The Catalytic Mechanism of DNA and RNA Polymerases

1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Sunderland

Perspective

The catalytic mechanism of DNA and RNA polymerases Vito Genna, Elisa Donati, and Marco De Vivo ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.8b03363 • Publication Date (Web): 18 Oct 2018 Downloaded from http://pubs.acs.org on October 19, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

The Catalytic Mechanism of DNA and RNA Polymerases Vito Genna,1 Elisa Donati1 and Marco De Vivo1

1. Laboratory of Molecular Modeling and Drug Discovery, Istituto Italiano di Tecnologia, Via Morego 30, 16163, Genoa, Italy

Corresponding author: Dr. Marco De Vivo Email: [email protected] Phone: +39 010 71781577

ACS Paragon Plus Environment

1

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 2 of 57

2

Page 3 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Abstract. DNA and RNA polymerases (Pols) catalyze nucleic acid biosynthesis in all domains of life, with implications for human diseases and health. Pols carry out nucleic acid extension through the addition of one incoming nucleotide trisphosphate at the 3’-OH terminus of the growing primer strand, at every catalytic cycle. Thus, Pol catalysis involves chemical reactions for nucleophile 3’OH deprotonation and nucleotide addition, as well as major protein conformational motions and structural rearrangements for nucleotide selection, binding, and nucleic acid translocation to complete the overall catalytic cycle. In this respect, quantum and molecular mechanics simulations, integrated with experimental data, have advanced our mechanistic understanding of how Pols operate at the atomic level. This Perspective outlines how modern molecular simulations can further deepen our understanding of Pol catalytic reactions and fidelity, which may help in devising strategies for designing drugs and artificial Pols for biotechnological and clinical purposes.

Keywords: DNA, RNA, Polymerases, Catalysis, QM/MM, Molecular Mechanics, Computations, Modeling

ACS Paragon Plus Environment

3

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 57

Introduction DNA and RNA polymerases (Pols) are enzymes that play a major role in crucial cellular processes, such as gene expression, regulation, transcription, and repair.1-3 As such, Pols are critical pharmacological targets for treating a wide range of diseases including viral and bacterial infection, neurodegenerative diseases, and cancer.4-6 In recent decades, this has prompted extensive efforts to better understand their mechanism of action.7-10 Indeed, an improved understanding of Pol catalysis has led to new drug discovery strategies for modulating Pol function, and to the engineering of innovative artificial Pols for biotechnological and therapeutic purposes.11-14 Pols elongate the nascent nucleic acid strand via a complex stepwise catalytic process with both chemical and physical steps. First, during each catalytic cycle, Pols extend the growing primer

strand

by

adding

one

incoming

nucleotide

at

the

3′

primer

terminus

(deoxyribonucleoside triphosphate, dNTP, for DNA Pols; ribonucleoside triphosphate, rNTP, for RNA Pols). This occurs via a well-characterized SN2-like phosphoryl-transfer reaction, which is assisted by the conserved two-metal-ion mechanism (typically involving Mg or Mn), as demonstrated by a wealth of structural data on ternary Pol/(R)DNA/(d)NTP complexes.15-17 Thus, Pol catalysis comprises chemical reactions for primer 3’-OH deprotonation (i.e. nucleophile activation) and nucleotide addition, which allow the nascent strand to be elongated, one incoming nucleotide at a time (Fig. 1). As part of the same catalytic cycle, Pols undergo functional large-amplitude protein conformational motions18-20 and structural rearrangements for nucleotide selection,21,

22

binding,23-25 and nucleic acid translocation.26-29 Together, these

chemical and physical steps constitute the overall catalytic cycle for processive nucleic acid biosynthesis catalyzed by Pols (Fig. 1). In this context, the rapidly growing body of Pol experimental data has revealed many key aspects of Pol catalysis.30 In particular, time-resolved X-ray crystallography has enabled researchers to capture and observe the chemical reaction for nucleotide addition, moving from the reactants to the products.31-33 One such example is the recent in crystallo catalysis observed

ACS Paragon Plus Environment

4

Page 5 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

for human DNA Pol- a eukaryotic Pol involved in DNA repair and recombination34, 35 (Fig. 2). These structural data have generated important atomistic insights into (d)NTP incorporation, which occurs via the nucleophilic attack of the primer terminal 3′-OH to the α-phosphorus of the incoming dNTP.34 That is, the phosphoryl-transfer reaction for nucleotide addition is catalyzed via the established two-metal-ion mechanism, where two Mg ions facilitate nucleophile formation, stabilize the transition state, and help the leaving group departure.36-39 Indeed, passing from the reactants to the products, Pol-’ crystals clearly show the typical inversion of the umbrella of the scissile phosphate during the phosphoryl-transfer reaction, as expected for a canonical SN2-like reaction.38 This leads to the formation of the pyrophosphate (PPi) product and the consequent conclusion of the catalytic cycle for nucleotide addition (Fig. 1).34,

40

However, the transient nature of the enzymatic transition state poses technical

limitations in determining the exact spatiotemporal location and structural features of this highenergy state. Moreover, there is debate over the mechanism and possible pathways of protontransfer events during catalysis, such as the one needed to deprotonate and activate the nucleophilic -3′O atom of the primer.41-45 In addition, recent ternary Pol/D(R)NA/d(r)NTP complexes have revealed the presence of an extra metal ion transiently bound to the catalytic site (Table 1).46, 47 This third cation has been observed in a few Pols structures, including in some Pol- It seems to be an additional key player for Pol catalysis, although its exact role is unclear. A few recent and comprehensive reviews have mostly focused on the structural and kinetics data of Pol enzymes.48-50 As mentioned above, in addition to the chemical steps for nucleotide incorporation, Pol function requires key physical steps for the nucleotide selection, binding, and nucleic acid translocation to complete the catalytic turnover. Importantly, these events also regulate Pol fidelity, which is the ability of Pols to faithfully replicate the template strand during catalysis. Pol fidelity requires the contribution of protein dynamical features for correct nucleotide recognition and binding.51-53 This occurs prior to nucleotide incorporation into the primer strand, with subsequent formation of the canonical Watson-Crick base pair.54 Yet the underlying molecular mechanism for Pol fidelity remains uncertain. Furthermore, for highly processive ACS Paragon Plus Environment

5

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 57

DNA/RNA synthesis defined by recurrent catalytic cycles for nucleotide addition, the Pols must efficiently translocate along the nucleic acid strand by one base pair. This catalytic step again requires large conformational rearrangements of the Pol-substrate complex. In this regard, the abundance of structural and kinetics data on Pol-, a small eukaryotic Pol for DNA repair,55 has greatly helped in understanding how Pol fidelity is enabled by major conformational changes during the overall catalytic cycle.56-58 In addition, nucleic acid translocation also requires the efficiency and coordination of large-amplitude motions of structural regions of the ternary Pol/D(R)NA/(d)NTP complex. This relates to Pol’s processivity, i.e. the number of incoming nucleotides incorporated along the growing primer strand during a single template-binding episode, before the Pol enzyme dissociates from the nucleic acid strand.15 Insights into these chemical and physical steps for Pol-assisted nucleic acid biosynthesis are critical to better understanding Pol catalysis. In this respect, quantum and molecular mechanics simulations have helped advance our atomistic understanding of Pol function.59-63 This Perspective analyzes the existing mechanistic hypotheses concerning Pol catalysis and fidelity in light of recent structural data and computational studies on Pol function. We conclude by examining artificial Pol enzymes, which have the potential to impact positively on human health. In sum, we outline how state-of-the-art multiscale molecular dynamics simulations, combined with experimental data, will help to deepen our understanding of nucleic acid synthesis performed by DNA/RNA Pols.

An intramolecular H-bond characterizes the typical geometry of the incoming nucleotide. To be added to the growing nucleic acid strand, the incoming nucleotide must enter into and bind to the two-metal-aided catalytic site of Pols (Fig. 1). Once bound to the (R)DNA-Pol binary complex, the conformation adopted by the nucleotide allows the SN2-like phosphoryl-transfer reaction for nucleotide addition, according to the two-metal-ion mechanism.36, 38 However, the preferred molecular conformation of nucleotides in water is radically different from the conformation adopted by nucleotides in ternary Pol/D(R)NA/(d)NTP complexes. Indeed, when freely dispersed in solution, the (d)NTP has been reported to prefer extended and relaxed ACS Paragon Plus Environment

6

Page 7 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

conformations, in which the triphosphate is bound to one divalent metal ion.64 This has been demonstrated by computational studies that used quantum mechanics/molecular mechanics (QM/MM) simulations to elucidate the mechanisms of ATP hydrolysis in water.65, 66 However, when the nucleotide is bound to the binary (R)DNA-Pol complex, it always shows a much more compacted conformation. This conformation is characterized by an intramolecular and stereospecific H-bond formed by the nucleophilic 3′-OH with the pro-S oxygen atom of the phosphate of the incoming nucleotide (Fig. 2). Notably, we recently demonstrated that this conformation is consistently preserved in all the nucleotides co-crystallized in ternary Pol/(R)DNA/(d)NTP complexes from all domains of life (Fig. 3A).41 In contrast to linear conformations observed in water, this typical (d)NTP architecture allows the incoming nucleotide to form base stacking with the preceding nucleobase along the nascent primer strand, while also forming base pairing with the templating nucleotide. Concomitantly, this conformation allows the proper placement of the (d)NTP -phosphate on top of the two metal ions within the catalytic site. In this way, the alkoxide nucleophilic species can now attack Pα thanks to the collinearity between the 3’O- group and the oxygen atom bridging the Pα and Pβ atoms, as expected for the canonical SN2-type reaction scheme.67-73 Therefore, it seems that a prerequisite for catalysis is the conformational change of the incoming nucleotide, which must move from an extended conformation in water to a much more bent geometry when in the catalytic pocket. The intramolecular H-bond helps in forming and maintaining this conformation. In addition, we have noted that this short H-bond co-exists with the reactive C3′-endo conformation of the sugar pucker adopted in the Michaelis-Menten complex by the incoming nucleobase (Fig. 3B).41 We have used these structural observations to propose a self-activated mechanism (SAM) for nucleic acid elongation catalyzed by Pols (vide infra). More recently, Wu et al. commented on the relevance and possible role of this conserved intramolecular H-bond.50 In particular, they observed that 2ʹ,3ʹ-dideoxy-NTP (ddNTP), which is usually a chain terminator in DNA sequencing, can be incorporated into DNA by polymerases even if it lacks hydroxyl groups on the sugar ring. That is, ddNTP cannot form such an intramolecular H-bond. This could explain ACS Paragon Plus Environment

7

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 57

the observed >100-fold reduction of the catalytic efficiency for ddNTP incorporation by Pols.50 Together, these indications suggest that this structurally conserved H-bond established between the nucleophilic 3′-OH and the -phosphate of the incoming nucleotide may indeed be functional for catalytic nucleotide addition. Intuitively, this conformational change of the nucleobase with intramolecular H-bond formation likely occurs during nucleotide recognition and binding. However, the energetics, dynamics, and spatiotemporal occurrence of this conformational rearrangement must still be elucidated.

Nucleophile formation and nucleotide addition for nucleic acid elongation. Pol-mediated nucleic acid elongation requires two chemical steps, i.e. nucleophile formation and subsequent nucleophilic attack to incorporate the incoming nucleotide into the growing primer strand, with consequent formation of the pyrophosphate (PPi) leaving group.38,

72

Over the years,

computational enzymology has generated and analyzed a few proposals for the mechanism used by Pols to perform nucleotide incorporation at the 3′ primer terminus, in a 5’→3’ direction. There are already comprehensive reviews and numerous experimental studies of the kinetics and structural data related to Pol catalysis.30,

50, 74, 75

Here, we focus on how computational

methods have been used to investigate the enzymatic mechanisms of Pol-mediated nucleic acid biosynthesis. A first mechanistic hypothesis was proposed by Warshel and coworkers, who investigated the catalytic mechanism for nucleotide incorporation by the DNA Pol of bacteriophage T7.44 In this mechanistic study, the nucleophilic oxygen is preferentially activated by a nearby aspartate (Asp654), when compared to mechanisms for nucleophile activation involving, as proton acceptor, either bulk water or one of the oxygens of the -phosphate of the incoming nucleobase (Fig. 4). Thus, the favored mechanism was the one where the ionized Asp654 acts as a general base, receiving the proton from the attacking 3′-OH of the primer strand. An analogous deprotonation mechanism was also proposed for DNA Pol-, with QM/MM calculations used to propose that an aspartate residue (Asp256) is the general base for activating the terminal primer deoxyribose -O3′.76 The same mechanism was also reported in ACS Paragon Plus Environment

8

Page 9 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

DNA Pol-λ, where computation supports that an aspartate (Asp490) acts as proton acceptor for nucleophile activation.77 A final example is a more recent computational study of the catalytic mechanism of the DNA Pol-I. This study proposed that a histidine residue (His829), instead of an aspartate, may accept the proton of the attacking 3′-OH, and thus activate the nucleophile.78 The common mechanistic feature of these proposals for nucleophile activation is the presence of a first coordination-shell residue that can act as a general base (i.e. proton acceptor) for the deprotonation of the attacking oxygen.78 Following this first protein-mediated chemical step for nucleophile activation, the reaction proceeds according to the two-metal-ion mechanism for nucleophilic attack on the -phosphate of the dNTP substrate, and departure of the PPi leaving group. An alternative mechanism for nucleophile formation and nucleotide addition was later proposed by Zhang and coworkers, who hypothesized that deprotonation of the attacking 3′OH of the primer strand may occur via a nearby water, which can act as a proton acceptor (Fig. 4).43 Eventually, this water shuttles the proton on the leaving PPi group. This is referred to as a water-mediated-substrate-assisted (WMSA) mechanism. It was proposed and examined via QM/MM simulations of the lesion-bypass DNA polymerase IV (Dpo4) catalysis for a nucleotidyl transfer reaction. The same research group also described the WMSA mechanism for the T7 DNA Pol42 and DNA Pol- enzymes.79 Later, Wang et al. reported and computationally examined a similar water-mediated mechanism for nucleophile formation in DNA polymerase IV80 (Dpo4) catalysis, during which two water molecules facilitate the initial deprotonation of the primer 3′OH, then the shuttling of the migrating proton to the leaving group. In the case of RNA polymerase II (RNA Pol-II), one hydroxide ion from the bulk was found to be the favored proton acceptor for nucleophile activation.81 Other possible pathways to shuttle the migrating proton to the PPi group were also proposed.81, 82 The PPi is then released via motions of a specific loop,83 as depicted via a Markov state model (MSM) constructed from extensive all-atom molecular dynamics (MD) simulations.84, 85 A final mechanistic scheme for Pol catalysis, which differs from the above hypotheses, is our recent proposal for a self-activated mechanism (SAM).41 It is ‘self-activated’ because the ACS Paragon Plus Environment

9

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 57

mechanism is initiated by the intramolecular H-bond found in the incoming nucleotide (as described above, Figs. 3 and 5). In SAM, this intramolecular H-bond prompts the in situ deprotonation of the 3′-OH in the incoming nucleotide. Our Car-Parrinello (CP) QM/MM simulations have shown that the proton of the 3′-OH of the entering nucleotide can be directly transferred onto the pro-S oxygen of the -phosphate as the nucleotide addition occurs. That is, the proton transfer occurs simultaneously with the detachment of the -phosphate from the -phosphate of the incoming nucleobase, leading to the formation and departure of the leaving (and protonated) PPi group (Fig. 5). Notably, in SAM, the newly formed 3’-hydroxide ion in the incoming nucleotide is therefore formed and located on top of MgA during nucleic acid translocation. In this way, the deprotonated 3’-hydroxide is ready for a new catalytic cycle for nucleotide addition. This mechanism for the deprotonation of the 3′-OH of the incoming nucleobase during nucleotide addition is one difference between SAM and the previously proposed mechanisms for Pol catalysis. However, in SAM, the nucleophile is formed as nucleotide addition occurs thanks to the activated 3’-hydroxide ion of the attacking group at the primer strand (formed via the same intramolecular proton transfer, during the previous catalytic cycle). Importantly, this key element of SAM allows the synergistic interconnection of a concerted closed-loop catalytic sequence of events that include the chemical steps for nucleophile deprotonation and nucleotide addition (as in the previous mechanistic proposals), and also the physical step for nucleic acid translocation (not considered in the previous mechanistic proposals). In this regard, at least in our case, ab initio QM/MM simulations (as opposed to static calculations) were instrumental in characterizing this concerted structural rearrangement associated to nucleophile activation and leaving group formation that occur simultaneously with, notably, partial nucleic acid translocation for Pol catalysis. Thus, the coupling between the chemical and physical steps, which form a closed-loop catalytic cycle for Pol catalysis, represents the novel conceptual aspect in SAM. However, the level of synchronicity and synergy of these concerted chemical and physical steps in SAM remains to be elucidated. As an aside, it is interesting that such concerted metal-aided mechanisms for proton transfer in biological ACS Paragon Plus Environment

10

Page 11 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

systems are somehow reminiscent of concepts used for homogeneous catalysis, with computationally characterized reaction mechanisms such as the “Concerted MetalationDeprotonation” (CMD) and “Ambiphilic Metal−Ligand Assistance” (AMLA) for carboxylateassisted C−H activation in organometallic systems.86-88

Different enzymatic mechanisms for the chemical steps in Pol activity. There are a few mechanistic hypotheses for Pol catalysis (see paragraph above), which are all equally plausible in principle (Fig. 4 and 5).41, 43, 44, 89 One main difference is the way the nucleophile is activated. Conveniently, one could argue that the chemical step for nucleophile activation is unlikely to affect the overall energetics of nucleotide addition. If so, one could start investigating the enzymatic reaction from an already activated nucleophile that was somehow deprotonated.89 However, even if that were the case, nucleophile activation is undoubtedly needed for catalysis to be initiated and for it to proceed with the subsequent chemical and physical steps. It is therefore informative to investigate and understand how this key event for nucleophile deprotonation occurs at the atomic level. In all the possible mechanisms discussed above, MgA facilitates 3′-OH deprotonation by lowering its pKa, thus favoring the ionized 3′-O- form. Concomitantly, MgB facilitates leaving group formation.40 Once activated, the 3’-hydroxide ion can perform the subsequent nucleophile attack at the P of the incoming nucleotide, with a dissociative or associative transition state (TS, defined in this case as a metastable pentacovalent phosphorane species) for metal-aided nucleotidyl transfer reaction supported by MgA and MgB, which act together to stabilize the TS along the reaction path.36 In light of the extensive literature on enzymatic metal-aided phosphoryl transfer reactions (see e.g. Ref.68, 69, 90, 91), the precise geometry of the TS for phosphoryl transfer is expected to depend on the specific Pol enzyme under investigation. Any of the mechanisms described above for Pol catalysis may be valid for a subset of Pol enzymes. It would be a stretch to generalize and present a specific TS geometry (associative vs. dissociative) as a template for all Pol catalytic mechanisms. In all this, there is still debate over how to experimentally test the mechanistic details of Pol catalysis. The energetics of the reactions can be compared to the ACS Paragon Plus Environment

11

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 57

experimental kcat using transition state theory. This comparison is commonly used to indirectly validate the mechanistic hypothesis.19,

92-94

Analogously, the computed energetics of the

enzyme-catalyzed reaction should be compared and validated by calculating the energetic barrier for the same uncatalyzed reaction in water, which is expected to be much higher.95-99 However, it is extremely difficult to experimentally demonstrate the specific pathways for nucleophile activation and proton migration, which lead the proton of the nucleophilic 3′-OH group to the final acceptor (bulk water, a protein residue, or the PPi leaving group). In this regard, for example, solvent deuterium isotope effects indicated two proton-transfer events for the RNA-dependent RNA polymerase (RdRp) from poliovirus, although their exact migration path is challenging to determine experimentally.100 However, such experimental studies are to be welcomed, because these are critical in order to examine and possibly verify mechanistic proposals about the pathways for proton migration, detected by computational investigations.

A third transient metal ion for nucleic acid biosynthesis. A recent new player in the twometal-ion mechanism for nucleic acid processing is an additional third metal ion (MgC), which was recently resolved close to and above MgB in Pol- structures.34, 47 Notably, a third solventexposed cation was provocatively proposed a few years ago as an active part of the two-metalion mechanism in ribonuclease H (RNase-H), based on evidence from MD simulations.101 It was recently found in a few structures of the human exonuclease 1 (hExo-1)102 and D. mobilis homing endonuclease (I-DmoI).103 Second-shell basic amino acids and cations have also been experimentally identified and computationally examined at structurally conserved positions in a large set of two-metal-ion enzymes that process DNA and RNA (including several Pols, nucleases and also ribozymes, Fig. 6), as well as metallolyase enzymes.103-108 Taken together, these experimental and computational results further support the idea that additional structural elements, located in a larger orbit of the metal-centered structural architecture, may be critical for the proper function of the two-metal-ion mechanism for nucleic-acid processing.104 In more detail, the third ion found in Pol- is precisely located within the enzymatic active site, on the opposite side of the 3’-OH group, and it chelates oxygens of the - and ACS Paragon Plus Environment

12

Page 13 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

phosphates of the dNTP. We have collected the available Pols structures resolved in the presence of the third ion (see Tab. 1). In addition to Pol-, there are structures from Pol-  and Pol-  enzymes, including the engineered Pol Kod-RI,111 an artificial threose nucleic acid (TNA) polymerase.112 Notably, in Kod-RI111 (and in Pol-,113 the third metal ion has a different location than in native Pols. Intriguingly, however, it is still in direct contact with the entering nucleotide. In all these structures, the catalytic metal ions are most often Mg and Mn, although Ca and Na are sometimes found in these structures, according to the specific conditions used to obtain the crystals (see Tab. 1). The surprising finding of a third metal ion in Pols has opened a stimulating and unresolved debate about the specific role of this additional ion for Pol catalysis.46 While the third metal ion seems functional,46, 114 there are at least three hypotheses about its mechanistic role, which are not necessarily mutually exclusive. Indeed, examining the structures containing this third ion (such as the MgC captured in Pol- during in crystallo reactions via time-resolved X-ray crystallography),47 it is interesting that this ion is mostly present bound in structures resembling the TS for nucleotide addition, or just after the chemical reaction occurred.34 Indeed, the diffusion and protein binding of this third ion is expected to occur on a timescale longer than that of the chemical reaction for nucleotide addition. It is thus unclear if metal binding indicates that this ion participates actively to the chemical reaction, stabilizes the products and/or facilitates group leaving departure. That is, the observed position of this transient third metal ion, and the time it appears along the reaction path resolved in crystallo, suggests that the third ion: i) may be necessary for the reaction to occur, further stabilizing the negative charge of the scissile phosphate during the nucleotidyl transfer, and/or ii) favors products formation, inhibiting the backward reaction from the products to the reagents, and/or iii) acts as an exit shuttle for the leaving PPi departure. In principle, these three different active roles for MgC during catalysis could be complementary or, at least, not necessarily unconnected. It is therefore challenging to ascertain whether the third ion performs one or more of those actions during catalysis.

ACS Paragon Plus Environment

13

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 57

Yang and collaborators maintain that the third ion is essential for nucleotidyl transfer in Pols,47 as suggested by the fact that a low metal ion concentration in experiments for reaction in crystallo hampers catalysis, i.e. the product is not experimentally detected in crystallo if a low metal ion concentration is used.46 This experimental observation supports the active role of the third ion in lowering the energetic barrier for the enzymatic phosphoryl-transfer in Pols, thus accelerating dNTP hydrolysis. However, the fact that low metal ion concentration does not lead to product detection in crystallo does not necessarily exclude the contribution of MgC in stabilizing the products when the metal is at physiological concentrations. Likewise, it remains plausible that the third ion plays a role in facilitating leaving group departure by acting as an exit shuttle. Indeed, the third ion appears to prevent the reverse enzymatic reaction (i.e. pyrophosphorolysis) by stabilizing the products. This has been demonstrated by QM/MM computations from Wilson and coworkers115 on the role of the third ion in Pol-, and by our CP QM/MM simulations of Pol-.41 More recently, Yoon and Warshel used empirical valence-bond (EVB) calculations on Pol- structures116 to obtain computed energy profiles that confirm the role of the third ion in stabilizing the reaction products. This was further confirmed via QM/MM simulations by Stevens and Hammes-Schiffer.114 In addition to these studies, we have previously reported forcefield-based MD simulations that describe the role of the third ion in lowering the barrier for PPi release, acting as the exit shuttle.21 In this regard, the idea that a third ion is needed for Pol catalysis is challenged by recent in crystallo reactions and new structural data, together with primer extension assays reported by Kottur and Nair for DNA polymerase IV (PolIV).117 These results support the hypothesis that the PPi group is hydrolyzed after nucleotidyl transfer. This energetically favors DNA synthesis via the conventional two-metal-ion mechanism, at least in the case of Pol IV.117 Based on the available structural,47 kinetics,48 and computational evidence,114,

118

it is

therefore tempting to believe that the third ion, characterized by its transient nature, actively participates in the chemical step for nucleotide incorporation in Pols.34 In addition, a few recent computational studies from different groups suggest that the third ion stabilizes product formation, and acts as an exit shuttle to facilitate the departure of the leaving PPi outside of the ACS Paragon Plus Environment

14

Page 15 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

catalytic site.21, 28, 82, 119 Certainly, this puzzling and novel aspect of the common two-metalmechanism in Pols merits future experimental and computational studies to further elucidate whether and how a third metal ion is an active player in this complex and fascinating biological process for nucleic acid biosynthesis.

Insights into what regulates fidelity in Pols. The fidelity of Pol enzymes is expressed as the error rate per nucleotide that is added during template-strand replication.120,

121

While the

chemical steps of nucleophile deprotonation and nucleotide addition for Pol-mediated nucleotidyl transfer are rather spatially restricted to the region of the catalytic site, fidelity also relates to physical steps for nucleotide selection, recognition, and binding, which necessarily involve major protein conformational motions and structural rearrangements. For example, this has been experimentally determined for the Dpo4 and Pol- enzymes, as well as for several other Pols.48, 51, 122-124 Each Pol is able to recognize the correct nucleobase in solution, which is selectively recruited and inserted along the primer strand, complementing the template base.125-127 These structural changes must eventually involve nucleic acid translocation to allow one nucleotide to be added to the growing strand at each catalytic cycle.84, 128-130 In addition, fidelity and processivity (i.e. the number of incoming nucleotides added to the primer strand for each single template-binding event) can be modulated by the proofreading 3’→5’ exonuclease. If there is incorrect incorporation, specialized exonucleases can resolve mispairing by removing the incorrect nucleotide, which is cleaved from the end of the strand one at a time.131-133 This is also true for highly processive Pols, like Pol- and Pol- in eukaryotes, and Pol-I, Pol-II and Pol-III in prokaryotes.132, 134-136 Once this correction is performed, the nucleic acid chain can be transferred back into the Pol enzyme to restart the polymerization process. Overall, this intricate yet elegant sequence of chemical and physical events defines Pol fidelity and processivity. In terms of computational studies, much effort has been devoted to examining the structural features and energetics of dynamical transformations and reactions that may contribute to fidelity and processivity in Pols. For example, a few studies have analyzed how Pols are able to ACS Paragon Plus Environment

15

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 57

avoid the addition of an incorrect nucleotide, if accidentally accommodated within the active site. One example is the MD-based investigation of fidelity of the T7 DNA Pol, where incorrect nucleotide binding generated a higher energetic barrier.19, 137 However, it remains unclear how this energy increment for incorrect nucleotide binding affects the energy for the subsequent nucleotide incorporation. In principle, one can hypothesize that the formation of a poorly reactive ternary Pol/(R)DNA/(d)NTP complex increases the energy barrier for nucleotide incorporation and, consequently, the chance for the incorrect nucleotide to dissociate from the Pol catalysis site, thus allowing binding of the correct nucleoside triphosphate.138 Moreover, the translocation step is critical for high-processivity, as shown by MD simulations of the DNA Pol-I, where this process is mediated by significant structural rearrangements regulated by conserved Pol residues.28,

139-141

All these events may also involve additional conserved second-shell

residues that have recently been reported to expand the two-metal-ion architecture in several nucleic-acid-processing enzymes.104 Then there are low-processive and low-fidelity Pols, such as Pol- which promotes misincorporation events in order to produce genome variability for antibody generation (a process named somatic hypermutation).35, 142-144 One hypothesis about these Pols is that the correct dNTP substrate remains preferentially bound to the binary Pol/dsDNA complex due to a more stable and better structured hydrogen bond pattern and base-stacking interaction.145 In particular, mutagenesis and kinetics studies have shown that the residue Arg61 in Pol- is critical for catalysis.146-148 The catalytic role of this specific residue was also confirmed via MD and QM/MM calculations, which revealed its contribution to stabilizing the TS for nucleotide addition, and facilitating PPi departure, acting in cooperation with MgC.21, 41 Forcefield-based MD simulations have shown that, compared to undamaged DNA, defined hydrogen bond patterns and stacking interaction favor binding of selective dNTPs to Pol- in the presence of DNA damage (thymine−thymine dimer, TTD).22, 149 The conserved Arg61 thus seems to affect Pol- fidelity and processivity. This is in line with experimental observations of this specific Pol.22, 149

This interpretation is also supported by other studies, which indicate that a specific enzyme

side chain close to the catalytic site plays a key role in selecting the correct dNTP in other lowACS Paragon Plus Environment

16

Page 17 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

fidelity Pols, such as Pol- and in T7-RNA Pol.128 A large dataset of Pol/DNA structures feature a positively charged residue (Lys or Arg) near the enzymatic two-metal active site.45, 151 In these structures, this conserved residue always interacts directly with the incoming nucleotide (Fig. 2). Recently, classical MD simulations and free-energy calculations152 have suggested that the Arg61-mediated stabilization of the incoming dNTP may favor the formation of the Watson−Crick (W−C) pairing in Pol- Fig. 7).151 Incidentally, distorted W−C or even Hoogsteen base-pairing conformations have been observed in some structures of low-fidelity human DNA Pol- and the African swine fever virus Pol-X.154,

155

In these structures, the

conserved positively charged residue (e.g. Arg61 in Pol- was either missing or slightly displaced.151 This positively charged residue is undoubtedly not the only element shaping the nascent nucleic acid geometry in Pols. Other structural elements may include recently identified and conserved second-shell residues.104 ‘Pre-chemistry protein conformational changes’ is the collective term for the catalytic sidechain rearrangements and/or major enzyme domain dynamics that occur prior to the catalytic step. The potential impact of pre-chemistry protein conformational changes on Pol catalysis is the subject of lively debate, with studies centered on Pol .20, 156-162 As thoroughly discussed by Mulholland et al.,157 the mechanistic implications of pre-chemistry conformational changes for Pol function depend on whether one considers protein dynamics in relation to the formation of a reactive protein conformation, or in terms of local atomic motions that may affect the reaction rate. Moreover, these arguments rely on specific assumptions about the rate-limiting step for Pol’s fidelity and reactivity, in the context of the complex kinetic behavior of the enzymatic system.157,163 In summary, we have described a stepwise and complex sequence of chemical and physical steps, which regulate nucleotide addition for nucleic acid elongation, and the fidelity and processivity of Pols in catalyzing nucleic acid biosynthesis. These steps include chemical reactions for the addition of a nucleotide to the growing primer, as well as large-amplitude conformational changes for substrate binding and unbinding, and the final nucleic acid translocation (or Pol dissociation from the template strand). But what is the rate-determining ACS Paragon Plus Environment

17

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 57

step for this multifactorial DNA/RNA extension process? This question is the subject of intense investigation and debate. One proposition is that the rate-determining step may be different for different Pols, and may vary according to whether the inserted nucleotide is correct or not.154, 164-167

Indeed, correct and incorrect nucleotide incorporation may involve different

conformational changes for the recognition and binding of the entering nucleotide in every given Pol enzyme (thus affecting its fidelity).121, 126, 168, 169 A few cases have been reported where the physical step for pre-chemistry conformational changes is suggested to be rate-liming for Pol function.18, 170-171 However, based also on computational studies,28, 137, 172, 173, 160 another valid hypothesis is that the chemical step for nucleotide addition is the actual rate-liming step for Pol-mediated nucleic acid biosynthesis.15,

46, 174, 175

To further complicate the issue, it is

arduous to assess consistently the energy contribution of each discrete chemical and physical step for Pol catalysis and fidelity, relative to the energetic and kinetics of the overall Pols catalytic cycle. While there is no general consensus on the rate-limiting step, a thorough analysis and discussion can be found in Raper et al.’s comprehensive review of the kinetics mechanisms of Pols.48 In this respect, computational simulations will continue to play a critical role in interpreting and enriching the kinetics and structural data on Pol catalysis.

Outlook and conclusions. Pol enzymes play a crucial role in the synthesis and processing of nucleic acids within cells. For this reason, their overall functioning and catalysis are the subject of intense research. Pols are essential for biotechnology applications including DNA amplification, sequencing, and polymerization of synthetic genetic polymers.176-178 In addition, Pols are often targeted for therapeutic intervention, especially in cancer.59, 179-183 It is therefore critical to understand their mechanism in order to modulate their function. Here, we have outlined how computation can help to elucidate the enzymatic mechanism of Pol enzymes. More generally, a computational researcher can nowadays choose from among many different codes and methods based on different levels of theory (e.g. forcefield-based, semi-

ACS Paragon Plus Environment

18

Page 19 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

empirical like in EVB, ab initio, DFT and even more costly/accurate quantum calculations).184-194 These can be coupled to one or more approaches for free-energy calculations (e.g. thermodynamic integration, metadynamics, umbrella sampling, free-energy perturbation, transition-path sampling, minimum free-energy path scheme, string methods, and others).60, 195-200

As we have shown, these different computational methods and approaches can be used

to interrogate enzymatic reaction mechanisms, and to reason for or against a particular mechanism.201,

202

Thus, the choice of the computational strategy depends on the specific

catalytic step in Pols to be investigated. For example, Markov state models used to investigate major protein conformational changes require extensive sampling (in the s-timescale, and longer),203 which is only accessible by means of force-field based MD simulations. Computationally expensive quantum-based approaches,186, 187 on the other hand, are necessary to examine chemical reactions like, in Pols, proton transfer events and phosphate hydrolysis. That is, each method has its strengths, weaknesses and limitations, usually related to accuracy and sampling. These two variables heavily define the quality of computational investigations of catalytic mechanisms (which often also depends on the researcher’s chemical intuition). Sampling is a major limitation when only a few reaction pathways are investigated. This can be problematic if the sampling omits meaningful regions of the catalytic conformational space, which is not always obvious a priori. Accuracy is also critical, especially when possible pathways return similar energy profiles, making it challenging to compare them. These limitations and challenges are even more difficult when compared to the sophisticated process of enzymemediated nucleic acid processing. With the advent of more powerful computers and algorithms for extended MD simulations and analyses, and advanced approaches to studying catalysis (e.g. machine-learning methods to investigate reaction mechanisms),204 we expect great progress in further elucidating the energetics and dynamics of the complex enzymatic mechanisms of Pol catalysis. Understanding Pol catalysis and mechanism of function is relevant for drug discovery to target native Pols involved in pathophysiological processes. However, it is also relevant in the fast-growing field of artificial Pols design. One such example is the development of new ACS Paragon Plus Environment

19

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 57

unnatural nucleotides to store and propagate genetic information, with modifications in the sugar ring and in the nucleobases.205 One active area is the design of unnatural nucleotides (referred to as xeno nucleic acids or XNA) to develop synthetic XNA polymers, which can store information and respond to external stimuli, while being immune to endogenous nuclease activity. Another example is XNA polymers that can perform catalysis (xenozymes) for practical applications such as diagnostic and therapeutic technology.206 In this context, computational studies will help increase our understanding of how these artificial objects operate. This will require the development of new force-field parameters and, most likely, new and smart computational approaches to realistically simulate how these synthetic molecules interact with endogenous enzymes. Notably, advances in computational methods and applications will advance the field toward the rational design of artificial Pols, which represent a new frontier in the era of precision genome editing. This Perspective has discussed representative studies that illustrate how computational modelling and simulations have been used to investigate Pol catalysis. These studies have generated insights into the metal-aided mechanisms of enzymatic chemical transformations, which occur in complex multistep reactions catalyzed by Pol enzymes. We conclude by emphasizing that these insights are not only intellectually rewarding but also necessary to interpret experimental data. We expect modern computational methods and approaches to continue to be of help in elucidating the atomic-level motions of Pol action, in enriching the interpretation of existing structural and mechanistic data, and in guiding the development of new experimental studies. For example, computational studies have generated information about the exact mechanism, geometry, and chemico-physical properties of the enzymatic TS of pharmacologically relevant Pols. This information can be used in the design of metal-dependent drugs207 and artificial enzymes that perform biomimetic chemistry, positively impacting human health and technology.208-210 We thus look forward to future computational endeavors that will shed new light on the enzymatic machinery behind the fascinating process of Pol-mediated nucleic acid biosynthesis.

ACS Paragon Plus Environment

20

Page 21 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Acknowledgments M.D.V. thanks the Italian Association for Cancer Research (AIRC) for financial support (IG 18883). V.G. thanks the European Molecular Biology Organization (EMBO) for financial support (ALTF 103-2018). We also thank Grace Fox for proofreading and copyediting the manuscript.

ACS Paragon Plus Environment

21

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 57

Figures

Figure 1. Schematic representation of nucleic acid synthesis catalyzed by RNA/DNA polymerases. Precatalytic state, nucleophile activation, nucleotide (d)NTP addition, and nucleic acid translocation with consequent postcatalytic product formation and liberation of pyrophosphate (PPi) leaving group. Green indicates the template strand (T), blue indicates the primer strand (P). The nucleophilic 3’O – is depicted in orange.

ACS Paragon Plus Environment

22

Page 23 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Figure 2. (Left) Overview of the general ternary Pol/(R)DNA/(d)NTP complex. Each Pol domain is differently colored: palm in yellow; thumb in blue; fingers in cyan; little finger in red. (Right) Close view of the two-metal-aided catalytic site commonly found in Pols. The two metal ions are in orange, nitrogen is in blue, carbon is in white, oxygen is in red, and phosphorus is in maroon. The residue K(R), depicted with cyan carbon, identifies the conserved Arg/Lys, which has been systematically found in this conserved position on top of the active site. The K1- and K2-like residues, with green carbon, are basic second-shell residues, which have been discovered in all biomolecules that perform nucleic acid editing.104

ACS Paragon Plus Environment

23

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 57

Figure 3. (A) Graph reporting the intramolecular H-bond (indicate as d-PT) in different polymerases. The length of d-PT is reported for structures of Pol families from each domain of life. The X-axis reports the protein name. The Y-axis reports d-PT length (Å). Green dots identify X-ray structures of Pols from prokaryotes, cyan from eukaryotes, and red from viruses. The background color indicates the enzyme commission number (EC number provided above). (B) Superimposition of (ribo)nucleotides co-crystallized in Pol reactive ternary complexes. Structures extracted from different crystals are superimposed following their species (A, C, G, T, U). The upper part indicates the conserved presence of d-PT in those (ribo/deoxy)nucleotides complexed with Pol/(R)DNA binary complexes. The lower part shows the C3′-endo sugar pucker ACS Paragon Plus Environment

24

Page 25 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

conformation that is always detected in these structures. Ribonucleotides (RNA) are cyan, deoxynucleotides (DNA) are white. The value reported for d-PT is the average value obtained for each type of (ribo/deoxy)nucleotide. Figure adapted from Ref.41.

Figure 4. Reaction mechanisms proposed for nucleotidyl transfer catalyzed by Pols. Horizontal shaded areas highlight the two proposed mechanisms: protein-mediated44 and watermediated and substrate-assisted43 (WMSA). Vertical shaded areas indicate the three major chemical steps for Pol catalysis (i.e. nucleophile activation in the prereactive state, nucleotide addition, and DNA translocation, followed by leaving group departure). Above: the proteinmediated mechanism, whereby a conserved Asp (often part of the widely conserved DED-motif, which is a structural feature of the two-metal-ion mechanism) receives the proton from the nucleophilic group. This proton transfer thus activates the nucleophile for the phosphoryltransfer reaction. The red arrow denotes possible proton transfer events in the proteinmediated mechanism. The dashed green arrow indicates a second plausible nucleophile ACS Paragon Plus Environment

25

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 57

activation pathway, whereby the proton directly migrates on the non-bridging oxygen of the dNTP -phosphate. Similarly, the dashed cyan arrow depicts an alternative nucleophile activation process mediated by a hydroxide anion in the bulk. Below: the WMSA requires at least two water molecules to mediate the proton shuttle and relay to the -phosphate via the -phosphate and the dNTP as the reaction proceeds.

ACS Paragon Plus Environment

26

Page 27 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Figure 5. Reaction scheme showing the self-activated mechanism (SAM) for nucleic acid polymerization.41 (A) Michaelis–Menten complex: This state leads to the two-metal-aided SN2type phosphoryl transfer with liberation of the pyrophosphate (PPi) leaving group. Notably, the nucleophilic oxygen is already activated here (i.e. deprotonated). (B) Products for nucleotide addition: here, the incoming nucleotide was added to the primer strand. Colored lines indicate selected distances taken as collective variables (CV1 = r1 – r2 and CV2 = r3 – r4 for QM/MM metadynamics) to investigate SAM. (C) Nucleophile formation and nucleic acid translocation: the nucleophile 3′-OH is activated through its deprotonation in favor of the leaving PPi (PT1), while r4 is progressively shortened, indicating initial nucleic acid translocation. (D) PPi exit: at this point, the newly formed 3′-hydroxide group of the incoming nucleotide is coordinated on top of metal A, while the leaving PPi departs from the catalytic site, helped by the transient third ACS Paragon Plus Environment

27

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 57

metal ion. (E) dNTP binding and catalytic site closure: the enzyme is ready for the subsequent polymerization cycle upon binding of a new nucleotide, with closure of the catalytic cycle. Figure from Ref.41.

ACS Paragon Plus Environment

28

Page 29 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Figure 6. (A, left) Overlap of the structures reported in (B, right) and manually aligned with Coot using the substrate and the two-metal-ion center as a guide.104 Substrate (green cartoon) and MA-MB metals (orange spheres) are reported for endonuclease BamHI only. Blue and red spheres depict the position of the K1- and K2-like elements from all structures, respectively. Grey circles represent spheres of a radius of 4.0 Å (K1) and 3.5 Å (K2) and identify sites 4 (K1) and 3 (K2), respectively. (B) Distances in angstroms of K1-like elements from the acidic residues that coordinate MA-MB (d K1-acidic, blue dots) and of K2-like elements from the substrate (d K2-substrate, red dots). The grey shade covers the optimal range of distances for hydrogen bonds and ionic interactions (Ippolito et al., 1990).211 Several outliers correspond to structures solved at low resolution (i.e. PDB: 5FJ8)212 or with no metals in the active site (i.e. PDB: 1IAW).213 For enzymes in which K1 contacts acidic residues indirectly (PDB: 2BAM,214 1DMU,215 1QPS, and 2ALZ216), we reported the closest distance to the linking residue. For structures where K2 is present,

but

the

substrate

is

not

resolved

in

the

PDB

file

(PDB: 3S1S,217 1FOK,218 4OGC,219 5B2O,220 and 5AXW221), we did not plot any data. In exo-λ structure PDB: 4WUZ,222 the “d K1-acidic” and “d K2-substrate” data points overlap and only the blue dot is actually visible. The names of the enzymes are on the x-axis, corresponding PDB codes are indicated on top of every data point. Enzymes are grouped by classes and their

ACS Paragon Plus Environment

29

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 57

respective Enzymatic Classification (E.C.) numbers are indicated on the top of the graph. Figure adapted from Ref.104.

ACS Paragon Plus Environment

30

Page 31 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Figure 7. (A-B) Free-energy surface for SAM in human DNA Pol-η. B, PT1, PT2, C, and D identify saddle points for SAM-catalyzed nucleic acid polymerization in DNA Pol-η, moving from point B of the catalytic cycle to an ensemble of global minima at point D (see reaction scheme and points B and D in Figure 5).41 (C-D) Free-energy simulations of Watson–Crick base pairing (W– C) stability in the presence and absence of K(R) side chain (i.e., Arg61 in Pol-η). Both free-energy surfaces (FESs) were reconstructed using ωPu and ωPy151 as collective variables to energetically describe dATP:dT base-pair stability. In particular, C depicts the FES of wild-type system showing two distinct energetic basins with different ωPu and ωPy values. The deepest corresponds to Watson–Crick base pairing (W–C) while the relative one corresponds to the Hoogsteen base pair (HG). In contrast, D displays the FES of R61A mutant system showing large and deep energetic basins with different ωPu and ωPy values. Here, the deepest energetic minimum

ACS Paragon Plus Environment

31

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 57

represents an ensemble of dATP:dT architectures far from the canonical W–C pairing. Of these, HG is the most sampled. In contrast, a relative minimum identifies a cluster of architectures close to the iconic W–C pairing. Figure adapted from Refs.41, 151 Pol Family

Species

Pols

S. cerevisiae Archaeal Archaeal Engineered

DNA Pol-δ KOD DNA-Pol 9°N DNA-Pol KOD-RI TNA Pol

Human

DNA Pol-β

B-family

X-family

DNA Pol-μ Human

DNA Pol-η

Y-family

PDB ID 3IAY113 5OMF111 5OMQ111 5VU8112 [4UAY, 4UB3, 4UB5, 4UBB, 4UBC],5 [4RPY, 4RPZ, 4RQ0, 4RQ2, 4RQ4, 4RQ5, 4RQ6, 4RQ8],109 [4KLG, 4KLH, 4KLI, 4KLJ, 4KLO, 4KLQ],31 [3RH4, 3RH5, 3RH6],125 3JPP223 4M0A,110 [5TYY, 5TYX, 5TYW, 5TYV, 5TYU],33 [5VZ9, 5VZC, 5VZF, 5VZI]224 [4ECT, 4ECU, 4ECV, 4ECW, 4ECX],34 [5KFH, 5KFI, 5KFJ, 5KFK, 5KFL, 5KFN, 5KFP, 5KFW, 5KFX, 5KFZ, 5KG0, 5KG1, 5KG2, 5KG3, 5KG4, 5KG5, 5KG6, 5KG7, 5L9X]47

Table 1. X-ray structures of ternary Pols/DNA/dNTP complexes with three metal ions in their active sites.

ACS Paragon Plus Environment

32

Page 33 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

References 1.

Potapov, V.; Fu, X.; Dai, N.; Correa, I. R., Jr.; Tanner, N. A.; Ong, J. L. Base modifications

affecting RNA polymerase and reverse transcriptase fidelity. Nucleic Acids Res. 2018, 46, 5753-5763. 2.

Loeb, L. A.; Monnat, R. J. DNA polymerases and human disease. Nat. Rev. Genet. 2008, 9, 594-

604. 3.

Cramer, P.; Armache, K. J.; Baumli, S.; Benkert, S.; Brueckner, E.; Buchen, C.; Damsma, G.

E.; Dengl, S.; Geiger, S. R.; Jaslak, A. J.; Jawhari, A.; Jennebach, S.; Kamenski, T.; Kettenberger, H.; Kuhn, C. D.; Lehmann, E.; Leike, K.; Sydow, J. E.; Vannini, A. Structure of eukaryotic RNA polymerases. Annu. Rev. Biophys. 2008, 37, 337-352. 4.

Dasari, A.; Deodhar, T.; Berdis, A. J. A Comparative Analysis of Translesion DNA Synthesis

Catalyzed by a High-Fidelity DNA Polymerase. J. Mol. Biol. 2017, 429, 2308-2323. 5.

Freudenthal, B. D.; Beard, W. A.; Perera, L.; Shock, D. D.; Kim, T.; Schlick, T.; Wilson, S.

H. Uncovering the polymerase-induced cytotoxicity of an oxidized nucleotide. Nature 2015, 517, 635639. 6.

Lange, S. S.; Takata, K.; Wood, R. D. DNA polymerases and cancer. Nat. Rev. Cancer 2011, 11,

96-110. 7.

Rai, D. K.; Diaz-San Segundo, F.; Campagnola, G.; Keith, A.; Schafer, E. A.; Kloc, A.; de

Los Santos, T.; Peersen, O.; Rieder, E. Attenuation of Foot-and-Mouth Disease Virus by Engineered Viral Polymerase Fidelity. J. Virol. 2017, e000081-17. 8.

Suzuki, T.; Gruz, P.; Honma, M.; Adachi, N.; Nohmi, T. Sensitivity of human cells expressing

low-fidelity or weak-catalytic-activity variants of DNA polymerase zeta to genotoxic stresses. DNA Repair (Amst) 2016, 45, 34-43. 9.

Xia, S. L.; Wang, J. M.; Konigsberg, W. H. DNA Mismatch Synthesis Complexes Provide

Insights into Base Selectivity of a B Family DNA Polymerase. J. Am. Chem. Soc. 2013, 135, 193-202.

ACS Paragon Plus Environment

33

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10.

Page 34 of 57

Goodman, M. F.; Tippin, B. The expanding polymerase universe. Nat. Rev. Mol. Cell Biol. 2000,

1, 101-109. 11.

Laos, R.; Thomson, J. M.; Benner, S. A. DNA polymerases engineered by directed evolution to

incorporate non-standard nucleotides. Front. Microbiol. 2014, 5, 565. 12.

Kranaster, R.; Marx, A. Engineered DNA polymerases in biotechnology. Chembiochem 2010,

11, 2077-2084. 13.

Chen, T.; Romesberg, F. E. Directed polymerase evolution. FEBS Lett. 2014, 588, 219-229.

14.

Ghadessy, F. J.; Ramsay, N.; Boudsocq, F.; Loakes, D.; Brown, A.; Iwai, S.; Vaisman, A.;

Woodgate, R.; Holliger, P. Generic expansion of the substrate spectrum of a DNA polymerase by directed evolution. Nat. Biotechnol. 2004, 22, 755-759. 15.

Yang, W. An overview of Y-Family DNA polymerases and a case study of human DNA

polymerase eta. Biochemistry 2014, 53, 2793-2803. 16.

Murakami, K. S. Structural biology of bacterial RNA polymerase. Biomolecules 2015, 5, 848-

864. 17.

Basu, R. S.; Warner, B. A.; Molodtsov, V.; Pupov, D.; Esyunina, D.; Fernandez-Tornero, C.;

Kulbachinskiy, A.; Murakami, K. S. Structural basis of transcription initiation by bacterial RNA polymerase holoenzyme. J. Biol. Chem. 2014, 289, 24549-24559. 18.

Meli, M.; Sustarsic, M.; Craggs, T. D. Kapanidis, A. N.; Colombo, G., DNA Polymerase

Conformational Dynamics and the Role of Fidelity-Conferring Residues: Insights from Computational Simulations. Front. Mol. Biosci. 2016, 3, 20. 19.

Xiang, Y.; Goodman, M. F.; Beard, W. A.; Wilson, S. H.; Warshel, A. Exploring the role of

large conformational changes in the fidelity of DNA polymerase beta. Proteins 2008, 70, 231-247. 20.

Schlick, T.; Arora, K.; Beard, W. A.; Wilson, S. H. Perspective: pre-chemistry conformational

changes in DNA polymerase mechanisms. Theor. Chem. Acc. 2012, 131, 1287-1295.

ACS Paragon Plus Environment

34

Page 35 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

21.

ACS Catalysis

Genna, V.; Gaspari, R.; Dal Peraro, M.; De Vivo, M. Cooperative motion of a key positively

charged residue and metal ions for DNA replication catalyzed by human DNA Polymerase-eta. Nucleic Acids Res. 2016, 44, 2827-2836. 22.

Ucisik, M. N.; Hammes-Schiffer, S. Relative Binding Free Energies of Adenine and Guanine to

Damaged and Undamaged DNA in Human DNA Polymerase eta: Clues for Fidelity and Overall Efficiency. J. Am. Chem. Soc. 2015, 137, 13240-13243. 23.

Batra, V. K.; Beard, W. A.; Shock, D. D.; Pedersen, L. C.; Wilson, S. H. Nucleotide-induced

DNA polymerase active site motions accommodating a mutagenic DNA intermediate. Structure 2005, 13, 1225-1233. 24.

Langer, A.;

Schraml, M.;

Strasser, R.;

Daub, H.;

Myers, T.;

Heindl, D.; Rant, U.

Polymerase/DNA interactions and enzymatic activity: multi-parameter analysis with electro-switchable biosurfaces. Sci. Rep. 2015, 5, 12066-12081. 25.

Hottin, A.; Marx, A. Structural Insights into the Processing of Nucleobase-Modified Nucleotides

by DNA Polymerases. Acc. Chem. Res. 2016, 49, 418-427. 26.

Berman, A. J.; Kamtekar, S.; Goodman, J. L.; Lazaro, J. M.; de Vega, M.; Blanco, L.; Salas,

M.; Steitz, T. A. Structures of phi29 DNA polymerase complexed with substrate: the mechanism of translocation in B-family polymerases. EMBO J. 2007, 26, 3494-3505. 27.

Kirby, T. W.; DeRose, E. F.; Cavanaugh, N. A.; Beard, W. A.; Shock, D. D.; Mueller, G. A.;

Wilson, S. H.; London, R. E. Metal-induced DNA translocation leads to DNA polymerase conformational activation. Nucleic Acids Res. 2012, 40, 2974-2983. 28.

Golosov, A. A.; Warren, J. J.; Beese, L. S.; Karplus, M. The Mechanism of the Translocation

Step in DNA Replication by DNA Polymerase I: A Computer Simulation Analysis. Structure 2010, 18, 83-93.

ACS Paragon Plus Environment

35

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

29.

Page 36 of 57

Dahl, J. M.; Mai, A. H.; Cherf, G. M.; Jetha, N. N.; Garalde, D. R.; Marziali, A.; Akeson, M.;

Wang, H.; Lieberman, K. R. Direct observation of translocation in individual DNA polymerase complexes. J. Biol. Chem. 2012, 287, 13407-13421. 30.

Wu, S.; Beard, W. A.; Pedersen, L. G.; Wilson, S. H. Structural comparison of DNA polymerase

architecture suggests a nucleotide gateway to the polymerase active site. Chem. Rev. 2014, 114, 27592774. 31.

Freudenthal, B. D.; Beard, W. A.; Shock, D. D.; Wilson, S. H. Observing a DNA Polymerase

Choose Right from Wrong. Cell 2013, 154, 157-168. 32.

Chen, Y.; Basu, R.; Gleghorn, M. L.; Murakami, K. S.; Carey, P. R. Time-resolved events on

the reaction pathway of transcript initiation by a single-subunit RNA polymerase: Raman crystallographic evidence. J. Am. Chem. Soc. 2011, 133, 12544-12555. 33.

Jamsen, J. A.; Beard, W. A.; Pedersen, L. C.; Shock, D. D.; Moon, A. F.; Krahn, J. M.;

Bebenek, K.; Kunkel, T. A.; Wilson, S. H. Time-lapse crystallography snapshots of a double-strand break repair polymerase in action. Nat. Commun. 2017, 8, 1-11 34.

Nakamura, T.; Zhao, Y.; Yamagata, Y.; Hua, Y. J.; Yang, W. Watching DNA polymerase eta

make a phosphodiester bond. Nature 2012, 487, 196-201. 35.

Zhao, Y.; Gregory, M. T.; Biertumpfel, C.; Hua, Y. J.; Hanaoka, F.; Yang, W. Mechanism of

somatic hypermutation at the WA motif by human DNA polymerase eta. Proc. Natl. Acad. Sci. U.S.A. 2013, 110, 8146-8151. 36.

Palermo, G.; Cavalli, A.; Klein, M. L.; Alfonso-Prieto, M.; Dal Peraro, M.; De Vivo, M.

Catalytic metal ions and enzymatic processing of DNA and RNA. Acc. Chem. Res. 2015, 48, 220-228. 37.

Steitz, T. A. DNA-Dependent and Rna-Dependent DNA-Polymerases. Curr. Opin. Struc. Biol.

1993, 3, 31-38.

ACS Paragon Plus Environment

36

Page 37 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

38.

ACS Catalysis

Steitz, T. A.; Steitz, J. A. A general two-metal-ion mechanism for catalytic RNA. Proc. Natl.

Acad. Sci. U.S.A. 1993, 90, 6498-6502. 39.

Vashishtha, A. K.; Wang, J.; Konigsberg, W. H. Different Divalent Cations Alter the Kinetics

and Fidelity of DNA Polymerases. J. Biol. Chem. 2016, 291, 20869-20875. 40.

Yang, W.; Lee, J. Y.; Nowotny, M. Making and breaking nucleic acids: two-Mg2+-ion catalysis

and substrate specificity. Mol. Cell 2006, 22, 5-13. 41.

Genna, V.; Vidossich, P.; Ippoliti, E.; Carloni, P.; De Vivo, M., A Self-Activated Mechanism

for Nucleic Acid Polymerization Catalyzed by DNA/RNA Polymerases. J. Am. Chem. Soc. 2016, 138, 14592-14598. 42.

Wang, L. H.; Broyde, S.; Zhang, Y. K. Polymerase-Tailored Variations in the Water-Mediated

and Substrate-Assisted Mechanism for Nucleotidyl Transfer: Insights from a Study of T7 DNA Polymerase. J. Mol. Biol. 2009, 389, 787-796. 43.

Wang, L. H.; Yu, X. Y.; Hu, P.; Broyde, S.; Zhang, Y. K. A water-mediated and substrate-

assisted catalytic mechanism for Sulfolobus solfataricus DNA polymerase IV. J. Am. Chem. Soc. 2007, 129, 4731-4737. 44.

Florian, J.; Goodman, M. F.; Warshel, A. Computer simulation of the chemical catalysis of DNA

polymerases: Discriminating between alternative nucleotide insertion mechanisms for T7 DNA polymerase. J. Am. Chem. Soc. 2003, 125, 8163-8177. 45.

Castro, C.; Smidansky, E. D.; Arnold, J. J.; Maksimchuk, K. R.; Moustafa, I.; Uchida, A.;

Gotte, M.; Konigsberg, W.; Cameron, C. E. Nucleic acid polymerases use a general acid for nucleotidyl transfer. Nat. Struct. Mol. Biol. 2009, 16, 212-218. 46.

Yang, W.; Weng, P. J.; Gao, Y. A new paradigm of DNA synthesis: three-metal-ion catalysis.

Cell Biosci. 2016, 6, 51-58.

ACS Paragon Plus Environment

37

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

47.

Page 38 of 57

Gao, Y.; Yang, W. Capture of a third Mg2+ is essential for catalyzing DNA synthesis. Science

2016, 352, 1334-1337. 48.

Raper, A. T.; Reed, A. J.; Suo, Z. Kinetic Mechanism of DNA Polymerases: Contributions of

Conformational Dynamics and a Third Divalent Metal Ion. Chem. Rev. 2018, 118, 6000-6025. 49.

Jain, R.; Aggarwal, A. K.; Rechkoblit, O. Eukaryotic DNA polymerases. Curr. Opin. Struct.

Biol. 2018, 53, 77-87. 50.

Wu, W. J.; Yang, W; Tsai, M. D. How DNA polymerases catalyse replication and repair with

contrasting fidelity. Nat. Rev. Chem. 2017, 1, 1-16. 51.

Joyce, C. M.; Potapova, O.; Delucia, A. M.; Huang, X.; Basu, V. P.; Grindley, N. D. Fingers-

closing and other rapid conformational changes in DNA polymerase I (Klenow fragment) and their role in nucleotide selectivity. Biochemistry 2008, 47, 6103-6116. 52.

Potapova, O.; Chan, C.; DeLucia, A. M.; Helquist, S. A.; Kool, E. T.; Grindley, N. D.; Joyce,

C. M. DNA polymerase catalysis in the absence of Watson-Crick hydrogen bonds: analysis by singleturnover kinetics. Biochemistry 2006, 45, 890-898. 53.

Joyce, C. M. Choosing the right sugar: how polymerases select a nucleotide substrate. Proc. Natl.

Acad. Sci. U.S.A. 1997, 94, 1619-1622. 54.

Echols, H.; Goodman, M. F. Fidelity mechanisms in DNA replication. Annu. Rev. Biochem. 1991,

60, 477-511. 55.

Matsumoto, Y.; Kim, K. Excision of deoxyribose phosphate residues by DNA polymerase beta

during DNA repair. Science 1995, 269, 699-702. 56.

Johnson, K. A. The kinetic and chemical mechanism of high-fidelity DNA polymerases. Biochim.

Biophys. Acta 2010, 1804, 1041-1048. 57.

Yamtich, J.; Sweasy, J. B. DNA polymerase family X: function, structure, and cellular roles.

Biochim. Biophys. Acta 2010, 1804, 1136-1150.

ACS Paragon Plus Environment

38

Page 39 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

58.

ACS Catalysis

Brown, J. A.; Pack, L. R.; Sanman, L. E.; Suo, Z. Efficiency and fidelity of human DNA

polymerases lambda and beta during gap-filling DNA synthesis. DNA Repair (Amst) 2011, 10, 24-33. 59.

Walker, A. R.; Cisneros, G. A. Computational Simulations of DNA Polymerases: Detailed

Insights on Structure/Function/Mechanism from Native Proteins to Cancer Variants. Chem. Res. Toxicol. 2017, 30, 1922-1935. 60.

Dreyer, J., Brancato, G., Ippoliti, E., Genna, V., De Vivo, M., Carloni, P., Rothlisberger, U. In

Simulating Enzyme Reactivity: Computational Methods in Enzyme Catalysis; Tuñón, I., Moliner, V., Eds.; Royal Society of Chemistry: London, 2016; p 294. 61.

Dans, P. D.; Walther, J.; Gomez, H.; Orozco, M. Multiscale simulation of DNA. Curr. Opin.

Struct. Biol. 2016, 37, 29-45. 62.

De Vivo, M.; Masetti, M.; Bottegoni, G.; Cavalli, A. Role of Molecular Dynamics and Related

Methods in Drug Discovery. J. Med. Chem. 2016, 59, 4035-4061. 63.

De Vivo, M.; Cavalli, A. Recent advances in dynamic docking for drug discovery. WIREs

Comput. Mol. Sci. 2017, e1320. 64.

Wang, P.; Oscarson, L.; Izatt, R. M.; Watt, G. D.; Larsen, C. D. Thermodynamic parameters for

the interaction of adenosine 5′-diphosphate, and adenosine 5′-triphosphate with Mg2+ from 323.15 to 398.15 K. J. Solution Chem. 1995, 24, 989-1012. 65.

Harrison, C. B.; Schulten, K. Quantum and Classical Dynamics Simulations of ATP Hydrolysis

in Solution. J. Chem. Theory. Comput. 2012, 8, 2328-2335. 66.

Akola, J.; Jones, R. O. ATP hydrolysis in water - A density functional study. J. Phys. Chem. B

2003, 107, 11774-11783. 67.

Ho, M. H.; De Vivo, M.; Dal Peraro, M.; Klein, M. L. Understanding the effect of magnesium

ion concentration on the catalytic activity of ribonuclease H through computation: does a third metal binding site modulate endonuclease catalysis? J. Am. Chem. Soc. 2010, 132, 13702-13712.

ACS Paragon Plus Environment

39

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

68.

Page 40 of 57

De Vivo, M.; Dal Peraro, M.; Klein, M. L. Phosphodiester cleavage in ribonuclease H occurs via

an associative two-metal-aided catalytic mechanism. J. Am. Chem. Soc. 2008, 130, 10955-10962. 69.

De Vivo, M.; Cavalli, A.; Carloni, P.; Recanatini, M. Computational study of the phosphoryl

transfer catalyzed by a cyclin-dependent kinase. Chemistry 2007, 13, 8437-8444. 70.

De Vivo, M.; Ensing, B.; Dal Peraro, M.; Gomez, G. A.; Christianson, D. W.; Klein, M. L.

Proton shuttles and phosphatase activity in soluble epoxide hydrolase. J. Am. Chem. Soc. 2007, 129, 387394. 71.

De Vivo, M.; Ensing, B.; Klein, M. L. Computational study of phosphatase activity in soluble

epoxide hydrolase: high efficiency through a water bridge mediated proton shuttle. J. Am. Chem. Soc. 2005, 127, 11226-11227. 72.

Steitz, T. A. DNA polymerases: structural diversity and common mechanisms. J. Biol. Chem.

1999, 274, 17395-17398. 73.

Beard, W. A.; Wilson, S. H. Structure and mechanism of DNA polymerase Beta. Chem. Rev.

2006, 106, 361-382. 74.

Berdis, A. J. Mechanisms of DNA polymerases. Chem. Rev. 2009, 109, 2862-2879.

75.

Oertell, K.; Chamberlain, B. T.; Wu, Y.; Ferri, E.; Kashemirov, B. A.; Beard, W. A.; Wilson,

S. H.; McKenna, C. E.; Goodman, M. F. Transition state in DNA polymerase beta catalysis: rate-limiting chemistry altered by base-pair configuration. Biochemistry 2014, 53, 1842-1848. 76.

Lin, P.; Pedersen, L. C.; Batra, V. K.; Beard, W. A.; Wilson, S. H.; Pedersen, L. G. Energy

analysis of chemistry for correct insertion by DNA polymerase beta. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 13294-13299. 77.

Cisneros, G.A.; Perera L.; García-Díaz M.; Bebenek K.; Kunkel T.A.; Pedersen L.G. Catalytic

mechanism of human DNA polymerase lambda with Mg2+ and Mn2+ from ab initio quantum mechanical/molecular mechanical studies. DNA Repair (Amst). 2008, 7(11), 1824-34.

ACS Paragon Plus Environment

40

Page 41 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

78.

ACS Catalysis

Miller, B. R.; Beese, L. S.; Parish, C. A.; Wu, E. Y. The Closing Mechanism of DNA Polymerase

I at Atomic Resolution. Structure 2015, 23, 1609-1620. 79.

Lior-Hoffmann, L.; Wang, L. H.; Wang, S. L.; Geacintov, N. E.; Broyde, S.; Zhang, Y. K.

Preferred WMSA catalytic mechanism of the nucleotidyl transfer reaction in human DNA polymerase kappa elucidates error-free bypass of a bulky DNA lesion. Nucleic Acids Res. 2012, 40, 9193-9205. 80.

Wang, Y.; Schlick, T. Quantum mechanics/molecular mechanics investigation of the chemical

reaction in Dpo4 reveals water-dependent pathways and requirements for active site reorganization. J. Am. Chem. Soc. 2008, 130, 13240-13250. 81.

Zhang, R.; Bhattacharjee, A.; Field, M. J.; Salahub, D. R. Multiple proton relay routes in the

reaction mechanism of RNAP II: assessing the effect of structural model. Proteins 2015, 83, 268-281. 82.

Da, L. T.; Pardo Avila, F.; Wang, D.; Huang, X. A two-state model for the dynamics of the

pyrophosphate ion release in bacterial RNA polymerase. PLoS Comput. Biol. 2013, 9, e1003020. 83.

Larson, M. H.; Zhou, J.; Kaplan, C. D.; Palangat, M.; Kornberg, R. D.; Landick, R.; Block, S.

M. Trigger loop dynamics mediate the balance between the transcriptional fidelity and speed of RNA polymerase II. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 6555-6560. 84.

Silva, D. A.; Weiss, D. R.; Avila, F. P.; Da, L. T.; Levitt, M.; Wang, D.; Huang, X. H.

Millisecond dynamics of RNA polymerase II translocation at atomic resolution. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 7665-7670. 85.

Huang, X. H. Simulating millisecond dynamics of RNA polymerase II translocation at atomic

resolution using Markov State models. Abstr. Pap. Am. Chem. Soc. 2014, 247. 86.

Davies, D. L.; Macgregor, S. A.; McMullin, C. L. Computational Studies of Carboxylate-

Assisted C-H Activation and Functionalization at Group 8-10 Transition Metal Centers. Chem. Rev. 2017, 117, 8649-8709.

ACS Paragon Plus Environment

41

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

87.

Page 42 of 57

Balcells, D.; Clot, E.; Eisenstein, O. C-H bond activation in transition metal species from a

computational perspective. Chem. Rev. 2010, 110, 749-823. 88.

Perrin, L.; Carr, K. J. T.; Mckay, D.; McMullin, C. L.; Macgregor, S. A.; Eisenstein, O. In

Computational Studies in Organometallic Chemistry; Macgregor, S. A., Eisenstein, O., Eds.; Springer: Berlin, 2015; Vol. 167, p 37. 89.

Ivanov, I.; Tainer, J. A.; McCammon, J. A. Unraveling the three-metal-ion catalytic mechanism

of the DNA repair enzyme endonuclease IV. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 1465-1470. 90.

Kamerlin, S. C. L.; Sharma, P. K.; Prasad, R. B.; Warshel, A. Why nature really chose phosphate.

Q. Rev. Biophys. 2013, 46, 1-132. 91.

Klahn, M.; Rosta, E.; Warshel, A., On the mechanism of hydrolysis of phosphate monoesters

dianions in solutions and proteins. J. Am. Chem. Soc. 2006, 128, 15310-15323. 92.

Lonsdale, R.; Ranaghan, K. E.; Mulholland, A. J. Computational enzymology. Chem. Commun.

(Camb) 2010, 46, 2354-2372. 93.

Kamerlin, S. C.; Haranczyk, M.; Warshel, A. Progress in ab initio QM/MM free-energy

simulations of electrostatic energies in proteins: accelerated QM/MM studies of pKa, redox reactions and solvation free energies. J. Phys. Chem. B 2009, 113, 1253-1272. 94.

Warshel, A.; Russell, S. T. Calculations of electrostatic interactions in biological systems and in

solutions. Q. Rev. Biophys. 1984, 17, 283-422. 95.

Warshel, A.; Sharma, P. K.; Kato, M.; Xiang, Y.; Liu, H.; Olsson, M. H. Electrostatic basis for

enzyme catalysis. Chem. Rev. 2006, 106, 3210-3235. 96.

Hanoian, P.; Liu, C. T.; Hammes-Schiffer, S.; Benkovic, S. Perspectives on electrostatics and

conformational motions in enzyme catalysis. Acc. Chem. Res. 2015, 48, 482-489. 97.

Hammes, G. G.; Benkovic, S. J.; Hammes-Schiffer, S. Flexibility, diversity, and cooperativity:

pillars of enzyme catalysis. Biochemistry 2011, 50, 10422-10430.

ACS Paragon Plus Environment

42

Page 43 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

98.

ACS Catalysis

Bhabha, G.; Lee, J.; Ekiert, D. C.; Gam, J.; Wilson, I. A.; Dyson, H. J.; Benkovic, S. J.;

Wright, P. E. A Dynamic Knockout Reveals That Conformational Fluctuations Influence the Chemical Step of Enzyme Catalysis. Science 2011, 332, 234-238. 99.

Branduardi, D.; De Vivo, M.; Rega, N.; Barone, V.; Cavalli, A. Methyl Phosphate Dianion

Hydrolysis in Solution Characterized by Path Collective Variables Coupled with DFT-Based Enhanced Sampling Simulations. J. Chem. Theory. Comput. 2011, 7, 539-543. 100.

Castro, C.; Smidansky, E.; Maksimchuk, K. R.; Arnold, J. J.; Korneeva, V. S.; Gotte, M.;

Konigsberg, W.; Cameron, C. E. Two proton transfers in the transition state for nucleotidyl transfer catalyzed by RNA- and DNA-dependent RNA and DNA polymerases. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 4267-4272. 101.

Nowotny, M.; Gaidamakov, S. A.; Crouch, R. J.; Yang, W. Crystal structures of RNase H bound

to an RNA/DNA hybrid: Substrate specificity and metal-dependent catalysis. Cell 2005, 121, 1005-1016. 102.

Shi, Y.; Hellinga, H. W.; Beese, L. S. Interplay of catalysis, fidelity, threading, and processivity

in the exo- and endonucleolytic reactions of human exonuclease I. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 6010-6015. 103.

Molina, R.; Stella, S.; Redondo, P.; Gomez, H.; Marcaida, M. J.; Orozco, M.; Prieto, J.;

Montoya, G. Visualizing phosphodiester-bond hydrolysis by an endonuclease. Nat. Struct. Mol. Biol. 2015, 22, 65-72. 104.

Genna, V.; Colombo, M.; De Vivo, M.; Marcia, M. Second-Shell Basic Residues Expand the

Two-Metal-Ion Architecture of DNA and RNA Processing Enzymes. Structure 2018, 26, 40-50 e42. 105.

Shishova, E. Y.; Yu, F.; Miller, D. J.; Faraldos, J. A.; Zhao, Y.; Coates, R. M.; Allemann, R.

K.; Cane, D. E.; Christianson, D. W. X-ray crystallographic studies of substrate binding to aristolochene synthase suggest a metal ion binding sequence for catalysis. J. Biol. Chem. 2008, 283, 15431-15439.

ACS Paragon Plus Environment

43

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

106.

Page 44 of 57

Hanna, R.; Doudna, J. A. Metal ions in ribozyme folding and catalysis. Curr. Opin. Chem. Biol.

2000, 4, 166-170. 107.

Syson, K.; Tomlinson, C. Chapados, B. R.; Sayers, J. R.; Tainer, J. A.; Williams, N. H.;

Grasby, J. A., Three metal ions participate in the reaction catalyzed by T5 flap endonuclease. J. Biol. Chem. 2008, 283, 28741-28746. 108.

Horton, N. C.; Perona, J. J. DNA cleavage by EcoRV endonuclease: Two metal ions in three

metal ion binding sites. Biochemistry 2004, 43, 6841-6857. 109.

Vyas, R.; Reed, A. J.; Tokarsky, E. J.; Suo, Z. C. Viewing Human DNA Polymerase beta

Faithfully and Unfaithfully Bypass an Oxidative Lesion by Time-Dependent Crystallography. J. Am. Chem. Soc. 2015, 137, 5225-5230. 110.

Moon, A. F.; Pryor, J. M.; Ramsden, D. A.; Kunkel, T. A.; Bebenek, K.; Pedersen, L. C.

Sustained active site rigidity during synthesis by human DNA polymerase mu. Nat. Struct. Mol. Biol. 2014, 21, 253-260. 111.

Kropp, H. M.; Betz, K.; Wirth, J.; Diederichs, K.; Marx, A., Crystal structures of ternary

complexes of archaeal B-family DNA polymerases. Plos One 2017, 12, e0188005. 112.

Chim, N.; Shi, C.; Sau, S. P.; Nikoomanzar, A.; Chaput, J. C. Structural basis for TNA synthesis

by an engineered TNA polymerase. Nat. Commun. 2017, 8, 1810-1821. 113.

Swan, M. K.; Johnson, R. E.; Prakash, L.; Prakash, S.; Aggarwal, A. K. Structural basis of high-

fidelity DNA synthesis by yeast DNA polymerase delta. Nat. Struct. Mol. Biol. 2009, 16, 979-986. 114.

Stevens, D. R.; Hammes-Schiffer, S. Exploring the Role of the Third Active Site Metal Ion in

DNA Polymerase eta with QM/MM Free Energy Simulations. J. Am. Chem. Soc. 2018, 140, 8965-8969. 115.

Perera, L.; Freudenthal, B. D.; Beard, W. A.; Shock, D. D.; Pedersen, L. G.; Wilson, S. H.

Requirement for transient metal ions revealed through computational analysis for DNA polymerase going in reverse. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, E5228-5236.

ACS Paragon Plus Environment

44

Page 45 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

116.

ACS Catalysis

Yoon, H.; Warshel, A. Simulating the fidelity and the three Mg mechanism of pol eta and

clarifying the validity of transition state theory in enzyme catalysis. Proteins 2017, 85, 1446-1453. 117.

Kottur, J.; Nair, D. T. Pyrophosphate hydrolysis is an intrinsic and critical step of the DNA

synthesis reaction Nucleic Acids Res. 2018, 46, 10. 118.

Perera, L.; Freudenthal, B. D.; Beard, W. A.; Pedersen, L. G.; Wilson, S. H. Revealing the role

of the product metal in DNA polymerase beta catalysis. Nucleic Acids Res. 2017, 45, 2736-2745. 119.

Da, L. T.; Wang, D.; Huang, X. H. Dynamics of pyrophosphate ion release and its coupled trigger

loop motion from closed to open state in RNA polymerase II. Abstr. Pap. Am. Chem. Soc. 2012, 243. 120.

Eckert, K. A.; Kunkel, T. A. DNA polymerase fidelity and the polymerase chain reaction. PCR

Methods Appl. 1991, 1, 17-24. 121.

Johnson, K. A. Conformational coupling in DNA polymerase fidelity. Annu. Rev. Biochem. 1993,

62, 685-713. 122.

Fiala, K. A.; Suo, Z. Mechanism of DNA polymerization catalyzed by Sulfolobus solfataricus P2

DNA polymerase IV. Biochemistry 2004, 43, 2116-2125. 123.

Santoso, Y.; Joyce, C. M.; Potapova, O.; Le Reste, L.; Hohlbein, J.; Torella, J. P.; Grindley,

N. D.; Kapanidis, A. N. Conformational transitions in DNA polymerase I revealed by single-molecule FRET. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 715-720. 124.

Showalter, A. K.; Tsai, M. D. A reexamination of the nucleotide incorporation fidelity of DNA

polymerases. Biochemistry 2002, 41, 10571-10576. 125.

Cavanaugh, N. A.; Beard, W. A.; Batra, V. K.; Perera, L.; Pedersen, L. G.; Wilson, S. H.

Molecular Insights into DNA Polymerase Deterrents for Ribonucleotide Insertion. J. Biol. Chem. 2011, 286, 31650-31660. 126.

Beckman, J. W.; Wang, Q. X.; Guengerich, F. P. Kinetic Analysis of Correct Nucleotide

Insertion by a Y-family DNA Polymerase Reveals Conformational Changes Both Prior to and following

ACS Paragon Plus Environment

45

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 57

Phosphodiester Bond Formation as Detected by Tryptophan Fluorescence. J. Biol. Chem. 2008, 283, 36711-36723. 127.

Li, Y.; Dutta, S.; Doublie, S.; Bdour, H. M.; Taylor, J. S.; Ellenberger, T. Nucleotide insertion

opposite a cis-syn thymine dimer by a replicative DNA polymerase from bacteriophage T7. Nat. Struct. Mol. Biol. 2004, 11, 784-790. 128.

Duan, B.; Wu, S.; Da, L. T.; Yu, J. A critical residue selectively recruits nucleotides for t7 RNA

polymerase transcription fidelity control. Biophys. J 2014, 107, 2130-2140. 129.

Nedialkov, Y. A.; Opron, K.; Assaf, F.; Artsimovitch, I.; Kireeva, M. L.; Kashlev, M.; Cukier,

R. I.; Nudler, E.; Burton, Z. F. The RNA polymerase bridge helix YFI motif in catalysis, fidelity and translocation. Biochim. Biophys. Acta 2013, 1829, 187-198. 130.

Malinen, A. M.; Turtola, M.; Parthiban, M.; Vainonen, L.; Johnson, M. S.; Belogurov, G. A.

Active site opening and closure control translocation of multisubunit RNA polymerase. Nucleic Acids Res. 2012, 40, 7442-7451. 131.

Subramanian, K.;

Rutvisuttinunt, W.;

Scott, W.; Myers, R. S. The enzymatic basis of

processivity in lambda exonuclease. Nucleic Acids Res. 2003, 31, 1585-1596. 132.

Khare, V.; Eckert, K. A. The proofreading 3'-->5' exonuclease activity of DNA polymerases: a

kinetic barrier to translesion DNA synthesis. Mutat. Res. 2002, 510, 45-54. 133.

Tran, H. T.; Gordenin, D. A.; Resnick, M. A. The 3'-->5' exonucleases of DNA polymerases

delta and epsilon and the 5'-->3' exonuclease Exo1 have major roles in postreplication mutation avoidance in Saccharomyces cerevisiae. Mol. Cell Biol. 1999, 19, 2000-2007. 134.

Miyabe, I.; Kunkel, T. A.; Carr, A. M. The major roles of DNA polymerases epsilon and delta

at the eukaryotic replication fork are evolutionarily conserved. PLoS Genet. 2011, 7, e1002407. 135.

Hindges, R.; Hubscher, U. DNA polymerase delta, an essential enzyme for DNA transactions.

Biol. Chem. 1997, 378, 345-362.

ACS Paragon Plus Environment

46

Page 47 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

136.

ACS Catalysis

Thommes, P.; Hubscher, U. Eukaryotic DNA replication. Enzymes and proteins acting at the

fork. Eur. J. Biochem. 1990, 194, 699-712. 137.

Florian, J.; Goodman, M. F.; Warshel, A. Computer simulations of protein functions: Searching

for the molecular origin of the replication fidelity of DNA polymerases. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 6819-6824. 138.

Ucisik, M. N.; Hammes-Schiffer, S. Relative Binding Free Energies of Adenine and Guanine to

Damaged and Undamaged DNA in Human DNA Polymerase eta: Clues for Fidelity and Overall Efficiency. J. Am. Chem. Soc. 2015, 137, 13240-13243. 139.

Wong, J. H. Y.; Brown, J. A.; Suo, Z.; Blum, P.; Nohmi, T.; Ling, H. Structural insight into

dynamic bypass of the major cisplatin-DNA adduct by Y-family polymerase Dpo4. Embo J 2010, 29, 2059-2069. 140.

Xu, C.; Maxwell, B. A.; Brown, J. A.; Zhang, L.; Suo, Z. Global conformational dynamics of

a Y-family DNA polymerase during catalysis. Plos Biol. 2009, 7, e1000225. 141.

Yin, Y. W.; Steitz, T. A. The structural mechanism of translocation and helicase activity in T7

RNA polymerase. Cell 2004, 116, 393-404. 142.

Sweasy, J. B.; Lauper, J. M.; Eckert, K. A. DNA polymerases and human diseases. Radiat. Res.

2006, 166, 693-714. 143.

Matthews, A. J.; Zheng, S.; DiMenna, L. J.; Chaudhuri, J. Regulation of immunoglobulin class-

switch recombination: choreography of noncoding transcription, targeted DNA deamination, and longrange DNA repair. Adv. Immunol. 2014, 122, 1-57. 144.

Roberts, S. A.; Gordenin, D. A. Hypermutation in human cancer genomes: footprints and

mechanisms. Nat. Rev. Cancer 2014, 14, 786-800. 145.

Goodman, M. F. Hydrogen bonding revisited: geometric selection as a principal determinant of

DNA replication fidelity. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 10493-10495.

ACS Paragon Plus Environment

47

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

146.

Page 48 of 57

Biertumpfel, C.; Zhao, Y.; Kondo, Y.; Ramon-Maiques, S.; Gregory, M.; Lee, J. Y.; Masutani,

C.; Lehmann, A. R.; Hanaoka, F.; Yang, W. Structure and mechanism of human DNA polymerase eta. Nature 2010, 465, 1044-1048. 147.

Su, Y.; Patra, A.; Harp, J. M.; Egli, M.; Guengerich, F. P. Roles of Residues Arg-61 and Gln-

38 of Human DNA Polymerase eta in Bypass of Deoxyguanosine and 7,8-Dihydro-8-oxo-2'deoxyguanosine. J. Biol. Chem. 2015, 290, 15921-15933. 148.

Katafuchi, A.; Sassa, A.; Niimi, N.; Gruz, P.; Fujimoto, H.; Masutani, C.; Hanaoka, F.; Ohta,

T.; Nohmi, T. Critical amino acids in human DNA polymerases eta and kappa involved in erroneous incorporation of oxidized nucleotides. Nucleic Acids Res. 2010, 38, 859-867. 149.

Ucisik, M. N.; Hammes-Schiffer, S. Comparative Molecular Dynamics Studies of Human DNA

Polymerase eta. J. Chem. Inf. Model. 2015, 55, 2672-2681. 150.

Liu, M. S.; Tsai, H. Y.; Liu, X. X.; Ho, M. C.; Wu, W. J.; Tsai, M. D. Structural Mechanism

for the Fidelity Modulation of DNA Polymerase lambda. J. Am. Chem. Soc. 2016, 138, 2389-2398. 151.

Genna, V.; Carloni, P.; De Vivo, M. A Strategically Located Arg/Lys Residue Promotes Correct

Base Paring During Nucleic Acid Biosynthesis in Polymerases. J. Am. Chem. Soc. 2018, 140, 3312-3321. 152.

Ensing, B.; De Vivo, M.; Liu, Z.; Moore, P.; Klein, M. L. Metadynamics as a Tool for Exploring

Free Energy Landscape of Chemical Reactions. Acc. Chem. Res. 2006, 39, 73-81. 153.

Nair, D. T.; Johnson, R. E.; Prakash, S.; Prakash, L.; Aggarwal, A. K. Replication by human

DNA polymerase-iota occurs by Hoogsteen base-pairing. Nature 2004, 430, 377-380. 154.

Chen, Y.; Zhang, J.; Liu, H.; Gao, Y.; Li, X.; Zheng, L.; Cui, R.; Yao, Q.; Rong, L.; Li, J.;

Huang, Z.; Ma, J.; Gan, J. Unique 5'-P recognition and basis for dG:dGTP misincorporation of ASFV DNA polymerase X. Plos Biol. 2017, 15, e1002599.

ACS Paragon Plus Environment

48

Page 49 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

155.

ACS Catalysis

Garcia-Escudero, R.; Garcia-Diaz, M.; Salas, M. L.; Blanco, L.; Salas, J. DNA polymerase X

of African swine fever virus: insertion fidelity on gapped DNA substrates and AP lyase activity support a role in base excision repair of viral DNA. J. Mol. Biol. 2003, 326, 1403-1412. 156.

Prasad, B. R.; Kamerlin, S. C. L.; Florian, J.; Warshel, A. Prechemistry barriers and checkpoints

do not contribute to fidelity and catalysis as long as they are not rate limiting. Theor. Chem. Acc. 2012, 131, 1-15. 157.

Mulholland, A. J; Roitbergm, A. E.; Tuñón, I. Enzyme dynamics and catalysis in the mechanism

of DNA polymerases. Theor. Chem. Acc. 2012, 131, 1-4. 158.

Pudney, C. R.; Johannissen, L. O.; Sutcliffe, M. J.; Hay, S.; Scrutton, N. S. Direct analysis of

donor-acceptor distance and relationship to isotope effects and the force constant for barrier compression in enzymatic H-tunneling reactions. J. Am. Chem. Soc. 2010, 132, 11329-11335. 159.

Pineda, J. R.; Antoniou, D.; Schwartz, S. D. Slow conformational motions that favor sub-

picosecond motions important for catalysis. J. Phys. Chem. B 2010, 114, 15985-15990. 160.

Hay, S.; Scrutton, N. S. Good vibrations in enzyme-catalysed reactions. Nat. Chem. 2012, 4, 161-

168. 161.

Olsson, M. H.; Parson, W. W.; Warshel, A. Dynamical contributions to enzyme catalysis: critical

tests of a popular hypothesis. Chem. Rev. 2006, 106, 1737-1756. 162.

Nashine, V. C.; Hammes-Schiffer, S.; Benkovic, S. J. Coupled motions in enzyme catalysis.

Curr. Opin. Chem. Biol. 2010, 14, 644-651. 163.

Garcia-Viloca, M.; Gao, J.; Karplus, M.; Truhlar, D. G. How enzymes work: analysis by modern

rate theory and computer simulations. Science 2004, 303, 186-195. 164.

Johnson, K. A. Conformational coupling in DNA polymerase information transfer. Philos. Trans.

R. Soc. Lond. B Biol. Sci. 1992, 336, 107-112.

ACS Paragon Plus Environment

49

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

165.

Page 50 of 57

Joyce, C. M.; Benkovic, S. J. DNA polymerase fidelity: kinetics, structure, and checkpoints.

Biochemistry 2004, 43, 14317-14324. 166.

Paschalis, V.; Le Chatelier, E.; Green, M.; Kepes, F.; Soultanas, P.; Janniere, L. Interactions

of the Bacillus subtilis DnaE polymerase with replisomal proteins modulate its activity and fidelity. Open Biol. 2017, 170146. 167.

Hoekstra, T. P.; Depken, M.; Lin, S. N.; Cabanas-Danes, J.; Gross, P.; Dame, R. T.; Peterman,

E. J. G.; Wuite, G. J. L. Switching between Exonucleolysis and Replication by T7 DNA Polymerase Ensures High Fidelity. Biophys J. 2017, 112, 575-583. 168.

Zhao, L. L.; Pence, M. G.; Eoff, R. L.; Yuan, S.; Fercu, C. A.; Guengerich, F. P. Elucidation

of kinetic mechanisms of human translesion DNA polymerase kappa using tryptophan mutants. Febs J. 2014, 281, 4394-4410. 169.

Zhang, H.; Guengerich, F. P. Effect of N2-guanyl modifications on early steps in catalysis of

polymerization by Sulfolobus solfataricus P2 DNA polymerase Dpo4 T239W. J. Mol. Biol. 2010, 395, 1007-1018. 170.

Baranovskiy, A. G.; Duong, V. N.; Babayeva, N. D.; Zhang, Y.; Pavlov, Y. I.; Anderson, K.

S.; Tahirov, T. H. Activity and fidelity of human DNA polymerase alpha depend on primer structure. J. Biol. Chem. 2018, 293, 6824-6843. 171.

Batra, V. K.; Perera, L.; Lin, P.; Shock, D. D.; Beard, W. A.; Pedersen, L. C.; Pedersen, L.

G.; Wilson, S. H. Amino acid substitution in the active site of DNA polymerase beta explains the energy barrier of the nucleotidyl transfer reaction. J. Am. Chem. Soc. 2013, 135, 8078-8088. 172.

Xiang, Y.; Oelschlaeger, P.; Florian, J.; Goodman, M. F.; Warshel, A. Simulating the effect of

DNA polymerase mutations on transition-state energetics and fidelity: evaluating amino acid group contribution and allosteric coupling for ionized residues in human pol beta. Biochemistry 2006, 45, 70367048.

ACS Paragon Plus Environment

50

Page 51 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

173.

ACS Catalysis

Dans, P. D.; Ivani, I.; Hospital, A.; Portella, G.; Gonzalez, C.; Orozco, M. How accurate are

accurate force-fields for B-DNA? Nucleic Acids Res. 2017, 45, 4217-4230. 174.

Klvaňa, M.; Bren, U.; Florián, J. Uniform Free-Energy Profiles of the P–O Bond Formation and

Cleavage Reactions Catalyzed by DNA Polymerases β and λ. J. Phys. Chem. B, 2016, 120, 13017–13030. 175.

Yang, W. Portraits of a Y-family DNA polymerase. FEBS Lett. 2005, 579, 868-872.

176.

Houlihan, G.; Arangundy-Franklin, S.; Holliger, P. Engineering and application of polymerases

for synthetic genetics. Curr. Opin. Biotechnol. 2017, 48, 168-179. 177.

Taylor, A. I.; Pinheiro, V. B.; Smola, M. J.; Morgunov, A. S.; Peak-Chew, S.; Cozens, C.;

Weeks, K. M.; Herdewijn, P.; Holliger, P. Catalysts from synthetic genetic polymers. Nature 2015, 518, 427-430. 178.

Pinheiro, V. B.; Taylor, A. I.; Cozens, C.; Abramov, M.; Renders, M.; Zhang, S.; Chaput, J.

C.; Wengel, J.; Peak-Chew, S. Y.; McLaughlin, S. H.; Herdewijn, P.; Holliger, P. Synthetic Genetic Polymers Capable of Heredity and Evolution. Science 2012, 336, 341-344. 179.

Srivastava, A. K.; Han, C.; Zhao, R.; Cui, T.; Dai, Y.; Mao, C.; Zhao, W.; Zhang, X.; Yu,

J.; Wang, Q. E. Enhanced expression of DNA polymerase eta contributes to cisplatin resistance of ovarian cancer stem cells. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 4411-4416.

180.

Patra, A.; Banerjee, S.; Johnson Salyard, T. L.; Malik, C. K.; Christov, P. P.; Rizzo, C. J.;

Stone, M. P.; Egli, M. Structural Basis for Error-Free Bypass of the 5-N-MethylformamidopyrimidinedG Lesion by Human DNA Polymerase eta and Sulfolobus solfataricus P2 Polymerase IV. J. Am. Chem. Soc. 2015, 137, 7011-7014.

ACS Paragon Plus Environment

51

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

181.

Page 52 of 57

Tomicic, M. T.; Aasland, D.; Naumann, S. C.; Meise, R.; Barckhausen, C.; Kaina, B.;

Christmann, M. Translesion polymerase eta is upregulated by cancer therapeutics and confers anticancer drug resistance. Cancer Res. 2014, 74, 5585-5596. 182.

Bergoglio, V.; Boyer, A. S.; Walsh, E.; Naim, V.; Legube, G.; Lee, M. Y.; Rey, L.; Rosselli,

F.; Cazaux, C.; Eckert, K. A.; Hoffmann, J. S. DNA synthesis by Pol eta promotes fragile site stability by preventing under-replicated DNA in mitosis. J. Cell Biol. 2013, 201, 395-408. 183.

Chou, K. M. DNA Polymerase Eta and Chemotherapeutic Agents. Antioxid. Redox Sign. 2011,

14, 2521-2529. 184.

Raugei, S.; Gervasio, F. L.; Carloni, P. DFT modeling of biological systems. Phys. Status Solidi

B 2006, 243, 2500-2515. 185.

Carloni, P.; Rothlisberger, U.; Parrinello, M. The role and perspective of ab initio molecular

dynamics in the study of biological systems. Acc. Chem. Res. 2002, 35, 455-464. 186.

Senn, H. M.; Thiel, W. QM/MM Methods for Biomolecular Systems. Angew. Chem. Int. Ed.

2009, 48, 1198-1229. 187.

Warshel, A. W., R. M. An empirical valence bond approach for comparing reactions in solutions

and in enzymes. J. Am. Chem. Soc. 1980, 102, 8. 188.

Warshel, A.; Weiss, R. M. Empirical valence bond calculations of enzyme catalysis. Ann. N. Y.

Acad. Sci. 1981, 367, 370-382. 189.

Kitaura, K.; Ikeo, E.; Asada, T.; Nakano, T.; Uebayasi, M. Fragment molecular orbital method:

an approximate computational method for large molecules. Chem. Phys. Lett. 1999, 313, 701-706. 190.

Kitaura, K.; Sin-Ichirou, S.; Nakano, T.; Komeiji, Y.; Uebayasi, M. Fragment molecular orbital

method: analytical energy gradients. Chem. Phys. Lett. 2001, 336, 163-170. 191.

Thiel, W. The MNDOC method, a correlated version of the MNDO model. J. Am. Chem. Soc.

1981, 103, 1413-1420.

ACS Paragon Plus Environment

52

Page 53 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

192.

ACS Catalysis

Thiel, W. MNDOC study of reactive intermediates and transition states. J. Am. Chem. Soc. 1981,

103, 1420-1425. 193.

Schweig, A.; Thiel, W. MNDOC study of excited states. J. Am. Chem. Soc. 1981, 103, 1425-

1431. 194.

Bennie, S. J.; van der Kamp, M. W.; Pennifold, R. C.; Stella, M.; Manby, F. R.; Mulholland,

A. J. A Projector-Embedding Approach for Multiscale Coupled-Cluster Calculations Applied to Citrate Synthase. J. Chem. Theory Comput. 2016, 12, 2689-2697. 195.

Hansen, N.; van Gunsteren, W. F. Practical Aspects of Free-Energy Calculations: A Review. J.

Chem. Theory Comput. 2014, 10, 2632-2647. 196.

Dal Peraro, M.; Ruggerone, P.; Raugei, S.; Gervasio, F. L.; Carloni, P. Investigating biological

systems using first principles Car-Parrinello molecular dynamics simulations. Curr. Opin. Struct. Biol. 2007, 17, 149-156. 197.

Kästner, J. Umbrella Sampling. WIREs Comput. Mol. Sci. 2011, 1, 932-942.

198.

Hu, H.; Yang, W. Free energies of chemical reactions in solution and in enzymes with ab initio

quantum mechanics/molecular mechanics methods. Annu. Rev. Phys. Chem. 2008, 59, 573-601. 199.

Dellago, C.; Bolhuis, P. G.; Csajka, F. S.; Chandler, D. Transition path sampling and the

calculation of rate constants. J. Chem. Phys. 1998, 108, 1964-1977. 200.

Hu, H.; Lu, Z.; Yang, W. QM/MM Minimum Free Energy Path: Methodology and Application

to Triosephosphate Isomerase. J. Chem. Theory Comput. 2007, 3, 390-406. 201.

van der Kamp, M. W.; Mulholland, A. J. Combined quantum mechanics/molecular mechanics

(QM/MM) methods in computational enzymology. Biochemistry 2013, 52, 2708-2728. 202.

Sousa, S. F.; Fernandes, P. A.; Ramos, M. J. Computational enzymatic catalysis - clarifying

enzymatic mechanisms with the help of computers. Phys. Chem. Chem. Phys. 2012, 14, 12431-12441.

ACS Paragon Plus Environment

53

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

203.

Page 54 of 57

Pande, V. S.; Beauchamp, K.; Bowman, G. R. Everything you wanted to know about Markov

State Models but were afraid to ask. Methods 2010, 52, 99-105. 204.

Kitchin, J. R. Machine learning in catalysis. Nature Catalysis 2018, 1, 230-232.

205.

Feldman, A. W.; Romesberg, F. E. Expansion of the Genetic Alphabet: A Chemist's Approach to

Synthetic Biology. Acc Chem Res 2018, 51, 394-403. 206.

Anosova, I.; Kowal, E. A.; Dunn, M. R.; Chaput, J. C.; Van Horn, W. D.; Egli, M. The structural

diversity of artificial genetic polymers. Nucleic Acids Res. 2016, 44, 1007-1021. 207.

Riccardi, L.; Genna, V.; De Vivo, M. Metal-ligand interactions in drug design. Nat. Rev. Chem.

2018, 2, 100-112. 208.

Arnold, F. H. Directed Evolution: Bringing New Chemistry to Life. Angew. Chem. Int. Ed. 2018,

57, 4143-4148. 209.

Yu, F.; Cangelosi, V. M.; Zastrow, M. L.; Tegoni, M.; Plegaria, J. S.; Tebo, A. G.; Mocny,

C. S.; Ruckthong, L.; Qayyum, H.; Pecoraro, V. L. Protein design: toward functional metalloenzymes. Chem. Rev. 2014, 114, 3495-3578. 210.

Muñoz-Robles, V.; Ortega-Carrasco, E.; Alonso-Cotchico, L.; Rodriguez-Guerra, J.; LLedós, A.;

Maréchal, J. D. Toward the Computational Design of Artificial Metalloenzymes: From Protein–Ligand Docking to Multiscale Approaches. ACS Catal. 2015, 5, 2469-2480. 211.

Ippolito, J. A.; Alexander, R. S.; Christianson, D. W. Hydrogen bond stereochemistry in protein

structure and function. J. Mol. Biol. 1990, 215, 457-471. 212.

Hoffmann, N. A.; Jakobi, A. J.; Moreno-Morcillo, M.; Glatt, S.; Kosinski, J.; Hagen, W. J.;

Sachse, C.; Muller, C. W. Molecular structures of unbound and transcribing RNA polymerase III. Nature 2015, 528, 231-236. 213.

Huai, Q.; Colandene, J. D.; Topal, M. D.; Ke, H. Structure of NaeI-DNA complex reveals dual-

mode DNA recognition and complete dimer rearrangement. Nat. Struct. Biol. 2001, 8, 665-669.

ACS Paragon Plus Environment

54

Page 55 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

214.

ACS Catalysis

Viadiu, H.; Aggarwal, A. K. The role of metals in catalysis by the restriction endonuclease

BamHI. Nat. Struct. Biol. 1998, 5, 910-916. 215.

Newman, M.; Lunnen, K.; Wilson, G.; Greci, J.; Schildkraut, I.; Phillips, S. E. Crystal structure

of restriction endonuclease BglI bound to its interrupted DNA recognition sequence. EMBO J. 1998, 17, 5466-5476. 216.

Nair, D. T.; Johnson, R. E.; Prakash, L.; Prakash, S.; Aggarwal, A. K. Human DNA polymerase

iota incorporates dCTP opposite template G via a G.C + Hoogsteen base pair. Structure 2005, 13, 15691577. 217.

Shen, B. W.; Xu, D.; Chan, S. H.; Zheng, Y.; Zhu, Z.; Xu, S. Y.; Stoddard, B. L.

Characterization and crystal structure of the type IIG restriction endonuclease RM.BpuSI. Nucleic Acids Res. 2011, 39, 8223-8236. 218.

Wah, D. A.; Hirsch, J. A.; Dorner, L. F.; Schildkraut, I.; Aggarwal, A. K. Structure of the

multimodular endonuclease FokI bound to DNA. Nature 1997, 388, 97-100. 219.

Jinek, M.; Jiang, F.; Taylor, D. W.; Sternberg, S. H.; Kaya, E.; Ma, E.; Anders, C.; Hauer,

M.; Zhou, K.; Lin, S.; Kaplan, M.; Iavarone, A. T.; Charpentier, E.; Nogales, E.; Doudna, J. A. Structures of Cas9 endonucleases reveal RNA-mediated conformational activation. Science 2014, 343, 1247997. 220.

Hirano, H.; Gootenberg, J. S.; Horii, T.; Abudayyeh, O. O.; Kimura, M.; Hsu, P. D.; Nakane,

T.; Ishitani, R.; Hatada, I.; Zhang, F.; Nishimasu, H.; Nureki, O. Structure and Engineering of Francisella novicida Cas9. Cell 2016, 164, 950-961. 221.

Nishimasu, H.; Cong, L.; Yan, W. X.; Ran, F. A.; Zetsche, B.; Li, Y.; Kurabayashi, A.;

Ishitani, R.; Zhang, F.; Nureki, O. Crystal Structure of Staphylococcus aureus Cas9. Cell 2015, 162, 1113-1126.

ACS Paragon Plus Environment

55

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

222.

Page 56 of 57

Zhang, J.; Pan, X.; Bell, C. E. Crystal structure of lambda exonuclease in complex with DNA

and Ca(2+). Biochemistry 2014, 53, 7415-7425. 223.

Batra, V. K.; Pedersen, L. C.; Beard, W. A.; Wilson, S. H.; Kashemirov, B. A.; Upton, T. G.;

Goodman, M. F.; McKenna, C. E. Halogenated beta,gamma-methylene- and ethylidene-dGTP-DNA ternary complexes with DNA polymerase beta: structural evidence for stereospecific binding of the fluoromethylene analogues. J. Am. Chem. Soc. 2010, 132, 7617-7625. 224.

Moon, A. F.; Pryor, J. M.; Ramsden, D. A.; Kunkel, T. A.; Bebenek, K.; Pedersen, L. C.

Structural accommodation of ribonucleotide incorporation by the DNA repair enzyme polymerase Mu. Nucleic Acids Res. 2017, 45, 9138-9148.

ACS Paragon Plus Environment

56

Page 57 of 57 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

TOC image

ACS Paragon Plus Environment

57