Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Cyanobacteria are key organisms in the global ecosystem, useful models for studying metabolic and physiological processes conserved in photosynthetic organisms, and potential renewable platforms for production of chemicals. Characterizing cyanobacterial metabolism and physiology is key to understanding their role in the environment and unlocking their potential for biotechnology applications. Many aspects of cyanobacterial biology differ from heterotrophic bacteria. For example, most cyanobacteria incorporate a series of internal thylakoid membranes where both oxygenic photosynthesis and respiration occur, while CO2 fixation takes place in specialized compartments termed carboxysomes. In this review, we provide a comprehensive summary of our knowledge on cyanobacterial physiology and the pathways in Synechocystis sp. PCC 6803 (Synechocystis) involved in biosynthesis of sugar-based metabolites, amino acids, nucleotides, lipids, cofactors, vitamins, isoprenoids, pigments and cell wall components, in addition to the proteins involved in metabolite transport. While some pathways are conserved between model cyanobacteria, such as Synechocystis, and model heterotrophic bacteria like Escherichia coli, many enzymes and/or pathways involved in the biosynthesis of key metabolites in cyanobacteria have not been completely characterized. These include pathways required for biosynthesis of chorismate and membrane lipids, nucleotides, several amino acids, vitamins and cofactors, and isoprenoids such as plastoquinone, carotenoids, and tocopherols. Moreover, our understanding of photorespiration, lipopolysaccharide assembly and transport, and degradation of lipids, sucrose, most vitamins and amino acids, and haem, is incomplete. We discuss tools that may aid our understanding of cyanobacterial metabolism, notably CyanoSource, a barcoded library of targeted Synechocystis mutants, which will significantly accelerate characterization of individual proteins.

Free full text 


Logo of bsrLink to Publisher's site
Biosci Rep. 2020 Apr 30; 40(4): BSR20193325.
Published online 2020 Apr 3. https://doi.org/10.1042/BSR20193325
PMCID: PMC7133116
PMID: 32149336

Current knowledge and recent advances in understanding metabolism of the model cyanobacterium Synechocystis sp. PCC 6803

Associated Data

Supplementary Materials

Abstract

Cyanobacteria are key organisms in the global ecosystem, useful models for studying metabolic and physiological processes conserved in photosynthetic organisms, and potential renewable platforms for production of chemicals. Characterizing cyanobacterial metabolism and physiology is key to understanding their role in the environment and unlocking their potential for biotechnology applications. Many aspects of cyanobacterial biology differ from heterotrophic bacteria. For example, most cyanobacteria incorporate a series of internal thylakoid membranes where both oxygenic photosynthesis and respiration occur, while CO2 fixation takes place in specialized compartments termed carboxysomes. In this review, we provide a comprehensive summary of our knowledge on cyanobacterial physiology and the pathways in Synechocystis sp. PCC 6803 (Synechocystis) involved in biosynthesis of sugar-based metabolites, amino acids, nucleotides, lipids, cofactors, vitamins, isoprenoids, pigments and cell wall components, in addition to the proteins involved in metabolite transport. While some pathways are conserved between model cyanobacteria, such as Synechocystis, and model heterotrophic bacteria like Escherichia coli, many enzymes and/or pathways involved in the biosynthesis of key metabolites in cyanobacteria have not been completely characterized. These include pathways required for biosynthesis of chorismate and membrane lipids, nucleotides, several amino acids, vitamins and cofactors, and isoprenoids such as plastoquinone, carotenoids, and tocopherols. Moreover, our understanding of photorespiration, lipopolysaccharide assembly and transport, and degradation of lipids, sucrose, most vitamins and amino acids, and haem, is incomplete. We discuss tools that may aid our understanding of cyanobacterial metabolism, notably CyanoSource, a barcoded library of targeted Synechocystis mutants, which will significantly accelerate characterization of individual proteins.

Keywords: comparative genomics, cyanobacteria, degradation, metabolism, Synechocystis

1. Introduction

Cyanobacteria are the only prokaryotes capable of oxygenic photosynthesis. Since their appearance >2.4 billion years ago [1], cyanobacteria have profoundly impacted Earth’s climate and ecosystem, most notably in generation of an oxygenic atmosphere [2]. In the current ecosystem, cyanobacteria are a diverse phylum of photosynthetic prokaryotes that account for approximately a quarter of global carbon fixation [3] and a high proportion of marine nitrogen fixation [4,5]. Some species also show great potential as biotechnology platforms for synthesis of pharmaceuticals, industrial compounds and biofuels, due to their highly efficient conversion of water and CO2 to biomass using solar energy [6–8]. Others are used in the food, dye, cosmetics and nutraceutical industries with their global market projected to be worth >£1.5 billion by 2026 [9]. Certain species are also sources of natural products, including antifungal, antibacterial and anti-cancer compounds, and toxins deleterious to human and animal health [10,11]. Chloroplasts probably descend from an internalized cyanobacterium [12], thus certain physiological and biochemical features are conserved in higher photosynthetic organisms, making cyanobacteria excellent chassis for production of plant-derived natural products, like terpenes. Many key processes conserved throughout the photosynthetic lineages were first characterized in cyanobacteria [13,14], and there is significant interest in engineering cyanobacterial enzymes and CO2-concentrating mechanisms into crop plants [15–19].

Despite their importance, our understanding of many key features of cyanobacterial physiology and biochemistry is poor. For example, in Synechocystis sp. PCC 6803 (Synechocystis), the most widely studied cyanobacterium, less than 1200 coding sequences (~30%) have assigned function (469 in metabolism and 115 in transport: Highlighted in red in Supplementary Table S1; ~558 in other cellular processes (including transposons and transposon related functions): Highlighted in red in Supplementary Table S3), which is less than half compared with Escherichia coli [20]. Of these coding sequences, only a small proportion have been characterized in a cyanobacterium [21], with the majority of assigned functions based on studies of homologues in other bacteria, even though the function, catalytic activity and importance of characterized genes may differ significantly between phototrophic and heterotrophic bacteria. It is also likely that a proportion of these coding sequences have incorrectly assigned functions. Several examples of Synechocystis genes that were experimentally validated as having functions different to the original assigned function, based on homology with genes from heterotrophic bacteria, are discussed throughout the review.

In this review we will provide a detailed overview of the metabolic biochemistry and transport processes found in cyanobacteria, with a focus on the model unicellular species Synechocystis and to a lesser degree, Synechococcus elongatus PCC 7942 (Synechococcus). In each section we will highlight recent findings pertaining to each particular metabolic pathway, including central carbon and sugar metabolism, amino acid, nucleotide, cofactor and vitamin, lipid and membrane components, isoprenoid and pigment biosynthesis, and the transporters localized in the different membrane compartments. While many cyanobacteria are filamentous, with some incorporating heterocysts (specialized nitrogen fixing cells), describing the additional level of physiological complexity in these species is beyond the scope of this review (for an excellent recent review see [22]). Other aspects of cyanobacteria, such as photosynthesis and electron transport, have also been the subject of a recent review [23], and will not be discussed, except when electron transport chain components are involved in metabolism.

In the interests of brevity, the majority of enzymatic steps will not be mentioned in the text but outlined in subsequent figures. Steps to which an enzyme from Synechocystis has not been assigned are indicated by only an arrow with no abbreviated protein name in close proximity. The discussion will primarily focus on reactions that differ in cyanobacteria compared with model heterotrophs, or have been specifically investigated in model cyanobacteria. In most cases, only the abbreviated protein name is included in the text, although full names are outlined in Supplementary Table S1 (Column C). We have also incorporated four tables, to help guide future work on identifying homologues and assigning putative protein function. Supplementary Table S1 lists the Synechocystis proteins in each metabolic process, in the order outlined in the text. Also shown are the E. coli K12 proteins demonstrating the highest sequence similarity to individual Synechocystis proteins. Supplementary Table S2 is in the opposite format, and includes a list of E. coli K12 proteins with assigned functions, and the Synechocystis proteins with the highest homology to each E. coli protein. Supplementary Table S3 includes a list of Synechocystis proteins potentially involved in processes other than metabolism and transport, while Supplementary Table S4 includes all remaining Synechocystis proteins that have no assigned function. We will also highlight the aspects of cyanobacterial physiology and biochemistry that have yet to be elucidated and some tools in development, most notably CyanoSource, a mutant library and plasmid resource for Synechocystis, which will accelerate research efforts in this field.

2. The physiology of Synechocystis sp. PCC 6803

In order to understand cyanobacterial metabolism, it is first necessary to describe their physiology, which is more complex than most other prokaryotes. The majority of cyanobacterial species incorporate an array of internal thylakoid membranes (TM) enclosing the thylakoid lumen, in addition to a cell envelope consisting of the plasma membrane (PM), peptidoglycan layer and outer membrane (OM) [24] (Figure 1). In Synechocystis and some other cyanobacteria, the S-layer, a paracrystalline protein layer, surrounds the OM [25]. TMs may contain perforations allowing transport of molecules or proteins through the array [26]. Cytoplasmic localized compartments such as the carboxysome, the site of carbon fixation, and various storage bodies accumulating glycogen, cyanophycin, polyhydroxybutyrate, lipids and polyphosphate, are predominantly distributed in the central area of the cell [27,28].

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g1.jpg
Schematic detailing the ultrastructure of Synechocystis sp. PCC 6803 showing various subcellular components

Schematic adapted from [32,34].

Only the primordial cyanobacterial species, Gloeobacter kilaueensis JS1 and Gloeobacter violaceus PCC 7421, both of which are extremely slow growing, lack TMs [29,30]. Therefore, there must be clear advantages in incorporating a series of internal membranes. The most obvious is the increased area available to accommodate photosynthetic complexes, in addition to incorporating a compartment that can be optimized for specialized functions. In Synechocystis, it has been demonstrated that the majority of characterized TM localized proteins are involved in photosynthetic and respiratory energy generation, suggesting that this is the primary function of this compartment [31,32]. In turn, these advantages must outweigh potential burdens arising from the additional complexity imposed on the cell. These burdens include the requirement for specialized cellular systems to target proteins and metabolites to the correct compartment, organize and pack TMs within the cell, and to partition TMs between daughter cells during division.

In Synechococcus, TMs are arranged in orderly sheets parallel to the PM with areas of convergence between the two compartments at various points [33]. Whether the TM and PM are two separate compartments is yet to be confirmed. TM arrangement in Synechocystis is more complicated with individual sheets often displaying disparate patterns. Three-dimensional imaging demonstrates that the majority of TMs arrange in stacks of parallel sheets that converge in distinct structures near the PM [34]. However, in contrast with earlier reports, the thylakoid and plasma membranes were shown to be two separate compartments, although the distance between them was sometimes as little as 2 nm. This suggests that processes occurring in the two compartments are spatially separated. A dense material was observed between this junction that may play a role in ‘attachment’ of the thylakoids to the cell wall but the exact process and the proteins/compounds involved, has not been determined.

3. Central metabolism

In this review, cyanobacterial central metabolism will include glycolysis/gluconeogenesis, the tricarboxylic acid (TCA) cycle, the pentose phosphate (PP) pathway and the Calvin–Benson–Bassham (CBB) cycle, including carbon fixation, in addition to pathways for production of storage compounds, fermentation products and chorismate, a key intermediate for other pathways (Figure 2). Many enzymes involved in these pathways are conserved between Synechocystis and E. coli (Supplementary Table S1). Therefore, research related to protein function has primarily focused on the processes and enzymatic steps that differ in cyanobacteria compared with model heterotrophs.

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g2.jpg
Schematic detailing the pathways involved in central metabolism

Biosynthetic steps involved in glycolysis and gluconeogenesis are highlighted in red and blue respectively. Steps in the Entner–Doudoroff pathway are highlighted in green. Steps involved in the oxidative pentose phosphate pathway and the Calvin–Benson–Bassham cycle are highlighted in orange and purple, respectively. Fermentation pathways are highlighted in pink. Photorespiration pathways are highlighted in olive. Where enzymes catalyse reactions in two pathways, the arrows are split between their respective colours. The carboyxsome is represented as a purple octagon. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

3.1 Catabolism of glucose and glycogen

Carbon based inputs into central metabolism can be derived from carbon fixation, catabolism of glycogen or via import of glucose. The ability to import glucose enables some cyanobacteria, including certain Synechocystis substrains, to grow heterotrophically or mixotrophically [35]. Glucose is imported into the cell via the transporter, GlcP [36]. There are three proposed degradation pathways, which may be active under different environmental conditions [37]. Enzymes in the first two, glycolysis (the Embden–Meyerhof–Parnas (EMP) pathway) and the oxidative PP pathway, are generally highly conserved between Synechocystis and E. coli (Supplementary Table S1), and consequently these processes have not been extensively investigated in cyanobacteria. However, there are some differences and additional enzymes found in cyanobacteria. For example, homology between the Synechocystis and E. coli PdhA and PdhB subunits of pyruvate dehydrogenase is low (E value = 0.007 and 5.66E-04, respectively), and this complex has not been characterized in a cyanobacterium. Escherichia coli encodes only a class II fructose-1,6-bisphophosphate aldolase (Fbp2) for glycolysis, while Synechocystis also encodes a class I isoform (Fbp1). While the role of Fbp1 has not been determined in Synechocystis, expression of Fbp1 from the cyanobacterium Halothece sp. PCC 7418 in Synechococcus has been demonstrated to confer salt tolerance on this species [38]. The Synechocystis genome also encodes a protein, OpcA, which is not present in E. coli, and has been suggested to be key for glucose-6-phosphate dehydrogenase (Zwf) activity, the first step of the oxidative PP pathway [39]. However, glucose-6-phosphate dehydrogenase activity was similar to wild-type when OpcA was deleted in Synechocystis [40]. Recently, a third glycolytic pathway was identified in Synechocystis (the Entner–Doudoroff (ED) pathway) [37]. This pathway allows conversion of glucose to the oxidative PP intermediate 6-P-gluconate, which is then converted into glyceraldehyde-3-P. The ED pathway is required for optimal photoautotrophic growth and glycogen catabolism, and possibly also optimal activity of the CBB cycle [41].

3.2 Carbon fixation and the Calvin–Benson–Bassham cycle

As the enzymes of the CBB cycle are not isolated in a sub-cellular organelle as in eukaryotes (i.e. the chloroplast), some reactions are shared with EMP and OPP pathways. The CBB cycle can be divided into two stages: (1) conversion of ribulose-1,5-P and CO2 into two molecules of glycerate-3-P via ribulose-1,5-P carboxylase/oxygenase (RuBisCO), which is located in carboxysomes; (2) regeneration of the precursor, ribulose-1,5-P, consuming ATP and NADPH predominantly derived from photosynthesis. The requirement to regenerate ribulose-1,5-P leads to one major difference in the EMP pathway between cyanobacteria and heterotrophs. In E. coli, glyceraldehyde-3-P dehydrogenase (Gap) catalyses the reversible oxidative phosphorylation of glyceraldehyde-3-P to glycerate-1,3-P, resulting in interconversion between NAD+ to NADH. In contrast, Synechocystis Gap1 displays only glycolytic activity and a strict affinity for NAD+. A second isoform, Gap2, catalyses the reverse reaction required for the CBB cycle using NADH and potentially also NADPH, which is generated in large amounts via photosynthesis [42].

3.3 Photorespiration

RuBisCO can assimilate O2 instead of CO2, resulting in the production of one molecule each of glycerate-3-P and glycolate-2-P. The latter product is toxic to chloroplast metabolism in photosynthetic eukaryotes and likely also to Synechocystis at high concentrations [43]. Therefore, glycolate-2-P is converted into glycerate-3-P via the photorespiratory salvage pathway, a multi-step process conserved in most organisms that perform oxygenic photosynthesis [44]. Glycolate-2-P is first converted into glyoxylate by GlcD1 or GlcD2. Three subsequent photorespiratory pathways for catabolism of glyoxylate have been proposed in Synechocystis and deletion of genes in each pathway results in a mutant that requires high CO2 conditions for survival [43]. The first involves conversion of glyoxylate to glycerate-3-P via tartonic semialdehyde biosynthesis, the second, conversion of glyoxylate to glycerate-3-P via glycine and L-serine interconversion, and the third conversion of glyoxylate to oxalate, which is subsequently converted into formate. The enzymes involved in several of these pathways have been predominantly identified in Arabidopsis thaliana, with putative homologs present in cyanobacteria [45]. Of these, Shm, involved in the second pathway, and GlcD1, have been shown to display similar enzymatic activity to their A. thaliana homologs [45]. Deletion of GlcD1 and GlcD2 in Synechocystis results in a complete loss of photorespiratory activity [43]. However, the role of the other putative cyanobacterial homologs has not been determined and many proteins currently assigned to photorespiration, as outlined in Eisenhut et al. [43], have been suggested to catalyse alternative reactions. Moreover, in the third pathway, only one putative enzyme, Odc, has been identified.

3.4 Synthesis of carbon storage compounds

Cyanobacteria require carbon storage compounds for periods when photosynthesis is not sufficient for the cells energy and metabolic requirements. In Synechocystis, under conditions where cells are accumulating excess sugars, a high proportion of glycerate-3-P generated via CO2 fixation is converted into glycogen (reviewed in [46]). In E. coli, ADP-glucose is used as the substrate to generate the primary, unbranched polymer via GlgA. However, two GlgA isoforms are present in Synechocystis with likely roles in elongating the polymer at varying length [47]. Glycogen catabolism in Synechocystis is catalysed by two isoforms of GlgX (GlgX1 and GlgX2) and GlgP (GlgP1 and GlgP2). The role of GlgX1 and GlgX2 has not been determined. The GlgP proteins perform the same catalytic activity under different environmental conditions, cleavage of glycogen to individual glucose-1-P residues [48]. When Synechocystis is exposed to certain stress conditions, an additional carbon storage compound, the polymer polyhydroxybutyrate, is synthesized from acetyl-CoA via PhaA, PhaB, and the PhaC/PhaE complex [49–51].

3.5 The tricarboxylic acid cycle

The tricarboxylic acid (TCA) cycle differs in cyanobacteria compared with heterotrophic bacteria, as highlighted by recent work in the last decade. Cyanobacteria lack the enzyme α-ketoglutarate dehydrogenase, which catalyses the fourth step of the TCA pathway in E. coli: conversion of α-ketoglutarate to succinyl-CoA. Instead, some cyanobacteria, including Synechocystis, have genes encoding two enzymes, α-ketoglutarate decarboxylase (2-OGDC) and succinic semialdehyde dehydrogenase (SSADH), which convert α-ketoglutarate into succinic semialdehyde, then succinic semialdehyde into succinate, respectively [52]. Compared with the standard TCA cycle, where conversion of α-ketoglutarate to succinate results in production of one NADH and one GTP, the 2-OGDC/SSADH pathway results in production of one NADPH [52]. Only the soluble subunits of succinate dehydrogenase, catalysing the sixth step, have been identified in cyanobacteria [23]. Succinate dehydrogenase is integrated into the thylakoid membrane interlinked photosynthetic and respiratory electron chain [53]. Synechocystis also encodes a succinyl-CoA synthetase complex (SucC/SucD), which probably catalyses the reversible conversion of succinate into succinyl-CoA in cyanobacteria [54], required for biosynthesis of methionine and lysine. Several recent papers have investigated the enzymatic properties of TCA enzymes conserved between cyanobacteria and heterotrophic bacteria [55–57]. In contrast with many heterotrophic bacteria, Synechocystis citrate synthase (GltA) was shown only to catalyse generation of citrate, not its cleavage. Synechocystis GltA has a lower substrate affinity and turnover rate than the E. coli homologue, is not inhibited by ATP and NADH, but is inhibited by phosphoenolpyruvate [55].

3.6 Alternate biosynthetic pathways linking metabolites of the tricarboxylic acid cycle, photorespiration and glycolysis

A range of additional pathways link the TCA cycle with glycolysis and photorespiration. Glyoxylate, produced via photorespiration, also plays a role in the glyoxylate cycle. This cycle consists of three TCA enzymes and two additional enzymes unique to this pathway: the first, isocitrate lyase (Icl), converts the TCA cycle intermediate isocitrate to succinate and glyoxylate; the second, malate synthase (Msy), converts glyoxylate and acetyl-CoA to the TCA cycle intermediate, malate. While activity of glyoxylate cycle enzymes has been detected in some cyanobacteria (reviewed in [58]), it is unclear whether Synechocystis encodes active variants of Icl and Msy.

Phosphoenolpyruvate carboxylase (PepC) catalyses the conversion of phosphoenolpyruvate, a glycolysis intermediate, and HCO3 to oxaloacetate, a TCA intermediate [59]. PepC can therefore be considered an inorganic carbon fixing enzyme (i.e. akin to RuBisCO). Metabolic flux analysis has shown that as much as 25% of all inorganic carbon fixation occurs via PepC in Synechocystis cultured under mixotrophic or heterotrophic conditions [60]. An additional protein, malic enzyme (ME), catalyses the reversible conversion of malate, a TCA intermediate, and pyruvate [61]. Deletion of ME in Synechocystis results in a mutant that displays poor growth when exposed to continuous but not diurnal light [62]. It was hypothesized that ME is required for pyruvate biosynthesis under continuous light.

3.7 Fermentation pathways

Three possible fermentation pathways are present in Synechocystis that generate D-lactate, acetate or succinate, respectively. Presumably fermentation plays a role in energy generation when cyanobacteria are exposed to long periods of darkness under anoxic conditions, but the importance of these pathways during changing environmental conditions has not been determined. D-lactate, acetate and succinate production has been observed in wild-type Synechocystis cells but only after 3 days growth under dark, anaerobic conditions [63]. A homolog of lactate dehydrogenase (Ddh), which converts pyruvate and NADH to lactate and NAD+, is encoded by Synechocystis. Two possible pathways for acetate production may be present in Synechocystis: (1) Conversion of acetyl-CoA to acetyl-P, then acetate, via phosphotransacetylase (Pta) and acetate kinase (Ack), respectively; (2) Direct reversible conversion of acetyl-CoA to acetate via acetyl-CoA synthetase (Acs) [63]. Production of succinate relies primarily on phosphoenolpyruvate as the initial substrate, which is subsequently converted to oxaloacetate via PepC and then fed into the reverse TCA cycle [64].

3.8 Chorismate biosynthesis

Chorismate is the precursor for biosynthesis of a range of amino acids and cofactors, and has further importance in cyanobacteria as the substrate for production of phylloquinone, plastoquinone, phenylalanine, tyrosine, folate and molybdopterin, in addition to tocopherols and carotenoids. The glycolytic and PP pathway intermediates phosphenolpyruvate and erythrose-4-P are the substrates for production of chorismate via a 7-step pathway in E. coli. However, the enzyme catalysing the first step, condensation of phosphoenolpyruvate and erythrose-4-P, has not been identified in Synechocystis [65]. Synechocystis proteins demonstrating high sequence similarity to five other enzymes in the E. coli pathway have been identified (Supplementary Table S1) with the exception of the third enzyme, AroQ (No BLAST match). It is unclear from the literature how function was assigned to Synechocystis AroQ, encoded by sll1112 in the KEGG database.

4. Metabolism and degradation of nucleotide sugars and sugar osmolytes

A range of nucleotide sugars required for lipopolysaccharide (LPS) biosynthesis or as cofactors for other reactions (i.e. UDP-glucose) are synthesized by Synechocystis (Figure 3). LPSs contain a range of sugar residues including rhamnose, galactose, glucosamine, mannose and fucose, which in Synechocystis are incorporated as 2,3-di-methyl-fucose and 2-methyl-fucose. 2-Methylxylose has also been reported in Synechocystis [66]. Only some of the biosynthetic pathways synthesizing the LPS sugar precursors have been identified in cyanobacteria, although predominantly on the basis of identifying proteins with high sequence similarity to characterized enzymes from heterotrophic bacteria. TDP-β-L-rhamnose is synthesized by a four step pathway from glucose-1-P. There are two potential homologs in Synechocystis for each of the last three enzymes in the pathway, RfbB, RfbC and RfbD, but the function of these isoenzymes has not been determined. UDP-N-acetylglucosamine is synthesized by a three step pathway from fructose-6-P and is the precursor not just for LPSs but also peptidoglycan. UDP-glucose is synthesized from glucose-1-P by CugP, a non-GalU UDP-glucose pyrophosphorylase, which differs from the GalU UDP-glucose pyrophosphorylase reaction conducted in most proteobacteria, including E. coli [67]. A UDP-glucose 4-epimerase (GalE) then catalyses the conversion of UDP-glucose to UDP-galactose. GDP-mannose is synthesized from fructose-6-P by a three step reaction and GDP-fucose from GDP-mannose by a two-step pathway. None of the proteins in these pathways have been characterized in cyanobacteria although deletion of the last gene in this pathway in Synechocystis, WcaG, resulted in production of carotenoids lacking fucose [68].

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g3.jpg
Metabolism and degradation of nucleotide sugars and sugar osmolytes

Compounds highlighted in blue are substrates for lipopolysaccharide biosynthesis. Steps highlighted in grey are compounds and reactions not involved in these pathways but detailed in Figure 1. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

Several sugars act as osmolytes, notably sucrose and glucosylglycerol. Osmolytes play a role in Synechocystis in salt tolerance [69,70]. In Synechocystis, sucrose is synthesized from UDP-glucose (or ADP-glucose) and fructose-6-P by two enzymes: SpsA and Spp [71,72]. Sucrose breakdown in Synechocystis is catalysed by an invertase (Inv) [73], resulting in production of glucose and fructose, which are likely phosphorylated to glucose-6-P by Glk and fructose-6-P by FrkA, and cycled back into glycolysis. A putative glucose kinase and fructose kinase are encoded in the Synechocystis genome, but have not been characterized. Glucosylglycerol is synthesized from ADP-glucose and glycerol-3-P via two enzymes: GgpS and GgpP [74]. Glycerol-3-P is derived from either the TCA cycle intermediate glycerine-3-P or possibly imported.

5. Amino acid biosynthesis and degradation

Synechocystis synthesizes 20 L-amino acids and two D-amino acids (Figure 4). The majority of enzymes involved in amino acid biosynthesis display high sequence similarity between Synechocystis and E. coli (Supplementary Table S1). Amino acids are synthesized from a range of substrates, including pyruvate, the TCA cycle intermediates α-ketoglutarate and oxaloacetate, chorismate, the nucleotide intermediate, 5-phosphoribosyl-1-pyrophosphate (discussed in section 6), and glycerate-3-P or glyoxylate. Biosynthesis of amino acids is divided into sections below based on the substrates utilized.

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g4.jpg
Metabolism of amino acids, cyanophycin, glutathione and iron–sulfur clusters

The 20 L-amino acids are highlighted in red while amino acids incorporated into peptidoglycan are highlighted in blue. The iron–sulfur biosynthetic pathways is highlighted in green. Steps highlighted in grey are compounds and reactions not involved in these pathways but detailed in Figure 1. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

5.1 Isoleucine, valine and leucine biosynthesis

α-Ketobutyrate (synthesized from L-threonine by IlvA) and pyruvate are the substrates for biosynthesis of L-isoleucine, while pyruvate is the sole substrate for L-valine and L-leucine biosynthesis. The enzymatic steps in Synechocystis are similar to those in E. coli, with the exception of the first step. In E. coli biosynthesis of α-acetolactate and α-aceto-β-hydroxybutyrate are typically catalysed by the IlvB/IlvN complex. However, in Synechocystis, the homologue for IlvB was identified as 2-OGDC in the TCA cycle (Section 3.5) [52]. An alternate acetolactate synthase, IlvG, demonstrates high sequence similarity to E. coli IlvG (E value = 0). IlvG may form a complex with IlvN and catalyse this step [75] but this requires further verification.

5.2 Glutamate, glutamine and proline biosynthesis

The TCA cycle intermediate α-ketoglutarate is the substrate for L-glutamate biosynthesis which in turn is the substrate for production of L-glutamine, D-glutamate and L-proline. D-glutamate is synthesized by MurI and is incorporated into peptidoglycan. Two different glutamine synthetases, GlnA and GlnN, convert L-glutamate into L-glutamine [76], and in the process incorporate ammonia into amino acid biosynthesis. Alternatively, several enzymes catalyse the opposite reaction where L-glutamine is converted into L-glutamate, including an NAD(P)H or possibly ferredoxin-dependent glutamate synthase (GltB/GltD) and a ferredoxin-dependent glutamate synthase (GlsF) [77]. L-proline is synthesized via three enzymes (ProA, ProB, ProC). Synechocystis also encodes a putative proline oxidase, PutA, which catabolized L-proline to L-glutamate, reducing NADP+ and possibly a quinone in the process [78].

5.3 Arginine biosynthesis

L-arginine is synthesized from L-glutamate via eight enzymatic steps, the sixth requiring carbomyl-P, which is synthesized from L-glutamine via CarA/CarB. This pathway is very similar to that in E. coli. However, Synechocystis does not encode ArgA or ArgE, catalysing the first and fifth steps of the pathway. Instead, it encodes ArgJ, a bifunctional enzyme that catalyses both these enzymatic reactions. Recently, an ornithine-ammonia cycle was identified in Synechocystis [79]. This cycle utilizes ArgF, ArgG, ArgH, and an additional enzyme, AgrE. AgrE converts L-arginine into L-ornithine, releasing ammonia in the process [80]. Synechocystis also encodes two putative SpeA and two putative SpeB proteins, which play a role in degradation of L-arginine to putrescine, a polyamine. In E. coli, putrescine can be used as a nitrogen and carbon source via conversion to succinate [80]. Whether putrescine has a similar role in cyanobacteria has not been determined.

5.4 Aspartate, cyanophycin and lysine biosynthesis

L-aspartate is synthesized from oxaloacetate and L-glutamate by AspC. L-aspartate and L-arginine are the substrates for cyanophycin, a nitrogen storage polymer. Cyanophycin is synthesized by CphA and then converted back to L-aspartate and L-arginine by CphB and LadC [81]. L-aspartate is converted into aspartate-4-semialdehyde, which is the substrate for biosynthesis of L-threonine and L-lysine. Synechocystis encodes all the enzymes in the five step diaminopimelate aminotransferase pathway required for L-lysine biosynthesis [82,83]. The third reaction, conversion of tetrahydrodipicolinate to L,L-diaminopimelate, is catalysed by DapL. In contrast, E. coli requires three enzymes, DapC, DapD and DapE, for this conversion. L-lysine is the substrate for production of the siderophore cadaverine by Cad. Three enzymes, ThrA, ThrB and ThrC, convert aspartate-4-semialdehyde into L-threonine by a pathway similar to that in E. coli.

5.5 Methionine biosynthesis

In E. coli, L-methionine is also synthesized from aspartate-4-semialdehyde. However, the Synechocystis genome does not encode homologues to MetA, MetB or MetC (Supplementary Table S2), the first three enzymes in the pathway. However, the genome does encode a putative MetH enzyme, which catalyses the last step, conversion of homocysteine to L-methionine. The enzymatic steps prior to this have not been determined, nor has the original substrate from which L-methionine is synthesized. The Synechocystis genome also encodes a putative MetK enzyme, which converts L-methionine into S-adenosyl-L-methionine, a cofactor utilized in many other reactions, most notably in biosynthesis of cyanocobalamin (Vitamin B12; Section 10.4). A putative AhcY enzyme is also encoded, which converts S-adenosyl-L-homocysteine, the product of reactions which use S-adenosyl-L-methionine as a cofactor, back to homocysteine.

5.6 Tryptophan, phenylalanine and tyrosine biosynthesis

Chorismate is the substrate for L-tryptophan, L-phenylalanine and L-tyrosine biosynthesis. The majority of enzymes involved in L-tryptophan biosynthesis are highly conserved between E. coli and Synechocystis. Attempts to generate an auxotrophic mutant of TrpB, one of the subunits catalysing the final step of L-tryptophan biosynthesesis, were unsuccessful [84], suggesting that it cannot be imported from the external environment. The pathway for L-phenylalanine and L-tyrosine biosynthesis differs between the two species and has not been completely determined in cyanobacteria. Both amino acids are synthesized from prephenate. However, only the second step of tyrosine biosynthesis, conversion of arogenate to L-tyrosine, has been determined, although sll1662 (PheA) has been speculated to catalyse the first step of L-phenylalanine biosynthesis, conversion of prephenate to prenylpyruvate [85].

5.7 Histidine biosynthesis

L-histidine, synthesized from the nucleotide precursor, 5-phosphoribosyl-1-pyrophosphate, is synthesized via a nine-step pathway in E. coli. Proteins demonstrating high sequence similarity to all characterized histidine biosynthetic enzymes in E. coli have been identified in Synechocystis. However, there are two putative HisC and HisD enzymes in Synechocystis. The function of these isoenzymes has not been determined.

5.8 Serine, glycine, cysteine and alanine biosynthesis

L-serine can potentially be synthesized via two routes. The first is via a three step light-independent pathway, which has been characterized in Synechocystis [86]. However, the second enzyme in this pathway, SerC has also been suggested to catalyse the transanimation reaction in photorespiration (Section 3.3) [43]. In the second pathway, L-serine (and also glycine) is synthesized from glyoxylate via the photorespiratory pathway or glyoxylate cycle in those species that encode the relevant enzymes. L-cysteine is then produced from L-serine via a two step pathway, the second of which could potentially be catalysed by either CysK or CysM. L-cysteine is subsequently desulfonated to produce L-alanine by Csd [87], which is subsequently converted to D-alanine, a component of peptidoglycan, via Alr.

5.9 Glutathione biosynthesis

L-cysteine and L-glutamate are the substrates for the first step of glutathione biosynthesis. Glutathione is a thiol that plays a key role in metal detoxification and tolerance of oxidative stress in Synechocystis [88]. The first step of glutathione biosynthesis is catalysed by GshA, encoded by an essential gene in Synechocystis [89]. In contrast, the enzyme catalysing the second step, GshB is non-essential, suggesting that glutathione is not required for Synechocystis viability but that the precursor, L-γ-glutamyl-L-cysteine, is [89].

5.10 Iron–sulfur cluster biosynthesis

Conversion of L-cysteine into L-alanine by Csd releases sulfur that is incorporated into iron-sulfur clusters. Two additional cysteine desulfarases have been identified in Synechocystis but unlike Csd, neither are essential [90–92]. Iron–sulfur clusters are incorporated into many proteins involved in photosynthesis, respiration and nitrogen fixation [93]. Figure 4 outlines iron–sulfur biosynthesis (highlighted in green) and subsequent transfer to proteins, based on characterization of proteins in other bacterial species [94]. SufE acts as a sulfur donor, and IscA as a Fe2+ donor to the scaffold proteins required for cluster formation (SufA/NifU) [95]. Additional subunits (SufB/SufC/SufD) aid in transfer of the iron–sulfur cluster to proteins. NifU is possibly involved in repairing iron–sulfur clusters in proteins but has not been characterized in cyanobacteria.

6. Nucleotide biosynthesis

Enzymes involved in nucleotide biosynthesis (Figure 5) are highly conserved between E. coli and Synechocystis (Supplementary Table S1), and therefore this pathway has not been investigated in great detail in cyanobacteria. Pyrimidines and purines require the same precursor, 5-phosophoribosyl-1-pyrophosphate, which is synthesized from the PP pathway intermediate, ribose-5P, after which the pathways diverge.

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g5.jpg
Metabolism of nucleotides

The purine and pyrimidine biosynthesis pathways are highlighted in red and blue respectively. Possible nucleotide salvage pathways are highlighted in green. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

6.1 Purine biosynthesis

In E. coli, purine biosynthesis requires eleven enzymatic steps for production of inosine monophosphate, the precursor of guanosine and adenosine based nucleotides (reviewed in [96]). Synechocystis encodes genes with high homology to all the purine biosynthetic enzymes required for inosine monophosphate in E. coli, including PurN and PurT, which are both capable of catalysing the third step (Supplementary Table S1). Both PurB and PurH catalyse two different steps in the pathway. In E. coli, inosine monophosphate is converted into guanosine diphosphate by GuaB, GuaA and GmpK, and adenosine diphosphate by PurA, PurB and AmpK [97]. All nucleoside-diphosphates are converted into nucleoside-triphosphates via NdkR [98] and to deoxyribonucleotides via the NrdA/NrdF complex [99]. All these enzymes are highly conserved between E. coli and Synechocystis (Supplementary Table S1).

6.2 Pyrimidine biosynthesis

In E. coli, pyrimidine biosynthesis requires six enzymatic steps for production of uridine diphosphate, the precursor of cytosine-, uridine- and thymidine-based nucleotides. Carbomyl-P, synthesized from glutamine and bicarbonate by CarA/CarB, is the initial substrate. Carbomyl-P is converted into orotate via a three-step pathway. Orotate phosophoribosyltransferase (PyrE) transfers a ribosyl group from 5-phosophoribosyl-1-pyrophosphate to orotate, forming oritidine-5-P, which is subsequently converted into uridine diphosphate via PyrF and PyrH. In E. coli, uridine diphosphate is converted into uridine triphosphate via NdkR, then cytosine triphosphate via PyrG [100]. The NrdA/NrdF complex then converts cytosine triphosphate into deoxycytosine triphosphate. The pathway for biosynthesis of deoxythymidine nucleotides has not been determined. However, enzymes homologous to those identified in the Lactococcus lactis pathway are conserved in Synechocystis [101]. Via this pathway, deoxycytosine triphosphate is converted into deoxyuridine monophosphate via Dcd, which is subsequently converted into deoxythymidine monophosphate via ThyX, which in turn is converted into deoxythymidine diphosphate via Tmk. However, experimental evidence is required to confirm whether this pathway is utilized by Synechocystis.

6.3 Nucleotide salvage pathways

Synechocystis also encodes a number of enzymes that display high sequence similarity to E. coli proteins involved in the nucleotide salvage pathway [100]. However, the role of the salvage pathway in cyanobacteria and how nucleotides are catabolized has not been investigated.

7. Cofactor biosynthesis

Unlike many cyanobacterial species, Synechocystis does not require the addition of any vitamins or cofactors for growth, suggesting that it encodes complete biosynthetic pathways for each essential compound. However, these pathways have not been extensively investigated. The majority of proteins in these pathways (Figure 6) have been assigned a function in cyanobacteria based on their homology to characterized enzymes from E. coli, with only a few enzymes characterized in Synechocystis or other model cyanobacterial species. Tocopherol biosynthesis is discussed in Section 9.4, since this cofactor is synthesized from the same initial substrates as other isoprenoids. Pseudocobalamin (Vitamin B12) biosynthesis is discussed in Section 10.4, since this cofactor is synthesized from the same initial substrates as bilins and chlorophyll.

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g6.jpg
Metabolism of vitamins and cofactors

Detailed are the pathways for biosynthesis of (A) Biotin, (B) NAD+ and NADP+, (C) folate, (D) molybdenum cofactors, (E) riboflavin and FAD, (F) thiamine, (G) pantothenate and coenzyme (A and H) pyridoxal-5P. Vitamins and cofactors are highlighted in blue. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

7.1 Biotin biosynthesis

In Synechocystis, biotin (vitamin B7) is an essential cofactor required by acetyl-CoA carboxylase (AccA/AccB/AccC/AccD; Section 8.1), which is involved in fatty acid biosynthesis [102]. The biotin biosynthetic pathway has been determined in E. coli [103]. In E. coli, biotin is synthesized from malonyl-ACP-methyl ester, which undergoes two cycles of fatty elongation to form pimeloyl-ACP-methyl ester. This is subsequently converted to biotin via five enzymatic steps. Synthesis of the pimeloyl-ACP precursor has not been determined in Synechocystis [104]. Putative homologues of only three enzymes in the biotin biosynthetic pathway, BioF, BioD and BioB (and not BioH, BioC and BioA) are encoded in the Synechocystis genome (Figure 6A) [103]. Recently, a novel enzyme, BioU, was demonstrated to catalyse the same reaction as BioA, conversion of 8-amino-7-oxononoate to 7,8-diaminononanoate [105]. The enzymatic activity of BioU is different from BioA. BioU utilizes then reforms NADPH, consumes CO2, and acts as a suicide enzyme, meaning it catalyses only a single reaction due to loss of a lysine group. Synechocystis also encodes a putative BirA protein, which reacts with biotin to form a biotin–BirA complex that represses biotin biosynthesis [104].

7.2 NAD+ and NADP+ biosynthesis

Nicotinamide adenine dinucleotide (NAD+) is synthesized in cyanobacteria from L-aspartate by a five-step pathway encoded by most bacterial species (Figure 6B) [106]. The last two enzymes in the pathway, NadD and NadE, have low sequence similarity to the equivalent E. coli proteins but the activity of the enzymes has been confirmed in Synechocystis [107]. A second two-step pathway for NAD+ biosynthesis from nicotinamide has also been proposed [107,108], although how nicotinamide is produced has not been determined. NAD+ is converted into NADP+, required as an electron acceptor in linear photosynthetic electron transport, by NAD kinases, of which two are present in Synechocystis (NadK1, NadK2) [109]. The NAD+/NADP+ ratio is regulated by pyridine nucleotide transhydrogenase (PntA/PntB), which catalyses electron transfer between the two compounds [110].

7.3 Folate biosynthesis

Folate (vitamin B9) based cofactors (e.g. tetrahydrofolate, 5-methyl tetrahydrofolate, 5,10-methylene tetrahydrofolate) are required in certain enzymatic reactions for biosynthesis of the amino acids L-methionine, L-serine and glycine (Figure 4), the cofactors pantothenate and coenzyme A (Figure 6G), purine nucleotides and thymidylate pyrimidines (Figure 5) and certain tRNAs [111]. Folate is synthesized from the precursors, chorismate and guanosine triphosphate (Figure 6C). A two-step pathway (PabB/PabC) results in conversion of chorismate to 4-aminobenzoate. A four-step pathway (FolE/FolB/FolK and possible FolQ) catalyses the conversion of guanosine triphosphate to 6-hydroxymethyl-7,8-dihydropteroate-PP, which together with 4-aminobenzoate, catalyses the formation of 7,8-dihydropteroate. FolQ (Designated as NudB in E. coli) [112] has not been characterized in Synechocystis but slr0920 shows low sequence similarity to NudB (e value = 4.56E-06) and may perform FolQ enzymatic activity (Supplementary Table S2). 7,8-Dihydropteroate is subsequently converted into the different folate variants, although only one enzyme catalysing these steps, FolC, has been identified. Whether 5-methyl tetrahydrofolate is synthesized by Synechocystis is unknown, since the genome does not encode MetF, which synthesizes this compound from 5,10-methylene tetrahydrofolate in E. coli [111].

7.4 Molybdenum cofactor biosynthesis

Molybdenum cofactors (molybdopterin guanine dinucleotide or molybdopterin-Mo) act as catalytic centres in a range of enzymes. In Synechococcus, a molybdenum cofactor is required for nitrate reductase (NarB; Section 11.1) activity [113]. If any other enzymes in cyanobacteria also require molybdenum cofactors has not been determined. Molybdenum cofactors are synthesized from guanosine triphosphate (Figure 6D). This pathway has been characterized in E. coli and proteins demonstrating high sequence similarity to each enzyme have been identified in Synechocystis [113]. Moreover, several enzymes in the pathway have been characterized in Synechococcus [113,114]. MoaC is likely a bifunctional enzyme catalysing the second step, formation of pyranopterin, and the fifth step, synthesis of the cofactor molybdopterin guanine dinucleotide. The third step, conversion of cyclic pyranopterin to molybdopterin is catalysed by MPT synthase (MoaD/MoaE), which is regenerated by MoeB [115].

7.5 Riboflavin and flavin adenine dinucleotide biosynthesis

Riboflavin (vitamin B2) and flavin adenine dinucleotide (FAD) are also synthesized from guanosine triphosphate (Figure 6E). In cyanobacteria, FAD is a cofactor involved in flavoprotein-mediated redox reactions. The pathway is similar between E. coli and Synechocystis and enzymes are highly conserved between the species (Supplementary Table S1). Three enzymes, RibA, RibD and RibF, catalyse two separate reactions in the pathway.

7.6 Thiamine biosynthesis

Thiamine diphosphate (vitamin B1) is a cofactor for several enzymes, including pyruvate dehydrogenase (Section 3.1), transketolase in the OPP/CBB pathways (TktA, Section 3.2), and acetolactate synthase, catalysing the first step of L-valine, L-leucine and L-isoleucine biosynthesis (IlvG/IlvN; Section 5.1) [116]. It is synthesized from the purine biosynthetic intermediate, 5-aminoimidazole ribonucleotide (Section 6.1; Figure 5), glycine and 1-deoxy-D-xylulose-5-P (Figure 6F). The pathway has been largely characterized in E. coli [117], but in Synechocystis, homologues have not been identified for every protein in the pathway (Supplementary Table S1). Notably, there is no protein in Synechocystis with high sequence similarity to ThiD (Supplementary Table S2), which catalyses the second biosynthetic step starting at 5-aminoimidazole ribonucleotide.

7.7 Pantothenate and coenzyme A biosynthesis

The majority of enzymes involved in biosynthesis of pantothenate (vitamin B5; Figure 6G) and coenzyme A are highly conserved between E. coli and Synechocystis (Supplementary Table S1). Coenzyme A is required for formation of acetyl-CoA and in fatty acid biosynthesis. Three enzymes convert α-ketoisovalerate, an intermediate required for L-valine and L-leucine biosynthesis (Section 5.1; Figure 4), to pantothenate. An additional enzyme, PanD, catalyses the third step, conversion of L-aspartate to β-alanine [118]. The second reaction can be catalysed by PanE, not encoded in the Synechocystis genome (Supplementary Table S2) or IlvC, which is also involved in L-isoleucine, L-valine and L-leucine biosynthesis (Figure 4). Coenzyme A is synthesized from pantothenate via five enzymatic steps [118]. Only the first step (conversion of pantophenate to 4-phosphopantophenate) is catalysed by a different enzyme from that in the E. coli pathway, specifically a type III pantophenate kinase (CoaX) [119].

7.8 Pyridoxal-5P biosyntheis

Pyridoxal-5-P (vitamin B6) is a cofactor required by a range of enzymes involved in amino acid biosynthesis and catabolism, iron, cell wall component and carbon metabolism, and biosynthesis of other cofactors (For a full list refer to [120]). Biosynthesis of pyridoxal-5-P in E. coli utilizes 1-deoxy-D-xylulose-5-P and 3-amino-2-oxopropyl phosphate as substrates, and is catalysed via PdxA/PdxJ, then PdxH (Figure 6H) [121]. PdxA, PdxJ and PdxH are conserved in Synechocystis but the three enzyme pathway for 3-amino-2-oxopropyl phosphate biosynthesis has not been determined.

8. Membrane and cell wall biosynthesis

Cyanobacterial membrane composition differs from that of heterotrophic bacteria. Five classes of lipids accumulate in Synechocystis plasma and thylakoid membranes: phosphatidylglycerol, monogalactosyl-diacylglycerol, digalactosyl-diacylglycerol, sulfoquinovosyl-diacylglycerol and hydrocarbons [122,123]. Like other Gram-negative prokaryotes, cyanobacteria are encompassed by a peptidoglycan layer and an OM containing lipopolysaccharides (LPSs).

8.1 Lipid biosynthesis

Cyanobacterial lipids are synthesized from acyl-ACPs (acyl carrier proteins), which in turn are synthesized from acetyl-CoA by a pathway similar to that in E. coli (Supplementary Table S1; Figure 7). Predominantly C16 and C18 acyl-ACPs are synthesized with various degrees of saturation catalysed by four possible desaturases (DesA, DesB, DesC, DesD) [124]. A PM associated protein, Aas [32], mediates import of acyl-ACPs and fatty acids from the PM and periplasm [125,126].

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g7.jpg
Metabolism of membrane lipids, peptidoglycan and lipopolysaccharides

Membrane lipids are highlighted in blue. Steps highlighted in grey are compounds and reactions not involved in these pathways but detailed in Figure 2 and Figure 3. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

Hydrocarbons are synthesized directly from acyl-ACPs [127,128], with the majority of cyanobacteria (including Synechocystis) producing C15 or C17 alkanes via a two-step pathway (Aar/Ado) [129], while the remainder produce C17 or C19 alkenes via a polyketide synthase (Ols) [130]. The other lipids are synthesized from 1,2-diacyl-glycerol-3-P, which is produced from acyl-ACPs via three enzymes (PlsX, PlsY, PlsC) [131]. A further three enzymatic steps are required for phosphatidylglycerol biosynthesis. The enzyme catalysing the second step, PgsA, is non-essential in Synechocystis, when the mutant is supplemented with phosphatidylglycerol [132]. There is no Synechocystis protein with any sequence similarity to PgpB, the enzyme in E. coli that catalyses the third step (Supplementary Table S2).

1,2-Diacyl-glycerol-3-P is likely converted into diacylglycerol, the common substrate for synthesis of the other membrane lipids. The enzyme catalysing this step has not been identified. The reverse reaction is likely catalysed by DgkA. MgdA catalyses conversion of diacylglycerol to monoglucosyl-diacylglycerol, which is likely converted into monogalactosyl-diacylglycerol by an unidentified epimerase [133]. Monogalactosyl-diacylglycerol is then converted into digalactosyl-diacylglycerol by DgdA [134]. Sulfoquinovose, synthesized from UDP-glucose and sulfate by SqdB [135,136], is reacted with diacylgycerol by SqdX to form sulfoquinovosyl-diacylglycerol [137].

The Synechocystis genome encodes no proteins with homology to enzymes involved in β-oxidation (Supplementary Table S2), although one report has suggested that the capacity for fatty acid catabolism is retained [138]. If so, there must be an alternate, uncharacterized pathway responsible for lipid degradation.

8.2 Lipoic acid biosynthesis

Lipoic acids are cofactors required for a range of enzymes, including pyruvate dehydrogenase and the glycine cleavage system (GcvH/GcvP/GcvT/GcvL; Figure 2) [139]. The biosynthestic pathway has been elucidated in E. coli [140]. Lipoic acids are covalently attached to enzymes via LipB and then sulfonated via LipA. In contrast to E. coli, there are two putative LipA proteins in Synechocystis (Supplementary Table S1).

8.3 Peptidoglycan biosynthesis and depolymerization

The structure of Synechocystis peptidoglycan has not been determined. However, peptidoglycan in the closely related species, Synechocystis sp. PCC 6714, incorporates L-alanine, D-alanine, D-glutamate and meso-diaminopimelate into peptide bridges, which are linked to polymers consisting of alternating acetylglucosamine and acetylmuramate monomers. The enzymes synthesizing peptidoglycan monomers (acetylglucosamine-N-acetylmuramate-pentapeptides) from UDP-N-acetylglucosamine are highly conserved between E. coli and Synechocystis (Supplementary Table S1). Surprisingly, the last two enzymes in the pathway, MraY and MurG have been localized to the TM in Synechocystis [31,32], suggesting that an additional protein or process must transport these monomers to the PM. The flippase involved in translocating the acetylglucosamine-N-acetylmuramate-pentapeptide monomers to the periplasmic side of the PM in E. coli (MurJ) has not been identified in cyanobacteria [141]. However, the protein encoded by slr0488 in Synechocystis demonstrates some sequence similarity to MurJ (E value = 1.06E-28; Supplementary Table S1) but its function needs to be confirmed experimentally.

Polymerization of peptidoglycan is catalysed by the penicillin-binding proteins (PBPs) 1–4 and FtsW [142], while depolymerization and recycling of peptidoglycan monomers is catalysed by PBPs 5-8 and AmiA-C [143]. Four proteins in E. coli have been implicated in importing depolymerized peptidoglycan components (NagE, MurP, AmpG, Opp) [144], but only Opp, an oligopeptide transporter consisting of four subunits, is encoded in the Synechocystis genome (Supplementary Table S2). A series of cytosolic enzymes conserved in Synechocystis (Mpl, NagZ, AnmK, NagK, MurQ) likely recycle depolymerized peptidoglycan components back into peptidoglycan biosynthesis [144]. Other E. coli enzymes involved in recycling (NagA, NagB, AmiD, AmpB) have no homologues in Synechocystis (Supplementary Table S2).

8.4 Lipopolysaccharide biosynthesis

LPSs are incorporated into the OM of cyanobacteria, including Synechocystis [66]. Four enzyme synthesize the Lipid A disaccharide core of the LPS and are highly conserved between E. coli and Synechocystis (Supplementary Table S1). The protein involved in translocating Lipid A disaccharide to the periplasmic side of the PM has not been identified, although four PM localized proteins with high sequence similarity to MsbA (slr2019: E value = 8.64E-91; sll1276: E value = 2.28E-84; sll1725: E value = 7.22E-83; slr1149: E value = 1.82E-73; Supplementary Table S2), the characterized Lipid A disaccharide flippase from E. coli [145], are encoded in the Synechocystis genome [32]. Biosynthesis of the polysaccharide portion of the LPS has not been determined in cyanobacteria [146]. Five PM-localized glycosyltransferases are encoded by the Synechocystis genome which may play a role in saccharide polymerization (Supplementary Table S1). However, the Synechocystis genome encodes no proteins with homology to those in E. coli involved in transporting polysaccharides across the PM (i.e. Wzm/Wzt or Wzx), ligation of the polysaccharide to the Lipid A disaccharide core (WaaL) or transport of the fully synthesized LPS to the OM (LptA, LptC, LptD, LptE), with the possible exception of LptB (Supplementary Table S2).

9. Isoprenoid, quinol and carotenoid biosynthesis

Isoprenoids play a key role in electron transport, photoprotection, light harvesting, membrane integrity and organization, and are incorporated into a range of compounds including LPSs, peptidoglycan and chlorophyll.

9.1 Isoprenoid biosynthesis

Isoprenoids, specifically undecaprenyl diphosphate, farnesyl diphosphate and geranylgeranyl diphosphate, are substrates required for biosynthesis of a wide range of compounds including hopenes, LPSs, peptidoglycan, carotenoids, phylloquinone, plastoquinone, chlorophyll and tocopherols. Geranylgeranyl diphosphate is synthesized from pyruvate and glyceraldehyde-3-P via eight enzymes, all of which are highly conserved between E. coli and Synechocystis (Supplementary Table S1; Figure 8) [147]. An additional enzyme, Ipi, is involved in isomerization of isopentenyl diphosphate and dimethylallyl diphosphate [148]. Synechocystis mutants lacking Ipi demonstrate deficient isoprenoid biosynthesis, smaller cell size and reduced TMs, and an altered cell wall [149].

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g8.jpg
Metabolism of isoprenoids, quinols and carotenoids

Carotenoids are highlighted in blue. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

9.2 Hopene biosynthesis

Hopenes are synthesized from farnesyl diphosphate in Synechocystis via two enzymes: Sqs and Shc [150]. While the exact role of hopenes has not been determined in cyanobacteria, they have been suggested to play a role in membrane integrity in non-sulfur purple photosynthetic bacteria [151]. Hopenes have been detected in the TM, PM and OM of Synechocystis sp. PCC 6714 [152]. Sqs and Shc are expressed under photoautotrophic conditions in Synechocystis [32].

9.3 Carotenoid biosynthesis

Geranylgeranyl diphosphate is the substrate for carotenoid biosynthesis. Carotenoids play a key role in assembly of photosynthetic complexes [153], membrane integrity and thylakoid organization [154], and as light harvesting and photoprotective pigments. Seven carotenoids have been detected in Synechocystis: synechoxanthin, myxol-2′-dimethylfucoside (myxoxanthophyll), zeaxanthin, 3′-hydroxy-echinenone, cis-zeaxanthin, echinenone and β-carotene [155]. The pathway has not been completely elucidated [156–158], but 12 enzymes have been demonstrated to play a role in carotenoid biosynthesis.

9.4 Tocopherol biosynthesis

Tocopherols (Vitamin E) play a role in protecting cyanobacteria from lipid peroxidation [159], cold tolerance [160] and potentially optimizing photosynthetic activity [161]. All tocopherols are synthesized from the precursor 6-methyl-6-phytyl-1,4-benzoquinol, which is synthesized by Hpt utilizing the substrates phytyl diphosphate and homogentisate [162–164]. Phytyl diphosphate is synthesized from geranylgeranyl diphosphate by ChlP [165]. Homogentisate is synthesized from 4-hydroxyphenyl-pyruvate [166], which is typically synthesized from prephenate by TyrA. However, Synechocystis TyrA demonstrates specificity only to arogenate [167], suggesting that 4-hydroxyphenyl-pyruvate may be synthesized by an alternate route. Four tocopherols (α, β, δ, γ) are produced by Synechocystis [168], although it has not been determined if each has separate roles in the cell. α- and γ-Tocopherols are synthesized from 6-methyl-6-phytyl-1,4-benzoquinol via VTE1, VTE3 and VTE4, while β and δ tocopherols are synthesized via VTE3 and VTE4 [169].

9.5 Phylloquinone and plastoquinone biosynthesis

Phylloquinone (Vitamin K1) and plastoquinone are synthesized from chorismate. Phylloquinone acts as an electron acceptor in photosystem I [170], and while not essential under photoautotrophic conditions, loss of this compound results in a severe growth defect when cells are exposed to high light conditions [171]. Phylloquinone is synthesized by ten enzymes of which several have been characterized in Synechocystis [171,172]. The majority have been identified based on homology with proteins synthesizing menaquinone (Vitamin K2) and characterized in other bacteria [173]. The second last enzyme in the pathway, MenA, utilizes phytyl diphosphate, while the last enzyme requires that dimethylphylloquinone be reduced via NAD(P)H dehydrogenase NdbB to dimethylphylloquinol, prior to synthesis of phylloquinone by MenG [174].

Plastoquinone is an essential electron carrier required for photosynthesis and respiration [23]. Despite the importance of plastoquinone, the entire biosynthetic pathway has not been determined [175]. Catalytic activity of only the first three enzymes in the pathway, UbiC, UbiA and UbiX, has been determined by expression of the Synechocystis genes in E. coli [175,176]. Deletion of a putative 4-hydroxy-3-solanesylbenzoate decarboxylases, encoded by sll0936, results in reduced plastoquinone levels [175], suggesting an uncharacterized role for this protein in its biosynthesis.

10. Chlorophyll, phycobilin and pseudocobalamin biosynthesis

Chlorophyll and phycobilins are the light harvesting pigments incorporated into photosystems and phycobilisomes, respectively. Pseudocobalamin (vitamin B12) is synthesized from the same precursor substrate, uroporphyrinogen III, and is therefore included in this section.

10.1 Haem biosynthesis

Haem, the precursor of phycobilins, is synthesized from L-glutamate and tRNAGlu via ten enzymatic steps (Figure 9). All enzymes, apart from HemJ, are highly conserved between E. coli and Synechocystis (Supplementary table 1) [177]. In contrast with E. coli, HemJ, not HemG or HemY, is the protophyrinogen IX oxidase most commonly found in cyanobacteria [178]. HemJ likely requires plastoquinone as an electron acceptor in Chlamydomonas reinhardtii [179] and localization of Synechocystis HemJ to the TM [32] suggests a similar enzymatic reaction. The Synechocystis genome also encodes additional enzymes expressed under micro-oxic conditions, including HemN1 (and possibly HemN2) [180], which can catalyse the eighth enzymatic step of haem biosynthesis, in addition to Ho2 [181,182] and ChlA2 [183], which are involved in bilin and chlorophyll biosynthesis, respectively. It should be noted that these enzymes still require oxygen for catalytic activity. However, they may bind oxygen with greater affinity than the enzymes catalysing the same step that are expressed under non-microoxic conditions. Haem does not accumulate in mutants deficient in Ho1 and Ho2, which catalyse the first steps in bilin biosynthesis, suggesting that haem is rapidly degraded by an uncharacterized pathway [182].

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g9.jpg
Metabolism of chlorophyll, phycobilin and pseudocobalamin

Proteins involved in anaerobic or low oxygen environment enzymatic steps are highlighted in blue. Cofactors in each reaction are shown with the exception of protons, water and inorganic phosphate.

10.2 Bilin biosynthesis

Haem is the substrate for biosynthesis of biliverdin, which in turn is the substrate for production of the pigments phycocyanobilin and phycoerythrobilin. These pigments are subsequently incorporated into the light harvesting phycobilisome complex [184]. Synechocystis only produces phycocyanobilin via the enzyme PcyA [184]. Synechocystis also encodes a biliverdin reductase, BvdR, resulting in production of bilirubin [185]. While the exact role of bilirubin has not been determined, deletion of BvdR results in a mutant with severely attenuated phycobilisomes.

10.3 Chlorophyll biosynthesis

Chlorophyll, the main pigment in photosystems, is synthesized from protoporphyrin IX, the immediate precursor of haem, via seven enzymatic steps. The complete pathway has been characterized in Synechocystis. The first step of chlorophyll biosynthesis is catalysed by three magnesium chelatase enzymes, ChlD, ChlH and ChlI, resulting in production of Mg-protoporphyrin IX [186]. GUN4 is also essential for magnesium chelatase activity [187–189]. The second step is catalysed by ChlM [190], while the third is catalysed via ChlA1 (AcsF) or ChlA2 [191]. Ycf54 may also be required for ChlA1 activity [192]. Two independent enzymes, a light-dependent NADPH:protochlorophyllide reductase (LPOR) or a ferredoxin-dependent DPOR complex, can catalyse the following step [193], while BciB catalyses the step after this [194,195]. Geranylgeranyl is incorporated into chlorophyll by ChlG in the second last step. In a landmark paper, expression of ChlDHI and GUN4, ChlM, ChlA1, LPOR, BciB, ChlG and ChlP in E. coli was sufficient for chlorophyll biosynthesis [196], demonstrating that no other enzymes are required in this pathway.

10.4 Pseudocobalamin biosynthesis

Cobalamin (Vitamin B12) is required for activity of MetH, involved in methionine biosynthesis (Figure 4), and may be required by certain enzymes in the quinone and folate biosynthesis pathways [197]. Cyanobacteria produce an alternate form of vitamin B12 termed pseudocobalamin [198]. Vitamin B12 is synthesized from the haem biosynthetic intermediate, uroporphyrinogen III. The cob(II)yrinate a,c-diamide component of vitamin B12 can be synthesize by either an aerobic or anaerobic pathway, which share certain enzymes [199]. These pathways have been characterized in a range of heterotrophic bacteria [199,200] but relatively few cyanobacterial enzymes have been investigated. Synechocystis encodes all the enzymes in the anaerobic pathway but is missing five in the aerobic pathway (CobG, CobF, CobK, CobS, CobT), suggesting that this biosynthetic route is not utilized (Supplementary Table S1). Several enzymes involved in converting cob(II)yrinate a,c-diamide to pseudocobalamin (CobO, CobQ, CbiB, CobU, CobV) are potentially encoded in the Synechocystis genome. However, the exact biosynthetic steps have not been determined and the pathway in Synechocystis can only be speculated based on characterized pathways from species that synthesize cobalamin [199].

Synechocystis also has the genetic potential to produce siroheme from the pseudocobalamin biosynthetic intermediate, sirohydrochlorin. Siroheme is a cofactor required for nitrite reductase [201] and possibly for other enzymes.

11. Transport systems

The majority of proteins potentially involved in metabolite transport localize to the PM (Figure 10) [32]. However, there are many putative transporters in Synechocystis with no assigned function (Supplementary Table S4), suggesting that our knowledge of cyanobacterial metabolite transport is still incomplete.

An external file that holds a picture, illustration, etc.
Object name is bsr-40-bsr20193325-g10.jpg
Proteins involved in metabolite transport and conversion of nitrogen, sulfur and phosphate based compounds

Localization of transporters in either the PM or TM is detailed. Subunits in each complex may not all be membrane localized but soluble. Cofactors in each reaction are shown with the exception of protons, water, oxygen and inorganic phosphate.

11.1 Ammonia, nitrate, nitrite and urea transport

A range of transporters are responsible for import of nitrogen-based compounds. Synechocystis encodes three ammonium transporters (Amt1, Amt2, Amt3), with Amt1 being responsible for the majority of uptake [202]. Another transporter complex, comprising four subunits, NrtA-D, imports nitrate and nitrite [203–205]. Nitrate is reduced to nitrite by NarB [206], while NirA converts nitrite to ammonium [207]. Both enzymes require electrons supplied by ferredoxin [201]. Synechocystis can also utilize urea, which is imported into the cell via a transporter complex composed of five subunits, UrtA-E [208]. Urea is converted into two molecules of ammonia via the urease complex comprising three subunits, UreA-C, which is assembled by four accessory proteins, UreD-G [209].

11.2 Amino acid transport

A range of permeases with affinity for different amino acids have been characterized in Synechocystis in an extensive study conducted by Quintero et al [210]. The basic amino acid transporter encoded by BgtA and BgtB mediates transport of L-arginine, L-histidine, L-lysine and L-glutamine. Two transporters, the Gtr complex composed of GtrA-C, and the single protein GltS system, mediate L-glutamate transport. The neutral amino acid system encoded by NatA-E mediates transport of L-alanine, glycine, L-leucine, L-proline, L-serine and L-histidine. A separate study also implicated this transporter in import of L-cysteine [211]. Whether these transporters can export amino acids or transport any of the other ten amino acids is unknown. It is also possible that uncharacterized permeases may play a role in transport of other amino acids.

11.3 Metal ion transport

The Synechocystis genome encodes a range of transporters mediating import of metal ions into the cytosol, and in the case of Cu+, into the thylakoid lumen. Additional transporters are also required for metal homeostasis and efflux.

11.3.1 Copper transport

Three copper (Cu+) transporters, CtaA, PacS and the Cop complex, have been characterized in Synechocystis. Cyanobacteria require Cu+ in the thylakoid lumen for the electron carrier plastocyanin. Proteome mapping of Synechocystis localized PacS to the PM and CtaA to the TM [32], suggesting these are the main Cu+ importers in each membrane [212]. A chaperone, Atx1, likely localizes to the cytosol but possibly also to the thylakoid lumen, binds Cu+ and delivers it to proteins requiring it for enzymatic activity [213,214]. The Cop complex, composed of CopA-C, is involved in Cu+ efflux [215]. An additional protein, CopM, binds Cu+ in the periplasm and mutants lacking this protein are highly sensitive to elevated levels of Cu+ [216].

11.3.2 Potassium transport

Synechocystis encodes two PM localized potassium (K+) uptake systems, Ktr (KtrA/KtrB) and Kdp (KdpA, KdpB, KdpC, KdpD) [217]. The Ktr system mediates rapid K+ uptake while the Kdp system maintains K+ levels under limiting conditions in the environment [217,218]. KtrC was initially incorrectly assigned as a subunit of the Ktr complex [219], but was later assigned to monoglucosyldiacylglycerol synthesis, not K+ import [134]. A third TM localized transporter, SynK [220], is responsible for K+ efflux from the thylakoid lumen [221]. An additional calcium activated, PM localized transporter, SynCak, may also be involved in potassium transport [222]. Deletion of SynCak in Synechocystis results in a mutant with altered membrane potential and greater resistance to zinc.

11.3.3 Calcium transport

Calcium (Ca2+) transport is not well understood in cyanobacteria. A putative Ca2+/H+ antiporter, SynCax, has been identified [223,224], and localizes to the TM [32]. A PM localized Ca2+ importer has not been identified. MscL has been proposed to be involved in Ca2+ export [225].

11.3.4 Iron transport

Iron is potentially imported into Synechocystis via multiple transporters, although only the Fut system is essential [226,227]. FeoB, which imports Fe2+, is the main iron transporter in Synechocystis [228]. In the Fut system, a periplasmic protein, FutA2, bind Fe3+ [229,230] prior to uptake by the FutB/FutC membrane transporter [227]. A second futA protein, FutA1, has been postulated to bind Fe3+ in the cytosol [228], although proteome mapping localized it to the PM [32]. Three ExbB–ExbD complexes identified in Synechocystis, possibly in association with TonB and one to three putative FhuA OM transporters, are also required for iron uptake [226,231]. Once imported, iron is stored in ferritin complexes (BfrA, BfrB) in the cytosol [232]. Synechocystis also encodes subunits of a putative Fe3+ dicitrate transporter, although this system is reportedly less important for iron import [104].

11.3.5 Manganese, molybdate, zinc and magnesium transport

Manganese (Mn2+) is imported into Synechocystis via the MntABC complex [233], although other low-affinity transport systems may be present. Mn2+ plays a key role in the oxygen evolving centre of photosystem II. Mnx is essential for tolerance of Synechocystis to high manganese levels and may play a role in exporting Mn2+ from the cytosol to the thylakoid lumen [234]. The Synechocystis genome encodes proteins (ModA and ModBC) with high homology to the characterized molybdate transporter of E. coli (E values = 6.32E-37 and 9.94E-51, respectively) [235], but this complex has not been characterized in a cyanobacterium. The zinc (Zn2+) transporter, composed of the ZnuA, ZnuB and ZnuC subunits, is highly conserved between E. coli and Synechocystis (Supplementary Table S1). Only the ZnuA protein has been characterized in Synechocystis [236]. A separate protein, ZiaA, is involved in Zn2+ export [237]. Atx1 may also act as a Zn2+ chaperone, in addition to its role as a Cu2+ chaperone [238]. The Synechocystis genome also encode two putative magnesium transport proteins, MgtC and MgtE [239], both of which localize to the PM [32].

11.3.6 Cation efflux systems

A number of cation efflux systems are encoded by the Synechocystis genome. The Nrs complex (NrsA, NrsB, NrsC, NrsD) was induced when cells were exposed to excess Ni2+, Co2+ and Zn2+, the CoaA transporter when cells were exposed to Co2+ and Zn2+, and the ArsB transporter by exposure to arsenic [240].

11.3.7 Sulfate transport

Sulfate is transported into cells by the SbpA/CysA/CysW/CysT system, which is highly conserved between E. coli and Synechocystis (Supplementary Table S1). Sulfate is converted into sulphide by the assimilatory pathway divided into four enzymatic steps. The enzymes catalysing the final three steps are conserved between E. coli and Synechocystis. The first enzyme in the pathway, Sat, is widely conserved in bacteria capable of sulfate reduction.

11.3.8 Phosphate transport

Synechocystis contains two systems for phosphate uptake, Pst1 and Pst2, each composed of four subunits [241,242]. The PstS subunits of each system, in addition to SphX, bind phosphate in the periplasm, prior to uptake [242]. Following uptake, phosphate can be stored in polyphosphate, which consists of polymers containing tens to hundreds of phosphates. Phosphate is converted into polyphosphate by polyphosphate kinase (Ppk1), via sequential addition of single residues [243]. A second Ppk enzyme, Ppk2, homologous to an enzyme characterized in Pseudomonas aeruginosa [244], likely synthesizes polyphosphate from ATP. Ppx catalyses depolymerization of polyphosphate, releasing inorganic phosphate [243]. Another enzyme, Ppa, converts diphosphate to phosphate and is essential in Synechocystis [243].

11.4 Sodium antiporters

Synechocystis encodes six putative sodium (Na+) antiporters [245], three of which localize to the TM (NhaS1, NhaS3, NhaS6) and two to the PM (NhaS2, NhaS5) [32]. Only NhaS3 is essential in Synechocystis [246]. NhaS3 has been suggested to play a role in maintaining not just H+ and Na+, but also K+ homeostasis [247]. Deletion of the remaining Nha antiporters did not affect growth, even when cells were exposed to high salt concentrations, suggesting that these proteins can compensate for loss of each other [246].

11.5 Organic and inorganic carbon transport

Synechocystis encodes transporters that import a range of organic carbon compounds. These include GlcP that imports glucose [36] and the Ggt complex, which imports glucosylglycerol and possibly sucrose and trehalose [248,249]. A number of transporters for inorganic carbon have been characterized in Synechocystis. These play a key role in the CO2-concentrating mechanism during photosynthesis, and include the Cmp complex (BCT1 transporter) [250,251], the SbtA transporter [252,253] and the BicA transporter [254].

11.6 Water transport

Synechocystis encodes an aquaporin water channel, aqpZ, which is required for regulating osmotic stress [255], and is essential for mixotrophic growth [256].

12. Future directions in understanding cyanobacterial metabolism

Gaining a complete understanding of cyanobacterial metabolism is dependent on optimizing the slow process of mutant generation and characterization, and developing bioinformatics tools that provide better insight into protein function, in order to easily develop enzyme assays. To bypass the laborious step of mutant generation, we are developing CyanoSource, a mutant library targeting every gene in Synechocystis. Construction of the library is outlined in Gale et al [257]. Building on our transformation and Modular Cloning (MoClo) techniques [258,259], we will collaborate with United Kingdom DNA Foundries in Norwich and Edinburgh to automate the generation of a whole genome library of gene insertion plasmids (representing 3456 coding sequences (CDSs)), and will transform Synechocystis to generate the largest available collection of known and novel cyanobacterial mutant strains.

Each CyanoSource plasmid will consist of a pUC19 based backbone into which two regions flanking the gene of interest will be inserted. Between these regions a positive selectable marker, a cassette conferring resistance to kanamycin (KanR), and a counter-selection negative selectable marker based on the cytosine deaminase protein CodA [260], will be inserted. Marked mutants will be generated by transformation of the plasmid into Synechocystis and growth of the mutant on increasing concentrations of kanamycin. If segregated mutants are not obtained on agar plates containing kanamycin concentrations of 400 μg/ml, the gene will be deemed essential. In this case, other growth conditions may be trialled, in addition to growth on different types of metabolites to generate auxotrophic mutants. Conditional mutants (i.e. specialized mutants that require an external stimulus to repress a gene) will be constructed for essential genes that cannot be removed by any of these mechanisms. Only marked mutants will be generated for CyanoSource. For generation of unmarked mutants, users can easily excise the kanR/CodA cassette and the plasmid containing just the backbone and flanking regions can be introduced into the marked mutant. Unmarked mutants are selected by growth of transformants on agar plates containing 5-fluorocytosine. CodA converts this chemical to 5-fluorouracil, which is highly toxic to many bacteria. All strains, including knockouts, partially segregated, conditional and auxotrophic mutants, and plasmids containing the flanking regions interspersed with the positive and negative selectable markers, will be made available to the academic and biotechnology community as these are constructed throughout 2020/21.

This library will allow us to determine the essential Synechocystis gene set, which can be compared with the one generated in Synechococcus via transposon mutagenesis [261]. This will provide insight into the essential gene set of the phylum. CyanoSource may also provide insights into the function of many proteins involved in metabolism. Generation of auxotrophic mutants will provide strong evidence that the encoded protein is involved in the same pathway as putative characterized homologues from other species. However, deletion of these genes may only be possible if the metabolite the encoded protein plays a role in synthesizing can be imported into the cell. Research groups with expertise in enzyme and pathway characterization but lacking expertise in generation of cyanobacterial mutants may also be encouraged to investigate the function and enzymatic activity of cyanobacterial proteins, especially in light of recent high-impact publications on characterization of Synechocystis enzymes and pathways [79,105].

A better understanding of Synechocystis metabolism will help to expand on current gaps in the metabolic biochemistry, as outlined in this review. Since it is likely that a high proportion of these pathways are conserved throughout the phylum, understanding Synechocystis metabolism will aid our understanding of cyanobacterial species that play a key role in the environment (e.g. marine Prochlorococcus and Synechococcus species) or which have characteristics ideal for biotechnology (e.g. the fast growing cyanobacteria, Synechococcus sp. PCC 11901 [262]). This will be critical in optimization of biotechnologically relevant species as renewable platforms for production of chemicals currently derived from fossil fuels.

Supplementary Material

Supplementary Refs. Material:
Supplementary Table S1:
Supplementary Table S2:
Supplementary Table S3:
Supplementary Table S4:

Abbreviations

2-OGDCα-ketoglutarate decarboxylase
CBBCalvin–Benson–Bassham
OMouter membrane
PBPpenicillin-binding protein
PMplasma membrane
PPpentose phosphate
SSADHsuccinic semialdehyde dehydrogenase
TCAtricarboxylic acid
TMthylakoid membranes

Competing Interests

The authors declare that there are no competing interests associated with the manuscript.

Funding

L.A.M. acknowledges funding support from the BBSRC Norwich Research Park Doctoral Training Partnership programme [grant number BB/S507404/1]; A.J.M and D.J.L.-S. acknowledge funding from the UK Biotechnology and Biological Sciences Research Council (BBSRC) [grant numbers BB/S020128/1 and BB/S020365/1], respectively.

Author Contribution

Conceptualization: D.J.L.-S.; Bioinformatics: L.A.M.; Writing – Original Draft Preparation: D.J.L.-S.; Writing – Review and Editing: all authors.

References

1. Rasmussen B., Fletcher I.R., Brocks J.J. and Kilburn M.R. (2008) Reassessing the first appearance of eukaryotes and cyanobacteria. Nature 455, 1101–U1109 10.1038/nature07381 [Abstract] [CrossRef] [Google Scholar]
2. Holland H.D. (2006) The oxygenation of the atmosphere and oceans. Philosophical Transact. Royal Society B-Biological Sci. 361, 903–915 10.1098/rstb.2006.1838 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
3. Zwirglmaier K. et al. . (2008) Global phylogeography of marine Synechococcus and Prochlorococcus reveals a distinct partitioning of lineages among oceanic biomes. Environ. Microbiol. 10, 147–161 [Abstract] [Google Scholar]
4. Galloway J.N. et al. . (2004) Nitrogen cycles: past, present, and future. Biogeochemistry 70, 153–226 10.1007/s10533-004-0370-0 [CrossRef] [Google Scholar]
5. Moisander P.H. et al. . (2017) Chasing after Non-cyanobacterial Nitrogen Fixation in Marine Pelagic Environments. Front. Microbiol. 8, 1736 10.3389/fmicb.2017.01736 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
6. Lea-Smith D.J. and Howe C.J. (2017) The Use of Cyanobacteria for Biofuel Production. In Biofuels and Bioenergy(Love J. and Bryant J.A., eds), Wiley [Google Scholar]
7. Ducat D.C., Way J.C. and Silver P.A. (2011) Engineering cyanobacteria to generate high-value products. Trends Biotechnol. 29, 95–103 10.1016/j.tibtech.2010.12.003 [Abstract] [CrossRef] [Google Scholar]
8. Santos-Merino M., Singh A.K. and Ducat D.C. (2019) New Applications of Synthetic Biology Tools for Cyanobacterial Metabolic Engineering. Front. Bioeng. Biotechnol. 7, 33 10.3389/fbioe.2019.00033 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
9. (2019) Algae Products - Global Market Outlook (2017-2026), Stratistics Market Research Consulting Pvt Ltd [Google Scholar]
10. Pearson L.A. et al. . (2016) The genetics, biosynthesis and regulation of toxic specialized metabolites of cyanobacteria. Harmful Algae 54, 98–111 10.1016/j.hal.2015.11.002 [Abstract] [CrossRef] [Google Scholar]
11. Tan L.T. (2007) Bioactive natural products from marine cyanobacteria for drug discovery. Phytochemistry 68, 954–979 10.1016/j.phytochem.2007.01.012 [Abstract] [CrossRef] [Google Scholar]
12. Howe C.J., Barbrook A.C., Nisbet R.E.R., Lockhart P.J. and Larkum A.W.D. (2008) The origin of plastids. Philos. Trans. R. Soc. Lond. B Biol. Sci. 363, 2675–2685 10.1098/rstb.2008.0050 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
13. Silva P. et al. . (2003) FtsH is involved in the early stages of repair of photosystem II in Synechocystis sp PCC 6803. Plant Cell 15, 2152–2164 10.1105/tpc.012609 [Abstract] [CrossRef] [Google Scholar]
14. Kurisu G., Zhang H., Smith J.L. and Cramer W.A. (2003) Structure of the cytochrome b6f complex of oxygenic photosynthesis: tuning the cavity. Science 302, 1009–1014 10.1126/science.1090165 [Abstract] [CrossRef] [Google Scholar]
15. Lin M.T., Occhialini A., Andralojc P.J., Parry M.A.J. and Hanson M.R. (2014) A faster Rubisco with potential to increase photosynthesis in crops. Nature 513, 547–+ 10.1038/nature13776 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
16. Zarzycki J., Axen S.D., Kinney J.N. and Kerfeld C.A. (2013) Cyanobacterial-based approaches to improving photosynthesis in plants. J. Exp. Bot. 64, 787–798 10.1093/jxb/ers294 [Abstract] [CrossRef] [Google Scholar]
17. Miyagawa Y., Tamoi M. and Shigeoka S. (2001) Overexpression of a cyanobacterial fructose-1,6-/sedoheptulose-1,7-bisphosphatase in tobacco enhances photosynthesis and growth. Nat. Biotechnol. 19, 965–969 10.1038/nbt1001-965 [Abstract] [CrossRef] [Google Scholar]
18. Long B.M. et al. . (2018) Carboxysome encapsulation of the CO2-fixing enzyme Rubisco in tobacco chloroplasts. Nat. Commun. 9, 3570 10.1038/s41467-018-06044-0 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
19. Orr D.J. et al. . (2020) Hybrid cyanobacterial-tobacco Rubisco supports autotrophic growth and pre-carboxysomal aggregation. Plant Physiol. 182, 807–818 [Abstract] [Google Scholar]
20. Ghatak S., King Z.A., Sastry A. and Palsson B.O. (2019) The y-ome defines the 35% of Escherichia coli genes that lack experimental evidence of function. Nucleic Acids Res. 47, 2446–2454 10.1093/nar/gkz030 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
21. Fujisawa T. et al. . (2017) CyanoBase: a large-scale update on its 20th anniversary. Nucleic Acids Res. 45, D551–D554 10.1093/nar/gkw1131 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
22. Flores E., Nieves-Morion M. and Mullineaux C.W. (2018) Cyanobacterial Septal Junctions: Properties and Regulation. Life-Basel 9, 1. [Europe PMC free article] [Abstract] [Google Scholar]
23. Lea-Smith D.J., Bombelli P., Vasudevan R. and Howe C.J. (2016) Photosynthetic, respiratory and extracellular electron transport pathways in cyanobacteria. Biochim. Biophys. Acta - Bioenergetics 1857, 247–255 10.1016/j.bbabio.2015.10.007 [Abstract] [CrossRef] [Google Scholar]
24. Stanier R.Y. and Cohen-Bazire G. (1977) Phototrophic prokaryotes: the cyanobacteria. Annu. Rev. Microbiol. 31, 225–274 10.1146/annurev.mi.31.100177.001301 [Abstract] [CrossRef] [Google Scholar]
25. Trautner C. and Vermaas W.F.J. (2013) The sll1951 gene encodes the surface layer protein of Synechocystis sp. strain PCC 6803. J. Bacteriol. 195, 5370–5380 10.1128/JB.00615-13 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
26. Nevo R. et al. . (2007) Thylakoid membrane perforations and connectivity enable intracellular traffic in cyanobacteria. EMBO J. 26, 1467–1473 10.1038/sj.emboj.7601594 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
27. van de Meene A.M.L., Hohmann-Marriott M.F., Vermaas W.F.J. and Roberson R.W. (2006) The three-dimensional structure of the cyanobacterium Synechocystis sp. PCC 6803. Arch. Microbiol. 184, 259–270 10.1007/s00203-005-0027-y [Abstract] [CrossRef] [Google Scholar]
28. Liberton M., Howard Berg R., Heuser J., Roth R. and Pakrasi H.B. (2006) Ultrastructure of the membrane systems in the unicellular cyanobacterium Synechocystis sp. strain PCC 6803. Protoplasma 227, 129–138 10.1007/s00709-006-0145-7 [Abstract] [CrossRef] [Google Scholar]
29. Saw J.H.W. et al. . (2013) Cultivation and Complete Genome Sequencing of Gloeobacter kilaueensis sp. nov., from a Lava Cave in Kīlauea Caldera, Hawai'i. PLoS ONE 8, e76376–e76376 10.1371/journal.pone.0076376 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
30. Rippka R., Waterbury J. and Cohenbazire G. (1974) Cyanobacterium Which Lacks Thylakoids. Arch. Microbiol. 100, 419–436 10.1007/BF00446333 [CrossRef] [Google Scholar]
31. Liberton M. et al. . (2016) Global proteomic analysis reveals an exclusive role of thylakoid membranes in bioenergetics of a model cyanobacterium. Mol. Cell. Proteomics 15, 2021–2032 10.1074/mcp.M115.057240 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
32. Baers L.L. et al. . (2019) Proteome mapping of a cyanobacterium reveals distinct compartment organisation and cell-dispersed metabolism. Plant Physiol. 181, 1721–1738 10.1104/pp.19.00897 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
33. Sarcina M. and Mullineaux C.W. (2000) Effects of tubulin assembly inhibitors on cell division in prokaryotes in vivo. FEMS Microbiol. Lett. 191, 25–29 10.1111/j.1574-6968.2000.tb09314.x [Abstract] [CrossRef] [Google Scholar]
34. Rast A. et al. . (2019) Biogenic regions of cyanobacterial thylakoids form contact sites with the plasma membrane. Nat. Plants 5, 436–446 10.1038/s41477-019-0399-7 [Abstract] [CrossRef] [Google Scholar]
35. Williams J.G.K. (1988) Construction of specific mutations in photosystem-II photosynthetic reaction center by genetic-engineering methods in Synechocystis-6803. Methods Enzymol. 167, 766–778 10.1016/0076-6879(88)67088-1 [CrossRef] [Google Scholar]
36. Schmetterer G.R. (1990) Sequence Conservation among the Glucose Transporter from the Cyanobacterium-Synechocystis Sp Pcc-6803 and Mammalian Glucose Transporters. Plant Mol. Biol. 14, 697–706 10.1007/BF00016502 [Abstract] [CrossRef] [Google Scholar]
37. Chen X. et al. . (2016) The Entner-Doudoroff pathway is an overlooked glycolytic route in cyanobacteria and plants. PNAS 113, 5441–5446 10.1073/pnas.1521916113 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
38. Patipong T. et al. . (2019) A class I fructose-1,6-bisphosphate aldolase is associated with salt stress tolerance in a halotolerant cyanobacterium Halothece sp. PCC 7418. Arch. Biochem. Biophys. 672, 108059 10.1016/j.abb.2019.07.024 [Abstract] [CrossRef] [Google Scholar]
39. Sundaram S., Karakaya H., Scanlan D.J. and Mann N.H. (1998) Multiple oligomeric forms of glucose-6-phosphate dehydrogenase in cyanobacteria and the role of OpcA in the assembly process. Microbiology 144, 1549–1556 10.1099/00221287-144-6-1549 [Abstract] [CrossRef] [Google Scholar]
40. Ozkul K. and Karakaya H. (2015) Characterisation of an opcA Mutant of the Unicellular Cyanobacterium Synechocystis sp. PCC 6803. Curr. Microbiol. 71, 572–578 10.1007/s00284-015-0889-4 [Abstract] [CrossRef] [Google Scholar]
41. Makowka A. et al. . (2020) Glycolytic Shunts Replenish the Calvin-Benson-Bassham Cycle as Anaplerotic Reactions in Cyanobacteria. Mol. Plant 13, 471–482 10.1016/j.molp.2020.02.002 [Abstract] [CrossRef] [Google Scholar]
42. Koksharova O., Schubert M., Shestakov S. and Cerff R. (1998) Genetic and biochemical evidence for distinct key functions of two highly divergent GAPDH genes in catabolic and anabolic carbon flow of the cyanobacterium Synechocystis sp. PCC 6803. Plant Mol. Biol. 36, 183–194 10.1023/A:1005925732743 [Abstract] [CrossRef] [Google Scholar]
43. Eisenhut M. et al. . (2008) The photorespiratory glycolate metabolism is essential for cyanobacteria and might have been conveyed endosymbiontically to plants. PNAS 105, 17199–17204 10.1073/pnas.0807043105 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
44. Eisenhut M., Roell M.S. and Weber A.P.M. (2019) Mechanistic understanding of photorespiration paves the way to a new green revolution. New Phytol. 223, 1762–1769 10.1111/nph.15872 [Abstract] [CrossRef] [Google Scholar]
45. Eisenhut M. et al. . (2006) The plant-like C2 glycolate cycle and the bacterial-like glycerate pathway cooperate in phosphoglycolate metabolism in cyanobacteria. Plant Physiol. 142, 333–342 10.1104/pp.106.082982 [Abstract] [CrossRef] [Google Scholar]
46. Zilliges Y. (2014) Glycogen, a dynamic cellular sink and reservoir for carbon. In Cell Biology of Cyanobacteria(Flores E., editor. , ed.), Caister Academic Press [Google Scholar]
47. Yoo S.H., Lee B.H., Moon Y., Spalding M.H. and Jane J.L. (2014) Glycogen synthase isoforms in Synechocystis sp. PCC6803: identification of different roles to produce glycogen by targeted mutagenesis. PLoS ONE 9, e91524–e91524 10.1371/journal.pone.0091524 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
48. Fu J. and Xu X.D. (2006) The functional divergence of two glgP homologues in Synechocystis sp PCC 6803. FEMS Microbiol. Lett. 260, 201–209 10.1111/j.1574-6968.2006.00312.x [Abstract] [CrossRef] [Google Scholar]
49. Taroncher-Oldenberg G., Nishina K. and Stephanopoulos G. (2000) Identification and analysis of the polyhydroxyalkanoate-specific beta-ketothiolase and acetoacetyl coenzyme A reductase genes in the cyanobacterium Synechocystis sp strain PCC6803. Appl. Environ. Microbiol. 66, 4440–4448 10.1128/AEM.66.10.4440-4448.2000 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
50. Hein S., Tran H. and Steinbuchel A. (1998) Synechocystis sp. PCC6803 possesses a two-component polyhydroxyalkanoic acid synthase similar to that of anoxygenic purple sulfur bacteria. Arch. Microbiol. 170, 162–170 10.1007/s002030050629 [Abstract] [CrossRef] [Google Scholar]
51. Koch M., Doello S., Gutekunst K. and Forchhammer K. (2019) PHB is Produced from Glycogen Turn-over during Nitrogen Starvation in Synechocystis sp. PCC 6803. Int. J. Mol. Sci. 20, E1942 10.3390/ijms20081942 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
52. Zhang S. and Bryant D.A. (2011) The tricarboxylic acid cycle in cyanobacteria. Science 334, 1551–1553 10.1126/science.1210858 [Abstract] [CrossRef] [Google Scholar]
53. Cooley J.W. and Vermaas W.F. (2001) Succinate dehydrogenase and other respiratory pathways in thylakoid membranes of Synechocystis sp. strain PCC 6803: capacity comparisons and physiological function. J. Bacteriol. 183, 4251–4258 10.1128/JB.183.14.4251-4258.2001 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
54. Zhang S., Qian X., Chang S., Dismukes G.C. and Bryant D.A. (2016) Natural and Synthetic Variants of the Tricarboxylic Acid Cycle in Cyanobacteria: Introduction of the GABA Shunt into Synechococcus sp. PCC 7002. Front. Microbiol. 7, 1972 10.3389/fmicb.2016.01972 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
55. Ito S., Koyama N. and Osanai T. (2019) Citrate synthase from Synechocystis is a distinct class of bacterial citrate synthase. Sci. Rep. 9, 6038 10.1038/s41598-019-42659-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
56. Katayama N., Takeya M. and Osanai T. (2019) Biochemical characterisation of fumarase C from a unicellular cyanobacterium demonstrating its substrate affinity, altered by an amino acid substitution. Sci. Rep. 9, 10629 10.1038/s41598-019-47025-7 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
57. Takeya M., Ito S., Sukigara H. and Osanai T. (2018) Purification and Characterisation of Malate Dehydrogenase From Synechocystis sp PCC 6803: Biochemical Barrier of the Oxidative Tricarboxylic Acid Cycle. Front. Plant Sci. 9, 947 10.3389/fpls.2018.00947 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
58. Knoop H. et al. . (2013) Flux Balance Analysis of Cyanobacterial Metabolism: The Metabolic Network of Synechocystis sp PCC 6803. PLoS Comput. Biol. 9, e100308110.1371/journal.pcbi.1003081 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
59. Takeya M., Hirai M.Y. and Osanai T. (2017) Allosteric Inhibition of Phosphoenolpyruvate Carboxylases is Determined by a Single Amino Acid Residue in Cyanobacteria. Sci. Rep. 7, 4108010.1038/srep41080 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
60. Yang C., Hua Q. and Shimizu K. (2002) Metabolic flux analysis in Synechocystis using isotope distribution from C-13-labeled glucose. Metab. Eng. 4, 202–216 10.1006/mben.2002.0226 [Abstract] [CrossRef] [Google Scholar]
61. Young J.D., Shastri A.A., Stephanopoulos G. and Morgan J.A. (2012) Mapping photoautotrophic metabolism with isotopically nonstationary C-13 flux analysis (vol 13, pg 656, 2011). Metab. Eng. 14, 185–185 10.1016/j.ymben.2012.01.005 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
62. Bricker T.M. et al. . (2004) The malic enzyme is required for optimal photo autotrophic growth of Synechocystis sp strain PCC 6803 under continuous light but not under a diurnal light regimen. J. Bacteriol. 186, 8144–8148 10.1128/JB.186.23.8144-8148.2004 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
63. Osanai T. et al. . (2015) Genetic manipulation of a metabolic enzyme and a transcriptional regulator increasing succinate excretion from unicellular cyanobacterium. Front. Microbiol. 6, 1064 10.3389/fmicb.2015.01064 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
64. Hasunuma T., Matsuda M., Kato Y., Vavricka C.J. and Kondo A. (2018) Temperature enhanced succinate production concurrent with increased central metabolism turnover in the cyanobacterium Synechocystis sp PCC 6803. Metab. Eng. 48, 109–120 10.1016/j.ymben.2018.05.013 [Abstract] [CrossRef] [Google Scholar]
65. Brey L.F. et al. . (2019) Metabolic engineering of Synechocystis sp. PCC 6803 for the production of aromatic amino acids and derived phenylpropanoids. Metab. Eng. 57, 129–139 10.1016/j.ymben.2019.11.002 [Abstract] [CrossRef] [Google Scholar]
66. Schmidt W., Weckesser J. and Mayer H. (1980) Lipopolysaccharides in four strains of the unicellular cyanobacterium Synechocystis. Arch. Microbiol. 127, 217–222 10.1007/BF00427196 [CrossRef] [Google Scholar]
67. Maeda K., Narikawa R. and Ikeuchi M. (2014) CugP Is a Novel Ubiquitous Non-GalU-Type Bacterial UDP-Glucose Pyrophosphorylase Found in Cyanobacteria. J. Bacteriol. 196, 2348–2354 10.1128/JB.01591-14 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
68. Mohamed H.E., van de Meene A.M.L., Roberson R.W. and Vermaas W.F.J. (2005) Myxoxanthophyll is required for normal cell wall structure and thylakoid organization in the cyanobacterium, Synechocystis sp strain PCC 6803. J. Bacteriol. 187, 6883–6892 10.1128/JB.187.20.6883-6892.2005 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
69. Marin K., Zuther E., Kerstan T., Kunert A. and Hagemann M. (1998) The ggpS gene from Synechocystis sp. strain PCC 6803 encoding glucosyl-glycerol-phosphate synthase is involved in osmolyte synthesis. J. Bacteriol. 180, 4843–4849 10.1128/JB.180.18.4843-4849.1998 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
70. Miao X.L., Wu Q.Y., Wu G.F. and Zhao N.M. (2003) Sucrose accumulation in salt-stressed cells of agp gene deletion-mutant in cyanobacterium Synechocystis sp PCC 6803. FEMS Microbiol. Lett. 218, 71–77 10.1111/j.1574-6968.2003.tb11500.x [Abstract] [CrossRef] [Google Scholar]
71. Curatti L. et al. . (1998) Sucrose-phosphate synthase from Synechocystis sp. Strain PCC 6803: Identification of the spsA gene and characterization of the enzyme expressed in Escherichia coli. J. Bacteriol. 180, 6776–6779 10.1128/JB.180.24.6776-6779.1998 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
72. Salerno G.L. and Curatti L. (2003) Origin of sucrose metabolism in higher plants: when, how and why? Trends Plant Sci. 8, 63–69 10.1016/S1360-1385(02)00029-8 [Abstract] [CrossRef] [Google Scholar]
73. Kirsch F., Luo Q., Lu X. and Hagemann M. (2018) Inactivation of invertase enhances sucrose production in the cyanobacterium Synechocystis sp. PCC 6803. Microbiology 164, 1220–1228 10.1099/mic.0.000708 [Abstract] [CrossRef] [Google Scholar]
74. Hagemann M. and Erdmann N. (1994) Activation and Pathway of Glucosylglycerol Synthesis in the Cyanobacterium-Synechocystis Sp Pcc-6803. Microbiology-Sgm 140, 1427–1431 10.1099/00221287-140-6-1427 [CrossRef] [Google Scholar]
75. Guardiola J., De Felice M., Lamberti A. and Iaccarino M. (1977) The acetolactate synthase isoenzymes of Escherichia coli K-12. Mol. Gen. Genet. 156, 17–25 10.1007/BF00272247 [Abstract] [CrossRef] [Google Scholar]
76. GarciaDominguez M., Reyes J.C. and Florencio F.J. (1997) Purification and characterization of a new type of glutamine synthetase from cyanobacteria. Eur. J. Biochem. 244, 258–264 10.1111/j.1432-1033.1997.00258.x [Abstract] [CrossRef] [Google Scholar]
77. Navarro F., Martin-Figueroa E., Candau P. and Florencio F.J. (2000) Ferredoxin-dependent iron-sulfur flavoprotein glutamate synthase (GlsF) from the cyanobacterium Synechocystis sp PCC 6803: Expression and assembly in Escherichia coli. Arch. Biochem. Biophys. 379, 267–276 10.1006/abbi.2000.1894 [Abstract] [CrossRef] [Google Scholar]
78. Liu L.K., Becker D.F. and Tanner J.J. (2017) Structure, function, and mechanism of proline utilization A (PutA). Arch. Biochem. Biophys. 632, 142–157 10.1016/j.abb.2017.07.005 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
79. Zhang H. et al. . (2018) The cyanobacterial ornithine-ammonia cycle involves an arginine dihydrolase. Nat. Chem. Biol. 14, 575–581 10.1038/s41589-018-0038-z [Abstract] [CrossRef] [Google Scholar]
80. Kurihara S. et al. . (2005) A novel putrescine utilization pathway involves gamma-glutamylated intermediates of Escherichia coli K-12. J. Biol. Chem. 280, 4602–4608 10.1074/jbc.M411114200 [Abstract] [CrossRef] [Google Scholar]
81. Flores E., Arevalo S. and Burnat M. (2019) Cyanophycin and arginine metabolism in cyanobacteria. Algal Res.-Biomass Biofuels Bioproducts 42, 101577 [Google Scholar]
82. Hudson A.O., Singh B.K., Leustek T. and Gilvarg C. (2006) An LL-diaminopimelate aminotransferase defines a novel variant of the lysine biosynthesis pathway in plants. Plant Physiol. 140, 292–301 10.1104/pp.105.072629 [Abstract] [CrossRef] [Google Scholar]
83. Hudson A.O., Gilvarg C. and Leustek T. (2008) Biochemical and phylogenetic characterization of a novel diaminopimelate biosynthesis pathway in prokaryotes identifies a diverged form of LL-diaminopimelate aminotransferase. J. Bacteriol. 190, 3256–3263 10.1128/JB.01381-07 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
84. Zhao G.P., Somerville R.L. and Chitnis P.R. (1994) Synechocystis Pcc-6803 Contains a Single-Gene for the Beta-Subunit of Tryptophan Synthase with Strong Homology to the Trpb Genes of Arabidopsis and Maize (Zea-Mays L). Plant Physiol. 104, 461–466 10.1104/pp.104.2.461 [Abstract] [CrossRef] [Google Scholar]
85. Legrand P. et al. . (2006) Biochemical characterization and crystal structure of Synechocystis arogenate dehydrogenase provide insights into catalytic reaction. Structure 14, 767–776 10.1016/j.str.2006.01.006 [Abstract] [CrossRef] [Google Scholar]
86. Klemke F. et al. . (2015) Identification of the light-independent phosphoserine pathway as an additional source of serine in the cyanobacterium Synechocystis sp. PCC 6803. Microbiology 161, 1050–1060 10.1099/mic.0.000055 [Abstract] [CrossRef] [Google Scholar]
87. Tirupati B., Vey J.L., Drennan C.L. and Bollinger J.M. Jr. (2004) Kinetic and structural characterization of Slr0077/SufS, the essential cysteine desulfurase from Synechocystis sp. PCC 6803. Biochemistry 43, 12210–12219 10.1021/bi0491447 [Abstract] [CrossRef] [Google Scholar]
88. Cameron J.C. and Pakrasi H.B. (2011) Glutathione in Synechocystis 6803: a closer look into the physiology of a gshB mutant. Plant Signal Behav. 6, 89–92 10.4161/psb.6.1.14145 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
89. Cameron J.C. and Pakrasi H.B. (2010) Essential Role of Glutathione in Acclimation to Environmental and Redox Perturbations in the Cyanobacterium Synechocystis sp. PCC 6803. Plant Physiol. 154, 1672–1685 10.1104/pp.110.162990 [Abstract] [CrossRef] [Google Scholar]
90. Kato S., Mihara H., Kurihara T., Yoshimura T. and Esaki N. (2000) Gene cloning, purification, and characterization of two cyanobacterial NifS homologs driving iron-sulfur cluster formation. Biosci. Biotechnol. Biochem. 64, 2412–2419 10.1271/bbb.64.2412 [Abstract] [CrossRef] [Google Scholar]
91. Jaschkowitz K. and Seidler A. (2000) Role of a NifS-like protein from the cyanobacterium Synechocystis PCC 6803 in the maturation of FeS proteins. Biochemistry 39, 3416–3423 10.1021/bi992147c [Abstract] [CrossRef] [Google Scholar]
92. Kessler D. (2004) Slr0077 of Synechocystis has cysteine desulfurase as well as cystine lyase activity. Biochem. Biophys. Res. Commun. 320, 571–577 10.1016/j.bbrc.2004.06.007 [Abstract] [CrossRef] [Google Scholar]
93. Kessler D. and Papenbrock J. (2005) Iron-sulfur cluster biosynthesis in photosynthetic organisms. Photosynth. Res. 86, 391–407 10.1007/s11120-005-5913-2 [Abstract] [CrossRef] [Google Scholar]
94. Fontecave M. and Ollagnier-de-Choudens S. (2008) Iron-sulfur cluster biosynthesis in bacteria: Mechanisms of cluster assembly and transfer. Arch. Biochem. Biophys. 474, 226–237 10.1016/j.abb.2007.12.014 [Abstract] [CrossRef] [Google Scholar]
95. Wollenberg M., Berndt C., Bill E., Schwenn J.D. and Seidler A. (2003) A dimer of the FeS cluster biosynthesis protein IscA from cyanobacteria binds a [2Fe2S] cluster between two protomers and transfers it to [2Fe2S] and [4Fe4S] apo proteins. Eur. J. Biochem. 270, 1662–1671 10.1046/j.1432-1033.2003.03522.x [Abstract] [CrossRef] [Google Scholar]
96. Kappock T.J., Ealick S.E. and Stubbe J. (2000) Modular evolution of the purine biosynthetic pathway. Curr. Opin. Chem. Biol. 4, 567–572 10.1016/S1367-5931(00)00133-2 [Abstract] [CrossRef] [Google Scholar]
97. Zhang Y., Morar M. and Ealick S.E. (2008) Structural biology of the purine biosynthetic pathway. Cell. Mol. Life Sci. 65, 3699–3724 10.1007/s00018-008-8295-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
98. Lascu I. and Gonin P. (2000) The catalytic mechanism of nucleoside diphosphate kinases. J. Bioenerg. Biomembr. 32, 237–246 10.1023/A:1005532912212 [Abstract] [CrossRef] [Google Scholar]
99. Larsson K.M. et al. . (2004) Structural mechanism of allosteric substrate specificity regulation in a ribonucleotide reductase. Nat. Struct. Mol. Biol. 11, 1142–1149 10.1038/nsmb838 [Abstract] [CrossRef] [Google Scholar]
100. Turnbough C.L. Jr and Switzer R.L. (2008) Regulation of pyrimidine biosynthetic gene expression in bacteria: repression without repressors. Microbiol. Mol. Biol. Rev. 72, 266–300, table of contents 10.1128/MMBR.00001-08 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
101. Kilstrup M., Hammer K., Ruhdal Jensen P. and Martinussen J. (2005) Nucleotide metabolism and its control in lactic acid bacteria. FEMS Microbiol. Rev. 29, 555–590 10.1016/j.fmrre.2005.04.006 [Abstract] [CrossRef] [Google Scholar]
102. Waldrop G.L., Holden H.M. and St Maurice M. (2012) The enzymes of biotin dependent CO2 metabolism: What structures reveal about their reaction mechanisms. Protein Sci. 21, 1597–1619 10.1002/pro.2156 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
103. Lin S. and Cronan J.E. (2011) Closing in on complete pathways of biotin biosynthesis. Mol. Biosyst. 7, 1811–1821 10.1039/c1mb05022b [Abstract] [CrossRef] [Google Scholar]
104. Rodionov D.A., Mironov A.A. and Gelfand M.S. (2002) Conservation of the biotin regulon and the BirA regulatory signal in Eubacteria and Archaea. Genome Res. 12, 1507–1516 10.1101/gr.314502 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
105. Sakaki K. et al. . (2020) A suicide enzyme catalyzes multiple reactions for biotin biosynthesis in cyanobacteria. Nat. Chem. Biol. 16, 415–422 10.1038/s41589-019-0461-9 [Abstract] [CrossRef] [Google Scholar]
106. Stancek M., Schnell R. and Ryden-Aulin M. (2005) Analysis of Escherichia coli nicotinate mononucleotide adenylyltransferase mutants in vivo and in vitro. BMC Biochem. 6, 16 10.1186/1471-2091-6-16 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
107. Gerdes S.Y. et al. . (2006) Comparative genomics of NAD biosynthesis in cyanobacteria. J. Bacteriol. 188, 3012–3023 10.1128/JB.188.8.3012-3023.2006 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
108. Raffaelli N. et al. . (1999) Synechocystis sp. slr0787 protein is a novel bifunctional enzyme endowed with both nicotinamide mononucleotide adenylyltransferase and ‘Nudix’ hydrolase activities. FEBS Lett. 444, 222–226 10.1016/S0014-5793(99)00068-X [Abstract] [CrossRef] [Google Scholar]
109. Ishikawa Y. and Kawai-Yamada M. (2019) Physiological Significance of NAD Kinases in Cyanobacteria. Front. Plant Sci. 10, 847 10.3389/fpls.2019.00847 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
110. Kämäräinen J. et al. . (2017) Pyridine nucleotide transhydrogenase PntAB is essential for optimal growth and photosynthetic integrity under low-light mixotrophic conditions in Synechocystis sp. PCC 6803. New Phytol. 214, 194–204 10.1111/nph.14353 [Abstract] [CrossRef] [Google Scholar]
111. de Crecy-Lagard V., El Yacoubi B., de la Garza R.D., Noiriel A. and Hanson A.D. (2007) Comparative genomics of bacterial and plant folate synthesis and salvage: predictions and validations. BMC Genomics 8, 245 10.1186/1471-2164-8-245 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
112. Hanson A.D., Pribat A., Waller J.C. and de Crecy-Lagard V. (2009) ‘Unknown’ proteins and ‘orphan’ enzymes: the missing half of the engineering parts list–and how to find it. Biochem. J. 425, 1–11 10.1042/BJ20091328 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
113. Rubio L.M., Flores E. and Herrero A. (1999) Molybdopterin guanine dinucleotide cofactor in Synechococcus sp. nitrate reductase: identification of mobA and isolation of a putative moeB gene. FEBS Lett. 462, 358–362 10.1016/S0014-5793(99)01556-2 [Abstract] [CrossRef] [Google Scholar]
114. Rubio L.M., Flores E. and Herrero A. (1998) The narA locus of Synechococcus sp. strain PCC 7942 consists of a cluster of molybdopterin biosynthesis genes. J. Bacteriol. 180, 1200–1206 10.1128/JB.180.5.1200-1206.1998 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
115. Wuebbens M.M. and Rajagopalan K.V. (2003) Mechanistic and mutational studies of Escherichia coli molybdopterin synthase clarify the final step of molybdopterin biosynthesis. J. Biol. Chem. 278, 14523–14532 10.1074/jbc.M300453200 [Abstract] [CrossRef] [Google Scholar]
116. Du Q.L., Wang H.H. and Xie J.P. (2011) Thiamin (Vitamin B1) Biosynthesis and Regulation: A Rich Source of Antimicrobial Drug Targets? Int. J. Biol. Sci. 7, 41–52 10.7150/ijbs.7.41 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
117. Rodionov D.A., Vitreschak A.G., Mironov A.A. and Gelfand M.S. (2002) Comparative genomics of thiamin biosynthesis in procaryotes. New genes and regulatory mechanisms. J. Biol. Chem. 277, 48949–48959 10.1074/jbc.M208965200 [Abstract] [CrossRef] [Google Scholar]
118. Leonardi R. and Jackowski S. (2007) Biosynthesis of Pantothenic Acid and Coenzyme A. EcoSal Plus 2, 1 10.1128/ecosalplus.3.6.3.4 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
119. Brand L.A. and Strauss E. (2005) Characterization of a new pantothenate kinase isoform from Helicobacter pylori. J. Biol. Chem. 280, 20185–20188 10.1074/jbc.C500044200 [Abstract] [CrossRef] [Google Scholar]
120. Richts B., Rosenberg J. and Commichau F.M. (2019) A Survey of Pyridoxal 5′-Phosphate-Dependent Proteins in the Gram-Positive Model Bacterium Bacillus subtilis. Front. Mol. Biosci. 6, 32 10.3389/fmolb.2019.00032 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
121. Laber B., Maurer W., Scharf S., Stepusin K. and Schmidt F.S. (1999) Vitamin B6 biosynthesis: formation of pyridoxine 5′-phosphate from 4-(phosphohydroxy)-L-threonine and 1-deoxy-D-xylulose-5-phosphate by PdxA and PdxJ protein. FEBS Lett. 449, 45–48 10.1016/S0014-5793(99)00393-2 [Abstract] [CrossRef] [Google Scholar]
122. Sheng J., Vannela R. and Rittrnann B.E. (2011) Evaluation of methods to extract and quantify lipids from Synechocystis PCC 6803. Bioresour. Technol. 102, 1697–1703 10.1016/j.biortech.2010.08.007 [Abstract] [CrossRef] [Google Scholar]
123. Lea-Smith D.J. et al. . (2016) Hydrocarbons are essential for optimal cell size, division and growth of cyanobacteria. Plant Physiol. 172, 1928–1940 10.1104/pp.16.01205 [Abstract] [CrossRef] [Google Scholar]
124. Tasaka Y. et al. . (1996) Targeted mutagenesis of acyl-lipid desaturases in Synechocystis: evidence for the important roles of polyunsaturated membrane lipids in growth, respiration and photosynthesis. EMBO J. 15, 6416–6425 10.1002/j.1460-2075.1996.tb01033.x [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
125. von Berlepsch S. et al. . (2012) The acyl-acyl carrier protein synthetase from Synechocystis sp. PCC 6803 mediates fatty acid import. Plant Physiol. 159, 606–617 10.1104/pp.112.195263 [Abstract] [CrossRef] [Google Scholar]
126. Kaczmarzyk D. and Fulda M. (2010) Fatty Acid Activation in Cyanobacteria Mediated by Acyl-Acyl Carrier Protein Synthetase Enables Fatty Acid Recycling. Plant Physiol. 152, 1598–1610 10.1104/pp.109.148007 [Abstract] [CrossRef] [Google Scholar]
127. Mendez-Perez D., Begemann M.B. and Pfleger B.F. (2011) Modular synthase-encoding gene involved in α-olefin biosynthesis in Synechococcus sp. strain PCC 7002. Appl. Environ. Microbiol. 77, 4264–4267 10.1128/AEM.00467-11 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
128. Schirmer A., Rude M.A., Li X., Popova E. and del Cardayre S.B. (2010) Microbial biosynthesis of alkanes. Science 329, 559–562 10.1126/science.1187936 [Abstract] [CrossRef] [Google Scholar]
129. Rajakovich L.J. et al. . (2015) Rapid Reduction of the Diferric-Peroxyhemiacetal Intermediate in Aldehyde-Deformylating Oxygenase by a Cyanobacterial Ferredoxin: Evidence for a Free-Radical Mechanism. J. Am. Chem. Soc. 137, 11695–11709 10.1021/jacs.5b06345 [Abstract] [CrossRef] [Google Scholar]
130. Lea-Smith D.J. et al. . (2015) Contribution of cyanobacterial alkane production to the ocean hydrocarbon cycle. PNAS 112, 13591–13596 10.1073/pnas.1507274112 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
131. Weier D., Muller C., Gaspers C. and Frentzen M. (2005) Characterisation of acyltransferases from Synechocystis sp PCC6803. Biochem. Biophys. Res. Commun. 334, 1127–1134 10.1016/j.bbrc.2005.06.197 [Abstract] [CrossRef] [Google Scholar]
132. Hagio M. et al. . (2000) Direct evidence for requirement of phosphatidylglycerol in photosystem II of photosynthesis. Plant Physiol. 124, 795–804 10.1104/pp.124.2.795 [Abstract] [CrossRef] [Google Scholar]
133. Awai K. et al. . (2006) Comparative genomic analysis revealed a gene for monoglucosyldiacylglycerol synthase, an enzyme for photosynthetic membrane lipid synthesis in cyanobacteria. Plant Physiol. 141, 1120–1127 10.1104/pp.106.082859 [Abstract] [CrossRef] [Google Scholar]
134. Awai K., Watanabe H., Benning C. and Nishida I. (2007) Digalactosyldiacylglycerol is required for better photosynthetic growth of Synechocystis sp PCC6803 under phosphate limitation. Plant Cell Physiol. 48, 1517–1523 10.1093/pcp/pcm134 [Abstract] [CrossRef] [Google Scholar]
135. Guler S., Seeliger A., Hartel H., Renger G. and Benning C. (1996) A null mutant of Synechococcus sp. PCC7942 deficient in the sulfolipid sulfoquinovosyl diacylglycerol. J. Biol. Chem. 271, 7501–7507 10.1074/jbc.271.13.7501 [Abstract] [CrossRef] [Google Scholar]
136. Aoki M., Sato N., Meguro A. and Tsuzuki M. (2004) Differing involvement of sulfoquinovosyl diacylglycerol in photosystem II in two species of unicellular cyanobacteria. Eur. J. Biochem. 271, 685–693 10.1111/j.1432-1033.2003.03970.x [Abstract] [CrossRef] [Google Scholar]
137. Guler S., Essigmann B. and Benning C. (2000) A cyanobacterial gene, sqdX, required for biosynthesis of the sulfolipid sulfoquinovosyldiacylglycerol. J. Bacteriol. 182, 543–545 10.1128/JB.182.2.543-545.2000 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
138. Beld J., Abbriano R., Finzel K., Hildebrand M. and Burkart M.D. (2016) Probing fatty acid metabolism in bacteria, cyanobacteria, green microalgae and diatoms with natural and unnatural fatty acids. Mol. Biosyst. 12, 1299–1312 10.1039/C5MB00804B [Abstract] [CrossRef] [Google Scholar]
139. Cronan J.E. (2016) Assembly of Lipoic Acid on Its Cognate Enzymes: an Extraordinary and Essential Biosynthetic Pathway. Microbiol. Mol. Biol. Rev. 80, 429–450 10.1128/MMBR.00073-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
140. Cronan J.E., Zhao X. and Jiang Y. (2005) Function, attachment and synthesis of lipoic acid in Escherichia coli. Adv. Microb. Physiol. 50, 103–146 10.1016/S0065-2911(05)50003-1 [Abstract] [CrossRef] [Google Scholar]
141. Sham L.T. et al. . (2014) MurJ is the flippase of lipid-linked precursors for peptidoglycan biogenesis. Science 345, 220–222 10.1126/science.1254522 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
142. Taguchi A. et al. . (2019) FtsW is a peptidoglycan polymerase that is functional only in complex with its cognate penicillin-binding protein. Nat. Microbiol. 4, 587–594 10.1038/s41564-018-0345-x [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
143. Marbouty M., Mazouni K., Saguez C., Cassier-Chauvat C. and Chauvat F. (2009) Characterization of the Synechocystis strain PCC 6803 penicillin-binding proteins and cytokinetic proteins FtsQ and FtsW and their network of interactions with ZipN. J. Bacteriol. 191, 5123–5133 10.1128/JB.00620-09 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
144. Reith J. and Mayer C. (2011) Peptidoglycan turnover and recycling in Gram-positive bacteria. Appl. Microbiol. Biotechnol. 92, 1–11 10.1007/s00253-011-3486-x [Abstract] [CrossRef] [Google Scholar]
145. Ruiz N., Kahne D. and Silhavy T.J. (2009) Transport of lipopolysaccharide across the cell envelope: the long road of discovery. Nat. Rev. Microbiol. 7, 677–683 10.1038/nrmicro2184 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
146. Kalynych S., Morona R. and Cygler M. (2014) Progress in understanding the assembly process of bacterial O-antigen. FEMS Microbiol. Rev. 38, 1048–1065 10.1111/1574-6976.12070 [Abstract] [CrossRef] [Google Scholar]
147. Englund E., Shabestary K., Hudson E.P. and Lindberg P. (2018) Systematic overexpression study to find target enzymes enhancing production of terpenes in Synechocystis PCC 6803, using isoprene as a model compound. Metab. Eng. 49, 164–177 10.1016/j.ymben.2018.07.004 [Abstract] [CrossRef] [Google Scholar]
148. Barkley S.J., Desai S.B. and Poulter C.D. (2004) Type II isopentenyl diphosphate isomerase from Synechocystis sp strain PCC 6803. J. Bacteriol. 186, 8156–8158 10.1128/JB.186.23.8156-8158.2004 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
149. Poliquin K. et al. . (2004) Inactivation of sll1556 in Synechocystis strain PCC 6803 impairs isoprenoid biosynthesis from pentose phosphate cycle substrates in vitro. J. Bacteriol. 186, 4685–4693 10.1128/JB.186.14.4685-4693.2004 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
150. Englund E. et al. . (2014) Production of squalene in Synechocystis sp. PCC 6803. PLoS ONE 9, e90270–e90270 10.1371/journal.pone.0090270 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
151. Welander P.V. et al. . (2009) Hopanoids play a role in membrane integrity and pH homeostasis in Rhodopseudomonas palustris TIE-1. J. Bacteriol. 191, 6145–6156 10.1128/JB.00460-09 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
152. Jurgens U.J., Simonin P. and Rohmer M. (1992) Localization and distribution of hopanoids in membrane systems of the cyanobacterium Synechocystis PCC 6714. FEMS Microbiol. Lett. 71, 285–288 10.1016/0378-1097(92)90723-2 [Abstract] [CrossRef] [Google Scholar]
153. Toth T.N. et al. . (2015) Carotenoids are essential for the assembly of cyanobacterial photosynthetic complexes. Biochim. Biophys. Acta 1847, 1153–1165 10.1016/j.bbabio.2015.05.020 [Abstract] [CrossRef] [Google Scholar]
154. Mohamed H.E., Vermaas W. and Myxoxanthophyll R. (2004) Slr1293 in Synechocystis sp. strain PCC 6803 Is the C-3′,4′ desaturase (CrtD) involved in myxoxanthophyll biosynthesis. J. Bacteriol. 186, 5621–5628 10.1128/JB.186.17.5621-5628.2004 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
155. Graham J.E. and Bryant D.A. (2008) The Biosynthetic Pathway for Synechoxanthin, an Aromatic Carotenoid Synthesized by the Euryhaline, Unicellular Cyanobacterium Synechococcus sp Strain PCC 7002. J. Bacteriol. 190, 7966–7974 10.1128/JB.00985-08 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
156. Maresca J.A., Graham J.E., Wu M., Eisen J.A. and Bryant D.A. (2007) Identification of a fourth family of lycopene cyclases in photosynthetic bacteria. PNAS 104, 11784–11789 10.1073/pnas.0702984104 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
157. Graham J.E. and Bryant D.A. (2009) The biosynthetic pathway for myxol-2′ fucoside (myxoxanthophyll) in the cyanobacterium Synechococcus sp. strain PCC 7002. J. Bacteriol. 191, 3292–3300 10.1128/JB.00050-09 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
158. Masamoto K., Wada H., Kaneko T. and Takaichi S. (2001) Identification of a gene required for cis-to-trans carotene isomerization in carotenogenesis of the cyanobacterium Synechocystis sp. PCC 6803. Plant Cell Physiol. 42, 1398–1402 10.1093/pcp/pce167 [Abstract] [CrossRef] [Google Scholar]
159. Maeda H., Sakuragi Y., Bryant D.A. and DellaPenna D. (2005) Tocopherols protect Synechocystis sp strain PCC 6803 from lipid peroxidation. Plant Physiol. 138, 1422–1435 10.1104/pp.105.061135 [Abstract] [CrossRef] [Google Scholar]
160. Yang Y., Yin C.T., Li W.Z. and Xu X.D. (2008) alpha-tocopherol is essential for acquired chill-light tolerance in the cyanobacterium Synechocystis sp strain PCC 6803. J. Bacteriol. 190, 1554–1560 10.1128/JB.01577-07 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
161. Sakuragi Y., Maeda H., DellaPenna D. and Bryant D.A. (2006) alpha-tocopherol plays a role in photosynthesis and macronutrient homeostasis of the cyanobacterium Synechocystis sp PCC 6803 that is independent of its antioxidant function. Plant Physiol. 141, 508–521 10.1104/pp.105.074765 [Abstract] [CrossRef] [Google Scholar]
162. Collakova E. and DellaPenna D. (2001) Isolation and functional analysis of homogentisate phytyltransferase from Synechocystis sp PCC 6803 and arabidopsis. Plant Physiol. 127, 1113–1124 10.1104/pp.010421 [Abstract] [CrossRef] [Google Scholar]
163. Savidge B. et al. . (2002) Isolation and characterization of homogentisate phytyltransferase genes from Synechocystis sp PCC 6803 and Arabidopsis. Plant Physiol. 129, 321–332 10.1104/pp.010747 [Abstract] [CrossRef] [Google Scholar]
164. Schledz M., Seidler A., Beyer P. and Neuhaus G. (2001) A novel phytyltransferase from Synechocystis sp PCC 6803 involved in tocopherol biosynthesis. FEBS Lett. 499, 15–20 10.1016/S0014-5793(01)02508-X [Abstract] [CrossRef] [Google Scholar]
165. Shpilyov A.V., Zinchenko V.V., Shestakov S.V., Grimm B. and Lokstein H. (2005) Inactivation of the geranylgeranyl reductase (ChlP) gene in the cyanobacterium Synechocystis sp. PCC 6803. Biochim. Biophys. Acta 1706, 195–203 10.1016/j.bbabio.2004.11.001 [Abstract] [CrossRef] [Google Scholar]
166. Dahnhardt D. et al. . (2002) The hydroxyphenylpyruvate dioxygenase from Synechocystis sp. PCC 6803 is not required for plastoquinone biosynthesis. FEBS Lett. 523, 177–181 10.1016/S0014-5793(02)02978-2 [Abstract] [CrossRef] [Google Scholar]
167. Bonner C.A., Jensen R.A., Gander J.E. and Keyhani N.O. (2004) A core catalytic domain of the TyrA protein family: arogenate dehydrogenase from Synechocystis. Biochem. J. 382, 279–291 10.1042/BJ20031809 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
168. Saw J. H. W. et al. . (2013) Cultivation and Complete Genome Sequencing of Gloeobacter kilaueensis sp. nov., from a Lava Cave in Kīlauea Caldera. PlosONE 8, e76376 [Europe PMC free article] [Abstract] [Google Scholar]
169. Shintani D.K., Cheng Z.G. and DellaPenna D. (2002) The role of 2-methyl-6-phytylbenzoquinone methyltransferase in determining tocopherol composition in Synechocystis sp. PCC6803. FEBS Lett. 511, 1–5 10.1016/S0014-5793(01)03223-9 [Abstract] [CrossRef] [Google Scholar]
170. Kozuleva M.A., Petrova A.A., Mamedov M.D., Semenov A.Y. and Ivanov B.N. (2014) O2 reduction by photosystem I involves phylloquinone under steady-state illumination. FEBS Lett. 588, 4364–4368 10.1016/j.febslet.2014.10.003 [Abstract] [CrossRef] [Google Scholar]
171. Johnson T.W. et al. . (2000) Recruitment of a foreign quinone into the A(1) site of photosystem I - I. Genetic and physiological characterization of phylloquinone biosynthetic pathway mutants in Synechocystis sp PCC 6803. J. Biol. Chem. 275, 8523–8530 10.1074/jbc.275.12.8523 [Abstract] [CrossRef] [Google Scholar]
172. Johnson T.W. et al. . (2003) The menD and menE homologs code for 2-succinyl-6-hydroxyl-2, 4-cyclohexadiene-1-carboxylate synthase and O-succinylbenzoic acid-CoA synthase in the phylloquinone biosynthetic pathway of Synechocystis sp PCC 6803. Biochim. Biophys. Acta-Bioenergetics 1557, 67–76 10.1016/S0005-2728(02)00396-1 [Abstract] [CrossRef] [Google Scholar]
173. Emonds-Alt B., Coosemans N., Gerards T., Remacle C. and Cardol P. (2017) Isolation and characterization of mutants corresponding to the MENA, MENB, MENC and MENE enzymatic steps of 5′-monohydroxyphylloquinone biosynthesis in Chlamydomonas reinhardtii. Plant J. 89, 141–154 10.1111/tpj.13352 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
174. Fatihi A. et al. . (2015) A dedicated Type II NADPH Dehydrogenase performs the penultimate step in the biosynthesis of Vitamin K1 in Synechocystis and Arabidopsis. Plant Cell 27, 1730–41 10.1105/tpc.15.00103 [Abstract] [CrossRef] [Google Scholar]
175. Pfaff C., Glindemann N., Gruber J., Frentzen M. and Sadre R. (2014) Chorismate pyruvate-lyase and 4-hydroxy-3-solanesylbenzoate decarboxylase are required for plastoquinone biosynthesis in the cyanobacterium Synechocystis sp. PCC6803. J. Biol. Chem. 289, 2675–2686 10.1074/jbc.M113.511709 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
176. Sadre R., Pfaff C. and Buchkremer S. (2012) Plastoquinone-9 biosynthesis in cyanobacteria differs from that in plants and involves a novel 4-hydroxybenzoate solanesyltransferase. Biochem. J. 442, 621–629 10.1042/BJ20111796 [Abstract] [CrossRef] [Google Scholar]
177. Anzaldi L.L. and Skaar E.P. (2010) Overcoming the heme paradox: heme toxicity and tolerance in bacterial pathogens. Infect. Immun. 78, 4977–4989 10.1128/IAI.00613-10 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
178. Skotnicova P. et al. . (2018) The cyanobacterial protoporphyrinogen oxidase HemJ is a new b-type heme protein functionally coupled with coproporphyrinogen III oxidase. J. Biol. Chem. 293, 12394–12404 10.1074/jbc.RA118.003441 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
179. Brzezowski P. et al. . (2019) The function of PROTOPORPHYRINOGEN IX OXIDASE in chlorophyll biosynthesis requires oxidised plastoquinone in Chlamydomonas reinhardtii. Commun. Biol. 2, 159 10.1038/s42003-019-0395-5 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
180. Goto T., Aoki R., Minamizaki K. and Fujita Y. (2010) Functional differentiation of two analogous coproporphyrinogen III oxidases for heme and chlorophyll biosynthesis pathways in the cyanobacterium Synechocystis sp. PCC 6803. Plant Cell Physiol. 51, 650–663 10.1093/pcp/pcq023 [Abstract] [CrossRef] [Google Scholar]
181. Yilmaz M., Kang I. and Beale S.I. (2010) Heme oxygenase 2 of the cyanobacterium Synechocystis sp PCC 6803 is induced under a microaerobic atmosphere and is required for microaerobic growth at high light intensity. Photosynth. Res. 103, 47–59 10.1007/s11120-009-9506-3 [Abstract] [CrossRef] [Google Scholar]
182. Aoki R., Goto T. and Fujita Y. (2011) A Heme Oxygenase Isoform is Essential for Aerobic Growth in the Cyanobacterium Synechocystis sp PCC 6803: Modes of Differential Operation of Two Isoforms/Enzymes to Adapt to Low Oxygen Environments in Cyanobacteria. Plant Cell Physiol. 52, 1744–1756 10.1093/pcp/pcr108 [Abstract] [CrossRef] [Google Scholar]
183. Peter E. et al. . (2009) Differential requirement of two homologous proteins encoded by sll1214 and sll1874 for the reaction of Mg protoporphyrin monomethylester oxidative cyclase under aerobic and micro-oxic growth conditions. Biochim. Biophys. Acta 1787, 1458–1467 10.1016/j.bbabio.2009.06.006 [Abstract] [CrossRef] [Google Scholar]
184. Frankenberg N., Mukougawa K., Kohchi T. and Lagarias J.C. (2001) Functional genomic analysis of the HY2 family of ferredoxin-dependent bilin reductases from oxygenic photosynthetic organisms. Plant Cell 13, 965–978 10.1105/tpc.13.4.965 [Abstract] [CrossRef] [Google Scholar]
185. Schluchter W.M. and Glazer A.N. (1997) Characterization of cyanobacterial biliverdin reductase. Conversion of biliverdin to bilirubin is important for normal phycobiliprotein biosynthesis. J. Biol. Chem. 272, 13562–13569 10.1074/jbc.272.21.13562 [Abstract] [CrossRef] [Google Scholar]
186. Jensen P.E., Gibson L.C.D., Henningsen K.W. and Hunter C.N. (1996) Expression of the chlI, chlD, and chlH genes from the cyanobacterium Synechocystis PCC6803 in Escherichia coli and demonstration that the three cognate proteins are required for magnesium-protoporphyrin chelatase activity. J. Biol. Chem. 271, 16662–16667 10.1074/jbc.271.28.16662 [Abstract] [CrossRef] [Google Scholar]
187. Wilde A., Mikolajczyk S., Alawady A., Lokstein H. and Grimm B. (2004) The gun4 gene is essential for cyanobacterial porphyrin metabolism. FEBS Lett. 571, 119–123 10.1016/j.febslet.2004.06.063 [Abstract] [CrossRef] [Google Scholar]
188. Davison P.A. et al. . (2005) Structural and biochemical characterization of Gun4 suggests a mechanism for its role in chlorophyll biosynthesis. Biochemistry 44, 7603–7612 10.1021/bi050240x [Abstract] [CrossRef] [Google Scholar]
189. Sobotka R. et al. . (2008) Importance of the cyanobacterial Gun4 protein for chlorophyll metabolism and assembly of photosynthetic complexes. J. Biol. Chem. 283, 25794–25802 10.1074/jbc.M803787200 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
190. Smith C.A., Suzuki J.Y. and Bauer C.E. (1996) Cloning and characterization of the chlorophyll biosynthesis gene chlM from Synechocystis PCC 6803 by complementation of a bacteriochlorophyll biosynthesis mutant of Rhodobacter capsulatus. Plant Mol. Biol. 30, 1307–1314 10.1007/BF00019561 [Abstract] [CrossRef] [Google Scholar]
191. Minamizaki K., Mizoguchi T., Goto T., Tamiaki H. and Fujita Y. (2008) Identification of two homologous genes, chlAI and chlAII, that are differentially involved in isocyclic ring formation of chlorophyll a in the cyanobacterium Synechocystis sp. PCC 6803. J. Biol. Chem. 283, 2684–2692 10.1074/jbc.M708954200 [Abstract] [CrossRef] [Google Scholar]
192. Chen G.E., Canniffe D.P. and Hunter C.N. (2017) Three classes of oxygen-dependent cyclase involved in chlorophyll and bacteriochlorophyll biosynthesis. PNAS 114, 6280–6285 10.1073/pnas.1701687114 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
193. Reinbothe C. et al. . (2010) Chlorophyll biosynthesis: spotlight on protochlorophyllide reduction. Trends Plant Sci. 15, 614–624 10.1016/j.tplants.2010.07.002 [Abstract] [CrossRef] [Google Scholar]
194. Islam M.R. et al. ., 2008) slr1923 of Synechocystis sp PCC6803 is essential for conversion of 3,8-divinyl(proto)chlorophyll(ide) to 3-monovinyl(proto)chlorophyll(ide). Plant Physiol. 148, 1068–1081 10.1104/pp.108.123117 [Abstract] [CrossRef] [Google Scholar]
195. Ito H., Yokono M., Tanaka R. and Tanaka A. (2008) Identification of a novel vinyl reductase gene essential for the biosynthesis of monovinyl chlorophyll in Synechocystis sp PCC6803. J. Biol. Chem. 283, 9002–9011 10.1074/jbc.M708369200 [Abstract] [CrossRef] [Google Scholar]
196. Chen G.E. et al. . (2018) Complete enzyme set for chlorophyll biosynthesis in Escherichia coli. Sci. Adv. 4, eaaq1407 10.1126/sciadv.aaq1407 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
197. Romine M.F. et al. . (2017) Elucidation of roles for vitamin B12 in regulation of folate, ubiquinone, and methionine metabolism. Proc. Natl. Acad. Sci. U.S.A. 114, E1205–E1214 10.1073/pnas.1612360114 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
198. Helliwell K.E. et al. . (2016) Cyanobacteria and Eukaryotic Algae Use Different Chemical Variants of Vitamin B12. Curr. Biol. 26, 999–1008 10.1016/j.cub.2016.02.041 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
199. Fang H., Kang J. and Zhang D. (2017) Microbial production of vitamin B12: a review and future perspectives. Microb. Cell Fact. 16, 15 10.1186/s12934-017-0631-y [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
200. Moore S.J. et al. . (2013) Elucidation of the anaerobic pathway for the corrin component of cobalamin (vitamin B12). Proc. Natl. Acad. Sci. U.S.A. 110, 14906–14911 10.1073/pnas.1308098110 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
201. Flores E., Frias J.E., Rubio L.M. and Herrero A. (2005) Photosynthetic nitrate assimilation in cyanobacteria. Photosynth. Res. 83, 117–133 10.1007/s11120-004-5830-9 [Abstract] [CrossRef] [Google Scholar]
202. Montesinos M.L., Muro-Pastor A.M., Herrero A. and Flores E. (1998) Ammonium/Methylammonium permeases of a cyanobacterium - Identification and analysis of three nitrogen-regulated amt genes in Synechocystis sp. PCC 6803. J. Biol. Chem. 273, 31463–31470 10.1074/jbc.273.47.31463 [Abstract] [CrossRef] [Google Scholar]
203. Omata T., Andriesse X. and Hirano A. (1993) Identification and characterization of a gene cluster involved in nitrate transport in the cyanobacterium Synechococcus sp. PCC7942. Mol. Gen. Genet. 236, 193–202 10.1007/BF00277112 [Abstract] [CrossRef] [Google Scholar]
204. Maeda S. and Omata T. (1997) Substrate-binding lipoprotein of the cyanobacterium Synechococcus sp. strain PCC 7942 involved in the transport of nitrate and nitrite. J. Biol. Chem. 272, 3036–3041 10.1074/jbc.272.5.3036 [Abstract] [CrossRef] [Google Scholar]
205. Luque I., Flores E. and Herrero A. (1994) Nitrate and nitrite transport in the cyanobacterium Synechococcus sp. PCC 7942 are mediated by the same permease. Biochem. Biophys. Acta- Bioenergetics 1184, 296–298 10.1016/0005-2728(94)90236-4 [CrossRef] [Google Scholar]
206. Rubio L.M., Herrero A. and Flores E. (1996) A cyanobacterial narB gene encodes a ferredoxin-dependent nitrate reductase. Plant Mol. Biol. 30, 845–850 10.1007/BF00019017 [Abstract] [CrossRef] [Google Scholar]
207. Suzuki I., Sugiyama T. and Omata T. (1993) Primary Structure and Transcriptional Regulation of the Gene for Nitrite Reductase from the Cyanobacterium Synechococcus PCC 7942. Plant Cell Physiol. 34, 1311–1320 [Google Scholar]
208. Valladares A., Montesinos M.L., Herrero A. and Flores E. (2002) An ABC-type, high-affinity urea permease identified in cyanobacteria. Mol. Microbiol. 43, 703–715 10.1046/j.1365-2958.2002.02778.x [Abstract] [CrossRef] [Google Scholar]
209. Veaudor T., Cassier-Chauvat C. and Chauvat F. (2019) Genomics of Urea Transport and Catabolism in Cyanobacteria: Biotechnological Implications. Front. Microbiol. 10, 2052 10.3389/fmicb.2019.02052 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
210. Quintero M.J., Montesinos M.L., Herrero A. and Flores E. (2001) Identification of genes encoding amino acid permeases by inactivation of selected ORFs from the Synechocystis genomic sequence. Genome Res. 11, 2034–2040 10.1101/gr.196301 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
211. Suginaka K. et al. . (1998) Cysteine Uptake for Accumulation of Glutathione by the Cyanobacterium Synechocystis strain PCC 6803. Biosci. Biotechnol. Biochem. 62, 424–428 10.1271/bbb.62.424 [Abstract] [CrossRef] [Google Scholar]
212. Tottey S., Rich P.R., Rondet S.A.M. and Robinson N.J. (2001) Two Menkes-type ATPases supply copper for photosynthesis in Synechocystis PCC 6803. J. Biol. Chem. 276, 19999–20004 10.1074/jbc.M011243200 [Abstract] [CrossRef] [Google Scholar]
213. Badarau A. and Dennison C. (2011) Thermodynamics of copper and zinc distribution in the cyanobacterium Synechocystis PCC 6803. PNAS 108, 13007–13012 10.1073/pnas.1101448108 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
214. Tottey S. et al. . (2012) Cyanobacterial metallochaperone inhibits deleterious side reactions of copper. PNAS 109, 95–100 10.1073/pnas.1117515109 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
215. Giner-Lamia J. et al. . (2012) The CopRS Two-Component System Is Responsible for Resistance to Copper in the Cyanobacterium Synechocystis sp PCC 6803. Plant Physiol. 159, 1806–1818 10.1104/pp.112.200659 [Abstract] [CrossRef] [Google Scholar]
216. Giner-Lamia J., Lopez-Maury L. and Florencio F.J. (2015) CopM is a novel copper-binding protein involved in copper resistance in Synechocystis sp PCC 6803. Microbiologyopen 4, 167–185 10.1002/mbo3.231 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
217. Matsuda N. and Uozumi N. (2006) Ktr-mediated potassium transport, a major pathway for potassium uptake, is coupled to a proton gradient across the membrane in Synechocystis sp. PCC 6803. Biosci. Biotechnol. Biochem. 70, 273–275 10.1271/bbb.70.273 [Abstract] [CrossRef] [Google Scholar]
218. Nanatani K. et al. . (2015) Comparative analysis of kdp and ktr mutants reveals distinct roles of the potassium transporters in the model cyanobacterium Synechocystis sp. strain PCC 6803. J. Bacteriol. 197, 676–687 10.1128/JB.02276-14 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
219. Matsuda N. et al. . (2004) Na+-dependent K+ uptake Ktr system from the cyanobacterium Synechocystis sp. PCC 6803 and its role in the early phases of cell adaptation to hyperosmotic shock. J. Biol. Chem. 279, 54952–54962 10.1074/jbc.M407268200 [Abstract] [CrossRef] [Google Scholar]
220. Zanetti M. et al. . (2010) A Novel Potassium Channel in Photosynthetic Cyanobacteria. PLoS ONE 5, e10118 10.1371/journal.pone.0010118 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
221. Checchetto V. et al. . (2012) Thylakoid potassium channel is required for efficient photosynthesis in cyanobacteria. PNAS 109, 11043–11048 10.1073/pnas.1205960109 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
222. Checchetto V. et al. . (2013) Functional characterization and determination of the physiological role of a calcium-dependent potassium channel from cyanobacteria. Plant Physiol. 162, 953–964 10.1104/pp.113.215129 [Abstract] [CrossRef] [Google Scholar]
223. Waditee R. et al. . (2004) Isolation and functional characterization of Ca2+/H+ antiporters from cyanobacteria. J. Biol. Chem. 279, 4330–4338 10.1074/jbc.M310282200 [Abstract] [CrossRef] [Google Scholar]
224. Jiang H.B., Cheng H.M., Gao K.S. and Qiu B.S. (2013) Inactivation of Ca(2+)/H(+) exchanger in Synechocystis sp. strain PCC 6803 promotes cyanobacterial calcification by upregulating CO(2)-concentrating mechanisms. Appl. Environ. Microbiol. 79, 4048–4055 10.1128/AEM.00681-13 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
225. Nazarenko L.V., Andreev I.M., Lyukevich A.A., Pisareva T.V. and Los D.A. (2003) Calcium release from Synechocystis cells induced by depolarization of the plasma membrane: MscL as an outward Ca2+ channel. Microbiology-Sgm 149, 1147–1153 10.1099/mic.0.26074-0 [Abstract] [CrossRef] [Google Scholar]
226. Katoh H., Hagino N., Grossman A.R. and Ogawa T. (2001) Genes essential to iron transport in the cyanobacterium Synechocystis sp. strain PCC 6803. J. Bacteriol. 183, 2779–2784 10.1128/JB.183.9.2779-2784.2001 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
227. Katoh H., Grossman A.R., Hagino N. and Ogawa T. (2000) A gene of Synechocystis sp strain PCC 6803 encoding a novel iron transporter. J. Bacteriol. 182, 6523–6524 10.1128/JB.182.22.6523-6524.2000 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
228. Kranzler C. et al. . (2014) Coordinated transporter activity shapes high-affinity iron acquisition in cyanobacteria. Isme J. 8, 409–417 10.1038/ismej.2013.161 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
229. Katoh H., Hagino N. and Ogawa T. (2001) Iron-binding activity of FutA1 subunit of an ABC-type iron transporter in the cyanobacterium Synechocystis sp strain PCC 6803. Plant Cell Physiol. 42, 823–827 10.1093/pcp/pce106 [Abstract] [CrossRef] [Google Scholar]
230. Badarau A. et al. . (2008) FutA2 is a ferric binding protein from Synechocystis PCC 6803. J. Biol. Chem. 283, 12520–12527 10.1074/jbc.M709907200 [Abstract] [CrossRef] [Google Scholar]
231. Jiang H.B. et al. . (2014) New insights into iron acquisition by cyanobacteria: an essential role for ExbB-ExbD complex in inorganic iron uptake. ISME J., 9, 297–309 [Europe PMC free article] [Abstract] [Google Scholar]
232. Keren N., Aurora R. and Pakrasi H.B. (2004) Critical roles of bacterioferritins in iron storage and proliferation of cyanobacteria. Plant Physiol. 135, 1666–1673 10.1104/pp.104.042770 [Abstract] [CrossRef] [Google Scholar]
233. Bartsevich V.V. and Pakrasi H.B. (1996) Manganese transport in the cyanobacterium Synechocystis sp. PCC 6803. J. Biol. Chem. 271, 26057–26061 10.1074/jbc.271.42.26057 [Abstract] [CrossRef] [Google Scholar]
234. Brandenburg F. et al. . (2017) The Synechocystis Manganese Exporter Mnx Is Essential for Manganese Homeostasis in Cyanobacteria. Plant Physiol. 173, 1798–1810 10.1104/pp.16.01895 [Abstract] [CrossRef] [Google Scholar]
235. Aguilar-Barajas E., Diaz-Perez C., Ramirez-Diaz M.I., Riveros-Rosas H. and Cervantes C. (2011) Bacterial transport of sulfate, molybdate, and related oxyanions. Biometals 24, 687–707 10.1007/s10534-011-9421-x [Abstract] [CrossRef] [Google Scholar]
236. Banerjee S., Wei B.X., Bhattacharyya-Pakrasi M., Pakrasi H.B. and Smith T.J. (2003) Structural determinants of metal specificity in the zinc transport protein ZnuA from Synechocystis 6803. J. Mol. Biol. 333, 1061–1069 10.1016/j.jmb.2003.09.008 [Abstract] [CrossRef] [Google Scholar]
237. Borrelly G.P., Rondet S.A., Tottey S. and Robinson N.J. (2004) Chimeras of P-type ATPases and their transcriptional regulators: contributions of a cytosolic amino-terminal domain to metal specificity. Mol. Microbiol. 53, 217–227 10.1111/j.1365-2958.2004.04106.x [Abstract] [CrossRef] [Google Scholar]
238. Dainty S.J., Patterson C.J., Waldron K.J. and Robinson N.J. (2010) Interaction between cyanobacterial copper chaperone Atx1 and zinc homeostasis. J. Biol. Inorg. Chem. 15, 77–85 10.1007/s00775-009-0555-z [Abstract] [CrossRef] [Google Scholar]
239. Pohland A.C. and Schneider D. (2019) Mg2+ homeostasis and transport in cyanobacteria - at the crossroads of bacterial and chloroplast Mg2+ import. Biol. Chem. 400, 1289–1301 10.1515/hsz-2018-0476 [Abstract] [CrossRef] [Google Scholar]
240. Peca L., Kos P.B. and Vass I. (2007) Characterization of the activity of heavy metal-responsive promoters in the cyanobacterium Synechocystis PCC 6803. Acta Biol. Hung. 58, 11–22 10.1556/ABiol.58.2007.Suppl.2 [Abstract] [CrossRef] [Google Scholar]
241. Burut-Archanai S., Eaton-Rye J.J. and Incharoensakdi A. (2011) Na+-stimulated phosphate uptake system in Synechocystis sp PCC 6803 with Pst1 as a main transporter. BMC Microbiol. 11, 225 10.1186/1471-2180-11-225 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
242. Pitt F.D., Mazard S., Humphreys L. and Scanlan D.J. (2010) Functional Characterization of Synechocystis sp Strain PCC 6803 pst1 and pst2 Gene Clusters Reveals a Novel Strategy for Phosphate Uptake in a Freshwater Cyanobacterium. J. Bacteriol. 192, 3512–3523 10.1128/JB.00258-10 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
243. Gomez-Garcia M.R., Losada M. and Serrano A. (2003) Concurrent transcriptional activation of ppa and ppx genes by phosphate deprivation in the cyanobacterium Synechocystis sp strain PCC 6803. Biochem. Biophys. Res. Commun. 302, 601–609 10.1016/S0006-291X(03)00162-1 [Abstract] [CrossRef] [Google Scholar]
244. Zhang H.Y., Ishige K. and Kornberg A. (2002) A polyphosphate kinase (PPK2) widely conserved in bacteria. PNAS 99, 16678–16683 10.1073/pnas.262655199 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
245. Inaba M., Sakamoto A. and Murata N. (2001) Functional expression in Escherichia coli of low-affinity and high-affinity Na(+)(Li(+))/H(+) antiporters of Synechocystis. J. Bacteriol. 183, 1376–1384 10.1128/JB.183.4.1376-1384.2001 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
246. Elanskaya I.V., Karandashova I.V., Bogachev A.V. and Hagemann M. (2002) Functional analysis of the Na+/H+ antiporter encoding genes of the cyanobacterium Synechocystis PCC 6803. Biochemistry (Mosc) 67, 432–440 10.1023/A:1015281906254 [Abstract] [CrossRef] [Google Scholar]
247. Tsunekawa K. et al. . (2009) Identification and characterization of the Na+/H+ antiporter Nhas3 from the thylakoid membrane of Synechocystis sp. PCC 6803. J. Biol. Chem. 284, 16513–16521 10.1074/jbc.M109.001875 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
248. Hagemann M., Richter S. and Mikkat S. (1997) The ggtA gene encodes a subunit of the transport system for the osmoprotective compound glucosylglycerol in Synechocystis sp. strain PCC 6803. J. Bacteriol. 179, 714–720 10.1128/JB.179.3.714-720.1997 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
249. Mikkat S. and Hagemann M. (2000) Molecular analysis of the ggtBCD gene cluster of Synechocystis sp strain PCC6803 encoding subunits of an ABC transporter for osmoprotective compounds. Arch. Microbiol. 174, 273–282 10.1007/s002030000201 [Abstract] [CrossRef] [Google Scholar]
250. Omata T. et al. . (1999) Identification of an ATP-binding cassette transporter involved in bicarbonate uptake in the cyanobacterium Synechococcus sp strain PCC 7942. PNAS 96, 13571–13576 10.1073/pnas.96.23.13571 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
251. Maeda S., Price G.D., Badger M.R., Enomoto C. and Omata T. (2000) Bicarbonate binding activity of the CmpA protein of the cyanobacterium Synechococcus sp strain PCC 7942 involved in active transport of bicarbonate. J. Biol. Chem. 275, 20551–20555 10.1074/jbc.M003034200 [Abstract] [CrossRef] [Google Scholar]
252. Du J., Forster B., Rourke L., Howitt S.M. and Price G.D. (2014) Characterisation of cyanobacterial bicarbonate transporters in E. coli shows that SbtA homologs are functional in this heterologous expression system. PLoS ONE 9, e115905 10.1371/journal.pone.0115905 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
253. Shibata M. et al. . (2002) Genes essential to sodium-dependent bicarbonate transport in cyanobacteria - Function and phylogenetic analysis. J. Biol. Chem. 277, 18658–18664 10.1074/jbc.M112468200 [Abstract] [CrossRef] [Google Scholar]
254. Price G.D., Woodger F.J., Badger M.R., Howitt S.M. and Tucker L. (2004) Identification of a SulP-type bicarbonate transporter in marine cyanobacteria. PNAS 101, 18228–18233 10.1073/pnas.0405211101 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
255. Akai M. et al. . (2012) Aquaporin AqpZ Is Involved in Cell Volume Regulation and Sensitivity to Osmotic Stress in Synechocystis sp Strain PCC 6803. J. Bacteriol. 194, 6828–6836 10.1128/JB.01665-12 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
256. Akai M. et al. . (2011) Plasma Membrane Aquaporin AqpZ Protein Is Essential for Glucose Metabolism during Photomixotrophic Growth of Synechocystis sp PCC 6803. J. Biol. Chem. 286, 25224–25235 10.1074/jbc.M111.236380 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
257. Gale G.A.R. et al. . (2019) Emerging Species and Genome Editing Tools: Future Prospects in Cyanobacterial Synthetic Biology. Microorganisms 7, E409 10.3390/microorganisms7100409 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
258. Lea-Smith D.J., Vasudevan R. and Howe C.J. (2016) Generation of marked and markerless mutants in model cyanobacterial species. J. Vis. Exp. 111, 10.3791/54001 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
259. Vasudevan R. et al. . (2019) CyanoGate: A Modular Cloning Suite for Engineering Cyanobacteria Based on the Plant MoClo Syntax. Plant Physiol. 180, 39–55 10.1104/pp.18.01401 [Abstract] [CrossRef] [Google Scholar]
260. Dubeau M.P. et al. . (2009) Cytosine Deaminase as a Negative Selection Marker for Gene Disruption and Replacement in the Genus Streptomyces and Other Actinobacteria. Appl. Environ. Microbiol. 75, 1211–1214 10.1128/AEM.02139-08 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
261. Rubin B.E. et al. . (2015) The essential gene set of a photosynthetic organism. PNAS 112, E6634–E6643 10.1073/pnas.1519220112 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
262. Wlodarczyk A., Selao T., Norling B. and Nixon P. (2019) Unprecedented biomass and fatty acid production by the newly discovered cyanobacterium Synechococcus sp. PCC 11901. BioRxiv 684944 [Google Scholar]

Articles from Bioscience Reports are provided here courtesy of Portland Press Ltd

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/92487209
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/92487209

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1042/bsr20193325

Supporting
Mentioning
Contrasting
2
48
0

Article citations


Go to all (26) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.


Funding 


Funders who supported this work.

Biotechnology and Biological Sciences Research Council (3)