seems that among the eukaryotes the division of these reductases is very clear, as algae and plants exclusively use APS-dependent enzymes for SO42– reduction, while yeast and fungi use PAPS reductase. Among prokaryotes, the boundary is not that simple: the APS-dependent pathway seems to be more widespread among different taxa, while the PAPS-dependent pathway is confined to γ-proteobacteria and cyanobacteria, in both of which however we may still find APS-reducing taxa (Patron et al., 2008). The PAPS-dependent reduction might have evolved from APR, possibly through an adaptation to Fe- and/or S-poor conditions. Limited Fe availability might have led to the replacement of the Fe–S-requiring mechanism with a costly alternative that would instead require one additional ATP for SO42– reduction. Supporting this hypothesis, the dissimilatory SO42–-reducing bacteria use APS reductase with Fe–S clusters.
Speciation of S metabolism at the species scale
S-containing specialized metabolites
The prime example of species-specific diversity of S metabolism are S-containing specialized metabolites. Plants produce various S-containing natural products—some highly specialized only in a few species, while some are more common (Fig. 6). S in these S-containing specialized metabolites can be present in oxidized or reduced form. The oxidized S modifications in various metabolites are products of biological sulfations catalyzed by a variety of SOTs (Hirschmann et al., 2014). These compounds have important and varied functions in different aspects of plant physiology and environmental interactions, and are thus often variable not only on the species level but also within ecotypes of the same species (Kliebenstein et al., 2001a). The full extent of S-containing plant metabolites is not fully described yet; even in the model plant species Arabidopsis thaliana, a large number of detected S-containing metabolites could not be structurally identified (Gläser et al., 2014).
Fig. 6.
Examples of S-containing specialized metabolites in cyanobacteria and plants. Sulfur atoms are marked green when they originate from reduced S assimilation or orange when they result from oxidized metabolism using sulfation.
Glucosinolates
Glucosinolates (GLSs) are the best characterized group of S-containing specialized metabolites specific to a small number of taxa as they are only found in the Brassicaceae (Halkier and Gershenzon, 2006). GLSs are amino acid-derived compounds with important roles in defense and S homeostasis in plants possessing S in both oxidized and reduced forms (Fig. 6). The ~170 GLS structures are divided into three classes depending on the amino acids they originate from; that is, the aliphatic GLS from amino acids Met, Ala, Leu, Ile, and Val (e.g. glucoraphanin, Fig. 6); the indolic GLSs from Trp (e.g. glucobrassicin); and the aromatic GLSs from Phe (e.g. gluconasturtiin) (Fahey et al., 2001; Sonderby et al., 2010). While the core structure of GLSs with an N-sulfated imine linked through a thioglucoside bond with glucose is identical, the modifications of the amino acid-derived side chain give the structural variety (Sonderby et al., 2010). One level of such variations occurs through chain elongation of Met before they enter the core pathway for GLS synthesis, while the others involve decorative modifications, such as hydroxylation of the amino acid-derived side chains and methylation of the hydroxyl group.
GLSs accumulated in plant organs remain inactive until exposed to and activated by myrosinases, which hydrolyze the thioglucoside bond, causing the aglycone to rearrange into the volatile isothiocyanate, nitrile, or epithionitrile (Halkier and Gershenzon, 2006). The chemical identity of the volatile products depends on the presence and nature of specifier proteins (Kuchernig et al., 2012). Traditionally, because of these volatile cocktails, GLSs were primarily seen as chemical defense compounds against insects and other herbivores, but also as attractants for specialized insects or as the basis for tritropic interactions (Ratzka et al., 2002). However, GLSs also serve an important role in plant immunity triggered by fungal and bacterial pathogens, and recently they were also shown to affect the plant microbiome (Bednarek et al., 2009; Clay et al., 2009; Fan et al., 2011; Jacoby et al., 2021). In particular, the indolic GLSs play an essential role in plant interactions with both pathogenic and beneficial fungi. Some GLSs provide clear health benefit to humans, while on the other hand, high GLS content in feed is detrimental for animal weight gains (Traka and Mithen, 2009). Correspondingly, a high GLS broccoli variety has been developed with increased content of glucoraphanin, the precursor of sulforaphane with anti-carcinogenic properties, while modern oilseed rape varieties have low GLS content instead enabling the use of the seed press cakes for animal feed (Mithen et al., 2003). Beside defense, herbivory, and microbe interactions, GLSs as S-rich compounds have been considered to represent the S storage in plants; however, the S recycling pathway from GLSs to primary metabolism has been shown only recently (Sugiyama et al., 2021).
GLS profiles differ not only across the plant species but also within the ecotypes of a single species. Such a difference has been instrumental in dissection of the regulation of their synthesis and identification of both regulatory and biosynthetic genes. The qualitative and quantitative variation in GLS accumulation in Arabidopsis, particularly of the aliphatic class of GLSs, is generally explainable by four genetic loci. GS-Elong, containing the methylthioalkylmalate synthase (MAM) genes, controls chain elongation of the aliphatic GLSs as they are responsible for reactions cyclically adding a carbon to Met and subsequently to the chain-elongated forms (Kroymann et al., 2001). In Arabidopsis and Brassica, two MAM genes control synthesis of the aliphatic GLSs with either three or four carbon molecules (Abrahams et al., 2020). However, long-chain aliphatic GLSs are also produced in Arabidopsis with C7 and C8, and in other Brassicaceae species with up to C10 and C11 side chains (Fahey et al., 2001). Another major quantitative trait locus (QTL) is caused by variation in 2-oxoacid-dependent dioxygenase AOP2 that converts methylsulfinyl GLSs to alkenyl GLSs (Kliebenstein et al., 2001b). Interestingly, AOP2 affects not only qualitative but also quantitative accumulation of GLSs (Chan et al., 2011). Variation of a neighboring gene AOP3 is responsible for production or absence of 3-hydroxypropyl GLS (Kliebenstein et al., 2001a). Interestingly, due to the duplications and rearrangements of the AOP3 locus in A. thaliana, 3-hydroxypropyl GLS accumulates in the seeds of all accessions but is much more limited in the leaves (Chan et al., 2011). The fourth QTL, GS-OH, is responsible for the ability to hydroxylate alkenyl GLS, for which the causal gene has been identified to encode 2-oxoacid-dependent dioxygenase unrelated to the AOP2/AOP3 enzymes (Hansen et al., 2008). The allelic configuration at these four loci can predict profiles of the aliphatic GLSs in Arabidopsis and partly in Brassicas, but other polymorphisms are responsible for further variation in other species, such as the BCMA locus in Boechera stricta, which controls production of the branched-chain aliphatic GLSs (Prasad et al., 2012). Interestingly, much less is known about the genes underlying variation in the indolic GLSs. Despite several QTLs being identified, only one gene, the CYP81F2 converting indol-3-yl-methylglucosinolate (I3M) to 4-hydroxy-indol-3-yl-methylglucosinolate (4OH-I3M), has been cloned as the causal gene for the Indole Glucosinolate Modifier1 (IGM1) locus (Pfalz et al., 2009).
Sulfoflavonoids and cysteine sulfoxides
Another large group of sulfated metabolites beside GLSs are the sulfoflavonoids (Fig. 6). Sulfoflavonoids have been identified in ~250 species in different forms and with different numbers of SO42– residues (Barron et al., 1988; Teles et al., 2018). Flavonol-specific SOTs have been described in A. thaliana (Gidda and Varin, 2006; Hashiguchi et al., 2013) and in several species of the genus Flaveria (Varin et al., 1992; Marsolais and Varin, 1998). Since different Flaveria species, however, produce different combinations of sulfoflavonoids, these metabolites were considered as possible markers differentiating the type of photosynthesis (e.g. C3 versus C4; Hannoufa et al., 1994). However, a more recent study revealed that there is no obvious correlation between the profiles of sulfoflavonoids and the photosynthetic types (Kleinenkuhnen et al., 2019). While flavonoids are known to represent numerous functions in plants, such as pigmentation, photoprotection, developmental regulation, and plant–microbe interactions, the biological function(s) of sulfated flavonoids is mostly unknown (Teles et al., 2018; Kleinenkuhnen et al., 2019). The sulfoflavonoids, such as quercetin 3-sulfate, can form a substantial S pool in plant leaves, as determined in Flaveria pringlei, and may thus serve as S storage compounds like GLSs (Sugiyama et al., 2021). However, the mechanism of the remobilization of the sulfate group is unclear since a sulfatase enzyme seems to be lacking in plants. Alternatively, sulfoflavonoids could represent an overflow mechanism to maintain S homeostasis during excessive SO42– supply. Whether the distribution of sulfoflavonoids in different taxa follows any ecological and/or evolutionary pattern, however, remains to be elucidated. It has to be noted that sulfoflavonoids are also detectable in human tissues, produced by sulfation of flavonoids during Phase II of xenobiotic metabolism (Pai et al., 2001).
Not all S-containing specialized metabolites are sulfated. The flavor of the Allium vegetables, for example, is formed by a mix of organic sulfides, disulfides, and other volatiles containing reduced S (Block, 1985; Marcinkowska and Jelen, 2022). Similar to GLSs, these flavor compounds are formed after enzymatic degradation of the non-volatile precursor sulfoxides (Block, 1985). These precursor sulfoxides accumulate in the cytosol and, upon cellular damage, are metabolized by alliinase, a vacuole-sequestered C-S lyase, to allicin and sulfenic acids. These primary products are then rapidly converted into a variety of volatile sulfides, disulfides, and trisulfides, or more complex cyclic dithiins and sulfines (Block, 1992). Four precursor sulfoxides have been detected in varying amounts in different Allium species and accessions: S-allyl cysteine sulfoxide (alliin), S-methyl cysteine sulfoxide (methiin), S-propyl cysteine sulfoxide (propiin), and S-trans-prop-1-enyl cysteine sulfoxide (isoalliin), converted to a much greater variation of organic S compounds in garlic oils or extracts (Block, 1992; Jones et al., 2004). These variations in accumulation of the S compounds detected among the onion or garlic accessions are associated with their varying flavors, and antimicrobial and health properties. However, the list of sulfoxide precursors and active volatile compounds appears far from complete, and possible involvement of additional protein factors in control of altering the chemical compositions of Allium S volatiles, which may be paralleled with the specifier proteins for GLS degradation (Eisenschmidt-Bonn et al., 2019), remains to be investigated (Block, 1992; Jones et al., 2004; Avgeri et al., 2020).
Specialized S metabolites of cyanobacteria
Cyanobacteria also produce a range of specialized metabolites with divergent structures, such as peptides, polyketides (PKs), peptide/PK hybrids, alkanes, terpenes, etc., that are often limited to a few strains (Nunnery et al., 2010; Kehr et al., 2011). Many cyanobacterial specialized metabolites have biological activities as cyanotoxins, while some may also serve as potential therapeutics, such as the cryptophycins (Nunnery et al., 2010). Prominent examples of S-containing specialized metabolites in cyanobacteria are thiohistidines, such as ovothiol (Fig. 6) and ergothioneine, found in several cyanobacterial strains such as Microcystis aeruginosa and Cylindrospermopsis raciborskii (Liao and Seebeck, 2017; Brancaccio et al., 2021). Although their exact physiological role is not known, thiohistidines seem to protect cyanobacteria against reactive oxygen species (Liao and Seebeck, 2017; Brancaccio et al., 2021). Sulfated specialized metabolites might also be present in cyanobacterial strains. In fact, SOT homologs have been found in several operons of metabolic genes in cyanobacteria. One example of a cyanobacterial sulfated metabolite is the cyanotoxin cylindrospermopsin produced by Cylindrospermopsis raciborskii, Umezakia natans, Aphanizomenon ovalisporum, and Raphidiopsis curvata (Yang et al., 2021) (Fig. 6). Cylindrospermopsin is composed of a guanidine modified by a SO42– group and an additional uracil side chain (Ohtani et al., 1992), and it functions most probably as an allelopathic metabolite to inhibit the growth of competing organisms (Teneva et al., 2023). Thus, also in cyanobacteria, different lineages utilize S for synthesis of metabolites with specialized functions, adding to the large variation of S biochemistry in the photosynthetic organisms.
Localization of S assimilation in bundle sheath versus mesophyll cells
A different kind of metabolic diversification of S assimilation has been observed between the plant lineages concerning the spatial distribution of the responsible enzymes. The cell type-specific localization of enzymes has become a prominent topic in plant physiology since the discovery of C4 photosynthesis, as this photosynthetic mechanism is dependent on separation of primary CO2 assimilation by the phosphoenolpyruvate carboxylase in the mesophyll cells and the decarboxylation of the resulting C4 acids and recapture of the CO2 by the Rubisco enzyme expressed in the bundle sheath cells (Hatch and Slack, 1966; Slack et al., 1969). However, not only is the pathway for the CO2 assimilation differentially distributed between the mesophyll and bundle sheath cells of C4 plants, but so are the reductive assimilation pathways for two major mineral nutrients, NO3– and SO42–, confined to these specific cell types (Mellor and Tregunna, 1971; Moore and Black, 1979; Gerwick et al., 1980; Passera and Ghisi, 1982). In contrast to the NO3– reduction pathway found specifically in the mesophyll, the SO42– reduction pathway is localized to the bundle sheath cells of a number of C4 species (Gerwick et al., 1980). While this cell type-specific localization seems to be conserved, at least in a few independent C4 lineages, its significance is still unknown (Weckopp and Kopriva, 2014). In attempts to clarify the importance of spatial separation of the SO42– assimilation pathway for the evolution of C4 photosynthesis, C3–C4 intermediate species of the model genus Flaveria were employed for a comparative analysis, from which the separation was found to be confined to the monocot C4 species (Koprivova et al., 2001). This has been confirmed further by RNA-seq analyses of another dicot C4 species, Cleome gynandra, which showed no differential expression of genes for the SO42– assimilation pathway between the mesophyll and bundle sheath cells (Kulahoglu et al., 2014).
Surprisingly, however, a translatome analysis of a C3 species Arabidopsis revealed a strong enrichment of pathway enzymes for SO42– assimilation as well as GLS biosynthesis in bundle sheath cells (Aubry et al., 2014). The bundle sheath cell-specific expression of the GLS biosynthetic genes seems to be controlled by the MYB transcription factors MYB28 and MYB29, both characterized previously as controlling aliphatic GLS biosynthesis, as well as the basic helix–loop–helix (bHLH) factors MYC2, MYC3, and MYC4 that are essential for this pathway and interacting with MYB28, MYB29, and other MYB transcription factors otherwise responsible for indolic GLS biosynthesis (Dickinson et al., 2020). Interestingly, this trait appears to be conserved in C3 plants, because the monocot rice also showed the same spatial gene expression pattern for SO42– assimilation (Hua et al., 2021). However, the bundle sheath localization of SO42– assimilation seems to not always be conserved ubiquitously among the C3 monocots, as demonstrated by activities for corresponding enzymes found in both cell types in wheat (Schmutz and Brunold, 1984). Thus, the variation in the cell type-specific localization of S assimilation does exist in plants, differentiating the C4 species with good evidence to support the exclusive presence of the entire pathway in the bundle sheath cells of the C4 monocots while possibly there is no such spatial separation in the C4 dicots. In contrast, for the C3 species, not enough evidence beyond the two model plant species exists and the jury is thus still out on whether the observed bundle sheath localization found in Arabidopsis and rice can be generalized.
Natural variation in genes for SO42– assimilation in control of S homeostasis
Natural accessions of a plant species can differ not only by distinct profiles of specialized metabolites, such as GLSs, but also by SO42– accumulation. The variation in SO42– accumulation has been explored extensively to dissect genetic control mechanisms associated with SO42– assimilation and general S homeostasis in Arabidopsis. Indeed, a >5-fold difference in SO42– accumulation was observed between the recombinant inbred lines (RILs) of the Arabidopsis accessions, Bay-0 and Shahdara (Loudet et al., 2007), and the subsequent QTL analysis resulted in identifying two key genes controlling a large part of the variation. The causal genes associated with these QTLs encoded the major isoforms of the S assimilation pathway enzymes, APR2 and ATPS1 (Loudet et al., 2007; Koprivova et al., 2013). Interestingly, the variation in SO42– accumulation was governed by two different mechanisms modifying the activity and the quantity of the corresponding enzymes in Arabidopsis accessions, Bay-0 and Shahdara. The first variation identified was associated with the Shahdara haplotype of APR2 with a non-synonymous single nucleotide polymorphism (SNP) that changes Ala399 into Glu and thus inhibits the binding of the enzyme to the electron donor GSH (Loudet et al., 2007). On the other hand, the second variation was associated with a deletion of an enhancer element in the first intron of the ATPS1 gene that led to a reduction in the ATPS1 transcript levels in Bay-0 relative to Shahdara, although the primary amino acid sequences of ATPS1 in Bay-0 and Shahdara were identical (Koprivova et al., 2013). These weak alleles of APR2 and ATPS1 lowered the enzymatic activity and the presence of the corresponding proteins, respectively, and thus reduced the S metabolic flux through the S assimilation pathway and ultimately led to the accumulation of the entry substrate SO42– (Koprivova et al., 2013; Chao et al., 2014). Interestingly, the less active form of APR2 from Shahdara had only a limited effect on steady-state levels of Cys and GSH accumulations, unless challenged by a stress, such as an exposure to selenate (Loudet et al., 2007; Grant et al., 2011).
An independent screen for genes causing higher accumulation of S and SO42– identified another weak APR2 allele that replaces Gly266 with Arg, inactivating the enzyme activity in the Hod accession of Arabidopsis (Chao et al., 2014). In a reverse approach, several accessions with various APR2 haplotypes were tested for SO42– accumulation, revealing a third natural variation with the Phe265 to Ser replacement in Lov-5 and three other Swedish accessions (Chao et al., 2014). Also, for ATPS1, an allele with a non-synonymous SNP replacing Gly342 with Asp was found to inactivate the enzyme in the Naes-2 accession (Herrmann et al., 2014). While these inactivating SNPs were either unique or rare among the Arabidopsis accessions, the SNPs associated with dysfunction of the transcriptional enhancer in the first intron of ATPS1 including the Bay-0 allele were more frequently observed in the population (Koprivova et al., 2013). These findings implicate conserved influence and prevalence of natural variations in APR2 and ATPS1 over SO42– and total S accumulations. However, current datasets suggest other mechanisms to be possibly involved in S homeostasis, since even within the four accessions with the APR2 Phe265–Ser haplotype, a large variation in both S and SO42– content was still observed (Chao et al., 2014). Indeed, recent detailed analysis of S homeostasis in Arabidopsis accessions revealed a tight connection of foliar S levels with SO42– uptake, but also a correlation with GSH and GLS synthesis (de Jager et al., 2023).
Natural variations in the mitochondrial OAS-TL—also known as OASC—suggest an influence of Cys biosynthesis on SO42– accumulation. These OASC variations were initially demonstrated using associative transcriptomics, a variant approach of a genome-wide association study (GWAS) with Brassica napus (Koprivova et al., 2014), while two different haplotypes were later identified from the Arabidopsis population (Koprivova et al., 2022). The SO42– accumulation profiles associated with the SNPs identified from B. napus and Arabidopsis suggested that they could alter the function of the OASC differently. In B. napus the causal SNP for the Gln355–Arg replacement led to an increase in SO42– accumulation, while in A. thaliana the Lys81–Arg polymorphism was causative of a decrease in SO42– accumulation (Koprivova et al., 2022). Unlike APR2 and ATPS1, however, the mechanism by which OASC affects SO42– accumulation is still not well established. OAS-TL is an essential catalytic component of the Cys synthase complex whose enzymatic function is strictly controlled by the presence or absence of the substrates OAS and sulfide, modulating the flux of Cys biosynthesis. Findings of the arsenite-tolerant astol1 mutant of rice and the causal mutation of the chloroplast-localizing OAS-TL stabilizing the assembly of the Cys synthase complex suggest a mechanism associated with an organelle-localized OAS-TL altering SO42– accumulation in plants (Sun et al., 2021). The mitochondrial pathway is the major route for synthesis of OAS (Watanabe et al., 2008; Wirtz et al., 2012). How naturally occurring polymorphisms of the mitochondrial OAS-TL influence the activity and composition of the Cys synthase complex and how would they modulate the mitochondrial OAS levels that may take control of the postulated S-sensing function of the enzyme complex await further investigation.
Conclusion
The biochemical properties of S make it pre-destined to be involved in a great variety of reactions that are indispensable. Changes in S availability in different chemical forms and redox states shaped life on Earth, such as the expansion of groups of microorganisms using S compounds as their energy source or alteration of the phytoplankton composition in the oceans. The extant organisms display functional innovations and extensive adaptive evolution that took place for the same S metabolic processes. Metabolic diversification is found across all scales, from diverse and complementary features of S-containing specialized metabolites in different accessions of the same plant species to a reverse flow of S between Cys and Met in plants and animals. While our comprehension of plant S metabolism and its integration with general S homeostasis has increased immensely alongside growing traction toward genetic and genomic research, we know very little about how and under what evolutionary drivers the ancestral species pivoted to establish the great variety of the S metabolic pathways and enzyme functions. Delving into the evolutionary history, however, holds the promise of unlocking the potential for optimizing these metabolic pathways for the future, paving the way for advancements in enhancing crop productivity and resilience.
Contributor Information
Stanislav Kopriva, Institute for Plant Sciences, Cluster of Excellence on Plant Sciences (CEPLAS), University of Cologne, Zülpicher Str. 47b, D-50674 Cologne, Germany.
Parisa Rahimzadeh Karvansara, Institute of Molecular Photosynthesis, Cluster of Excellence on Plant Sciences (CEPLAS), Heinrich Heine University Düsseldorf, D-40225 Düsseldorf, Germany.
Hideki Takahashi, Department of Biochemistry and Molecular Biology, Michigan State University, East Lansing, MI 48824, USA.
Donald Ort, University of Illinois, USA.
Conflict of interest
The authors have no conflicts to declare.
Funding
The work of SK and PRK is supported by the Deutsche Forschungsgemeinschaft (DFG) under Germany’s Excellence Strategy—EXC 2048/1—project 390686111. SK is additionally funded by a DFG grant no. 426501900. HT acknowledges funding support from AgBioResearch USDA-NIFA (Hatch Project 1018575).
References
- Abola AP, Willits MG, Wang RC, Long SR.. 1999. Reduction of adenosine-5'-phosphosulfate instead of 3'-phosphoadenosine-5'-phosphosulfate in cysteine biosynthesis by Rhizobium meliloti and other members of the family Rhizobiaceae. Journal of Bacteriology 181, 5280–5287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abrahams RS, Pires JC, Schranz ME.. 2020. Genomic origin and diversification of the glucosinolate MAM locus. Frontiers in Plant Science 11, 711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anantharaman K, Hausmann B, Jungbluth SP, et al. 2018. Expanded diversity of microbial groups that shape the dissimilatory sulfur cycle. The ISME Journal 12, 1715–1728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aroca A, Gotor C, Bassham DC, Romero LC.. 2020. Hydrogen sulfide: from a toxic molecule to a key molecule of cell life. Antioxidants 9, 621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ashykhmina N, Lorenz M, Frerigmann H, Koprivova A, Hofsetz E, Stührwohldt N, Flügge U-I, Haferkamp I, Kopriva S, Gigolashvili T.. 2019. PAPST2 plays critical roles in removing the stress signaling molecule 3ʹ-phosphoadenosine 5ʹ-phosphate from the cytosol and its subsequent degradation in plastids and mitochondria. The Plant Cell 31, 231–249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aubry S, Smith-Unna RD, Boursnell CM, Kopriva S, Hibberd JM.. 2014. Transcript residency on ribosomes reveals a key role for the Arabidopsis thaliana bundle sheath in sulfur and glucosinolate metabolism. The Plant Journal 78, 659–673. [DOI] [PubMed] [Google Scholar]
- Avgeri I, Zeliou K, Petropoulos SA, Bebeli PJ, Papasotiropoulos V, Lamari FN.. 2020. Variability in bulb organosulfur compounds, sugars, phenolics, and pyruvate among Greek garlic genotypes: association with antioxidant properties. Antioxidants 9, 967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bangash SAK, Müller-Schüssele SJ, Solbach D, Jansen M, Fiorani F, Schwarzländer M, Kopriva S, Meyer AJ.. 2019. Low-glutathione mutants are impaired in growth but do not show an increased sensitivity to moderate water deficit. PLoS One 14, e0220589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barron D, Varin L, Ibrahim RK, Harborne JB, Williams CA.. 1988. Sulphated flavonoids—an update. Phytochemistry 27, 2375–2395. [Google Scholar]
- Bednarek P, Pislewska-Bednarek M, Svatos A, et al. 2009. A glucosinolate metabolism pathway in living plant cells mediates broad-spectrum antifungal defense. Science 323, 101–106. [DOI] [PubMed] [Google Scholar]
- Bick JA, Aslund F, Chen Y, Leustek T.. 1998. Glutaredoxin function for the carboxyl-terminal domain of the plant-type 5'-adenylylsulfate reductase. Proceedings of the National Academy of Sciences, USA 95, 8404–8409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bick JA, Dennis JJ, Zylstra GJ, Nowack J, Leustek T.. 2000. Identification of a new class of 5'-adenylylsulfate (APS) reductases from sulfate-assimilating bacteria. Journal of Bacteriology 182, 135–142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Birke H, Haas FH, De Kok LJ, Balk J, Wirtz M, Hell R.. 2012. Cysteine biosynthesis, in concert with a novel mechanism, contributes to sulfide detoxification in mitochondria of Arabidopsis thaliana. The Biochemical Journal 445, 275–283. [DOI] [PubMed] [Google Scholar]
- Block E. 1985. The chemistry of garlic and onions. Scientific American 252, 114–119. [DOI] [PubMed] [Google Scholar]
- Block E. 1992. The organosulfur chemistry of the genus allium—implications for the organic chemistry of sulfur. Angewandte Chemie 31, 1135–1178. [Google Scholar]
- Brancaccio M, Tangherlini M, Danovaro R, Castellano I.. 2021. Metabolic adaptations to marine environments: molecular diversity and evolution of ovothiol biosynthesis in bacteria. Genome Biology and Evolution 13, evab169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Breitinger U, Clausen T, Ehlert S, Huber R, Laber B, Schmidt F, Pohl E, Messerschmidt A.. 2001. The three-dimensional structure of cystathionine beta-lyase from Arabidopsis and its substrate specificity. Plant Physiology 126, 631–642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brosnan JT, Brosnan ME.. 2006. The sulfur-containing amino acids: an overview. Journal of Nutrition 136, 1636S–1640S. [DOI] [PubMed] [Google Scholar]
- Brunold C. 1990. Reduction of sulfate to sulfide. In: Rennenberg H, Brunold C, De Kok LJ, Stulen I, eds. Sulfur nutrition and sulfur assimilation in higher plants. The Hague, The Netherlands: SPB Academic Publishing, 13–31. [Google Scholar]
- Burini RC, Kano HT, Yu Y.. 2018. The life evolution on the sulfur cycle: from ancient elemental sulfur reduction and sulfide oxidation to the contemporary thiol-redox challenges. In: Erkekoglu P, Kocer-Gumusel B, eds. Glutathione in health and disease. London: IntechOpen. [Google Scholar]
- Cairns NG, Pasternak M, Wachter A, Cobbett CS, Meyer AJ.. 2006. Maturation of Arabidopsis seeds is dependent on glutathione biosynthesis within the embryo. Plant Physiology 141, 446–455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calderwood A, Kopriva S.. 2014. Hydrogen sulfide in plants: from dissipation of excess sulfur to signaling molecule. Nitric Oxide 41, 72–78. [DOI] [PubMed] [Google Scholar]
- Canfield DE. 1998. A new model for Proterozoic ocean chemistry. Nature 396, 450–453. [Google Scholar]
- Canfield DE, Raiswell R.. 1999. The evolution of the sulfur cycle. American Journal of Science 299, 697–723. [Google Scholar]
- Carroll KS, Gao H, Chen H, Stout CD, Leary JA, Bertozzi CR.. 2005. A conserved mechanism for sulfonucleotide reduction. PLoS Biology 3, e250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chan EK, Rowe HC, Corwin JA, Joseph B, Kliebenstein DJ.. 2011. Combining genome-wide association mapping and transcriptional networks to identify novel genes controlling glucosinolates in Arabidopsis thaliana. PLoS Biology 9, e1001125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chan KX, Mabbitt PD, Phua SY, et al. 2016. Sensing and signaling of oxidative stress in chloroplasts by inactivation of the SAL1 phosphoadenosine phosphatase. Proceedings of the National Academy of Sciences, USA 113, E4567–E4576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chan KX, Phua SY, Van Breusegem F.. 2019. Secondary sulfur metabolism in cellular signalling and oxidative stress responses. Journal of Experimental Botany 70, 4237–4250. [DOI] [PubMed] [Google Scholar]
- Chao DY, Baraniecka P, Danku J, Koprivova A, Lahner B, Luo H, Yakubova E, Dilkes B, Kopriva S, Salt DE.. 2014. Variation in sulfur and selenium accumulation is controlled by naturally occurring isoforms of the key sulfur assimilation enzyme ADENOSINE 5'-PHOSPHOSULFATE REDUCTASE2 across the Arabidopsis species range. Plant Physiology 166, 1593–1608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clay NK, Adio AM, Denoux C, Jander G, Ausubel FM.. 2009. Glucosinolate metabolites required for an Arabidopsis innate immune response. Science 323, 95–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cobbett C, Goldsbrough P.. 2002. Phytochelatins and metallothioneins: roles in heavy metal detoxification and homeostasis. Annual Review of Plant Biology 53, 159–182. [DOI] [PubMed] [Google Scholar]
- Cohen Y, Jorgensen BB, Revsbech NP, Poplawski R.. 1986. Adaptation to hydrogen sulfide of oxygenic and anoxygenic photosynthesis among cyanobacteria. Applied and Environmental Microbiology 51, 398–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coughtrie MWH. 2016. Function and organization of the human cytosolic sulfotransferase (SULT) family. Chemico-Biological Interactions 259, 2–7. [DOI] [PubMed] [Google Scholar]
- Cuevasanta E, Moller MN, Alvarez B.. 2017. Biological chemistry of hydrogen sulfide and persulfides. Archives of Biochemistry and Biophysics 617, 9–25. [DOI] [PubMed] [Google Scholar]
- Dahl C, Engels S, Pott-Sperling AS, Schulte A, Sander J, Lubbe Y, Deuster O, Brune DC.. 2005. Novel genes of the dsr gene cluster and evidence for close interaction of Dsr proteins during sulfur oxidation in the phototrophic sulfur bacterium Allochromatium vinosum. Journal of Bacteriology 187, 1392–1404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dahl C, Friedrich C, Kletzin A.. 2008. Sulfur oxidation in prokaryotes. In: Encyclopedia of life sciences. Chichester, UK: John Wiley & Sons, Ltd. [Google Scholar]
- de Beer D, Weber M, Chennu A, Hamilton T, Lott C, Macalady J, Klatt JM.. 2017. Oxygenic and anoxygenic photosynthesis in a microbial mat from an anoxic and sulfidic spring. Enviromental Microbiology 19, 1251–1265. [DOI] [PubMed] [Google Scholar]
- de Jager N, Shukla V, Koprivova A, Lycka M, Bilalli L, You Y, Zeier J, Kopriva S, Ristova D.. 2023. Traits linked to natural variation of sulfur content in A. thaliana leaf. Journal of Experimental Botany 75, 1036–1050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dickinson PJ, Knerova J, Szecowka M, Stevenson SR, Burgess SJ, Mulvey H, Bagman AM, Gaudinier A, Brady SM, Hibberd JM.. 2020. A bipartite transcription factor module controlling expression in the bundle sheath of Arabidopsis thaliana. Nature Plants 6, 1468–1479. [DOI] [PubMed] [Google Scholar]
- Doblas VG, Smakowska-Luzan E, Fujita S, Alassimone J, Barberon M, Madalinski M, Belkhadir Y, Geldner N.. 2017. Root diffusion barrier control by a vasculature-derived peptide binding to the SGN3 receptor. Science 355, 280–284. [DOI] [PubMed] [Google Scholar]
- Doll NM, Royek S, Fujita S, et al. 2020. A two-way molecular dialogue between embryo and endosperm is required for seed development. Science 367, 431–435. [DOI] [PubMed] [Google Scholar]
- Droux M, Ruffet ML, Douce R, Job D.. 1998. Interactions between serine acetyltransferase and O-acetylserine (thiol) lyase in higher plants—structural and kinetic properties of the free and bound enzymes. European Journal of Biochemistry 255, 235–245. [DOI] [PubMed] [Google Scholar]
- Eisenschmidt-Bonn D, Schneegans N, Backenkohler A, Wittstock U, Brandt W.. 2019. Structural diversification during glucosinolate breakdown: mechanisms of thiocyanate, epithionitrile and simple nitrile formation. The Plant Journal 99, 329–343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Estavillo GM, Crisp PA, Pornsiriwong W, et al. 2011. Evidence for a SAL1–PAP chloroplast retrograde pathway that functions in drought and high light signaling in Arabidopsis. The Plant Cell 23, 3992–4012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fahey JW, Zalcmann AT, Talalay P.. 2001. The chemical diversity and distribution of glucosinolates and isothiocyanates among plants. Phytochemistry 56, 5–51. [DOI] [PubMed] [Google Scholar]
- Fan J, Crooks C, Creissen G, Hill L, Fairhurst S, Doerner P, Lamb C.. 2011. Pseudomonas sax genes overcome aliphatic isothiocyanate-mediated non-host resistance in Arabidopsis. Science 331, 1185–1188. [DOI] [PubMed] [Google Scholar]
- Friedrich CG, Rother D, Bardischewsky F, Quentmeier A, Fischer J.. 2001. Oxidation of reduced inorganic sulfur compounds by bacteria: emergence of a common mechanism? Applied and Environmental Microbiology 67, 2873–2882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fritz G, Buchert T, Huber H, Stetter KO, Kroneck PM.. 2000. Adenylylsulfate reductases from archaea and bacteria are 1:1 alphabeta-heterodimeric iron–sulfur flavoenzymes—high similarity of molecular properties emphasizes their central role in sulfur metabolism. FEBS Letters 473, 63–66. [DOI] [PubMed] [Google Scholar]
- Gerwick BC, Ku SB, Black CC.. 1980. Initiation of sulfate activation: a variation in C4 photosynthesis plants. Science 209, 513–515. [DOI] [PubMed] [Google Scholar]
- Gidda SK, Varin L.. 2006. Biochemical and molecular characterization of flavonoid 7-sulfotransferase from Arabidopsis thaliana. Plant Physiology and Biochemistry 44, 628–636. [DOI] [PubMed] [Google Scholar]
- Gigolashvili T, Geier M, Ashykhmina N, Frerigmann H, Wulfert S, Krueger S, Mugford SG, Kopriva S, Haferkamp I, Flügge U-I.. 2012. The Arabidopsis thylakoid ADP/ATP carrier TAAC has an additional role in supplying plastidic phosphoadenosine 5'-phosphosulfate to the cytosol. The Plant Cell 24, 4187–4204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giovanelli J, Mudd SH, Datko AH.. 1985. Quantitative analysis of pathways of methionine metabolism and their regulation in Lemna. Plant Physiology 78, 555–560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gläser K, Kanawati B, Kubo T, Schmitt-Kopplin P, Grill E.. 2014. Exploring the Arabidopsis sulfur metabolome. The Plant Journal 77, 31–45. [DOI] [PubMed] [Google Scholar]
- Grant K, Carey NM, Mendoza M, Schulze J, Pilon M, Pilon-Smits EA, van Hoewyk D.. 2011. Adenosine 5'-phosphosulfate reductase (APR2) mutation in Arabidopsis implicates glutathione deficiency in selenate toxicity. The Biochemical Journal 438, 325–335. [DOI] [PubMed] [Google Scholar]
- Gunal S, Hardman R, Kopriva S, Mueller JW.. 2019. Sulfation pathways from red to green. Journal of Biological Chemistry 294, 12293–12312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gutierrez-Marcos JF, Roberts MA, Campbell EI, Wray JL.. 1996. Three members of a novel small gene-family from Arabidopsis thaliana able to complement functionally an Escherichia coli mutant defective in PAPS reductase activity encode proteins with a thioredoxin-like domain and ‘APS reductase’ activity. Proceedings of the National Academy of Sciences, USA 93, 13377–13382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Halkier BA, Gershenzon J.. 2006. Biology and biochemistry of glucosinolates. Annual Review of Plant Biology 57, 303–333. [DOI] [PubMed] [Google Scholar]
- Hannoufa A, Brown RHH, Ibrahim RK.. 1994. Variations in flavonoid sulphate patterns in relation to photosynthetic types of five Flaveria species. Phytochemistry 36, 353–356. [Google Scholar]
- Hansen BG, Kerwin RE, Ober JA, Lambrix VM, Mitchell-Olds T, Gershenzon J, Halkier BA, Kliebenstein DJ.. 2008. A novel 2-oxoacid-dependent dioxygenase involved in the formation of the goiterogenic 2-hydroxybut-3-enyl glucosinolate and generalist insect resistance in Arabidopsis. Plant Physiology 148, 2096–2108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hashiguchi T, Sakakibara Y, Hara Y, Shimohira T, Kurogi K, Akashi R, Liu MC, Suiko M.. 2013. Identification and characterization of a novel kaempferol sulfotransferase from Arabidopsis thaliana. Biochemical and Biophysical Research Communications 434, 829–835. [DOI] [PubMed] [Google Scholar]
- Hatch MD, Slack CR.. 1966. Photosynthesis by sugar-cane leaves. A new carboxylation reaction and the pathway of sugar formation. The Biochemical Journal 101, 103–111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hell R, Wirtz M.. 2011. Molecular biology, biochemistry and cellular physiology of cysteine metabolism in Arabidopsis thaliana. The Arabidopsis Book 9, e0154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herrmann J, Ravilious GE, McKinney SE, Westfall CS, Lee SG, Baraniecka P, Giovannetti M, Kopriva S, Krishnan HB, Jez JM.. 2014. Structure and mechanism of soybean ATP sulfurylase and the committed step in plant sulfur assimilation. Journal of Biological Chemistry 289, 10919–10929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hesse H, Hoefgen R.. 2003. Molecular aspects of methionine biosynthesis. Trends in Plant Science 8, 259–262. [DOI] [PubMed] [Google Scholar]
- Hesse H, Kreft O, Maimann S, Zeh M, Hoefgen R.. 2004. Current understanding of the regulation of methionine biosynthesis in plants. Journal of Experimental Botany 55, 1799–1808. [DOI] [PubMed] [Google Scholar]
- Hirschmann F, Krause F, Papenbrock J.. 2014. The multi-protein family of sulfotransferases in plants: composition, occurrence, substrate specificity, and functions. Frontiers in Plant Science 5, 556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hua L, Stevenson SR, Reyna-Llorens I, Xiong H, Kopriva S, Hibberd JM.. 2021. The bundle sheath of rice is conditioned to play an active role in water transport as well as sulfur assimilation and jasmonic acid synthesis. The Plant Journal 107, 268–286. [DOI] [PubMed] [Google Scholar]
- Jacoby RP, Koprivova A, Kopriva S.. 2021. Pinpointing secondary metabolites that shape the composition and function of the plant microbiome. Journal of Experimental Botany 72, 57–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones MG, Hughes J, Tregova A, Milne J, Tomsett AB, Collin HA.. 2004. Biosynthesis of the flavour precursors of onion and garlic. Journal of Experimental Botany 55, 1903–1918. [DOI] [PubMed] [Google Scholar]
- Kaufmann C, Sauter M.. 2019. Sulfated plant peptide hormones. Journal of Experimental Botany 70, 4267–4277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kehr JC, Gatte Picchi D, Dittmann E.. 2011. Natural product biosyntheses in cyanobacteria: a treasure trove of unique enzymes. Beilstein Journal of Organic Chemistry 7, 1622–1635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klein M, Papenbrock J.. 2004. The multi-protein family of Arabidopsis sulphotransferases and their relatives in other plant species. Journal of Experimental Botany 55, 1809–1820. [DOI] [PubMed] [Google Scholar]
- Kleinenkuhnen N, Buchel F, Gerlich SC, Kopriva S, Metzger S.. 2019. A novel method for identification and quantification of sulfated flavonoids in plants by neutral loss scan mass spectrometry. Frontiers in Plant Science 10, 885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kliebenstein DJ, Kroymann J, Brown P, Figuth A, Pedersen D, Gershenzon J, Mitchell-Olds T.. 2001a. Genetic control of natural variation in Arabidopsis glucosinolate accumulation. Plant Physiology 126, 811–825. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kliebenstein DJ, Lambrix VM, Reichelt M, Gershenzon J, Mitchell-Olds T.. 2001b. Gene duplication in the diversification of secondary metabolism: tandem 2-oxoglutarate-dependent dioxygenases control glucosinolate biosynthesis in Arabidopsis. The Plant Cell 13, 681–693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kopriva S, Buchert T, Fritz G, et al. 2001. Plant adenosine 5'-phospho sulfate reductase is a novel iron–sulfur protein. Journal of Biological Chemistry 276, 42881–42886. [DOI] [PubMed] [Google Scholar]
- Kopriva S, Buchert T, Fritz G, et al. 2002. The presence of an iron–sulfur cluster in adenosine 5'-phosphosulfate reductase separates organisms utilizing adenosine 5'-phosphosulfate and phosphoadenosine 5'-phosphosulfate for sulfate assimilation. Journal of Biological Chemistry 277, 21786–21791. [DOI] [PubMed] [Google Scholar]
- Kopriva S, Fritzemeier K, Wiedemann G, Reski R.. 2007. The putative moss 3'-phosphoadenosine-5'-phosphosulfate reductase is a novel form of adenosine-5'-phosphosulfate reductase without an iron–sulfur cluster. Journal of Biological Chemistry 282, 22930–22938. [DOI] [PubMed] [Google Scholar]
- Kopriva S, Koprivova A.. 2004. Plant adenosine 5'-phosphosulphate reductase: the past, the present, and the future. Journal of Experimental Botany 55, 1775–1783. [DOI] [PubMed] [Google Scholar]
- Kopriva S, Mugford SG, Baraniecka P, Lee BR, Matthewman CA, Koprivova A.. 2012. Control of sulfur partitioning between primary and secondary metabolism in Arabidopsis. Frontiers in Plant Science 3, 163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koprivova A, Elkatmis B, Gerlich SC, Trick M, Harper AL, Bancroft I, Kopriva S.. 2022. Natural variation in OASC gene for mitochondrial O-acetylserine thiollyase affects sulfate levels in Arabidopsis. Plants 12, 35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koprivova A, Giovannetti M, Baraniecka P, Lee BR, Grondin C, Loudet O, Kopriva S.. 2013. Natural variation in the ATPS1 isoform of ATP sulfurylase contributes to the control of sulfate levels in Arabidopsis. Plant Physiology 163, 1133–1141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koprivova A, Harper AL, Trick M, Bancroft I, Kopriva S.. 2014. Dissection of the control of anion homeostasis by associative transcriptomics in Brassica napus. Plant Physiology 166, 442–450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koprivova A, Melzer M, von Ballmoos P, Mandel T, Brunold C, Kopriva S.. 2001. Assimilatory sulfate reduction in C3, C3–C4, and C4 species of Flaveria. Plant Physiology 127, 543–550. [PMC free article] [PubMed] [Google Scholar]
- Kroymann J, Textor S, Tokuhisa JG, Falk KL, Bartram S, Gershenzon J, Mitchell-Olds T.. 2001. A gene controlling variation in Arabidopsis glucosinolate composition is part of the methionine chain elongation pathway. Plant Physiology 127, 1077–1088. [PMC free article] [PubMed] [Google Scholar]
- Krüßel L, Junemann J, Wirtz M, et al. 2014. The mitochondrial sulfur dioxygenase ETHYLMALONIC ENCEPHALOPATHY PROTEIN1 is required for amino acid catabolism during carbohydrate starvation and embryo development in Arabidopsis. Plant Physiology 165, 92–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuchernig JC, Burow M, Wittstock U.. 2012. Evolution of specifier proteins in glucosinolate-containing plants. BMC Evolutionary Biology 12, 127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kulahoglu C, Denton AK, Sommer M, et al. 2014. Comparative transcriptome atlases reveal altered gene expression modules between two Cleomaceae C3 and C4 plant species. The Plant Cell 26, 3243–3260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laureano-Marín AM, Aroca A, Pérez-Pérez ME, Yruela I, Jurado-Flores A, Moreno I, Crespo JL, Romero LC, Gotor C.. 2020. Abscisic acid-triggered persulfidation of the cys protease ATG4 mediates regulation of autophagy by sulfide. The Plant Cell 32, 3902–3920. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liao C, Seebeck FP.. 2017. Convergent evolution of ergothioneine biosynthesis in cyanobacteria. ChemBioChem 18, 2115–2118. [DOI] [PubMed] [Google Scholar]
- Loudet O, Saliba-Colombani V, Camilleri C, Calenge F, Gaudon V, Koprivova A, North KA, Kopriva S, Daniel-Vedele F.. 2007. Natural variation for sulfate content in Arabidopsis thaliana is highly controlled by APR2. Nature Genetics 39, 896–900. [DOI] [PubMed] [Google Scholar]
- Marcinkowska MA, Jelen HH.. 2022. Role of sulfur compounds in vegetable and mushroom aroma. Molecules 27, 6116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marsolais F, Varin L.. 1998. Recent developments in the study of the structure–function relationship of flavonol sulfotransferases. Chemico-Biological Interactions 109, 117–122. [DOI] [PubMed] [Google Scholar]
- Maruyama-Nakashita A. 2017. Metabolic changes sustain the plant life in low-sulfur environments. Current Opinion in Plant Biology 39, 144–151. [DOI] [PubMed] [Google Scholar]
- Matsuzaki Y, Ogawa-Ohnishi M, Mori A, Matsubayashi Y.. 2010. Secreted peptide signals required for maintenance of root stem cell niche in Arabidopsis. Science 329, 1065–1067. [DOI] [PubMed] [Google Scholar]
- Mellor GE, Tregunna EB.. 1971. The localization of nitrate-assimilating enzymes in leaves of plants with the C4-pathway of photosynthesis. Canadian Journal of Botany 49, 137–142. [Google Scholar]
- Mithen R, Faulkner K, Magrath R, Rose P, Williamson G, Marquez J.. 2003. Development of isothiocyanate-enriched broccoli, and its enhanced ability to induce phase 2 detoxification enzymes in mammalian cells. Theoretical and Applied Genetics 106, 727–734. [DOI] [PubMed] [Google Scholar]
- Moore R, Black CC.. 1979. Nitrogen assimilation pathways in leaf mesophyll and bundle sheath cells of C4 photosynthesis plants formulated from comparative studies with Digitaria sanguinalis (L.) Scop. Plant Physiology 64, 309–313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mugford SG, Yoshimoto N, Reichelt M, et al. 2009. Disruption of adenosine-5'-phosphosulfate kinase in Arabidopsis reduces levels of sulfated secondary metabolites. The Plant Cell 21, 910–927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mulo P, Sicora C, Aro EM.. 2009. Cyanobacterial psbA gene family: optimization of oxygenic photosynthesis. Cellular and Molecular Life Sciences 66, 3697–3710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakayama T, Shinohara H, Tanaka M, Baba K, Ogawa-Ohnishi M, Matsubayashi Y.. 2017. A peptide hormone required for Casparian strip diffusion barrier formation in Arabidopsis roots. Science 355, 284–286. [DOI] [PubMed] [Google Scholar]
- Noctor G, Foyer CH.. 1998. Ascorbate and glutathione: keeping active oxygen under control. Annual Review of Plant Physiology and Plant Molecular Biology 49, 249–279. [DOI] [PubMed] [Google Scholar]
- Noctor G, Mhamdi A, Chaouch S, Han Y, Neukermans J, Marquez-Garcia B, Queval G, Foyer CH.. 2012. Glutathione in plants: an integrated overview. Plant, Cell & Environment 35, 454–484. [DOI] [PubMed] [Google Scholar]
- Nunnery JK, Mevers E, Gerwick WH.. 2010. Biologically active secondary metabolites from marine cyanobacteria. Current Opinion in Biotechnology 21, 787–793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ogata S, Matsunaga T, Jung M, Barayeu U, Morita M, Akaike T.. 2023. Persulfide biosynthesis conserved evolutionarily in all organisms. Antioxidants & Redox Signaling 39, 983–999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohtani I, Moore RE, Runnegar MTC.. 1992. Cylindrospermopsin: a potent hepatotoxin from the blue-green alga Cylindrospermopsis raciborskii. Journal of the American Chemical Society 114, 7941–7942. [Google Scholar]
- Oliveira TF, Franklin E, Afonso JP, Khan AR, Oldham NJ, Pereira IA, Archer M.. 2011. Structural insights into dissimilatory sulfite reductases: structure of desulforubidin from Desulfomicrobium norvegicum. Frontiers in Microbiology 2, 71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Olson KR. 2021. The biological legacy of sulfur: a roadmap to the future. Comparative Biochemistry and Physiology A 252, 110824. [DOI] [PubMed] [Google Scholar]
- Pai TG, Suiko M, Sakakibara Y, Liu MC.. 2001. Sulfation of flavonoids and other phenolic dietary compounds by the human cytosolic sulfotransferases. Biochemical and Biophysical Research Communications 285, 1175–1179. [DOI] [PubMed] [Google Scholar]
- Passera C, Ghisi R.. 1982. ATP sulphurylase and O-acetylserine sulphydrylase in isolated mesophyll protoplasts and bundle sheath strands of S-deprived maize leaves. Journal of Experimental Botany 33, 432–438. [Google Scholar]
- Patron NJ, Durnford DG, Kopriva S.. 2008. Sulfate assimilation in eukaryotes: fusions, relocations and lateral transfers. BMC Evolutionary Biology 8, 39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peck HD Jr. 1961. Enzymatic basis for assimilatory and dissimilatory sulfate reduction. Journal of Bacteriology 82, 933–939. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pfalz M, Vogel H, Kroymann J.. 2009. The gene controlling the indole glucosinolate modifier1 quantitative trait locus alters indole glucosinolate structures and aphid resistance in Arabidopsis. The Plant Cell 21, 985–999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prasad KV, Song BH, Olson-Manning C, et al. 2012. A gain-of-function polymorphism controlling complex traits and fitness in nature. Science 337, 1081–1084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rahimzadeh Karvansara P, Komenda J, Kopriva S.. 2023. Sulfur metabolism in cyanobacteria. In: Mishra AK, Singh SS, eds. Cyanobacteria: metabolisms to molecules. London: Elsevier, 117–157. [Google Scholar]
- Ratti S, Knoll AH, Giordano M.. 2011. Did sulfate availability facilitate the evolutionary expansion of chlorophyll a+c phytoplankton in the oceans? Geobiology 9, 301–312. [DOI] [PubMed] [Google Scholar]
- Ratzka A, Vogel H, Kliebenstein DJ, Mitchell-Olds T, Kroymann J.. 2002. Disarming the mustard oil bomb. Proceedings of the National Academy of Sciences, USA 99, 11223–11228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ravilious GE, Nguyen A, Francois JA, Jez JM.. 2012. Structural basis and evolution of redox regulation in plant adenosine-5'-phosphosulfate kinase. Proceedings of the National Academy of Sciences, USA 109, 309–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saito K. 2004. Sulfur assimilatory metabolism. The long and smelling road. Plant Physiology 136, 2443–2450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sauter M, Moffatt B, Saechao MC, Hell R, Wirtz M.. 2013. Methionine salvage and S-adenosylmethionine: essential links between sulfur, ethylene and polyamine biosynthesis. The Biochemical Journal 451, 145–154. [DOI] [PubMed] [Google Scholar]
- Schmidt A. 1972. On the mechanism of photosynthetic sulfate reduction. An APS-sulfotransferase from Chlorella. Archiv fur Mikrobiologie 84, 77–86. [DOI] [PubMed] [Google Scholar]
- Schmutz D, Brunold C.. 1984. Intercellular localization of assimilatory sulfate reduction in leaves of Zea mays and Triticum aestivum. Plant Physiology 74, 866–870. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scuffi D, Álvarez C, Laspina N, Gotor C, Lamattina L, García-Mata C.. 2014. Hydrogen sulfide generated by l-cysteine desulfhydrase acts upstream of nitric oxide to modulate abscisic acid-dependent stomatal closure. Plant Physiology 166, 2065–2076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Setya A, Murillo M, Leustek T.. 1996. Sulfate reduction in higher plants: molecular evidence for a novel 5'-adenylylsulfate reductase. Proceedings of the National Academy of Sciences, USA 93, 13383–13388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Slack CR, Hatch MD, Goodchild DJ.. 1969. Distribution of enzymes in mesophyll and parenchyma-sheath chloroplasts of maize leaves in relation to the C4-dicarboxylic acid pathway of photosynthesis. The Biochemical Journal 114, 489–498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sonderby IE, Geu-Flores F, Halkier BA.. 2010. Biosynthesis of glucosinolates—gene discovery and beyond. Trends in Plant Science 15, 283–290. [DOI] [PubMed] [Google Scholar]
- Stevenson CE, Hughes RK, McManus MT, Lawson DM, Kopriva S.. 2013. The X-ray crystal structure of APR-B, an atypical adenosine 5'-phosphosulfate reductase from Physcomitrella patens. FEBS Letters 587, 3626–3632. [DOI] [PubMed] [Google Scholar]
- Stewart V, Ronald PC.. 2022. Sulfotyrosine residues: interaction specificity determinants for extracellular protein-protein interactions. Journal of Biological Chemistry 298, 102232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sugiyama R, Li R, Kuwahara A, et al. 2021. Retrograde sulfur flow from glucosinolates to cysteine in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA 118, e2017890118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun SK, Xu X, Tang Z, Tang Z, Huang XY, Wirtz M, Hell R, Zhao FJ.. 2021. A molecular switch in sulfur metabolism to reduce arsenic and enrich selenium in rice grain. Nature Communications 12, 1392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi H, Kopriva S, Giordano M, Saito K, Hell R.. 2011. Sulfur assimilation in photosynthetic organisms: molecular functions and regulations of transporters and assimilatory enzymes. Annual Review of Plant Biology 62, 157–184. [DOI] [PubMed] [Google Scholar]
- Teles YCF, Souza MSR, Souza MFV.. 2018. Sulphated flavonoids: biosynthesis, structures, and biological activities. Molecules 23, 480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teneva I, Velikova V, Belkinova D, Moten D, Dzhambazov B.. 2023. Allelopathic potential of the cyanotoxins microcystin-LR and cylindrospermopsin on green algae. Plants 12, 1403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Traka M, Mithen R.. 2009. Glucosinolates, isothiocyanates and human health. Phytochemistry Reviews 8, 269–282. [Google Scholar]
- Varin L, DeLuca V, Ibrahim RK, Brisson N.. 1992. Molecular characterization of two plant flavonol sulfotransferases. Proceedings of the National Academy of Sciences, USA 89, 1286–1290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vernoux T, Wilson RC, Seeley KA, et al. 2000. The ROOT MERISTEMLESS1/CADMIUM SENSITIVE2 gene defines a glutathione-dependent pathway involved in initiation and maintenance of cell division during postembryonic root development. The Plant Cell 12, 97–110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wagner W, Follmann H, Schmidt A.. 1978. Multiple functions of thioredoxins. Zeitschrift fur Naturforschung. Section C, Biosciences 33, 517–520. [DOI] [PubMed] [Google Scholar]
- Watanabe M, Mochida K, Kato T, Tabata S, Yoshimoto N, Noji M, Saito K.. 2008. Comparative genomics and reverse genetics analysis reveal indispensable functions of the serine acetyltransferase gene family in Arabidopsis. The Plant Cell 20, 2484–2496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weber M, Suter M, Brunold C, Kopriva S.. 2000. Sulfate assimilation in higher plants—characterization of a stable intermediate in the adenosine 5'-phosphosulfate reductase reaction. European Journal of Biochemistry 267, 3647–3653. [DOI] [PubMed] [Google Scholar]
- Weckopp SC, Kopriva S.. 2014. Are changes in sulfate assimilation pathway needed for evolution of C4 photosynthesis? Frontiers in Plant Science 5, 773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wirtz M, Beard KFM, Lee CP, et al. 2012. Mitochondrial cysteine synthase complex regulates O-acetylserine biosynthesis in plants. Journal of Biological Chemistry 287, 27941–27947. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wirtz M, Droux M, Hell R.. 2004. O-acetylserine (thiol) lyase: an enigmatic enzyme of plant cysteine biosynthesis revisited in Arabidopsis thaliana. Journal of Experimental Botany 55, 1785–1798. [DOI] [PubMed] [Google Scholar]
- Xia Y, Lu C, Hou N, Xin Y, Liu J, Liu H, Xun L.. 2017. Sulfide production and oxidation by heterotrophic bacteria under aerobic conditions. The ISME Journal 11, 2754–2766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Y, Yu G, Chen Y, Jia N, Li R.. 2021. Four decades of progress in cylindrospermopsin research: the ins and outs of a potent cyanotoxin. Journal of Hazardous Materials 406, 124653. [DOI] [PubMed] [Google Scholar]






