Abstract
Thymine DNA Glycosylase (TDG) performs essential functions in maintaining genetic integrity and epigenetic regulation. Initiating base excision repair, TDG removes thymine from mutagenic G·T mispairs caused by 5-methylcytosine (mC) deamination and other lesions including uracil (U) and 5-hydroxymethyluracil (hmU). In DNA demethylation, TDG excises 5-formylcytosine (fC) and 5-carboxylcytosine (caC), which are generated from mC by Tet (ten–eleven translocation) enzymes. Using improved crystallization conditions, we solved high-resolution (up to 1.45 Å) structures of TDG enzyme–product complexes generated from substrates including G·U, G·T, G·hmU, G·fC and G·caC. The structures reveal many new features, including key water-mediated enzyme–substrate interactions. Together with nuclear magnetic resonance experiments, the structures demonstrate that TDG releases the excised base from its tight product complex with abasic DNA, contrary to previous reports. Moreover, DNA-free TDG exhibits no significant binding to free nucleobases (U, T, hmU), indicating a Kd >> 10 mM. The structures reveal a solvent-filled channel to the active site, which might facilitate dissociation of the excised base and enable caC excision, which involves solvent-mediated acid catalysis. Dissociation of the excised base allows TDG to bind the beta rather than the alpha anomer of the abasic sugar, which might stabilize the enzyme–product complex.
INTRODUCTION
Thymine DNA glycosylase (TDG) is a base excision repair (BER) enzyme that acts on derivatives of 5-methylcytosine (mC) arising from deamination or oxidation. TDG excises thymine from G·T mispairs, an activity that is needed to protect against C→T transition mutations caused by deamination of mC to T (1,2). It also plays an essential role in active DNA demethylation, which likely accounts for findings that its depletion in mice causes embryonic lethality (3,4). One established pathway for active DNA demethylation involves TDG excision of 5-formylcytosine or 5-carboxylcytosine (5,6), oxidation derivatives of mC generated by TET (ten–eleven translocation) enzymes (6–10). In addition to these key substrates, TDG removes many other bases (in vitro), including uracil, 5-halogenated uracils (5FU, 5ClU, 5BrU, 5IU) and 5-hydroxymethyl-U (hmU) (11,12). Human TDG is comprised of a catalytic domain of about 190 residues, flanked by two regions (about 110 residues each) that are disordered but important for certain functions, interactions with other proteins and regulation by post-translational modifications such as acetylation, phosphorylation and SUMO conjugation (13–18).
Crystal structures of TDG have revealed many details regarding its specificity and mechanism of catalysis and how SUMO proteins bind and alter its structure. The first reported structures were of SUMO-conjugated TDG (residues 117–332) (19,20); which defined its SUMO-interacting motif (SIM) and suggested a mechanism to explain findings that SUMO conjugation weakens its affinity for abasic DNA (18,21). The first structure of DNA-bound TDG was of the catalytic domain (TDGcat; residues 111–308), with a tetrahydrofuran (THF) abasic site analog flipped into its active site (22). This structure of the enzyme–product (E·P) complex revealed contacts with the mismatched guanine that may confer some specificity for G·T mispairs, and contacts with a flanking (3′) guanine that may impart specificity for a CpG dinucleotide context (22). However, because the E·P complex was prepared using the THF analog, the structure did not inform potential TDG interactions with the C1’ hydroxyl of the natural abasic site, nor did it reveal whether the excised base is retained in the product complex. Structures were subsequently reported for TDGcat bound to a G·U mispair (23) and to a G·caC pair (24), with non-cleavable (2′-fluoroarabino) deoxynucleotide analogs of U and caC flipped into the active site. These structures reveal contacts with the flipped base in the productive enzyme–substrate (E·S) complex.
Reported more recently was the structure of an E·P complex resulting from TDGcat action on a G·hmU mispair in DNA (25), which is remarkable in two respects. First, the excised hmU is reportedly trapped in the active site, despite its relatively low concentration, and, second, hmU is displaced from the abasic sugar, suggesting it moves to a new active-site location after bond cleavage (Figure 1). Given the conditions for the crystallization sample (0.2 mM G·hmU substrate, 0.35 mM TDGcat) (25) and the activity of TDG for hmU excision (kmax = 2 min-1) (11), the substrate was fully converted to products prior to crystallization, giving a 0.2 mM concentration of hmU and abasic DNA. If hmU remains bound in a ternary product complex, as reported, it would suggest that TDG possesses unusually high affinity for the excised base, given the relatively low hmU concentration in the crystallization sample and that crystals were cryoprotected by soaking in a solution that lacked hmU (25). Notably, for the bacterial MUG (mismatch-specific uracil glycosylase) enzymes, which are the closest homologs to TDG (32% sequence identity with Escherichia coli MUG) (22,26), the excised uracil is not observed in a structure (2.35 Å) of the E·P complex derived from MUG and a G·U substrate (27). While uracil DNA glycosylase (UNG) traps the excised uracil in the E·P complex (28,29), it binds uracil with unusually high affinity, due in part to a strong hydrogen bond (30,31), which is not observed for TDG. Structures of other glycosylases that feature the excised base in an E·P complex, including SMUG1 (single-strand selective monofunctional uracil-DNA glycosylase) (32) and TAG (3-methyladenine DNA glycosylase I) (33), were generated by growing or soaking crystals in the presence of a large molar excess of the nucleobase.
The reported structure of the TDGcat G·hmU product complex also suggests that, in a discrete step after glycosidic bond cleavage, the hmU base rotates and moves 5 Å relative to its expected pre-cleavage position, forming new active-site contacts at O2, N3 and O4 (25) (Figure 1). Although a structure of the E·S complex for G·hmU has not been reported, the pre-cleavage position of hmU likely resembles that of uracil in the E·S complex for a G·U mispair (Figure 1) (23,24). For UNG, the position of uracil and its active-site contacts are largely conserved in the pre- and post-excision (E·S and E·P) complexes (29,34). For TAG, contacts with 3-methyladenine (m3A) are the same in the binary complex (TAG·m3A) and a ternary complex that also includes abasic DNA (33,35,36). Similar findings are reported for MutY (37,38) and 8-oxoguanine DNA glycosylase (39,40). Thus, to our knowledge, the dramatic post-cleavage relocation of the hmU base proposed for TDG (25) would be unprecedented for a DNA glycosylase.
However, it is important to consider this possibility because a mechanism whereby the enzyme separates the excised base from the abasic sugar after C-N bond cleavage could potentially contribute to catalysis. DNA glycosylase reactions investigated to date follow a stepwise mechanism, where cleavage of the N-glycosidic bond yields a discrete though short-lived oxacarbenium ion intermediate, followed by nucleophile addition (41–45). For some glycosylases the first step is reversible and the C-N bond breaks and reforms repeatedly prior to irreversible nucleophile addition (41). As such, sequestration of the leaving group (LG) could potentially suppress reformation of the C-N bond and favor nucleophile addition, depending on the rate of LG displacement relative to that of nucleophile addition. Likewise, release of the excised base could also favor nucleophile addition, again depending on the rate. Thus, it is important to establish whether TDG retains the excised base in its product complex, and if the base is displaced away from the abasic sugar, for hmU and other substrates.
To address this problem, we developed improved crystallization conditions for TDG and solved high-resolution (up to 1.45 Å) structures of product complexes generated from substrates including G·hmU, G·T, G·U, G·fC and G·caC. The new crystallization conditions yield structures of DNA-bound TDG that are of much higher resolution than those previously reported (up to 2.49 Å), revealing details that were not observed in previous structures. These improved crystallization conditions are expected to facilitate future structural studies of TDG. We also studied the product complex in aqueous solution, using nuclear magnetic resonance (NMR) spectroscopy and investigated the affinity of TDG for isolated nucleobases (U, T, hmU). The results define the constituents and nature of the product complex for TDGcat, revealing previously unobserved features and informing its mechanism of catalysis. We propose a model whereby release of the excised base could tighten TDG binding to abasic DNA.
MATERIALS AND METHODS
Materials
Human TDG (full length), and its catalytic domain (TDGcat; residues 111–308), were expressed in E. coli and purified essentially as described (22,46). The enzymes were >99% pure as judged by sodium dodecyl sulphate-polyacrylamide gel electrophoresis (Coomassie stained gel) and concentrations were determined by absorbance (47,48). Uniformly 15N-labeled TDGcat was produced by expression in MOPS minimal media with 99% [15N]-NH4Cl (1 g/l) (C.I.L.) (49,50). Briefly, transformed BL21 (DE3) cells (Novagen) were grown overnight on an LB plate (37°C), then several colonies were used to inoculate 0.2 l of LB medium, and grown (37°C) to OD600 = 0.6. Cells were harvested, suspended in 2 l of MOPS minimal media, and grown to OD600 = 0.7. The temperature was reduced to 15°C, expression was induced with IPTG (0.4 mM) overnight (∼16 h).
Oligodeoxynucleotides (ODNs) were obtained (trityl-off) from the Keck Foundation Biotechnology Resource Laboratory, Yale University, or IDT. ODNs were purified by reverse phase HPLC using an XBridge OST C18 column (Waters Corp.), with mobile phases of 0.1 M TEAA pH 7.0, 5% acetonitrile (A) and 0.1 M TEAA pH 7 15% acetonitrile (B), a flow rate of 5 ml/min, and a gradient of 25 to 55% B over 16 min. Purified ODNs were exchanged into 0.02 M Tris–HCl pH 7.5, 0.04 M NaCl, and their concentration was determined by absorbance (46). ODNs containing the 2′-fluoroarabino analogs of deoxyuridine or deoxythymidine, referred to as UF and TF, respectively, were synthesized at Yale using phosphoramidites obtained from Glen Research (UF) or Link Technologies (TF) (51). TDG binds productively to UF and to TF but these analogs are fully resistant to glycosidic bond cleavage (48,51–53), because the subtle, single-atom substitution destabilizes the chemical transition-state.
DNA for crystal structures included a 28mer strand, 5′-AGCTGTCCATCGCTCAxGTACAGAGCTG, where x is the base excised by TDG, and a complement, 5′-CAGCTCTGTACGTGAGCGATGGACAGCT, that pairs G with the abasic site. The target base (x) is in a CpG context (underlined), consistent with TDG specificity (46,54). This construct was also used for electrophoretic mobility shift assays (EMSAs), with a 3′ 6-FAM on the complementary strand. DNA for NMR included a 23mer, 5′-CCACTGCTCAxGTACAGAGCTGT, where x is the abasic site resulting from TDGcat activity, and a complement, 5′- CAGCTCTGTACGTGAGCAGTGGA-3′, which pairs G with the abasic site, giving a 22-bp duplex with 3′ A or T overhangs and with the AP site in a CpG context (22).
X-ray crystallography
Samples used for crystallization contained 0.35 mM TDGcat and 0.42 mM DNA in a buffer of 5 mM Tris–HCl pH 7.5, 0.13 M NaCl, 0.2 mM dithiothreitol, 0.2 mM ethylenediaminetetraacetic acid. E·P complexes were produced by incubating TDGcat with DNA substrate for a sufficient time to ensure full conversion to product, confirmed by HPLC (11). Crystals were grown at room temperature (∼22°C) by sitting drop vapor diffusion, using 1 μl of the TDGcat-DNA sample and 1 or 2 µl of mother liquor, which was 30% (w/v) PEG 4000, 0.2 M ammonium acetate, 0.1 M sodium acetate, pH 6.0. Crystals typically appeared within in a few days. Crystals were cryo-protected using mother liquor supplemented with 18% ethylene glycol (and other components as noted), and flash cooled in liquid nitrogen.
X-ray diffraction data were collected at the Stanford Synchrotron Radiation Lightsource (SSRL; beamlines 7–1 and 12–2). The images were processed and scaled using Mosflm (55) and Aimless (56) from the CCP4 program suite (57). Alternatively, data processing was performed using the autoxds script developed by Ana Gonzalez and Yingssu Tsai (58) (http://smb.slac.stanford.edu/facilities/software/xds). Whenever possible, we took advantage of the shape of these crystals (long thin blades) to collect multiple datasets that could later be merged to increase resolution. Resolution cutoff was determined based on CC1/2 values (59). The structure of the product complex resulting from the G-hmU substrate (PDB ID: 4XEG) was solved by molecular replacement using Phaser (60) and a previous structure of DNA-bound TDGcat as the search model (PDB ID: 4FNC). All other structures reported were determined by molecular replacement using our new structure (PDB ID: 4XEG) as the search model. Refinement was performed using BUSTER-TNT (61) or REFMAC5 (62), and model building was performed using Coot (63). Specifically, every structural model was initially refined with REFMAC5 (restrained to ideal geometry and using isotropic B factors). Possible TLS groups were identified using the TLSMD server (64,65), and Rfree values were used to select the best model. The structure of an E·P complex derived from a G·U substrate (PDB ID: 4Z47) was deemed amenable to anisotropic B factor refinement, given that it produced similar reduction in both R and Rfree and that the sphericity distribution was consistent with expectations as predicted by the PARVATI server (66). Final models were refined using both REFMAC5 and BUSTER-TNT, and refinement protocol chosen based on lower Rfree values. While BUSTER-TNT refinement produced models with systematically better geometry, electron density maps did not show any significant differences that might have altered conclusions reached in this work. The structural figures were made with PyMOL (http://www.pymol.org).
NMR spectroscopy
15N-TROSY and 15N-HSQC experiments were collected on a 950 MHz Bruker Avance III NMR spectrometer, and the data were processed and analyzed using NMRPipe and NMRDraw (67). The sample conditions are provided in the relevant figure legends. For samples of TDGcat bound to DNA containing a pre-existing abasic (AP) site, the abasic site was generated by treating DNA containing a G·U mispair with a catalytic amount of UNG (1:500 molar ratio UNG:DNA). For all other E·P complexes, samples were generated by adding TDGcat to substrate DNA (G·U, G·T or G·hmU) and incubating sufficiently to ensure full conversion to product, as confirmed by HPLC (11).
Equilibrium binding assays
Equilibrium binding of TDG to a G·TF substrate analog, in the presence or absence of nucleobases (U, hmU or T), was analyzed using EMSAs, performed essentially as described (21). Note that TF is the 2′-fluoroarabino analogue of dT (described above). Samples contained 10 nM G·TF DNA, 5 nM to 1000 nM TDG and 10 mM of the nucleobase (where used). Samples were incubated at room temperature for 30 min, loaded onto a 6% native denaturing polyacrylamide gel and run for 60 min, 100 V at 4°C. Gels were imaged using a Typhoon 9400 variable mode imager (GE Healthcare).
RESULTS
New conditions for crystalizing DNA-bound TDG
The first structures of DNA-bound TDG were of E·S or E·P complexes, with the catalytic domain of human TDG (TDGcat) (22–24). Crystals were obtained under neutral pH conditions and the enzyme bound to the DNA with a stoichiometry of 2:1 (TDGcat:DNA), one subunit at the target site and a neighboring subunit bound to non-specific DNA (Condition I, Table 1). However, biochemical studies indicate that TDG binds DNA with a stoichiometry of 1:1 under limiting enzyme conditions and that a single TDG subunit possesses full catalytic activity (22,48,68). It seems likely that the 2:1 stoichiometry is attributable to the DNA construct used for crystallization, a 22-bp duplex with 3′-dA and 3′-dT overhangs that forms a long contiguous strand in the crystals. These conditions yielded crystals that were mostly of fairly poor quality, and many had to be screened to identify a few that diffracted with acceptable resolution (2.8–3.0 Å). A different approach for crystallizing DNA-bound TDGcat was reported by Hashimoto et al., yielding structures with improved resolution, up to 2.49 Å (Condition II, Table 1) (25,69). While these conditions yield a binding stoichiometry of 1:1, which is likely more biologically relevant than 2:1 binding (22,48,68), the pH is 4.6, and TDG activity is not detected at pH values below ∼5.5 (70,71).
Table 1. Crystallization conditions used for obtaining structures of DNA-bound TDGcat.
Condition I | Condition II | Condition III | |
---|---|---|---|
(PDB ID: 2RBA) | (PDB ID: 4FNC) | (PDB ID: 4Z47) | |
Space group | P65 | C2 | C2 |
Resolution (Å) | 2.79 | 2.49 | 1.45 |
Rwork/Rfree | 0.230/0.276 | 0.224/0.267 | 0.140/0.195 |
Wilson B-factor (Å2) | 87.2 | 58.7 | 23.5 |
No. of atoms, protein and DNA | 3847 | 2716 | 2767 |
No. of water molecules | 0 | 38 | 281 |
Stoichiometry (TDGcat:DNA) | 2:1 | 1:1 | 1:1 |
pH (mother liquor) | 7.0 | 4.6 | 6.0 |
DNA construct | 22 bp A/T overhang | 28 bp T overhang | 28 bp no overhang |
R.m.s.d to new structure (Å) | 0.40 | 0.27 |
For each type of crystallization condition, the PDB ID is given for the highest-resolution structure reported to date. R.m.s.d. values were obtained from alignment of the protein using PyMOL; for the 2:1 complex (PDB ID: 2RBA), the TDGcat subunit bound to the abasic analog was used for alignment.
Seeking to obtain higher quality crystals and conditions that are more biologically relevant, we modified the approach of Hashimoto et al., by altering the DNA and screening for conditions at pH values above 4.6. We obtained high-quality crystals at pH 5.5 and 6.0, and optimized the conditions at pH 6.0, where TDG exhibits full glycosylase activity (70,71). We solved structures of E·P complexes resulting from TDGcat action on DNA substrates including G·T, G·U, G·hmU, G·fC and G·caC. The resolution of these structures, up to 1.45 Å (Supplementary Table S1), is generally much better than that of any previously reported structure (Table 1). The new structures reveal hundreds of water molecules (Figure 2A, Table 1), far more than observed in previous structures of TDGcat; some mediate substrate binding and may function in catalysis (vide infra). As expected, these structures adopt a very similar overall fold relative to structures obtained from previous crystallization conditions (see RMSDs, Table 1).
We note that the 28 bp DNA used for these new structures differs slightly but significantly from that used by Hashimoto et al. (25,69), which had a 3′ dT overhang on the strand containing the target base, and a different sequence (Figure 2B). One or both of these differences, and/or the change in pH, may account for the improved resolution of our structures. Notably, the terminal dG of the target strand, and its dC partner in the complementary strand, are contacted by Arg213 and Arg209, respectively, from a symmetry-related molecule of TDGcat (Figure 2C and D). These contacts are also seen in the structure reported by Hashimito et al. (25), but the interaction might be somewhat destabilized by the 3′-dT overhang (only the 5'-phosphate of the dT overhang is present in the electron density for their structure). There is also a symmetry-related contact involving the side chain of His179 and a DNA backbone phosphate; the dinucleotide sequence containing this phosphate differs for our DNA (TpC) relative to that used by Hashimoto et al. (CpT) (Figure 2B) (25). Notably, structures reported here exhibit two slightly different conformations for the dC nucleotide contacted by His179, and for its 3′ (dG) neighbor.
The excised base dissociates from the product complex
To address the proposal that TDG traps the excised hmU base in the product complex, we consider our new structure of this complex, solved at 1.72 Å resolution (Figure 3A). The electron density is excellent and clearly defines the constituents of the active site, which include the flipped abasic nucleotide, as expected, and many water molecules, most not detected in previous structures. Importantly, the electron density demonstrates unambiguously that an acetate molecule, rather than the excised hmU base, resides in the TDGcat active site. Thus, hmU has dissociated from the product complex. The presence of acetate in the active site is likely explained by its interaction with three ordered water molecules and its high concentration (300 mM) in the mother liquor used for crystallization. Notably, the same acetate concentration was present in the mother liquor used by Hashimoto et al. for their structure of the E·P complex that was reported to contain the excised hmU base (PDB ID: 4FNC) (25). The electron density for PDB entry 4FNC, as reported by the electron density server (EDS) (72) and PDB_REDO (73) validation servers, is compatible with acetate and neighboring water molecules, though given the lower resolution (2.49 Å), interpretation may not be straightforward.
Given that hmU is clearly absent from the enzyme–product complex, we also determined the structure of product complexes resulting from the action of TDGcat on other substrates. Structures of E·P complexes generated from G·fC and G·caC substrates, solved at 2.02 Å and 2.45 Å resolution, respectively, give the same result (Supplementary Table S1. The active site contains the abasic sugar, several water molecules and an acetate molecule, but not the excised base (not shown).
We also investigated whether the nucleobase, if present at high concentration, might occupy the product complex. We crystallized E·P complexes that were generated by incubating TDGcat with a G·T, G·U, or G·hmU substrate in buffer that also contained a 10 mM concentration of the relevant nucleobase (T, U or hmU). Importantly, the same concentration of the nucleobase was also present in the mother liquor for crystallization and the solution used to cryoprotect the crystals. The resulting structures reveal that the nucleobase is not bound in the product complex, despite its high concentration during sample preparation, crystallization and cryoprotection. This is clearly demonstrated by the E·P complex generated from a G·U substrate (Figure 3B), solved at 1.45 Å resolution (Supplementary Table S1). The active site contains an acetate molecule, the abasic sugar and water molecules, but not the excised uracil base. Likewise, a 1.83 Å structure of the E·P complex generated from a G·T substrate (Supplementary Table S1) shows that thymine is absent from the active site, despite the presence of 10 mM thymine at all stages of crystal production (not shown). We note that the 10 mM concentration of the nucleobases used here for TDG is much higher than that used for structures of glycosylases that do bind their cognizant nucleobase in the E·P complex, including TAG, UNG and SMUG1 (28,29,32,33). Thus, TDG has negligible affinity for the excised base in the product complex.
Notably, all of the structures described above were generated from crystals that were grown using mother liquor that contained 300 mM acetate. To rule out the possibility that acetate displaces the excised base in the product complex, we grew crystals of a TDGcat product complex, generated from a G·U substrate, using mother liquor that did not contain acetate and solved a structure at 1.75 Å resolution (Supplementary Table S1). The electron density demonstrates that the active site does not contain acetate or the excised uracil base (Figure 3C). Rather, it contains only water molecules and the abasic sugar.
NMR studies indicate that the excised base dissociates from the product complex in solution
We also sought to determine whether the excised base is trapped in the TDGcat product complex in aqueous solution. Given the exquisite sensitivity of backbone amide 1H-15N resonances to structural and environmental changes, NMR chemical-shift-perturbation experiments are a powerful and widely used method to monitor protein–ligand interactions (74–76). Our approach was to compare the backbone 1H-15N chemical shifts of TDGcat in two different samples. The first sample included TDGcat with DNA containing a pre-existing abasic site (G·AP); the second sample was generated by incubating TDGcat with DNA substrate to give complete conversion to product (confirmed by HPLC). If the excised base dissociates from the product complex then there should be no substantial chemical shift differences because the resulting complexes would be identical (TDGcat bound to G·AP DNA). However, if the excised base is trapped in a ternary product complex and forms multiple interactions with active-site groups, as suggested (25), then substantial and numerous chemical shift perturbations should be observed relative to the binary complex of TDGcat bound to abasic DNA.
To obtain the baseline NMR spectrum, we collected a 15N-TROSY experiment for TDGcat bound to abasic DNA (Figure 4A, black peaks). We collected the same data for a product complex generated by incubating TDGcat with otherwise identical DNA that contained a G·hmU mispair (Figure 4A, red peaks). These two NMR spectra exhibit no substantial differences in backbone 1H-15N chemical shifts, indicating that the excised hmU base dissociates from the product complex. The same NMR experiments were collected for product complexes generated by incubating TDGcat with DNA containing a G·T mispair (Figure 4B, red peaks) or a G·U mispair (Figure 4C, red peaks). These spectra also exhibit no substantial differences relative to the baseline spectra of TDGcat bound to DNA containing a pre-existing abasic site (Figure 4B and C, black peaks), indicating that the excised base dissociates from the product complex generated by TDGcat action on a G·T or a G·U substrate. Thus, the NMR results are fully consistent with our crystallographic findings.
To illustrate the potential extent of chemical shift changes that might be observed if the excised base was retained in the product complex, we collected a TROSY spectrum for TDGcat bound to the identical DNA construct containing a G·UF mispair; UF is an analog of deoxyuridine that flips into the active site but is not excised by TDG (23,48,51,52). We find substantial chemical shift changes for dozens of backbone 1H-15N resonances of TDGcat bound to G·UF DNA (Figure 4D, red peaks) versus G·AP DNA (Figure 4D, black peaks). The predominant difference between these DNAs is the presence of uracil for G·UF but not G·AP DNA (Figure 4D). Two other differences that may account for some chemical shift changes are the hydroxyl at C1′ for the abasic nucleotide, and the 2′-fluoro of UF. Nevertheless, it is likely that at least some of the extensive shift perturbations are due to the presence of the uracil base in G·UF but not G·AP DNA, indicating that if the excised base were retained in the product complex then chemical shift changes would be expected in Figures 4A–C.
DNA-free TDG also exhibits negligible affinity for nucleobases that it excises from DNA
Given that that the nucleobases excised by TDG do not bind with significant affinity to the enzyme–product complex, we investigated their potential binding to DNA-free TDG. We performed NMR experiments for TDGcat in the absence and presence of uracil, thymine, or hmU at a concentration of 10 mM (Supplementary Figure S1). Remarkably, the NMR spectra reveal no significant differences, indicating that TDGcat has very weak affinity for these nucleobases, Kd >> 10 mM, even tough it readily excises them from DNA. We also investigated the potential interaction of these bases with full-length TDG, by examining their effect on TDG binding to DNA containing a G·TF mispair, where TF is an analog of deoxythymidine that flips into the active site but is not excised by TDG (23,48,51,52) (Figure 4D shows UF). Binding to G·T mispairs is weak relative to other substrates and is more sensitive to agents that perturb substrate binding (23,24,48,51,70). TDG binding to G·TF DNA is not detectably altered by the presence of uracil, thymine or hmU at high (10 mM) concentration (Supplementary Figure S2). Together, the results indicate that free TDG has negligible affinity for U, T or hmU.
Solvent-filled channel to the enzyme active site
The high-resolution structures reported here, which reveal hundreds of water molecules, indicate that the active site of DNA-bound enzyme is accessible to solvent. Specifically, the structures reveal a network of water molecules extending from the active site down along the target strand of the DNA toward the 5′ end (Figure 5). It seems reasonable that this channel could provide an escape route for the excised base, though this may require movement of the enzyme or DNA (or both). The channel may also be important for catalysis, given the previous finding that TDG excision of caC is acid catalyzed, involving a proton derived from solvent rather than a general acid of the enzyme (71). The solvent-filled channel revealed here for product complexes suggests a mechanism by which the caC base can be protonated (via solvent) in the enzyme–substrate complex.
Anomeric form of the abasic sugar in product complexes
Interestingly, our new structures indicate that absence of the excised base (or acetate) in the active site impacts the equilibrium between the two potential anomeric forms of the abasic sugar produced by TDG. Abasic sites adopt several potential forms in free DNA (Figure 6A), including a cyclic hemiacetal, a ring-opened aldehyde and a hydrated aldehyde (latter not shown). The cyclic hemiacetal predominates (99%), existing as a mix of α and β anomers that interconvert through the aldehyde (77–79). The α anomer is the initial form produced by TDG and other monofunctional DNA glycosylases that use water as the nucleophile (Figure 6A) (23,44). For product complexes crystallized in the presence of acetate, the α anomer clearly predominates (Figure 6B). However, for those crystallized under acetate-free conditions, the abasic sugar exists in a roughly equal mix of α and β anomers, each forming different interactions involving C1’-OH (Figure 6C). The β anomer contacts several ordered water molecules in the active-site, while the α anomer contacts the Oδ of Asn140 (Figure 6C), the side chain that coordinates the putative water nucleophile in the E·S complex (23). Interestingly, refinement indicates that a water molecule may occupy the site vacated by C1’-OH of the α but not the β anomer (not shown). This water molecule would interact with Asn140-Oδ, the backbone O of Thr197, and O4’ of the abasic sugar, some of which bind the putative water nucleophile in the E·S complex (23). Our findings show that dissociation of the excised base is required for TDG binding to the β anomer of the abasic sugar. They also raise the possibility that the abasic sugar equilibrates between the α and β anomers in the TDG active site, and that this equilibrium is strongly shifted towards α by the presence of acetate.
DISCUSSION
Improved crystallization conditions for TDG
Our new conditions for crystallizing DNA-bound TDGcat yield structures of unprecedented resolution for this enzyme, maintain the 1:1 binding stoichiometry observed by Hashimoto et al. (25,69), and are at pH 6.0, where TDG is fully active (70,71). The conditions yield crystals in a matter of days and only a few are typically needed to obtain high-quality data. As such, the approach may allow the determination of yet unsolved structures of DNA-bound TDG, and should make it feasible to obtain structures that could be mechanistically informative, i.e., of an active-site mutation, but which might not have been pursued using previous conditions, due to unpredictable crystal quality or moderate resolution. The new structures reveal hundreds of water molecules, some mediating enzyme–substrate interactions. For example, water molecules bridge enzyme contacts with the 5′-phosphate of the abasic nucleotide and the two phosphates 3′ of the abasic site (not shown), and a network of water molecules populate the active site. The structures reveal a solvent-filled channel to the active site (Figure 5), providing a potential pathway for proton transfer from bulk solution, which may be needed for acid catalysis of caC excision given that no enzyme group appears to perform general acid catalysis (71). Notably, most of the water-mediated enzyme–DNA contacts observed in our new structures were not detected in previous structures of DNA-bound TDG. Indeed, except for the putative water nucleophile, no water molecules were observed in the four structures obtained using the initial crystallization conditions (Type I, Table 1) (22–24).
The excised base dissociates from the product complex
Our structures demonstrate that the excised base is not retained in the enzyme–product complex, regardless of the substrate from which it is derived (G·hmU, G·U, G·T, G·fC or G·caC). These crystallographic findings are confirmed in solution by NMR, for product complexes generated from G·hmU, G·U and G·T substrates. Moreover, the same result is obtained even when the relevant base is supplemented to a concentration of 10 mM during sample preparation, crystal growth and cryoprotection. Our results are consistent with previous findings that the excised base (uracil) is not observed in a structure (2.35 Å) of the enzyme–product complex derived from a G·U substrate for MUG from E. coli (27), the most closely related glycosylase to TDG (32% identity) (22). Our structures also demonstrate that acetate resides in the active site of the E·P complex for crystals grown in the presence of acetate (300 mM). Moreover, crystals grown under acetate-free conditions yield structures of the E·P complex that contain only water molecules in the active site. Our findings conflict with a previous report that the excised hmU base remains trapped in the TDGcat product complex generated from a G·hmU substrate (25). The crystallization conditions used in the previous studies also included acetate (300 mM), and are very similar to those used here except for differences in pH and the DNA construct (Table 1, Figure 2). The high-resolution (up to 1.45 Å) structures here suggest that acetate and water molecules account for the electron density that had been attributed to the excised hmU base in the previous structure (PDB ID: 4FNC), solved at 2.49 Å resolution.
How does the excised base depart the product complex?
Observation that the excised base is absent from the product complex raises the question of how it is released. One plausible mechanism is that the excised base escapes through the solvent-filled channel revealed by our new structures, which could involve movement of the enzyme or DNA (or both). Notably, MUG has a similar though seemingly smaller channel, and it releases the excised uracil base from an E·P complex (27). By contrast, UNG, which traps the excised uracil in its product complex, lacks the channel found in TDG and MUG (29,34). UNG also forms a strong hydrogen bond with the excised uracil (30,31), which likely helps to retain it in the E·P complex. Alternatively, it is possible that TDG releases the excised base through a different region of the active site, but this would require substantial structural changes given that no other obvious escape route is evident (22,25). Another possibility is that release of the excised base depends on partial or full dissociation of abasic DNA, as we previously suggested (80). However, our new findings indicate that dissociation of the excised base might precede that of abasic DNA, by escaping through the channel. Additional studies are needed to establish the mechanism of base release from TDG. It will also be of interest to determine the kinetic parameters for dissociation of the excised base and abasic DNA.
TDG has negligible affinity for isolated nucleobases
Our findings reveal that TDG has no significant affinity for isolated nucleobases, even though it binds specifically to DNA containing these bases and readily excises them from DNA. TDG does not bind uracil, thymine or hmU, even at a 10 mM concentration, indicating very low affinity (Kd >> 10 mM). Likewise, enzyme–product complexes also lack significant affinity for free nucleobases. These results are consistent with previous findings that TDG is not inhibited by thymine, uracil, 5-fluorouracil and 3,N4-ethenocytosine (ϵC), at concentrations of 5 mM (80,81). Unlike TDG, many DNA glycosylases exhibit substantial affinity for free nucleobases. Fellow superfamily member UNG binds uracil with a Kd of about 0.1 mM (at pH 7.5) (30,82) and TAG binds m3A with a Kd of 0.04 mM (35). Our findings for TDG are also remarkable given that it exhibits high specificity for binding to DNA that contains the substrate base pairs examined here. For example, binding is nearly 500-fold tighter for DNA containing a G·U mispair (Kd = 0.6 nM) relative to non-specific DNA (Kd = ∼0.3 uM) (48). Studies of TDGcat indicate that, relative to G·U, binding is 4- and 10-fold tighter to G·fC and G·hmU, respectively (24). While some of this specificity is probably attributable to base-pairing properties, much is undoubtedly due to interactions with the flipped target base. Previous structures of E·S complexes (G·U, G·caC or A·caC) show that TDG forms contacts with the flipped base at positions including O2, N3, O4 or N4H2, and the carboxyl oxygen of caC (23,24), suggesting similar contacts for other substrates (hmU, T, 5-halo-U, fC). These contacts seem likely to help retain the flipped base in the E·S complex and/or stabilize the anionic form of the base that serves as the leaving group in the chemical step of the reaction. Yet, despite these contacts, TDG possesses no significant affinity for substrate bases that are isolated from DNA.
Dissociation of the excised base might stabilize the enzyme–product complex with abasic DNA
Our results, together with previous findings for other glycosylases, indicate that release of the excised base impacts the anomeric form adopted by the abasic sugar, which could potentially strengthen the product complex. Findings here indicate that the anomeric form of the flipped abasic sugar depends on whether the E·P complex is occupied with acetate. Structures with acetate exhibit only the α conformation, while those lacking acetate exhibit a roughly equal mix of α and β anomers. This is potentially relevant because the β anomer forms more contacts with TDG (including water mediated ones) than the α anomer (Figure 6), suggesting that TDG might bind tighter to the β versus α anomer. In support of this idea, the abasic sugar adopts only the β anomer in new structures of E·P complexes obtained using a construct of TDG that contains additional N-terminal residues relative to TDGcat (Malik, S.S., Coey, C.T., Pozharski, E. et al. manuscript in preparation). Our new structures here indicate that release of the excised base (and the absence of acetate) is essential for adoption of the β anomer, and this idea is supported by structures of other DNA glycosylases. A structure of the MUG product complex (2.35 Å) shows that the excised uracil is absent and the abasic sugar adopts the β anomer (27). Structures of E·P complexes for UNG show that the anomeric form of the abasic sugar depends on whether the excised base remains in the active site. Two structures that retain the excised uracil feature only the α-anomer (29,34), while the β anomer predominates in a structure for which the excised uracil base has dissociated (L272A-UNG variant) (29). Taken together, these observations raise the possibility that release of the excised base could potentially confer tighter binding of TDG to its abasic DNA product, by allowing formation of the β anomer.
ACCESSION NUMBERS
Coordinates and structure factors have been deposited in the Protein Data Bank (http://www.rcsb.org/) with accession numbers 4XEG, 4Z7B, 4Z7Z, 4Z47, 4Z3A and 5CYS.
Supplementary Material
Acknowledgments
We thank Braden Roth for collecting some of the NMR experiments. Portions of this research were carried out at the Stanford Synchrotron Radiation Lightsource, a Directorate of SLAC National Accelerator Laboratory and an Office of Science User Facility operated for the U.S. Department of Energy Office (DOE) by Stanford University.
SUPPLEMENTARY DATA
Supplementary Data are available at NAR Online.
FUNDING
Supported in part by National Institutes of Health (NIH) [grant R01GM72711 to ACD]. The SSRL Structural Molecular Biology Program is supported by the DOE Office of Biological and Environmental Research, the National Institute of General Medical Sciences [including P41GM103393]; and the National Center for Research Resources [P41RR001209]. The contents of this publication are solely the responsibility of the authors and do not necessarily represent the official views of NIGMS, NCRR or NIH. Funding for open access charge: NIH.
Conflict of interest statement. None declared.
REFERENCES
- 1.Neddermann P., Jiricny J. The purification of a mismatch-specific thymine-DNA glycosylase from HeLa cells. J. Biol. Chem. 1993;268:21218–21224. [PubMed] [Google Scholar]
- 2.Neddermann P., Gallinari P., Lettieri T., Schmid D., Truong O., Hsuan J.J., Wiebauer K., Jiricny J. Cloning and expression of human G/T mismatch-specific thymine-DNA glycosylase. J. Biol. Chem. 1996;271:12767–12774. doi: 10.1074/jbc.271.22.12767. [DOI] [PubMed] [Google Scholar]
- 3.Cortellino S., Xu J., Sannai M., Moore R., Caretti E., Cigliano A., Le Coz M., Devarajan K., Wessels A., Soprano D., et al. Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair. Cell. 2011;146:67–79. doi: 10.1016/j.cell.2011.06.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Cortazar D., Kunz C., Selfridge J., Lettieri T., Saito Y., Macdougall E., Wirz A., Schuermann D., Jacobs A.L., Siegrist F., et al. Embryonic lethal phenotype reveals a function of TDG in maintaining epigenetic stability. Nature. 2011;470:419–423. doi: 10.1038/nature09672. [DOI] [PubMed] [Google Scholar]
- 5.Maiti A., Drohat A.C. Thymine DNA glycosylase can rapidly excise 5-formylcytosine and 5-carboxylcytosine: potential implications for active demethylation of CpG sites. J. Biol. Chem. 2011;286:35334–35338. doi: 10.1074/jbc.C111.284620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.He Y.F., Li B.Z., Li Z., Liu P., Wang Y., Tang Q., Ding J., Jia Y., Chen Z., Li L., et al. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science. 2011;333:1303–1307. doi: 10.1126/science.1210944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Ito S., Shen L., Dai Q., Wu S.C., Collins L.B., Swenberg J.A., He C., Zhang Y. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science. 2011;333:1300–1303. doi: 10.1126/science.1210597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Pfaffeneder T., Hackner B., Truss M., Munzel M., Muller M., Deiml C.A., Hagemeier C., Carell T. The discovery of 5-formylcytosine in embryonic stem cell DNA. Angew. Chem. Int. Ed. Engl. 2011;50:7008–7012. doi: 10.1002/anie.201103899. [DOI] [PubMed] [Google Scholar]
- 9.Song C.X., Szulwach K.E., Dai Q., Fu Y., Mao S.Q., Lin L., Street C., Li Y., Poidevin M., Wu H., et al. Genome-wide profiling of 5-formylcytosine reveals its roles in epigenetic priming. Cell. 2013;153:678–691. doi: 10.1016/j.cell.2013.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Shen L., Wu H., Diep D., Yamaguchi S., D'Alessio A.C., Fung H.L., Zhang K., Zhang Y. Genome-wide analysis reveals TET- and TDG-dependent 5-methylcytosine oxidation dynamics. Cell. 2013;153:692–706. doi: 10.1016/j.cell.2013.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Bennett M.T., Rodgers M.T., Hebert A.S., Ruslander L.E., Eisele L., Drohat A.C. Specificity of human thymine DNA glycosylase depends on N-glycosidic bond stability. J. Am. Chem. Soc. 2006;128:12510–12519. doi: 10.1021/ja0634829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Cortazar D., Kunz C., Saito Y., Steinacher R., Schar P. The enigmatic thymine DNA glycosylase. DNA Repair (Amst) 2007;6:489–504. doi: 10.1016/j.dnarep.2006.10.013. [DOI] [PubMed] [Google Scholar]
- 13.Tini M., Benecke A., Um S.J., Torchia J., Evans R.M., Chambon P. Association of CBP/p300 acetylase and thymine DNA glycosylase links DNA repair and transcription. Mol. Cell. 2002;9:265–277. doi: 10.1016/s1097-2765(02)00453-7. [DOI] [PubMed] [Google Scholar]
- 14.Mohan R.D., Rao A., Gagliardi J., Tini M. SUMO-1-dependent allosteric regulation of thymine DNA glycosylase alters subnuclear localization and CBP/p300 recruitment. Mol. Cell. Biol. 2007;27:229–243. doi: 10.1128/MCB.00323-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Guan X., Madabushi A., Chang D.Y., Fitzgerald M., Shi G., Drohat A.C., Lu A.L. The human checkpoint sensor Rad9-Rad1-Hus1 interacts with and stimulates DNA repair enzyme TDG glycosylase. Nucleic Acids Res. 2007;35:6207–6218. doi: 10.1093/nar/gkm678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Madabushi A., Hwang B.J., Jin J., Lu A.L. Histone deacetylase SIRT1 modulates and deacetylates DNA base excision repair enzyme thymine DNA glycosylase. Biochem. J. 2013;456:89–98. doi: 10.1042/BJ20130670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Smet-Nocca C., Wieruszeski J.M., Chaar V., Leroy A., Benecke A. The thymine-DNA glycosylase regulatory domain: residual structure and DNA binding. Biochemistry. 2008;47:6519–6530. doi: 10.1021/bi7022283. [DOI] [PubMed] [Google Scholar]
- 18.Hardeland U., Steinacher R., Jiricny J., Schar P. Modification of the human thymine-DNA glycosylase by ubiquitin-like proteins facilitates enzymatic turnover. EMBO J. 2002;21:1456–1464. doi: 10.1093/emboj/21.6.1456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Baba D., Maita N., Jee J.-G., Uchimura Y., Saitoh H., Sugasawa K., Hanaoka F., Tochio H., Hiroaki H., Shirakawa M. Crystal structure of thymine DNA glycosylase conjugated to SUMO-1. Nature. 2005;435:979–982. doi: 10.1038/nature03634. [DOI] [PubMed] [Google Scholar]
- 20.Baba D., Maita N., Jee J.G., Uchimura Y., Saitoh H., Sugasawa K., Hanaoka F., Tochio H., Hiroaki H., Shirakawa M. Crystal structure of SUMO-3-modified thymine-DNA glycosylase. J. Mol. Biol. 2006;359:137–147. doi: 10.1016/j.jmb.2006.03.036. [DOI] [PubMed] [Google Scholar]
- 21.Coey C.T., Fitzgerald M.E., Maiti A., Reiter K.H., Guzzo C.M., Matunis M.J., Drohat A.C. E2-mediated small ubiquitin-like modifier (SUMO) modification of thymine DNA glycosylase is efficient but not selective for the enzyme-product complex. J. Biol. Chem. 2014;289:15810–15819. doi: 10.1074/jbc.M114.572081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Maiti A., Morgan M.T., Pozharski E., Drohat A.C. Crystal structure of human thymine DNA glycosylase bound to DNA elucidates sequence-specific mismatch recognition. Proc. Natl. Acad. Sci. U.S.A. 2008;105:8890–8895. doi: 10.1073/pnas.0711061105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Maiti A., Noon M.S., Mackerell A.D. Jr, Pozharski E., Drohat A.C. Lesion processing by a repair enzyme is severely curtailed by residues needed to prevent aberrant activity on undamaged DNA. Proc. Natl. Acad. Sci. U.S.A. 2012;109:8091–8096. doi: 10.1073/pnas.1201010109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Zhang L., Lu X., Lu J., Liang H., Dai Q., Xu G.L., Luo C., Jiang H., He C. Thymine DNA glycosylase specifically recognizes 5-carboxylcytosine-modified DNA. Nat. Chem. Biol. 2012;8:328–330. doi: 10.1038/nchembio.914. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Hashimoto H., Hong S., Bhagwat A.S., Zhang X., Cheng X. Excision of 5-hydroxymethyluracil and 5-carboxylcytosine by the thymine DNA glycosylase domain: its structural basis and implications for active DNA demethylation. Nucleic Acids Res. 2012;40:10203–10214. doi: 10.1093/nar/gks845. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Gallinari P., Jiricny J. A new class of uracil-DNA glycosylases related to human thymine-DNA glycosylase. Nature. 1996;383:735–738. doi: 10.1038/383735a0. [DOI] [PubMed] [Google Scholar]
- 27.Barrett T.E., Savva R., Panayotou G., Barlow T., Brown T., Jiricny J., Pearl L.H. Crystal structure of a G:T/U mismatch-specific DNA glycosylase: mismatch recognition by complementary-strand interactions. Cell. 1998;92:117–129. doi: 10.1016/s0092-8674(00)80904-6. [DOI] [PubMed] [Google Scholar]
- 28.Slupphaug G., Mol C.D., Kavli B., Arvai A.S., Krokan H.E., Tainer J.A. A nucleotide-flipping mechanism from the structure of human uracil-DNA glycosylase bound to DNA. Nature. 1996;384:87–92. doi: 10.1038/384087a0. [DOI] [PubMed] [Google Scholar]
- 29.Parikh S.S., Mol C.D., Slupphaug G., Bharati S., Krokan H.E., Tainer J.A. Base excision repair initiation revealed by crystal structures and binding kinetics of human uracil-DNA glycosylase with DNA. EMBO J. 1998;17:5214–5226. doi: 10.1093/emboj/17.17.5214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Drohat A.C., Stivers J.T. Escherichia coli uracil DNA glycosylase: NMR characterization of the short hydrogen bond from His187 to uracil O2. Biochemistry. 2000;39:11865–11875. doi: 10.1021/bi000922e. [DOI] [PubMed] [Google Scholar]
- 31.Drohat A.C., Stivers J.T. NMR evidence for an unusually low N1 pKa for uracil bound to uracil DNA glycosylase: implications for catalysis. J. Am. Chem. Soc. 2000;122:1840–1841. [Google Scholar]
- 32.Wibley J.E., Waters T.R., Haushalter K., Verdine G.L., Pearl L.H. Structure and specificity of the vertebrate anti-mutator uracil-DNA glycosylase SMUG1. Mol. Cell. 2003;11:1647–1659. doi: 10.1016/s1097-2765(03)00235-1. [DOI] [PubMed] [Google Scholar]
- 33.Metz A.H., Hollis T., Eichman B.F. DNA damage recognition and repair by 3-methyladenine DNA glycosylase I (TAG) EMBO J. 2007;26:2411–2420. doi: 10.1038/sj.emboj.7601649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Parikh S.S., Walcher G., Jones G.D., Slupphaug G., Krokan H.E., Blackburn G.M., Tainer J.A. Uracil-DNA glycosylase-DNA substrate and product structures: conformational strain promotes catalytic efficiency by coupled stereoelectronic effects. Proc. Natl. Acad. Sci. U.S.A. 2000;97:5083–5088. doi: 10.1073/pnas.97.10.5083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Cao C., Kwon K., Jiang Y.L., Drohat A.C., Stivers J.T. Solution structure and base perturbation studies reveal a novel mode of alkylated base recognition by 3-methyladenine DNA glycosylase I. J. Biol. Chem. 2003;278:48012–48020. doi: 10.1074/jbc.M307500200. [DOI] [PubMed] [Google Scholar]
- 36.Brooks S.C., Adhikary S., Rubinson E.H., Eichman B.F. Recent advances in the structural mechanisms of DNA glycosylases. Biochim. Biophys. Acta. 2013;1834:247–271. doi: 10.1016/j.bbapap.2012.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Guan Y., Manuel R.C., Arvai A.S., Parikh S.S., Mol C.D., Miller J.H., Lloyd S., Tainer J.A. MutY catalytic core, mutant and bound adenine structures define specificity for DNA repair enzyme superfamily. Nat. Struct. Biol. 1998;5:1058–1064. doi: 10.1038/4168. [DOI] [PubMed] [Google Scholar]
- 38.Lee S., Verdine G.L. Atomic substitution reveals the structural basis for substrate adenine recognition and removal by adenine DNA glycosylase. Proc. Natl. Acad. Sci. U.S.A. 2009;106:18497–18502. doi: 10.1073/pnas.0902908106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Bruner S.D., Norman D.P., Verdine G.L. Structural basis for recognition and repair of the endogenous mutagen 8-oxoguanine in DNA. Nature. 2000;403:859–866. doi: 10.1038/35002510. [DOI] [PubMed] [Google Scholar]
- 40.Fromme J.C., Bruner S.D., Yang W., Karplus M., Verdine G.L. Product-assisted catalysis in base-excision DNA repair. Nat. Struct. Biol. 2003;10:204–211. doi: 10.1038/nsb902. [DOI] [PubMed] [Google Scholar]
- 41.McCann J.A.B., Berti P.J. Transition-state analysis of the DNA repair enzyme MutY. J. Am. Chem. Soc. 2008;130:5789–5797. doi: 10.1021/ja711363s. [DOI] [PubMed] [Google Scholar]
- 42.Werner R.M., Stivers J.T. Kinetic isotope effect studies of the reaction catalyzed by uracil DNA glycosylase: evidence for an oxocarbenium ion-uracil anion intermediate. Biochemistry. 2000;39:14054–14064. doi: 10.1021/bi0018178. [DOI] [PubMed] [Google Scholar]
- 43.Berti P.J., McCann J.A. Toward a detailed understanding of base excision repair enzymes: transition state and mechanistic analyses of N-glycoside hydrolysis and N-glycoside transfer. Chem. Rev. 2006;106:506–555. doi: 10.1021/cr040461t. [DOI] [PubMed] [Google Scholar]
- 44.Drohat A.C., Maiti A. Mechanisms for enzymatic cleavage of the N-glycosidic bond in DNA. Org. Biomol. Chem. 2014;12:8367–8378. doi: 10.1039/c4ob01063a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Dinner A.R., Blackburn G.M., Karplus M. Uracil-DNA glycosylase acts by substrate autocatalysis. Nature. 2001;413:752–755. doi: 10.1038/35099587. [DOI] [PubMed] [Google Scholar]
- 46.Morgan M.T., Bennett M.T., Drohat A.C. Excision of 5-halogenated uracils by human thymine DNA glycosylase: robust activity for DNA contexts other than CpG. J. Biol. Chem. 2007;282:27578–27586. doi: 10.1074/jbc.M704253200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Gill S.C., von Hippel P.H. Calculation of protein extinction coefficients from amino acid sequence data. Anal. Biochem. 1989;182:319–326. doi: 10.1016/0003-2697(89)90602-7. [DOI] [PubMed] [Google Scholar]
- 48.Morgan M.T., Maiti A., Fitzgerald M.E., Drohat A.C. Stoichiometry and affinity for thymine DNA glycosylase binding to specific and nonspecific DNA. Nucleic Acids Res. 2011;39:2319–2329. doi: 10.1093/nar/gkq1164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Manvilla B.A., Varney K.M., Drohat A.C. Chemical shift assignments for human apurinic/apyrimidinic endonuclease 1. Biomol. NMR Assign. 2009;4:5–8. doi: 10.1007/s12104-009-9196-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Amburgey J.C., Abildgaard F., Starich M.R., Shah S., Hilt D.C., Weber D.J. 1H, 13C and 15N NMR assignments and solution secondary structure of rat Apo-S100 beta. J. Biomol. NMR. 1995;6:171–179. doi: 10.1007/BF00211781. [DOI] [PubMed] [Google Scholar]
- 51.Maiti A., Morgan M.T., Drohat A.C. Role of two strictly conserved residues in nucleotide flipping and N-glycosylic bond cleavage by human thymine DNA glycosylase. J. Biol. Chem. 2009;284:36680–36688. doi: 10.1074/jbc.M109.062356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Scharer O.D., Kawate T., Gallinari P., Jiricny J., Verdine G.L. Investigation of the mechanisms of DNA binding of the human G/T glycosylase using designed inhibitors. Proc. Natl. Acad. Sci. U.S.A. 1997;94:4878–4883. doi: 10.1073/pnas.94.10.4878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Barrett T.E., Scharer O.D., Savva R., Brown T., Jiricny J., Verdine G.L., Pearl L.H. Crystal structure of a thwarted mismatch glycosylase DNA repair complex. EMBO J. 1999;18:6599–6609. doi: 10.1093/emboj/18.23.6599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Waters T.R., Swann P.F. Kinetics of the action of thymine DNA glycosylase. J. Biol. Chem. 1998;273:20007–20014. doi: 10.1074/jbc.273.32.20007. [DOI] [PubMed] [Google Scholar]
- 55.Leslie A.W., Powell H. In: Evolving Methods for Macromolecular Crystallography. Read R, Sussman J, editors. Vol. 245. Netherlands: Springer; 2007. pp. 41–51. [Google Scholar]
- 56.Evans P.R. An introduction to data reduction: space-group determination, scaling and intensity statistics. Acta Crystallogr. D Biol. Crystallogr. 2011;67:282–292. doi: 10.1107/S090744491003982X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Winn M.D., Ballard C.C., Cowtan K.D., Dodson E.J., Emsley P., Evans P.R., Keegan R.M., Krissinel E.B., Leslie A.G., McCoy A., et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D Biol. Crystallogr. 2011;67:235–242. doi: 10.1107/S0907444910045749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Kabsch W. Xds. Acta Crystallogr. D Biol. Crystallogr. 2010;66:125–132. doi: 10.1107/S0907444909047337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Karplus P.A., Diederichs K. Linking crystallographic model and data quality. Science. 2012;336:1030–1033. doi: 10.1126/science.1218231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.McCoy A.J., Grosse-Kunstleve R.W., Storoni L.C., Read R.J. Likelihood-enhanced fast translation functions. Acta Crystallogr. D. 2005;61:458–464. doi: 10.1107/S0907444905001617. [DOI] [PubMed] [Google Scholar]
- 61.Bricogne G., Blanc E., Brandl M., Flensburg C., Keller P., Paciorek W., Rovers i.P., Sharff A., Smart O.S., Vonrhein C., et al. Cambridge: Global Phasing Ltd; 2011. [Google Scholar]
- 62.Winn M.D., Isupov M.N., Murshudov G.N. Use of TLS parameters to model anisotropic displacements in macromolecular refinement. Acta Crystallogr. D. 2001;57:122–133. doi: 10.1107/s0907444900014736. [DOI] [PubMed] [Google Scholar]
- 63.Emsley P., Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D. 2004;60:2126–2132. doi: 10.1107/S0907444904019158. [DOI] [PubMed] [Google Scholar]
- 64.Painter J., Merritt E.A. Optimal description of a protein structure in terms of multiple groups undergoing TLS motion. Acta Crystallogr. D Biol. Crystallogr. 2006;62:439–450. doi: 10.1107/S0907444906005270. [DOI] [PubMed] [Google Scholar]
- 65.Painter J., Merritt E.A. TLSMD web server for the generation of multi-group TLS models. J. Appl. Crystallogr. 2006;39:109–111. [Google Scholar]
- 66.Zucker F., Champ P.C., Merritt E.A. Validation of crystallographic models containing TLS or other descriptions of anisotropy. Acta Crystallogr. D. 2010;66:889–900. doi: 10.1107/S0907444910020421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Delaglio F., Grzesiek S., Vuister G.W., Zhu G., Pfeifer J., Bax A. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J. Biomol. NMR. 1995;6:277–293. doi: 10.1007/BF00197809. [DOI] [PubMed] [Google Scholar]
- 68.Buechner C.N., Maiti A., Drohat A.C., Tessmer I. Lesion search and recognition by thymine DNA glycosylase revealed by single molecule imaging. Nucleic Acids Res. 2015;43:2716–2729. doi: 10.1093/nar/gkv139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Hashimoto H., Zhang X., Cheng X. Activity and crystal structure of human thymine DNA glycosylase mutant N140A with 5-carboxylcytosine DNA at low pH. DNA Repair (Amst) 2013;12:535–540. doi: 10.1016/j.dnarep.2013.04.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Maiti A., Drohat A.C. Dependence of substrate binding and catalysis on pH, ionic strength, and temperature for thymine DNA glycosylase: Insights into recognition and processing of G.T mispairs. DNA Repair. 2011;10:545–553. doi: 10.1016/j.dnarep.2011.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Maiti A., Michelson A.Z., Armwood C.J., Lee J.K., Drohat A.C. Divergent mechanisms for enzymatic excision of 5-formylcytosine and 5-carboxylcytosine from DNA. J. Am. Chem. Soc. 2013;135:15813–15822. doi: 10.1021/ja406444x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Kleywegt G.J., Harris M.R., Zou J.Y., Taylor T.C., Wahlby A., Jones T.A. The Uppsala Electron-Density Server. Acta Crystallogr. D. 2004;60:2240–2249. doi: 10.1107/S0907444904013253. [DOI] [PubMed] [Google Scholar]
- 73.Joosten R.P., Long F., Murshudov G.N., Perrakis A. The PDB_REDO server for macromolecular structure model optimization. IUCrJ. 2014;1:213–220. doi: 10.1107/S2052252514009324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Bonvin A.M., Boelens R., Kaptein R. NMR analysis of protein interactions. Curr. Opin. Chem. Biol. 2005;9:501–508. doi: 10.1016/j.cbpa.2005.08.011. [DOI] [PubMed] [Google Scholar]
- 75.Zuiderweg E.R. Mapping protein-protein interactions in solution by NMR spectroscopy. Biochemistry. 2002;41:1–7. doi: 10.1021/bi011870b. [DOI] [PubMed] [Google Scholar]
- 76.Williamson M.P. Using chemical shift perturbation to characterise ligand binding. Prog. Nucl. Magn. Reson. Spectrosc. 2013;73:1–16. doi: 10.1016/j.pnmrs.2013.02.001. [DOI] [PubMed] [Google Scholar]
- 77.Wilde J.A., Bolton P.H., Mazumder A., Manoharan M., Gerlt J.A. Characterization of the equilibrating forms of the aldehydic abasic site in duplex DNA by oxygen-17 NMR. J. Am. Chem. Soc. 1989;111:1894–1896. [Google Scholar]
- 78.Manoharan M., Ransom S.C., Mazumder A., Gerlt J.A., Wilde J.A., Withka J.A., Bolton P.H. The characterization of abasic sites in DNA heteroduplexes by site specific labeling with carbon-13. J. Am. Chem. Soc. 1988;110:1620–1622. [Google Scholar]
- 79.Beger R.D., Bolton P.H. Structures of apurinic and apyrimidinic sites in duplex DNAs. J. Biol. Chem. 1998;273:15565–15573. doi: 10.1074/jbc.273.25.15565. [DOI] [PubMed] [Google Scholar]
- 80.Fitzgerald M.E., Drohat A.C. Coordinating the initial steps of base excision repair. Apurinic/apyrimidinic endonuclease 1 actively stimulates thymine DNA glycosylase by disrupting the product complex. J. Biol. Chem. 2008;283:32680–32690. doi: 10.1074/jbc.M805504200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Abu M., Waters T.R. The main role of human thymine-DNA glycosylase is removal of thymine produced by deamination of 5-methylcytosine and not removal of ethenocytosine. J. Biol. Chem. 2003;278:8739–8744. doi: 10.1074/jbc.M211084200. [DOI] [PubMed] [Google Scholar]
- 82.Drohat A.C., Jagadeesh J., Ferguson E., Stivers J.T. Role of electrophilic and general base catalysis in the mechanism of Escherichia coli uracil DNA glycosylase. Biochemistry. 1999;38:11866–11875. doi: 10.1021/bi9910878. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.