* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download Probing the origins of glutathione biosynthesis through biochemical
Expression vector wikipedia , lookup
Ultrasensitivity wikipedia , lookup
Clinical neurochemistry wikipedia , lookup
Restriction enzyme wikipedia , lookup
Proteolysis wikipedia , lookup
Photosynthetic reaction centre wikipedia , lookup
Light-dependent reactions wikipedia , lookup
Metabolic network modelling wikipedia , lookup
Two-hybrid screening wikipedia , lookup
Deoxyribozyme wikipedia , lookup
Microbial metabolism wikipedia , lookup
Ribosomally synthesized and post-translationally modified peptides wikipedia , lookup
NADH:ubiquinone oxidoreductase (H+-translocating) wikipedia , lookup
Photosynthesis wikipedia , lookup
Biochemistry wikipedia , lookup
Cyanobacteria wikipedia , lookup
Citric acid cycle wikipedia , lookup
Enzyme inhibitor wikipedia , lookup
Metalloprotein wikipedia , lookup
Catalytic triad wikipedia , lookup
Adenosine triphosphate wikipedia , lookup
Oxidative phosphorylation wikipedia , lookup
Amino acid synthesis wikipedia , lookup
Biosynthesis wikipedia , lookup
Evolution of metal ions in biological systems wikipedia , lookup
Biochem. J. (2013) 450, 63–72 (Printed in Great Britain) 63 doi:10.1042/BJ20121332 Probing the origins of glutathione biosynthesis through biochemical analysis of glutamate-cysteine ligase and glutathione synthetase from a model photosynthetic prokaryote William B. MUSGRAVE, Hankuil YI, Dustin KLINE, Jeffrey C. CAMERON, Jonathan WIGNES, Sanghamitra DEY, Himadri B. PAKRASI and Joseph M. JEZ1 Department of Biology, Washington University, One Brookings Drive, Campus Box 1137, St. Louis, MO 63130, U.S.A. Glutathione biosynthesis catalysed by GCL (glutamate-cysteine ligase) and GS (glutathione synthetase) is essential for maintaining redox homoeostasis and protection against oxidative damage in diverse eukaroytes and bacteria. This biosynthetic pathway probably evolved in cyanobacteria with the advent of oxygenic photosynthesis, but the biochemical characteristics of progenitor GCLs and GSs in these organisms are largely unexplored. In the present study we examined SynGCL and SynGS from Synechocystis sp. PCC 6803 using steady-state kinetics. Although SynGCL shares ∼ 15 % sequence identity with the enzyme from plants and α-proteobacteria, sequence comparison suggests that these enzymes share similar active site residues. Biochemically, SynGCL lacks the redox regulation associated with the plant enzymes and functions as a monomeric protein, indicating that evolution of redox regulation occurred later in the green lineage. Site-directed mutagenesis of SynGCL establishes this enzyme as part of the plant-like GCL family and identifies a catalytically essential arginine residue, which is structurally conserved across all forms of GCLs, including those from non-plant eukaryotes and γ -proteobacteria. A reaction mechanism for the synthesis of γ -glutamylcysteine by GCLs is proposed. Biochemical and kinetic analysis of SynGS reveals that this enzyme shares properties with other prokaryotic GSs. Initial velocity and product inhibition studies used to examine the kinetic mechanism of SynGS suggest that it and other prokaryotic GSs uses a random ter-reactant mechanism for the synthesis of glutathione. The present study provides new insight on the molecular mechanisms and evolution of glutathione biosynthesis; a key process required for enhancing bioenergy production in photosynthetic organisms. INTRODUCTION proteobacteria (e.g. Xanthomonas campestris and Agrobacterium tumefaciens) form three distinct phylogentic groups, which are unrelated by sequence, but share common three-dimensional folds [10–13]. A defining feature of each group is the oligomeric organization of GCL as a monomeric enzyme in γ - and αproteobacteria, a redox-sensitive heterodimer of catalytic and regulatory subunits in non-plant eukaryotes, and as a redoxsensitive homodimer in plants [12,14–17]. Similarly, GSs can be grouped into prokaryotic and eukaryotic forms that function as tetrameric and dimeric enzymes respectively [9,18–23]. Glutathione biosynthesis and its biological roles in bacteria, yeast, humans and plants have been well studied, but comparatively little is known about the enzymes of this pathway in cyanobacteria, which is surprising considering that glutathione metabolism probably evolved in these organisms with the advent of oxygenic photosynthesis [9]. In contrast with eukaryotes and microbes, where the ascorbate/glutathione cycle functions as a central antioxidant system [1,7,8], glutathione alone may act as the primary soluble redox buffer in cyanobacteria. Cyanobacteria contain high levels of glutathione that can be further increased by providing exogenous cysteine [24–27]. The ascorbate levels in cyanobacteria are nearly 250-fold lower than those reported for plant chloroplasts and genes encoding for ascorbate peroxidase have not been identified in cyanobacterium, although enzymatic activity has been reported in Nostoc muscorum 7119 and Synechococcus 6311 and 7942 [25,28,29]. Metabolically, glutathione in Synechocystis is linked Glutathione (L-γ -glutamyl-L-cysteinyl-glycine) is the central redox buffer in eukaryotes and many prokaryotes and plays multiple roles in maintaining cellular homoeostasis in response to a range of oxidative stresses [1]. As part of these protection systems, glutathione is critical for redox buffering, detoxification of xenobiotics, sulfur storage and transport, and as a modifier of protein function [1–6]. Photosynthetic organisms use multiple enzymatic and small molecule anti-oxidant and redox-buffering systems to protect themselves from oxidative damage [7,8]; however, phylogenetic studies suggest that the evolution of the enzymes required for glutathione biosynthesis arose in cyanobacteria as protection against ROS (reactive oxygen species) generated as by-products from photosynthesis and metabolism [8,9]. The enzymes of the cyanobacterial glutathione biosynthesis pathway offer a glimpse at the origins of this critical metabolic system and an understanding of their biochemical properties is critical in organisms with a potential for sustainable bioenergy production. Glutathione synthesis occurs in two sequential ATP-dependent steps catalysed by GCL (glutamate-cysteine ligase; also known as γ -glutamylcysteine synthetase; EC 6.3.2.2) and GS (glutathione synthetase, EC 6.3.2.3) (Figure 1). The GCLs of (i) γ -proteobacteria (e.g. Escherichia coli), (ii) non-plant eukaryotes (e.g. human, yeast and Drosophila), and (iii) plants (e.g. Arabidopsis thaliana, Brassica juncea and Physcomitrella patens) and α- Key words: biosynthesis, cyanobacterium, glutamate-cysteine ligase (GCL), glutathione, glutathione synthetase (GS), Synechocystis. Abbreviations used: BjGCL, Brassica juncea glutamate-cysteine ligase; BSO, buthionine sulfoxime; GCL, glutamate-cysteine ligase; GS, glutathione synthetase; ROS, reactive oxygen species; SynGCL, Synechocystis sp. PCC 6803 glutamate-cysteine ligase; SynGS, Synechocystis sp. PCC 6803 glutathione synthetase. 1 To whom correspondence should be addressed (email [email protected]). c The Authors Journal compilation c 2013 Biochemical Society 64 Figure 1 W. B. Musgrave and others Overall reactions of glutathione biosynthesis The reactions catalysed by GCL (gshA in Synechocystis sp. PCC6803) and GS (gshB in Synechocystis sp. PCC6803) are shown. to selenate tolerance [30], iron–sulfur cluster delivery [31,32] and arsenate tolerance [33]. To date, only limited examinations of GCL and GS, encoded by the gshA and gshB genes respectively, from Synechocystis and other cyanobacteria have been reported. Cloning and expression of the gshA gene from Anabena sp. PCC 7120 confirmed that it encodes a GCL [34]. Moreover, analysis of a gshA deletion mutant in Synechocystis sp. PCC 6803 suggested that this gene is essential for growth, as a homozygous mutant could not be isolated [35]. A gshB mutant strain was previously isolated from a screen for pigment biosynthesis in Synechococcus sp. PCC 7942 [36]. In Synechocystis sp. PCC 6803, generation of a gshB deletion mutant resulted in accumulation of γ -glutamylcysteine [35]. Additional studies using the Synechocystis sp. PCC 6803 gshB deletion mutant established that synthesis of glutathione is beneficial during acclimation to both environmental and redox perturbations, especially during extreme oxidative stresses, and is important for photosynthetic electron transport efficiency, antibiotic resistance and photosystem stability [35,37,38]. In the present study we examine the biochemical function of the GCL (SynGCL) and GS (SynGS) from the glutathione biosynthesis pathway of Synechocystis sp. PCC 6803. Previous crystallographic studies of the GCL from Indian mustard (B. juncea) provided information about the active site architecture of a group 3 enzyme [16], but no functional studies examining the contributions of active-site residues for either these or any GCLs have been reported. Identification of amino acid residues that alter the kinetic properties of SynGCL indicates that this enzyme shares active site features with the GCL from plants, even though these enzymes show less than 15 % sequence identity. Mutagenesis of SynGCL also suggests a common chemical reaction mechanism for the synthesis of γ -glutamylcysteine in the first committed step of the pathway in all organisms. Kinetic analysis of SynGS provides the first examination of the steadystate reaction sequence of a prokaryotic GS. Overall, these studies reveal the molecular details of the glutathione biosynthesis pathway in the organism believed to be the evolutionary precursor of chloroplasts in the green plant lineage. Comparison with the enzymes from other organisms suggests that biochemical optimization of these enzymes occurred early with the later addition of specialized regulatory controls. The generation and detoxification of ROS is a central factor during many cellular perturbations with implications in health and biotechnology. Since glutathione is a key soluble antioxidant and redox buffer, a mechanistic and kinetic understanding of its synthesis is critical to understand and modulate cellular redox environment. EXPERIMENTAL Materials E. coli BL21(DE3) cells were purchased from Novagen. Ni2 + -nitrilotriacetic acid agarose was bought from Qiagen. c The Authors Journal compilation c 2013 Biochemical Society Benzamidine–Sepharose and the HiLoad 26/60 Superdex-200 FPLC column were from Amersham. The BioMol Green reagent used for the detection of inorganic phosphate was purchased from BioMol Research Laboratories. For the GS assays, γ glutamylcysteine was purchased from Bachem. All other reagents were purchased from Sigma–Aldrich and were of ACS reagent quality or better. Expression construct generation and site-directed mutagenesis Synechocystis sp. PCC 6803 gshA (slr0990) was PCR-amplified from the previously described pSL2083 construct [35] with 5 -dTTTCATATGCAATTGACTAAAGGGTTAGAAGT-3 as the forward primer (NdeI site is underlined and start codon is in bold) and 5 -dATAGTCGACTCACACCAAACTTAGTATTTCATCCCGGGC-3 as the reverse primer (SalI site is underlined and stop codon is in bold). A similar approach was used for Synechocystis sp. PCC 6803 gshB (slr1238) from the previously described pSL2085 construct [35] with 5 -dTTTCATATGAAACTGGCTTTTATTATCGATCC-3 as the forward primer (NdeI site is underlined and start codon is in bold) and 5 -dATAGTCGACCTAAAATTGTTTTTCCAACCAGCAAATTAC-3 as the reverse primer (SalI site is underlined and stop codon is in bold). The resulting PCR fragments were digested with NdeI and SalI and then sub-cloned into pET-28a digested with the same restriction enzymes to generate the pET-28a-SynGCL and pET-28a-SynGS constructs. Site-directed mutagenesis of SynGCL used the ® QuikChange PCR mutagenesis method (Agilent Technologies) with pET-28a-SynGCL as a template to generate the following mutants: E37Q, E44Q, T117S, T117A, R167K, R167A, H121Q, H121A, R248K and R248A. Protein expression and purification Protein expression and purification for SynGCL (wild-type and mutants) and SynGS used similar methods. Expression constructs were transformed into E. coli BL21(DE3) cells. Transformed E. coli were grown at 37 ◦ C in Terrific broth containing 50 μg/ml kanamycin until D600 ≈ 0.8. After induction with 1 mM isopropyl β-D-thiogalactopyranoside, the cultures were grown at 20 ◦ C for 18 h. Cells were pelleted by centrifugation (4000 g for 10 min) and resuspended in lysis buffer [50 mM Tris, pH 8.0, 500 mM NaCl, 25 mM imidazole, 5 mM MgCl2 , 10 % (v/v) glycerol and 1 % (v/v) Tween 20]. After sonication (Fisher Brand Sonicator; 6 × 30 s total time 0.5 s on/off cycle; on ice) and centrifugation (13 000 g for 30 min), the supernatant was passed over a Ni2 + nitrilotriacetic acid agarose column equilibrated with lysis buffer. His-tagged protein was eluted with elution buffer (lysis buffer without Tween 20 and with 250 mM imidazole) and then loaded on to a Superdex S-200 26/60 size-exclusion FPLC column equilibrated in 50 mM Hepes, pH 7.5, 5 mM MgCl2 , 100 mM NaCl and 10 % glycerol. Protein concentration was determined by the Bradford method (Protein Assay, Bio-Rad Laboratories) with BSA as standard. Glutathione biosynthesis in cyanobacteria Enzyme assays RESULTS The enzymatic activities of SynGCL and SynGS were determined spectrophotometrically at 25 ◦ C by measuring the rate of ADP formation using a coupled assay with pyruvate kinase and lactate dehydrogenase [11,21]. A common reaction mixture (0.5 ml) was used for both proteins and contained 100 mM Hepes, pH 7.5, 150 mM NaCl, 10 mM MgCl2 , 2 mM sodium phosphoenolpyruvate, 0.2 mM NADH, 5 units of type III rabbit muscle pyruvate kinase and 10 units of type II rabbit muscle lactate dehydrogenase. For SynGCL, standard substrate concentrations were 1 mM cysteine, 10 mM glutamate and 2 mM ATP. For SynGS, standard substrate concentrations were 1 mM γ -glutamylcysteine, 5 mM glycine and 1 mM ATP. Reactions were initiated by addition of protein and the rate of decrease in A340 was followed using a Beckman DU800 spectrophotometer. For determination of the steady-state kinetic parameters, two substrates were fixed at saturation and the third substrate varied in concentration and the resulting initial velocity data was fitted to v = kcat [S]/(K m + [S]) using Kaleidagraph (Synergy Software). Expression and purification of Synechocystis glutathione biosynthesis enzymes Initial velocity analysis of the SynGS kinetic mechanism Analysis of the kinetic mechanism used global data fitting analysis [11,21,39]. Reaction rates were measured using standard assay conditions in a matrix of substrate concentrations (γ glutamylcysteine: 0.05–0.5 mM; ATP: 0.1–1.2 mM; glycine: 0.1–1.2 mM). Global curve fitting in SigmaPlot (Systat Software) was used for modelling of the kinetic data to rapid equilibrium rate equations for the 16 possible ter-reactant kinetic mechanisms [11,21,39]. The best fit of the data was to v/V max = ([Glu][Cys][ATP]/αβγ K Glu K Cys K ATP )/(1 + [Glu]/K Glu + [Cys]/K Cys + [ATP]/K ATP + [Glu][Cys]/γ K Glu K Cys + [Glu][ATP]/ β K Glu K ATP + [Cys][ATP]/αK ATP K Cys + [Glu][Cys][ATP]/αβ γ K Glu K Cys K ATP ), which describes a random, rapid-equilibrium ter-reactant system. Product inhibition assays for SynGS Initial velocities in the presence of either glutathione or phosphate were monitored spectrophotometrically using the standard assay system. Enzymatic activity of SynGS was determined after addition of protein to solutions containing either glutathione (0, 15, 30 and 60 mM) or phosphate (0, 25, 50 and 100 mM) and various concentrations of γ -glutamylcysteine (0.01–0.15 mM), glycine (0.1–1 mM) or ATP (0.01–0.3 mM). Assays with added phosphate used MnCl2 at a 1:1 ratio with ATP in place of MgCl2 to prevent precipitation of magnesium phosphate. For assays with added ADP, initial rates were measured by colorimetric determination of inorganic phosphate. Each 0.1 ml assay mixture contained 100 mM Hepes, pH 7.5, 150 mM NaCl, 10 mM MgCl2 , various substrate concentrations and ADP (0, 0.25, 0.50, 1 and 2.5 mM). The reaction was initiated by addition of protein and quenched at several time intervals (0– 5 min) with 1 ml of BioMol Green reagent. Reaction rates were linear over the time course monitored. Control reactions were used to correct for background levels of inorganic phosphate in the reaction mixture. Colour was measured at A620 and quantified with a standard curve for inorganic phosphate (0– 8 nmol). Global fitting analysis of initial velocity data was performed with SigmaPlot (Systat Software) using the equations for competitive inhibition, v = V max /(1 + {(K m /[S])(1 + [I]/K i )}); non-competitive inhibition, v = V max /{(1 + [I]/K i )(1 + K m /[S])}; and uncompetitive inhibition, v = V max /(1 + [I]/K i + K m /[S]). 65 The predicted function of the Synechocystis gshA and gshB genes as GCL and GS respectively was reported previously [35]; however, biochemical analysis of these proteins has not been performed. The nucleotide sequence of Synechocystis gshA (1170 bp) encodes a 389 amino acid protein with a predicted molecular mass of 44.1 kDa and a pI value of 5.35 (Figure 2). Amino acid sequence comparisons show that it shares 64 % sequence identity with the previously identified GCL from Anabena [34], 15 % identity with the GCL from A. thaliana and B. juncea, and less than 12 % identity with the enzyme from α-proteobacteria (Figure 2). As discussed below, although the overall similarity between the cyanobacterial and plant GCL is limited, a number of active site residues described in the crystal structure of BjGCL (B. juncea GCL) are conserved in the cyanobacterial enzymes [16]. No significant homology was detected between the cyanobacterial GCL and the bacterial, fungal and human enzymes. The nucleotide sequence of Synechocystis gshB (936 bp) encodes a 320 amino acid protein with a predicted molecular mass of 35.1 kDa and a pI of 5.39 (Figure 3). Comparison of SynGS with the enzyme from E. coli revealed 39 % sequence identity, whereas comparison with the GS from Arabidopsis, human and yeast showed less than 15 % sequence identity. Active site residues identified in the structure of the E. coli enzyme in complex with ADP and glutathione are nearly invariant in SynGS (Figure 3) [40]. For biochemical analysis of the Synechocystis glutathione biosynthesis enzymes, both were overexpressed in E. coli as N-terminal hexahistidine-tagged proteins and purified using nickel-affinity and size-exclusion chromatographies. SDS/PAGE analysis of the purified recombinant proteins showed that SynGCL and SynGS migrated with molecular masses of ∼ 46 and ∼ 38 kDa respectively, which corresponded to the predicted mass of each expressed protein (Supplementary Figure S1 at http://www.biochemj.org/bj/450/bj4500063add.htm). Gelfiltration chromatography of SynGCL showed that the protein eluted as a monomeric 45 kDa form, as observed for the GCL from α- and γ -proteobacteria [12,17]. SynGS eluted from the size-exclusion column as a 150 kDa species that represents a tetramer similar to that reported for the E. coli enzyme [18]. Steady-state kinetic analysis of SynGCL The steady-state kinetic parameters of the SynGCL for glutamate, cysteine and ATP were determined using a spectrophotometric assay (Table 1). The kinetic values were comparable with those reported for the GCL from Anabena [34]. Enzymatic activity required the presence of Mg2 + ; however, screening of other divalent ions (10 mM CaCl2 , 10 mM MnCl2 or 10 mM CuSO4 ) showed that only Mn2 + could substitute for Mg2 + with a 5-fold reduction in specific activity. Redox sensitivity in the plant GCL is mediated by two cysteine residues [11,15– 17], which are absent in the sequences of the cyanobacterial GCL (Figure 2). Consistent with the lack of both regulatory cysteine residues in SynGCL, incubation of the protein with either 5 mM 2-mercaptoethanol or dithiothreitol did not alter enzymatic activity, unlike in plants [11,15–17]. SynGCL was inhibited by glutathione (IC50 = 7.4 + − 0.3 mM), which is comparable with the inhibition of the Arabidopsis and Anabena enzymes [11,34]. BSO (buthionine sulfoxime), which rapidly inactivates the bacterial c The Authors Journal compilation c 2013 Biochemical Society 66 Figure 2 W. B. Musgrave and others Sequence comparison of GCLs from cyanobacteria and plants Alignment of the sequence of the GCLs from Synechocystis sp. PCC 6803 (SynGCL; accession number slr0990), Anabena sp. PCC 7120 (AnaGCL; accession number NP_487391.1), A. thaliana (AtGCL; accession number AEE84706.1), B. juncea (BjGCL; accession number CAD91712), P. patens (PpGCL; accession number Phypa1_1:173526), X. campestris (XcGCL; accession number NP_638742) and A. tumefaciens (AgGCL; accession number NP_353679). For clarity, the alignment does not include the N-terminal regions of the plant and α-proteobacteria sequences that do not match the cyanobacterial enzymes. Numbering corresponds to SynGCL with every tenth residue marked above the alignment. Invariant residues are highlighted in black with grey text and similar residues are highlighted in grey with white. Residues that correspond to those in the active site of the BjGCL crystal structure [16] are indicated with an asterisk. Residues in the regulatory disulfide bond of the plant GCL [15–16] are indicated with a #. Figure 3 Sequence comparison of GSs from cyanobacteria and bacteria Alignment of the amino acid sequences of the GSs from Synechocystis sp. PCC 6803 (SynGS; slr1238) and E. coli (EcGS; BAE77010.1). Numbering corresponds to SynGS with every tenth residue marked above the alignment. Conserved residues are highlighted in grey with white text. Active site residues identified in the E. coli GS crystal structure [40] are indicated with an asterisk. Table 1 Steady-state kinetic parameters of SynGCL Active site mutagenesis of SynGCL Reactions were performed as described in the Experimental section. All V /E tot and K m values are expressed as a means + − S.E.M. (n = 3). Substrate V /E tot (min − 1 ) K m (μM) V /K m (M − 1 ·s − 1 ) Glutamate Cysteine ATP 21.0 + − 1.3 25.2 + − 1.7 30.6 + − 1.9 953 + − 16 59 + − 0.5 145 + − 13 367 7120 3520 and mammalian enzymes [41,42], is a poor inhibitor of SynGCL (IC50 = 33.7 + − 4.7 mM). c The Authors Journal compilation c 2013 Biochemical Society Although the cyanobacterial and plant GCLs share limited sequence identity, alignment of the GCLs from plants, α-proteobacteria and cyanobacteria shows conservation of residues across the length of the proteins, including active site residues identified in the crystal structure of the BjGCL in complex with the inhibitor BSO [16] (Figures 2 and 4). In the B. juncea active site (Figure 4), the portion of BSO corresponding to glutamate is bound by Thr242 , Gln246 , Arg292 and Trp296 , which correspond to Thr117 , His121 , Arg167 and Phe171 in SynGCL. Residues forming the proposed cysteine-binding site in the plant GCLs (i.e. Arg220 , Tyr221 , Met224 , Met239 , Tyr330 and Phe375 ) are varied in SynGCL (Pro99 , Gln100 , Val116 , Met200 Glutathione biosynthesis in cyanobacteria Table 2 67 Steady-state kinetic parameters of mutant SynGCL Reactions were performed as described in the Experimental section. Reactions performed at sub-saturating (50 mM) glutamate are indicated by an asterisk. V /E tot and K m values were expressed as a means + − S.E.M. (n = 3). Mutant Substrate V /E tot (min − 1 ) K m (μM) V /K m (M − 1 ·s − 1 ) T117S Glutamate Cysteine ATP Glutamate Cysteine ATP Glutamate Cysteine* ATP* Glutamate Cysteine* ATP* Glutamate Cysteine ATP 8.2 + − 0.6 8.0 + − 1.3 13.5 + − 1.5 11.3 + − 1.4 20.1 + − 2.7 22.9 + − 2.2 7.6 + − 2.4 5.4 + − 0.1 5.9 + − 0.2 2.7 + − 0.7 1.5 + − 0.2 2.3 + − 0.3 0.5 + − 0.1 0.6 + − 0.1 0.5 + − 0.1 6100 + − 150 186 + − 6.1 352 + − 19 1180 + − 135 116 + − 16 279 + − 66 15200 + − 4900 707 + − 49 310 + − 18 >50 mM 1950 + − 180 445 + − 71 3760 + − 427 857 + − 10 969 + − 77 22.4 717 639 160 2890 1370 8.3 127 317 <0.9 12.8 86.1 2.2 11.7 8.6 H121Q H121A R167K Figure 4 GCL active site Residues forming the active site of the BjGCL (PDB code 2GWC [16]) are shown. BSO is labelled with the cysteine (cys) and glutamate (glu) portions of the molecule indicated. The proposed position of the ATP-binding site in BjGCL is also shown. Residues (in single letter code) are numbered according to BjGCL with the corresponding residues of SynGCL shown in parentheses. and Leu244 ). Two glutamate residues in the Mg2 + -binding site of BjGCL (Glu159 and Glu165 ) are also found in SynGCL (Glu37 and Glu44 ). Moreover, an invariant arginine residue (Arg379 in BjGCL; Arg248 in SynGCL) is positioned between the dipeptide-binding site and the proposed ATP-binding site in BjGCL. Due to the low sequence identity between the cyanobacterial and plant enzymes, we examined the functional contributions of putative active site residues in SynGCL using a series of sitedirected mutants (E37Q, E44Q, T117S, T117A, H121Q, H121A, R167K, R167A, R248K and R248A). Each mutant was expressed in E. coli cells and purified as a soluble monomeric protein, as determined by size-exclusion chromatography. The SynGCL E37Q, E44Q, T117A, R167A and R248A mutants displayed no detectable activity. Other mutants exhibited a range of effects on the steady-state kinetic parameters (Table 2). The T117S and H121Q mutants displayed less than 5-fold reductions in turnover rate and K m values for all three substrates that were increased 3- to 22-fold. Mutation of His121 to an alanine and Arg167 to a lysine primarily affected the K m value of each mutant for glutamate, although the cysteine K m value was also strongly affected. As the concentration for this substrate in assays with either varied cysteine or varied ATP was not at saturation, the kinetic parameters of the H121A and R167K mutants should be considered as estimates. The R248K mutation largely affected the turnover rate with 40- to 60-fold reductions, along with 15- and 7fold increases in the K m values for cysteine and ATP respectively. Overall, the effect of mutations in SynGCL were consistent with the location of these residues in an active site similar to that of BjGCL and suggest a possible chemical reaction mechanism for these enzymes. Steady-state kinetic analysis of SynGS The steady-state kinetic parameters for γ -glutamylcysteine, glycine and ATP were determined for SynGS (Table 3). The K m values of SynGS for the three substrates differ by less than 3-fold compared with the E. coli enzyme, although the turnover rate of the cyanobacterial enzyme is ∼ 50-fold slower than the bacterial GS [43,44]. SynGS is a Mg2 + -dependent enzyme; however, Mn2 + could substitute with a 2.2-fold reduction in turnover rate at a 10 mM ion concentration. In plants, glutathione analogues with R248K Table 3 Steady-state kinetic parameters of SynGS Reactions were performed as described in the Experimental section. All V /E tot and K m values were expressed as means + − S.E.M. (n = 3). Substrate V /E tot (min − 1 ) K m (μM) V /K m (M − 1 ·s − 1 ) γ -Glutamylcysteine Glycine ATP 277 + − 29 253 + − 35 324 + − 18 99 + − 14 407 + − 90 136 + − 12 46 630 10 360 39 710 substitutions of the glycine moiety are formed by GS-related enzymes [23]. To test for possible substrate variation, β-alanine, serine and glutamate were assayed with SynGS and yielded no detectable enzymatic activity. Initial velocity and product inhibition analyses of the kinetic mechanism of SynGS To determine the kinetic mechanism of a bacterial GS, initial velocity studies using SynGS were performed to generate data in which for a saturating concentration of one substrate the reaction rates were measured over the entire range of the other two substrates (Supplementary Figure S2 at http://www.biochemj. org/bj/450/bj4500063add.htm). The lack of parallel lines in the double-reciprocal plots eliminated potential ping-pong kinetic mechanisms from consideration. Global fitting of the initial velocity data for all three substrates to each of the sixteen rate equations describing possible ter-reactant systems, which include six potential ordered mechanisms, nine partially ordered/random mechanisms and a random kinetic mechanism, yielded a best fit (r2 = 0.962) using the random ter-reactant equation [11,21,39]. Analysis using other kinetic mechanism models yielded unsatisfactory fits (r2 = 0.881–0.920). Fitting of the data provides estimates of turnover rate, substrate equilibrium dissociation constants (K ATP , K Gly and K γ EC ), and the interaction factors between glycine and γ -glutamylcysteine (α), ATP and γ -glutamylcysteine (β), and ATP and glycine (γ ) (Table 4). The kinetic parameters determined by global fitting (Table 4), in which βγ K ATP , αγ K Gly and αβK γ EC correspond to the apparent K m values for ATP, γ -glutamylcysteine and glycine respectively, are consistent with those determined by steady-state kinetic analysis (Table 2). The model-derived parameters suggest c The Authors Journal compilation c 2013 Biochemical Society 68 W. B. Musgrave and others Table 4 Kinetic constants for a random ter-reactant model of the SynGS kinetic mechanism K ATP , K Gly and K γ EC are the calculated equilibrium dissociation constants for the binding of substrate with the free enzyme. The interaction factors between glycine and γ -glutamylcysteine (γ EC), ATP and γ EC, and ATP and glycine are represented by α, β and γ respectively. Values are means + − S.E.M. (i) Fitted parameters Parameter Value V /E tot (min − 1 ) K ATP (μM) K Gly (μM) K γ EC (μM) α β γ 184 + − 25 170 + − 28 403 + − 75 211 + − 33 0.42 + − 0.12 0.70 + − 0.20 1.93 + − 0.37 (ii)Substrate dependences Kinetic constant Value K ATP (μM) β K ATP γ K ATP βγ K ATP K Gly (μM) α K Gly γ K Gly αγ K Gly K γ EC (μM) α K γ EC β K γ EC αβ K γ EC 179 125 346 242 403 169 778 1.5 211 89 148 62 Figure 5 Kinetic mechanism of SynGS A random ter-reactant system best describes the formation of a catalytic ternary complex. K ATP , K Gly and K γ EC are the equilibrium constants for substrate binding. Interaction factors between glycine and γ -glutamylcysteine (γ EC), ATP and γ EC, and ATP and glycine are represented by α, β and γ respectively. Bold arrows indicate the preferred route to the ternary complex with broken arrows showing steps with negative interactions between binding sites. that modest positive and negative interactions occur between different binding sites. Positive interactions occur between the ATP and γ -glutamylcysteine (β = 0.70) and the glycine and γ -glutamylcysteine binding sites (α = 0.42) with a negative interaction between ATP and glycine (γ = 1.93). The fitted parameters (Table 4) and the resulting kinetic model (Figure 5) suggest a random ter-reactant system with a preferred set of interactions between binding sites for formation of the catalytic ternary complex in SynGS. To examine the order of product release from SynGS, the inhibition patterns for ADP, glutathione and phosphate against each substrate were analysed (Table 5). ADP was a competitive c The Authors Journal compilation c 2013 Biochemical Society Table 5 Product inhibition patterns of SynGS All assays (n = 3) were performed as described in the Experimental section. γ EC, γ -glutamylcysteine. The inhibition patterns are as follows: C, competitive; NC, noncompetitive. Varied substrate Inhibitor Fixed substrate Inhibition K i (mM) ATP γ EC Glycine ATP γ EC Glycine ATP γ EC Glycine ADP ADP ADP Glutathione Glutathione Glutathione Pi Pi Pi γ EC (1 mM), glycine (5 mM) Glycine (5 mM), ATP (2 mM) γ EC (1 mM), ATP (2 mM) γ EC (2 mM), glycine (5 mM) Glycine (5 mM), ATP (2 mM) γ EC (1 mM), ATP (2 mM) γ EC (1 mM), glycine (5 mM) Glycine (5 mM), ATP (2 mM) γ EC (1 mM), ATP (2 mM) C NC NC NC NC NC NC NC NC 1.12 + − 0.19 1.41 + − 0.20 2.19 + − 0.22 29.9 + − 5.7 27.3 + − 6.0 38.6 + − 3.7 54.0 + − 5.3 42.8 + − 7.3 68.2 + − 9.3 inhibitor with respect to ATP and a non-competitive inhibitor against both γ -glutamylcysteine and glycine. Both glutathione and phosphate were non-competitive inhibitors of ATP, γ -glutamylcysteine and glycine. DISCUSSION Although glutathione plays a central role in maintaining cellular redox homoeostasis, only recently have biochemical and structural studies provided insight on how the enzymes required for its biosynthesis function and are regulated. The synthesis of thiol-containing redox buffers, especially in photosynthetic organisms, also may have evolved to overcome the instability of cysteine [43]. The role of GCLs as the first step of glutathione biosynthesis is thought to have originated in cyanobacteria and then spread to other bacteria and eukaryotes leading to three distinct groups of GCLs [9]; however, sequence similarity between the various types of GCLs is limited. Moreover, little molecular information on this enzyme from cyanobacteria is available [34]. Previous bioinformatic analysis showed that the cyanobacterial GCL share less than 15 % identity with the plant enzymes and lacked any detectable homology with the GCLs from γ -proteobacteria (group 1) and non-plant (group 2) eukaryotes. Interestingly, structure-guided sequence comparison of SynGCL with the GCLs from A. thaliana and B. juncea (Figures 2 and 4) suggests that the cyanobacterial and plant GCLs may share similar active site residues. To investigate this possible relationship, we first functionally examined SynGCL and then tested the contribution of conserved residues to catalysis. Biochemical characterization of SynGCL revealed that the protein functioned as a monomer, which is similar to that of the enzyme from the α- and γ -proteobacteria [12,17]. Steadystate kinetic analysis of SynGCL (Table 1) demonstrated that the cyanobacterial enzyme was insensitive to changes in redox state, which is consistent with the lack of residues forming the regulatory disulfide found in the plant GCL [11,15–17]. As reported for the GCLs from E. coli, yeast, A. thaliana and Anabena [11–13,34], glutathione was a millimolar inhibitor of SynGCL. The lower K i of glutathione for GCL compared with GS also suggests that this enzyme is more sensitive to control by feedback inhibition. Moreover, the poor inhibition of the cyanobacterial enzyme by BSO, a potent inactivator of the bacterial and mammalian homologues [41,42], suggests probable differences in active site structure between the different groups of GCLs. As the limiting step in the pathway [44], regulation of GCLs through feedback inhibition by glutathione appears to be the original control mechanism, with thiol-based redox-regulation of oligomerization occuring later in plants and the recrutiment of Glutathione biosynthesis in cyanobacteria Figure 6 69 Proposed reaction mechanism of SynGCL a regulatory subunit in the GCLs of non-plant eukaryotes [12,14– 17]. Previous crystallographic studies of the GCL from B. juncea provided information about the active site architecture of a group 3 enzyme [16] (Figure 4), but no functional studies examining the contributions of active site residues for these GCL have been reported. In the BjGCL structure, the position of the inhibitor BSO delineated residues in proximity to the glutamate- and cysteinederived portions of the ligand and residues forming a Mg2 + binding site [16]. Sequence comparison of the cyanobacterial and plant GCLs implies that these distantly related enzymes retain a common active site, as key residues in the active site are invariant between the two proteins (Figures 2 and 4). On the basis of the BjGCL structure (Figure 4), Glu37 and Glu44 in SynGCL would be positioned to bind a catalytically essential Mg2 + ion and Thr117 , Arg167 and His121 would interact with glutamate. A role for the invariant arginine residue located at the interface of the dipeptide- and ATP-binding sites in BjGCL has not been previously proposed; however, this residue (Arg248 in SynGCL) is ideally positioned to stabilize the reaction catalysed by the enzyme (Figure 6). In the present study we used site-directed mutagenesis to examine the role of invariant residues (i.e. Glu37 , Glu44 , Thr117 , Arg167 , His121 and Arg248 ) in SynGCL. Kinetic analysis of point mutants targeting these residues in SynGCL suggests that the plant and cyanobacterial GCLs share common features (Table 2). Mutation of either Glu37 or Glu44 in the putative Mg2 + -binding site to a glutamine abrogated enzymatic activity with the side-chain charge being as important to maintain the position of the metal ion. The effects of the T117S/A, H121Q/A and R167K/A mutants indicate that these residues, which are conserved between the plant and cyanobacterial GCL, are important for efficient formation of γ -glutamylcysteine, most likely for binding of the glutamate substrate. The T117A and R167A mutants lacked detectable activity with all the other mutants exhibiting altered kinetic parameters compared with the wild-type enzyme. Mutation of His121 to a glutamine residue yielded less than 3-fold changes in kinetic parameters, whereas mutation to an alanine residue primarily altered the K m of glutamate. The T117S mutant displayed a 10- to 15-fold reduction in catalytic efficiency with all three substrates. The largest changes were observed with the R167K mutant, which did not show saturation kinetics using up to 50 mM glutamate. On the basis of the crystal structure of BjGCL (Figure 4), Arg167 in SynGCL is predicted to form a critical ionic interaction with the glutamate backbone carboxy group. Mutagenesis of Arg248 , which is an invariant residue across the group 3 GCLs (Figure 2), to an alanine residue results in an inactive enzyme and the R248K mutation resulted in an up to 610-fold decrease in V/K m . On the basis of the position of this arginine residue in the BjGCL crystal structure at the interface of the putative ATP- and dipeptide-binding sites and the steady-state kinetic results observed for the SynGCL R248K and R248A mutants, this residue plays a critical role in catalysis. The overall reaction catalysed by all GCLs requires the ATP-dependent phosphorylation of glutamate to form an acyl-phosphate intermediate. Next, nucleophilic attack of the cysteine amine group leads to release of inorganic phosphate and formation of γ -glutamylcysteine [3,11]. We propose a reaction mechanism for SynGCL (and the other group 3 enzymes) in which the guanidyl group of Arg248 stabilizes formation of the pentavalent transition state that results in a phosphorylated glutamate intermediate and the formation of the transition state after nucleophilic attack by cysteine (Figure 6). The catalytically essential Mg2 + , which is co-ordinated by Glu37 and Glu44 , increases the nucleophilicity of the γ -carboxylate during formation of the acylphosphate intermediate and stabilizes the phosphate group in the active site [42,45– 47]. The order of chemical steps also agrees with the proposed kinetic mechanisms for the GCLs from A. thaliana and Trypanosoma [11,39]. Although the three groups of GCLs lack sequence homology [10], the structures of the GCLs from E. coli, yeast and B. juncea reveal a common three-dimensional fold [12,13,16,47]. Moreover, each of these enzymes have an arginine residue positioned between the nucleotide- and amino-acid-binding sites, where this residue probably plays a similar mechanistic role across all of the GCLs. Interestingly, the basic features of this reaction are also conserved across other ATP-dependent peptide synthetases, including GS [19,20,22,48–50]. GS, the second enzyme in glutathione biosynthesis, has been extensively studied from eukaryotic sources, including humans, plants and yeast [8,19–23], and multiple structural studies of the E. coli GS have been performed [18,40,45,46]; however, no cyanobacterial GSs have been characterized to date and the kinetic mechanism of the prokaryotic GSs remains unexamined. Multiple biochemical features of SynGS, including sequence homology with E. coli GS (Figure 3), its tetrameric oligomerization and kinetic properties (Table 3), establish this enzyme as a member of the prokaryotic GS family. Analysis of its kinetic mechanism using initial velocity and product inhibition studies are most consistent with a random ter-reactant system that displays a preference for the order of substrate addition. The observed intersection of the initial velocity patterns (Supplementary Figure S2) indicates that no irreversible steps occur between the binding of any substrate and removes c The Authors Journal compilation c 2013 Biochemical Society 70 W. B. Musgrave and others possible ping-pong type mechanisms from consideration. These results also suggest the formation of a ternary complex before catalysis. Global fitting of the initial velocity data to each of the 16 possible rate equations for ter-reactant kinetic mechanisms showed that a random mechanism best describes the observed data (Table 4 and Supplementary Figure S2). The fitted parameters and substrate dependencies indicate a preference for substrate addition that leads to the pre-catalytic ternary complex. The positive interactions between the binding of either ATP/γ -glutamylcysteine or glycine/γ -glutamylcysteine and the negative interaction with ATP/glycine suggest a possible kinetic model for SynGS (Figure 5). In comparison with the kinetic mechanism of A. thaliana GS [21], the substrate dependencies for the cyanobacterial enzyme are much weaker and may reflect differences in the active sites of the prokaryotic and eukaryotic GSs that influence substrate binding. Product inhibition experiments with SynGS also suggest a random mechanism (Table 5). If the kinetic mechanism of SynGS were sequentially ordered, then the observed inhibition patterns would all be uncompetitive, except for the competitive inhibition pattern of the last product against the first substrate [51]. In contrast, a completely random mechanism would lack any abortive complexes and yield competitive inhibition patterns. With SynGS, only ADP is a competitive inhibitor compared with ATP and non-competitive compared with the other two substrates. These results suggest that both ADP and ATP bind to the same enzyme form, which suggests that ADP is the last product released and that ATP is the most likely first substrate to bind. The noncompetitive inhibition of γ -glutamylcysteine and glycine by ADP indicates that ADP is released from an enzyme form that differs from the form that binds either substrate. Inhibition by glutathione and phosphate yielded non-competitive inhibition patterns against all three substrates and suggested that the reaction products are either randomly released, form dead-end complexes with free enzyme and/or an E•S complex, or bind at an allosteric site on the enzyme [52]. Crystal structures of the E. coli GS [18,40,45,46] indicate that there is no allosteric binding site present and suggest that the SynGS active site would not prevent binding of substrates and products as dead-end complexes. Moreover, based on the product inhibition data, the release of phosphate and glutathione from SynGS appears to be random and may occur before release of ADP. Overall, the initial velocity and product inhibition profiles are comparable with those described for multiple other enzymes that use a common set of steps in their reaction sequences, including the GS from A. thaliana [21], GCL from A. thaliana and Trypanosoma [11,34], UDPN-acetylmuramate:L-alanine ligase [53], glutamine synthetase [54], D-alanyl-D-alanine ligase [55] and UDP-N-acetylmuramatetripeptide:D-alanyl-D-alanine-adding enzyme [56]. All of these enzymes transfer the γ -phosphate group of ATP to a substrate C-terminal carboxylate followed by nucleophilic attack on the acylphosphate intermediate by a second substrate leads to product formation with the release of phosphate, peptide product and ADP. The evolution of glutathione biosynthesis in cyanobacteria coincides with the appearance of photosynthesis. Earlier bioinformatic analysis suggested that GCL evolved in cyanobacteria and was later transfered into plants and αproteobacteria [9]; however, the distant homology even among group 3 GCLs from these organisms made functional comparisons difficult. The biochemical and mechanistic analyses of SynGCL for the first time establish a direct link between the GCLs in cyanobacteria and plants. Although the three distinct groups of GCLs lack any sequence homology, they share a common three-dimensional architecture [12,13,16] and, as shown in the c The Authors Journal compilation c 2013 Biochemical Society present study, a structurally conserved arginine residue for catalysis. Similarly, the GSs from prokaryotes and eukaryotes are two distinct phylogenies [9]; however, both types of enzymes use a random ter-reactant kinetic mechanism for the synthesis of a tripeptide critical for protection against oxidative stresses [21]. Cyanobacteria, such as Synechocystis sp. PCC 6803, are excellent model systems for understanding photosynthesis and for testing strategies to enhance its metabolic efficiency for biofuel applications [57–62]. With engineering and optimization of oxygenic photosynthesis, alterations in redox environment can impact electron transport efficiency and photosystem stability and become a major limiting factor in energy utilization and biofuel production [59]. Moreover, metabolic engineering for synthesis of biofuels can also lead to the generation of ROS. For example, pre-treatment of lignocellulosic biomass for biological conversion processes releases toxic furan aldehydes, such as furfural, that promote ROS generation [63]. Similarly, production of biofuel hexane in Synechocystis sp. PCC 6803 results in increased expression of enzymes linked to glutathione metabolism [64]. Recent work suggests that enhancing glutathione biosynthesis can protect against ROS during biofuel production, as the introduction of GCL and GS into Clostridium acetobutylicum, a microbe used for butanol production, led to improved survival upon aeration and increased butanol production [65]. Understanding the molecular basis of glutathione biosynthesis in cyanobacteria is an important step toward optimizing these organisms for future sustainable energy applications and for providing a diverse repertoire of characterized enzymes with favourable kinetic and stability properties that might be useful for tailoring the redox environment of diverse production organisms. AUTHOR CONTRIBUTION William Musgrave, Hankuil Yi and Jeffrey Cameron designed the experiments. William Musgrave, Hankuil Yi, Dustin Kline, Jeffrey Cameron, Jonathan Wignes and Sanghamitra Dey performed the experiments. William Musgrave, Hankuil Yi, Jeffrey Cameron, Sanghamitra Dey, Himadri Pakrasi and Joseph Jez wrote the paper. FUNDING This work was supported by the National Science Foundation [grant number MCB-0904215 (to J.M.J.)] and the Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Science, U.S. Department of Energy [grant number DE-FG02-99ER20350 (to H.B.P.)]. REFERENCES 1 Meister, A. and Anderson, M. E. (1983) Glutathione. Annu. Rev. Biochem. 52, 711–760 2 Rouhier, N., Lemaire, S. D. and Jacquot, J. P. (2008) The role of glutathione in photosynthetic organisms: emerging functions for glutaredoxins and glutathionylation. Annu. Rev. Plant Biol. 59, 143–166 3 Yi, H., Galant, A., Ravilious, G. E., Preuss, M. L. and Jez, J. M. (2010) Sensing sulfur conditions: simple to complex biochemical regulatory mechanisms in plant thiol metabolism. Mol. Plant 3, 269–279 4 Yi, H., Ravilious, G. E., Galant, A., Krishnan, H. B. and Jez, J. M. (2010) Thiol metabolism in soybean: sulfur to homoglutathione. Amino Acids 39, 963–978 5 Foyer, C. H. and Noctor, G. (2011) Ascorbate and glutathione: the heart of the redox hub. Plant Physiol. 155, 2–18 6 Galant, A., Preuss, M. L., Cameron, J. C. and Jez, J. M. (2011) Plant glutathione biosynthesis – diversity in biochemical regulation and reaction products. Front. Plant Sci. 2, 45 7 Asada, K. (1999) The water-water cycle in chloroplasts: scavenging of active oxygens and dissipation of excess photons. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50, 601–639 Glutathione biosynthesis in cyanobacteria 8 Latifi, A., Ruiz, M. and Zhang, C. C. (2009) Oxidative stress in cyanobacteria. FEMS Microbiol. Rev. 33, 256–278 9 Copley, S. D. and Dhillon, J. K. (2002) Lateral gene transfer and parallel evolution in the history of glutathione biosynthesis genes. Genome Biol. 3, 25.1–25.16 10 May, M. J. and Leaver, C. J. (1994) Arabidopsis thaliana γ -glutamylcysteine synthetase is structurally unrelated to mammalian, yeast, and Escherichia coli homologs. Proc. Natl. Acad. Sci. U.S.A. 91, 10059–10063 11 Jez, J. M., Cahoon, R. E. and Chen, S. (2004) Arabidopsis thaliana glutamate-cysteine ligase: functional properties, kinetic mechanism, and regulation of activity. J. Biol. Chem. 279, 33463–33470 12 Hibi, T., Nii, H., Nakatsu, T., Kimura, A., Kato, H., Hiratake, J. and Oda, J. (2004) Crystal structure of γ -glutamylcysteine synthetase: insights into the mechanism of catalysis by a key enzyme from gluathione homeostasis. Proc. Natl. Acad. Sci. U.S.A. 101, 15052–15057 13 Biterova, E. I. and Barycki, J. J. (2009) Mechanistic details of glutathione biosynthesis revealed by crystal structures of Saccharomyces cerevisiae glutamate-cysteine ligase. J. Biol. Chem. 284, 32700–32708 14 Seelig, G. F., Simondsen, R. P. and Meister, A. (1984) Reversible dissociation of γ -glutamylcysteine synthetase into two subunits. J. Biol. Chem. 259, 9345–9347 15 Hicks, L. M., Cahoon, R. E., Bonner, E. R., Rivard, R. S., Sheffield, J. and Jez, J. M. (2007) Thiol-based regulation of redox-active glutamate-cysteine ligase from Arabidopsis thaliana . Plant Cell 19, 2653–2661 16 Hothorn, M., Wachter, A., Gromes, R., Stuwe, T., Rausch, T. and Scheffzek, K. (2006) Structural basis for the redox control of plant glutamate cysteine ligase. J. Biol. Chem. 281, 27557–27565 17 Gromes, R., Hothorn, M., Lenherr, E. D., Rybin, V., Scheffzek, K. and Rausch, T. (2008) The redox switch of γ -glutamylcysteine ligase via a reversible monomer-dimer transition is a mechanism unique to plants. Plant J. 54, 1063–1075 18 Yamaguchi, H., Kato, H., Hata, Y., Nishioka, T., Kimura, A., Oda, J. and Katsube, Y. (1993) Three-dimensional structure of the glutathione synthetase from Escherichia coli B at 2.0 Å resolution. J. Mol. Biol. 229, 1083–1100 19 Polekhina, G., Board, P. G., Gali, R. R., Rossjohn, J. and Parker, M. W. (1999) Molecular basis of glutathione synthetase deficiency and a rare gene permutation event. EMBO J. 18, 3204–3213 20 Gogos, A. and Shapiro, L. (2002) Large conformational changes in the catalytic cycle of glutathione synthase. Structure 10, 1669–1676 21 Jez, J. M. and Cahoon, R. E. (2004) Kinetic mechanism of glutathione synthetase from Arabidopsis thaliana . J. Biol. Chem. 279, 42726–42731 22 Herrera, K., Cahoon, R. E., Kumaran, S. and Jez, J. M. (2007) Reaction mechanism of glutathione synthetase from Arabidopsis thaliana : site-directed mutagenesis of active site residues. J. Biol. Chem. 282, 17157–17165 23 Galant, A., Arkus, K. A., Zubieta, C., Cahoon, R. E. and Jez, J. M. (2009) Structural basis for evolution of product diversity in soybean glutathione biosynthesis. Plant Cell 21, 3450–3458 24 Fahey, R. C., Brown, W. C., Adams, W. B. and Worsham, M. B. (1978) Occurrence of glutathione in bacteria. J. Bacteriol. 133, 1126–1129 25 Tel-Or, E., Huffedji, M. and Packer, L. (1985) The role of glutathione and ascorbate in hydroperoxide removal in cyanobacteria. Biochem. Biophys. Res. Comm. 132, 533–539 26 Suginaka, K., Yamamoto, K., Ashida, H. and Kono, Y. (1998) Cysteine uptake for accumulation of glutathione by the cyanobacterium Synechocystis PCC 6803. Biosci. Biotechnol. Biochem. 62, 424–428 27 Suginaka, K., Yamamoto, K., Ashida, H. and Sawa, Y. (1999) Effect of intracellular glutathione on heat-induced cell death in the cyanobacterium, Synechocystis PCC 6803. Biosci. Biotechnol. Biochem. 63, 1112–1125 28 Mittler, R. and Tel-Or, E. (1991) Oxidative stress responses in the unicellular cyanobacterium Synechococcus PCC7942. Free Radical Res. Comm. 12–13, 845–850 29 Stork, T., Michel, K. P., Pistorius, E. K. and Dietz, K. J. (2005) Bioinformatic analysis of the genomes of the cyanobacteria Synechocystis sp. PCC 6803 and Synechococcus elongatus PCC 7942 for the presence of peroxiredoxins and their transcript regulation under stress. J. Exp. Bot. 56, 3193–3206 30 Marteyn, B., Domain, F., Legrain, P., Chauvat, F. and Cassier-Chauvat, C. (2009) The thioredoxin reductase-glutaredoxins-ferredoxin crossroad pathway for selenate tolerance in Synechocystis PCC6803. Mol. Microbiol. 71, 520–532 31 Picciocchi, A., Saguez, C., Boussac, A., Cassier-Chauvat, C. and Chauvat, F. (2007) CGFS-type monothiol glutaredoxins from the cyanobacterium Synechocystis PCC6803 and other evolutionary distant model organisms possess a glutathione-ligated [2Fe-2S] cluster. Biochemistry 46, 15018–15026 32 Iwema, T., Picciocchi, A., Traore, D. A., Ferrer, J. L., Chauvat, F. and Jacquamet, L. (2009) Structural basis for delivery of the intact [Fe2 S2 ] cluster by monothiol glutaredoxin. Biochemistry 48, 6041–6043 33 Lopez-Maury, L., Sanchez-Riego, A. M., Reyes, J. C. and Florencio, F. J. (2009) The glutathione/glutaredoxin system is essential for arsentate reduction in Synechocystis sp. strain PCC 6803. J. Bacteriol. 191, 3534–3543 71 34 Ashida, H., Sawa, Y. and Shibata, H. (2005) Cloning, biochemical, and phylogenetic characterizations of γ -glutamylcysteine synthetase from Anabena sp. PCC 7120. Plant Cell Physiol. 46, 557–562 35 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 36 Okumura, N., Masamoto, K. and Wada, H. (1997) The gshB gene in the cyanobacterium Synechococcus sp. PCC 7942 encodes a functional glutathione synthetase. Microbiology 143, 2883–2890 37 Cameron, J. C. and Pakrasi, H. B. (2011) Glutathione in Synechocystis 6803: a closer look into the physiology of a gshB mutant. Plant Signaling Behav. 6, 89–92 38 Cameron, J. C. and Pakrasi, H. B. (2011) Glutathione facilitates antibiotic resistance and photosystem I stability during exposure to gentamicin in cyanobacteria. Appl. Environ. Microbiol. 77, 3547–3550 39 Brekken, D. L. and Phillips, M. A. (1998) Trypanosoma brucei γ -glutamylcysteine synthetase: characterization of the kinetic mechanism and the role of Cys-319 in cystamine inactivation. J. Biol. Chem. 273, 26317–26322 40 Hara, T., Kato, H., Katsube, Y. and Oda, J. (1996) A pseudo-michaelis quaternary complex in the reverse reaction of a ligase: structure of Escherichia coli B glutathione synthetase complexed with ADP, glutathione, and sulfate at 2.0 Å resolution. Biochemistry 35, 11967–11974 41 Orlowski, M. and Meister, A. (1971) Partial reactions catalyzed by γ -glutamylcysteine synthetase and evidence for an activated glutamate intermediate. J. Biol. Chem. 246, 7095–7105 42 Kelly, B. S., Antholine, W. E. and Griffith, O. W. (2002) Escherichia coli γ -glutamylcysteine synthetase: two active site metal ions affect substrate and inhibitor binding. J. Biol. Chem. 277, 50–58 43 Fahey, R. C. (2012) Glutathione analogs in prokaryotes. Biochim. Biophys. Acta, doi:10.1016/j.bbagen.2012.10.006 44 Richman, P. G. and Meister, A. (1975) Regulation of γ -glutamyl-cysteine synthetase by nonallosteric feedback inhibition by glutathione. J. Biol. Chem. 250, 1422–1426 45 Tanaka, T., Kato, H., Nishioka, T. and Oda, J. (1992) Mutational and proteolytic studies on a flexible loop in glutathione synthetase from Escherichia coli B: the loop and arginine 233 are critical for the catalytic reaction. Biochemistry 31, 2259–2265 46 Kato, H., Tanaka, T., Yamaguchi, H., Hara, T., Nishioka, T., Katsube, Y. and Oda, J. (1994) Flexible loop that is novel catalytic machinery in a ligase: atomic structure and function of the loopless glutathione synthetase. Biochemistry 33, 4995–4999 47 Biterova, E. I. and Barycki, J. J. (2010) Structural basis for feedback and pharmacological inhibition of Sacchromyces cerevisiae glutmate cysteine ligase. J. Biol. Chem. 285, 14459–14566 48 Abbott, J. J., Ford, J. L. and Phillips, M. A. (2002) Substrate binding determinants of Trypanosoma brucei γ -glutamylcysteine synthetase. Biochemistry 41, 2741–2750 49 Liaw, S. H. and Eisenberg, D. (1994) Structural model for the reaction mechanims of glutamine synthetase, based on five crystal structures of enzyme-substrate complexes. Biochemistry 33, 675–681 50 Krajewski, W. W., Jones, T. A. and Mowbray, S. L. (2005) Structure of Mycobacterium tuberculosis glutamine synthetase in complex with a transition-state mimic provides functional insights. Proc. Natl. Acad. Sci. U.S.A. 102, 10499–10504 51 Segal, I. H. (1975) Enzyme Kinetics: Behavior and Analysis of Rapid Equilibrium and Steady-State Enzyme Systems, John Wiley & Sons, New York 52 Cooper, B. F. and Rudolph, F. B. (1995) Product inhibition applications. Methods Enzymol. 249, 188–211 53 Emanuele, J. J., Jin, H., Yanchunas, J. and Villafranca, J. J. (1997) Evaluation of the kinetic mechanism of Escherichia coli uridine diphosphate-N -acetylmuramate:L-alanine ligase. Biochemistry 36, 7264–7271 54 Meek, T. D. and Villafranca, J. J. (1980) Kinetic mechanism of Escherichia coli glutamine synthetase. Biochemistry 19, 5513–5519 55 Mullins, L. S., Zawadzke, L. E., Walsh, C. T. and Raushel, F. M. (1990) Kinetic evidence for the formation of D-alanyl phosphate in the mechanism of D-alanyl-D-alanine ligase. J. Biol. Chem. 265, 8993–8998 56 Anderson, M. S., Eveland, S. S., Onishi, H. R. and Pompliano, D. L. (1996) Kinetic mechanism of the Escherichia coli UDPMurNAc-tripeptide D-alanyl-D-alanine-adding enzyme: use of a glutathione S-transferase fusion. Biochemistry 35, 16264–16269 57 Wang, B., Wang, J., Zhang, W. and Meldrum, D. R. (2012) Application of synthetic biology in cyanobacteria and algae. Front. Microbiol. 3, 344 58 Machado, I. M. and Atsumi, S. (2012) Cyanobacterial biofuel production. J. Biotechnol. 162, 50–56 59 Rosgaard, L., de Porcellinis, A. J., Jacobsen, J. H., Frigaard, N. U. and Sakuragi, Y. (2012) Bioengineering of carbon fixation, biofuels, and biochemicals in cyanobacteria and plants. J. Biotechnol. 162, 134–147 c The Authors Journal compilation c 2013 Biochemical Society 72 W. B. Musgrave and others 60 Nogales, J., Gudmundsson, S., Knight, E. M., Palsson, B. O. and Thiele, I. (2012) Detailing the optimality of photosynthesis in cyanobacteria through systems biology analysis. Proc. Natl. Acad. Sci. U.S.A. 109, 2678–2683 61 Jones, C. S. and Mayfield, S. P. (2012) Algae biofuels: versatility for the future of bioenergy. Curr. Opin. Biotech. 23, 346–351 62 Singh, A. K., Elvitigala, T., Cameron, J. C., Ghosh, B. K., Bhattacharyya-Pakrasi, M. and Pakrasi, H. B. (2010) Integrative analysis of large scale expression profiles reveals core transcriptional response and coordination between multiple cellular processes in a cyanobacterium. BMC Syst. Biol. 4, 105 Received 24 August 2012/20 November 2012; accepted 21 November 2012 Published as BJ Immediate Publication 21 November 2012, doi:10.1042/BJ20121332 c The Authors Journal compilation c 2013 Biochemical Society 63 Allen, S. A., Clark, W., McCaffery, J. M., Cai, Z., Lanctot, A., Slininger, P. J., Liu, Z. L. and Gorisch, S. W. (2010) Furfural induces reactive oxygen species accumulation and cellular damage in Saccharomyces cerevisiae . Biotech. Biofuels 3, 2 64 Liu, J., Chen, L., Wang, J., Qiao, J. and Zhang, W. (2012) Proteomic analysis reveals resistance mechanism against biofuel hexane in Synechocystis sp. PCC 6803. Biotechnol. Biofuels 5, 68 65 Zhu, L., Dong, H., Zhang, Y. and Li, Y. (2011) Engineering the robustness of Clostridium acetobutylicum by introducing glutathione biosynthetic capacity. Metab. Eng. 13, 426–434 Biochem. J. (2013) 450, 63–72 (Printed in Great Britain) doi:10.1042/BJ20121332 SUPPLEMENTARY ONLINE DATA Probing the origins of glutathione biosynthesis through biochemical analysis of glutamate-cysteine ligase and glutathione synthetase from a model photosynthetic prokaryote William B. MUSGRAVE, Hankuil YI, Dustin KLINE, Jeffrey C. CAMERON, Jonathan WIGNES, Sanghamitra DEY, Himadri B. PAKRASI and Joseph M. JEZ1 Department of Biology, Washington University, One Brookings Drive, Campus Box 1137, St. Louis, MO 63130, U.S.A. Figure S1 SDS/PAGE analysis of SynGCL and SynGS Samples were stained for total protein using Coomassie Blue. Lanes 1 (MW), 2 (A) and 3 (B) show the molecular marker (molecular masses in kDa to the left-hand side), purified SynGCL and purified SynGS respectively. 1 To whom correspondence should be addressed (email [email protected]). c The Authors Journal compilation c 2013 Biochemical Society W. B. Musgrave and others Figure S2 Initial velocity analysis of the kinetic mechanism of SynGS Experimental data is indicated by the symbols in the reciprocal plots of the initial velocity experiments. The lines shown represent the global fit of all data to the equation for a random ter-reactant system. Parameters of the fit are listed in Table 4 of the main text. The following conditions were used for each set of assays: (A) 0.5 mM γ -glutamylcysteine (γ EC); 0.12–1.2 mM ATP; and 0.12 (䉱), 0.175 (䊐), 0.35 (䊏), 0.7 (䊊) or 1.2 (䊉) mM glycine; (B) 0.5 mM γ -glutamylcysteine; 0.12–1.2 mM glycine; and 0.12 (䉱), 0.175 (䊐), 0.35 (䊏), 0.7 (䊊) or 1.2 (䊉) mM ATP; (C) 1.2 mM glycine; 0.12–1.2 mM ATP; and 0.05 (䉱), 0.075 (䊐), 0.15 (䊏), 0.3 (䊊) or 0.45 (䊉) mM γ -glutamylcysteine; (D) 1.2 mM glycine; 0.05–0.45 mM γ -glutamylcysteine; and 0.12 (䉱), 0.175 (䊐), 0.35 (䊏), 0.7 (䊊) or 1.2 (䊉) mM ATP; (E) 1.2 mM ATP; 0.05–0.45 mM γ -glutamylcysteine; and 0.12 (䉱), 0.175 (䊐), 0.35 (䊏), 0.7 (䊊) or 1.2 (䊉) mM glycine; and (F) 1.2 mM ATP; 0.12–1.2 mM glycine; and 0.05 (䉱), 0.075 (䊐), 0.15 (䊏), 0.3 (䊊) or 0.45 (䊉) mM γ -glutamylcysteine. Received 24 August 2012/20 November 2012; accepted 21 November 2012 Published as BJ Immediate Publication 21 November 2012, doi:10.1042/BJ20121332 c The Authors Journal compilation c 2013 Biochemical Society