Watching DNA polymerase [eegr] make a phosphodiester bond
Teruya Nakamura,Ye Zhao,Wei Yang,Teruya Nakamura,Yuriko Yamagata,Ye Zhao,Yue-jin Hua
DOI: https://doi.org/10.1038/nature11181
IF: 64.8
2012-01-01
Nature
Abstract:DNA synthesis has been extensively studied, but the chemical reaction itself has not been visualized. Here we follow the course of phosphodiester bond formation using time-resolved X-ray crystallography. Native human DNA polymerase η, DNA and dATP were co-crystallized at pH 6.0 without Mg2+. The polymerization reaction was initiated by exposing crystals to 1 mM Mg2+ at pH 7.0, and stopped by freezing at desired time points for structural analysis. The substrates and two Mg2+ ions are aligned within 40 s, but the bond formation is not evident until 80 s. From 80 to 300 s structures show a mixture of decreasing substrate and increasing product of the nucleotidyl-transfer reaction. Transient electron densities indicate that deprotonation and an accompanying C2′-endo to C3′-endo conversion of the nucleophile 3′-OH are rate limiting. A third Mg2+ ion, which arrives with the new bond and stabilizes the intermediate state, may be an unappreciated feature of the two-metal-ion mechanism. Formation and breakage of chemical bonds underlie all life processes. DNA replication, which is essential for cell proliferation, is but one example1, 2. In each reaction cycle, a dNTP complementary to the templating base is incorporated into DNA in a nucleotidyl-transfer reaction catalysed by a polymerase, during which a new bond is formed between the 3′-OH of the primer strand and the α-phosphate of the dNTP, and the phosphodiester bond between the α- and β-phosphates of dNTP is broken (Fig. 1a). As a result the primer strand is extended by one nucleotide and a pyrophosphate is released. This reaction has been shown to require two Mg2+ ions (A and B) and is inferred to be SN2-type, forming a pentacovalent phosphate intermediate3, 4. A similar two-metal-ion-dependent mechanism is believed to be shared by all DNA and RNA polymerases and many nucleases5. DNA synthesis has been analysed by kinetic measurements, dynamic simulations and structural studies2, 6, 7. Crystal structures of a number of DNA polymerases in substrate-bound forms have been determined using non-reactive substrate analogues or with Ca2+ instead of Mg2+ to prevent the nucleotidyl-transfer reaction5, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18. These structures show active-site configurations in an array of pre-reaction ground states. Pre-steady-state kinetic studies of several DNA polymerases indicate that, after large conformational changes induced by substrate binding, unspecified subtle changes in the active site are the rate-limiting step19, 20, 21, 22, 23, 24, 25, 26, 27. The actual process of DNA synthesis has, to our knowledge, never been visualized. Human DNA polymerase η (Pol η) is specialized in lesion bypass and has a preformed catalytic centre that undergoes limited conformational change during DNA synthesis28. Crystal structures of the catalytic domain of Pol η complexed with DNA and a non-reactive dNTP analogue poised for catalysis have been determined at 1.8 Å resolution29, 30. Taking advantage of the slow reaction rate and hindered post-reaction DNA translocation in crystallo, we follow the course of DNA synthesis by X-ray crystallography and report here the structural changes and transient elements that are associated with the reaction. The Pol η activity in a reaction buffer containing 0.9 mM Mg2+ and 0.1 mM Ca2+ was only one-third of that with 1 mM Mg2+ (Fig. 1b), indicating that both active-site metal ions have to be Mg2+ for the catalysis. In accordance with the bell-shaped pH-dependence curve characteristic for the acid–base catalysis17, 31, the catalytic rate of Pol η is extremely low at pH 6.0 and rises with increasing pH from 6.0 to 8.0 (Fig. 1c). To prevent the nucleotidyl-transfer reaction, crystals of Pol η–substrate complexes were grown at pH 6.0 with only one Ca2+ per protein–DNA–dATP complex (Methods). After soaking in MES (pH 6.0–7.2) or HEPES buffer (pH 7.0–7.5) with Na+ or K+ but no divalent cation for 5 to 30 min, these crystals maintained excellent diffraction up to pH 7.0, but decayed at pH 7.5 with reduced resolution and increased diffuse scattering. The structure of Pol η crystal equilibrated at pH 6.8 was refined at 1.50 Å. A Ca2+ ion clearly occupies the B metal-ion-binding site and is coordinated in the octahedral geometry by oxygen atoms of the dATP and the active-site residues (Fig. 1d). The A metal-ion-binding site has low occupancy of a monovalent cation. The occupancy of Na+ or K+ in the A site increases with the pH (Supplementary Fig. 1a, b), which is probably correlated with deprotonation of the active-site carboxylates. Pol η binds Na+ more readily than K+, particularly below pH 7.0, probably because Na+ is similar to Mg2+ in size and is smaller than K+. As expected, there is no nucleotidyl-transfer reaction, and the structure is termed the ground-state ternary complex. The ground-state structure is similar to the Pol η ternary complex with a non-reactive dATP analogue dAMPNPP (2′-deoxyadenosine-5′-[(α, β)-imido]triphosophate) and two Mg2+ ions (Protein Data Bank ID 3MR2)29. However, without an A-site divalent cation, the 3′-OH of the primer strand shifts away from the dATP and forms hydrogen bonds with the side chains of the S113 and D115 residues, and the active-site carboxylates D13 and E116 adopt different rotamer conformations (Fig. 1d). Replacement of dAMPNPP with dATP leads to stabilization of R61, which forms bidentate hydrogen bonds with the α-phosphate. The nucleotidyl transfer was initiated by transferring Pol η crystals to a pH 6.8 or 7.0 reaction buffer containing 1 mM Mg2+ but no dATP (Methods). After incubation at 293 K for 40 to 300 s, the reaction was terminated in ~40-s intervals by freezing crystals in liquid nitrogen at 77 K. Diffraction data were collected to 1.50–1.95 Å Bragg spacings (Supplementary Table 1). The reaction process was monitored by the electron density corresponding to the new chemical bond in the Fo − Fc map compared with the refined ground-state structure (Fig. 2a and Supplementary Fig. 2). The reaction time courses at pH 6.8 and 7.0 in crystallo (Fig. 2b) are approximately 20–100-fold slower than in solution (Fig. 1c), probably owing to the reduced thermal motion. By 40 s, the A site was fully occupied with Mg2+ (Fig. 2a and Supplementary Fig. 1c, d). About 50% of Ca2+ in the B site was replaced by Mg2+, and exchange of the remaining Ca2+ took place slowly (Supplementary Fig. 3). Binding of two Mg2+ ions leads to the alignment of the 3′-OH and dATP18, 32. The refined structure is nearly identical to that with dAMPNPP, except for tighter dATP coordination by R61 and closer proximity of the Mg2+ ions (3.4 Å apart versus 3.6 Å). As there is no sign of bond formation, the structure is termed the reactant state. Electron density corresponding to a new bond between the 3′-OH and α-phosphorus of dATP begun to emerge at 80 s, increased quickly over the next 60 s and reached a maximum after 200 s, when the reaction was 60–70% complete (Fig. 2b). The seemingly reduced rate after 140 s is probably due to the reverse reaction in crystallo. The slight decline of product after 250 s (Fig. 2b) is due to a sideway product translocation, which is clear with longer incubation (Supplementary Fig. 4a). To alleviate an impediment to proper DNA translocation by the crystal lattice, we replaced an AT base pair with a mismatch at the DNA end that formed lattice contacts (Methods and Supplementary Table 2). The DNA with a TG mismatch led to isomorphous crystals and proper translocation of the DNA product (Supplementary Fig. 4b). Interestingly, the time courses of nucleotidyl transfer in the AT and TG crystals are nearly identical and are unaffected by the lattice contacts (Fig. 2b). At the peak of chemical-bond formation between 200 and 250 s, the scissile phosphate can be refined in a pentacovalent transition state without restraints (Fig. 2c); however, the bond distances between the phosphorus and the attacking or leaving oxygen atoms are 2.2–2.5 Å (Fig. 2c), much longer than the expected 2.0 Å observed with transition-state mimics such as AlF4 or MgF3 (refs 33–35). Moreover, in the Fo − Fc map residual electron densities are observed around the new and scissile phosphodiester bonds (Fig. 2c). The same diffraction data, however, can be well fitted as a mixture of the reactant state and the product state immediately before and after the nucleotidyl transfer (Fig. 2d). Because the transition state is transient and unstable, the structures obtained between 80 and 300 s are refined as a mixture of the reactant state and product state at different ratios (Fig. 2b and Supplementary Table 1a). For instance, the 1.52 Å structure at 230 s consists of 40% substrate and 60% product. Between the reactant and product state, the protein, DNA and dATP are superimposable except for atoms in the reaction centre. Most notably, the α-phosphorus of dATP moves 1.4 Å along a straight line between the attacking and leaving oxygen atoms (separated by 4.6 Å; Fig. 2d). The shift of the α-phosphate is accompanied by alteration of residue R61 of Pol η, which flips away from the scissile phosphate and is replaced by a new metal ion and water molecules (Fig. 2a, d; see details below). On the primer strand, changes are confined to the 3′ nucleotide. The 3′-OH, together with the deoxyribose, moves towards the α-phosphate by 0.5 Å, and the sugar pucker changes from C2′-endo in the reactant state to C3′-endo in the product state (Fig. 2d). With the loss of the nucleophile and α-phosphate as ligands, the A-site Mg2+ dissociates in the product state, as evidenced by the declining occupancy (Supplementary Fig. 1c, d). Concomitantly, D13 assumes a second conformation and forms a hydrogen bond with K224. The C3′-endo (A form) conformation at the 3′ primer end was observed in ternary complexes with the A-, B- and X-family DNA polymerases and is thought to be important for forming a shallow minor groove for dNTP selection9, 11, 12, 14, 15, 17, 18, 32. Among the Y-family polymerases, the primer end has always been observed as C2′-endo28. In the Pol η ground-state and reactant-state complexes, the dATP and the nucleotide 5′ to the primer end have the A-form conformation, but only in the product-state structure does the primer end adopt the C3′-endo conformation to avoid clashes between its C2′ atom and the non-bridging oxygen of dATP during and immediately after the nucleophilic attack (Figs 2d and 3a). Because the electron density is weak for the sugar moiety at the 3′-primer end, we tested the effect of the A-form conformation by using a primer with a ribonucleotide at its 3′ end. The catalytic efficiency (kcat/Km) of Pol η is comparable whether a ribonucleotide or deoxyribonucleotide is at the primer end (Fig. 3b), as observed for DNA pol β (ref. 36), indicating that the A-form conformation is probably necessary for DNA synthesis in general. Two unexpected spheres of electron density appear in the Fo − Fc maps in the course of new bond formation (Fig. 2a). The first is within hydrogen-bonding distance of the 3′-OH. Its electron density peaks at 80 s and declines considerably after 140 s when the product state becomes prominent. Although absent in the ground state and the 40-second reactant-state structure, the electron density is superimposable with a water molecule observed in the Pol η–dAMPNPP ternary complex. This water molecule is hydrogen-bonded with the 3′-OH, the O4′ of dATP and another water molecule in the 80-second reactant-state structure, but it is incompatible with the C3′-endo conformation in the product state (Fig. 3a). The 3′-OH is also hydrogen-bonded to a water molecule in an X-family DNA polymerase crystallized with dUMPNPP (2′-deoxyuridine-5′-[(α, β)-imido]triphosophate)32. A water-mediated and substrate-assisted catalytic mechanism has previously been proposed37, 38. In addition to the water molecule, S113 is hydrogen-bonded with the 3′-OH in the ground state and may pick up the proton and pass it to E116 in the reactant state (Fig. 1d). However, the S113A-mutant Pol η retains 95% of kcat with a threefold increase of Km (Fig. 3b and Supplementary Table 3). Interestingly, mutations of the S113 equivalent in the A- and B-family DNA polymerases (highly conserved histidine and threonine, respectively) also have limited impact on catalytic efficiency39, 40, 41. We therefore propose that 3′-OH is prone to deprotonation when coordinated by Mg2+ and aligned with the incoming nucleotide4. The proton can be passed on to the transient water molecule and then to bulk solvent. When the 3′-end of the primer is a ribonucleotide, the O2′ may relay the proton out instead. An incoming dNTP most probably participates in deprotonation of the nucleophile because, with the nitrogen substitution in dAMPNPP, the water molecule is bound stably to the 3′-OH29, 30. The second emerging sphere of electron density is very close to the α-phosphate on the opposite side of the A- and B-site Mg2+ ions (Fig. 2a). It starts to appear at 140 s and intensifies with the reaction time. By 230 s with the structure refined at 1.52 Å resolution, the octahedral geometry and the short coordination distances indicate that this is a divalent cation and most probably Mg2+ (Fig. 4). This third metal ion is liganded by four water molecules, two of which occupy the space of the departed R61 (Fig. 2d). Two additional ligands are the leaving oxygen (bridging between the α- and β-phosphate) and the non-bridging oxygen of the α-phosphate (Fig. 4). This Mg2+ thus bridges the two reaction products destined to separate and prevents DNA translocation. In the TG crystals with 1 mM Mg2+ in the reaction buffer, the mixed reactant-state and product-state intermediate stays at equilibrium for up to 15 min, but with reduced Mg2+ (5 µM) and addition of dATP (5 µM), product release occurs much faster (Supplementary Information). The physiological role of the third metal ion is probably to neutralize the negative charge built up in the transition state and may also facilitate protonation of the pyrophosphate. Three metal ions have been structurally observed in a number of enzymes catalysing phosphoryl-transfer reactions, including alkaline phosphatase, P1 nuclease and endonuclease IV33, 42, 43, 44. In the case of endonuclease IV, the three Zn2+ ions observed in both the substrate and product state44 are reminiscent of the Mg2+ with Pol η (Supplementary Fig. 5). The two metal ions essential for catalysis flank the scissile phosphate on one side, and the third metal ion bridges the reaction products on the other. A similar arrangement of three metal ions was proposed for group-I introns based on chemical probing45. However, the third metal ion was absent in a crystal structure of a splicing intermediate, probably owing to its transient nature46. Previously, flash–freeze was used to trap covalent intermediates and conformational changes associated with a chemical reaction47, 48, 49. Here, we extend the technology to record DNA synthesis in real time and at atomic resolution (Fig. 5 and Supplementary Movie 1). The arrival of a water molecule to deprotonate the 3′-OH occurs only after two-Mg2+-ion-induced substrate alignment and in the presence of a correct dNTP but not a non-reactive dAMPNPP. Mg2+-dependent and water-mediated deprotonation may occur in all DNA polymerases. The third metal ion that replaces the side chain of R61 during the nucleotidyl transfer probably stabilizes the transition state and facilitates product release. In the one-metal-ion-dependent TraI-like nucleases and topoisomerases a conserved Lys or Arg also occupies the position equivalent to R61 in Pol η (ref. 50). The involvement of an additional transient metal ion may be a general feature of both the one- and two-metal-ion mechanism. S113A-mutant Pol η was made using QuikChange (Stratagene). Wild-type Pol η (residues 1–432) and S113A-mutant proteins were expressed and purified as described29. The ternary complexes of Pol η, DNA and dATP were mixed with Ca2+ at 1:1 molar ratio and concentrated to a protein concentration of ~3 mg ml−1. Oligonucleotides used for original (AT pair at the DNA end) and three ‘loosely packed’ crystals (TG or GT mispaired and an unpaired A) are shown in Supplementary Table 2. An unpaired nucleotide at the primer 5′ end was disordered in the original Pol η structures29, and was therefore removed to improve the diffraction quality. All crystals were obtained using the hanging-drop vapour-diffusion method against a reservoir solution containing 0.1 M MES (pH 6.0) and 9–17% (w/v) PEG2K-MME and streak seeding. Based on diffraction qualities, the complexes containing the original AT and TG-end DNA were selected for chemical reaction in crystallo. These crystals were first transferred and incubated in a pre-reaction buffer containing 0.1 M MES (pH 6.8–7.2), 5 μM dATP, 20% (w/v) PEG2K-MME and 1 mM dithiothreitol (DTT) for ~30 min. For the pH 6.8 time course (Fig. 2b, Supplementary Table 1b) and translocation in the TG crystal, the pre-reaction buffer also contained 5 µM MgCl2. The different pre-reaction buffer might cause the larger variations of the pH 6.8 reactions. Chemical reaction was initiated by transferring the crystals with a nylon loop into a reaction buffer containing 0.1 M MES (pH 6.8–7.2), 1 mM MgCl2, 20% (w/v) PEG2K-MME and 1 mM DTT. After incubation for a desired time point (40 s and intervals of ~40 s), the crystals were quickly dipped in a cryo-solution supplemented with 20% (w/v) glycerol and flash-cooled in liquid nitrogen. For in crystallo reactions, MES buffer was titrated by KOH and not NaOH to avoid binding of Na+ ion at the A site. To observe product release, TG crystals were soaked in 5 µM dATP and 5 µM MgCl2 for 30 min and chased with addition of 1 mM MgCl2 for 1–8 min. Diffraction data were collected at 100 K on beam lines 22ID and 22BM of the advanced photon source. Data were processed with HKL200051 or XDS52, and converted to structure factors by TRUNCATE53. All data were isomorphic in the P61 space group and processed with identical unit-cell parameters a = b = 98.8 Å and c = 82.4 Å for structure and electron-density comparison. All structural figures were drawn using PyMOL (http://www.pymol.org). To monitor the new bond formation (Fig. 2a), Fo − Fc maps are calculated with PHENIX54 using the Fo of 40 s, 80 s, 140 s and 230 s after addition of 1 mM Mg2+ (pH 7.0) and the Fc of the refined reactant-state structure (40 s) with the 3′-OH of the primer end, the A-site Mg2+ and the α-phosphate of dATP omitted. To determine the reaction time courses (Fig. 2b and Supplementary Figs 1–3), all data were scaled against and compared to the refined ground-state structure, and electron densities for the new bond (between the 3′-OH of the primer strand and the α-phosphorus of dATP) and the metal ions were measured in the Fo − Fc map (with the B-site metal ion omitted) at each time point using programs in CCP453. After refinement of the ground-state structure, the intermediate structures were refined as a mixture of reactant state and product state using PHENIX54 and Coot55. The ratio of reactant state and product state at each time point was determined according to the time course of the new bond formation (Fig. 2b) and difference (Fo − Fc) in electron-density maps. For high-resolution refinement (~1.5 Å resolution), the translation–libration–screw refinement was applied. Data collection and refinement statistics are summarized in Supplementary Table 1. Kinetic measurements including Km and kcat were done using the template/primer shown in Supplementary Table 3. The basic reaction mixture contained 2.5 nM Pol η, 5 μM 5′-fluorescein-labelled primer and template, and 0–80 µM dNTP in 40 mM Tris–HCl (pH 7.5), 5 mM MgCl2, 10 mM dithiothreitol, 100 mM KCl, 0.1 mg ml−1 bovine serum albumin and 5% glycerol. For the metal-ion-competition assay, Mg2+ was increased from 0 to 1.0 mM in 0.1-mM steps and Ca2+ was decreased to keep the combined concentration of 1.0 mM. The effects of pH were screened from pH 6.0 to pH 8.0. All reactions took place at room temperature for 4 or 8 min and were stopped by addition of 10× formamide loading buffer to the final concentrations of 8% formamide, 1 mM EDTA (pH 8.0) and 0.1 mg ml−1 xylene cyanol. After heating to 90 °C for 3 min and immediately placing on ice, products were resolved on 20% polyacrylamide sequencing gels containing 5.5 M urea. Quantification and curve fitting were carried out as described20. Km, kcat and (kcat/Km) of wild-type and S113A-mutant Pol η catalysing single-nucleotide incorporation using deoxyribonucleotide versus ribonucleotide primer are summarized in Supplementary Table 3 and plotted in Fig. 3b. Download references We thank D. Leahy, M. Gellert and R. Craigie for editing the manuscript. The research was supported by the intramural research program of the National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health (W.Y., T.N. and Y.Z.); the Japan Society for the Promotion of Science Institutional Program for Young Researcher Overseas visits, Kumamoto University, and the Kumayaku Alumni Research Fund (T.N.); Chinese Ministry of Education scholarship (Y.Z.); National Natural Science Foundation of China (Y.-J.H.); and Grant-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science, and Technology of Japan (Y.Y.). Atomic coordinates and structure factors for the reported crystal structures have been deposited with the Protein Data Bank with accession codes from 4ECQ to 4ECZ, 4ED0 to 4ED3, and 4ED6 to 4ED8.
What problem does this paper attempt to address?
-
Watching DNA polymerase η make a phosphodiester bond
Teruya Nakamura,Ye Zhao,Yuriko Yamagata,Yue-jin Hua,Wei Yang
DOI: https://doi.org/10.1038/nature11181
IF: 64.8
2012-01-01
Nature
Abstract:DNA synthesis has been extensively studied, but the chemical reaction itself has not been visualized. Here we follow the course of phosphodiester bond formation using time-resolved X-ray crystallography. Native human DNA polymerase η, DNA and dATP were co-crystallized at pH 6.0 without Mg 2+ . The polymerization reaction was initiated by exposing crystals to 1 mM Mg 2+ at pH 7.0, and stopped by freezing at desired time points for structural analysis. The substrates and two Mg 2+ ions are aligned within 40 s, but the bond formation is not evident until 80 s. From 80 to 300 s structures show a mixture of decreasing substrate and increasing product of the nucleotidyl-transfer reaction. Transient electron densities indicate that deprotonation and an accompanying C2′-endo to C3′-endo conversion of the nucleophile 3′-OH are rate limiting. A third Mg 2+ ion, which arrives with the new bond and stabilizes the intermediate state, may be an unappreciated feature of the two-metal-ion mechanism.
-
Mechanism of the Nucleotidyl-Transfer Reaction in DNA Polymerase Revealed by Time-Resolved Protein Crystallography
Teruya Nakamura,Ye Zhao,Yuriko Yamagata,Yue-Jin Hua,Wei Yang
DOI: https://doi.org/10.2142/biophysics.9.31
2013-01-01
BIOPHYSICS
Abstract:Nucleotidyl-transfer reaction catalyzed by DNA polymerase is a fundamental enzymatic reaction for DNA synthesis. Until now, a number of structural and kinetic studies on DNA polymerases have proposed a two-metalion mechanism of the nucleotidyl-transfer reaction. However, the actual reaction process has never been visualized. Recently, we have followed the nucleotidyl-transfer reaction process by human DNA polymerase η using time-resolved protein crystallography. In sequence, two Mg(2+) ions bind to the active site, the nucleophile 3'-OH is deprotonated, the deoxyribose at the primer end converts from C2'-endo to C3'-endo, and the nucleophile and the α-phosphate of the substrate dATP approach each other to form the new bond. In this process, we observed transient elements, which are a water molecule to deprotonate the 3'-OH and an additional Mg(2+) ion to stabilize the intermediate state. Particularly, the third Mg(2+) ion observed in this study may be a general feature of the two-metalion mechanism.
-
Primer terminal ribonucleotide alters the active site dynamics of DNA polymerase η and reduces DNA synthesis fidelity
Caleb Chang,Christie Lee Luo,Sarah Eleraky,Aaron Lin,Grace Zhou,Yang Gao
DOI: https://doi.org/10.1016/j.jbc.2023.102938
IF: 5.485
2023-03-01
Journal of Biological Chemistry
Abstract:DNA polymerases catalyze DNA synthesis with high efficiency, which is essential for all life. Extensive kinetic and structural efforts have been executed in exploring mechanisms of DNA polymerases, surrounding their kinetic pathway, catalytic mechanisms, and factors that dictate polymerase fidelity. Recent time-resolved crystallography studies on DNA polymerase η (Pol η) and β have revealed essential transient events during the DNA synthesis reaction, such as mechanisms of primer deprotonation, separated roles of the three metal ions, and conformational changes that disfavor incorporation of the incorrect substrate. DNA-embedded ribonucleotides (rNs) are the most common lesion on DNA and a major threat to genome integrity. While kinetics of rN incorporation has been explored and structural studies have revealed that DNA polymerases have a steric gate that destabilizes ribonucleotide triphosphate binding, the mechanism of extension upon rN addition remains poorly characterized. Using steady-state kinetics, static and time-resolved X-ray crystallography with Pol η as a model system, we showed that the extra hydroxyl group on the primer terminus does alter the dynamics of the polymerase active site as well as the catalysis and fidelity of DNA synthesis. During rN extension, Pol η error incorporation efficiency increases significantly across different sequence contexts. Finally, our systematic structural studies suggest that the rN at the primer end improves primer alignment and reduces barriers in C2'-endo to C3'-endo sugar conformational change. Overall, our work provides further mechanistic insights into the effects of rN incorporation on DNA synthesis.
biochemistry & molecular biology
-
Multiple deprotonation paths of the nucleophile 3′-OH in the DNA synthesis reaction
Mark T Gregory,Yang Gao,Qiang Cui,Wei Yang,Mark T. Gregory
DOI: https://doi.org/10.1073/pnas.2103990118
IF: 11.1
2021-06-04
Proceedings of the National Academy of Sciences
Abstract:Significance To determine if there is a general base in the DNA synthesis reaction that deprotonates the nucleophile, we systematically removed potential hydrogen-bond acceptors of the nucleophile, the 3′-OH of the primer strand. We then characterized the activity of human DNA Pol η by kinetic, structural, and molecular dynamics simulation analyses. We found that no single or combined perturbations eliminate catalysis. Moreover, removal of two potential proton acceptors of the 3′-OH by mutating the conserved S113 to Ala and addition of 2′-F to the primer end rescued the defects of the S113A mutation alone. Our results support that there is no specific general base and the proton is prone to leave the O3′ upon activation by the three Mg 2+ ions.
multidisciplinary sciences
-
DNA synthesis from diphosphate substrates by DNA polymerases
Cassandra R. Burke,Andrej Lupták
DOI: https://doi.org/10.1073/pnas.1712193115
IF: 11.1
2018-01-16
Proceedings of the National Academy of Sciences
Abstract:Significance All extant cellular organisms are thought to replicate their genomes using triphosphorylated substrates (dNTPs). Because only the α-phosphate is retained in the DNA backbone, both dNTPs and diphosphates (dNDPs) can in principle drive DNA synthesis. The activation barrier for the transphosphorylation is expected to be higher for dNDPs than for dNTPs, rendering the dNDP reactions slower; however, at elevated temperatures this penalty may be less prohibitive. We demonstrate DNA synthesis from dNDPs for a number of DNA polymerases, including bacterial and archaeal replicative and repair enzymes. Activation energy analysis of the forward (DNA synthesis) and reverse (phosphorolysis of DNA) reactions catalyzed by the Taq DNA polymerase shows that DNA synthesis is strongly favored, allowing surprisingly robust replication from low-energy substrates.
-
Structure of eukaryotic DNA polymerase δ bound to the PCNA clamp while encircling DNA
Fengwei Zheng,Roxana E Georgescu,Huilin Li,Michael E O'Donnell,Roxana E. Georgescu,Michael E. O’Donnell
DOI: https://doi.org/10.1073/pnas.2017637117
IF: 11.1
2020-11-17
Proceedings of the National Academy of Sciences
Abstract:Significance The structure of the eukaryotic chromosomal replicase, DNA polymerase (Pol) δ, was determined in complex with its cognate proliferating cell nuclear antigen (PCNA) sliding clamp on primed DNA. The results show that the Pol3 catalytic subunit binds atop the PCNA ring, and the two regulatory subunits of Pol δ, Pol31, and Pol32, are positioned off to the side of the Pol3 clamp. The catalytic Pol3 binds DNA and PCNA such as to thread the DNA straight through the circular PCNA clamp. Considering the large diameter of the PCNA clamp, there is room for water between DNA and the inner walls of PCNA, indicating the clamp “waterskates” on DNA during function with polymerase.
-
Real-time single-molecule studies of the motions of DNA polymerase fingers illuminate DNA synthesis mechanisms
Geraint W. Evans,Johannes Hohlbein,Timothy Craggs,Louise Aigrain,Achillefs N. Kapanidis
DOI: https://doi.org/10.1093/nar/gkv547
IF: 14.9
2015-05-26
Nucleic Acids Research
Abstract:DNA polymerases maintain genomic integrity by copying DNA with high fidelity. A conformational change important for fidelity is the motion of the polymerase fingers subdomain from an open to a closed conformation upon binding of a complementary nucleotide. We previously employed intra-protein single-molecule FRET on diffusing molecules to observe fingers conformations in polymerase-DNA complexes. Here, we used the same FRET ruler on surface-immobilized complexes to observe fingers-opening and closing of individual polymerase molecules in real time. Our results revealed the presence of intrinsic dynamics in the binary complex, characterized by slow fingers-closing and fast fingers-opening. When binary complexes were incubated with increasing concentrations of complementary nucleotide, the fingers-closing rate increased, strongly supporting an induced-fit model for nucleotide recognition. Meanwhile, the opening rate in ternary complexes with complementary nucleotide was 6 s(-1), much slower than either fingers closing or the rate-limiting step in the forward direction; this rate balance ensures that, after nucleotide binding and fingers-closing, nucleotide incorporation is overwhelmingly likely to occur. Our results for ternary complexes with a non-complementary dNTP confirmed the presence of a state corresponding to partially closed fingers and suggested a radically different rate balance regarding fingers transitions, which allows polymerase to achieve high fidelity.
biochemistry & molecular biology
-
Single base resolution in tunneling reads of DNA composition
Shuo Huang,Jin He,Shuai Chang,Peiming Zhang,Feng Liang,Shengqin Li,Michael Tuchband,Alexander Fuhrman,Robert Ros,Stuart Lindsay
2014-01-01
Abstract:Single-molecule DNA sequencing based on measuring the physical properties of bases as they pass through a nanopore1,2 eliminates the need for the enzymes and reagents used in other approaches. Theoretical calculations indicate that electron tunneling could identify bases in singlestranded DNA, yielding long reads and eliminating enzymatic processing.3–5 It was shown recently that tunneling can sense individual nucleotides6 and nucleosides.7 Here, we show that tunneling electrodes functionalized with recognition reagents can identify a single base flanked by other bases in a short DNA oligomer. The residence time of a single base in a recognition junction is on the order of a second, but pulling the DNA through the junction with a force of tens of piconewtons would yield reading speeds of tens of bases per second. Changes in the ion current through a nanopore can be used to identify translocating nucleotides. This opens the way to DNA sequencing if an exonuclease can pass each cleaved nucleotide into the pore sequentially.8 As an alternative, it has been proposed that the high spatial resolution of electron tunneling would allow direct reading of bases in an intact DNA polymer. 3–5 Recent progress in measuring electron tunneling through nucleotides or nucleosides shows that they can be identified by means of characteristic current signals.6,7 Recognition tunneling7,9 is an approach in which electrodes are functionalized with reagents that bind the target DNA bases. Contact via molecular adsorbates has been used to produce extraordinarily high spatial resolution in atomic force microscopy10 and, as we show here, single bases can be resolved in a DNA polymer when read by means of a selective chemical contact. Correpsondence and requests for material should be addressed to SL. Web Summary: Electron tunneling via functionalized electrodes can resolve and identify a single DNA base embedded in an oligomer. Author Contributions SH, SC and JH carried out tunneling measurements and characterized the samples. PZ, FL and Sq. L designed, synthesized and characterized reagents. MT prepared tunneling probes. AF and RR carried out force spectroscopy. SL designed experiments, analyzed data and wrote the paper. Competing Financial Interests SL, PZ and JH are named as inventors in patent applications. NIH Public Access Author Manuscript Nat Nanotechnol. Author manuscript; available in PMC 2014 August 04. Published in final edited form as: Nat Nanotechnol. 2010 December ; 5(12): 868–873. doi:10.1038/nnano.2010.213. N IH -P A A uhor M anscript N IH -P A A uhor M anscript N IH -P A A uhor M anscript To extend recognition tunneling to reads in buffered aqueous electrolyte, we synthesized the reagent 4-mercaptobenzamide (Fig. 1a and Methods) which presents two hydrogen-bond donor sites (on the nitrogen) and one hydrogen-bond acceptor site (the carbonyl). Likely binding modes to the four bases are shown in Fig. 2a.7 A gold (111) substrate and a partially-insulated gold STM probe were functionalized with this reagent (Methods and online supporting information) and characterized in an electron tunneling junction formed in a scanning tunneling microscope (PicoSPM, Agilent, Chandler, AZ). Fig. 1a shows a d(CCACC) oligomer trapped in a tunnel gap through hydrogen bonding to one mercaptobenzamide molecule on the probe and another on the substrate. In reality, the oligomer is probably held by many contacts, but only those that complete a short tunneling path (highlighted) will contribute significantly to the current. In our measurements, the probe is not deliberately scanned, but moves over the substrate as the microscope drifts. Alternatively, molecules may diffuse through the gap. Characteristic bursts of current are observed, and an example is shown in Fig. 1b. As we show below, the low frequency, large amplitude pulses indicate a C, while the high frequency, small amplitude pulses signal an A. Fig. 1c shows a sliding average of the spike amplitudes – values below the red line identify an A base unambiguously. Figure 1d shows a sliding average over the pulse frequencies (as defined for each adjacent pair of spikes) – the low frequency regions at each end enhance the confidence with which those regions can be assigned to a C base. The probability of an assignment to A (red line) or C (blue line) is shown in Fig. 1e. Calculation of these probabilities is based on our study of nucleotides, homopolymers and heteropolymers as described below. This example clearly shows that a single A base can be identified with high confidence when flanked by C bases in an intact DNA molecule. We first characterized the tunnel gap using doubly-distilled water and 0.1 mM phosphate buffer (PB – pH=7.4). Small signals were observed from buffer alone with bare electrodes, but they were much rarer when both electrodes were functionalized and the tunnel gap conductance set to 20 pS or less. (Fig. 2b and online supporting information). The tunnel decay was much more rapid (decay constant, β = 14.2±3.2 nm−1) with both electrodes functionalized than is the case in water alone (β ~ 6.1±0.7 nm−1 – 11 and online supporting information) and we estimate that the tunnel gap at i=10 pA and V = +0.5V is a little over the length of two benzamide molecules (i.e. a little greater than 2 nm). Introducing DNA nucleotides (10 μM in PB) into the tunnel gap yielded characteristic noise spikes as shown in Figs. 2c–f. The signal count rate (defined in Fig. 2k) varied considerably from 25 counts/s (5-methyl-deoxycytidine 5’-monophophate, dmCMP) to less than 1 c/s (deoxycytidine 5’-monophophate, dCMP). No signals were recorded at all with thymidine 5’-monophophate (dTMP), the signal looking exactly like the control (Fig. 2b). STM images suggest that this nucleotide binds to the surface (and presumably the probe) very strongly, blocking interactions in which a single molecule spans the junction. The current occurs in bursts of spikes (longer signal runs are given in online supporting information) and distributions of the spike heights were quite well fitted with two Gaussians distributions of the logarithm of current7 as shown in Figs 2 g–j (fitting parameters are given Supplementary Information accompanies this paper. Huang et al. Page 2 Nat Nanotechnol. Author manuscript; available in PMC 2014 August 04. N IH -P A A uhor M anscript N IH -P A A uhor M anscript N IH -P A A uhor M anscript in the online supporting information). These histograms were generated by counting only pulses that exceeded 1.5× the SD of the local noise background – i.e., typically pulses above 6 pA (a full description of the analysis procedure is given by Chang et el.7). dCMP generates the highest signals and the lowest count rate while deoxyadenosine 5’monophophate (dAMP) and dmCMP produce the smallest signals and the highest count rate (we found little difference between cytidine and 5-methylcytidine in organic solvent7 – supporting online information). The three bases with narrower pulse height distributions (dAMP, dmCMP and GMP) often show bursts of “telegraph-noise” characteristic of sources that fluctuate between two levels9 (particularly marked for dAMP). Such a two-level distribution is a strong indication that the tunneling signals are generated by a single molecule trapped in the tunnel junction.9 The characteristics of the tunneling noise from the nucleotides are summarized in Table 1. dAMP signals are well-separated from dCMP signals, and dmCMP signals are well separated from dCMP signals in spike amplitude and in the time distribution of their signals (Table 1 and online supporting information). For this reason, we chose to investigate DNA oligomers composed of A, C and mC bases. Figs. 3a,c and e show representative tunneling noise traces for d(A)5, d(C)5 and d(C)5 with the corresponding current peak distributions shown in Figs. 3b, d and f. Comparing Fig. 3b (d(A)5) with Fig. 2g (dAMP), Fig. 3d (d(C)5) with Fig. 2h (dCMP) and Fig. 3f (d(C)5) with Fig. 2i (dmCMP) leads to the following startling conclusion: most of the polymer binding events in the tunnel junction generate signals that resemble those generated by single nucleotides. That this should be so is not obvious. It requires (1) that single bases are being read and (2) that steric constraints owing to the polymer backbone do not prevent basebinding events from dominating the signals. There are some (small) differences between nucleotide and oligomer signals: (1) Peak positions, widths and relative intensities are altered somewhat (see online supporting information for details of the fits, and the also the nucleotide distributions which have been replotted on top of the homopolymer distributions as the black lines on Figs 3b,d and f.). (2) Almost all of the signals generated by nucleotides are less than 0.1 nA at 0.5V bias (Table 1). In contrast, 20% of the total signals generated by d(A)5 and d(C)5 are larger than 0.1 nA at this bias (Table 2 this is not obvious in Figure 2 where distributions are plotted only up to 0.1nA – the high current regions are shown in the online supporting information). These high current (>0.1 nA) features in d(A)5 and d(C)5 are continuously distributed so they do not represent parallel reads of more than one base at a time (where currents would be distributed in multiples of the single molecule values12). Rather, they are new features associated with the presence of the polymeric structure in the tunnel gap. Such a nonspecific, large amplitude spike is labeled by an asterisk in Fig. 1b. Features at I > 0.1 nA appear much less frequently in oligomers of mixed sequence, suggesting that they are associated with base-stacking in the homopolymers. Fig. 3h shows a current distribution for d(ACACA) where 95% of events are below 0.1 nA. Fig. 3j shows a current distribution for d(CmCCmCC) where 99% of events are below 0.1 nA. The solid red Huang et al. Page 3 Nat Nanotechnol. Author manuscript; available in PMC 2014 August
-
Unveiling the Mechanism of Deprotonation and Proton Transfer of DNA Polymerase Catalysis Via Single-Molecule Conductance.
Lihua Zhao,Yang Xu,Zhiheng Yang,Wenzhe Liu,Shichao Zhong,Jingwei Bai,Xuefeng Guo
DOI: https://doi.org/10.1002/advs.202408112
2024-01-01
Abstract:DNA polymerases (Pols) play important roles in the transmission of genetic information. Although the function and (de)regulation of Pols are linked to many human diseases, the key mechanism of 3'-OH deprotonation and the PPi formation are not totally clear. In this work, a method is presented to detect the full catalytic cycle of human Pol (hPol β) in graphene-molecule-graphene single-molecule junctions. Real-time in situ monitoring successfully revealed the spatial and temporal properties of the open and closed conformation states of hPol β, distinguishing the reaction states in the Pols catalytic cycle and unveiling 3'-OH deprotonation and pyrophosphate (PPi) formation mechanism of hPol β. Proton inventory experiment demonstrated that the rate-limiting step of PPi formation is deprotonation, which occurs before a reverse conformational change. Additionally, by detecting the acidity (pKa), it is found that MgA-bound OH- acted as a general base and activated the nucleophile of 3'-OH, and that acidic residue D190 or D192 coordinated with MgB as a proton donor to PPi. This work provides useful insights into a fundamental chemical reaction that impacts genome synthesis efficiency and Pol fidelity, which the discovery of Pol-targeting drugs and design of artificial Pols for DNA synthetic applications are expected to accelerated.
-
Unusual Base-Pairing Interactions in Monomer-Template Complexes.
Wen Zhang,Chun Pong Tam,Jiawei Wang,Jack W Szostak
DOI: https://doi.org/10.1021/acscentsci.6b00278
IF: 18.2
2016-01-01
ACS Central Science
Abstract:Many high-resolution crystal structures have contributed to our understanding of the reaction pathway for catalysis by DNA and RNA polymerases, but the structural basis of nonenzymatic template-directed RNA replication has not been studied in comparable detail. Here we present crystallographic studies of the binding of ribonucleotide monomers to RNA primer-template complexes, with the goal of improving our understanding of the mechanism of nonenzymatic RNA copying, and of catalysis by polymerases. To explore how activated ribonucleotides recognize and bind to RNA templates, we synthesized an unreactive phosphonate-linked pyrazole analogue of guanosine 5'-phosphoro-2-methylimidazolide (2-MeImpG), a highly activated nucleotide that has been used extensively to study nonenzymatic primer extension. We cocrystallized this analogue with structurally rigidified RNA primer-template complexes carrying single or multiple monomer binding sites, and obtained high-resolution X-ray structures of these complexes. In addition to Watson-Crick base pairing, we repeatedly observed noncanonical guanine:cytidine base pairs in our crystal structures. In most structures, the phosphate and leaving group moieties of the monomers were highly disordered, while in others the distance from O3' of the primer to the phosphorus of the incoming monomer was too great to allow for reaction. We suggest that these effects significantly influence the rate and fidelity of nonenzymatic RNA replication, and that even primitive ribozyme polymerases could enhance RNA replication by enforcing Watson-Crick base pairing between monomers and primer-template complexes, and by bringing the reactive functional groups into closer proximity.
-
Replication protein A dynamically re-organizes on primer/template junctions to permit DNA polymerase δ holoenzyme assembly and initiation of DNA synthesis
Jessica L Norris,Lindsey O Rogers,Kara G Pytko,Rachel L Dannenberg,Samuel Perreault,Vikas Kaushik,Sahiti Kuppa,Edwin Antony,Mark Hedglin
DOI: https://doi.org/10.1093/nar/gkae475
IF: 14.9
2024-06-07
Nucleic Acids Research
Abstract:DNA polymerase δ (pol δ) holoenzymes, comprised of pol δ and the processivity sliding clamp, PCNA, carry out DNA synthesis during lagging strand replication, initiation of leading strand replication, and the major DNA damage repair and tolerance pathways. Pol δ holoenzymes are assembled at primer/template (P/T) junctions and initiate DNA synthesis in a stepwise process involving the major single strand DNA (ssDNA)-binding protein complex, RPA, the processivity sliding clamp loader, RFC, PCNA and pol δ. During this process, the interactions of RPA, RFC and pol δ with a P/T junction all significantly overlap. A burning issue that has yet to be resolved is how these overlapping interactions are accommodated during this process. To address this, we design and utilize novel, ensemble FRET assays that continuously monitor the interactions of RPA, RFC, PCNA and pol δ with DNA as pol δ holoenzymes are assembled and initiate DNA synthesis. Results from the present study reveal that RPA remains engaged with P/T junctions throughout this process and the RPA•DNA complexes dynamically re-organize to allow successive binding of RFC and pol δ. These results have broad implications as they highlight and distinguish the functional consequences of dynamic RPA•DNA interactions in RPA-dependent DNA metabolic processes.
biochemistry & molecular biology
-
Control of DNA polymerase gp5 chain substitution by DNA double strand annealing pressure
Qi Jia,Qin-Kai Fan,Wen-Qing Hou,Chen-Guang Yang,Li-Bang Wang,Hao Wang,Chun-Hua Xu,Ming Li,Ying Lu,,
DOI: https://doi.org/10.7498/aps.70.20210707
IF: 0.906
2021-01-01
Acta Physica Sinica
Abstract:DNA polymerase is essential for DNA replication and repair. As it only performs the 5′-3′ polymerization, there are two kinds of DNA replication. One of them is called strand-displacement synthesis: DNA polymerase opens the double-strand (ds) DNA to attain the 3′-5′strand (leading strand) and copy this template in a continuous way, and the other is extension synthesis: DNA polymerase copies the newly separated 5′-3′ strand (lagging strand) in a discontinuous manner. The replication complex of T7 phage is an optimal model to investigate the mechanism of replication because it is only constituted by 4 terms of protein which are DNA helicase gp4, DNA polymerase gp5 with co-factor thioredoxin (Trx), and single-strand (ss) DNA-binding protein gp2.5. The replication complex of T7 encounters both strand-displacement synthesis and extension synthesis. Previous researches reported that gp5 can have rapid extension synthesis but lacks the ability to attain strand-displacement synthesis. It also reported that gp4 translocates on ssDNA at a rapid speed but unwinds dsDNA at a very low speed. However, gp5 and gp4 together can attain rapid and processive strand-displacement synthesis. Although extensively studied, this mechanism remains unclear. Here in this work, the dynamic of strand-displacement synthesis by gp5 is investigated with single-molecule Förster (fluorescence) resonance energy transfer (smFRET). It is found that gp5, without the help of external tension, can open dsDNA but only attain strand-displacement synthesis about 4 base pairs (bp), because its exonuclease activity excises the nascent nucleotides. Therefore gp5 repeats in the synthesis-excision cycle which results in the less production of strand-displacement synthesis. We conduct another control experiment by nano-tensioner, a high precision smFRET setup which can exert a tension on dsDNA, to change the dsDNA regression pressure on gp5. It is observed that reduced dsDNA regression pressure can increase the length of strand-displacement synthesis and reduce the length of excision which indicates that the dsDNA regression pressure can regulate the strand-displacement synthesis of gp5. The further experiment shows that after gp5 and gp4 are assembled into a replisome, it can have a processive strand-displacement synthesis and barely any excision presented. The speed of replisome is a little higher than gp5 alone but much higher than gp4 alone. Additionally, the length of strand-displacement synthesis by replisome is much longer than gp5 alone. Therefore it is indicated that the gp4 can reduce dsDNA regression pressure to enables gp5 to attain processive strand-displacement synthesis. On the other hand, the gp5 facilitates gp4 to unwind the dsDNA.
physics, multidisciplinary
-
Structural basis for the increased processivity of D-family DNA polymerases in complex with PCNA
Clément Madru,Ghislaine Henneke,Pierre Raia,Inès Hugonneau-Beaufet,Gérard Pehau-Arnaudet,Patrick England,Erik Lindahl,Marc Delarue,Marta Carroni,Ludovic Sauguet
DOI: https://doi.org/10.1038/s41467-020-15392-9
IF: 16.6
2020-03-27
Nature Communications
Abstract:Abstract Replicative DNA polymerases (DNAPs) have evolved the ability to copy the genome with high processivity and fidelity. In Eukarya and Archaea, the processivity of replicative DNAPs is greatly enhanced by its binding to the proliferative cell nuclear antigen (PCNA) that encircles the DNA. We determined the cryo-EM structure of the DNA-bound PolD–PCNA complex from Pyrococcus abyssi at 3.77 Å. Using an integrative structural biology approach — combining cryo-EM, X-ray crystallography, protein–protein interaction measurements, and activity assays — we describe the molecular basis for the interaction and cooperativity between a replicative DNAP and PCNA. PolD recruits PCNA via a complex mechanism, which requires two different PIP-boxes. We infer that the second PIP-box, which is shared with the eukaryotic Polα replicative DNAP, plays a dual role in binding either PCNA or primase, and could be a master switch between an initiation and a processive phase during replication.
multidisciplinary sciences
-
New Experimental Strategies for Estimating Rates of Conformational Changes in DNA Polymerases
MN Wang,GB Lu,XS Xie,WH Konigsberg
DOI: https://doi.org/10.1096/fasebj.20.4.a512
2006-01-01
Abstract:To explore the mechanism used by RB69 DNA polymerase (RB69 pol), several new experimental strategies were applied to determine rates of conformational changes by stopped-flow fluorescence (sfF). 2-Aminopurine, a fluorescent analog of adenine, was used in the template strand of primer-template (P/T) as a probe to detect conformational changes in an RB69 pol · P/T · dNTP complex as reflected in the extent of base stacking. When a ddP/T was employed, conformational changes could be detected after dNTP binding but before nucleotidyl transfer since this step cannot occur with a ddP/T. “Exchange-inert” metal ion complexes such as Rh(III)dNTP, were also used to determine conformational changes in the absence of chemistry when Mg2+ was not present in the reaction. Another approach involved using a P/T labeled with a cyanine dye (Cy3) appended at the 5′-end of the template strand. A single molecule fluorescence technique was developed for studying the kinetics and mechanism of nucleotide addition and adapted for RB69 pol. The bimolecular rate constant for the association of RB69 pol to P/T was ~2.2 X 107 M−1s−1. The maximum rates for two steps of conformational changes observed by sfF were 264±21 and 19.5±0.4 per second respectively with the second rate corresponding to the rate of nucleotidyl transfer as determined by chemical quench. The conformational change before chemistry, induced by dNTP binding, reflects the movement of DNA template as well as the closing of the fingers domain, which we have shown not be the rate-limiting step in RB69pol catalyzed primer-extension reactions. This research was supported by GM63276-04.
-
In crystallo observation of three metal ion promoted DNA polymerase misincorporation
Caleb Chang,Christie Lee Luo,Yang Gao
DOI: https://doi.org/10.1038/s41467-022-30005-3
IF: 16.6
2022-04-29
Nature Communications
Abstract:Abstract Error-free replication of DNA is essential for life. Despite the proofreading capability of several polymerases, intrinsic polymerase fidelity is in general much higher than what base-pairing energies can provide. Although researchers have investigated this long-standing question with kinetics, structural determination, and computational simulations, the structural factors that dictate polymerase fidelity are not fully resolved. Time-resolved crystallography has elucidated correct nucleotide incorporation and established a three-metal-ion-dependent catalytic mechanism for polymerases. Using X-ray time-resolved crystallography, we visualize the complete DNA misincorporation process catalyzed by DNA polymerase η. The resulting molecular snapshots suggest primer 3 ́-OH alignment mediated by A-site metal ion binding is the key step in substrate discrimination. Moreover, we observe that C-site metal ion binding preceded the nucleotidyl transfer reaction and demonstrate that the C-site metal ion is strictly required for misincorporation. Our results highlight the essential but separate roles of the three metal ions in DNA synthesis.
multidisciplinary sciences
-
DNA Replication across α-l-(3'-2')-Threofuranosyl Nucleotides Mediated by Human DNA Polymerase η
Rachana Tomar,Pratibha P Ghodke,Amritraj Patra,Elizabeth Smyth,Alexander Pontarelli,William Copp,F Peter Guengerich,John C Chaput,Christopher J Wilds,Michael P Stone,Martin Egli
DOI: https://doi.org/10.1021/acs.biochem.4c00387
IF: 3.321
2024-09-11
Biochemistry
Abstract:α-l-(3'-2')-Threofuranosyl nucleic acid (TNA) pairs with itself, cross-pairs with DNA and RNA, and shows promise as a tool in synthetic genetics, diagnostics, and oligonucleotide therapeutics. We studied in vitro primer insertion and extension reactions catalyzed by human trans-lesion synthesis (TLS) DNA polymerase η (hPol η) opposite a TNA-modified template strand without and in combination with O4-alkyl thymine lesions. Across TNA-T (tT), hPol η inserted mostly dAMP and dGMP, dTMP and dCMP with lower efficiencies, followed by extension of the primer to a full-length product. hPol η inserted dAMP opposite O4-methyl and -ethyl analogs of tT, albeit with reduced efficiencies relative to tT. Crystal structures of ternary hPol η complexes with template tT and O4-methyl tT at the insertion and extension stages demonstrated that the shorter backbone and different connectivity of TNA compared to DNA (3' → 2' versus 5' → 3', respectively) result in local differences in sugar orientations, adjacent phosphate spacings, and directions of glycosidic bonds. The 3'-OH of the primer's terminal thymine was positioned at 3.4 Å on average from the α-phosphate of the incoming dNTP, consistent with insertion opposite and extension past the TNA residue by hPol η. Conversely, the crystal structure of a ternary hPol η·DNA·tTTP complex revealed that the primer's terminal 3'-OH was too distant from the tTTP α-phosphate, consistent with the inability of the polymerase to incorporate TNA. Overall, our study provides a better understanding of the tolerance of a TLS DNA polymerase vis-à-vis unnatural nucleotides in the template and as the incoming nucleoside triphosphate.
-
Two-Metal-Ion Catalysis: Inhibition of DNA Polymerase Activity by a Third Divalent Metal Ion
Jimin Wang,William H. Konigsberg
DOI: https://doi.org/10.3389/fmolb.2022.824794
IF: 6.113
2022-03-01
Frontiers in Molecular Biosciences
Abstract:Almost all DNA polymerases (pols) exhibit bell-shaped activity curves as a function of both pH and Mg 2+ concentration. The pol activity is reduced when the pH deviates from the optimal value. When the pH is too low the concentration of a deprotonated general base (namely, the attacking 3′-hydroxyl of the 3′ terminal residue of the primer strand) is reduced exponentially. When the pH is too high the concentration of a protonated general acid (i.e., the leaving pyrophosphate group) is reduced. Similarly, the pol activity also decreases when the concentration of the divalent metal ions deviates from its optimal value: when it is too low, the binding of the two catalytic divalent metal ions required for the full activity is incomplete, and when it is too high a third divalent metal ion binds to pyrophosphate, keeping it in the replication complex longer and serving as a substrate for pyrophosphorylysis within the complex. Currently, there is a controversy about the role of the third metal ion which we will address in this review.
biochemistry & molecular biology
-
Bidentate and tridentate metal‐ion coordination states within ternary complexes of RB69 DNA polymerase
W. Konigsberg,S. Eom,Jimin Wang,S. Xia
DOI: https://doi.org/10.1002/pro.2026
IF: 8
2012-03-01
Protein Science
Abstract:Two divalent metal ions are required for primer‐extension catalyzed by DNA polymerases. One metal ion brings the 3′‐hydroxyl of the primer terminus and the α‐phosphorus atom of incoming dNTP together for bond formation so that the catalytically relevant conformation of the triphosphate tail of the dNTP is in an α,β,γ‐tridentate coordination complex with the second metal ion required for proper substrate alignment. A probable base selectivity mechanism derived from structural studies on Dpo4 suggests that the inability of mispaired dNTPs to form a substrate‐aligned, tridentate coordination complex could effectively cause the mispaired dNTPs to be rejected before catalysis. Nevertheless, we found that mispaired dNTPs can actually form a properly aligned tridentate coordination complex. However, complementary dNTPs occasionally form misaligned complexes with mutant RB69 DNA polymerases (RB69pols) that are not in a tridentate coordination state. Here, we report finding a β,γ‐bidentate coordination complex that contained the complementary dUpNpp opposite dA in the structure of a ternary complex formed by the wild type RB69pol at 1.88 Å resolution. Our observations suggest that several distinct metal‐ion coordination states can exist at the ground state in the polymerase active site and that base selectivity is unlikely to be based on metal‐ion coordination alone.
Chemistry,Medicine,Biology
-
A hand-off of DNA between archaeal polymerases allows high-fidelity replication to resume at a discrete intermediate three bases past 8-oxoguanine
Matthew T Cranford,Joseph D Kaszubowski,Michael A Trakselis
DOI: https://doi.org/10.1093/nar/gkaa803
IF: 14.9
2020-09-30
Nucleic Acids Research
Abstract:Abstract During DNA replication, the presence of 8-oxoguanine (8-oxoG) lesions in the template strand cause the high-fidelity (HiFi) DNA polymerase (Pol) to stall. An early response to 8-oxoG lesions involves ‘on-the-fly’ translesion synthesis (TLS), in which a specialized TLS Pol is recruited and replaces the stalled HiFi Pol for lesion bypass. The length of TLS must be long enough for effective bypass, but it must also be regulated to minimize replication errors by the TLS Pol. The exact position where the TLS Pol ends and the HiFi Pol resumes (i.e. the length of the TLS patch) has not been described. We use steady-state and pre-steady-state kinetic assays to characterize lesion bypass intermediates formed by different archaeal polymerase holoenzyme complexes that include PCNA123 and RFC. After bypass of 8-oxoG by TLS PolY, products accumulate at the template position three base pairs beyond the lesion. PolY is catalytically poor for subsequent extension from this +3 position beyond 8-oxoG, but this inefficiency is overcome by rapid extension of HiFi PolB1. The reciprocation of Pol activities at this intermediate indicates a defined position where TLS Pol extension is limited and where the DNA substrate is handed back to the HiFi Pol after bypass of 8-oxoG.
biochemistry & molecular biology
-
A new paradigm of DNA synthesis: three-metal-ion catalysis
Wei Yang,Peter J. Weng,Yang Gao
DOI: https://doi.org/10.1186/s13578-016-0118-2
2016-09-06
Abstract:Enzyme catalysis has been studied for over a century. How it actually occurs has not been visualized until recently. By combining in crystallo reaction and X-ray diffraction analysis of reaction intermediates, we have obtained unprecedented atomic details of the DNA synthesis process. Contrary to the established theory that enzyme-substrate complexes and transition states have identical atomic composition and catalysis occurs by the two-metal-ion mechanism, we have discovered that an additional divalent cation has to be captured en route to product formation. Unlike the canonical two metal ions, which are coordinated by DNA polymerases, this third metal ion is free of enzyme coordination. Its location between the α- and β-phosphates of dNTP suggests that the third metal ion may drive the phosphoryltransfer from the leaving group opposite to the 3′-OH nucleophile. Experimental data indicate that binding of the third metal ion may be the rate-limiting step in DNA synthesis and the free energy associated with the metal-ion binding can overcome the activation barrier to the DNA synthesis reaction.
biochemistry & molecular biology