Citation: Ashwani Kumar Vashishtha, William H. Konigsberg. Effect of different divalent cations on the kinetics and fidelity of DNA polymerases[J]. AIMS Biophysics, 2018, 5(4): 272-289. doi: 10.3934/biophy.2018.4.272
[1] | Ashwani Kumar Vashishtha, William H. Konigsberg . The effect of different divalent cations on the kinetics and fidelity of Bacillus stearothermophilus DNA polymerase. AIMS Biophysics, 2018, 5(2): 125-143. doi: 10.3934/biophy.2018.2.125 |
[2] | Thomas D. Christian, William H. Konigsberg . Single-molecule FRET reveals proofreading complexes in the large fragment of Bacillus stearothermophilus DNA polymerase I. AIMS Biophysics, 2018, 5(2): 144-154. doi: 10.3934/biophy.2018.2.144 |
[3] | Roman Marks, Piotr H. Pawłowski . Rotational-electric principles of RNA/DNA and viability. AIMS Biophysics, 2023, 10(3): 385-400. doi: 10.3934/biophy.2023023 |
[4] | Arsha Moorthy, Alireza Sarvestani, Whitney Massock, Chamaree de Silva . Twisting and extension: Application of magnetic tweezers to DNA studies. AIMS Biophysics, 2023, 10(3): 317-346. doi: 10.3934/biophy.2023020 |
[5] | Davood Norouzi, Ataur Katebi, Feng Cui, Victor B. Zhurkin . Topological diversity of chromatin fibers: Interplay between nucleosome repeat length, DNA linking number and the level of transcription. AIMS Biophysics, 2015, 2(4): 613-629. doi: 10.3934/biophy.2015.4.613 |
[6] | Théo Lebeaupin, Hafida Sellou, Gyula Timinszky, Sébastien Huet . Chromatin dynamics at DNA breaks: what, how and why?. AIMS Biophysics, 2015, 2(4): 458-475. doi: 10.3934/biophy.2015.4.458 |
[7] | Raghvendra P Singh, Guillaume Brysbaert, Marc F Lensink, Fabrizio Cleri, Ralf Blossey . Kinetic proofreading of chromatin remodeling: from gene activation to gene repression and back. AIMS Biophysics, 2015, 2(4): 398-411. doi: 10.3934/biophy.2015.4.398 |
[8] | Roman Marks . Bubble mediated polymerization of RNA and DNA. AIMS Biophysics, 2022, 9(2): 96-107. doi: 10.3934/biophy.2022009 |
[9] | Larisa I. Fedoreyeva, Boris F. Vanyushin . Regulation of DNA methyltransferase gene expression by short peptides in nicotiana tabacum regenerants. AIMS Biophysics, 2021, 8(1): 66-79. doi: 10.3934/biophy.2021005 |
[10] | Nily Dan . Lipid-Nucleic Acid Supramolecular Complexes: Lipoplex Structure and the Kinetics of Formation. AIMS Biophysics, 2015, 2(2): 163-183. doi: 10.3934/biophy.2015.2.163 |
DNA polymerases are responsible for faithfully copying the genome during replication and repair [1,2,3,4,5]. DNA pols require two divalent metal ions Mg2+ or Mn2+ for primer extension and for removing the misincorporated dNTPs via the 3′-5′ exonuclease activity associated with certain DNA pols [6,7,8,9,10]. A two metal ion mechanism is used by all DNA pols to catalyze nucleotide addition to a growing primer strand [11]. Although DNA polymerases employ the physiologically relevant Mg2+, other divalent metal ions can substitute for Mg2+ though they tend to reduce the fidelity of DNA replication [12,13,14,15]. The effect of metal ion cofactors on the fidelity of DNA replication has been studied for various DNA pols including E. coli DNA pol I [16], AMV DNA pol [17], Klenow fragment of E. coli DNA pol I [18], T4 pol [12], T7 pol [12], human pol α [12], pol β [12] and Dpo4 [13]. Some metal ions have been shown to be mutagens and carcinogens probably because they reduce the base selectivity of DNA pols [12,13,16,17,18,19,20]. Different divalent cations influence fidelity check points in the minimal kinetic scheme for the nucleotidyl transfer reaction (Scheme 1). Cations that can substitute for Mg2+ affect DNA pols by: 1) altering the ground-state binding affinity of incoming dNTPs to pol:DNA binary complexes [21]; 2) decreasing base selectivity by promoting misincorporation during primer-extension [13]; 3) decreasing the rate of base excision [22]; 4) altering primer-extension past a mismatch at the primer-template (P/T) terminus [22]. This review will address the way in which various metal ions increase misincorporation based on their physical properties. Our emphasis will be on the effect of different cations on the behavior of DNA pols [12,13,16,17,18,23]. Previous reviews have summarized the effect of different divalent cations on the structure and function of pol β [24,25,26].
All DNA pols known to date have a basic requirement of two divalent metal ions in order to catalyze primer extension [11]. The metal ion present in the “A” site assists in lowering the pKa of the terminal 3′-OH group on the primer and coordinates both 3′-OH of the primer strand and the α-phosphate of the incoming dNTP which facilitates its nucleophilic attack on the incoming dNTP's α-phosphorous atom [11]. The metal ion, occupying the “B” site, coordinates the α-, β-, and γ- non-bridging phosphate oxygens of the incoming dNTP, helping to neutralize the developing negative charge as the ternary complex approaches the transition state in the nucleotidyl transfer reaction, and assists in the departure of the PPi product. Yang et al. [27] have shown that for pol β, both A and B metal ions are required to prepare the active site for nucleotidyl transfer and formation of binary dNTP/metal ion complex in the “B” site is unable to induce fingers closing, a necessary step for the phosphodiester bond formation to take place. They also showed that release of the catalytic metal ion triggers the opening of the fingers. Bakhtina et al. [28] have replaced Mg:dNTP binary complex with Rh(III)dNTP (an exchange-inert complex) and selectively studied the effect of filling the “A” site with Mg. Their results showed that for pol β, fingers closing could take place even in the absence of the “A” metal ion but the “A” metal ion is readily able to diffuse into the A site even with fingers in the closed state [28]. In contrast to the results obtained with pol β, similar studies carried out on RB69pol showed that, analogous to pol β, the fingers can close in the absence of an “A” metal ion but unlike pol β the metal ion A is unable to diffuse into the “A” site [21]. The fingers must reopen before the A metal ion can diffuse into the “A” site. Moreover, catalysis can take place only upon the binding of both A and B metal ions. Nakamura et al. [29] used time-resolved X-ray crystallography with a ternary complex containing human pol η, DNA, and dATP and observed that once the metal ions in the A and B site are bound, the 3′-OH group and the α-phosphate of the incoming dNTP get aligned for catalysis followed by the formation of a new bond. Interestingly, these authors showed that a third Mg2+ ion appears after the nucleotidyl transfer reaction was initiated with Mg2+ but before release of the products [29]. Yang et al. proposed that the third Mg2+ ion plays a crucial role in stabilizing the transition state by neutralizing the negative charges built up in the transition state and is likely involved in facilitating the protonation of the pyrophosphate leaving group [29]. The third metal ion was shown to be coordinated to four water molecules in addition to a bridging oxygen atom between the α- and β-phosphorous atoms and to the non-bridging oxygen of the α-phosphate [30]. Freudenthal et al. [31] reported similar results with pol β whereby they confirmed the presence of a third metal ion in the “C site” and showed that this presence is short lived. Recent studies by Gao et al. have clearly showed the occupancy of the C site with a Mn2+ ion and that the third metal ion was proposed to be required for catalysis by pol η [30].
A number of crystal structures have been solved with DNA pols bound to the incoming correct or incorrect dNTPs as well as the DNA substrate and different metal ions [11,14,30,32,33,34]. Xia et al. obtained the structure of the ternary complex of RB69pol in the presence of Mn2+ using dUpNpp which is a non-hydrolyzable dNTP analogue using the triple mutant (tm) variant (L561A/S565G/Y567A) [11]. The ternary complex structure showed that Mn2+ bound in both the “A” and “B” sites has octahedral geometry (Figure 1). Even though Mn2+ is in octahedral geometry in both sites, Mn2+ present in the “B” site is present in perfect octahedral geometry and is coordinated by the carboxylate side-chains of D411 and D623, and the backbone carbonyl oxygen of L412 along with the triphosphate tail of the incoming dUpNpp. On the other hand, Mn2+ bound in the “A” site, is present in a highly distorted octahedral geometry as Mn2+ in this structure is coordinated with the 3′-OH group of the primer, the oxygen of the incoming dNTP's α-phosphate and the two carboxylate side-chains of D411 and D623. In this structure, the distance between the Pα of the incoming dUpNpp and the 3′-OH group is too large to allow phosphodiester bond formation [11]. These authors proposed that as the reaction proceeds to the transition state, Mn2+ present in the “A” site helps to reduce the distance between the 3′-OH and the Pα of the incoming dNTP, allowing bond formation to occur [11]. In the same study, Xia et al. showed that when Mn2+ was replaced with Mg2+, Mg2+ present in the “B” site was present in perfect octahedral while the Mg2+ occupying the “A” site was present in a distorted octahedral geometry, results were very similar to those obtained with Mn2+ [11].
Johnson et al. have determined the crystal structures of mismatched base-pairs with BST pol (an A family pol) in the presence of Mg2+ and Mn2+ [35]. Besides BST pol, the structures of all 12 mispairs have also been reported for RB69pol (a B family pol) in the presence of Mg2+ [32]. These authors used a quadruple mutant (L415A/L561A/S565G/Y567A) to obtain the ternary complexes containing all 12 mispairs. In addition to RB69pol [11,36], T7 pol [37,38], Klenow fragment [39,40,41], and Dpo4 [13,15,42], pol β has also been extensively studied [14,24,25,26,28]. In summary, structural studies using various DNA pols shows that all DNA pols from different families have the same coordination geometry of the A and B metal ions in the active site.
Sirover et al. [20,43] screened a number of metal ions for their ability to act as cofactors for AMV DNA pol lacking the exonuclease activity. Following this, they also studied the effect of these metal ions on single base substitutions directly measuring how these metal ions affect the fidelity of AMV DNA pol. They utilized poly(A).oligo(dT) as the DNA substrate for their assays and found that of all the metal ions tested, eight metal ions including Ag+, Be2+, Cd2+, Co2+, Cr3+, Mn2+, Ni2+, and Pb2+ decreased the fidelity of AMV DNA pol by 30% or more [20]. These metal ions were identified as being mutagenic or carcinogenic [20]. Subsequent studies carried out by these authors showed that Mn2+ and Co2+ could activate E. coli DNA pol I [16,44]. Despite their ability to act as cofactors, these divalent cations enhanced misincorporation. For example, the efficiency of incorporation of dGMP opposite poly[d(A-T)] DNA substrate was invariant at various [Mg2+] while the efficiency was 2-3 fold higher at all concentrations of Co2+. When Mg2+ was replaced with Mn2+, the efficiency was 3-5 fold higher suggesting that Mn2+ readily permits misincorporation. Further studies by Seal et al. on human pol α and pol β [45] showed that both these pols can be activated by Mn2+, and Co2+ but at activating concentrations, these divalent cations enhance the frequency of misincorporation. Interestingly, at higher concentrations these metal ions actually inhibit the incorporation of the complementary nucleotide further affecting the fidelity of these pols. Later studies by Snow et al. showed that Ni2+ activates several different pols including T4 pol, Klenow fragment, AMV pol, human pol α, and T7 pol [12]. In the same study, they found that the base selectivity of different pols could be affected to different extents in the presence of Ni2+ [12].
Pelletier et al. [14] have crystallized pol β in the presence of blunt-end DNA, incoming dNTP and various metal ions. The structures of these ternary complexes in the presence of several different metal ions showed primer extension in crystals in the presence of Mn2+, Cd2+, and Zn2+ suggesting that these divalent cations were able to activate pol β. Egli et al. [42] carried out detailed steady-state kinetic studies using Dpo4 pol and tested a number of divalent cations and found that apart from Mg2+, only Mn2+ and Ca2+ could support primer extension with Dpo4, other divalent cations tested including Sr2+, Ba2+, Zn2+, Cu2+, Ni2+, and Co2+ failed to show any extension products. Later studies by Vashishtha et al. tested the competence of several divalent cations to determine their ability to activate RB69 pol [36]. These studies showed that, apart from Mg2+ and Mn2+, Co2+ and to a lesser extent Ni2+, were the only divalent cations which could support both pol and exo activities [36]. Recent work by Vashishtha et al. on BST pol showed that Mn2+, Co2+ and Cd2+ could replace Mg2+ in supporting the polymerase activity [46]. These studies showed that BST pol and pol β [14] are the only exceptions among all DNA pols which have the ability to utilize Cd2+. The metal ion preferences for different DNA pols are summarized in Table 1. From these studies it can be concluded that DNA pols from different families have the ability to utilize different metal ions for catalysis, but they do not follow a pattern that can be predicted from their physical properties.
DNA Polymerase (Pol Family) | Metal Ion | ||||||||||
Mn2+ | Co2+ | Fe2+ | Ni2+ | Zn2+ | Cd2+ | Sr2+ | Ba2+ | Cu2+ | Cr3+ | Ca2+ | |
DNA pol I (A) | + | + | - | + | - | - | - | - | - | - | - |
Human pol α (B) | + | + | - | - | + |
- | - | - | - | - | - |
pol β (X) | + | + |
- | - | + | + | - | - | - | - | - |
AMV DNA pol (RT) | + | + | - | + | - | - | - | - | - | - | - |
T4 pol (B) | + | + | - | + | - | - | - | - | - | - | - |
T7 pol (A) | + | + | - | + | - | - | - | - | - | - | - |
RB69pol (B) | + | + | - | + | - | - | - | - | - | - | - |
BST pol (A) | + | + |
- | + |
- |
+ |
- | - | - | - |
- |
Dpo4 pol (Y) | + | + |
- | - | - | - | - | - | - | - | + |
a Metal ion which activate the respective DNA polymerase are shown with + and those that are not able to support the polymerase activity are shown with -. b Studies by Pelletier et al. and Egli et al. previously claimed that human pol β and Dpo4 were not able to utilize Co2+ as cofactor however, recent studies by Vashishtha et al. showed that both pol β and Dpo4 can catalyze primer extension in the presence of Co2+ as explained in the text. c Vashishtha et al. [46] d Zhang et al. [64] |
The various steps involved in the phosphoryl transfer reaction involve binding of DNA pol to DNA to form the binary ED complex followed by addition of the incoming dNTP (correct or incorrect) to give the open ternary EDN complex (fingers open). This open complex then undergoes conformational changes to give the FDN closed complex (fingers closed) followed by the phosphoryl transfer reaction resulting in products FDn+1 and PPi. In this minimal kinetic scheme (Scheme 1), there are various fidelity checkpoints (steps in the minimal kinetic scheme) which the DNA pols employ to ensure accurate copying of template DNA minimizing mutations in the resulting DNA. These fidelity checkpoints include: 1) the conformational change of polymerase from open to a closed conformation when a correct incoming nucleotide is detected in the nascent base-pair binding pocket. 2) The chemistry step involving the formation of the phosphodiester bond which is much faster for a correct incoming nucleotide as compared to an incorrect incoming nucleotide and 3) removal of the incorrect dNMP incorporated in the growing primer strand via the exonuclease activity. Divalent metal ions can potentially affect these fidelity checkpoints and alter the fidelity of DNA replication by: 1) altering the ground-state binding affinity of incoming dNTPs (correct and incorrect) to DNA pol/P/T binary complexes [21]; 2) enhancing misincorporation by increasing the rate of incorporation of incorrect incoming dNTPs during primer-extension [13]; and 3) affecting the exonuclease activity [22]. In following sections, we will discuss the effect of divalent cations on these fidelity checkpoints.
Scheme 2 shows the various steps along the reaction pathway for the nucleotidyl transfer reaction. Kd,g indicates the ground-state binding affinity involving all the steps up to the formation of the closed FDN complex. A lower Kd,g value indicates higher binding affinity of the incoming dNTP towards the ED binary complex and vice versa. Typically, the Kd,g values are higher for an incoming incorrect dNTP as compared to the correct incoming dNTP in the presence of a given divalent cation [47]. Different divalent cations can potentially affect the Kd,g values for incoming dNTPs (correct or incorrect). Zhang et al. measured the Kd,g for dTTP binding opposite 2AP (which represents a correct nucleotide binding event) as the templating base (2AP at the n position) using a dideoxy-terminated P/T with RB69pol and they reported a Kd,g value of 9 µM in the presence of Mg2+ [47]. The basis of this equilibrium binding assay was the fact that 2AP in the ED binary complex exists in the unstacked form (high fluorescence state) and during the equilibrium fluorescence titration involving the formation of the EDN complex, 2AP becomes stacked resulting in quenching of the fluorescence signal [48,49]. The quenching of 2AP fluorescence is a function of [dNTP] and follows a hyperbolic function [48]. When 2AP was shifted to the n + 1 position, the fluorescence signal did not change during the titration of dCTP opposite dG as the templating base [47], consequently the Kd,g for binding of the incoming dNTPs could not be determined. Wang et al. [21] also carried out a similar study using RB69pol in the presence of Ca2+ with ddP/T containing 2AP at the n position and reported a Kd,g value of 53 nM for dTTP (correct). In contrast, the Kd,g value for the incorrect incoming dCTP was found to be 53 µM, representing a difference of 1000-fold in the ground-state binding affinities of the correct and incorrect dNTPs. These data suggest that binding of the incoming dNTP in the ground-state acts as a crucial fidelity checkpoint. Similar studies on T4 pol carried out by Hariharan et al. [48] in the presence of Mg2+ showed that the Kd,g value for dTTP binding opposite 2AP was 31 µM, very similar to that reported for RB69pol [21]. Vashishtha et al. [36] carried out detailed studies deciphering the effect of various divalent cations on ground-state binding affinity for incoming dNTPs. They reported the Kd,g values in the presence of different divalent cations which were shown to activate RB69pol including Co2+, and Mn2+, apart from Mg2+. Kd,g values were also reported in the presence of Ca2+ . Their data clearly showed that the identity of the divalent cation has a dramatic effect on the Kd,g values of correct and incorrect incoming dNTPs. The order of binding affinities for incoming dTTP opposite 2AP observed with different cations was: Ca2+ > Mn2+ > Co2+ > Mg2+ suggesting that the binding affinity was highest in the presence of Ca2+ and lowest in the presence of Mg2+. The Kd,g values showed an identical pattern when dTTP was replaced with dCTP as the incorrect incoming dNTP.
In a recent study by Vashishtha et al. [46] where they measured the Kd,g values in the presence of different divalent cations with BST pol, showed contrasting behavior as compared to their previous study carried out using RB69pol [36]. For example, with RB69pol, the Kd,g value for dTTP binding opposite 2AP was 5-fold lower as compared to Mg2+ but with BST pol, this value was 3-fold higher as compared to Mg2+ suggesting that with RB69pol dTTP binds much more tightly in the presence of Co2+ as compared to Mg2+ but surprisingly this trend is reversed with BST pol. When dTTP was replaced with dCTP (incorrect incoming nucleotide), Kd,g values were enhanced by 1400-fold in the presence of Mg2+, a behavior very similar to RB69pol but the difference in the Kd,g values for the correct and incorrect incoming dNTPs is much more pronounced with BST pol as compared to RB69pol (a difference of 1400-fold with BST pol compared to 40-fold for RB69pol) suggesting that in the presence of Mg2+, BST pol possesses a higher discrimination factor between the correct and incorrect incoming dNTPs as compared to RB69pol. Interestingly, the Kd,g value in the presence of Cd2+ was in between that of Mg2+ and Mn2+ [46]. These authors observed a very different behavior with Co2+ and Mn2+ where the Kd,g values for the correct and incorrect incoming dNTPs differed by only 8-fold and 3-fold with Co2+ and Mn2+ respectively. Overall, the Kd,g values obtained with BST pol showed a varied pattern as compared to similar values with RB69pol except with Mn2+ where the Kd,g values were substantially lower than Mg2+ for both pols. The variation in trends observed with these pols can be rationalized based on the fact that BST pol belongs to the A-family of DNA pols while RB69pol belongs to the B-family of DNA pols and the sequence diversity of these two DNA pols could be accountable for the difference in their ground-state binding affinities.
Studies in the past have shown that different divalent cations have a profound effect on the base selectivity of various pols and this effect varies with the nature of the pol as well as the identity of the divalent cation [18,19,50,51,52,53]. In general, when Mg2+ is substituted by Mn2+ the fidelity of DNA replication decreases, and this has been shown to be true for numerous DNA pols including T4 pol [19,54], T7 pol [37], E. coli DNA pol I [18,55], AMV DNA pol [52] and pol β [14]. Interestingly, the effect of Mn2+ on the fidelity of replication varies as a function of the metal ion concentration for example, studies by Beckman et al. showed that the fidelity of DNA replication is similar to that observed with Mg2+ at very low [Mn2+] (< 1 µM), while the fidelity decreased as the concentration of Mn2+ was elevated (< 100 µM) [55]. These authors proposed that the plausible reason for this dependence of fidelity on the metal ion concentration was that at elevated [Mn2+], Mn2+ could potentially bind to the DNA template affecting fidelity. Sirover et al. carried out similar studies using AMV DNA pol and showed that Co2+ and Ni2+ also affect the fidelity as a function of the metal ion concentration [17]. Studies by Hays et al. using T4 DNA pol showed that replacement of Mg2+ by Mn2+ enhanced the rate of dNMP incorporation opposite an abasic site by 11-34 fold [54]. Johnson et al. determined the structures of all 12 mismatched base-pairs for BST pol in the presence of Mg2+ and Mn2+ [35]. Moreover, in the presence of Mn2+ some mismatches were more readily incorporated than others leading to primer extension in crystals. Similar results were obtained by Bebenek et al. in solution [5]. Vashishtha et al. [36] used pre-steady-state kinetic studies to determine the effect of various divalent cations on the fidelity of replication of RB69pol which is a B-family pol and showed that apart from Mg2+, Co2+ and Mn2+ could act as cofactors for RB69pol catalyzed reactions. These authors showed that as compared to Mg2+ the incorporation efficiency for dTMP incorporation opposite dA was 5-fold higher in the presence of Co2+ and 3-fold higher in the presence of Mn2+ respectively. In the same study, the effect of divalent cations on base selectivity was studied using pyrimidine:pyrimidine, purine:pyrimidine and purine:purine mispairs. Surprisingly, base selectivity was also lower when Mg2+ was replaced with Co2+ but this decrease was still much less pronounced as compared to when Mn2+ replaced Mg2+ [36].
Recent studies by Vashishtha et al. [46] with BST pol which is an A-family pol showed that apart from Mg2+, Mn2+, and Co2+, Cd2+ could also serve as a cofactor for the polymerization reaction which is in contrast to similar studies using RB69pol [36] where Cd2+ failed to support nucleotidyl transfer reaction. On the other hand, Ni2+ which poorly activated RB69pol was not able to support nucleotidyl transfer with BST pol. They reported that the incorporation efficiency for the correct nucleotide incorporation (dTMP opposite dA) was 6-fold higher in the presence of Co2+ as compared to Mg2+ while this value was 8-fold higher with Mn2+ and slightly higher with Cd2+. Irrespective of the variation in kpol values, the Kd,app values were very similar irrespective of the identity of the divalent cation. Interestingly, the incorporation efficiency with both these pols representing two different pol families is much higher with Co2+ as compared to Mg2+.
The effect of different divalent cations on base selectivity was also studied using pyrimidine:pyrimidine and purine:pyrimidine mispairs. In contrast to RB69pol [36], the incorporation efficiencies for incorrect incoming nucleotides were ∼10-50 fold higher when Mg2+ was replaced with Co2+ suggesting that the base discrimination is greatly impacted with Co2+. Interestingly, the Kd,app values in the presence of Co2+ were lower as compared to Mg2+ with BST pol but the Kd,app values did not follow a clear trend with RB69pol [36]. A drastic impact on base selectivity was observed when Mg2+ was replaced with Mn2+ (kpol/Kd,app values were 13-1300 fold higher as compared to Mg2+). Interestingly, replacement of Mg2+ with Cd2+ resulted in a decrease in base selectivity [46]. These studies in conjunction with previous studies using long DNA substrates clearly show that among all divalent cations, Mn2+ is the most highly mutagenic metal ion and this behavior is a direct result of a sharp increase in the rate of incorporation (kpol or Vmax) accompanied by a simultaneous decrease in the Kd,app or Km values for incorrect incoming dNTPs [19,36].
During primer extension, if an incorrect nucleotide gets incorporated in the growing primer strand, the associated exonuclease activity removes the incorrect nucleotide. If the incorrect nucleotide is not removed, this results in errors being preserved during DNA replication. Loeb et al. and Snow et al. carried out extensive studies on the effect of metal ions on misincorporation with various DNA pols but they did not study the effect of metal ions on the efficiency of bypass past a mismatch [12,20,43]. Vashishtha et al. carried out comprehensive studies using RB69pol and determined the effect of different divalent cations on the ability of this pol to bury a mismatch [36]. They used DNA substrates containing either a dA/dC mismatch (representing a purine:pyrimidine mismatch which can form hydrogen bonds) or a dA/dG mismatch (representing a purine:purine bulky mispair incapable of forming hydrogen bonds between bases). In the presence of Mg2+, there was a substantial increase in the Kd,app value from 56 µM to 800 µM while the kpol value decreased sharply from > 300 s−1 to 1 s−1 for extension past the dA/dC mismatch, suggesting that the addition of a correct incoming nucleotide past the mismatch was very slow. With Mn2+, the Kd,app was similar to Mg2+, while the kpol was 17-fold higher, as compared to Mg2+. In the presence of Co2+, the Kd,app value was much higher than those obtained with Mg2+ and Mn2+. They observed similar trends with DNA containing the dA/dG mismatch in the presence of Mg2+, Mn2+, and Co2+, except that the Kd,app values were quite similar with all three divalent cations (300-500 µM). In general, the kpol values were 40-380-fold lower than those obtained with a dA/dC mispair.
Recent studies by Vashishtha et al. using BST pol examined the efficiency of mismatch bypass extension by this pol in the presence of Mg2+, Mn2+, Co2+ and Cd2+ [46]. They utilized a duplex DNA containing a dA/dC mismatch. In the presence of Mg2+, the kpol value dropped by 200-fold and the Kd,app value increased by 24-fold compared to the extension past a DNA containing matched DNA. This results in ∼4800-fold decrease in the efficiency of incorporation when a mismatch is encountered at the P/T terminus. When Mg2+ was replaced with Co2+, the incorporation efficiency past the mismatched DNA was higher than that obtained in the presence of Mg2+. Interestingly when Mg2+ was replaced by Mn2+ in the assay, the incorporation efficiency was 100-fold higher compared to Mg2+. Majority of this enhancement was a direct result of a sharp increase in the kpol value accompanied by a substantial decrease in the Kd,app value. Surprisingly, when Cd2+ was used in the assay, the incorporation efficiency past a dA/dC mismatch was very similar to that obtained with Mg2+ suggesting that Cd2+ does not promote misincorporation past a mismatch and behaves in a similar fashion to Mg2+ [46]. Based on the results obtained with RB69pol (a B-family pol) and BST pol (an A-family pol), in general, Co2+ and Mn2+ would be expected to increase the ability of DNA pols to bury a mismatch during DNA replication.
Similar to the polymerase active site, the exonuclease site also requires divalent cations to catalyze the 3′-5′ exonuclease activity associated with several DNA pols [5,36,38,41,56,57]. Divalent cations are required but have a different effect on the exonuclease activity. Results with RB69pol [36] showed that, compared to Mn2+ and Co2+, Mg2+ was most effective in promoting base excision but the exo rates varied only slightly among these three metal ions. Ni2+ on the other hand caused a dramatic decrease in exo activity (33-fold with Ni2+ vs. Mg2+). Similarly, the rates of base excision were reported to be nearly identical for E. coli DNA pol I with Mg2+, Mn2+ and Co2+ [16].
High fidelity DNA pols usually possess high base selectivity preventing misincorporation events during primer extension. Despite this, sometimes they do incorporate an incorrect nucleotide but the mechanism for this misincorporation is still unclear. In order to answer this question as to how DNA pols actually incorporate an incorrect incoming nucleotide, several groups have successfully crystallized ternary complexes containing the mismatched bases mimicking the cognate base-pairs [58,59]. Bebenek et al. have used a deletion mutant of pol λ which is an X-family DNA pol and crystallized the pol λ:DNA:dGTP ternary complex using a non-hydrolyzable analog of dGTP (dGMPCPP) opposite dT [59]. In this structure, the A and B metal ions overlay quite well with the corresponding structures containing cognate base-pairs. Also, the G:T primer-terminal base-pair occupies the same position corresponding to the correct base-pair with the exception of a minor twist in the template base position.
Tautomerization is a well-known phenomenon among certain base-pairs during the nucleotidyl transfer reaction [59,60,61]. Replicative DNA pols occasionally incorporate incorrect dNMPs [5] during the tautomerization reaction leading to the formation of high-energy tautomers [59,60,61]. Interestingly, structural evidence for these rare tautomers has been provided by Beese et al. whereby they observed these tautomers using a D598A/F710Y double mutant of BST pol in the presence of Mn2+ [58]. These authors showed that when ternary complex formation (BST pol:DNA:dNTP) takes place in the presence of Mn2+, the C/A mismatched base pair at the primer terminus adopts a tautomeric cognate base-pair shape, that is virtually indistinguishable from the canonical, Watson-crick base-pair in double stranded DNA at the insertion site [58]. Moreover, in the presence of Mn2+, the triphosphate tail was also properly aligned for catalytic reaction to take place, as well as BST pol was present in the “closed” conformation normally observed during correct nucleotide incorporation facilitating the misincorporation. When the reaction was studied in the presence of Mg2+, contrasting results were obtained as compared to Mn2+. For example, the C/A mismatch formed a non-cognate wobble base pair, and the BST pol was found to be in an “ajar” or partially closed conformation which hindered the incorporation of the incorrect dNMP. In addition, the triphosphate tail was also distorted misaligning the geometry required for the attack on the alpha phosphate of the incoming dNTP by the 3′-OH group of the primer. Together, these factors prevent the incorporation of an incorrect incoming nucleotide into the growing primer strand in the presence of Mg2+ [58].
Recent studies by Vashishtha et al. using wild type BST pol provided kinetic evidence for this rare tautomer hypothesis whereby they observed enhanced rates of incorporation of dAMP opposite dC in the presence of Mn2+ as compared to Mg2+ (kpol value was 130-fold higher in the presence of Mn2+) [46]. The formation of cognate base-pair mimicking the Watson-Crick base pairing along with the proper alignment of the triphosphate tail for nucleophilic attack as shown by Beese et al. could account for these enhanced rates obtained with incorrect dNMPS in the presence of Mn2+ as opposed to Mg2+ where the A/C mismatch forms a wobble base-pair and the pol is present in the Ajar confirmation resulting in misalignment of the residues involved in catalyzing the nucleotidyl transfer. The enhanced kpol values observed for dAMP incorporation opposite dC in the presence of Co2+ and Cd2+ could similarly be explained based on the speculation that the A/C mismatch adopts a tautomeric cognate base-pair shape and the triphosphate tail is also likely present in the proper alignment for nucleotidyl transfer. This proposal awaits further confirmation which will be tested once the Co2+ and Cd2+ bound ternary crystal structures become available. Indeed, the formation of cognate base-pair in the presence of these metal ions could explain the lower Kd,app values obtained with these divalent cations as opposed to Mg2+. These results provide a structural rationale for the mutagenic behavior of Mn2+.
In addition to Mg2+ most other DNA pols can also use Mn2+and Co2+, albeit with reduced fidelity [16,17,18,20,36]. Recent studies by Vashishtha et al. on BST pol have shown that this pol is also able to utilize Co2+ to extend the primer terminus [46]. Pelletier et al. and Egli et al. have shown that pol β [14] and Dpo4 [42] are the two known exceptions where the DNA pols cannot be activated by Co2+. In contrast, Vashishtha et al. [36] showed that these DNA pols can actually catalyze primer extension in the presence of Co2+. The apparent conflicting results can be rationalized based on the fact that different assay conditions were used by each of these groups. For example Pelletier et al. [14] used blunt-ended DNA with pol β, while Vashishtha et al. [36] used a P/T with a four base overhang 5′ to the templating base. In the Egli et al. experiments [42] 2 mM DTT was included in their assay with Dpo4 which reduced Co2+ to Co1+ (E0 = −0.33V for DTT vs. −0.28V for Co2+). DTT was omitted by Vashishtha et al. in their assays of Dpo4 so cobalt remained as Co2+ [36]. Based on these results it appears that Co2+ can support catalysis for all DNA pols that have been studied to date.
Several studies have shown that DNA pols from different families can potentially utilize various divalent cations as cofactors in order to carry out nucleotidyl transfer reaction [14,18,19,37,52,54,55,62]. It is quite surprising and puzzling as to why different DNA pols are able to utilize only certain divalent cations despite the fact that all DNA pols share a common active site containing the metal ion occupying the “A” site in distorted octahedral geometry while the one occupying the “B” site in perfect octahedral geometry [6,8,9,10,19,20,21,23,30,31,32,33,34,35,36,63].
Certain properties of divalent cations including their coordination geometry, their ionic radii, and their ability to lower the pKa of the 3′-OH group help to determine if a given divalent cation could serve as a cofactor for DNA pols (Table 2). A crucial factor in this regard is the ability of the given divalent cation to make the 3′-OH group of the primer strand more nucleophilic by lowering its pKa and facilitating its attack on the α-phosphate atom of the incoming dNTP. By comparing the ability of different divalent cations to lower the pKa of bound water (and analogously the pKa of the 3′-OH group) it appears that the pKa value is lowest for Fe2+ (8.4), similar for Cd2+, Co2+, and Zn2+ (9.8, 9.4, and 9.6 respectively), slightly higher for Ni2+, and Mn2+ (10.6 and 10.1), and much higher for Ca2+ and Mg2+ (12.8 and 11.4) (Table 2). Based on these values, Fe2+ is expected to support catalysis with DNA pols but surprisingly, Fe2+ fails to act as a cofactor for DNA pols including RB69pol [36] and BST pol [46]. Similarly, based on the values in Table 2, Co2+, Cd2+, and Zn2+ would be expected to be more effective as cofactors but this prediction is not borne out as is clear from the results obtained with various DNA pols [16,17,18,20,36,42,43,44]. In two separate studies, Vashishtha et al. showed that Zn2+ was not able to catalyze nucleotidyl transfer reaction with RB69pol [36] and BST pol [46]. Moreover, Ni2+ was also not able to support the polymerase activity of BST pol [46] despite having a comparable pKa value to Mn2+ but was a weak activator of RB69pol [36]. Based of its inability to lower the pKa of the 3′-OH group of the primer, Ca2+ is not able to support catalysis with DNA pols and Dpo4 is the only exception in this regard which is able to utilize Ca2+ as a cofactor albeit with poor incorporation efficiency compared to Mg2+ [11]. This suggests that the ability of divalent cations to lower the pKa of the 3′-OH group is not the sole factor for the ability of a given divalent cation to support catalysis for DNA pols, therefore other factors must be considered.
Besides the ability to lower the pKa of the 3′-OH group, the ionic radii of a given divalent cation plays a crucial role in determining its ability to support catalysis. The divalent cation present in the “A site” essentially determines the proximal distance between the 3′-hydroxyl group of the growing primer strand and the α-phosphorous atom of the incoming dNTP as the transition state is being approached leading to phosphodiester bond formation. The ionic radii of various divalent cations are summarized in Table 2. The ionic radii of Mg2+ is 0.86 Å, and Mg2+ is a universal activator of all DNA pols known to date. Based on this fact it would be expected that other divalent cations whose ionic radii are close to that of Mg2+ should also be able to support catalysis with various DNA pols. The ionic radii of Mn2+, Co2+, Ni2+ and Zn2+ are comparable to that of Mg2+, hence these divalent cations should be able to bring the 3′-hydroxyl group and α-phosphate atom of the incoming dNTP within the required distance for phosphodiester formation but the results show that Ni2+ and Zn2+ fail to activate
Parameter | Metal Ion | ||||||
Mg2+ | Mn2+ | Co+ | Ni2+ | Zn2+ | Cd2+ | Ca2+ | |
Ionic radius (Å) | 0.86 | 0.81 | 0.89 | 0.83 | 0.88 | 0.95 | 1.1 |
Coordination | Oct | Oct | Oct | Oct | Oct | Oct | Oct |
Geometry | Td | Td | Td | Td | Td |
Td | PBP |
Sq | Sq | Sq | TBP | TBP | HBP | ||
TBP | TBP | ||||||
pKa of the water molecule | 11.4 | 11.5 | 10.0 | 10.6 | 7.0 | 9.0 | 12.8 |
a Td represents Tetrahedral, Sq represents Square planar, TBP represents Trigonal bipyramidal, Oct represents Octahedral, PBP represents pentagonal bipyramid, and HBP represents hexagonal bipyramidal. b Even though Zn2+ can form Octahedral complexes, majority of Zn2+ complexes are tetrahedral. |
BST pol and Ni2+ is only able to weakly activate RB69pol [36]. The ionic radii of Fe2+ and Cd2+ (0.92 and 0.95 Å) are slightly larger than that of Mg2+ but surprisingly none of these metal ions support catalysis with RB69pol [36]. In contrast, Cd2+ is able to activate BST pol [46]. Moreover, with BST pol, the kpol value obtained in the presence of Cd2+ is comparable to that obtained with Mg2+ while Fe2+ whose ionic radii is very similar to Cd2+ fails to support phosphodiester bond formation. Ca2+ fails to activate all DNA pols except Dpo4 which is the only pol which can be weakly activated by this divalent cation [42]. The inability of Ca2+ to activate DNA pols can be rationalized on the basis of its larger ionic radii as compared to other divalent cations (1.1 Å for Ca2+ vs. 0.86 Å for Mg2+). This large ionic radii prevents Ca2+ occupying the “A” site from allowing the 3′-hydroxyl group and α-phosphate atom of the incoming dNTP to come close to the proximal distance for the phosphodiester bond formation [11].
Another important factor is the preference of different divalent cations for a given coordination geometry. Studies by Xia et al. [11] using metal ion (Mg2+ or Mn2+) bound complexes of a variant of RB69pol showed that the metal ion occupying the “B” site is present in a perfect octahedral geometry while the metal ion present in the “A” site is present in a distorted octahedral geometry. It is well known that other divalent cations including Co2+, Cd2+, Fe2+, and Ni2+ can also form octahedral complexes and hence can theoretically support the phosphoryl transfer reaction with DNA pols. Contrary to this expectation, studies by Vashishtha et al. have shown that only Co2+, and Ni2+ can support the reaction with RB69pol [36]. In contrast, studies with BST pol showed that Co2+, and Cd2+ could support catalysis but Fe2+, and Ni2+ are not able to act as cofactors for BST pol [46]. Interestingly, Ca2+ can also form octahedral complexes but since the ionic radii is much larger Ca2+ prefers pentagonal bipyramidal and hexagonal bipyramidal geometries over octahedral geometry which helps to explain the inability of Ca2+ to support catalysis with various pols except Dpo4 where Ca2+ acts as a weak activator [42]. Based on its ability to effectively lower the pKa of hydroxyl group and its similar ionic radii compared to Mg2+, Zn2+ is expected to act as a cofactor for DNA pols but Zn2+ fails to catalyze the phosphoryl transfer reaction except with pol β [14] and pol α [64]. The inability of Zn2+ to act as a cofactor for DNA pols could be explained on the basis of its preference for tetrahedral geometry as opposed to octahedral geometry required for metal ions bound in the “A” and “B” sites. The ability of only selected divalent cations (including Mg2+, Co2+, Mn2+, and Cd2+ for BST pol and Mg2+, Co2+, Mn2+, and Ni2+ for RB69pol) to act as cofactors for these pols is consistent with: 1) the ability of these metal ions to form octahedral complexes [34]; 2) similar ionic radii of these metal ions and; 3) their ability to effectively lower the pKa of the water molecule (and presumably the 3′-hydroxyl group of the primer).
As discussed above, DNA pols are susceptible to making errors during DNA replication resulting in mutations being preserved in the DNA if not corrected by the exonuclease activity. It is important to study these mutations in cases such as tumors and several types of cancers [65]. Frederico et al. have used the M13 reversion assay to determine the rate of cytosine deamination in DNA [66]. The basis of this assay is the reversion of a mutant in the lacZα gene coding sequence of bacteriophage M13mp2 [67]. Recent developments in DNA sequencing have led to the evolution of a new era of DNA sequencing known as Next-generation DNA sequencing (NGS) which has revolutionized sequencing both in terms of the cost as well as the amount of data generated from sequencing [65,68]. Schmitt et al. have developed a method called “Duplex Sequencing” which is based on individually tagging and sequencing each of the two DNA strands resulting in a theoretical background error rate of less than one mutation per billion nucleotides sequenced [65]. These authors have used this method to determine the frequency and pattern of random mutagenesis in mitochondrial DNA from human cells underlying the importance of this technique. Interestingly, Yasukawa et al. have also developed a simple and rapid method to determine the fidelity of reverse transcriptase using NGS [68,69].
DNA pols from different families are able to catalyze primer extension utilizing different divalent cations. Both Cd2+, and Zn2+ fail to support catalysis with RB69pol [36], while Zn2+ is unable to activate BST pol [46]. Interestingly, crystal soaking experiments with pol β have shown that Cd2+, and Zn2+ could support catalysis leading to primer extension with blunt-end DNA [14]. Ni2+ has been shown to support primer- extension with all DNA pols albeit with greatly reduced activity except for Dpo4 [42], human pol α [50], pol β [14] and BST pol [46]. On the other hand, Co2+ is able to activate all DNA pols known till date and detailed pre-steady-state kinetic studies showed that Co2+ is in fact better than Mg2+ in terms of its ability to support correct nucleotide incorporation. Moreover, Ca2+ is unable to support catalysis with all DNA pols known except for Dpo4, although, Ca2+ is very poor compared to Mg2+ [42]. Thus, it is difficult to generalize as to which DNA pol can be activated by specific divalent cations solely based on their properties.
This work was supported by the Frank and Suzanne Konigsberg Research Fund. We would like to thank Suzanne Fields for her help with MS preparation.
The authors declare no conflict of interest.
[1] |
Echols H, Goodman MF (1991) Fidelity mechanisms in DNA replication. Annu Rev Biochem 60: 477–511. doi: 10.1146/annurev.bi.60.070191.002401
![]() |
[2] |
Johnson KA (1993) Conformational coupling in DNA polymerase fidelity. Annu Rev Biochem 62: 685–713. doi: 10.1146/annurev.bi.62.070193.003345
![]() |
[3] |
Joyce CM, Benkovic SJ (2004) DNA polymerase fidelity: kinetics, structure, and checkpoints. Biochemistry 43: 14317–14324. doi: 10.1021/bi048422z
![]() |
[4] |
Kunkel T, Bebenek K (1988) Recent studies of the fidelity of DNA synthesis. BBA-Gene Struct Expr 951: 1–15. doi: 10.1016/0167-4781(88)90020-6
![]() |
[5] |
Kunkel TA, Bebenek K (2000) DNA replication fidelity. Annu Rev Biochem 69: 497–529. doi: 10.1146/annurev.biochem.69.1.497
![]() |
[6] |
Drake JW (1969) Comparative rates of spontaneous mutation. Nature 221: 1132. doi: 10.1038/2211132a0
![]() |
[7] | Goodman MF, Tippin B (2000) The expanding polymerase universe. Nat Rev Mol Cell Biol 1: 101–109. |
[8] |
Filee J, Forterre P, Sen-Lin T, et al. (2002) Evolution of DNA polymerase families: evidences for multiple gene exchange between cellular and viral proteins. J Mol Evol 54: 763–773. doi: 10.1007/s00239-001-0078-x
![]() |
[9] |
Vashishtha AK, Kuchta RD (2016) Effects of acyclovir, foscarnet, and ribonucleotides on Herpes Simplex Virus-1 DNA Polymerase: Mechanistic insights and a novel mechanism for preventing stable incorporation of ribonucleotides into DNA. Biochemistry 55: 1168–1177. doi: 10.1021/acs.biochem.6b00065
![]() |
[10] |
Vashishtha AK, Kuchta RD (2015) Polymerase and exonuclease activities in herpes simplex virus type 1 DNA polymerase are not highly coordinated. Biochemistry 54: 240–249. doi: 10.1021/bi500840v
![]() |
[11] |
Xia S, Wang M, Blaha G, et al. (2011) Structural insights into complete metal ion coordination from ternary complexes of B family RB69 DNA polymerase. Biochemistry 50: 9114–9124. doi: 10.1021/bi201260h
![]() |
[12] |
Snow ET, Xu LS, Kinney PL (1993) Effects of nickel ions on polymerase activity and fidelity during DNA replication in vitro. Chem Biol Interact 88: 155–173. doi: 10.1016/0009-2797(93)90089-H
![]() |
[13] |
Vaisman A, Ling H, Woodgate R, et al. (2005) Fidelity of Dpo4: effect of metal ions, nucleotide selection and pyrophosphorolysis. EMBO J 24: 2957–2967. doi: 10.1038/sj.emboj.7600786
![]() |
[14] |
Pelletier H, Sawaya MR, Wolfle W, et al. (1996) A structural basis for metal ion mutagenicity and nucleotide selectivity in human DNA polymerase beta. Biochemistry 35: 12762–12777. doi: 10.1021/bi9529566
![]() |
[15] |
Irimia A, Loukachevitch LV, Eoff RL, et al. (2010) Metal-ion dependence of the active-site conformation of the translesion DNA polymerase Dpo4 from Sulfolobus sulfataricus. Acta Crystallogr Sect F Struct Biol Commun 66: 1013–1018. doi: 10.1107/S1744309110029374
![]() |
[16] | Sirover MA, Dube DK, Loeb LA (1979) On the fidelity of DNA replication. Metal activation of Escherichia coli DNA polymerase I. J Biol Chem 254: 107–111. |
[17] | Sirover MA, Loeb LA (1977) On the fidelity of DNA replication. Effect of metal activators during synthesis with avian myeloblastosis virus DNA polymerase. J Biol Chem 252: 3605–3610. |
[18] |
Miyaki M, Murata I, Osabe M, et al. (1977) Effect of metal cations on misincorporation by E. coli DNA polymerases. Biochem Biophys Res Commun 77: 854–860. doi: 10.1016/S0006-291X(77)80056-9
![]() |
[19] | Goodman MF, Keener S, Guidotti S, et al. (1983) On the enzymatic basis for mutagenesis by manganese. J Biol Chem 258: 3469–3475. |
[20] |
Sirover MA, Loeb LA (1976) Infidelity of DNA synthesis in vitro: screening for potential metal mutagens or carcinogens. Science 194: 1434–1436. doi: 10.1126/science.1006310
![]() |
[21] |
Lee HR, Wang M, Konigsberg W (2009) The reopening rate of the fingers domain is a determinant of base selectivity for RB69 DNA polymerase. Biochemistry 48: 2087–2098. doi: 10.1021/bi8016284
![]() |
[22] |
Villani G, Tanquy Le Gac N, Wasungu L, et al. (2002) Effect of manganese on in vitro replication of damaged DNA catalyzed by the herpes simplex virus type-1 DNA polymerase. Nucleic Acids Res 30: 3323–3332. doi: 10.1093/nar/gkf463
![]() |
[23] |
Xia S, Konigsberg WH (2014) RB69 DNA polymerase structure, kinetics, and fidelity. Biochemistry 53: 2752–2767. doi: 10.1021/bi4014215
![]() |
[24] |
Beard WA, Wilson SH (2014) Structure and mechanism of DNA polymerase β. Biochemistry 53: 2768–2780. doi: 10.1021/bi500139h
![]() |
[25] |
Sobol RW, Wilson SH (2001) Mammalian DNA beta-polymerase in base excision repair of alkylation damage. Prog Nucleic Acid Res Mol Biol 68: 57–74. doi: 10.1016/S0079-6603(01)68090-5
![]() |
[26] |
Beard WA, Wilson SH (2000) Structural design of a eukaryotic DNA repair polymerase: DNA polymerase beta. Mutat Res 460: 231–244. doi: 10.1016/S0921-8777(00)00029-X
![]() |
[27] |
Yang L, Arora K, Beard WA, et al. (2004) Critical role of magnesium ions in DNA polymerase beta's closing and active site assembly. J Am Chem Soc 126: 8441–8453. doi: 10.1021/ja049412o
![]() |
[28] |
Bakhtina M, Lee S, Wang Y, et al. (2005) Use of viscogens, dNTPalphaS, and rhodium(III) as probes in stopped-flow experiments to obtain new evidence for the mechanism of catalysis by DNA polymerase beta. Biochemistry 44: 5177–5187. doi: 10.1021/bi047664w
![]() |
[29] |
Nakamura T, Zhao Y, Yamagata Y, et al. (2012) Watching DNA polymerase eta make a phosphodiester bond. Nature 487: 196–201. doi: 10.1038/nature11181
![]() |
[30] |
Gao Y, Yang W (2016) Capture of a third Mg2+ is essential for catalyzing DNA synthesis. Science 352:1334–1337. doi: 10.1126/science.aad9633
![]() |
[31] |
Freudenthal BD, Beard WA, Shock DD, et al. (2013) Observing a DNA polymerase choose right from wrong. Cell 154: 157–168. doi: 10.1016/j.cell.2013.05.048
![]() |
[32] |
Xia S, Wang J, Konigsberg WH (2013) DNA mismatch synthesis complexes provide insights into base selectivity of a B family DNA polymerase. J Am Chem Soc 135: 193–202. doi: 10.1021/ja3079048
![]() |
[33] | Doublie S, Tabor S, Long AM, et al. (1998) Crystal structure of a bacteriophage T7 DNA replication complex at 2.2 A resolution. Nature 391: 251–258. |
[34] |
Johnson SJ, Taylor JS, Beese LS (2003) Processive DNA synthesis observed in a polymerase crystal suggests a mechanism for the prevention of frameshift mutations. Proc Natl Acad Sci USA 100: 3895–3900. doi: 10.1073/pnas.0630532100
![]() |
[35] |
Johnson SJ, Beese LS (2004) Structures of mismatch replication errors observed in a DNA polymerase. Cell 116: 803–816. doi: 10.1016/S0092-8674(04)00252-1
![]() |
[36] |
Vashishtha AK, Konigsberg WH (2016) Effect of Different Divalent Cations on the Kinetics and Fidelity of RB69 DNA Polymerase. Biochemistry 55: 2661–2670. doi: 10.1021/acs.biochem.5b01350
![]() |
[37] |
Tabor S, Richardson CC (1989) Effect of manganese ions on the incorporation of dideoxynucleotides by bacteriophage T7 DNA polymerase and Escherichia coli DNA polymerase I. Proc Natl Acad Sci USA 86: 4076–4080. doi: 10.1073/pnas.86.11.4076
![]() |
[38] | Hori K, Mark DF, Richardson CC (1979) Deoxyribonucleic acid polymerase of bacteriophage T7. Characterization of the exonuclease activities of the gene 5 protein and the reconstituted polymerase. J Biol Chem 254: 11598–11604. |
[39] |
Kuchta RD, Mizrahi V, Benkovic PA, et al. (1987) Kinetic mechanism of DNA polymerase I (Klenow). Biochemistry 26: 8410–8417. doi: 10.1021/bi00399a057
![]() |
[40] | Venkitaraman AR (1989) Use of modified T7 DNA polymerase (sequenase version 2.0) for oligonucleotide site-directed mutagenesis. Nucleic Acids Res 17: 3314. |
[41] | Joyce CM (1989) How DNA travels between the separate polymerase and 3'-5'-exonuclease sites of DNA polymerase I (Klenow fragment). J Biol Chem 264: 10858–10866. |
[42] |
Irimia A, Zang H, Loukachevitch LV, et al. (2006) Calcium is a cofactor of polymerization but inhibits pyrophosphorolysis by the Sulfolobus solfataricus DNA polymerase Dpo4. Biochemistry 45: 5949–5956. doi: 10.1021/bi052511+
![]() |
[43] |
Sirover MA, Loeb LA (1976) Metal-induced infidelity during DNA synthesis. Proc Natl Acad Sci USA 73: 2331–2335. doi: 10.1073/pnas.73.7.2331
![]() |
[44] |
Sirover MA, Loeb LA (1976) Metal activation of DNA synthesis. Biochem Biophys Res Commun 70: 812–817. doi: 10.1016/0006-291X(76)90664-1
![]() |
[45] | Seal G, Shearman CW, Loeb LA (1979) On the fidelity of DNA replication. Studies with human placenta DNA polymerases. J Biol Chem 254: 5229–5237. |
[46] |
Vashishtha AK, Konigsberg WH (2018) The effect of different divalent cations on the kinetics and fidelity of Bacillus stearothermophilus DNA polymerase. AIMS Biophys 5: 125–143. doi: 10.3934/biophy.2018.2.125
![]() |
[47] |
Zhang H, Cao W, Zakharova E, et al. (2007) Fluorescence of 2-aminopurine reveals rapid conformational changes in the RB69 DNA polymerase-primer/template complexes upon binding incorporati of matched deoxynucleoside triphosphates. Nucleic Acids Res 35: 6052–6062. doi: 10.1093/nar/gkm587
![]() |
[48] |
Hariharan C, Bloom LB, Helquist SA, et al. (2006) Dynamics of nucleotide incorporation: snapshots revealed by 2-aminopurine fluorescence studies. Biochemistry 45: 2836–2844. doi: 10.1021/bi051644s
![]() |
[49] |
Frey MW, Sowers LC, Millar DP, et al. (1995) The nucleotide analog 2-aminopurine as a spectroscopic probe of nucleotide incorporation by the Klenow fragment of Escherichia coli polymerase I and bacteriophage T4 DNA polymerase. Biochemistry 34: 9185–9192. doi: 10.1021/bi00028a031
![]() |
[50] | Chin YE, Snow ET, Cohen MD, et al. (1994) The effect of divalent nickel (Ni2+) on in vitro DNA replication by DNA polymerase alpha. Cancer Res 54: 2337–2341. |
[51] |
Bock CW, Katz AK, Markham GD, et al. (1999) Manganese as a replacement for magnesium and zinc: Functional comparison of the divalent ions. J Am Chem Soc 121: 7360–7372. doi: 10.1021/ja9906960
![]() |
[52] |
Dube DK, Loeb LA (1975) Manganese as a mutagenic agent during in vitro DNA synthesis. Biochem Biophys Res Commun 67: 1041–1046. doi: 10.1016/0006-291X(75)90779-2
![]() |
[53] |
Jin YH, Clark AB, Slebos RJC, et al. (2003) Cadmium is a mutagen that acts by inhibiting mismatch repair. Nat Genet 34: 326–329. doi: 10.1038/ng1172
![]() |
[54] |
Hays H, Berdis AJ (2002) Manganese substantially alters the dynamics of translesion DNA synthesis. Biochemistry 41: 4771–4778. doi: 10.1021/bi0120648
![]() |
[55] |
Beckman RA, Mildvan AS, Loeb LA (1985) On the fidelity of DNA replication: manganese mutagenesis in vitro. Biochemistry 24: 5810–5817. doi: 10.1021/bi00342a019
![]() |
[56] |
Doetsch PW, Chan GL, Haseltine WA (1985) T4 DNA polymerase (3'–5') exonuclease, an enzyme for the detection and quantitation of stable DNA lesions: the ultraviolet light example. Nucleic Acids Res 13: 3285–3304. doi: 10.1093/nar/13.9.3285
![]() |
[57] | Kunkel TA, Soni A (1988) Exonucleolytic proofreading enhances the fidelity of DNA synthesis by chick embryo DNA polymerase-gamma. J Biol Chem 263: 4450–4459. |
[58] |
Wang W, Hellinga HW, Beese LS (2011) Structural evidence for the rare tautomer hypothesis of spontaneous mutagenesis. Proc Natl Acad Sci USA 108: 17644–17648. doi: 10.1073/pnas.1114496108
![]() |
[59] |
Bebenek K, Pedersen LC, Kunkel TA (2011) Replication infidelity via a mismatch with Watson-Crick geometry. Proc Natl Acad Sci USA 108: 1862–1867. doi: 10.1073/pnas.1012825108
![]() |
[60] |
Harris VH, Smith CL, Cummins WJ, et al. (2003) The effect of tautomeric constant on the specificity of nucleotide incorporation during DNA replication: support for the rare tautomer hypothesis of substitution mutagenesis. J Mol Biol 326: 1389–1401. doi: 10.1016/S0022-2836(03)00051-2
![]() |
[61] |
Topal MD, Fresco JR (1976) Complementary base pairing and the origin of substitution mutations. Nature 263: 285–289. doi: 10.1038/263285a0
![]() |
[62] |
Vashishtha AK, Wang J, Konigsberg WH (2016) Different Divalent Cations Alter the Kinetics and Fidelity of DNA Polymerases. J Biol Chem 291: 20869–20875. doi: 10.1074/jbc.R116.742494
![]() |
[63] |
Xia S, Vashishtha A, Bulkley D, et al. (2012) Contribution of partial charge interactions and base stacking to the efficiency of primer extension at and beyond abasic sites in DNA. Biochemistry 51: 4922–4931. doi: 10.1021/bi300296q
![]() |
[64] |
Zhang Y, Baranovskiy AG, Tahirov ET (2016) Divalent ions attenuate DNA synthesis by human DNA polymerase alpha by changing the structure of the template/primer or by perturbing the polymerase reaction. DNA Repair 43:24–33. doi: 10.1016/j.dnarep.2016.05.017
![]() |
[65] |
Schmitt MW, Kennedy SR, Salk JJ, et al. (2012) Detection of ultra-rare mutations by next-generation sequencing. Proc Natl Acad Sci USA 109: 14508–14513. doi: 10.1073/pnas.1208715109
![]() |
[66] | Frederico LA, Kunkel TA, Shaw BR (1990) A sensitive genetic assay for the detection of cytosine deamination: determination of rate constants and the activation energy. Biochemistry 29: 2532–2537. |
[67] | Bebenek K, Kunkel TA (1995) Analyzing fidelity of DNA polymerase. Methods Enz 262 217–232. |
[68] |
Yasukawa K, Iida K, Okano H, et al. (2017) Next-generation sequencing-based analysis of reverse transcriptase fidelity. Biochem Biophys Res Commun 492: 147–153. doi: 10.1016/j.bbrc.2017.07.169
![]() |
[69] |
Ellefson JW, Gollihar J, Shroff R, et al. (2016) Synthetic evolutionary origin of a proofreading reverse transcriptase. Science 352: 1590–1593. doi: 10.1126/science.aaf5409
![]() |
1. | Mengyao Zhang, Jing Ye, Jin-song He, Fang Zhang, Jianfeng Ping, Cheng Qian, Jian Wu, Visual detection for nucleic acid-based techniques as potential on-site detection methods. A review, 2020, 1099, 00032670, 1, 10.1016/j.aca.2019.11.056 | |
2. | Carolina Diaz Arenas, Aleksandra Ardaševa, Jonathan Miller, Alexander S. Mikheyev, Yohei Yokobayashi, Ribozyme Mutagenic Evolution: Mechanisms of Survival, 2021, 51, 0169-6149, 321, 10.1007/s11084-021-09617-0 | |
3. | Said Laatri, Soufiane El Khayari, Zidane Qriouet, Exploring the molecular aspect and updating evolutionary approaches to the DNA polymerase enzymes for biotechnological needs: A comprehensive review, 2024, 276, 01418130, 133924, 10.1016/j.ijbiomac.2024.133924 |
DNA Polymerase (Pol Family) | Metal Ion | ||||||||||
Mn2+ | Co2+ | Fe2+ | Ni2+ | Zn2+ | Cd2+ | Sr2+ | Ba2+ | Cu2+ | Cr3+ | Ca2+ | |
DNA pol I (A) | + | + | - | + | - | - | - | - | - | - | - |
Human pol α (B) | + | + | - | - | + |
- | - | - | - | - | - |
pol β (X) | + | + |
- | - | + | + | - | - | - | - | - |
AMV DNA pol (RT) | + | + | - | + | - | - | - | - | - | - | - |
T4 pol (B) | + | + | - | + | - | - | - | - | - | - | - |
T7 pol (A) | + | + | - | + | - | - | - | - | - | - | - |
RB69pol (B) | + | + | - | + | - | - | - | - | - | - | - |
BST pol (A) | + | + |
- | + |
- |
+ |
- | - | - | - |
- |
Dpo4 pol (Y) | + | + |
- | - | - | - | - | - | - | - | + |
a Metal ion which activate the respective DNA polymerase are shown with + and those that are not able to support the polymerase activity are shown with -. b Studies by Pelletier et al. and Egli et al. previously claimed that human pol β and Dpo4 were not able to utilize Co2+ as cofactor however, recent studies by Vashishtha et al. showed that both pol β and Dpo4 can catalyze primer extension in the presence of Co2+ as explained in the text. c Vashishtha et al. [46] d Zhang et al. [64] |
Parameter | Metal Ion | ||||||
Mg2+ | Mn2+ | Co+ | Ni2+ | Zn2+ | Cd2+ | Ca2+ | |
Ionic radius (Å) | 0.86 | 0.81 | 0.89 | 0.83 | 0.88 | 0.95 | 1.1 |
Coordination | Oct | Oct | Oct | Oct | Oct | Oct | Oct |
Geometry | Td | Td | Td | Td | Td |
Td | PBP |
Sq | Sq | Sq | TBP | TBP | HBP | ||
TBP | TBP | ||||||
pKa of the water molecule | 11.4 | 11.5 | 10.0 | 10.6 | 7.0 | 9.0 | 12.8 |
a Td represents Tetrahedral, Sq represents Square planar, TBP represents Trigonal bipyramidal, Oct represents Octahedral, PBP represents pentagonal bipyramid, and HBP represents hexagonal bipyramidal. b Even though Zn2+ can form Octahedral complexes, majority of Zn2+ complexes are tetrahedral. |
DNA Polymerase (Pol Family) | Metal Ion | ||||||||||
Mn2+ | Co2+ | Fe2+ | Ni2+ | Zn2+ | Cd2+ | Sr2+ | Ba2+ | Cu2+ | Cr3+ | Ca2+ | |
DNA pol I (A) | + | + | - | + | - | - | - | - | - | - | - |
Human pol α (B) | + | + | - | - | + |
- | - | - | - | - | - |
pol β (X) | + | + |
- | - | + | + | - | - | - | - | - |
AMV DNA pol (RT) | + | + | - | + | - | - | - | - | - | - | - |
T4 pol (B) | + | + | - | + | - | - | - | - | - | - | - |
T7 pol (A) | + | + | - | + | - | - | - | - | - | - | - |
RB69pol (B) | + | + | - | + | - | - | - | - | - | - | - |
BST pol (A) | + | + |
- | + |
- |
+ |
- | - | - | - |
- |
Dpo4 pol (Y) | + | + |
- | - | - | - | - | - | - | - | + |
a Metal ion which activate the respective DNA polymerase are shown with + and those that are not able to support the polymerase activity are shown with -. b Studies by Pelletier et al. and Egli et al. previously claimed that human pol β and Dpo4 were not able to utilize Co2+ as cofactor however, recent studies by Vashishtha et al. showed that both pol β and Dpo4 can catalyze primer extension in the presence of Co2+ as explained in the text. c Vashishtha et al. [46] d Zhang et al. [64] |
Parameter | Metal Ion | ||||||
Mg2+ | Mn2+ | Co+ | Ni2+ | Zn2+ | Cd2+ | Ca2+ | |
Ionic radius (Å) | 0.86 | 0.81 | 0.89 | 0.83 | 0.88 | 0.95 | 1.1 |
Coordination | Oct | Oct | Oct | Oct | Oct | Oct | Oct |
Geometry | Td | Td | Td | Td | Td |
Td | PBP |
Sq | Sq | Sq | TBP | TBP | HBP | ||
TBP | TBP | ||||||
pKa of the water molecule | 11.4 | 11.5 | 10.0 | 10.6 | 7.0 | 9.0 | 12.8 |
a Td represents Tetrahedral, Sq represents Square planar, TBP represents Trigonal bipyramidal, Oct represents Octahedral, PBP represents pentagonal bipyramid, and HBP represents hexagonal bipyramidal. b Even though Zn2+ can form Octahedral complexes, majority of Zn2+ complexes are tetrahedral. |