Skip to main content
The Journal of Biological Chemistry logoLink to The Journal of Biological Chemistry
. 2013 Dec 31;289(9):6259–6272. doi: 10.1074/jbc.M113.533554

Streptococcus sanguinis Class Ib Ribonucleotide Reductase

HIGH ACTIVITY WITH BOTH IRON AND MANGANESE COFACTORS AND STRUCTURAL INSIGHTS*

Olga Makhlynets , Amie K Boal §,1, DeLacy V Rhodes , Todd Kitten ¶,2, Amy C Rosenzweig §,3, JoAnne Stubbe ‡,‖,4
PMCID: PMC3937692  PMID: 24381172

Background: Class Ib ribonucleotide reductase (RNR) is an essential enzyme for aerobic growth of S. sanguinis.

Results: Its manganese form is 3.5-fold more active than the iron form when assayed with NrdH and thioredoxin reductase.

Conclusion: This specific activity is the highest reported to date for the class Ib RNR.

Significance: Our studies suggest why manganese is important in streptococcal pathogenesis.

Keywords: Enzyme Catalysis, Iron, Manganese, Pathogenesis, Ribonucleotide Reductase, Streptococcus, Activity, Endocarditis, Flavodoxin

Abstract

Streptococcus sanguinis is a causative agent of infective endocarditis. Deletion of SsaB, a manganese transporter, drastically reduces S. sanguinis virulence. Many pathogenic organisms require class Ib ribonucleotide reductase (RNR) to catalyze the conversion of nucleotides to deoxynucleotides under aerobic conditions, and recent studies demonstrate that this enzyme uses a dimanganese-tyrosyl radical (MnIII2-Y) cofactor in vivo. The proteins required for S. sanguinis ribonucleotide reduction (NrdE and NrdF, α and β subunits of RNR; NrdH and TrxR, a glutaredoxin-like thioredoxin and a thioredoxin reductase; and NrdI, a flavodoxin essential for assembly of the RNR metallo-cofactor) have been identified and characterized. Apo-NrdF with FeII and O2 can self-assemble a diferric-tyrosyl radical (FeIII2-Y) cofactor (1.2 Y2) and with the help of NrdI can assemble a MnIII2-Y cofactor (0.9 Y2). The activity of RNR with its endogenous reductants, NrdH and TrxR, is 5,000 and 1,500 units/mg for the Mn- and Fe-NrdFs (Fe-loaded NrdF), respectively. X-ray structures of S. sanguinis NrdIox and MnII2-NrdF are reported and provide a possible rationale for the weak affinity (2.9 μm) between them. These streptococcal proteins form a structurally distinct subclass relative to other Ib proteins with unique features likely important in cluster assembly, including a long and negatively charged loop near the NrdI flavin and a bulky residue (Thr) at a constriction in the oxidant channel to the NrdI interface. These studies set the stage for identifying the active form of S. sanguinis class Ib RNR in an animal model for infective endocarditis and establishing whether the manganese requirement for pathogenesis is associated with RNR.

Introduction

Ribonucleotide reductases (RNRs)5 catalyze the conversion of nucleotides to deoxynucleotides, providing the monomeric precursors required for DNA replication and repair in all organisms (1). All RNRs rely on a metallo-cofactor to oxidize a conserved cysteine in the active site of the enzyme into a thiyl radical, which then initiates reduction of nucleotides (2). In the case of the class I RNRs, two subunits are required, The β subunit contains a dinuclear metallo-cofactor that oxidizes this cysteine in the α subunit where nucleotide reduction occurs. The oxidation is reversible and occurs over a 35-Å distance (3). Recently, two new cofactors of the class I RNRs have been characterized, which has led to their subclassification (Ia, Ib, and Ic). The class Ia RNRs use a FeIII2-Y cofactor in vitro and in vivo. In contrast, the class Ib RNRs are active with both MnIII2-Y and FeIII2-Y (48) cofactors, although the class Ic RNRs require a MnIVFeIII cofactor (9).

In most organisms, the genes for the class Ib RNRs are found in operons. The subunits for these RNRs, α and β, are designated NrdE and NrdF, respectively. Two additional gene products are often found within this operon as follows: NrdI, a flavodoxin that we have recently shown behaves as a flavin oxidase (10), and NrdH, a glutaredoxin-like thioredoxin (11, 12). NrdF has been known for some time to be able to self-assemble an active FeIII2-Y cofactor from apo-NrdF, FeII, and O2 in a manner analogous to the corresponding β (NrdB) in the class Ia RNRs (13, 14). We have recently demonstrated with the Escherichia coli and Bacillus subtilis NrdFs (4, 5), as have Sjöberg and co-workers (8) with Bacillus anthracis NrdF, that a MnIII2-Y cofactor can be assembled from a MnII2-NrdF, but only in the presence of the reduced form of NrdI (NrdIhq) and O2. An x-ray structure of a complex of E. coli NrdIhq and MnII2-NrdF, in conjunction with biochemical studies, recently demonstrated that the oxidant produced by NrdIhq is channeled directly to one of the manganese ions in the MnII2-NrdF (15). Recent mechanistic studies on the B. subtilis MnIII2-Y cofactor assembly further established that O2 is the oxidant required to convert MnII2-NrdF to the MnIII2-Y (16).

In many, but not all, of the class Ib RNRs, NrdH can re-reduce the disulfide in α produced concomitant with dNDP formation (Fig. 1) and thus is required for multiple turnovers (6, 11, 1719). NrdH itself can be re-reduced in vitro by TrxR and NADPH in E. coli, B. anthracis, Staphylococcus aureus, and Corynebacterium glutamicum (6, 11, 20, 21).

FIGURE 1.

FIGURE 1.

RNR catalyzes the conversion of NDPs into dNDPs concomitant with disulfide bond formation. In the S. sanguinis class Ib RNR, the reducing equivalents to re-reduce this disulfide in vivo are supplied by NrdH/TrxR/NADPH.

Identification of the metal cofactor (manganese and/or iron) in class Ib RNR in vivo under different growth conditions is an active area of investigation. Recently, Corynebacterium ammoniagenes, E. coli, and B. subtilis NrdFs isolated from their endogenous host organisms were all shown to contain a MnIII2-Y cofactor (5, 7, 22, 23). Human and most eukaryotic RNRs utilize class Ia RNRs exclusively, whereas pathogenic organisms, including Streptococcus sanguinis, Streptococcus pneumoniae, B. anthracis, S. aureus, and Corynebacterium diphtheriae depend on class Ib RNR for aerobic production of deoxynucleotides. Knowledge of how the biosynthetic pathways for the class Ia and class Ib RNRs differ could thus potentially lead to new targets for antimicrobial therapeutics (24).

S. sanguinis was chosen to understand the role of the manganese RNR in pathogenesis for several reasons. First, it has recently been shown in a screen for virulence factors for this organism that deletion of SsaB, a manganese transporter, severely reduces infectivity of this organism (25). Second, deletion of SodA (the manganese-dependent superoxide dismutase) does not adequately explain the reduced virulence of the ssaB mutant (26). Third, S. sanguinis contains a class Ib and a class III RNR, essential for aerobic and anaerobic growth, respectively (12). This organism can thus also serve as a model for many pathogenic organisms with class Ib RNR as the only aerobic source of deoxynucleotides (27). Fourth, based on phylogenetic analysis (28), S. sanguinis NrdF belongs to a subclass distinct from E. coli, B. subtilis, and C. ammoniagenes, and thus these studies may be informative as to whether all NrdFs utilize a MnIII2-Y cofactor.

In this study, we report the cloning, expression, and isolation of NrdE, NrdF, NrdI paralogs (NrdI, FmnI, and FmnG), NrdH, and two putative thioredoxin reductases TrxR1 and TrxR2. We establish in vitro that an active cofactor of S. sanguinis class Ib RNR can be assembled with both manganese and iron and that only NrdI, not FmnI or FmnG, is essential in MnIII2-Y cofactor formation in NrdF. The MnIII2-Y-NrdF has activity of 6.1 s−1, 3.5-fold higher than for the FeIII2-Y-NrdF when assayed with NrdH and TrxR1. The MnIII2-Y-NrdF turnover number is very similar to the E. coli class Ia RNR (FeIII2-Y) (29). Finally, an animal model for infective endocarditis is available. Data in the accompanying paper (30) show that in vivo only the manganese NrdF appears to be active. Because mammalian host organisms use a FeIII2-Y for RNR activity, finding an inhibitor of NrdI/NrdF interactions required for MnIII2-Y cluster assembly could provide a new target for therapeutic intervention.

MATERIALS AND METHODS

General

All chemical reagents were purchased from Sigma, unless otherwise indicated. FMN was 73–79% FMN, <6% free riboflavin, <6% riboflavin diphosphates. FAD was 94% pure. MnCl2·4H2O (>98% pure, <5 ppm iron) and (NH4)2Fe(SO4)2·6H2O (99%) were used as a source of MnII and FeII, respectively. Primers were synthesized by Integrated DNA Technologies. All restriction enzymes were from New England Biolabs. His6-tagged SUMO protease was expressed in codon+ Rosetta cells (Novagen) from pTB145 provided by Prof. Bradley Pentelute (Chemistry Department, Massachusetts Institute of Technology). Buffers in this study are as follows: buffer A, 50 mm Hepes, 5% (v/v) glycerol, pH 7.6; buffer B, buffer A and 5 mm β-mercaptoethanol; buffer C, 50 mm sodium phosphate, 10% (v/v) glycerol, pH 7.6; buffer D, 50 mm Hepes, pH 7.6, 15 mm MgSO4, 1 mm EDTA. O2-saturated buffer A was prepared immediately prior to use by sparging on ice with O2 for at least 30 min. [3H]CDP was obtained from ViTrax and diluted with CDP in buffer A to 6,000–12,000 cpm/nmol before use. Radioactive samples were analyzed using a Beckman Coulter LS6500 scintillation counter. EPR spectra were acquired on a Bruker EMX X-band spectrometer at 77 K using a finger Dewar filled with liquid nitrogen and 706-PQ Wilmad EPR tubes or at room temperature (RT) using a Wilmad flat cell (150 μl). UV-visible spectra of anaerobic samples were acquired on a Varian Cary 3 UV-visible spectrophotometer using anaerobic cuvettes (Starna) with Teflon silicon septa (Thermo Scientific) and anaerobic titrations used a 50-μl gas tight syringe with a repeat dispenser (Hamilton). Fluorescence measurements were carried out using a Photon Technology International QM-4-SE spectrofluorometer and FELIX software. Manganese concentrations were measured using a PerkinElmer Life Sciences AAnalyst 600 atomic absorption spectrometer, and FeII concentrations were measured by the ferrozine assay (31). DNA sequencing and MALDI-TOF mass spectrometry were performed at the Biopolymers Laboratory, Massachusetts Institute of Technology. All anaerobic procedures were carried out in a glove box (MBraun), and all proteins were purified at 4 °C. For each protein molecular mass and the following extinction coefficients (ϵ) were calculated using ExPASy: NrdF or β2280 = 133,620 m−1 cm−1); NrdE or α2280 = 161,140 m−1 cm−1); NrdH (ϵ280 = 2,980 m−1 cm−1). Concentrations of NrdF and NrdE are reported for dimers. Concentration of NrdIox was determined based on A451 of FMNox cofactor (ϵ451 = 12,170 m−1 cm−1). Concentrations of TrxR1 and TrxR2 were estimated based on absorbance of bound FAD at 455 and 463 nm, respectively. As an extinction coefficient at these wavelengths, we used a number previously measured for E. coli TrxR, ϵ456 = 11,300 m−1 cm−1 (32).

Cloning of the S. sanguinis Genes Required for RNR Activity

Platinum Pfx DNA polymerase (Invitrogen), sense and antisense primers (supplemental Table S1) containing restriction sites (underlined), were used to amplify nrdE, nrdF, nrdI (SSA_2263), fmnG (SSA_1668), fmnI (SSA_1683), trxR2 (SSA_0813) and trxR1 (SSA_1865) from WT S. sanguinis SK36 genomic DNA as a template. Taq DNA polymerase (Promega), sense and antisense primers without restriction sites, was used to amplify nrdH. All PCRs were performed using 60 ng of genomic DNA, 0.25 μg of each primer in a total volume of 50 μl, and a thermocycler program that was optimized for long AT-rich primers (supplemental Table S2). nrdE, nrdF, nrdI, fmnG, fmnI, trxR1, and trxR2 were cloned into pET28a (Novagen) and nrdF into pET24a via NdeI and BamHI or NdeI and XhoI restriction sites using T4 DNA ligase (Promega). nrdH was cloned into pETSUMO using T4 DNA ligase as described in the manufacturer's protocol (Invitrogen).

Expression and Purification of NrdI, FmnG, FmnI, Apo-NrdF, and NrdE

pET28a-nrdI was transformed into BL21(DE3) cells and expressed in LB in the presence of 50 μg/ml kanamycin. The culture was grown at 37 °C with shaking at 200 rpm to OD600 = 0.5, and then riboflavin was added to a final concentration of 10 mg/liter. NrdI culture was induced at A600 = 0.7–0.8 with isopropyl β-d-1-thiogalactopyranoside (Promega) to a final concentration of 0.4 mm and incubated at 30 °C. In 4–5 h, cells were harvested by centrifugation at 3,000 × g for 10 min at 4 °C, frozen in liquid nitrogen, and stored at −80 °C. Typical yield was 1.8 g wet cell paste/liter of culture.

Cell pellet (18 g) was resuspended in 90 ml of buffer C, containing 1.7 mm FMN and two Complete protease inhibitor mixture tablets (Roche Applied Science). The cell suspension was passed twice through a French pressure cell at 14,000 p.s.i. and then centrifuged at 30,000 × g for 20 min. Nucleic acids were precipitated by addition of streptomycin sulfate solution to a final concentration of 1.3% (w/v) with stirring for 15 min. After the solution was spun down at 30,000 × g for 30 min, the supernatant was loaded onto Ni-NTA-agarose column (Qiagen, 1.5 × 3.4 cm, 6 ml) pre-equilibrated with buffer C, and the column was washed with buffer C containing 100 mm NaCl and 20 mm imidazole, pH 7.6, until A280 of the flow-through was <0.02. The protein was eluted with a 70 × 70-ml linear gradient of 20–250 mm imidazole in buffer C. NrdI-containing fractions, identified by A280 and A415, were pooled, loaded onto a Q-Sepharose Fast Flow column (GE Healthcare, 2.5 × 6.5 cm, 32 ml) pre-equilibrated with 100 mm NaCl in buffer C, and the column was washed with the same buffer. Flow-through fractions containing NrdI were pooled, and the protein was concentrated using an Amicon YM-10 centrifugal filter (Millipore), desalted on Sephadex G-25 column (Sigma, 1.5 × 36.5 cm, 64 ml), and further concentrated to 1–1.5 mm. A typical yield was 3 mg/liter of culture, and the protein was >95% homogeneous by SDS-PAGE analysis. FmnI and FmnG were purified as described above for NrdI.

NrdF and NrdE were expressed in BL21(DE3) and purified using protocols previously optimized for E. coli class Ib RNR (10). NrdF and DNA bands partially overlap (judged by A260/A280 and SDS-PAGE analysis); to remove DNA, the pooled NrdF fractions were chromatographed twice on a Q-Sepharose Fast Flow column.

Expression and Purification of NrdH, TrxR1, and TrxR2

pETSUMO-nrdH was transformed into BL21(DE3) cells and expressed in LB in the presence of 50 μg/ml kanamycin. The culture was grown at 37 °C with shaking at 200 rpm to OD600 = 0.6, and then the temperature was lowered to 30 °C, and the culture was induced with isopropyl β-d-1-thiogalactopyranoside to a final concentration of 0.4 mm. In 4 h, cells were harvested by centrifugation at 3,000 × g for 10 min at 4 °C, frozen in liquid nitrogen, and stored at −80 °C. Typical yield was 2.3 g of cell paste/liter of culture.

Cell pellet (9.2 g) was resuspended in 46 ml of buffer B containing one Complete protease inhibitor mixture tablet. The cell suspension was passed twice through a French pressure cell at 14,000 p.s.i. and then centrifuged at 20,000 × g for 25 min. Nucleic acids were precipitated by the addition of a streptomycin sulfate solution to a final concentration of 0.5% (w/v) with stirring over 20 min. After the solution was spun down at 20,000 × g for 30 min, the supernatant was loaded onto Ni-NTA-agarose column (1.5 × 2.5 cm, 4.4 ml) pre-equilibrated with buffer B containing 30 mm imidazole, and the column was washed with the same buffer until A280 of the flow-through was <0.02. The protein was eluted with 50 ml of 200 mm imidazole in buffer B, and 2-ml fractions were collected. Fractions containing SUMO-NrdH (21.6 kDa) were identified by SDS-PAGE (12% (w/v) acrylamide), pooled, and loaded onto a Q-Sepharose Fast Flow column (2.5 × 5.5 cm, 27 ml) pre-equilibrated with 100 mm NaCl in buffer B. The column was then washed with the same buffer (300 ml) then SUMO-NrdH was eluted with 100 × 100-ml linear gradient of 100–500 mm NaCl in buffer B, and 3.6-ml fractions were collected. Fractions with high A280 were analyzed by SDS-PAGE, pooled, and concentrated to 470 μm using an Amicon YM-10 filter.

SUMO-NrdH (470 μm) was divided into 0.6-ml aliquots, and each was incubated with SUMO protease (150 μl, 50 μm) overnight at 4 °C. About 60% of the protein was cleaved under these optimized conditions. The digested SUMO-NrdH was loaded directly onto a Ni-NTA-agarose column (1.5 × 2.5 cm, 4.4 ml) pre-equilibrated with 30 mm imidazole in buffer B. The column was washed with the same buffer, and 1-ml fractions were collected. Fractions containing NrdH (8.2 kDa), assessed by SDS-PAGE analysis (16% (w/v) Tricine gel, Invitrogen), were pooled and concentrated using an Amicon YM-3 filter. NrdH (2 ml) was then loaded onto Sephadex G-25 column (1.5 × 33 cm, 58 ml) pre-equilibrated with buffer A containing 10 mm DTT and 1 mm EDTA, eluted with the same buffer, and concentrated to 1 mm using an Amicon YM-3 filter. The ratio A280/A260 of homogeneous NrdH is 1.3. Successful removal of the tag was confirmed by MALDI-TOF MS.

To remove DTT from NrdH, required for the DTNB assay, NrdH (400 μl, 1 mm) was loaded onto Sephadex G-25 column (1 × 6.5 cm, 5 ml) pre-equilibrated with buffer A, and 0.5-ml fractions were collected. Fractions containing NrdH, as judged by A280/A260, were pooled and concentrated to 450 μm as described above.

pET28a-trxR1 (or pET28a-trxR2) was transformed into BL21(DE3) cells and trxR1 (trxR2) overexpressed in the presence of 50 μg/ml kanamycin. The culture was grown at 37 °C to OD600 = 0.5, and then riboflavin (10 mg/liter) was added, and the temperature was lowered to 30 °C; 10 min later, the culture was induced with isopropyl β-d-1-thiogalactopyranoside to a final concentration of 0.4 mm. In 4 h, the cells were harvested, as above to give typical yields of 2.5 g wet cell paste/liter of culture.

Cell pellet (7.5 g) was resuspended in 38 ml of buffer A containing a Complete protease inhibitor tablet and 1.7 mm FAD. The cell suspension was passed twice through a French pressure cell at 14,000 p.s.i. and then centrifuged at 20,000 × g for 25 min. Nucleic acids were precipitated by the addition of streptomycin sulfate solution to a final concentration of 1% (w/v). After the solution was spun down at 20,000 × g for 30 min, the supernatant was loaded onto a Ni-NTA-agarose column (1 × 4.2 cm, 3 ml), pre-equilibrated with buffer A with 30 mm imidazole, and the column was washed with the same buffer until A280 was <0.02. The protein was eluted with 20 ml of 200 mm imidazole in buffer A. The yellow fractions were pooled, diluted 4-fold with buffer A, and loaded onto Q-Sepharose Fast Flow column (2.5 × 5.5 cm, 27 ml), pre-equilibrated with buffer A and 100 mm NaCl, and the column was washed with the same buffer (∼150 ml). TrxR1 (TrxR2) was eluted with 100 × 100-ml linear gradient of 100–550 mm NaCl in buffer A and concentrated using an Amicon YM-30 filter; NaCl was removed by further dilution/concentration in buffer A.

Optimized Cofactor Assembly from Apo-NrdF Loaded with FeII or MnII

Buffers and proteins were degassed on a Schlenk line with at least five cycles of evacuation and refilling with argon. Concentrated NrdIox (0.5–1 mm) and apo-NrdF (0.8 mm) were stable at 4 °C for at least 5 days. To prepare MnII2-NrdF, apo-NrdF (45 μl, 800 μm) was incubated with 6 eq of MnII (2 mm solution, freshly prepared) in buffer A at 37 °C for 15 min. A mixture of MnII2-NrdF (60 μm) and NrdIox (120 μm) in buffer A in a total volume of 600 μl was titrated anaerobically with dithionite (∼3 mm, standardized using potassium ferricyanide (33)) until NrdIox was completely reduced to NrdIhq (judged by the disappearance of the band at 580 nm associated with NrdIsq). Cluster assembly was initiated by bubbling O2 through the MnII2-NrdF/NrdIhq solution for ∼1 min. This protocol typically gave 0.6 Y2. Increased yields of Y2 were obtained by recycling NrdI. After cluster assembly as described above, the mixture of MnIII2-Y (0.6 Y2) and NrdIox was again degassed on a Schlenck line (six cycles) and transferred inside the glove box to an anaerobic cuvette. The UV-visible spectrum was recorded to ensure that radical content remained intact, and then the mixture was titrated with dithionite (∼3 mm) to reduce NrdIox to NrdIhq. O2 was then bubbled through the sample for ∼1 min. To remove free MnII and NrdI, the mixture was incubated with EDTA (5 mm) at 4 °C for ∼30 min and loaded onto a Q-Sepharose Fast Flow column (1 × 3 cm, 2.5 ml) pre-equilibrated with 200 mm NaCl in buffer A. The column was washed with 200 mm NaCl (buffer A, 25 ml), and MnIII2-Y-NrdF was eluted with 500 mm NaCl (buffer A, 20 ml); 1-ml fractions were collected, and the protein was concentrated using an Amicon YM-30 filter. Typical radical content was 0.9–1.0 Y2.

For cluster assembly with iron, apo-NrdF (∼150 μl, 900 μm) was incubated anaerobically with 6 eq of FeII at 37 °C for 15 min and then diluted with buffer A to 60 μm. Cluster assembly was initiated by addition of an equal volume of oxygenated buffer A. To remove excess iron, ferrozine (80-fold excess over NrdF) and sodium dithionite (40-fold excess) were added, and the mixture was incubated on ice for 5 min. The protein was then desalted on a Sephadex G-25 column (1 × 11 cm, 8.5 ml) in buffer A. NrdF-containing fractions were pooled and concentrated. The amount of NrdF-bound FeII was measured by the ferrozine assay (31).

Y Quantitation

All EPR spectra used for spin quantitation were acquired under nonsaturating conditions at 77 K (4). Spin quantitation was performed by double integration of the signal and comparison with a standard of E. coli FeIII2-Y-NrdB.

Kd for the NrdIhq and MnII2-NrdF Interaction

The procedure is a minor modification of that previously reported (16). Under anaerobic conditions at 23 °C, NrdIhq (240 or 360 μm, ∼ 40 μl) in buffer C was added in 2-μl portions using an air-tight Hamilton syringe into MnII2-NrdF (1 or 3 μm, 700 μl) in the same buffer. To ensure that NrdIhq remains reduced throughout titration, buffer C also contained 100 μm dithionite. After each injection, the cuvette was inverted several times, and the solution was allowed to equilibrate at RT in the dark for 1 min, and the spectrum was recorded. The excitation wavelength was 380 nm, and the emission was measured from 475 to 625 nm in 1 nm steps with a 1-s integration time. The excitation and emission bandwidth slit was 1.5 and 0.75 mm, respectively. No photobleaching was detected using these settings. Data were analyzed by the method of Eftink (34). The titration was performed four times using different concentrations of MnII2-NrdF and NrdIhq, and the data were fit in IgorPro to obtain the stoichiometry of NrdIhq binding to NrdF (n) and the dissociation constant (Kd).

General Crystallographic Methods

All crystallographic datasets were collected at the Life Sciences Collaborative Access Team (LS-CAT) or General Medical Sciences and Cancer Institutes Collaborative Access Team (GM/CA-CAT) beamlines at the Advanced Photon Source and processed using the HKL2000 software package (35). Iterative rounds of refinement and model building were performed using Refmac5 (36) and Coot (37). Ramachandran plots were calculated with Molprobity (38) and figures were generated with the PyMOL Molecular Graphics System (Schrödinger, LLC). Internal channel calculations were performed with HOLLOW (39) using a 1.4 Å probe radius. Electrostatic surface potential calculations were carried out using the PyMOL APBS plugin (40). Electron density maps were calculated with FFT (41). Table 1 reports all data collection and refinement statistics.

TABLE 1.

Data collection and refinement statistics for the S. sanguinis MnII2-NrdF and NrdIox x-ray structures

S. sanguinis MnII2-NrdF S. sanguinis NrdIox
Data collection
    Space group P21 C2
    Resolution 2.65 Å (2.70 to 2.65 Å) 1.88 Å (1.91 to 1.88 Å)
    Rsyma,b 0.166 (0.747) 0.084 (0.705)
    〈II 10.8 (2.2) 16.9 (2.4)
    Completeness 99.9% (98.7%) 100% (100%)

Refinement
    Rworkc/Rfreed 0.237/0.276 0.207/0.238
    Average B-factor 28.7 24.2
    Root mean square deviations
        Bond lengths 0.006 Å 0.005 Å
        Bond angles 0.849° 1.008°

a Rsym = Σ|IobsIav|/ΣFobs, where the summation is over all reflections.

b Values in parentheses refer to the highest resolution shell.

c Rwork = Σ|FobsFcalc|/ΣFobs.

d For calculation of Rfree, 5% of the reflections was reserved.

X-ray Structure Determination of S. sanguinis MnII2-NrdF

Crystals of S. sanguinis MnII2-NrdF (25 mg/ml in 20 mm Hepes buffer, 5% (v/v) glycerol, pH 7.6) were generated using the hanging drop vapor diffusion method with 25% (w/v) PEG 3000, 250 mm magnesium formate, and 100 mm Hepes, pH 7.6, as the precipitating solution. Crystals appeared after 2 weeks of incubation at RT and were prepared for data collection by mounting on rayon loops and flash freezing in liquid nitrogen following cryoprotection by brief soaking in well solution containing 25% (v/v) glycerol. X-ray diffraction datasets were processed as described above with additional scaling performed using the UCLA MBI Diffraction Anisotropy Server (42). The structure was solved by molecular replacement using BALBES (43) with the Salmonella typhimurium FeIII2-NrdF structure (PDB accession code 1R2F) as the search model. Eight copies of NrdF, arranged into four β2 dimers, are present in the asymmetric unit. The quality of the electron density map varies widely between the eight monomers. The electron density is the least well defined for two of the monomers (chains A and G) and is of the highest quality for chains B and H. The latter subunits were used to draw conclusions about the structural features of the metal-binding site and oxidant channel. To aid in model building, tight noncrystallographic symmetry restraints were used in the initial phases of model refinement and released in the final rounds. The final model consists of residues 3–287 for chain A, residues 3–286 for chains B–F, residues 4–286 for chains G and H, two MnII ions per NrdF monomer, and 64 water molecules. Ramachandran plots show that 99.9% of residues are in allowed and generously allowed regions.

X-ray Structure Determination of S. sanguinis NrdIox

Crystals of S. sanguinis NrdIox (25 mg/ml in 20 mm Hepes buffer, 5% (v/v) glycerol, pH 7.6) were generated from a commercial screen (Qiagen) using the sitting drop vapor diffusion method with 30% (w/v) PEG 4000, 200 mm ammonium sulfate, and 100 mm sodium acetate, pH 5.6 as the precipitating solution. Crystals appeared after 1 week of incubation at RT and were prepared for data collection by addition of a well solution, in a 1:1 ratio, containing 50% (v/v) glycerol to the crystallization drop followed by mounting on rayon loops and flash freezing in liquid nitrogen. The structure was solved by molecular replacement using PHASER (44) with B. subtilis NrdIox (PDB accession code 1RLJ) as the search model. Ramachandran plots indicate that 100% of residues are in allowed and generously allowed regions. The asymmetric unit contains two NrdI monomers and the final model consists of residues 2–66, 72–154 in chain A, residues 2–154 in chain B, two FMN molecules, two sulfate molecules, and 204 water molecules. Residues 67–71 in chain A are disordered and could not be modeled. Attempts to determine the structure of reduced forms of NrdI by soaking NrdIox crystals in 10–100 mm solutions of dithionite produced color changes in the crystals, but structures obtained from the resulting diffraction datasets did not exhibit any significant conformational changes in response to FMN reduction.

DTNB Assay for TrxR1/TrxR2

In a final volume of 290 μl NADPH (300 μm), variable amounts of DTT-free NrdH (0.06–5 μm), 100 mm Tris (pH 7.5 at 20 °C), and 2 mm EDTA were mixed. DTNB was added to a final concentration of 1 mm, and the mixture was equilibrated to 25 °C in a cuvette. The reaction was initiated by addition of TrxR1 (17.5 nm/dimer) and monitored by change in A412. The turnover number was calculated as described previously (45). A similar experiment was carried out with TrxR2 (17.5 nm/dimer) and NrdH (5-30 μm); no change in A412 was observed.

Activity Assays

Three sets of conditions optimized to assay S. sanguinis RNR using DTT, NrdH/DTT, or NrdH/TrxR1/NADPH as the reductant are described. 1) For DTT a typical activity assay contained in a final volume of 170 μl the following: reconstituted NrdF (0.2 μm), NrdE (2 μm), dATP (100 μm), DTT (20 mm), [3H]CDP (0.5 mm, 11,700 cpm/nmol) in buffer D at 37 °C. Aliquots (30 μl) were removed at 0, 3, 6, 9, and 12 min, and the reaction was stopped by heating at 100 °C for 2 min. 2) For NrdH/DTT, a typical assay contained in a final volume of 170 μl the following: NrdF (0.07 μm), NrdE (0.14 μm), dATP (100 μm), DTT (20 mm), NrdH (10 μm), [3H]CDP (0.5 mm, 6,644 cpm/nmol) in buffer D at 37 °C, and aliquots (30 μl) were taken at 0, 1, 2, 3, 4 min. 3) For NrdH/TrxR1/NADPH, a typical assay contained in a final volume of 170 μl the following: NrdF (0.07 μm), NrdE (0.14 μm), dATP (100 μm), NrdH (DTT-free, 10 μm), TrxR1 (0.5 μm), NADPH (1 mm), [3H]CDP (0.5 mm, 6,644 cpm/nmol) in buffer D at 37 °C, and aliquots (30 μl) were taken at 0, 1, 2, 3, 4 min. [3H]dCDP was quantitated by the method of Steeper and Steuart (46). One unit of activity is defined as the production of 1 nmol of dCDP per min.

Km for NrdE-NrdF Interaction (47)

The reactions contained in a final volume of 170 μl the following: dATP (100 μm), DTT (20 mm), NrdH (10 μm), [3H]CDP (0.5 mm), and NrdF/NrdE in a ratio 1:1 (1 to 80 nm) in buffer D at 37 °C. Protein solutions below 0.1 μm also contained BSA (0.2 mg/ml). Aliquots (30 μl) were taken over 12 min for samples containing 1–5 nm, over 8 min for samples with 7–10 nm, and over 4 min for samples containing 20–80 nm NrdF and NrdE.

RESULTS

Identification of the Genes for Cluster Assembly and Activity of S. sanguinis RNR

The genome of S. sanguinis SK36 has been sequenced, and nrdH-nrdE-nrdK-nrdF were annotated in a single operon (48). A search for nrdI using B. subtilis nrdI as the query sequence revealed three candidate genes: SSA_2263 (nrdI), SSA_1668 (fmnG), and SSA_1683 (fmnI) (Fig. 2A). A general screen for genes essential under aerobic conditions identified only SSA_2263 (49). Thus, SSA_2263 was tentatively annotated as NrdI. Further analysis of this nrdI sequence, specifically the spacing between the ribosomal binding site and start codon, and ClustalW sequence alignments of characterized NrdIs suggested that Met5 is the actual start site and that the annotated start site is incorrect (Fig. 2B and supplemental Fig. S1). SSA_1668 and SSA_1683 were shown to bind FMN, and their genes were named fmnG and fmnI, respectively. Finally, to identify candidates for TrxR, B. anthracis thioredoxin reductase (BA5387) was chosen for a BLAST search, and two candidate genes, SSA_1865 or trxR1 (TrxR1) and SSA_0813 or trxR2 (TrxR2), were identified. Studies in the accompanying paper (30) reveal that only SSA_1865 is essential under aerobic growth conditions.

FIGURE 2.

FIGURE 2.

Location and organization of the genes of S. sanguinis associated with the class Ib RNR and its metallo-cofactor assembly and the identification of the nrdI start codon and ribosomal binding site. A, three annotated NrdIs and the RNR operon containing the genes for the redoxin that re-reduces the active site disulfide (nrdH), the two subunits of RNR (nrdE and nrdF), and an unknown open reading frame nrdK. B, two possible start codons for NrdI. Sequences upstream of the putative nrdI start codons are in black, the start codons and nrdI are in red. RBS, ribosomal binding site.

Expression and Purification of Apo-NrdF, NrdE, NrdI, FmnI, FmnG, NrdH, and TrxRs

Genes encoding NrdE, NrdF, NrdI, FmnI, FmnG, NrdH, TrxR1, and TrxR2 were amplified by PCR using S. sanguinis SK36 genomic DNA as a template. The gene for each protein was cloned into pET vectors for expression in BL21(DE3), and the sequences were verified. Apo-NrdF was expressed without a tag in pET24a. NrdE, NrdI, FmnI, FmnG, TrxR1, and TrxR2 all were expressed in pET28a with an N-terminal His6 tag and a 10-amino acid linker (MGSSHHHHHHSSGLVPRGSH). To obtain high yields of soluble protein, NrdH was cloned into pETSUMO and expressed as a fusion with a His6-SUMO tag, and subsequent to protein purification, the His6-SUMO tag was removed using SUMO protease. All of the proteins were purified to >95% homogeneity by conventional methods (Ni-NTA and ion exchange chromatography).

Characterization of NrdI

Phylogenetic analysis suggests that S. sanguinis NrdI is likely distinct from previously characterized NrdIs (5, 10, 16). Thus, NrdI (full-length and truncated by four amino acids, Fig. 2B) along with FmnI and FmnG were expressed and purified. As discussed subsequently, neither FmnI nor FmnG support MnIII2-Y assembly in NrdF, whereas both the full-length and truncated NrdIs support MnIII2-Y assembly in NrdF to the same extent. Given our re-annotation of NrdI, the truncated NrdI was used in subsequent experiments.

The characterization of S. sanguinis NrdI followed our recent protocols for E. coli and B. subtilis NrdIs (5, 10). NrdI as isolated contained 0.8 FMN/NrdI (A274/A451 = 3.4), and the extinction coefficient (ϵ451ox) of NrdI-bound FMN was calculated to be 12.17 mm−1 cm−1. The amount of FMN semiquinone was previously shown to be a distinguishing factor between generic flavodoxins and the NrdIs. Although flavodoxins have very distinct half-reaction reduction potentials, allowing accumulation of 100% semiquinone upon FMN reduction, the NrdIs have similar half-reaction reduction potentials. To determine the extent of semiquinone formation, anaerobic titration of NrdIox with dithionite was performed (data not shown). The results indicate that S. sanguinis NrdI stabilizes 35% of a neutral semiquinone FMN (ϵ580sq = 5.3 mm−1 cm−1, Fig. 3A) at RT, similar to E. coli (30%) (4) and B. subtilis (24%) (5) but distinct from B. cereus (100%) (50) and B. anthracis (60%) (51) NrdIs.

FIGURE 3.

FIGURE 3.

Spectroscopic characterization of S. sanguinis NrdI, MnII2-NrdF, and assembled clusters. A, spectra of NrdIox (red), NrdIhq (black), and NrdIsq (blue) in the presence (dotted) or absence (solid) of NrdF in buffer C. B, comparison of the EPR spectra (77 K) of S. sanguinis FeIII2-Y-NrdF (1.2 Y2, acquired at 50 microwatt power), and MnIII2-Y-NrdF (0.9 Y2, 1 milliwatt power). Spectra were normalized for NrdF concentration and power. Other spectrometer settings were 9.45 GHz, 1.5 G modulation amplitude, and 10.24 ms time constant. C, EPR spectrum of MnII2-NrdF (380 μm) at 9 K, 0.1 milliwatt. To prepare MnII2-NrdF, apo-NrdF (210 μl, 900 μm) was incubated with 6 eq of MnII in buffer A at 37 °C for 15 min. Unbound MnII was removed using a Sephadex G-25 column (36.5 × 1.5 cm, 64 ml) in buffer A, and the protein was concentrated on an Amicon YM-30 filter. The resulting protein contained 3.8 MnIII2. D, visible spectra of line 1, 100 μm NrdIox (dotted line); line 2, sample prepared by mixing 100 μm NrdIhq, 50 μm MnII2-NrdF, and O2; and line 3, sample resulting from titration of line 2 with dithionite to re-reduce NrdIox to NrdIhq followed by addition of O2. All samples are in buffer A.

Assembly of an Active NrdF Cofactor with Iron and Manganese

For FeIII2-Y, apo-NrdF was incubated anaerobically with FeII for 15 min at 37 °C followed by O2 addition. The spectrum of the resulting product revealed a sharp feature at 408 nm and broad bands at 325 and 370 nm corresponding to the Y (1.2 Y2) and the diferric cluster, respectively. The EPR spectrum of FeIII2-Y is similar to those previously reported (Fig. 3B, red) for E. coli, C. ammoniagenes, and S. typhimurium (4, 14, 52).

MnIII2-Y

Apo-NrdF was mixed aerobically with MnII (6 eq), incubated at 37 °C for 15 min, and the free MnII removed by Sephadex G-25 chromatography. Atomic absorption analysis typically gave 3.6–3.8 MnII per β2. The EPR spectrum of MnII2-NrdF at 9 K (Fig. 3C) is very similar to E. coli and S. typhimurium MnII2-NrdFs and distinct from the B. subtilis MnII2-NrdF (4, 14, 16).

Typically in cluster assembly studies, MnII2-NrdF is generated as above and incubated anaerobically with NrdIhq, without MnII removal, followed by exposure to O2. The UV-visible spectrum of the resulting NrdF (Fig. 3D, line 2) reveals a typical Y feature at 408 nm and absorption features at 550 and 330 nm associated with the MnIII2 cluster. Quantitative analysis typically gives 0.5–0.6 Y2 and 3.1–3.4 MnIII2. Many variables (pH, temperature, and ratio of manganese/NrdF) were examined in an effort to increase the amount of Y2; all conditions resulted in similar yields. A higher yield of Y2, however, was achieved by recycling the NrdI. The NrdIox-NrdF mixture resulting from the first effort to assemble cluster (Fig. 3D, line 2) was made anaerobic and titrated with dithionite to reduce NrdIox into NrdIhq. The sample was then re-exposed to O2 (Fig. 3D, line 3). This recycling process typically gives 0.9 Y2 and 3.7 MnIII2. The EPR spectrum of the MnIII2-Y at 77 K is a broad signal with a line width of 150 G and is shown in Fig. 3B (black line). Furthermore, as in the case of the other MnIII2-Y-NrdFs (4, 5, 22), the spectrum dramatically sharpens at 4 K (data not shown). Attempts to assemble MnIII2-Y cluster using FmnIhq or FmnGhq and conditions optimized for NrdI were unsuccessful.

The increase in active MnIII2-Y-NrdF cluster by NrdI recycling likely mimics the assembly process in vivo where a flavodoxin reductase would recycle catalytic amounts of NrdI (5, 23). However, a search of the S. sanguinis genome failed to reveal any candidate homologous to E. coli flavodoxin reductase (FlxR).

Quantitative Characterization of the Association of NrdIhq with MnII2-NrdF

To assess the affinity of NrdIhq for MnII2-NrdF, we took advantage of the previous observation in E. coli and B. subtilis systems that NrdIhq fluorescence is altered by the presence of NrdF (16). Similarly, fluorescence of S. sanguinis NrdIhq increases 8-fold in the presence of S. sanguinis MnII2-NrdF. Thus, titration studies were carried out as described previously (16), and the analysis reveals a Kd of 2.9 ± 1 μm with 1.5 ± 0.4 NrdIhq per NrdF. This Kd value is higher than those we previously reported for the NrdI-NrdF interactions in E. coli (Kd < 0.05 μm) and B. subtilis (Kd = 0.6 μm) (16). Given the Kd value and the concentrations of MnII2-NrdF (60 μm) and NrdIhq (120 μm) used to study cluster assembly described above, >95% of MnII2-NrdF is complexed.

To further explore the similarities/differences among NrdIs from various organisms, an additional experiment was carried out. Previous studies in the E. coli system showed that reduction of NrdIox in the presence of apo-NrdF (1 eq) led to formation of an anionic FMN semiquinone (∼30%). In a similar experiment with B. subtilis NrdF, no anionic form was detected (16). We evaluated the FMN form that accumulated upon reduction of S. sanguinis NrdI in the presence of apo-NrdF, and no anionic semiquinone flavin was observed.

Crystallographic Analysis of the S. sanguinis MnII2-NrdF

Because the streptococcal NrdFs are in a phylogenetic group that has not been previously characterized, we determined the structure MnII2-NrdF (2.65 Å resolution, Table 1) to compare with the corresponding structures of the E. coli and B. subtilis NrdFs. As shown in Fig. 4, the MnII2-binding site strongly resembles that of the E. coli MnII2-NrdF structure (15). Each MnII ion is six-coordinate with three bridging carboxylates, including the unusual bridging mode for Glu157 (S. sanguinis numbering), two His ligands, and two coordinated water molecules. The side chain of Asp66 hydrogen bonds to Tyr104, the site of Y formation in the active metallo-cofactor, and also resembles the linkage between the corresponding Tyr residue and the metal-binding site in E. coli NrdF. Although the moderate resolution prevents conclusive determination of the orientation of Glu191, the residue is modeled in a μ-η12 binding mode, as observed in E. coli MnII2-NrdF, and the refined model is consistent with this assignment. The structural similarity to E. coli MnII2-NrdF is consistent with similarities observed in the EPR spectra of these proteins.

FIGURE 4.

FIGURE 4.

X-ray structure of S. sanguinis (Ss) MnII2-NrdF (left) in comparison with E. coli (Ec) (middle, PDB accession code 3N37) and B. subtilis (Bs) (right, PDB accession code 4DR0) MnII2-NrdF structures. MnII ions (magenta) and water molecules (red) are shown as spheres, and selected amino acid side chains are shown in stick format. An FoFc omit map for the side chain of Glu157 (chain B) in S. sanguinis MnII2-NrdF is shown in green mesh contoured at 3.5σ.

The unusual configuration of the carboxylate side chains in the E. coli MnII2-NrdF structure opens a solvent-lined channel near Mn2 to the surface of protein (Fig. 5A). The channel is further accommodated by small hydrophobic residues near Mn2 (Ala75 and Ile94 in E. coli NrdF (53)). B. subtilis MnII2-NrdF, which has a different MnII2 coordination environment and larger Met residues in place of the Ala/Ile pair adjacent to the Mn2 site, does not contain a solvent-occupied channel (53). Because S. sanguinis MnII2-NrdF shares a similar coordination environment with E. coli MnII2-NrdF and retains the hydrophobic residues near Mn2 (Val74 and Ile93), it is not surprising that it contains a similar channel (Fig. 5B). The overall cavity shape and volume are nearly identical to the E. coli NrdF channel, and electron density for water molecules is also observed (Fig. 5, C and D). This solvent-lined channel in E. coli MnII2-NrdF is predicted to function as a conduit for the oxidant (O2 or HOO(H)) produced by O2/NrdIhq to the Mn2 site in MnII2-NrdF. This proposal is based on the structure of the E. coli NrdI-NrdF complex, which revealed a channel extending along the NrdI-NrdF interface, connecting the FMN of NrdI to Mn2 in NrdF. While this manuscript was being revised, the first structures of B. cereus MnII2-NrdF were reported (54) revealing partial occupancy of a bridging μ-1,3 binding mode for the Glu157 equivalent in this system, contrary to expectation based on the B. subtilis MnII2-NrdF structure (53). These new observations illustrate the capacity for dynamic conformational change in the MnII metal-binding site in NrdF. For instance, the B. subtilis MnII2-NrdF coordination environment may transiently switch to an arrangement resembling the E. coli/S. sanguinis MnII2-NrdF structures to facilitate oxidant transport or MnII loading.

FIGURE 5.

FIGURE 5.

Solvent channel to the Mn2-binding site in MnII2-NrdF structures. A, E. coli (Ec); B–D, S. sanguinis (Ss). In Ec (A) and Ss (B) MnII2-NrdF structures, the solvent-accessible channel is modeled as a cyan transparent surface. C, 2FoFc electron density map (cyan mesh) contoured at 1.2σ reveals ordered water molecules in the channel in the chain B NrdF monomer. D, zoomed-in view of the modeled waters near the MnII2 site and associated hydrogen-bonding interactions. Selected amino acid side chains and backbone atoms are shown in stick format. MnII ions (magenta) and water molecules (cyan) are shown as spheres.

Another interesting conserved feature in the S. sanguinis and E. coli MnII2-NrdF structures involves a constriction in the channel immediately above Mn2. In the E. coli system, the constriction is formed by the side chain of Ser159 and the backbone carbonyl of the bridging ligand Glu192, whereas in the S. sanguinis system it involves Thr158 and Glu191 (Fig. 5, A and B, supplemental Fig. S2). Because oxidant passage through this constriction may be a slow step in metallo-cofactor activation, substitution to a more sterically bulky residue could translate into a slower rate of reaction with the oxidant in the S. sanguinis system.

Crystal Structure of S. sanguinis NrdIox

NrdIs have been structurally characterized thus far from several species of Bacillus (50, 51) and E. coli (in complex with MnII2-NrdF) (15). As noted above, streptococcal NrdIs represent a third phylogenetic group for which no structural information exists. Sequence alignments (supplemental Fig. S3) reveal that S. sanguinis NrdI has three distinct insertions. The 1.88 Å resolution crystal structure supports this prediction showing an extension of helix α1 (residues 14–30), insertion of a new helix between the β-strand of β2 and β3 (α1*, residues 40–47), and extension of the loop between β3 and α2 near the FMN cofactor (Fig. 6A). This extended loop is 11 amino acids long (residues 65–76), making it significantly larger than the corresponding loops in E. coli NrdI (eight residues) and the Bacillus NrdIs (three residues).

FIGURE 6.

FIGURE 6.

X-ray structure of S. sanguinis NrdIox. A, comparison of the structures of B. subtilis (Bs) NrdIox (left, PDB accession code 1RLJ) and S. sanguinis (Ss) NrdIox (right). B, close-up view of the fully ordered 70s loop in chain B in S. sanguinis NrdIox. C, close-up view of the partially disordered S. sanguinis NrdIox chain A 70s loop (disordered residues 67–71 are shown as a heavy gray dashed line) in an alternate conformation. Selected amino acids and the flavin cofactor (yellow) are shown in stick format. Hydrogen bonding interactions are illustrated with thin dashed black lines. The flavin C4a position is marked with an asterisk.

The S. sanguinis NrdIox structure contains two copies of the protein in the asymmetric unit and thus provides two independent views of this extended loop's conformation. In one monomer, it is fully ordered and extends directly away from the FMN (Fig. 6B). In the second monomer, it is partially disordered (residues 67–71 are not modeled) and folds in toward the si face of the flavin (Fig. 6C). The fully ordered view of the loop (residues 65–76) shows that it adopts a complex secondary structure composed of an extension of the β-sheet secondary structure from which the loop emanates followed by a series of successive turns. Participation of the loop in an extended β-sheet hydrogen-bonding pattern prevents the 70s loop backbone from interacting directly with the FMN cofactor in the oxidized state. This feature is distinct from what is observed in the E. coli NrdIox structure in which the corresponding loop is flexible and glycine-rich, allowing for hydrogen bonding interactions with the N5 of the FMN via a backbone amide NH (15). The short loop (XFG) in the Bacillus NrdIs is similar to S. sanguinis NrdIox in that neither can form this interaction. The Bacillus and E. coli NrdIs have also been structurally characterized in the sq- or hq-reduced states (15, 50, 51). All of these structures reveal a backbone amide flip that positions a carbonyl group near the N5 position to interact with the now protonated form of the cofactor.

To understand whether a similar conformational change occurs upon reduction of S. sanguinis NrdI, diffraction datasets were collected on dithionite-soaked NrdIox crystals. This treatment reduced the FMN to the sq or hq state, based on observation of color change in the crystals, but the resulting electron density maps did not reveal any significant conformational changes in the 70s loop backbone. The protonated FMN cofactor (FMNH or FMNH) interacts instead with a water molecule (not observed in the NrdIox structure). The observed outcome in the dithionite-soaked S. sanguinis NrdI crystals could be the result of crystal packing interactions that preclude conformational changes in the loop or it could indicate that the 70s loop backbone cannot, in any oxidation state, bind directly to the FMN N5 position due to the β-sheet hydrogen bonding pattern in the loop near the FMN (Fig. 6B).

The S. sanguinis NrdI 70s loop is also unusual in that it contains many more bulky and negatively charged residues than the corresponding regions in other NrdIs. These residues could affect the electrostatic properties of the flavin and its accessibility to O2. However, evaluation of the electrostatic environment in NrdI near the FMN isoalloxazine ring shows that the interior of the FMN pocket remains positively charged, similar to what is observed in E. coli and the Bacillus NrdIs (Fig. 7). The negatively charged residues in the S. sanguinis NrdI 70s loop instead generate a strong negative patch on the exterior of the protein. We propose that this patch may be involved in interaction with a flavin reductase required to recycle NrdIox in vivo.

FIGURE 7.

FIGURE 7.

Electrostatic surface potential diagram of the S. sanguinis NrdIox (chain B) 70s loop and flavin binding pocket is shown contoured at −10 kBT (red) and +10 kBT (blue). The FMN cofactor is shown in stick format.

Physiological Reductant for NrdE

The reduction of NDPs to dNDPs is accompanied by oxidation of two active site cysteines in NrdE to a disulfide (Fig. 1) (2, 55). A number of artificial and endogenous systems are capable of mediating this re-reduction step, including dithiothreitol (DTT),Trx/TrxR/NADPH, Grx/GSH/GR/NADPH, and NrdH/TrxR/NADPH, where Trx is thioredoxin, Grx is glutaredoxin, and GR is glutaredoxin reductase (6, 11, 20, 21, 56). The observation that nrdH is colocalized in the same operon with nrdE and nrdF in S. sanguinis and many other organisms makes this protein the most reasonable candidate for the endogenous reductant. Our initial attempts to isolate untagged NrdH were unsuccessful due to its low solubility. To overcome this problem, NrdH was fused to a His6-SUMO tag and then isolated by nickel-affinity chromatography. Subsequent removal of the tag with SUMO protease gave soluble NrdH, which has been used with TrxR to assay RNR. Two candidate genes (trxR1 and trxR2) for thioredoxin reductases were identified, and the corresponding proteins, TrxR1 and TrxR2, were overexpressed and purified to homogeneity by nickel-affinity chromatography. To ensure that the isolated TrxRs were fully loaded with cofactor, FAD was added to crude cell lysates prior to purification (6, 57). This protocol gave homogeneous FAD-bound TrxR1 and TrxR2 with the ratio A272/A455 of 6.0 and 6.4, respectively. Complete cofactor loading was further confirmed by anion exchange chromatography.

The turnover number for each TrxR was measured using NADPH, NrdH, and the DTNB assay (45). The Km value of TrxR1 for NrdH was 0.09 μm and kcat = 3.5 s−1 giving kcat/Km of 4.03 × 107 m−1 s−1. TrxR2, however, showed no activity under the same conditions. Thus, the results suggest that TrxR1 is the reductant for NrdH in vivo.

Activity of FeIII2-Y and MnIII2-Y NrdFs Using DTT, NrdH/DTT, and NrdH/TrxR1/NADPH

The activity of the class Ib RNRs has been predominantly reported for the Fe-loaded NrdFs (14, 21, 58, 59), as the role of NrdI in generation of active Mn-loaded RNRs was not elucidated until 2010. Although a number of recent studies have reported activities for MnIII2-Y NrdFs, in most cases the proteins contained substoichiometric manganese loading and Y and/or the endogenous reductant was not used (46, 8). As a starting point to identify the metallo-cofactor required for S. sanguinis class Ib in cultures and in an animal model for S. sanguinis-mediated endocarditis, we have measured the activity of pure reconstituted MnIII2-Y and FeIII2-Y with DTT, DTT/NrdH, and the endogenous reductants NrdH and TrxR1 (Fig. 1). Some representative results of optimization of the concentrations of NrdE, NrdF, and variable reductants are shown in Fig. 8, A–C.

FIGURE 8.

FIGURE 8.

SA of MnIII2-Y-NrdF (0.9 Y2). A, SA of NrdF in the presence of variable amounts of NrdE using DTT (red) or NrdH/DTT (black) as a reductant. B, SA measured with variable [NrdH], 70 nm [NrdE], [NrdF], 0.1 mm [dATP], and 20 mm [DTT]; C, Km value for the interaction between NrdF and NrdE using a 1:1 ratio of subunits, 0.1 mm [dATP], 10 μm [NrdH], and 20 mm [DTT]; D, SA measured in the presence of variable dATP concentrations, 70 nm [NrdF], 0.14 μm [NrdE], 10 μm [NrdH], and 20 mm [DTT]. NrdE, specific activity of 3,000 units/mg, was used in all experiments. All plots were fit to Michaelis-Menten equation using IgorPro.

The Kd value for NrdE-NrdF interactions in the class Ib RNRs still remains largely unknown. However, recent studies on B. subtilis class Ib RNR demonstrated that the active form is a 1:1 complex of subunits (5, 47). Thus, our initial studies focused on establishing the ratio of NrdE to NrdF for maximum activity using DTT and NrdH/DTT as reductant. First, MnIII2-Y-NrdF (0.9 Y2) at 0.2 μm was assayed with increasing concentrations of NrdE (0.4–4 μm) and DTT in excess. Under these conditions, maximum activity is observed at 5 eq of NrdE (Fig. 8A, inset). When the assay was carried out with MnIII2-Y-NrdF (0.07 μm) using NrdH as the reductant, the SA of MnIII2-Y was 20-fold higher relative to DTT and only 1 to 2 eq of NrdE were required for maximum activity (Fig. 8A). Differences in the NrdE/NrdF ratios required for maximal activity with DTT and NrdH could in part reflect the higher efficiency of NrdH in recycling NrdE, which results in higher concentrations of reduced NrdE available for each turnover. In almost all subsequent assays, a 1:2 ratio of NrdF to NrdE was used.

The effect of NrdH on RNR activity was also examined by maintaining a 1:1 ratio of NrdF/NrdE while increasing the concentration of reduced NrdH. The apparent Km value for NrdH is 0.4 μm (Fig. 8B). Studies of the DTT requirement for NrdH cycling revealed that levels as low as 1 mm were sufficient for optimal activity. Thus with 0.07 μm MnIII2-Y, 10 μm NrdH, and 20 mm DTT, a specific activity of 5,700 units/mg was observed. Although DTT is routinely used as a reductant to assay RNR directly and, as just noted, can recycle oxidized NrdH, the endogenous reductant for NrdH is TrxR1. Thus, DTT was replaced with 0.5 μm TrxR1 and 1 mm NADPH and a specific activity of 5,000 units/mg was measured (Table 2). The observed turnover numbers with NrdH/DTT and NrdH/TrxR1/NADPH are very similar and considerably higher than the specific activity of MnIII2-Y-NrdF reported for other organisms with their endogenous reductants (6, 8). One additional variable, the concentration of 1:1 ratio NrdE and NrdF, was examined. The results, shown in Fig. 8C, reveal a Vmax of 6,000 units/mg and a Km of 6.4 nm.

TABLE 2.

Activity of MnIII2-Y-NrdF and FeIII2-Y-NrdF reconstituted in pure proteins with different reductants

Metal cofactor Y2 SA, units/mg
MeIII/β2
DTT NrdH, DTT NrdH, TrxR, NADPH
Mn 0.9 260 ± 30 5,700 ± 460 5,000 ± 175 3.7 ± 0.2
Fe 1.2 170 ± 40 1,500 ± 360 1,500 ± 100 3 ± 0.5

The SA of FeIII2-Y-NrdF (1.2 Y2) was also measured using the conditions optimized for the MnIII2-Y and gave an activity of 170 units/mg with DTT, 1,500 units/mg with NrdH/DTT, and 1,500 units/mg with NrdH/TrxR1/NADPH. A comparison of the Mn- and Fe-loaded NrdF activities indicates that the former has a turnover number that is 3.5-fold higher than the Fe-loaded enzyme, establishing that Mn-loaded NrdF is a better catalyst.

In the assays described in Fig. 8, CDP was used as a substrate and dATP (100 μm) as an allosteric effector. Our recent studies of B. subtilis class Ib RNR showed that dATP inhibits RNR activity at concentrations >5 μm (47). This result was surprising because class Ib RNRs do not have an ATP cone or activity domain, the typical binding site for dATP that facilitates class Ia RNR inhibition (52, 60, 61). To determine whether the S. sanguinis class Ib RNR is inhibited by dATP, the activity of MnIII2-Y-NrdF was measured in the presence of increasing dATP concentrations (Fig. 8D). The apparent Km is 2.4 μm, and no inhibition was observed even at 1 mm dATP.

DISCUSSION

Since our discovery that NrdI plays an essential role in MnIII2-Y formation, many reports have appeared describing the activity and properties of manganese-loaded class Ib RNRs (48, 22). Although the identification of the endogenous reductant has been reported in some of the studies, in most cases the activity of the class Ib RNR remained very low even with the endogenous reductant (6, 8). We have chosen to examine the importance of manganese- versus iron-loaded cofactor in S. sanguinis class Ib RNR because this organism causes infective endocarditis, and deletion of a manganese transporter results in loss of virulence (25). In addition, its NrdI and NrdF proteins belong to a group phylogenetically distinct from the E. coli, C. ammoniagenes, and B. subtilis class Ib enzymes, the only reported manganese-RNRs at the time we began this work.

Studies of the class Ib RNR have been hampered by the poor efficiency of MnIII2-Y cluster assembly and consequently low catalytic activity of the enzyme. Our recent work on the B. subtilis class Ib RNR revealed that chromatographic removal of apo- and substoichiometrically metallated NrdF and identification and use of the endogenous reductant (Trx/TrxR) increased the activity of MnIII2-Y-NrdF to 1,475 units/mg, 18 times that measured for RNR isolated directly from B. subtilis. Furthermore, the activity of the MnIII2-Y-NrdF was 12 times that observed for FeIII2-Y (125 units/mg) (47). In a similar study of the B. anthracis class Ib RNR, using the endogenous reductant Trx/DTT, the activity of the MnIII2-Y was 10 times higher than the FeIII2-Y. However, the overall activity of MnIII2-Y-NrdF was only 65 units/mg, possibly associated with the inefficiency of manganese cluster assembly (6). With S. sanguinis RNR, the activity of 6,000 units/mg is very high, but the activity difference between the manganese- and iron-loaded NrdFs is only 3.5-fold using the endogenous reductant, NrdH/TrxR1/NADPH. This result is distinct from all other systems studied to date and raises an important question of which metal is used as a cofactor in vivo and how the organism's environment might influence the enzyme's metallation state.

Class Ib RNRs are found in actinobacteria, firmicutes, and α- and γ-proteobacteria, and recent studies have used bioinformatic analysis of the genomes of interest to identify candidate genes for the endogenous reductant(s) (6, 47). In some organisms, nrdH is found in an operon with nrdE and nrdF, whereas in the Bacillus genus, the nrdH is in a distant location (12). Recent studies of B. anthracis class Ib RNR identified several candidates for its endogenous reductant system as follows: two Trxs, three putative NrdHs, and three TrxRs. Biochemical analysis of the reductant's efficiency (kcat/Km) in recycling NrdE, accompanied by Western blot analysis of concentrations of the most interesting reductants, led to the conclusion that both Trx1 and NrdH could function in this capacity but that the former is the most likely candidate in vivo (6). Importantly, the reported Vmax for the B. anthracis MnIII2-Y (45 to 65 units/mg) was 100-fold lower than our S. sanguinis turnover number and independent of the reductants Trx1/DTT, Trx1/TrxR1, NrdH/DTT, and NrdH/TrxR1. Finally, Trx1 had no stimulatory effect at all on the turnover number of FeIII2-Y, and therefore it was concluded that the manganese-loaded NrdF was the likely form of the class Ib RNR in vivo. Our studies of the B. subtilis class Ib RNR also support this hypothesis by showing that thioredoxins TrxA and YosR (NrdH-like) give a 10- and 5-fold stimulation of MnIII2-Y activity, respectively, relative to DTT, but they stimulate FeIII2-Y activity only 2-fold. In contrast, S. sanguinis NrdH significantly increased activity of both MnIII2-Y and FeIII2-Y 20-fold and 9-fold, respectively, relative to DTT.

Recently, we have measured an Km of 25 nm for the B. subtilis class Ib NrdE-NrdF (αβ) subunit interactions, which is 4–10-fold lower than the interactions in the E. coli class Ia RNR (Kd of 0.06 to 0.2 μm (62)). These results suggested that class Ib RNR can be assayed using a 1:1 ratio of subunits (47), instead of with an excess of one subunit over the other to ensure complete αβ complex formation (29). Our studies of S. sanguinis NrdF-NrdE, conducted using a 1:1 ratio of subunits in the concentration range from 0.001 to 0.1 μm (Fig. 8C), gave a Km of 6.4 nm, similar to the B. subtilis class Ib RNR. The NrdE-NrdF interaction presents an opportunity to crystallize the active αβ complex, and this work is in progress.

The interaction of S. sanguinis NrdF with its NrdI is also of interest because they belong to a third phylogenetic subgroup that remained uncharacterized until our studies (24, 28, 51). This subclassification based on bioinformatics is supported by the measured Kd of 2.9 μm for S. sanguinis NrdI-NrdF interactions, distinct from tighter interactions measured for E. coli (<0.05 μm) and B. subtilis (0.6 μm) systems (16). Although our crystallization efforts have not yielded a structure of the S. sanguinis NrdF-NrdI complex, crystallization of the individual proteins has been successful. As suggested by sequence alignments (supplemental Fig. S2), the coordination environment of the S. sanguinis MnII2-NrdF is much more similar to that of E. coli NrdF than the B. subtilis protein. The three distinguishing features between the E. coli and B. subtilis NrdFs are the H-bonding interactions between the Tyr residue to be oxidized in the active cofactor and the aspartate residue coordinated to Mn1, the presence of a H2O molecule coordinated to Mn1, and the unusual bridging coordination of Glu158 to Mn1 and Mn2 (15, 53). Using all three criteria, S. sanguinis MnII2-NrdF is similar to E. coli MnII2-NrdF (Fig. 4). Another distinguishing feature between E. coli and B. subtilis NrdFs is the presence of a water-lined channel linking Mn2 to the FMN in NrdI containing a constriction created by Ser159 (Fig. 5A). The channel has been proposed to provide a pathway for the oxidant from the flavin in NrdI to the metal-binding site in NrdF. The crystal structure of S. sanguinis MnII2-NrdF reveals the presence of a similar channel with a Thr158 counterpart to Ser159 at the constriction (Fig. 5B). The substitution for a bulkier side chain at this site may provide a more stringent selectivity filter for the oxidant. This property may be particularly critical in streptococci because these organisms accumulate high concentrations of H2O2 (63, 64), and the observed constriction may restrict NrdF MnII2 site access by this molecule. In addition, the presence of this constriction might explain why MnII2-NrdF cannot be activated directly with H2O2 itself (4, 14).

The x-ray structure of S. sanguinis NrdI confirms predictions about the unique features of the streptococcal class Ib proteins. The longer sequence of S. sanguinis NrdI translates into an extended helix α1, an additional helix α1*, and an extended 70s loop when compared with structures of E. coli and B. subtilis NrdIs (Fig. 6A). These extensions could be involved in interaction with another protein, such as a NrdI reductase. A crystal structure of E. coli flavodoxin reductase (FlxR) (65) was used to construct a docking model with S. sanguinis NrdI (models generated with ClusPro (66)). As shown in Fig. 9A, a model was produced that places the long and bulky 70s loop and inserted helix α1* of NrdI in close contact with FlxR near its FAD cofactor. A similar model for interactions between FlxR and flavodoxin suggested that flavodoxin binds in a bowl-shaped pocket close to FAD with positively charged residues of FlxR positioned to interact with conserved negatively charged residues in the flavodoxin (65). A negative electrostatic surface potential near the FMN is a defining characteristic of flavodoxins, but NrdIs instead use positively charged residues to facilitate reduction of FMNH to FMNH (67, 68). Based on analysis of the S. sanguinis NrdI electrostatic surface potential near its FMN cofactor, the negatively charged 70s loop does not seem to influence the positive electrostatic environment around the FMN and instead is localized at the outer surface of the protein (Fig. 7). Thus, we hypothesize that the negatively charged 70s loop may play a role in facilitating interaction between the positively charged pocket of FlxR (Fig. 9B) and the positively charged patch around FMN in NrdI.

FIGURE 9.

FIGURE 9.

Model of an interaction between E. coli FlxR (PDB accession code 1FDR) and S. sanguinis NrdI. A, docking model shows that the 70s loop and α1* of NrdI may interact with FlxR. FlxR is in green; NrdI is in cyan with the 70s loop and α1* in blue, and the FMN (red) and FAD (magenta) cofactors are shown as sticks. The model was generated using ClusPro (66). B, electrostatic surface potential diagram of FlxR near the modeled interface with NrdI is shown contoured at −15 kBT (red) and +15 kBT (blue). The FAD cofactor is shown in stick format.

Using the E. coli crystal structure of the NrdF-NrdI complex and sequence alignments (supplemental Fig. S2 and S3) with representative class Ib enzymes from the other subgroups, residues involved in forming the NrdF-NrdI interface were predicted for the B. subtilis and S. sanguinis systems. Interestingly, the NrdF portion of the protein-protein interface is more conserved than the NrdI portion. Therefore, differences in the Kd values measured for E. coli, B. subtilis, and S. sanguinis may be due to variations in the NrdI surface (supplemental Table S3) rather than differences in the NrdF component. Moreover, electrostatic analysis of S. sanguinis NrdF revealed that the region surrounding the predicted NrdI-binding site is negative, which, in combination with the negatively charged 70s loop, could account for the particularly low affinity between these proteins.

An active cluster, MnIII2-Y-NrdF and FeIII2-Y-NrdF, can be assembled with 1 Y2, and both forms exhibit relatively high activity with the physiological reductant NrdH/TrxR. However, the 3.5-fold difference in activities relative to the 10- and 12-fold differences between the MnIII2-Y and FeIII2-Y observed with the B. anthracis and B. subtilis NrdFs, respectively, suggests that this organism might be able to stay active in vivo with either cofactor, with loading dependent on the growth environment. The accompanying paper by Rhodes et al. (30) demonstrates in a rabbit model for infective endocarditis that a strain of S. sanguinis in which nrdI has been deleted does not colonize heart valves, unlike the WT-strain. These studies thus provide the first evidence that a class Ib NrdF requires manganese under conditions in which the organism is pathogenic. Considering that streptococci and other pathogens, including enterococci, staphylococci, and Bacillus sp., contain class Ib RNR as their only aerobic RNR, prevention of MnIII2-Y formation in the class Ib RNR may be an attractive target for new antimicrobials.

Supplementary Material

Supplemental Data

Acknowledgments

We thank Prof. S. J. Lippard (Department of Chemistry, Massachusetts Institute of Technology) for access to the atomic absorption spectrometer, Dr. J. Wilson for assistance in atomic absorption data acquisition, and Prof. Bradley Pentelute (Massachusetts Institute of Technology) for providing the plasmid that codes for SUMO protease. We also thank Sara Sperling for assistance with crystallization of NrdI. Use of the Advanced Photon Source, an Office of Science User Facility operated for the United States Department of Energy Office of Science by Argonne National Laboratory, was supported by the United States Department of Energy under Contract DE-AC02-06CH11357. Use of LS-CAT Sector 21 was supported by the Michigan Economic Development Corp. and Michigan Technology Tri-Corridor Grant 085P1000817. GM/CA-CAT at APS has been funded in whole or in part by National Institutes of Health Grant Y1-CO-1020 from NCI and Grant Y1-GM-1104 from NIGMS.

*

This work was supported, in whole or in part, by National Institutes of Health Grants GM81393 (to J. S.), GM58518 (to A. C. R.), R56AI085195 (to T. K.), and K12GM093857 (to Paul B. Fisher for D. V. R.) and Pathway to Independence Award K99/R00 (to A. K. B.).

The atomic coordinates and structure factors (codes 4N83 and 4N82) have been deposited in the Protein Data Bank (http://wwpdb.org/).

5
The abbreviations used are:
RNR
ribonucleotide reductase
DTNB
5,5′-dithiobis-(2-nitrobenzoic acid)
FlxR
E. coli flavodoxin reductase
Grx
glutaredoxin
hq
FMN hydroquinone form
Ni-NTA
nickel nitrilotriacetic acid
NrdE
α subunit of class Ib ribonucleotide reductase
NrdF
β subunit of class Ib ribonucleotide reductase
NrdH
glutaredoxin-like protein that reduces NrdE
NrdI
flavodoxin-like protein essential for cluster assembly with manganese cofactor
ox
FMN oxidized form
PDB
Protein Data Bank
SA
specific activity
sq
FMN semiquinone form
TrxR
thioredoxin reductase
Tricine
N-[2-hydroxy-1,1-bis(hydroxymethyl)ethyl]glycine
Y
tyrosyl radical.

REFERENCES

  • 1. Hofer A., Crona M., Logan D. T., Sjöberg B.-M. (2012) DNA building blocks: keeping control of manufacture. Crit. Rev. Biochem. Mol. Biol. 47, 50–63 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Stubbe J., van Der Donk W. A. (1998) Protein radicals in enzyme catalysis. Chem. Rev. 98, 705–762 [DOI] [PubMed] [Google Scholar]
  • 3. Minnihan E. C., Nocera D. G., Stubbe J. (2013) Reversible, long-range radical transfer in E. coli class Ia ribonucleotide reductase. Acc. Chem. Res. 46, 2524–2535 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Cotruvo J. A., Jr., Stubbe J. (2010) An active dimanganese(III)-tyrosyl radical cofactor in Escherichia coli class Ib RNR. Biochemistry 49, 1297–1309 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Zhang Y., Stubbe J. (2011) Bacillus subtilis class Ib RNR is a dimanganese(III)-tyrosyl radical enzyme. Biochemistry 50, 5615–5623 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Gustafsson T. N., Sahlin M., Lu J., Sjöberg B.-M., Holmgren A. (2012) Bacillus anthracis thioredoxin systems, characterization and role as electron donors for ribonucleotide reductase. J. Biol. Chem. 287, 39686–39697 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Stolle P., Barckhausen O., Oehlmann W., Knobbe N., Vogt C., Pierik A. J., Cox N., Schmidt P. P., Reijerse E. J., Lubitz W., Auling G. (2010) Homologous expression of the nrdF gene of Corynebacterium ammoniagenes strain ATCC 6872 generates a manganese-metallocofactor (R2F) and a stable tyrosyl radical (Y) involved in ribonucleotide reduction. FEBS J. 277, 4849–4862 [DOI] [PubMed] [Google Scholar]
  • 8. Crona M., Torrents E., Røhr A. K., Hofer A., Furrer E., Tomter A. B., Andersson K. K., Sahlin M., Sjöberg B.-M. (2011) NrdH-redoxin protein mediates high enzyme activity in manganese-reconstituted ribonucleotide reductase from Bacillus anthracis. J. Biol. Chem. 286, 33053–33060 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Jiang W., Yun D., Saleh L., Barr E. W., Xing G., Hoffart L. M., Maslak M.-A., Krebs C., Bollinger J. M., Jr. (2007) A manganese(IV)/iron(III) cofactor in Chlamydia trachomatis ribonucleotide reductase. Science 316, 1188–1191 [DOI] [PubMed] [Google Scholar]
  • 10. Cotruvo J. A., Jr., Stubbe J. (2008) NrdI, a flavodoxin involved in maintenance of the diferric-tyrosyl radical cofactor in Escherichia coli class Ib RNR. Proc. Natl. Acad. Sci. U.S.A. 105, 14383–14388 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Jordan A., Aslund F., Pontis E., Reichard P., Holmgren A. (1997) Characterization of Escherichia coli NrdH. J. Biol. Chem. 272, 18044–18050 [DOI] [PubMed] [Google Scholar]
  • 12. Lundin D., Torrents E., Poole A. M., Sjöberg B.-M. (2009) RNRdb, a curated database of the universal enzyme family ribonucleotide reductase, reveals a high level of misannotation in sequences deposited to GenBank. BMC Genomics 10, 589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Atkin C. L., Thelander L., Reichard P. (1973) Iron and free radical in ribonucleotide reductase. J. Biol. Chem. 248, 7464–7472 [PubMed] [Google Scholar]
  • 14. Huque Y., Fieschi F., Torrents E., Gibert I., Eliasson R., Reichard P., Sahlin M., Sjöberg B.-M. (2000) The active form of the R2F protein of class Ib ribonucleotide reductase from Corynebacterium ammoniagenes is a diferric protein. J. Biol. Chem. 275, 25365–25371 [DOI] [PubMed] [Google Scholar]
  • 15. Boal A. K., Cotruvo J. A., Jr., Stubbe J., Rosenzweig A. C. (2010) Structural basis for activation of class Ib RNR. Science 329, 1526–1530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Cotruvo J. A., Jr., Stich T. A., Britt R. D., Stubbe J. (2013) Mechanism of assembly of the dimanganese-tyrosyl radical cofactor of class Ib ribonucleotide reductase: enzymatic generation of superoxide is required for tyrosine oxidation via Mn(III)Mn(IV) intermediate. J. Am. Chem. Soc. 135, 4027–4039 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Jordan A., Pontis E., Aslund F., Hellman U., Gibert I., Reichard P. (1996) The ribonucleotide reductase system of Lactococcus lactis. J. Biol. Chem. 271, 8779–8785 [DOI] [PubMed] [Google Scholar]
  • 18. Gon S., Faulkner M. J., Beckwith J. (2006) In vivo requirement for glutaredoxins and thioredoxins in the reduction of the ribonucleotide reductases of Escherichia coli. Antioxid. Redox Signal. 8, 735–742 [DOI] [PubMed] [Google Scholar]
  • 19. Lu J., Holmgren A. (2013) The thioredoxin antioxidant system. Free Radic. Biol. Med. January 8:66:75–87 [DOI] [PubMed] [Google Scholar]
  • 20. Van Laer K., Dziewulska A. M., Fislage M., Wahni K., Hbeddou A., Collet J.-F., Versées W., Mateos L. M., Tamu Dufe V., Messens J. (2013) NrdH redoxin of Mycobacterium tuberculosis and Corynebacterium glutamicum dimerizes at high protein concentration and exclusively receives electrons from thioredoxin reductase. J. Biol. Chem. 288, 7942–7955 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Rabinovitch I., Yanku M., Yeheskel A., Cohen G., Borovok I., Aharonowitz Y. (2010) Staphylococcus aureus NrdH redoxin is a reductant of the class Ib ribonucleotide reductase. J. Bacteriol. 192, 4963–4972 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Cox N., Ogata H., Stolle P., Reijerse E., Auling G., Lubitz W. (2010) A tyrosyl-dimanganese coupled spin system is the native metalloradical cofactor of the R2F subunit of the ribonucleotide reductase of Corynebacterium ammoniagenes. J. Am. Chem. Soc. 132, 11197–11213 [DOI] [PubMed] [Google Scholar]
  • 23. Cotruvo J. A., Stubbe J. (2011) Escherichia coli class Ib RNR contains a dimanganese(III)-tyrosyl radical cofactor in vivo. Biochemistry 50, 1672–1681 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Cotruvo J. A., Stubbe J. (2011) Class I ribonucleotide reductases: metallocofactor assembly and repair in vitro and in vivo. Annu. Rev. Biochem. 80, 733–767 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Das S., Kanamoto T., Ge X., Xu P., Unoki T., Munro C. L., Kitten T. (2009) Contribution of lipoproteins and lipoprotein processing to endocarditis virulence in Streptococcus sanguinis. J. Bacteriol. 191, 4166–4179 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Crump K. E., Bainbridge B., Brusko S., Turner L., Ge X., Stone V., Xu P., Kitten T. (2013) The relationship of the lipoprotein SsaB, manganese, and superoxide dismutase in Streptococcus sanguinis virulence for endocarditis. Mol. Microbiol., submitted [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Jordan A., Reichard P. (1998) Ribonucleotide reductases. Annu. Rev. Biochem. 67, 71–98 [DOI] [PubMed] [Google Scholar]
  • 28. Torrents E., Sahlin M., Biglino D., Gräslund A., Sjöberg B.-M. (2005) Efficient growth inhibition of Bacillus anthracis by knocking out the ribonucleotide reductase tyrosyl radical. Proc. Natl. Acad. Sci. U.S.A. 102, 17946–17951 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Ge J., Yu G., Ator M. A., Stubbe J. (2003) Pre-steady-state and steady-state kinetic analysis of E. coli class I ribonucleotide reductase. Biochemistry 42, 10071–10083 [DOI] [PubMed] [Google Scholar]
  • 30. Rhodes D. V., Crump K. E., Makhlynets O. V., Snyder M., Ge X., Xu P., Stubbe J., Kitten T. (2014) Genetic characterization and role in virulence of the ribonucleotide reductases of Streptococcus sanguinis. J. Biol. Chem. 289, 6273–6287 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Fish W. W. (1988) Rapid colorimetric micromethod for the quantitation of complexed iron in biological samples. Methods Enzymol. 158, 357–364 [DOI] [PubMed] [Google Scholar]
  • 32. Williams C. H., Jr., Zanetti G., Arscott L. D., McAllister J. K. (1967) Lipoamide dehydrogenase, glutathione reductase, thioredoxin reductase, and thioredoxin. J. Biol. Chem. 242, 5226–5231 [PubMed] [Google Scholar]
  • 33. Schellenberg K. H., Hellerman L. (1958) Oxidation of reduced diphosphopyridine nucleotide. J. Biol. Chem. 231, 547–556 [PubMed] [Google Scholar]
  • 34. Eftink M. R. (1997) Fluorescence methods for studying equilibrium macromolecule-ligand interactions. Methods Enzymol. 278, 221–257 [DOI] [PubMed] [Google Scholar]
  • 35. Otwinowski Z., Minor W. (1997) Processing of x-ray diffraction data collected in oscillation mode. Methods Enzymol. (Part A) 276, 307–326 [DOI] [PubMed] [Google Scholar]
  • 36. Murshudov G. N., Vagin A. A., Dodson E. J. (1997) Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr. D Biol. Crystallogr. 53, 240–255 [DOI] [PubMed] [Google Scholar]
  • 37. Emsley P., Cowtan K. (2004) Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 [DOI] [PubMed] [Google Scholar]
  • 38. Chen V. B., Arendall W. B., 3rd, Headd J. J., Keedy D. A., Immormino R. M., Kapral G. J., Murray L. W., Richardson J. S., Richardson D. C. (2010) MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D Biol. Crystallogr. 66, 12–21 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Ho B. K., Gruswitz F. (2008) HOLLOW: generating accurate representations of channel and interior surfaces in molecular structures. BMC Struct. Biol. 8, 49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Baker N. A., Sept D., Joseph S., Holst M. J., McCammon J. A. (2001) Electrostatics of nanosystems: application to microtubules and the ribosome. Proc. Natl. Acad. Sci. U.S.A. 98, 10037–10041 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Ten Eyck L. F. (1985) Fast Fourier transform calculation of electron density maps. Methods Enzymol. 115, 324–337 [DOI] [PubMed] [Google Scholar]
  • 42. Strong M., Sawaya M. R., Wang S., Phillips M., Cascio D., Eisenberg D. (2006) Toward the structural genomics of complexes: crystal structure of a PE/PPE protein complex from Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. U.S.A. 103, 8060–8065 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Long F., Vagin A. A., Young P., Murshudov G. N. (2008) BALBES: a molecular-replacement pipeline. Acta Crystallogr. D Biol. Crystallogr. 64, 125–132 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. McCoy A. J. (2007) Solving structures of protein complexes by molecular replacement with Phaser. Acta Crystallogr. D Biol. Crystallogr. 63, 32–41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Arnér E. S., Holmgren A. (2000) in Current Protocols in Toxicology (Maines M., Costa L., Reed D., Sassa S., eds) pp. 7.4.1–7.4.14, John Wiley & Sons, Inc., New York [Google Scholar]
  • 46. Steeper J. R., Steuart C. D. (1970) A rapid assay for CDP reductase activity in mammalian cell extracts. Anal. Biochem. 34, 123–130 [DOI] [PubMed] [Google Scholar]
  • 47. Parker M. J., Zhu X., Stubbe J. (2014) Bacillus subtilis class Ib ribonucleotide reductase: high activity and dynamic subunit interaction. Biochemistry, 10.1021/bi401056e [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Xu P., Alves J. M., Kitten T., Brown A., Chen Z., Ozaki L. S., Manque P., Ge X., Serrano M. G., Puiu D., Hendricks S., Wang Y., Chaplin M. D., Akan D., Paik S., Peterson D. L., Macrina F. L., Buck G. A. (2007) Genome of the opportunistic pathogen Streptococcus sanguinis. J. Bacteriol. 189, 3166–3175 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Xu P., Ge X., Chen L., Wang X., Dou Y., Xu J. Z., Patel J. R., Stone V., Trinh M., Evans K., Kitten T., Bonchev D., Buck G. A. (2011) Genome-wide essential gene identification in Streptococcus sanguinis. Sci. Rep. 1, 125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Røhr A. K., Hersleth H.-P., Andersson K. K. (2010) Tracking flavin conformations in protein crystal structures with Raman spectroscopy and QM/MM calculations. Angew. Chem. Int. Ed. Engl. 49, 2324–2327 [DOI] [PubMed] [Google Scholar]
  • 51. Johansson R., Torrents E., Lundin D., Sprenger J., Sahlin M., Sjöberg B.-M., Logan D. T. (2010) High-resolution crystal structures of the flavoprotein NrdI in oxidized and reduced states–an unusual flavodoxin. FEBS J. 277, 4265–4277 [DOI] [PubMed] [Google Scholar]
  • 52. Jordan A., Pontis E., Atta M., Krook M., Gibert I., Barbé J., Reichard P. (1994) A second class I ribonucleotide reductase in Enterobacteriaceae: characterization of the Salmonella typhimurium enzyme. Proc. Natl. Acad. Sci. U.S.A. 91, 12892–12896 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Boal A. K., Cotruvo J. A., Jr., Stubbe J., Rosenzweig A. C. (2012) The dimanganese (II) site of Bacillus subtilis class Ib ribonucleotide reductase. Biochemistry 51, 3861–3871 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Hammerstad M., Hersleth H.-P., Tomter A., B., Røhr Å. K., Anderson K. K. (2013) Crystal structure of Bacillus cereus class Ib ribonucleotide reductase di-iron NrdF in complex with NrdI. ACS Chem. Biol., 10.1021/cb400757h [DOI] [PubMed] [Google Scholar]
  • 55. Thelander L., Reichard P. (1979) Reduction of ribonucleotides. Annu. Rev. Biochem. 48, 133–158 [DOI] [PubMed] [Google Scholar]
  • 56. Phulera S., Mande S. C. (2013) The crystal structure of Mycobacterium tuberculosis NrdH at 0.87 Å suggests a possible mode of its activity. Biochemistry 52, 4056–4065 [DOI] [PubMed] [Google Scholar]
  • 57. Mulrooney S. B. (1997) Application of a single-plasmid vector for mutagenesis and high-level expression of thioredoxin reductase and its use to examine flavin cofactor incorporation. Protein Expr. Purif. 9, 372–378 [DOI] [PubMed] [Google Scholar]
  • 58. Yang F., Curran S. C., Li L.-S., Avarbock D., Graf J. D., Chua M.-M., Lu G., Salem J., Rubin H. (1997) Characterization of two genes encoding the Mycobacterium tuberculosis ribonucleotide reductase small subunit. J. Bacteriol. 179, 6408–6415 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Roca I., Torrents E., Sahlin M., Gibert I., Sjöberg B.-M. (2008) NrdI essentiality for class Ib ribonucleotide reduction in Streptococcus pyogenes. J. Bacteriol. 190, 4849–4858 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Eliasson R., Pontis E., Jordan A., Reichard P. (1996) Allosteric regulation of the third ribonucleotide reductase (NrdEF enzyme) from Enterobacteriaceae. J. Biol. Chem. 271, 26582–26587 [DOI] [PubMed] [Google Scholar]
  • 61. Aravind L., Wolf Y. I., Koonin E. V. (2000) The ATP-cone: an evolutionarily mobile, ATP-binding regulatory domain. J. Mol. Microbiol. Biotechnol. 2, 191–194 [PubMed] [Google Scholar]
  • 62. Climent I., Sjöberg B.-M., Huang C. Y. (1991) Carboxyl-terminal peptides as probes for Escherichia coli ribonucleotide reductase subunit interaction: kinetic analysis of inhibition studies. Biochemistry 30, 5164–5171 [DOI] [PubMed] [Google Scholar]
  • 63. Chen L., Ge X., Dou Y., Wang X., Patel J. R., Xu P. (2011) Identification of hydrogen peroxide production-related genes in Streptococcus sanguinis and their functional relationship with pyruvate oxidase. Microbiology 157, 13–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Kreth J., Merritt J., Shi W., Qi F. (2005) Competition and coexistence between Streptococcus mutans and Streptococcus sanguinis in the dental biofilm. J. Bacteriol. 187, 7193–7203 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Ingelman M., Bianchi V., Eklund H. (1997) The three-dimensional structure of flavodoxin reductase from Escherichia coli at 1.7 Å resolution. J. Mol. Biol. 268, 147–157 [DOI] [PubMed] [Google Scholar]
  • 66. Comeau S. R., Gatchell D. W., Vajda S., Camacho C. J. (2004) ClusPro: an automated docking and discrimination method for the prediction of protein complexes. Bioinformatics 20, 45–50 [DOI] [PubMed] [Google Scholar]
  • 67. Zhou Z., Swenson R. P. (1995) Electrostatic effects of surface acidic amino acid residues on the oxidation-reduction potentials of the flavodoxin from Desulfovibrio vulgaris (Hildenborough). Biochemistry 34, 3183–3192 [DOI] [PubMed] [Google Scholar]
  • 68. Ludwig M. L., Schopfer L. M., Metzger A. L., Pattridge K. A., Massey V. (1990) Structure and oxidation-reduction behavior of 1-deaza-FMN flavodoxins: modulation of redox potentials in flavodoxins. Biochemistry 29, 10364–10375 [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supplemental Data

Articles from The Journal of Biological Chemistry are provided here courtesy of American Society for Biochemistry and Molecular Biology

RESOURCES