Background: Sequence-based annotation of novel-type phosphoserine phosphatases (iPSPs) is difficult because of their diverse substrate preferences.
Results: Crystal structure and mutation analyses of iPSP1 from Hydrogenobacter thermophilus identified the residues involved in substrate recognition.
Conclusion: The His85 side chain and C-terminal region of iPSP are important for its phosphoserine phosphatase activity.
Significance: These findings will help to identify iPSPs from sequence databases.
Keywords: Amino Acid, Bacterial Metabolism, Crystal Structure, Metabolism, Phosphatase, Amino Acid Biosynthesis, Autotroph, Histidine Phosphatase Superfamily, Substrate Specificity, Phosphoserine Phosphatase
Abstract
Novel-type serine-synthesizing enzymes, termed metal-independent phosphoserine phosphatases (iPSPs), were recently identified and characterized from Hydrogenobacter thermophilus, a chemolithoautotrophic bacterium belonging to the order Aquificales. iPSPs are cofactor-dependent phosphoglycerate mutase (dPGM)-like phosphatases that have significant amino acid sequence similarity to dPGMs but lack phosphoglycerate mutase activity. Genes coding dPGM-like phosphatases have been identified in a broad range of organisms; however, predicting the function of the corresponding proteins based on sequence information alone is difficult due to their diverse substrate preferences. Here, we determined the crystal structure of iPSP1 from H. thermophilus in the apo-form and in complex with its substrate l-phosphoserine to find structural units important for its phosphatase activity toward l-phosphoserine. Structural and biochemical characterization of iPSP1 revealed that the side chains of His85 and C-terminal region characteristic of iPSP1 are responsible for the PSP activity. The importance of these structural units for PSP activity was confirmed by high PSP activity observed in two novel dPGM-like proteins from Cyanobacteria and Chloroflexus in which the two structural units were conserved. We anticipate that our present findings will facilitate understanding of the serine biosynthesis pathways of organisms that lack gene(s) encoding conventional PSPs, as the structural information revealed here will help to identify iPSP from sequence databases.
Introduction
Novel-type serine-synthesizing enzymes, termed metal-independent phosphoserine phosphatases (iPSPs5; EC 3.1.3.3; l-O-phosphoserine + H2O → l-serine + phosphate), were recently identified and characterized from Hydrogenobacter thermophilus TK-6, an obligately chemolithoautotrophic, thermophilic bacterium belonging to the order Aquificales (1–3). Although the kinetic parameters of iPSPs for l-phosphoserine are similar to those of conventional Mg2+-dependent phosphoserine phosphatases (dPSPs), iPSPs do not require divalent cations for their activity. In addition, iPSPs show no amino acid sequence homology to dPSPs. Therefore, the identification and properties of iPSPs suggest that these enzymes connect the missing link between phosphoserine and serine in the serine biosynthetic pathway of H. thermophilus, which lacks a gene encoding a conventional dPSP.
To date, two types of iPSPs have been identified in H. thermophilus, iPSP1 and iPSP2. The former is a homodimer of PspA subunits, and the latter is a heterodimer of PspA and PspB subunits. In contrast to dPSPs, which belong to the haloacid dehalogenase-like hydrolase superfamily, PspA and PspB belong to the histidine phosphatase superfamily, which is composed of a functionally diverse group of phosphatases and mutases, including cofactor-dependentphosphoglycerate mutases (dPGMs; EC 5.4.2.1), fructose-2,6-bisphosphatase (EC 3.1.3.46), and phytases (EC 3.1.3.–) (4). Members of the histidine phosphatase superfamily share four conserved catalytic core residues consisting of two histidine and two arginine residues (e.g. Arg9, His10, Arg59, and His151 in Bacillus stearothermophilus PhoE). In PhoE, His10 acts as the base to extract the phosphate group from the substrate, whereas the other three core residues interact electrostatically with the phosphate group during the reaction. In addition to the four conserved residues, an aspartate or glutamate residue (e.g. Glu83 in PhoE), which acts as a proton donor, is also required for the dephosphorylation reaction (4–6).
When compared with proteins with known function, PspA and PspB, which share 35% sequence identity, have the highest amino acid sequence similarity with dPGM (22 and 18% identities, respectively, to E. coli dPGM). In addition, the five catalytic core residues (four conserved residues and an additional aspartate/glutamate) that are characteristic of the histidine phosphatase superfamily are completely conserved in both iPSP subunits and dPGMs. For this reason, iPSPs were originally annotated as dPGMs; however, iPSPs lack the residues proposed to be important for mutase activity and show no phosphoglycerate mutase activity (1, 7, 8). As iPSPs exhibit substantial phosphatase activity specific to l-phosphoserine, these enzymes are now thought to function as PSPs in vivo.
Proteins with homology to dPGM, but without mutase activity, have been found in a broad range of organisms and are predicted to act as phosphatases, although their physiological roles remain unknown (8–10). For example, PhoE from B. stearothermophilus is a homodimeric phosphatase with broad substrate specificity, and it has the highest specificity toward 3-phosphoglyceric acid and α-naphthylphosphate (8). Our previous study revealed that the PspA and PspB subunits of iPSPs exhibit completely different substrate specificities from that of PhoE (1). Namely, PspA has strict substrate specificity toward l-phosphoserine, whereas PspB appears to have much lower phosphatase activity toward l-phosphoserine and higher activity toward p-nitrophenyl phosphate. We also found that PspA homologs from Thermus thermophilus and Arabidopsis thaliana had slight PSP activity. However, the T. thermophilus PspA homolog had higher phosphatase activity toward dl-phosphothreonine and 3-phosphoglycerate than toward l-phosphoserine (1). Taken together, these results indicate that dPGM-like phosphatases have diverse substrate preferences, and their function cannot be predicted based on the whole amino acid sequence similarity alone. Therefore, identification of the structural units in the PspA subunit of H. thermophilus iPSP1 that is important for the PSP activity are expected to enable the identification of iPSPs among dPGM-like proteins with increased accuracy.
Here, we report the crystal structures of the apo- and the substrate-bound forms of iPSP1. Based on these crystal structures and mutagenesis data, we determined the structural units that would aid in the identification of iPSPs among dPGM-like proteins.
EXPERIMENTAL PROCEDURES
Gene Cloning and Protein Expression and Purification
The genes encoding the PspA and PspB subunits of H. thermophilus TK-6 (IAM 12695, DSM 6534), pspA (YP_003431771, HTH0103), and pspB (YP_003431851, HTH0183) were cloned into the expression vectors pCDFDuet-1 and pET21c (Novagen, Darmstadt, Germany), respectively, as described previously (1). The constructed plasmids were then mutated for expressing H9A mutants of the PspA and PspB subunits in which His9 was converted to Ala, and H85A, ΔC1, and ΔC7 mutants of PspA, in which His85 was converted to Ala (iPSP1 (H85A)); Ile211 was deleted (iPSP1 (ΔC1)), and Val2050–Ile211 were deleted (iPSP1 (ΔC7)), using PrimeSTAR mutagenesis basal kit (Takara Bio, Otsu, Japan) according to the manufacturer's instructions. The genes encoding iPSP homologs from Roseiflexus castenholzii DSM 13941 (YP_001432425, RCAS2324; YP_001430323, RCAS0171; and YP_001433513, RCAS3445) and Thermosynechococcus elongatus BP-1 (NP_682322, TLR1532) were amplified by PCR from the genomic DNAs of respective species. The amplified fragments were inserted into the NdeI/XhoI (R. castenholzii ones) or NdeI/EcoRI (T. elongatus one) sites of pET-21c. The nucleotide sequences of the above-mentioned genes were verified by sequencing. The primers used in this study are listed in Table 1.
TABLE 1.
Primers used in this study
| Name | DNA sequence |
|---|---|
| PspA H9A F | 5′-GTGCGAGCTGCGGAGAGCGAGTGGAA-3′ |
| PspA H9A R | 5′-CTCCGCAGCTCGCACCAGTATAAGCTT-3′ |
| PspB H9A F | 5′-GGTCAGGGCTGCGCAGAGCGAGTATAA-3′ |
| PspB H9A R | 5′-TGCGCAGCCCTGACCAAATACAAACG-3′ |
| PspA H85A F | 5′-ATAGACGCCGGTATGTGGTCGGGC-3′ |
| PspA H85A R | 5′-CATACCGGCGTCTATTTCTATTATC-3′ |
| PspA ΔC1 F | 5′-AAGGCTTAGTAGCTTGCTGGCATG-3′ |
| PspA ΔC1 R | 5′-AAGCTACTAAGCCTTGTGAGCTTC-3′ |
| PspA ΔC7 F | 5′-GTTTTATTAGCTTGCTGGCATGTAC-3′ |
| PspA ΔC7 R | 5′-GCAAGCTAATAAAACTCTCCCAGATG-3′ |
| TLR1532 F | 5′-GTCCATATGGCACAGTGGG-3′ |
| TLR1532 R | 5′-ACAGAATTCCCTATAGCGCTCC-3′ |
| RCAS2324 F | 5′-ATGCATATGCGCCTGATCATC-3′ |
| RCAS2324 R | 5′-CGTCTCGAGTTCAACCTTTAC-3′ |
| RCAS0171 F | 5′-TATCCATATGCGCACCGCTC-3′ |
| RCAS0171 R | 5′-GGTCTCGAGGGAAGGTCAAAG-3′ |
| RCAS3445 F | 5′-AACCATATGACCACCTTCTACATC-3′ |
| RCAS3445 R | 5′-TCGCTCGAGTTCTCATACCC-3′ |
Escherichia coli BL21-Codon Plus (DE3)-RIL was used for the individual and co-expression of the PspA and PspB subunits, and their mutants. E. coli BL21 (DE3) was used for the expression of iPSP homologs from R. castenholzii and T. elongatus. The transformed E. coli cells harboring the expression plasmids were inoculated into Luria-Bertani medium containing the appropriate antibiotic(s) as follows: 50 μg ml−1 streptomycin, 50 μg ml−1 ampicillin, and/or 34 μg ml−1 chloramphenicol. After cultivating the cells aerobically at 37 °C until the absorbance at 600 nm reached ∼0.6, the expressions of iPSP1, iPSP2, and their mutants were induced by the addition of 1 mm (final concentration) isopropyl thio-β-d-galactopyranoside to the medium, followed by further cultivation for 3 h at 37 °C. The expressions of the R. castenholzii and T. elongatus iPSP homologs were induced with 0.5 mm isopropyl thio-β-d-galactopyranoside for 4 h at 37 °C. The cells were then harvested by centrifugation, resuspended in 20 mm Tris-HCl buffer (pH 8.0) containing 1 mm EDTA (4 ml/g of wet cells), sonicated, and subjected to centrifugation (100,000 × g for 1 h) to obtain supernatants. iPSP1s from H. thermophilus and its mutants were purified from the supernatants using butyl-Toyopearl (Tosoh, Tokyo, Japan) and Mono Q columns (GE Healthcare) after heat treatment at 80 °C for 10 min, as described previously (1). For the purification of iPSP homologs from R. castenholzii and T. elongatus, nonheat-treated supernatants were loaded onto a Q-Sepharose fast-flow column (GE Healthcare) and partially purified, as described previously (11).
Enzyme Assays
PSP activity was assayed by measuring the production of inorganic phosphate, as described previously with minor modifications (1). Briefly, the reaction mixture contained 10 μl of 1.0 m HEPES-NaOH (pH 8.0 at room temperature), 5 μl of 50 mm l-phosphoserine, 0.5 μl of 100 mm EDTA (pH 8.0), 5 μl of enzyme solution in 20 mm Tris-HCl (pH 8.0), and 29.5 μl of water (total volume = 50 μl). For the determination of kinetic parameters, reaction mixtures containing 0.5, 1.0, 3.0, 5.0, 10.0, and 20.0 mm l-phosphoserine were used. Note that in our previous study, the reaction mixture contained 0.5–10 mm l-phosphoserine for the determination of kinetic parameters (1) because of the limited amount of purified native enzymes. Here, we used up to 20 mm l-phosphoserine to more accurately determine the kinetic parameters. Therefore, the final buffer concentration of the reaction mixture was increased from 20 to 200 mm to maintain a constant pH. The reaction mixture was incubated for 5–20 min at 70 °C for H. thermophilus proteins and 55 °C for R. castenholzii and T. elongatus proteins, which were close to the optimal growth temperatures of the host organisms. One unit of PSP activity was defined as the amount of enzyme producing 1 μmol of inorganic phosphate/min.
Protein Assay
Protein concentrations were measured using Bradford protein assay (Bio-Rad) with bovine serum albumin as the standard.
Crystallization and Data Collection for iPSP1 and iPSP1 (H9A)-l-Phosphoserine
The crystallization and x-ray diffraction data collection of wild-type iPSP1 (iPSP1 (WT)) were performed as described previously (12). In brief, iPSP1 (WT) was crystallized by the sitting-drop vapor diffusion method. A 1.5-μl drop was first prepared by mixing 0.75 μl of an iPSP1 (WT) solution (10 mg ml−1 in 5 mm Tris-HCl (pH 8.0)) and a 0.75-μl reservoir solution (100 mm HEPES-NaOH (pH 7.5), 10% (v/v) 2-propanol, and 20% (w/v) polyethylene glycol 4000). Obtained crystals were cryoprotected by soaking in a 3:1 (v/v) mixture of reservoir solution and ethylene glycol. Two x-ray diffraction data sets from a single crystal of iPSP1 (WT) were collected as described previously (12). The obtained data sets were indexed, integrated, scaled, and merged with the XDS program (13).
iPSP1 (H9A) complexed with l-phosphoserine was crystallized by the sitting-drop vapor diffusion method. A 1.5-μl drop was first prepared by mixing 0.75 μl of an iPSP1 (H9A) solution (10 mg/ml in 5 mm Tris-HCl (pH 8.0) containing 12-molar excess of l-phosphoserine) and 0.75 μl of reservoir solution (100 mm sodium acetate trihydrate (pH 4.6), 20 mm calcium chloride dihydrate, and 30% (v/v) (±)-2-methyl-2,4-pentanediol). Obtained crystals were cryoprotected by soaking in a 4:1 (v/v) mixture of reservoir solution and ethylene glycol. X-ray diffraction experiments for iPSP1 (H9A)·l-phosphoserine crystals were performed at beamline NE3A at Photon Factory-Advanced Ring (Ibaraki, Japan). A data set for the iPSP1 (H9A)·l-phosphoserine crystal was collected using an ADSC Quantum 270 detector under the following conditions: wavelength, 1.0000 Å; distance between the crystal and the detector, 213.9 mm; oscillation angle, 1.0°; and exposure time, 1 s per image. The collected dataset was indexed, integrated, scaled, and merged with the programs XDS (13).
Structural Solution and Refinement of iPSP1 (WT) and iPSP1 (H9A)-l-Phosphoserine
The initial structure of iPSP1 (WT) was obtained by the molecular replacement method using the program Molrep (14) in the CCP4 suite. The crystal structure of PhoE from B. stearothermophilus (25% amino acid sequence identity with iPSP1; PDB code 1H2E) was used as the search model. The initial structure was refined using Coot (15) and Refmac5 in the CCP4 suite (16). The stereochemistry of the structure was checked by the program RAMPAGE (17). The initial structure of the iPSP1 (H9A)·l-phosphoserine complex was obtained by the molecular replacement method using the program Molrep in the CCP4 suite using iPSP1 (WT) as a template. l-Phosphoserine and water molecules were modeled in the final stages of refinement based on the Fobs − Fcalc electron density map. The refined structures were visualized with PyMOL (Version 1.5.0.1, Schrödinger, LLC.). The structural similarity between any two protomers was evaluated with the root mean square deviation of Cα atoms calculated with the DaliLite server (18). Interface analysis was performed using the Protein Interfaces, Surfaces, and Assemblies (PISA) server of the European Bioinformatics Institute (19–21).
Homology Modeling
The tertiary structure of the PspB subunit was modeled using Swiss-Model (automated mode) (22–24) and the crystal structure of chain A of iPSP1 (WT) as the template. The tertiary structures of TLR1532 and RCAS2324 were modeled using Modeler 9.10 and the same template structure as above. Here, the pairwise sequence alignments obtained with MUSCLE (25) were used (see Fig. 1). The residues after Thr249 and entire amino acid sequence were used for the modeling of TLR1532 and RCAS2324, respectively.
FIGURE 1.
Multiple sequence alignment of PspA and its homologs constructed using the MUSCLE program. A, gaps after Gly201 (PspA subunit numbering) were manually removed. Amino acids surrounded by a black box and shaded indicate residues that can form hydrogen bonds with the phospho group and other parts of l-phosphoserine in PspA, respectively, and their corresponding amino acids conserved in other sequences. Residue numbers of PspA are shown above the sequences. B, alignment used for structure modeling. The alignments shown here, other than that of the C terminus, are the same as those shown in A.
Data Deposition
The atomic coordinates and structural factors of H. thermophilus iPSP1 have been deposited in the Protein Data Bank under PDB codes 4IJ5 and 4IJ6.
RESULTS
PSP Activity of H9A Mutants
Our previous studies suggested that the PSP activities of H. thermophilus iPSP1 (PspA-PspA) and iPSP2 (PspA-PspB) are mainly catalyzed by the PspA subunit, because the Vmax of iPSP2 is nearly half that of iPSP1, whereas their Km values are similar (1). To determine whether the PspB subunit had detectable PSP activity, we constructed H9A mutants of the PspA and PspB subunits, where His9, a putative active-site residue predicted to act as a base, was mutated to Ala.
The heat-treated culture supernatant of an E. coli strain overexpressing a homodimer of PspA (H9A)-PspA (H9A) showed almost no PSP activity (≤0.01 units mg−1, Table 2). We confirmed by SDS-PAGE analysis that the expression level of the mutant protein was similar to that of PspA (WT) (data not shown). This result strongly suggests that His9 of the PspA subunit acts as a base, as found in other histidine phosphatase superfamily proteins. PSP activity was also not detected in culture supernatant containing the PspA (H9A)-PspB (WT) heterodimer; however, culture supernatant from E. coli cells overexpressing the PspA (WT)-PspB (H9A) heterodimer (Fig. 2) exhibited PSP activity (39 ± 0 units mg−1; Table 2). These results indicate that the PspB subunit has no or undetectable PSP activity.
TABLE 2.
PSP activity of H9A mutants
| Component | PSP activity |
|---|---|
| units mg−1 | |
| PspA (H9A)-PspA (H9A) | ≤0.01 |
| PspA (WT)-PspB (H9A) | 39 ± 0 |
| PspA (H9A)-PspB (WT) | ≤0.01 |
| PspA (H9A)-PspB (H9A) | ≤0.01 |
FIGURE 2.

12% SDS-PAGE analysis of partially purified iPSPs used for the enzymatic activity measurements. Lane 1, Q-Sepharose fraction of RCAS2324; lane 2, Q-Sepharose fraction of TLR1532; lanes 3–5, supernatant of iPSP2-expressed and heat-treated CFE. Lane 3, PspA (WT)·PspB (H9A); lane 4, PspA (H9A)·PspB (WT); lane 5, PspA (H9A)·PspB (H9A). M, molecular mass markers. The arrowheads indicate the bands corresponding to the expressed proteins.
Structural Determination of iPSP1 (WT) without a Ligand and iPSP1 (H9A) Complexed with l-Phosphoserine
The crystal structure of iPSP1 (WT) with no ligand was solved at 1.5-Å resolution by molecular replacement using the atomic coordinates of PhoE from B. stearothermophilus as a search model. The iPSP1 crystal belonged to the orthorhombic space group P212121 with unit cell parameters a = 49.8 Å, b = 73.6 Å, and c = 124.3 Å, and it contained two PspA subunits in an asymmetric unit. The final R and Rfree values of iPSP1 (WT) were 18.4 and 19.7%, respectively. The Ramachandran plot indicated that 98.6 and 1.4% of the residues were in the most favored and allowed regions, respectively. The data collection and refinement statistics are presented in Table 3.
TABLE 3.
Crystallographic data collection and refinement statistics
| Protein name | iPSP1 (WT) | iPSP1 (H9A)·l-phosphoserine |
|---|---|---|
| X-ray data collection | ||
| Beamline | SPring-8 BL-32XU | PF-AR NE3A |
| Wavelength | 1.0000 Å | 1.0000 Å |
| Resolutiona | 45.0 to 1.50 Å (1.54 to 1.50 Å) | 38.2 to 1.80 Å (1.85 to 1.80 Å) |
| Space group | P212121 | P212121 |
| Unit cell parameter | a 49.8 Å, b 73.6 Å, c 124.3 Å | a 49.3 Å, b 68.9 Å, c 123.6 Å |
| VM | 2.32 Å3 Da−1 | 2.14 Å3 Da−1 |
| Total reflectionsa | 630,969 (32,454) | 277,086 (16,180) |
| Unique reflectionsa | 73,612 (5300) | 39,566 (2740) |
| Redundancya | 8.6 (6.1) | 7.0 (5.9) |
| Completenessa | 99.9% (99.0%) | 99.3% (94.1%) |
| Rsyma,b | 9.1% (58.7%) | 4.2% (16.6%) |
| 〈I/σI〉a | 14.8 (3.3) | 33.6 (10.4) |
| Refinement | ||
| Protein Data Bank code | 4IJ5 | 4IJ6 |
| Resolution | 36.1 to 1.50 Å | 38.2 to 1.80 Å |
| Rwork/Rfreec | 18.4/19.7% | 19.0/21.8% |
| Root mean square deviations | ||
| Bonds | 0.0084 Å | 0.0053 Å |
| Angles | 1.2716° | 1.0986° |
| No. of atoms (average B-factors (Å2)) | ||
| Protein | 3481 (16.0) | 3365 (17.3) |
| l-Phosphoserine | 22 (11.5) | |
| Water | 242 (23.2) | 147 (23.9) |
| Ethylene glycol | 16 (22.3) | 28 (23.1) |
| Chloride | 1 (13.0) | 2 (19.8) |
| Observed residue range(s) | ||
| Chain A | 1 to 211 | 1 to 207 |
| Chain B | 1 to 211 | 1 to 203 |
| Ramachandran plot | ||
| Most favored region | 98.6% | 98.0% |
| Additionally favored regions | 1.4% | 2.0% |
a Values for the highest resolution shells are shown in parentheses for x-ray data collection.
b Rsym = Σhkl ((Σi|Ii − 〈I〉|)/Σi|Ii|).
c Rwork = (Σhkl‖Fo| − |Fc‖)/Σhkl|Fo|. Rwork was calculated with 95% of the data used for refinement, and Rfree was calculated with the remaining 5% of the data.
The crystal structure of the iPSP1 (H9A)·l-phosphoserine complex was solved at 1.8-Å resolution by molecular replacement using the atomic coordinates of iPSP1 (WT) as a template. The iPSP1 (H9A)·l-phosphoserine complex crystal belonged to the orthorhombic space group P212121 with unit cell parameters a = 49.3 Å, b = 68.9 Å, and c = 123.6 Å, and it consisted of two PspA (H9A) subunit l-phosphoserine complexes in an asymmetric unit. The final R and Rfree values of iPSP1 (H9A)·l-phosphoserine were 19.0 and 21.8%, respectively. The Ramachandran plot indicated that 98.0 and 2.0% of the residues were in the most favored and allowed regions, respectively. The data collection and refinement statistics are shown in Table 3.
Overall Structures of iPSP1s
The final model of iPSP1 (WT) contained all of the amino acid residues (residues 1–211) for the two PspA subunits (chains A and B) in an asymmetric unit, and it consisted of 3481 protein atoms, one chloride ion, four ethylene glycol molecules, and 242 water molecules. The final model of the iPSP1 (H9A)·l-phosphoserine complex contained all of the amino acid residues, with the exception of several C-terminal residues (residues 1–207 and 1–203 for chains A and B, respectively), and it consisted of 3365 protein atoms, two chloride ions, seven ethylene glycol molecules, and 147 water molecules.
The PspA subunit adopted a dPGM-like fold (6, 26) containing six α-helices and six β-strands that formed an α/β/α sandwich of the central mixed β-sheet flanked by α-helices on both sides (Fig. 3, A and B). An additional smaller domain was predominantly composed of two α-helices (α3 and α4). Approximately 11% of the surface area of PspA (1,082 Å2 per protomer) was buried in the dimer interface, 40% of which was occupied by hydrophobic residues. At the dimer interfaces that were mainly formed by the β6 strand and a loop connecting α6 and β5, 24 hydrogen bonds, eight salt bridges, and a disulfide bond were predicted to exist as well as hydrophobic interactions (Fig. 3A and Table 4). These findings were consistent with the gel filtration data of iPSP1 indicating that the protein behaves as a homodimer of the PspA subunit in solution (1). The dimerization behavior of the PspA subunit in iPSP1 was similar to that of the dPGM-like protein Rv3124 from Mycobacterium tuberculosis (PDB code 2A6P) (27).
FIGURE 3.
Structure of iPSP1 and its homologs. Overall structure of homodimeric iPSP1 (WT; A) and its protomer (B). The conserved catalytic core residues (Arg8, His9, Arg58, Glu82, and His150) are shown as sticks. C, superimposition of active-site residues of iPSP1 (WT) (gray) and PhoE (PDB code 1H2E) (cyan). The PspA subunit numberings are shown. D, Fobs − Fcalc omit map of iPSP1 (H9A)·l-phosphoserine. E, superimposition of iPSP1 (WT) (gray) and iPSP1 (H9A)·l-phosphoserine (pink). F, 2Fobs − Fcalc electron density map contoured at 1.0σ (blue mesh) of the C-terminal residues region (residues 205–211) of iPSP1 (WT) (shown as sticks). l-Phosphoserine, also shown as sticks, is modeled by superposing iPSP1 (WT) and iPSP1 (H9A) complexed with l-phosphoserine. G, superimposition of d-phosphoserine onto the structure of PspA (H9A)·l-phosphoserine. Carbon atoms of iPSP1 (WT), iPSP1 (H9A), and d- and l-phosphoserine are shown in gray, pink, cyan, and light green, respectively. Phosphorus, oxygen, and nitrogen atoms are shown in orange, red, and blue, respectively. Images were generated with PyMOL.
TABLE 4.
Interface of iPSP1 (WT) homodimer
| Chain A | Chain B | Distance |
|---|---|---|
| Å | ||
| Disulfide bond | ||
| Cys198/Sγ | Cys198/Sγ | 2.07 |
| Hydrogen bonds | ||
| Val108/O | Arg188/Nη2 | 2.85 |
| Glu109/O | Arg188/Nη2 | 3.82 |
| Leu162/O | Lys168/Nζ | 2.75 |
| Ser171/O | Ile191/N | 2.81 |
| Asp175/Oδ2 | Lys193/Nζ | 2.93 |
| Asn189/O | Ser171/Oγ | 2.80 |
| Ile191/O | Gly173/N | 2.87 |
| Leu192/O | Asn195/Nδ2 | 2.79 |
| Leu194/O | Leu194/N | 3.12 |
| Asn195/O | Lys193/Nζ | 2.82 |
| Thr197/Oγ1 | Lys193/Nζ | 2.87 |
| Arg188/Nη2 | Val108/O | 2.74 |
| Arg187/Nη2 | Glu109/Oϵ2 | 3.08 |
| Arg188/Nη2 | Glu109/O | 3.71 |
| Lys168/Nζ | Leu162/O | 2.79 |
| Ile191/N | Ser171/O | 2.82 |
| Lys193/Nζ | Asp175/Oδ2 | 2.90 |
| Ser171/Oγ | Asn189/O | 2.77 |
| Gly173/N | Ile191/O | 2.83 |
| Asn195/Nδ2 | Leu192/O | 2.83 |
| Leu194/N | Leu194/O | 3.09 |
| Lys193/Nζ | Asn195/O | 2.74 |
| Lys193/Nζ | Thr197/Oγ1 | 2.84 |
| Salt bridges | ||
| Glu109/Oϵ1 | Arg187/Nη2 | 3.74 |
| Glu109/Oϵ1 | Arg187/Nη1 | 3.81 |
| Glu109/Oϵ2 | Arg187/Nη1 | 3.64 |
| Glu109/Oϵ2 | Arg187/Nϵ | 3.98 |
| Asp175/Oδ2 | Lys193/Nζ | 2.93 |
| Arg187/Nη2 | Glu109/Oϵ2 | 3.08 |
| Lys193/Nζ | Asp175/Oδ2 | 2.90 |
The root mean square deviations between PspA (WT), PspA (H9A), and PhoE ranged from 0.2 to 1.6 Å, suggesting that their overall structures were nearly identical. The spatial positions and orientations were highly conserved between the essential residues of the PspA subunit (Arg8, His9, Asn15, Gln21, Gly22, Arg58, Glu82, and His150) and those of PhoE (Arg9, His10, Asn16, Gln22, Gly23, Arg59, Glu83, and His151) (Fig. 3C).
Binding Manner of l-Phosphoserine to iPSP1
Electron densities of l-phosphoserine were observed at the catalytic pocket when the inactive iPSP 1 (H9A) mutant was crystallized in the presence of this substrate (Fig. 3D). In the crystal structure of the enzyme·substrate complex, oxygen atoms of the carboxyl group of l-phosphoserine formed hydrogen bonds with main-chain amides of Gln21 and Gly22, and Nϵ2 of Gln21, and partly formed a hydrogen or an ionic bond with Nϵ2 of His85. Nitrogen atom of amino group of l-phosphoserine formed a hydrogen or an ionic bond with oxygen atoms of the side-chain carboxyl group of Glu82, and formed hydrogen bonds with Nϵ2 of His85 and Oγ1 of Thr151. As the phosphate group of l-phosphoserine was considered to be negatively charged, the crystal structure of the iPSP1 (H9A)·l-phosphoserine complex indicated that this group interacted with iPSP1 residues as follows: (i) O1P of the phosphate group formed an ionic bond with Nη2 of Arg8, and hydrogen bonds with side chains Nδ2 of Asn15 and Nϵ2 of Gln21; (ii) O2P of the phosphate group formed an ionic bond with Nϵ and Nη1 of Arg58 and formed a hydrogen or an ionic bond with Nϵ2 of Glu21 and Nδ1 of His150; and (iii) O3P of the phosphate group formed an ionic bond with Nϵ of Arg8 and a hydrogen bond with main-chain amide of Thr151 (Fig. 3E and Table 5).
TABLE 5.
Hydrogen bonds formed between iPSP1 (H9A) and l-phosphoserine
| Molecule/atom name | Molecule/residue/Atom name | Distance |
|---|---|---|
| Å | ||
| l-Phosphoserine/O1P | PspA/Arg8/Nη2 | 2.99 |
| l-Phosphoserine/O1P | PspA/Asn15/Nδ2 | 2.88 |
| l-Phosphoserine/O1P | PspA/Gln21/Nϵ2 | 2.99 |
| l-Phosphoserine/O2P | PspA/Gln21/Nϵ2 | 3.44 |
| l-Phosphoserine/O2P | PspA/Arg58/Nη1 | 3.01 |
| l-Phosphoserine/O2P | PspA/Arg58/Nϵ | 3.40 |
| l-Phosphoserine/O2P | PspA/His150/Nδ1 | 3.70 |
| l-Phosphoserine/O3P | PspA/Arg8/Nϵ | 3.12 |
| l-Phosphoserine/O3P | PspA/Thr151/N | 2.97 |
| l-Phosphoserine/Oγ | PspA/Thr151/Oγ1 | 2.84 |
| l-Phosphoserine/N | PspA/His85/Nϵ2 | 3.09 |
| l-Phosphoserine/N | PspA/Glu82/Oϵ2 | 2.84 |
| l-Phosphoserine/N | PspA/Thr151/Oγ1 | 3.23 |
| l-Phosphoserine/OXT | PspA/Gln21/N | 2.92 |
| l-Phosphoserine/OXT | PspA/Gln21/Nϵ2 | 2.83 |
| l-Phosphoserine/O | PspA/Gly22/N | 2.96 |
| l-Phosphoserine/O | PspA/His85/Nϵ2 | 3.40 |
| l-Phosphoserine/O | PspA/Arg58/Nη1 | 3.95 |
To predict the substrate-binding manner of iPSP1 (WT), we modeled the structure of iPSP (WT) in complex with l-phosphoserine by superposing the crystal structure of iPSP1 (WT) onto that of iPSP1 (H9A) (Fig. 3E). The modeled structure of the protein·ligand complex suggested that hydrogen bonds could be formed by oxygen atom of the phosphate group of l-phosphoserine and Nϵ2 of His9, in addition to the interactions observed in the crystal structure of the iPSP1 (H9A)·l-phosphoserine complex. The iPSP1 (WT)·substrate model suggested that the invariant catalytic residues (Arg8, His9, Arg58, and His150; PspA subunit numbering) and the proton donor Glu (Glu82) in dPGMs and their relatives are also essential for the phosphatase activity of iPSP1.
Multiple amino acid sequence alignment indicated that the PspA subunit residues that interact with l-phosphoserine were highly conserved in the PspB subunit and PhoE (Fig. 1A). Among the PspA residues involved in forming hydrogen and/or ionic bonds with the phosphate group of l-phosphoserine (Arg8, His9, Asn15, Gln21, Arg58, His150, and Thr151), only Thr151 was replaced by Gly in both the PspB subunit and PhoE. These findings suggest that the phosphate group binding and dephosphorylation mechanisms of the PspA subunit in iPSP1 are similar to those of the PspB subunit, PhoE, and other proteins in the histidine phosphatase superfamily. The alignment analysis also indicated that the residues forming hydrogen and/or ionic bonds with the l-serine moiety of the substrate were also conserved in PspB and PhoE, with the exception of His85, which was replaced by Phe85 and Leu86 in PspB and PhoE, respectively. Thus, His85 was suggested to be important for the substrate specificity of iPSP1.
The crystal structure-based modeling also revealed that several C-terminal residues in iPSP1 (WT) formed a lid covering the active site cleft. In the iPSP1 (WT)·l-phosphoserine model, the C-terminal Ile211 residue in iPSP1 (WT) was found to be close enough to form a hydrophobic interaction with l-phosphoserine (Fig. 3F), whereas the orientation of the side chain of Ile211 in iPSP1 (H9A) complexed with l-phosphoserine was unclear because no electron density was observed for Ile211 in iPSP1 (H9A). The residue corresponding to Ile211 was absent in the PspB subunit and PhoE, whose C termini were seven and five residues shorter, respectively, than that of PspA (Fig. 1A), suggesting that the C-terminal residues of PspA, PspB, and PhoE interact differently with the bound substrate. Thus, the C-terminal residues of the PspA subunit were also suggested to be important for the PSP activity of iPSP1.
Kinetic Analyses of iPSP1 Mutants
The kinetic parameters of several iPSP1 mutants were determined and compared with those of iPSP1 (WT) to confirm the validity of the residues suggested to be important for the PSP activity based on the crystal structure analyses. The iPSP1 mutants, iPSP1 (H85A), iPSP1 (ΔC1), and iPSP1 (ΔC7), were purified by the same method as used for iPSP1 (WT), and their activities were then assayed using 0.5–20 mm l-phosphoserine (Table 6 and Fig. 4). iPSP1 (H85A) exhibited a markedly decreased affinity to l-phosphoserine (Km > 20 mm) when compared with iPSP1 (WT) (Km = 1.5 mm). In contrast, iPSP1 (ΔC1) displayed a drastically decreased Vmax (9.5 units mg protein−1) when compared with iPSP1 (WT) (217 units mg protein−1), whereas its Km was almost the same level as that of iPSP (WT). iPSP1 (ΔC7) showed the lowest PSP activity of the three examined mutants, displaying higher Km and lower Vmax values than iPSP (WT). The PSP activities of iPSP1 (H85A), iPSP1 (ΔC1), and iPSP1 (ΔC7) in the presence of 5 mm l-phosphoserine (15 ± 0.1, 5.7 ± 0.1, and 1.3 ± 0.1 units mg protein−1, respectively) were markedly lower than that of iPSP1 (WT) (173 ± 9 units mg protein−1). These results confirmed that both His85 and the C-terminal residues were critical for the PSP activity of iPSP1.
TABLE 6.
Kinetic parameters for wild-type and mutant iPSP1
| iPSP1 | Km | Vmax | kcat/Km |
|---|---|---|---|
| mm | units mg−1 | s−1 mm−1 | |
| WT | 1.5 ± 0.1 | 217 ± 5 | 98 |
| H85A | ≥20 | 173 ± 52 | |
| ΔC1 | 3.3 ± 0.1 | 9.5 ± 0.1 | 2.0 |
| ΔC7 | 15.2 ± 1.4 | 5.2 ± 0.3 | 0.24 |
FIGURE 4.
Kinetic plots of iPSP1 and its mutants for l-phosphoserine.
Detection of PSP Activity in iPSP Homologs
Proteins with high amino acid sequence similarity to the PspA subunit of iPSP1, namely TLR1532 from T. elongatus and RCAS0170, RCAS2324, and RCAS3445 from R. castenholzii (Fig. 1A), were expressed in E. coli BL21 (DE3) and subjected to PSP activity assays. TLR1532, RCAS2324, and RCAS3445 were obtained in the soluble fraction, whereas RCAS0170 formed inclusion bodies. PSP activity was detected in the cell-free extracts of TLR1532- and RCAS2324-expressing E. coli but was not observed in that of RCAS3445-expressing E. coli. Partially purified TLR1532 and RCAS2324 (8.9- and 2.3-fold purified, respectively, from E. coli cell-free extract; Fig. 2) showed PSP activities of 63 ± 1 and 9.8 ± 1 units mg protein−1, respectively, suggesting that these proteins possessed substantial PSP activities.
DISCUSSION
In this study, we solved the crystal structures of the substrate-free and -bound forms of iPSP1 from H. thermophilus. An inactive mutant iPSP1 (H9A) was used for the structural analysis of the bound form. Moreover, we identified that two structural units are necessary for the PSP activity of iPSP1 based on both structural and functional data.
His85 in the PspA subunit was not conserved in the PspB subunit nor in PhoE and was found to be critical for the high PSP activity of iPSP1, because iPSP1 (H85A) possesses only 8.7% of the PSP activity of iPSP1 (WT) in the presence of 5 mm l-phosphoserine. His85 very likely contributes to substrate affinity, as iPSP1 (H85A) has a markedly increased Km value compared with that of iPSP1 (WT). The successful determination of the crystal structure of iPSP1 (H9A)·l-phosphoserine has demonstrated that the Nϵ2 atom of His85 forms a hydrogen bond with the nitrogen of l-phosphoserine and allows us to speculate on the molecular mechanism of l-phosphoserine recognition by iPSP1. The imidazole group of His85 is perpendicular to the aromatic ring of Phe116, which is suggested to form an aromatic hydrogen bond with hydrogenated Nδ1 of His85 (28). In addition, as the amino group of l-phosphoserine was positively charged (−NH3+) around physiological pH values (29, 30), the nonhydrogenated Nϵ2 of His85 may be appropriately oriented for recognizing the amino group of l-phosphoserine by the interaction between the hydrogenated Nδ1 of His85 and the side chain of Phe116. It is possible that the PspB subunit in iPSP2 lacks iPSP activity due to the replacement of His85 with Phe85 in this subunit, because the side chain of Phe85 cannot form a hydrogen bond with the amino group of l-phosphoserine.
The importance of His85 for substrate affinity also sheds light on the stereo selectivity of iPSP1. Our previous study revealed that iPSP1 has extremely low activity toward d-phosphoserine (1). This property can be explained by the formation of hydrogen bonds between the amino group of l-phosphoserine and the side chain of His85, and the carboxyl group of l-phosphoserine and the main-chain amides in Gln21 and Gly22, which are considered to contribute to substrate binding. When d-phosphoserine is superimposed onto the PspA (H9A)·l-phosphoserine structure by fitting the α-carbon and carboxyl groups of l- and d-phosphoserine, the amino group of d-phosphoserine cannot form a hydrogen bond with the side chain of His85 or any other residue (Fig. 3G). Therefore, the affinity of PspA to d-phosphoserine is expected to be lower than that to l-phosphoserine, resulting in a lower phosphatase activity to d-phosphoserine. This explanation is supported by the fact that iPSP1 also shows no activity toward 3-phosphoglyceric acid, in which the amino group of d-phosphoserine is substituted by a hydroxyl group (1).
The C-terminal region of the PspA subunit was also revealed to affect the PSP activity of iPSP1, because ΔC1 and ΔC7 mutants of iPSP1 have drastically reduced activities toward l-phosphoserine as a target substrate. The deletion of the C-terminal Ile211 residue in iPSP1 (ΔC1), which could interact with l-phosphoserine in the active cleft, clearly reduced Vmax but did not significantly affect Km values. These results suggest that Ile211 contributes to the PSP activity of iPSP1 by stabilizing the transition state or increasing the pre-exponential factor of the reaction. However, it remains unclear whether the deletion of a single C-terminal residue at position 211 or the substitution of it by a residue without a hydrophobic side chain can cause the drastic decrease of PSP activity. Although further study is needed, we speculate that not only the length of the C terminus but the hydrophobicity of the side chain may influence the activity of iPSPs because the C-terminal residues of all proteins from which substantial levels of PSP activity have been detected and iPSP1 orthologs from Aquificae (1) are Ile or Leu.
The C-terminal regions of histidine phosphatase superfamily proteins, including dPGM and fructose-2,6-bisphosphatase, have been studied because these regions protrude over the active-site cleft and are presumed to affect substrate selectivity (8, 31–33). Several C-terminal residues in iPSP1 also form a lid that covers the active site cleft, as shown in the wild-type structure (Fig. 5). This lid is expected to have a certain degree of flexibility, as in the case of dPGM (34), because the entrance to the active site of iPSP1 (WT) is too small for l-phosphoserine (Fig. 5) and the B-factor of the C terminus is relatively high. The C-terminal lid of iPSP1 likely functions to hold l-phosphoserine in the active site, as the Km value of iPSP1 (ΔC7), in which the lid is completely lost and the active site cleft is fully exposed to the solvent (Fig. 5), was increased ∼10-fold when compared with that of iPSP1 (WT). It has been reported that deletion of the C-terminal region (residues 235–246) of dPGM from Saccharomyces cerevisiae reduces the mutase activity but has no effect on phosphatase activity (31, 32). However, the C-terminal length of yeast dPGM was found to be 12 residues longer than that of PspA (data not shown). In addition, the phosphatase activity of dPGM is only a side reaction and is considerably lower than that of iPSP. Therefore, there is no conflict between the contribution of the C termini of dPGMs and iPSPs on their phosphatase activities.
FIGURE 5.
Entrance to the catalytic site of iPSP1 and its homologs. The structures of the PspA subunit (chain A of iPSP1 (WT)), modeled PspB, and PhoE were aligned and compared from the same angle. l-Phosphoserine was superimposed on each structure at the orientation determined using iPSP1 (H9A)·l-phosphoserine (chain A). The structures of PspA (ΔC1) and PspA (ΔC7) were constructed using the structure of PspA (WT). Images were generated with PyMOL.
The active site of the PspB subunit in iPSP2 is predicted to be wide open and to lack a lid based on the homology model constructed here using the crystal structure of the PspA subunit in iPSP1 (WT) as the template, because the C terminus of the PspB subunit is seven residues shorter than that of the PspA subunit (Fig. 5). The molecular model of the PspB subunit helps explain why PspB lacks PSP activity and has stronger phosphatase activity toward p-nitrophenyl phosphate than PspA (1); the wider entrance of the active-site pocket of PspB enables access to larger substrates, as in the case of PhoE (Fig. 5) (8).
Interestingly, dPGM-like proteins from both Cyanobacteria and Chloroflexi (TLR1532 and RCAS2324, respectively, which contain conserved histidine residues corresponding to His85 of iPSP1) exhibit substantial PSP activity. In contrast, dPGM-like proteins that lack this conserved histidine residue (PspB from H. thermophilus, RCAS3445, TTHA0368, At5g04120, and At3g05170 (1)) possess no activity or very weak PSP activity. Moreover, the substrate selectivity of PhoE, in which His85 is not conserved, markedly differs from that of iPSP1, although the PSP activity of PhoE has not been tested (1, 8). Although PhoE and dPGM-like proteins with low PSP activity (TTHA0368 and At5g04120) have a histidine residue at the position corresponding to Asp84 (adjacent to His85) of iPSP1, the crystal structure of PhoE (PDB code 1H2E) shows that the imidazole ring of this histidine residue is oriented in the opposite direction of the active core. Based on this finding, it is unlikely that the histidine residue at position 84 of PhoE, or other similar dPGM-like proteins, can form hydrogen bonds with the substrate.
We also determined that the C-terminal regions of TLR1532 and RCAS2324, which have high PSP activities, are similar to that of iPSP1 (Fig. 1B). These proteins contain conserved “-199HLG201-” and “-210A(I/L)211-COOH” (PspA subunit numbering) sequences that are connected by a loop consisting of 8–10 residues that are not conserved (Fig. 1). Homology models of TLR1532 and RCAS2324 suggest that the interactions between these conserved sequences and neighboring residues are similar to those that occur in iPSP1 (Fig. 6). In addition, the connecting loops are bound to the entrance region of the active sites by a comparable number of interactions as those identified in iPSP1, even though the sequences of the loops are not conserved. Therefore, the C-terminal Leu residue of TLR1532 and RCAS2324 will also be able to thrust into the active-site cleft in the same manner as the Ile residue of iPSP1. These observations strongly suggest that the C termini of dPGM-like proteins from Cyanobacteria and Chloroflexi also interact with the substrate, product, and/or the reaction intermediates in the active site and are important for PSP activity.
FIGURE 6.
C-terminal lids of iPSP1 (A), TLR1532 (B), and RCAS2324 (C).
The amino acid sequence identities of TLR1532 and RCAS2324 with the PspA subunit in iPSP1 are 32 and 36%, respectively, although that between the PspA and PspB subunits is 35%. The similar levels of sequence identity despite the differences in substrate specificity highlight the difficulty in predicting high PSP activity from the primary structure of a dPGM-like protein. However, our present findings indicate that the presence of His85 and the characteristic C-terminal sequence and/or length are useful indicators of high PSP activity of a dPGM-like protein. Using these indicators, we can predict that several Firmicutes species, in addition to most Cyanobacteria and Chloroflexi species, which lack conventional dPSP gene(s) (1), possess iPSP gene(s).
The identification of iPSP gene(s) is important for understanding the serine and glycine anabolic pathways that function in organisms lacking dPSP gene(s) (1). For example, nearly all Cyanobacteria species have an iPSP gene instead of a dPSP gene and synthesize serine from l-phosphoserine as indicated by 14C-labeling analysis (35). Therefore, iPSP may actually function as a serine-synthesizing enzyme in Cyanobacteria.
In conclusion, the crystal structures of iPSP1 (WT) and (H9A)·l-phosphoserine from H. thermophilus determined in this study and the kinetic analyses of mutated proteins have revealed that the presence of a His residue at position 85 and the sequence and/or length of the C-terminal region are both necessary for the PSP activity of iPSP1. These two structural features can be used as indicators to identify dPGM-like proteins with high PSP activity.
Acknowledgments
We thank Masahiko Ikeuchi (University of Tokyo), Shin Haruta, Katsumi Matsuura, and Hiromi Fukushima (Tokyo Metropolitan University) for providing T. elongatus and R. castenholzii cells. Homology modeling was supported by Tohru Terada and Kentaro Shimizu (University of Tokyo). We also thank Masafumi Kameya (Toyama Prefectural University) for frequent and fruitful discussions. The synchrotron radiation experiments were performed at beamline BL-32XU at SPring-8 (Hyogo, Japan) with the approval of the Targeted Proteins Research Program Office at RIKEN/SPring-8 and at beamline NE3A at Photon Factory-Advanced Ring (Ibaraki, Japan) under the approval of the Photon Factory Program Advisory Committee (Proposal No. 2008S2-001).
This work was supported in part by the Targeted Proteins Research Program of the Ministry of Education, Culture, Sports, Science, and Technology of Japan, the Platform for Drug Discovery, Informatics and Structural Life Science of the Ministry of Education, Culture, Sports, Science, and Technology of Japan, Grant-in-aid for Scientific Research (A) 21248010 from the Japan Society for the Promotion of Science, and Grant-in-aid for Japan Society for the Promotion of Science Fellows 23-3030.
The atomic coordinates and structure factors (codes 4IJ5 and 4IJ6) have been deposited in the Protein Data Bank (http://wwpdb.org/).
- iPSP
- metal-independent phosphoserine phosphatase
- PSP
- phosphoserine phosphatase
- dPSP
- metal-dependent PSP
- dPGM
- dependent phosphoglycerate mutase
- PDB
- Protein Data Bank.
REFERENCES
- 1. Chiba Y., Oshima K., Arai H., Ishii M., Igarashi Y. (2012) Discovery and analysis of cofactor-dependent phosphoglycerate mutase homologs as novel phosphoserine phosphatases in Hydrogenobacter thermophilus. J. Biol. Chem. 287, 11934–11941 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Kawasumi T., Igarashi Y., Kodama T., Minoda Y. (1980) Isolation of strictly thermophilic and obligately autotrophic hydrogen bacteria. Agric. Biol. Chem. 44, 1985–1986 [Google Scholar]
- 3. Kawasumi T., Igarashi Y., Kodama T., Minoda Y. (1984) Hydrogenobacter thermophilus gen. nov., sp. nov., an extremely thermophilic, aerobic, hydrogen-oxidizing bacterium. Int. J. Syst. Bacteriol. 34, 5–10 [Google Scholar]
- 4. Rigden D. (2008) The histidine phosphatase superfamily: structure and function. Biochem. J. 409, 333–348 [DOI] [PubMed] [Google Scholar]
- 5. Hamada K., Kato M., Shimizu T., Ihara K., Mizuno T., Hakoshima T. (2005) Crystal structure of the protein histidine phosphatase SixA in the multistep His-Asp phosphorelay. Genes Cells 10, 1–11 [DOI] [PubMed] [Google Scholar]
- 6. Rigden D. J., Littlejohn J. E., Henderson K., Jedrzejas M. J. (2003) Structures of phosphate and trivanadate complexes of Bacillus stearothermophilus phosphatase PhoE: structural and functional analysis in the cofactor-dependent phosphoglycerate mutase superfamily. J. Mol. Biol. 325, 411–420 [DOI] [PubMed] [Google Scholar]
- 7. Rigden D. J., Walter R. A., Phillips S. E., Fothergill-Gilmore L. A. (1999) Sulphate ions observed in the 2.12 Å structure of a new crystal form of S. cerevisiae phosphoglycerate mutase provide insights into understanding the catalytic mechanism. J. Mol. Biol. 286, 1507–1517 [DOI] [PubMed] [Google Scholar]
- 8. Rigden D. J., Bagyan I., Lamani E., Setlow P., Jedrzejas M. J. (2001) A cofactor-dependent phosphoglycerate mutase homolog from Bacillus stearothermophilus is actually a broad specificity phosphatase. Protein Sci. 10, 1835–1846 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Bourgis F., Botha F. C., Mani S., Hiten F. N., Rigden D. J., Verbruggen N. (2005) Characterization and functional investigation of an Arabidopsis cDNA encoding a homologue to the d-PGMase superfamily. J. Exp. Bot. 56, 1129–1142 [DOI] [PubMed] [Google Scholar]
- 10. Hills T., Srivastava A., Ayi K., Wernimont A. K., Kain K., Waters A. P., Hui R., Pizarro J. C. (2011) Characterization of a new phosphatase from Plasmodium. Mol. Biochem. Parasitol. 179, 69–79 [DOI] [PubMed] [Google Scholar]
- 11. Chiba Y., Terada T., Kameya M., Shimizu K., Arai H., Ishii M., Igarashi Y. (2012) Mechanism for folate-independent aldolase reaction catalyzed by serine hydroxymethyltransferase. FEBS J. 279, 504–514 [DOI] [PubMed] [Google Scholar]
- 12. Chiba Y., Horita S., Ohtsuka J., Arai H., Nagata K., Igarashi Y., Tanokura M., Ishii M. (2012) Crystallization and preliminary x-ray diffraction analysis of a novel type of phosphoserine phosphatase from Hydrogenobacter thermophilus TK-6. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 68, 911–913 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Kabsch W. (2010) XDS. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Vagin A., Teplyakov A. (2000) An approach to multi-copy search in molecular replacement. Acta Crystallogr. D Biol. Crystallogr. 56, 1622–1624 [DOI] [PubMed] [Google Scholar]
- 15. Emsley P., Lohkamp B., Scott W. G., Cowtan K. (2010) Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Vagin A. A., Steiner R. A., Lebedev A. A., Potterton L., McNicholas S., Long F., Murshudov G. N. (2004) REFMAC5 dictionary: organization of prior chemical knowledge and guidelines for its use. Acta Crystallogr. D Biol. Crystallogr. 60, 2184–2195 [DOI] [PubMed] [Google Scholar]
- 17. Lovell S. C., Davis I. W., Arendall W. B., 3rd, de Bakker P. I., Word J. M., Prisant M. G., Richardson J. S., Richardson D. C. (2003) Structure validation by Cα geometry: φ,ψ and Cβ deviation. Proteins 50, 437–450 [DOI] [PubMed] [Google Scholar]
- 18. Holm L., Park J. (2000) DaliLite workbench for protein structure comparison. Bioinformatics 16, 566–567 [DOI] [PubMed] [Google Scholar]
- 19. Krissinel E., Henrick K. (2005) Detection of protein assemblies in crystals. Comput. Life Sci. 3695, 163–174 [Google Scholar]
- 20. Krissinel E., Henrick K. (2007) Inference of macromolecular assemblies from crystalline state. J. Mol. Biol. 372, 774–797 [DOI] [PubMed] [Google Scholar]
- 21. Krissinel E. (2010) Crystal contacts as nature's docking solutions. J. Comput. Chem. 31, 133–143 [DOI] [PubMed] [Google Scholar]
- 22. Arnold K., Bordoli L., Kopp J., Schwede T. (2006) The SWISS-MODEL workspace: a web-based environment for protein structure homology modelling. Bioinformatics 22, 195–201 [DOI] [PubMed] [Google Scholar]
- 23. Schwede T., Kopp J., Guex N., Peitsch M. C. (2003) SWISS-MODEL: An automated protein homology-modeling server. Nucleic Acids Res. 31, 3381–3385 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Guex N., Peitsch M. C. (1997) SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modeling. Electrophoresis 18, 2714–2723 [DOI] [PubMed] [Google Scholar]
- 25. Edgar R. C. (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Rigden D. J., Mello L. V., Setlow P., Jedrzejas M. J. (2002) Structure and mechanism of action of a cofactor-dependent phosphoglycerate mutase homolog from Bacillus stearothermophilus with broad specificity phosphatase activity. J. Mol. Biol. 315, 1129–1143 [DOI] [PubMed] [Google Scholar]
- 27. Watkins H. A., Baker E. N. (2006) Structural and functional analysis of Rv3214 from Mycobacterium tuberculosis, a protein with conflicting functional annotations, leads to its characterization as a phosphatase. J. Bacteriol. 188, 3589–3599 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Parkinson G., Gunasekera A., Vojtechovsky J., Zhang X., Kunkel T. A., Berman H., Ebright R. H. (1996) Aromatic hydrogen bond in sequence-specific protein DNA recognition. Nat. Struct. Mol. Biol. 3, 837–841 [DOI] [PubMed] [Google Scholar]
- 29. Śmiechowski M. (2010) Theoretical pKa prediction of O-phosphoserine in aqueous solution. Chem. Phys. Lett. 501, 123–129 [Google Scholar]
- 30. Zachariou M., Traverso I., Spiccia L., Hearn M. T. (1996) Potentiometric investigations into the acid-base and metal ion binding properties of immobilized metal ion affinity chromatographic (IMAC) adsorbents. J. Phys. Chem. 100, 12680–12690 [Google Scholar]
- 31. Sasaki R., Sugimoto R., Chiba H. (1966) Yeast phosphoglyceric acid mutase-modifying enzyme. Arch. Biochem. Biophys. 115, 53–61 [DOI] [PubMed] [Google Scholar]
- 32. Walter R. A., Nairn J., Duncan D., Price N. C., Kelly S. M., Rigden D. J., Fothergill-Gilmore L. A. (1999) The role of the C-terminal region in phosphoglycerate mutase. Biochem. J. 337, 89–95 [PMC free article] [PubMed] [Google Scholar]
- 33. Bond C. S., White M. F., Hunter W. N. (2001) High resolution structure of the phosphohistidine-activated form of Escherichia coli cofactor-dependent phosphoglycerate mutase. J. Biol. Chem. 276, 3247–3253 [DOI] [PubMed] [Google Scholar]
- 34. White M. F., Fothergill-Gilmore L. A. (1992) Development of a mutagenesis, expression, and purification system for yeast phosphoglycerate mutase. Investigation of the role of active-site His-181. Eur. J. Biochem. 207, 709–714 [DOI] [PubMed] [Google Scholar]
- 35. Colman B., Norman E. G. (1997) Serine synthesis in cyanobacteria by a nonphotorespiratory pathway. Physiol. Plant. 100, 133–136 [Google Scholar]





