logo

Restriction Enzymes: A History

By Wil A.M. Loenen, Leiden University Medical Center
April 2019 · 346 pages, illustrated (38 color and 26 B&W)
ISBN 978-1-621821-05-2

<<  Chapter 7   —   Chapter 9  >>

Chapter 8

Chapter doi:10.1101/restrictionenzymes_8

Improved Detection Methods, Single-Molecule Studies, and Whole-Genome Analyses Result in Novel Insights on Structures, Functions, and Applications of Type I, II, III, and IV Restriction Enzymes: ∼2004–2016

INTRODUCTION

As mentioned in Chapter 7, with the arrival of whole-genome sequencing projects it has become clear that the Type I subclasses IA–IE and Type III R-M systems are common in bacteria and archaea (http://rebase.neb.com/rebase/rebase.html). The subdivision of Type II REases in 2003 into 11 subtypes based on behavior and cleavage properties (Roberts et al. 2003; Chapter 7) was helpful, but sometimes puzzling: Some REases would fit in more than one category or in none properly; unrelated proteins could be assigned to one or more of these subtypes; and some REases restricted DNA/RNA hybrids (a useful property to study small regulatory RNAs!) (Roberts et al. 2003; Murray et al. 2010; Loenen et al. 2014b). Although this subdivision remains useful, enzyme structures and/or domain organizations can be used as an alternative for classification (Niv et al. 2007; Pingoud et al. 2014). The notion that REases are evolutionarily related despite the lack of sequence similarity has grown more and more compelling, especially because of the increase in crystal structures with the PD···(D/E)XK fold (called the “PD fold”). As discussed in Chapter 7 (and depicted in Figs. 6 and 7 in that chapter), the structural studies by the group of Virginjuis Šikšnys provided formal evidence for a conserved sequence motif in the active site of REases, called the PD (D/E)XK site. The active site residues of EcoRI (G↓AATTC), NgoMIV (G↓CCGGC), and Cfr10I (R↓CCGGY) appeared to be common to 11 other REases belonging to Type IIP, IIE, and IIF (Kovall and Matthews 1999; Pingoud et al. 2002; Šikšnys et al. 2004). The almost simultaneous appearance of the structures of BamHI, PvuII, and EcoRV elicited much excitement (Winkler et al. 1993; Cheng et al. 1994; Newman et al. 1994), and both the papers on BamHI (Newman et al. 1994) and PvuII (Cheng et al. 1994) discussed the core structural motifs identified in the paper by Venclovas et al. (1994) when the structures of EcoRI and EcoRV were compared.

Early attempts to change specificity had not been very successful (Wolfes et al. 1986; Jeltsch et al. 1996; Lukacs et al. 2000; Pingoud et al. 2014). Substitutions usually resulted in a decrease in activity, but without exception failed to produce substantial changes in specificity. These findings led to the important lesson that recognition did not simply involve amino acids in direct contact with the bases and the backbone but also required water molecules and a complex network of other interactions (Pingoud et al. 2014). Sequence-specific DNA recognition by REases often involved binding to B-DNA in the major groove, with or without DNA distortion, similar to many regulatory proteins (see e.g., Rohs et al. 2010; Pingoud et al. 2014). In contrast, recognition by base flipping was used by enzymes that do chemistry: MTases, DNA repair enzymes (Roberts and Cheng 1998; Cheng and Roberts 2001), but also some REases (Bochtler et al. 2006; Horton et al. 2006; Tamulaitis et al. 2007; Szczepanowski et al. 2008; Miyazono et al. 2014; Manakova et al. 2015).

The new class of Type IV REases, defined in 2003 (Roberts et al. 2003), are modification-dependent enzymes that recognize modified Cs and As (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003). Early findings in the history of modified DNA date back quite a while, long before its importance became known. Modified DNA containing m5C was discovered in 1925 (Johnson and Coghill 1925), followed by m6A (Dunn and Smith 1955a,b), and hm5C in the 1950s (Hershey et al. 1953; Wyatt and Cohen 1953). The analysis of T* mutant phages in 1952 (Chapter 1; Luria and Human 1952) led to the discovery of enzymes that glycosylate hm5C (ghm5C) and of host genes that enable restriction of nonglucosylated phage DNA: rglA and rglB (restricts glucose-less DNA; Luria and Human 1952; Revel and Luria 1970). The rglA and rglB genes were renamed mcrA and mcrBC (modified cytosine restriction) (Noyer-Weidner et al. 1986; Raleigh and Wilson 1986) and were the first designated Type IV REases (Roberts et al. 2003).

Current interest in modified bases is high, as research into the dynamics of DNA modifications (“epigenetic phenomena”) have become of paramount importance for research into all kingdoms (Loenen and Raleigh 2014). Interestingly, hm5C was already discovered in eukaryotic (rat) brain and liver in 1972 (Penn et al. 1972) but did not receive much attention until the discovery of the role of Tet (ten-eleven translocation) proteins, a topic outside the scope of this book (Tet proteins are involved in m5C conversion and hence in control of normal and malignant cell differentiation) (see e.g., Veron and Peters 2011; Pastor et al. 2013; Baumann 2014; Stower 2014; Lu et al. 2015; Hendrickson and Cairns 2016; Jeschke et al. 2016). By 1980, some eight types of modified bases in phage DNA had been described (Warren 1980). A little later the m4C modification was found in Bacillus (Janulaitis et al. 1983), which is often present in thermophilic and mesophilic bacteria (Ehrlich et al. 1985, 1987). For an extensive description of the techniques used to detect and analyze such modified bases, see Weigele and Raleigh (2016), whose review discusses the initial harsh chemical treatments, physiological methods, paper chromatography, anion exchange columns, high-performance liquid chromatography (HPLC), mass spectrometry (MS), and SMRT. SMRT technology analyzes fluorescently labeled nucleotides that are incorporated slightly slower when encountering modified bases in the template strand than unmodified template bases during the sequencing procedure. This method thus allows the analysis of the “methylome” (i.e., the distribution of methylated bases in the DNA of different organisms).

This chapter uses the reviews that appeared in 2014 in Nucleic Acids Research as starting material, plus selected talks and posters presented at the 7th NEB meeting in Gdansk in 2015 (Loenen and Raleigh 2014; Loenen et al. 2014a,b; Mruk and Kobayashi 2014; Pingoud et al. 2014; Rao et al. 2014). Groups in Atlanta (Cheng), Bangalore (Rao and Nagaraja), Baltimore (Chandrasegaran), Berlin (Reuter and Kruger), Bristol (Halford and Szczelkun), Delft (Dekker), Edinburgh (Dryden), Gdansk (Mruk and Skowron), Giessen (Pingoud), Moscow (Zavil'gel'skii), New York (Aggarwal), Piscataway (Bogdanova), Pittsburgh (Jen-Jacobson), Portsmouth (Kneale), Seattle (Stoddard), Tokyo (Kobayashi), Tucson (Horton), Vilnius (Lubys and Šikšnys), and Warsaw (Piekarowicz, Bujnicki, and Bochtler) and at NEB (Roberts, Raleigh, Morgan, and Wilson) made important contributions to the field, as discussed throughout this chapter. Control (C) proteins of Type II enzymes were studied by the groups of Bob Blumenthal in Toledo and Geoff Kneale in Portsmouth. Data about the four types from approximately 2004 onward will be discussed, including structures of some Type II REases, as well as the very first structures of the other types, which together reveal many new, unexpected, and amazing details about the mechanisms employed to prevent indiscriminate restriction by the REase (subunit). Other types of control of restriction were elucidated, via transcription regulation, DNA mimics, C proteins, or the cognate MTase. R-M genes and lone MTase genes in pathogenic organisms also became of great interest because they are linked to virulence via “phase variation” (Piekarowicz 2013). Yet another breakthrough was the discovery of the family of the Type II REase MmeI, which would finally allow the generation of the new specificities so long hoped for. As in the previous chapters, this final chapter starts with the Type II REases (Part A), followed by the ATP-dependent Type I (Part B) and III (Part C) R-M systems, and the modification-dependent Type IV REases (Part D). The final section discusses the phenomenon of phase variation used by pathogenic bacteria to combat phage and evade host immunity (Part E).

PART A: TYPE II ENZYMES

Introduction

By 2014, approximately 4000 REases had been identified belonging to more than 350 different prototype Type II REases (i.e., biochemically different) (Roberts et al. 2010; Pingoud et al. 2014). The majority of these prototypes had characterized or putative relatives in sequenced genomes, resulting in more than 8000 publications (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2010; Pingoud et al. 2014). Most Type II REases shared little amino acid sequence similarity, with the exception of, for example, EcoRI and RsrI, an early example of “neutral drift”: EcoRI and RsrI (recognition site G↓AATTC) are identical in places with 50% overall identity (Aiken et al. 1986; Stephenson et al. 1989), allowing the construction of active hybrids (Chuluunbaatar et al. 2007). Also cases of mosaicism occur—for example, EcoRI, MunI (C↓AATTG), and MluCI (↓AATT) (Pingoud et al. 2014). In this large family, “compelling examples” (Pingoud et al. 2014) could be found of convergent (e.g., HaeIII [GG↓CC] and BsuRI [GG↓CC] [Wilson and Murray 1991]) and of divergent evolution (e.g., Bsu36I [CC↓TNAGG], BlpI [GC↓TNAGC], Bpu10I [CCTNAGC], and BbvCI [CCTCAGC] [Heiter et al. 2005]). In addition to the divalent cations Mg2+ and Mn2+ , discussed earlier, some Type II REases use Zn2+ (BslI [CCNNNNN↓NNGG], PacI [TTAAT↓TAA] [Vanamee et al. 2003; Shen et al. 2010; Horton 2015]) and Co2+, Ni2+, and Cu2+ (Pingoud et al. 2014). In the case of the well-known 8-bp cutter NotI (GC↓GGCCGC), the enzyme is dependent on Fe2+, but this Fe2+ is incorporated in a structural Cys4 cluster (Lambert et al. 2008; Pingoud et al. 2014). Despite the high specificity of all enzymes, star activity on noncognate sites does occur, which can be partially inhibited by, for example, spermidine, hydrostatic pressure, mutation, or lowering enzyme concentrations (Pingoud et al. 2014).

The number of crystal structures rose from 16 in 2004 (Chapter 7; summarized in Horton et al. 2004) to 35 by 2014 (Pingoud et al. 2014) and to more than 50 new “de novo” (i.e., the first structure of a particular enzyme) enzyme structures in 2017 (Horton 2015). Figure 1A shows the REase structures in the Protein Data Base (PDB) by February 2, 2018 (http://rebase.neb.com/cgi-bin/cryyearbar 2017). Figure 1B shows a graph displaying the difference between the total number of REase structures and the de novo structures in the PDB (due to follow-up structures with ligands and/or mutations of a particular enzyme that are also deposited in the PDB by April 2017 [Horton 2015]). Most of these enzymes carry the PD fold, but, in addition, structures of REases with PLD, GIY-YIG, HNH, and half-pipe folds have been elucidated (see page 191). The REases that were the first of their (sub)type (adapted from Horton 2015) are indicated in bold in Appendix 1, which lists selected REases studied from approximately 2004 to 2017, including some earlier references, where appropriate. Although these structures would greatly aid modeling studies, even for well-characterized REases the properties that determine specificity and selectivity remain difficult to predict, because the enzyme is fixed in the crystal and changes conformation during the catalysis, and the additional interactions involved in this “transition” state are not evident in the crystal structure (Lanio et al. 2000; Pingoud et al. 2014).

FIGURE 1 

FIGURE 1. (A) The REase structures in the Protein Data Base (PDB) by February 2, 2008 (http://rebase.neb.com/cgi-bin/cryyearbar 2017). (B) Comparison of the total number of REase structures deposited in the PDB with de novo REase structures. (Courtesy of Nancy Horton.)

 

The picture emerging from all these publications is that (similar to other protein families) the various domains involved in DNA binding, specific recognition, restriction, ATP binding and hydrolysis, and methylation have been fused or separated in all sorts of ways during the course of evolution. As a result, enzymes may have one or two catalytic sites and cleave DNA in one or two steps, with or without sliding and detaching from their DNA and with or without looping (Embleton et al. 2004; Halford and Marko 2004; Halford et al. 2004; Pingoud et al. 2014). Several studies addressed the question of the contribution of 1D and 3D movements of the REases along the DNA in order to find their recognition site (Gowers and Halford 2003; Gowers et al. 2005). Often multimers would bind to two sites rather than acquiring a second catalytic domain, which would be evolutionarily simple. One interesting study by the Bristol group concerns the reaction mechanism of seven REases that recognize GGCGCC and cut at different positions (Gowers et al. 2004). Using plasmids with one or two copies of this sequence revealed five distinct mechanisms, much larger than generally thought at the time (Gowers et al. 2004). Another example includes enzymes specific for the CCNGG sequences (Fig. 1 in Sasnauskas et al. 2015a, adapted in Fig. 3). Nearly 70% of all Type II REases belong to three families; the rest remain “mysteries”: They may be fringe members, or examples of new folds and DNA degradation mechanisms (see Pingoud et al. 2014 for further discussion).

Catalytic Domains of Type II REases

The PD···(D/E)XK Structural Fold

The PD···(D/E)XK fold (called the “PD” domain in this chapter) is present with variations in almost all Type II REases whose structures have been determined and is classified in the SCOP (Structural Classification of Proteins) database (http://scop.mrc-lmb.cam.ac.uk) as the REase-like fold (Niv et al. 2007; Steczkiewicz et al. 2012). The PD motif is often not easy to identify without information from a crystal structure, as the motif may vary and the amino acids involved are often in different locations along the polypeptide chain (Pingoud et al. 2014). Among 289 characterized Type II enzymes, 69% belonged to the PD superfamily (Orlowski and Bujnicki 2008) that includes the four nucleases mentioned in Chapter 7 (lambda exonuclease, MutH, VSR, and TnsA), but also, for example, RecB, Sulfolobus solfataricus Holliday-junction resolvase, and T7 endo I. The mechanism of catalysis continues to be the subject of study and debate. For example, the number of Mg2+ ions needed during catalysis remains uncertain (see Pingoud et al. 2014 for details and discussion).

The HNH and GIY-YIG Structural Domains

Other endonucleolytic motifs have been identified, including HNH and GIY-YIG motifs, found in homing endonucleases (HEases), Holliday-junction resolvases, exonucleases, nonspecific Serratia nuclease, and colicins (Friedhoff et al. 1999; Galburt et al. 1999; Jurica and Stoddard 1999; Pingoud et al. 2005a; Stoddard 2005; Kleinstiver et al. 2011, 2013). HNH examples are, for example, KpnI (GGTAC↓C) (Saravanan et al. 2004, 2007b; Vasu et al. 2013), Hpy99I (CGWCG↓), and PacI (TTAAT↓TAA), whereas two GIY-YIG REases, Eco29kI (CCGC↓GG) and Hpy188I (TCN↓GA), have been crystallized (Pertzev et al. 1997; Xu et al. 2000b; Bujnicki et al. 2001; Bujnicki 2004; Ibryashkina et al. 2007; Gasiunas et al. 2008; Kaminska et al. 2008; Orlowski and Bujnicki 2008; Mak et al. 2010; Mokrishcheva et al. 2011; Sokolowska et al. 2011). HNH motifs are often difficult to recognize because of the weak connection between the HNH and the residues that form the active site (Sokolowska et al. 2009). HNH enzymes use Mg2+ or Mn2+, but also other ions (Ni2+, Co2+, Zn2+, or Ca2+), sometimes with Cys4-Zn2+ binding elements (called ββα-metal fold), although many Cys4-Zn2+ motifs are not associated with catalytic sites but perform structural roles (Saravanan et al. 2004; Orlowski and Bujnicki 2008; Sokolowska et al. 2009; Shen et al. 2010; Pingoud et al. 2014).

Other Endonuclease Structural Domains

Thought unusual at the time, BfiI (ACTGGG [5/4]) was the first REase found that did not belong to the PD family: It carries the PLD nuclease domain and does not require Mg2+ for restriction (Sapranauskas et al. 2000). BfiI is a homodimer with a carboxy-terminal “B3-like” DNA-binding domain (DBD), which resembles B3 domains of some plant transcription factors. The catalytic site is formed at the interface of the two amino-terminal domains (similar to that of Nuc endonuclease from S. typhimurium), and although it binds to two sites at once, it cleaves only one strand at a time via an unusual covalent enzyme–DNA intermediate. BfiI appears to swivel the catalytic site by 180° and the same residues perform the same reaction on both DNA strands (Lagunavicius et al. 2003; Sasnauskas et al. 2003, 2007, 2010; Gražulis et al. 2005; Golovenko et al. 2014; Pingoud et al. 2014). Using the classification into 11 subtypes, this enzyme may be assigned to six or more of these subtypes (Marshall and Halford 2010; Pingoud et al. 2014). The enzyme AspCNI (GCCGC [9/5]) has a PLD-like domain and cleaves poorly at high concentrations (Heiter et al. 2015). PLD REases are not as rare as previously thought (Sapranauskas et al. 2000), as REBASE BLAST identified more than 40 other putatives (Pingoud et al. 2014). Some ATP-dependent enzymes (e.g., NgoAVII and CglI) contain a B3-like DNA recognition domain and a PLD catalytic domain (Tamulaitienė et al. 2014).

Type II REase Subtypes

This Type II section gives an overview and update with examples of the 11 subtypes, using two reviews (Roberts et al. 2003; Pingoud et al. 2014) as starting material. Figure 2 shows the subunit composition and cleavage mechanism of selected subtypes of Type II REases. Note that Pingoud et al. (2014) do not always follow the REBASE classification (http://rebase.neb.com/rebase/rebase.html). The reason for this is that different subtypes do not necessarily group with the different branches of the REase evolutionary tree, as exemplified by, for example, members of the EcoRII “CCGG family” studied by the Vilnius group (Table 1), which all cut at the same site (in contrast to the site studied by the Bristol group mentioned above [Gowers et al. 2004]): SsoII (↓CCNGG, Type IIP), EcoRII (↓CCWGG, Type IIE), and NgoMIV (G↓CCGGC, Type IIF) have similar DNA-binding sites and catalytic centers (Pingoud et al. 2002; Niv et al. 2007). Specificities for partly related, and even unrelated, sequences can nevertheless depend on the same structural framework: ↓CCNGG (SsoII), ↓CCWGG (PspGI/EcoRII), G↓CCGGC (NgoMIV), R↓CCGGY (Cfr10I), and MboI (↓GATC) (Pingoud et al. 2005c).

FIGURE 2 

FIGURE 2. Subunit composition and cleavage mechanism of selected subtypes of Type II REases. Type IIP enzymes act mainly as homodimers (top) and cleave both DNA strands at once. Some act as dimers of dimers (homotetramers) instead and do the same. Still others act as monomers (bottom) and cleave the DNA strands separately, one after the other. Bright triangles represent catalytic sites. Type IIS enzymes generally bind as monomers but cleave as “transient” homodimers. Type IIB enzymes cleave on both sides of their bipartite recognition sequences. Their subunit/domain stoichiometry and polypeptide chain continuity varies. Three examples of primary forms are shown: BcgI, AloI, and HaeIV. These forms assemble in higher-order oligomers for cleavage. Type IIB enzymes display bilateral symmetry with respect to their methylation and cleavage positions. It is not clear whether they cleave to the left or to the right of the half-sequence bound. Type IIG enzymes (e.g., BcgI) might cleave upstream (left) of their bound recognition half-site. All other Type IIG enzymes (e.g., MmeI) cleave downstream from the site, often with the same geometry. These proteins have very similar amino acid sequences, however, suggesting that somehow the reactions are the same. Type IIT enzymes cleave within or close to asymmetric sequences. Composition varies; they have two different catalytic sites: top-strand-specific and bottom-strand-specific. In some, both subunits/domains interact with the recognition sequence (left cartoons). In others, only the larger subunit/domain recognizes the DNA. (Reprinted from Pingoud et al. 2014.)

 

Table 1. The CCGG family studied by the Vilnius group of Virgis Šikšnys

Enzyme Recognition site Type Structure PDB IDa Reference(s)

Ecl18kI

↓CCNGG

IIF

Dimer/tetramer

2FQZ, 2GB7

Bochtler et al. 2006

EcoRII

↓CCWGG

IIE

Dimer

3HQF, 3HQG

Zhou et al. 2004; Golovenko et al. 2009

PspGI

↓CCWGG

IIP

Dimer

3BM3

Szczepanowski et al. 2008

PfoI

T↓CCNGGA

IIP

Dimer

Ms. in prep.

Manakova et al. 2015

Kpn2I

T↓CCGGA

IIP

Dimer

Ms. in prep.

Manakova et al. 2015

Cfr10I

R↓CCGGY

IIF

Tetramer

1CFR

Bozic et al. 1996

Bse634I

R↓CCGGY

IIF

Tetramer

3V1Z, 3V20, 3V21, 1KNV

Gražulis et al. 2002; Manakova et al. 2012

NgoMIV

G↓CCGGC

IIF

Tetramer

4ABT (cited in Manakova et al. 2012)

Deibert et al. 2000

BsaWI

W↓CCGGW

IIF

Dimer/tetramer/oligomer

4ZSF

Tamulaitis et al. 2015

AgeI

A↓CCGGT

IIP

Monomer/dimer

5DWA, 5DWB, 5DWC

Manakova et al. 2015; Tamulaitienė et al. 2017

SgrAI

CR↓CCGGYG

IIF

Dimer/tetramer/oligomer

4C3G cryoEM
3MQY, 3N78, 3N7B, 3MQ6, 3DVO, 3DW9, 3DPG

Lyumkis et al. 2013; Little et al. 2011; Park et al. 2010; Dunten et al. 2008

UbaLAI

CC↓WGG

IIE

Monomer

5O63

Sasnauskas et al. 2015, 2017

Updated by Elena Manak, Gintautas Tamulaitis, and Giedrius Sasnauskas (November, 2017).

a PDB ID is the identification number in the Protein Data Bank.

 

Type IIA

Type IIA enzymes usually have separate R and S domains, recognize asymmetric sequences, and cleave within or at a defined position in or close to this site (Roberts et al. 2003; Pingoud et al. 2014). Many have two MTases each modifying one strand of the recognition sequence, rather than a single MTase. Others are combined R-M enzymes, some with separate MTases. Kinetic studies indicate that Type IIA enzymes transiently dimerize for cooperative cleavage. Examples are BbvCI, which uses two different catalytic sites from different subunits (Bellamy et al. 2005; Heiter et al. 2005), and Mva1269I (GAATGC [1/−1], IIA/IIS), which uses two sites from different domains within the same protein (Armalyte et al. 2005).

Type IIB

Type IIB enzymes cleave on both sides of a bipartite site releasing ∼34 bp (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003; Marshall et al. 2007; Pingoud et al. 2014). Some enzymes comprise a large single RMS polypeptide with features in common with Type I enzymes. Sometimes SAM acts as the cofactor for R as well as for S. Most IIB enzymes can only restrict when bound to two sites, preferably in cis, or in trans on concatenates. The first IIB R-M system, BcgI (cloned in 1994 [Kong et al. 1994] and extensively studied by the Halford group), concertedly cleaves two double-strand bonds [(10/12) CGANNNNNNTGC (12/10)] (Kong et al. 1993; Marshall and Halford 2010; Sasnauskas et al. 2010; Marshall et al. 2011; Smith et al. 2013a,b; Pingoud et al. 2014). It can be considered IIB/G/H/S (Kong and Smith 1998; Jurenaite-Urbanaviciene et al. 2007; Marshall and Halford 2010; Marshall et al. 2011; Smith et al. 2013a,b; Pingoud et al. 2014), like some other Type IIB enzymes (http://rebase.neb.com/rebase/rebase.html). BcgI comprises two subunits, RM and S, for cleavage and methylation with a stoichiometry of (RM)2S1, comparable to the Type I pentamer R2M2S1, but cutting at fixed positions (Kong et al. 1994; Kong and Smith 1997; Kong 1998). Other enzymes include, for example, BaeI [(10/15) ACNNNNGTAYC (12/7)], BsaXI [(9/12) ACNNNNCTCC (10/7)], and NgoAVIII [(12/14) GACNNNNNTGA (13/11)] (Sears et al. 1996; Marshall and Halford 2010). The BcgI-like enzymes modify both strands of their recognition sequences without additional MTases, and cleavage requires multiple (RM)2S1 complexes for double-strand cleavage on both sides of the recognition site (Marshall et al. 2007, 2011). The exact mechanism requires further investigation (see Marshall and Halford 2010 for discussion). Other IIB enzymes are, for example, AloI ([7/12] GAACNNNNNNTC [12/7]), PpiI ([7/12] GAACNNNNNCTC [13/8]), CjeI ([8/14] CCANNNNNNGT [15/9]), and TstI ([8/13] CACNNNNNNTCC [12/7]) (Jurenaite-Urbanaviciene et al. 2007; Smith et al. 2014). Domain-swapping experiments suggest that, like Type I enzymes, TRD swapping may also be used to generate hybrid specificities of Type II enzymes (Jurenaite-Urbanaviciene et al. 2007). Domain swapping and circular permutation of subdomains of BsaXI ([9/12] ACNNNNNCTCC [10/7]), or deletion, resulted in either active protein with altered specificity, poor protein yields, or inactive enzymes, which allowed mapping of critical amino acids for the interaction between the RM subunit and the TRD of the S subunit (Xu et al. 2015).

Type IIC

Type IIC are combined RM enzymes (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003; Pingoud et al. 2014). Most IIC bind as monomers to continuous and asymmetric sequences and cleave on one side of the recognition site at 1 turn, 1½ turn, or 2 turns away, whereas others cleave on both sides (i.e., IIB). Cleavage via transient dimerization is likely and is more efficient on DNA with multiple recognition sites or on addition of oligonucleotides. Examples are Eco57I (CTGAAG [16/14]), MmeI (TCCRAC [20/18]), and BpuSI (also called RM·BpuSI, GGGAC [10/14]), which are also considered to be IIC as well as IIE or IIG, respectively (see pages 202–203).

Type IIE

The prototype Type IIE enzymes are EcoRII and NaeI with separate domains for cleavage and allosteric activation (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003; Pingoud et al. 2014), as discussed in Chapter 7 and, for example, Reuter et al. (2004). The Type IIE enzymes prove to be diverse in structure (Fig. 3). Figure 3 shows a comparison of the structures of NaeI (Fig. 3A), EcoRII (Fig. 3B), and a new Type IIE enzyme from an unknown bacterium, named UbaLAI (Sasnauskas et al. 2017) by the CCGG group in Vilnius (Fig. 4).

FIGURE 3 

FIGURE 3. Diversity of Type IIE REases. In all panels PD · · · (D/E)XK subunits are colored in different shades of green, monomeric MvaI-like PD · · · (D/E)XK domains are red, catabolite activator protein (CAP)-like domains are orange and light brown, and B3-like domains are blue. Yellow diamonds in the cartoon representations denote the catalytic center(s) present in each enzyme. (A) NaeI (GCC↓GGC) is a Type II homodimer that simultaneously binds two recognition sites. One is cleaved by the EcoRV-like dimer of the catalytic N domains (Endo domains), whereas the second one bound to the CAP DNA-binding motif in the carboxy-terminal domain (Topo domain) stimulates cleavage of the first site (PDB ID: 1IAW) (Embleton et al. 2001; Huai et al. 2001). (B) EcoRII (↓CCWGG) is a Type IIE homodimer capable of simultaneous binding of three recognition sites. One is cleaved by the PspGI-like dimer of the catalytic C domains, whereas two others, one per EcoRII-N effector domain, stimulate cleavage of the first site (PDB IDs: 3hqf and 3hqg) (Tamulaitis et al. 2006a,b; Golovenko et al. 2009). (C) UbaLAI (CC↓WGG) is a novel monomeric REase consisting of an MvaI-like catalytic domain (red) and an EcoRII-N-like effector domain (blue; PDB ID to be published). UbaLAI requires two recognition sites for optimal activity, and, like NaeI and EcoRII, uses one copy of a recognition site to stimulate cleavage of a second copy. UbaLAI-N acts as a handle that tethers the monomeric UbaLAI-C domain to the DNA, thereby helping UbaLAI-C to perform two sequential DNA nicking reactions on the second recognition site during a single DNA-binding event (Sasnauskas et al. 2017). The structure of the UbaLAI-C domain is a model built using Modeller (Webb and Sali 2016). (Portions reprinted from Sasnauskas et al. 2017, courtesy of Gintautas Tamulaitis.)

 

FIGURE 4 

FIGURE 4 

FIGURE 4. (A) CCGG group photo. From left to right: Inga Songailiene, Gintautas Tamulaitis, Elena Mankova, Saulius Gražulis, Virgis Šikšnys, Giedrė Tamulaitienė, Giedrius Sasnauskas, and Mindaugas Zaremba. (B) Graduate students and postdoctoral colleagues from 1977 through 2011 at Steve Halford’s retirement party (2011). His group, from left to right: Stuart Bellamy, Dave Scott, Rachel Smith, Kelly Sanders, Niall Gormley, Tim Nobbs, Mark Watson, Panos Soultanos, Geoff Baldwin, Steve Halford (in his DNA jumper), Darren Gowers, Mark Szczelkun, Neil Stanford, Jacqui Marshall, Barry Vipond, Yana Kovacheva, Katie Wood, Tony Maxwell, Isobel Kingston, John Taylor, Sophie Castell, Michelle Embleton, Christian Vermote, Alistair Jacklin, Alison Ackroyd, Fiona Preece, Susan Retter, Lucy Catto, Shelley Williams. Christian Parker and Denzil Bilcock were at the party but not in the photo. (Absent: Pete Luke, Paul Bennett, Samantha Hall, Lois Wenztell, Symon Erskine, Mark Oram, Abigail Bath, David Rusling, and Sumita Ganguly). (C) Werner Arber, Noreen Murray, and D.N. Rao at the 6th NEB meeting in Bremen (2010). (D) Participants of the CSHL meeting in 2013: History of Restriction Enzymes. (D, Courtesy Cold Spring Harbor Laboratory Archives.)

 

Nearly a dozen papers were published on EcoRII in collaborations between experts in the field of crystallography, AFM, and single-molecule studies (Zhou et al. 2002, 2003, 2004; Kruger and Reuter 2005; Tamulaitis et al. 2006a,b, 2008; Shlyakhtenko et al. 2007; Gilmore et al. 2009; Golovenko et al. 2009; Szczepek et al. 2009). A high-resolution crystal structure of the dimeric EcoRII was published in 2004, which revealed a hinge loop connecting the catalytic and allosteric activation domains (Zhou et al. 2002, 2003, 2004). The catalytic domain (comprised of two copies of the carboxy-terminal domain) had the PD fold, whereas the two amino-terminal regulatory/effector domains had a different DNA recognition fold with a large cleft. This fold was novel at the time, but is in fact the B3-like fold mentioned above and present in BfiI and NgoAVII (more specifically, it is a SCOP double-split β-barrel fold, of the DNA-binding pseudobarrel domain superfamily). The structure explained the mechanism of autoinhibition/activation of EcoRII, which was novel in REases, but similar to that described for various transcription factors (Zhou et al. 2004). This structure contained three possible DNA-binding regions, and in line with this, only a plasmid with three recognition sites yielded linear DNA during a single turnover, whereas the same plasmid with only one or two sites did not (Tamulaitis et al. 2006b). AFM studies showed two-loop structures with an EcoRII dimer at the core of the three-site synaptosome (Shlyakhtenko et al. 2007). A variant of AFM (called high-speed AFM) allowed single-molecule imaging of the EcoRII protein (Gilmore et al. 2009). In this way, binding, translocation, and dissociation could be monitored, and they indicated that EcoRII can translocate along the DNA to search for a second binding site, after finding the first site. Dissociation from the loop structure resulted in either two monomers bound to the two sites or one dimer to one site (Gilmore et al. 2009). Further experiments showed the very different ways in which the enzyme interacted with the effector and substrate DNA. The carboxy-terminal domain flipped the central T:A base pair out, and interacted with the CC:GG half-sites, whereas the effector domain bound asymmetrically without pushing out the T:A base pair (Golovenko et al. 2009). Interestingly, the 7-bp cutter PfoI (T↓CCNGGA) also uses base flipping as part of its DNA recognition mechanism. But in this case the extrahelical bases are captured in binding pockets that are quite different from those in the related structurally characterized enzymes Ecl18kI, PspGI, and EcoRII-C (Manakova et al. 2015). PspGI (↓CCWGG) and Ecl18kI/SsoII (↓CCNGG) flip the central A and T (W) bases out of the helix, compressing the recognition sequence in effect to just CC-GG (Bochtler et al. 2006; Tamulaitis et al. 2007; Szczepanowski et al. 2008). Repression of catalysis by the amino-terminal domain was further analyzed by site-directed mutagenesis and addition of soluble peptides in trans, which revealed the structural elements essential for autoinhibition (Szczepek et al. 2009). The crystal structure of MvaI identified MvaI as a monomer that recognizes its pseudosymmetric target sequence (CC↓WGG) asymmetrically (Kaus-Drobek et al. 2007). The enzyme has two lobes: a catalytic one that contacts the bases from the minor groove side, and the other that contacts those from the major groove. MvaI resembles BcnI (CC↓SGG), and also MutH, which nicks DNA rather than cutting both strands. The reason for this is clear: MvaI, BcnI, and MutH have a single catalytic site and just nick their substrates upon binding. Because the substrates of MvaI and BcnI are symmetric, these two enzymes can then bind in the opposite orientation and nick the other strand resulting in double-strand cleavage. The substrate of MutH (hemimethylated GATC) is asymmetric, and so MutH can only bind in one orientation and thus cannot cut the second strand. Different responses to slight substrate asymmetries, which could be altered by protein engineering, determine whether these monomeric REases make single-strand nicks or double-strand breaks (Sokolowska et al. 2007a; see Kaus-Drobek et al. 2007 for further details). For some other studies on the EcoRII and CCGG family, see Kubareva et al. (1992, 2000); Šikšnys et al. (2004); Pingoud et al. (2005b); Sud'ina et al. (2005); Zaremba et al. (2006); Fedotova et al. (2009); and Abrosimova et al. (2013), and the Type IIF section.

Type IIF

Type IIF bind two recognition sites and cleave all four strands at once as pairs of back-to-back dimers (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003; Šikšnys et al. 2004; Zaremba et al. 2005; Pingoud et al. 2014). The structures of Cfr10I (R↓CCGGY), Bse634I (R↓CCGGY), and NgoMIV (G↓CCGGY) (Chapter 7) and the observed transient tetramerization of Ecl18kI (↓CCNGG) indicated that the boundaries between IIE and IIF are not strict (Šikšnys et al. 2004; Zaremba et al. 2005, 2010; Pingoud et al. 2014). Work on Bse634I continued in Vilnius (Zaremba et al. 2005, 2006, 2012; Manakova et al. 2012). The tetramer could be converted to a dimeric enzyme by mutation, and kinetic studies indicated two types of communication signals via the dimer–dimer interface in the tetramer: an inhibitory and an activating signal, which somehow control the catalytic and regulatory properties of the Bse634I and mutant proteins (Zaremba et al. 2005, 2006). Contrary to expectation, dimeric enzymes have the same fidelity toward their recognition site as the tetramer, because they act concertedly at two sites, thus providing a safety catch against cleavage at a single unmodified site (Zaremba et al. 2012). The structures of the SfiI (GGCCNNNN↓NGGCC) tetramer in complex with cognate DNA provided details on how SfiI recognized and cleaved its target DNA sites (Viadiu et al. 2003; Vanamee et al. 2005). Some other Type IIF enzymes are PluTI (GGCGC↓C) (Khan et al. 2010; Pingoud et al. 2014) and SgrAI (CR↓CCGGYG). SgrAI (Laue et al. 1990; Tautz et al. 1990; Capoluongo et al. 2000; Bitinaite and Schildkraut 2002; Daniels et al. 2003; Hingorani-Varma and Bitinaite 2003; Wood et al. 2005; Dunten et al. 2008, 2009; Park et al. 2010; Little et al. 2011; Lyumkis et al. 2013; Ma et al. 2013b; Horton 2015) is also a member of the CCGG family and preferentially cleaves concertedly at two sites. Interestingly, SgrAI assembles into homotetramers, and then other molecules join to generate helical structures with one DNA-bound homodimer after another. Adjacent homodimers are not back-to-back (i.e., 180°), but at ∼90°, and four homodimers form almost one turn of a left-hand spiral of 18 homodimers or perhaps even more. These SgrAI filaments have some star activity, probably as a result of asymmetry generated by the multimerization process (Fig. 5). Another interesting enzyme is the Type IIF homotetrameric GIY-YIG Cfr42I enzyme that is rather similar to the monomeric/dimeric Eco29kI enzyme, which supports the notion of convergent evolution of REases belonging to unrelated nuclease families toward homotetramers with a “safety catch” (Gasiunas et al. 2008).

FIGURE 5 

FIGURE 5. CryoEM structure of SgrAI bound to DNA. Each SgrAI dimer is colored uniquely. This picture was made using PDB coordinates and surface rendering. (The original figure in the paper by Lyumkis et al. [2013] was the actual cryoEM envelope, carved up into different subunits.) (Adapted from Lyumkis et al. 2013, with permission from Elsevier.)

 

Type IIG

Type IIG are Type I-like combined RM systems, with an amino-terminal PD domain, and a γ-class MTase domain in a single protein (http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003; Niv et al. 2007; Loenen et al. 2014a; Pingoud et al. 2014). The S specificity subunit may be present as a separate subunit or as a domain attached to the carboxyl terminus of RM. IIG are stimulated by SAM or are SAM-dependent. This definition includes most IIB and IIC REases (Loenen et al. 2014a; Pingoud et al. 2014). Only one catalytic site is present in these domains, and cleavage of duplex DNA is thought to occur by the transient dimerization of neighboring enzyme molecules. Examples are Eco57I, MmeI, and BpuSI.

Eco57I was the first member of a new class of monomeric enzymes, initially called Type IV (like BspLU11III [GGGAC (10/14)] from Bacillus sp. LU11 [Lepikhov et al. 2001]), but renamed Type IIG enzymes (Janulaitis et al. 1992a,b), although it can also be considered IIE as it is accompanied by one additional MTase. It is a large RMS protein, cuts one and one-half turns away, and is useful for engineering (Janulaitis et al. 1992b; Rimseliene et al. 2003; Pingoud et al. 2014). It methylates the top strand of its asymmetric recognition site (CTGAAG [16/14]), whereas a separate MTase, M·Eco57I, methylates the adenine in the bottom strand (Janulaitis et al. 1992a). M·Eco57I can also methylate the same adenine in the top strand as Eco57I (Janulaitis et al. 1992a; Loenen et al. 2014a). Some other monomeric Type IIG have accompanying MTases that methylate m5C (BpuSI) or m4C (BseRI, GAGGAG [10/8]) (Loenen et al. 2014a).

MmeI is IIE/IIG/IIC and cuts two turns away (TCCRAC [20/18]). It was the first Type IIG enzyme to be purified and belongs to a large family of closely related enzymes with many different specificities (Boyd et al. 1986; Morgan et al. 2008; Loenen et al. 2014a). Based on in vitro studies, MmeI has also been named Type IIL, for lone-strand DNA modification (Morgan et al. 2009). As the enzyme does not require a head-to-head approach in vitro, there is disagreement on its mode of action: Does in vivo MmeI act on two inverted (head-to-head) recognition sequences like Type III enzymes (Dryden et al. 2011; Schwarz et al. 2011; Loenen et al. 2014a), using sliding or DNA looping between adjacent sites (Halford et al. 1999; Halford 2001), or perhaps bind DNA as a monomer and then form dimers or multimers before methylation or cleavage, similar to Type I enzymes (Loenen et al. 2014a)? But why would MmeI slide along the DNA, as the adenine that will eventually be methylated is likely to be flipped into the binding pocket on specific site recognition (Cooper et al. 2017; Bogdanove et al. 2018)? MmeI requires at least two bound specificity sites for cutting. Unlike FokI, adding excess enzyme in solution, without a specific site, does not stimulate cutting. Richard (Rick) Morgan suggests a model that includes the requirement for enzyme bound at two (or possibly four) sites to come together for cutting (Cooper et al. 2017; Bogdanove et al. 2018). As methylation is effective at single sites, this process does not require dimerization of the enzyme.

MmeI has been well characterized (Boyd et al. 1986; Tucholski et al. 1995, 1998; Nakonieczna et al. 2007, 2009; Morgan et al. 2008, 2009; Callahan et al. 2011, 2016), and rational engineering based on sequence alignments and mutational analysis led to altered specificities that could be predicted (Morgan and Luyten 2009; Morgan et al. 2009). Changes in the S domain alter the recognition site for both R and M (like Type I enzymes), and hence members of the MmeI family have been able to diverge widely in the course of evolution (Morgan et al. 2008, 2009; Morgan and Luyten 2009). Certain different pairs of amino acids are specific for alternative base pairs in the recognition sequence: for example, Glu806···Arg808 in MmeI (TCCRAC) specifies the third C, whereas Lys806···Asp808 specifies G at that position (TCCRAG). The crystal structure has been solved (Callahan et al. 2011, 2016). Together with the structure of MmeI in complex with DNA (and SAM-analog sinefungin) (Callahan et al. 2011), these data on the MmeI family allowed the construction of REases with novel predictable DNA recognition and restriction properties, which had “long been a goal of modern biology” (Callahan et al. 2016) and previously denied for EcoRI and EcoRV. With this in mind, Geoff Wilson pondered whether one could predict and design new specificities of other enzymes (e.g., Type I HsdS) or even predict those of putative HsdS subunits in REBASE based on sequence data alone (Loenen et al. 2014a). The answer to this is yes, as in recent times Rick Morgan has predicted and made specificity changes in Type I HsdS systems (R Morgan, in prep.).

BpuSI (GGGAC [10/14]) is IIG or IIS and has two MTases (Shen et al. 2011; Sarrade-Loucheur et al. 2013; Pingoud et al. 2014). The crystal structure indicates that it resembles the well-characterized carboxy-terminal cleavage domain of FokI (GGATG [9/13]) and produces 5′ sticky ends (Wah et al. 1997, 1998; Shen et al. 2011). This is unusual because most Type IIG enzymes create 3′ overhangs, indicating that their catalytic sites cleave across the minor groove of DNA rather than across the major groove. BpuSI was crystallized without DNA and evidently must undergo significant structural rearrangements to bind DNA and carry out catalysis (Shen et al. 2011). This means that the carboxy-terminal S domain must rotate with respect to the R and M domains and reorganize in order to bind DNA (also seen with other REases) (Shen et al. 2011; Sarrade-Loucheur et al. 2013; Pingoud et al. 2014).

Type IIH

Type IIH are hybrid Type IIP-like (e.g., GACNNN↓NNGTC) REases with an m6A MTase (Pingoud et al. 2014). M·AhdI is a tetramer of M and S subunits, suggestive of the ancestral form of Type I MTases. As such, they have been called “Type 1½” RM systems, and a “missing link” between Type I and II IIH enzymes, but as they have proved rather common, this distinction may no longer be relevant (Marks et al. 2003; Pingoud et al. 2014).

Type IIM

Type IIM enzymes recognize methylated DNA. The well-known DpnI (Lacks and Greenberg 1975; Pingoud et al. 2014) cuts Gm6A↓TC as a monomer, one strand at a time. The complementary specificities of DpnI and DpnII have been useful for site-directed mutagenesis, as DpnII cuts unmethylated ↓GATC) sites (Lacks and Greenberg 1977). DpnI has the amino-terminal PD domain and a carboxy-terminal winged-helix (wH) allosteric activator domain. Both domains bind methylated DNA with sequence specificity (Lacks and Greenberg 1975; Siwek et al. 2012; Mierzejewska et al. 2014; Pingoud et al. 2014). A new addition to this subtype is the BisI (Gm5CNGC) enzyme and its relatives (Xu et al. 2016). Some enzymes that recognize methylated DNA and are classified as Type IV enzymes would also fit into the IIM subtype, if they cut at specific sites (see the section Part D: Type IV Enzymes).

Type IIP

The best-known orthodox Type IIP palindromic REases are, of course, EcoRI and EcoRV. Type IIP cleave symmetric recognition sequences and have a single domain in which recognition and cleavage functions are integrated (Pingoud et al. 2014). They tend to have a single cognate MTase, although some have two MTases. The IIP REases can be monomeric but most are homodimers or homotetramers. Multimers usually cleave both DNA strands in one binding event, whereas monomers need to cleave sequentially first one strand, then the other, because of the opposite 5′ to 3′ polarity of the DNA strands (Gowers et al. 2004; Pingoud et al. 2014). In line with this prediction, the BcnI (CC↓SGG) monomer first localizes the recognition site by 1D and 3D diffusion, and nicks one DNA strand; it then diffuses from the nicked site, turns 180°, diffuses back, and cleaves the other (unnicked) strand (Sokolowska et al. 2007b; Kostiuk et al. 2011, 2015, 2017; Sasnauskas et al. 2011).

Single-molecule studies with EcoRV provided evidence for fast 3D sliding and jumping of EcoRV on nonspecific DNA following a slow initial 1D diffusion (Bonnet et al. 2008). Using optical tweezers with fluorescence tracking, it became clear that the enzyme stays in close contact with the DNA during sliding (Bonnet et al. 2008; Biebricher et al. 2009). Aneel Aggarwal and coworkers analyzed the structure of BstYI, a thermophilic REase that cleaves 5′-Pu/GATCPy-3′, a degenerate version of the BamHI (G↓GATCC) and BglII (A↓GATCT) recognition sites. A comparison of free BstYI with BamHI and BglII revealed a strong structural likeness between these enzymes, but in addition, BstYI also contained an extra “arm” domain possibly related to the thermostability of BstYI (Townson et al. 2004). The cocrystal structure with DNA revealed a mechanism of degenerate DNA recognition, which will stimulate thoughts about the possibilities and limitations in altering specificities of closely related REases (Townson et al. 2005). Interestingly, an isoschizomer of BamHI, OkrAI (G↓GATCC), is a much smaller version of BamHI, which recognizes the DNA in a similar manner, “a rare opportunity to compare two REases that work on exactly the same DNA substrate” (Vanamee et al. 2011).

The group of Ichizo Kobayashi studied regulation of the EcoRI operon (see page 214) (Liu and Kobayashi 2007; Liu et al. 2007), whereas the group of Linda Jen-Jacobsen in Pittsburgh continued studies on EcoRI with respect to the mechanism of coupling between DNA recognition specificity and catalysis (Kurpiewski et al. 2004), the inhibition by Cu2+ ions of Mg2+-catalyzed DNA cleavage (Ji et al. 2014), and the relaxed specificity and structure of promiscuous mutants of EcoRI that cleave at EcoRI* sites (Sapienza et al. 2005, 2007, 2014). As EcoRI* sites are not protected by M·EcoRI, promiscuous mutants are deleterious to the host. They encountered “unanticipated and counterintuitive observations” that three EcoRI mutants with such relaxed specificity in vivo nevertheless bound more tightly than wild-type EcoRI to the cognate site (GAATTC) in vitro and even preferred that site to EcoRI* sites (Sapienza et al. 2005). How could this be? Using structural and thermodynamic analyses, this question was addressed further (Sapienza et al. 2007, 2014). The crystal structure of the promiscuous mutant A138T homodimer in complex with the cognate site was nearly identical to that of the wild-type complex, except that the threonine138 side chains interacted with bases 5′ to the GAATTC site. This would enable A138T to form complexes with EcoRI* sites that structurally resembled the specific wild-type complex with GAATTC (Sapienza et al. 2007). The importance of these flanking bases was also confirmed by the finding that AAATTC sites with an adjacent 5′-purine-pyrimidine (5′-RY) were cleaved much faster (up to 170× faster!). This and further thermodynamic analyses supported the notion that specificity relied on a series of cooperative events that were “uniquely associated with specific recognition” (Sapienza et al. 2014).

SwaI (ATTT↓AAAT) (Dedkov and Degtyarev 1998) and PacI (TTAAT↓ TAA) both recognize AT-rich DNA sequences, but their protein structures are completely different (Shen et al. 2010). In the case of PacI, the normal base-pairing is completely disrupted in the bound structure: “two bases on each strand are unpaired, four are engaged in noncanonical A:A and T:T base pairs, and the remaining two bases are matched with new Watson–Crick partners.” This suggests that PacI is an unusual REase that recognizes its target site via contacts not visualized in the DNA-bound cocrystal structure (Shen et al. 2010). Whereas PacI is elongated and follows the track of the DNA helix (Shen et al. 2010), SwaI is flattened and horseshoe-shaped (Shen et al. 2015). SwaI has an open conformation with the DNA-binding surface accessible from the outside. When bound to DNA, the enzyme is closed and completely encircles the DNA. Like PacI, SwaI profoundly distorts the DNA on binding, but in a different way (Shen et al. 2010). In SwaI, the central T:A and A:T bases are unpaired, and the two adenines switch positions and stack on each other in the reverse order. This is accompanied by a ∼50° bend in the helix and severe compression of the major groove, much as is seen in EcoRV (GAT↓ATC) (Winkler et al. 1993). The authors had no idea “how this surprising reversal in base order takes place” (Shen et al. 2015, 2017). Like EcoP15I, which has been used to count Huntington's disease CAG repeats, TseI (G↓CWGC) is also useful for the analysis of A:A and T:T mismatches in CAG and CTG repeats in this dreadful disease (Moncke-Buchner et al. 2002; Ma et al. 2013a).

Type IIS

Type IIS cut at a fixed distance from the recognition site (http://rebase.neb.com/rebase/rebase.html; Szybalski et al. 1991; Roberts et al. 2003; Welsh et al. 2004; Niv et al. 2007; Pingoud et al. 2014). The recognition and cleavage domains are separated by a linker region allowing fusion of the cleavage domain to other recognition modules, thus generating novel specificities. They usually have two MTases, each methylating one of the two strands (m6A or m5C). The name Type IIS (for “shifted”) enzymes was first coined by Wacław Szybalski and coworkers at the University of Wisconsin, who devised “ingenious applications” in the 1980s (Hasan et al. 1986; Kim et al. 1988; Pingoud et al. 2014). All Type IIB, IIC, and IIG REases can be considered IIS (cut outside their recognition sites), but all share the integral γ-class MTase as described above.

FokI (GGATG [9/13]) is one of the earliest, most-studied, Type IIS enzymes with a DNA recognition domain and a separate cleavage domain, which has been used extensively for genome engineering (Sugisaki and Kanazawa 1981; Nwankwo and Wilson 1987; Mandecki and Bolling 1988; Kaczorowski et al. 1989; Kita et al. 1989a,b; Landry et al. 1989; Looney et al. 1989; Sugisaki et al. 1989; Goszczynski and McGhee 1991; Szybalski et al. 1991; Li et al. 1993; Skowron et al. 1993; Kim et al. 1994; Waugh and Sauer 1994; Yonezawa and Sugiura 1994; Kim et al. 1996a, 1997, 1998; Skowron et al. 1996; Hirsch et al. 1997; Wah et al. 1997, 1998; Bitinaite et al. 1998; Leismann et al. 1998; Chandrasegaran and Smith 1999; Friedrich et al. 2000; Vanamee et al. 2001; Catto et al. 2006; Laurens et al. 2012; Pernstich and Halford 2012; Rusling et al. 2012; Guilinger et al. 2014b; Mino et al. 2014; Pingoud et al. 2014). The accompanying two MTases are fused into a single protein. Single-particle EM studies provided new insights into the activation mechanism of FokI and avoidance of aspecific cleavage (Vanamee et al. 2007). FokI crystals show the catalytic domain to be hidden behind the DNA recognition domain, which will require a substantial conformational change before cutting can take place after dimerization of two catalytic domains (Bitinaite et al. 1998; Pingoud et al. 2014). Details on the cleavage mechanism still need to be sorted out, but the need for two enzyme molecules for catalysis appears to be quite common (Embleton et al. 2001; Welsh et al. 2004; Catto et al. 2006, 2008; Gemmen et al. 2006; Sanders et al. 2009; Pingoud et al. 2014). Other enzymes also have a FokI-like domain—for example, StsI (GGATG [10/14]) (Kita et al. 1992a,b) and Mva1269I (GAATGC [1/−1]) (Armalyte et al. 2005).

Type IIT

Type IIT enzymes are heterodimers with two subunits (e.g., Bpu10I, BbvCI) or heterotetramers (e.g., BslI [CCNNNNN↓NNGG] [http://rebase.neb.com/rebase/rebase.html; Roberts et al. 2003; Pingoud et al. 2014]). IIT use two different catalytic sites for cleavage. Some enzymes are single chain (e.g., Mva1269I uses an EcoRI-like domain and a FokI-like domain) (Armalyte et al. 2005; Pingoud et al. 2014). Type IIT systems usually have two MTases (either separate proteins or fused as a single protein) that each modify one strand. They are useful after conversion to strand-specific nicking enzymes (see Chan et al. 2011 for a review, and page 209)—for example, BbvCI has two catalytic sites from different subunits, each cleaving its own strand (Bellamy et al. 2005; Heiter et al. 2005).

The “Half-Pipe”

PabI of Pyrococcus abyssi (GTA↓C) was thought to be a bona fide REase, as it was found near a MTase gene (Pingoud et al. 2014). However, it is a homodimeric DNA glycosylase with a unique structure and flips all four purines out of the helix, leaving the pyrimidines as intrahelical “orphans” (Ishikawa et al. 2005; Watanabe et al. 2006; Miyazono et al. 2007, 2014; Pingoud et al. 2014; Kojima and Kobayashi 2015). Close isoschizomers of PabI are ubiquitous in Helicobacter pylori strains. Whether PabI is involved in genetic rearrangements remains to be investigated (Pingoud et al. 2014).

Type II Enzymes as Tools for Gene Targeting

Fusions

As briefly discussed in Chapter 7, Srinivasan Chandrasegaran at Johns Hopkins School of Medicine pioneered what is now termed gene targeting by fusing the REase endonuclease domain of FokI to a zinc-finger protein to create a novel engineered zinc-finger nuclease (ZFN) (Li and Chandrasegaran 1993). ZFNs usually contain three to six Zn fingers (each ∼30 aa) with a ββα fold that binds one Zn2+ via 2Cys + 2His (Miller et al. 1985; Klug 2010a,b). Each finger recognizes a 3-bp target sequence via four amino acids that project from the α-helix into the major DNA groove (Durai et al. 2005; Wu et al. 2007). Two different three-finger ZFNs will recognize an 18-bp sequence, sufficient to be unique in the human genome. Such constructs have been used with considerable success, although they tend to be less specific than expected (Urnov et al. 2005, 2010; Carroll 2011a,b; Gabriel et al. 2011; Handel and Cathomen 2011; Pattanayak et al. 2011; Perez-Pinera et al. 2012a; see also Carroll 2014; Carroll and Beumer 2014; Hendel et al. 2015).

ZNF-based engineered highly specific REases can be used for gene targeting by introducing a dsDNA break into a complex genome and thereby stimulating homologous recombination (Yanik et al. 2013; Carroll 2014). With the exception of engineered homing endonucleases (“meganucleases”) with integrated DBD and catalytic domains (Galetto et al. 2009), the other engineered nucleases have distinct DBD and catalytic domains. ZFNs usually have the cleavage domain of FokI (Li et al. 1992, 1993; Li and Chandrasegaran 1993; Waugh and Sauer 1993; Kim et al. 1994, 1996b; Chandrasegaran and Smith 1999; Bibikova et al. 2002; Urnov et al. 2005; Miller et al. 2007; Szczepek et al. 2007; Mino et al. 2009; Mori et al. 2009; Carroll 2011a,b; Gabriel et al. 2011; Handel and Cathomen 2011; Pattanayak et al. 2011; Ramalingam et al. 2011, 2013; Handel et al. 2012; Bhakta et al. 2013; Pingoud et al. 2014), but PvuII (CAG↓CTG) has also been used for this purpose (Schierling et al. 2012).

Nonspecific (“off-target”) cleavage can be reduced by mutations in the dimerization surface (Miller et al. 2007; Szczepek et al. 2007), but according to Steve Halford the off-target problem might well be due to dimerization between a specific and a nonspecific ZFN (Halford et al. 2011).

The fusion construct with PvuII (CAG↓CTG) was slightly better than that with FokI (Schierling et al. 2012; Pingoud et al. 2014), but fusions with the transcription activator-like effector (TALE) proteins, where one module recognizes one base (Fig. 6A,B; Pingoud et al. 2014), were an improvement on engineered nuclease contructs. These proteins contain many (up to 35) nearly identical repeats of ∼34 aa. The 13th residue in each repeat recognizes the DNA base. The repeats form a superhelix around the DNA, following the track of the major groove for several turns. The individual repeats are left-handed two-helix bundles that, one after the other, juxtapose the 13th amino acid of each repeat to adjacent bases in one strand of the DNA (Deng et al. 2012; Mak et al. 2012, 2013).

FIGURE 6 

FIGURE 6. Engineering of Type II REases as tools for gene targeting. (A) Engineered highly specific endonucleases that can be used for gene targeting by introducing a double-strand break into a complex genome and thereby stimulating homologous recombination (Yanik et al. 2013). With the exception of engineered homing endonucleases (“meganucleases”) in which the function of DNA binding and DNA cleavage is present in the same polypeptide chain (Galetto et al. 2009), the other engineered nucleases consist of separate DNA-binding (green) and DNA-cleavage (blue) modules. ZFNs and TALENs usually have the nonspecific cleavage domain of the restriction endonuclease FokI as DNA-cleavage module, but the restriction endonuclease PvuII can also be used for this purpose (Schierling et al. 2012; Yanik et al. 2013). PvuII has also been employed in triple-helix-forming oligonucleotide (TFO)-linked nucleases (Eisenschmidt et al. 2005) and in protein fusions (with catalytically inactive I-SceI) (Fonfara et al. 2012) as DNA-cleavage module. ZFNs, TALENs, and TFO-linked nucleases are programmable, as are the RNA-mediated nucleases (Jinek et al. 2012) modified after Pingoud and Wende (2011). (Reprinted from Yanik et al. 2013.) (B) TALE-PvuII fusion proteins. (a) Scheme of the architecture of TALE–PvuII fusion proteins. (Left) wtPvuII, a homodimer in which the DNA-binding module of a TALE protein is fused via a linker of defined length. (Right) scPvuII, a monomeric nuclease in which the DNA-binding module of a TALE protein is fused via a linker of defined length. (b) Model of a TALE–wtPvuII fusion protein. The fusion protein is a dimer of identical subunits, each composed of a PvuII subunit and a TALE protein. This model was constructed by aligning the structures of the individual proteins PDB 1pvi (Cheng et al. 1994) and PDB 3ugm (Mak et al. 2012) on a DNA composed of the PvuII recognition site and two TALE target sites upstream of and downstream from the PvuII recognition site, separated by 6 bp. The carboxyl termini of the PvuII subunits and the amino termini of the TALE protein are separated by ∼3 nm. This distance must be covered by a peptide linker of suitable length. The image was generated with PyMol. (Reprinted from Yanik et al. 2013.)

 

In the case of PvuII, the DBD of a TALE protein is fused via a linker of defined length to the homodimeric REase (Fig. 6B). Wild-type PvuII (wtPvuII) is shown on the left in Figure 6Ba, and a variant of PvuII as a TALE-linked monomer (scPvuII) on the right in Figure 6Ba. A model of a TALE-PvuII fusion protein was constructed using the structures PDB 1pvi (Cheng et al. 1994) and PDB 3ugm (Mak et al. 2012) on a DNA composed of the PvuII recognition site and two TALE target sites upstream of and downstream from the PvuII recognition site, separated by 6 bp (Fig. 6Bb). The fusion protein is a dimer of identical subunits, each composed of a PvuII subunit and a TALE protein.

TALE-based nucleases (engineered TALE nucleases [TALENs]), based on FokI and PvuII, proved much better tools for genome manipulations than did ZFNs (Miller et al. 2011; Perez-Pinera et al. 2012b; Joung and Sander 2013; Yanik et al. 2013), but they also have some off-target activity. Profiling of 30 different unique TALENs for the ability of potential off-target cleavage using in vitro selection and high-throughput sequencing resulted in 76 predicted off-target substrates in the human genome, 16 of which were accessible and modified by TALENs in human cells (Guilinger et al. 2014a). This analysis allowed the construction of a TALEN variant with ∼10× lower off-target activity in human cells (Guilinger et al. 2014a).

In 2014, FokI was fused to Cas9, which cleaves dsDNA at a sequence programmed by a short single-stranded guide RNA (Guilinger et al. 2014b). Unfortunately, genome editing by Cas9 can also result in off-target DNA recognition. Fusions of catalytically inactive Cas9 and FokI nuclease (fCas9) modified target DNA sites with >140-fold higher specificity than wild-type Cas9 and with an efficiency similar to that of paired Cas9 “nickases” that cleave only one DNA strand each. The specificity of fCas9 was at least fourfold higher than that of paired nickases and may be a good strategy for highly specific genome-wide editing (Guilinger et al. 2014b). Use of very long (up to 10 kb) homologous flanking arms for break repair also improves targeting (Baker et al. 2017).

Nickases (Nicking Enzymes)

Another approach to gene targeting has been the use of nickases. Precise incisions in genomic DNA are required for (faithful) homologous recombination, but dsDNA breaks would activate the error-prone, nonhomologous end-joining (NHEJ) pathway. This led to the idea that a nicking domain that would cut only one DNA strand might work better than a cleavage domain, and could be used for DNA repair studies and other DNA manipulations (e.g., terminal labeling, genome mapping, and DNA amplification) (Chan et al. 2011; Xiao et al. 2011). The large subunits of some heterodimeric REases (e.g., some Type IIT and IIS) can function as nicking enzymes when separated from their normal partner (Higgins et al. 2001; Heiter et al. 2005; Yunusova et al. 2006; Xu et al. 2007), whereas dimeric enzymes can be mutated to generate a single catalytic site (Stahl et al. 1996; Wende et al. 1996; Morgan et al. 2000; Simoncsits et al. 2001; Heiter et al. 2005). Examples are BbvCI (Heiter et al. 2005), BspD6I (GACTC [4/6]) (Kachalova et al. 2008), BsrDI (GCAATG [2/0]) (Xu et al. 2007), Mva1269I (Armalyte et al. 2005), and BtsCI (GGATG [2/0]) (Too et al. 2010). Such nickases have been used in fusions with zinc fingers, TALE proteins, and methyl CpG binding domains (for further details, see Boch et al. 2009; Moscou and Bogdanove 2009; Hockemeyer et al. 2011; Gabsalilow et al. 2013; Mussolino et al. 2014; Pingoud et al. 2014; Ramalingam et al. 2014; Thanisch et al. 2014; Dreyer et al. 2015; Rogers et al. 2015).

Control of Restriction of Type II Enzymes

Control by C Proteins

Expression of the MTase gene and methylation of the host DNA before synthesis of the REase is essential after entry of a Type II system into the cell. In 1992, the Blumenthal laboratory provided the first evidence for temporal control in a subset of R-M systems, the plasmid-based PvuII system of Proteus vulgaris (Tao et al. 1991; Tao and Blumenthal 1992), soon followed by that in the BamHI system (Ives et al. 1992; Sohail et al. 1995).

A small C gene upstream of, and partially overlapping with, the REase gene is coexpressed from pres1, located within the MTase gene, at low level with the REase after entry of the self-transmissible PvuII plasmid into a new host, whereas the MTase gene is expressed at normal levels from its own two promoters pmod1 and pmod2 located within the C gene (Fig. 7).

FIGURE 7 

FIGURE 7. Intricate control of restriction in the operons of the Type II R-M systems of PvuII and Esp1396I by controlling C proteins (Loenen et al. 2014b). A small C gene upstream of, and partially overlapping with, R is coexpressed from pres1, located within the M gene, at low level with R after entry of the self-transmissible PvuII plasmid into a new host, whereas M is expressed at normal levels from its own two promoters pmod1 and pmod2 located within the C gene. A similar C protein operates in Esp1396I, but in this case the genes are convergently transcribed with transcription terminator structures in between, and M is expressed from a promoter under negative control of operator OR, when engaged by the C protein in a manner similar to that of the PvuII system. C proteins keep both R and M under control and have been tentatively identified in more than 300 R-M systems. See the text for further details. (Reprinted from Loenen et al. 2014b.)

 

The C protein binds to two palindromic DNA sequences (C boxes) upstream of the C and REase genes: OL, associated with activation, and OR, associated with repression. Low basal expression from the pvuIIC promoter leads to accumulation of the activator, which enhances transcription of the C and REase genes (Tao et al. 1991; Tao and Blumenthal 1992; Bart et al. 1999; Knowle et al. 2005; Williams et al. 2013). After this initial low-level expression of C·PvuII protein from the weak promoter pres1, positive feedback by high-affinity binding of a C protein dimer to the distal OL site later stimulates expression from the second promoter pres, resulting in a leaderless transcript and more C and R protein. The proximal site OR is a much weaker binding site, but C protein bound at OL enhances the affinity of OR for C protein, and at high levels of C protein, the protein–OR complex down-regulates expression of C and R. In this way, C protein is both an activator and negative regulator of its own transcription.

The regulation is similar to gene control in phage lambda: Differential binding affinities for the promoters in turn depend on differential DNA sequence and dual symmetry recognition. C proteins belong to the helix-turn-helix family of transcriptional regulators that include the cI and cro repressor proteins of lambdoid phages. In the wake of PvuII and BamHI, other R-M systems were discovered that were controlled by C proteins, including BglII (A↓GATCT) (Anton et al. 1997), Eco72I (CAC↓GTG) (Rimseliene et al. 1995), EcoRV (Zheleznaya et al. 2003), Esp1396I (CCANNNN↓NTGG) (Cesnaviciene et al. 2003; Bogdanova et al. 2009), SmaI (CCC↓GGG) (Heidmann et al. 1989), and AhdI (McGeehan et al. 2005). In the case of Esp1396I, the genes are convergently transcribed with transcription terminator structures in between, and the MTase gene is expressed from a promoter under negative control of operator OR, when engaged by C protein in a manner similar to that of the PvuII system (Fig. 7). C·Esp1396I controls OR, OL, and OM in a similar manner as described above. In this way, C proteins keep both R and M under control. This delay of REase expression depends on the rate of C-protein accumulation, and this may help explain the ability of C-regulated R-M systems to spread widely (Williams et al. 2013). By September 2013, REBASE listed 19 characterized C proteins, as well as 432 putatives (http://rebase.neb.com/rebase/rebase.html). The organization of the genes in the system and regulatory details differ from system to system, and some C proteins are fused to their REase genes (http://rebase.neb.com/rebase/rebase.html; Tao et al. 1991; Tao and Blumenthal 1992; Semenova et al. 2005; Bogdanova et al. 2009; Liang and Blumenthal 2013). Whether R-M systems as a whole evolved in concert with C proteins remains to be investigated.

The first structures of C proteins without DNA appeared in 2005: C·AhdI from Geoff Kneale's laboratory (McGeehan et al. 2005) and C·BclI from Ganesaratinam (Bali) K. Balendiran's laboratory in collaboration with NEB (Sawaya et al. 2005). These structures resembled those of helix-turn-helix DNA-binding proteins, as expected. The details of the interactions between C proteins and their C boxes in the DNA came later with the studies on the AhdI operon, and the structures of C·AhdI, C·Esp1396I, and C·BclI (Marks et al. 2003; McGeehan et al. 2004, 2005, 2006; Streeter et al. 2004; Sawaya et al. 2005; Callow et al. 2007; Papapanagiotou et al. 2007; Bogdanova et al. 2008; Ball et al. 2009). With the structure and further experiments, the mechanism behind the genetic switch could be elucidated (McGeehan et al. 2008, 2012; Ball et al. 2009, 2012). C·Esp1396I bound as a tetramer, with two dimers bound adjacently on the 35-bp operator sequence OL + OR (McGeehan et al. 2008). This cooperative binding of dimers to the DNA operator controls the switch from activation to repression of the C and R genes. The existence of C proteins explained the difficulty to introduce some R-M genes into E. coli (e.g., BamHI [Brooks et al. 1989]; see Loenen et al. 2014b for further details). The C genes belong to different incompatibility groups, which exclude unrelated R-M systems (called “apoptotic mutual exclusion”): For example, the pvuIIC and bamHIC genes define one exclusion group and prevent entry of ecoRVC due to premature activation of the EcoRV REase gene (Nakayama and Kobayashi 1998).

In 2016, the group of Iwona Mruk in Gdansk reported an unexpected regulatory variation on the above theme (Rezulak et al. 2016). The C·Csp231I gene regulates expression of the REase gene like other C-regulated R-M systems, but there is additional novel control. Separate tandem promoters drive most transcription of the Csp231I REase gene, a distinctive property not seen in other tested C-linked R-M systems. Further, the C protein only partially controls REase expression, yet plays a role in viability of the cells within the population by affecting stability and propagation. Deletion of the C gene led to high REase activity and resulted in loss of these cells in mixed cultures with wild-type R-M cells.

Transcriptional Control: The Case of EcoRI

The transcriptional control discussed above via C proteins has been found for many Type II enzymes, but not all Type II enzymes have such multiple (convergent) promoters and controlling C proteins. A prime example is EcoRI, whose enzymatic activity is controlled in a different way until methylation is complete. The Tokyo group of Ichizo Kobayashi investigated the intricate control of the EcoRI gene, ecoRIR (Liu and Kobayashi 2007; Liu et al. 2007; Mruk et al. 2011). This gene is upstream of the modification gene, ecoRIM. The M gene can be transcribed from two promoters within ecoRIR, allowing expression of the MTase gene with and without ecoRIR, as there is no transcription terminator between the two genes. In addition, the ecoRIR gene has two reverse promoters. These convergent promoters negatively affect each other, as in lambda (Ward and Murray 1979). Transcription from the reverse promoter is terminated by the forward promoters and generates a small antisense RNA. The presence of the antisense RNA gene in trans reduced lethality mediated by cleavage of undermethylated chromosomes after loss of the EcoRI plasmid (postsegregational killing) (Heitman et al. 1989; Mruk et al. 2011). This can be viewed as programmed cell death in prokaryotes. Kobayashi compares R-M systems with toxin/antitoxin (TA) systems composed of an intracellular toxin (the REase) and an antitoxin (the MTase) that neutralizes its effect. These systems would limit the genetic flux between lineages with different sequence-specific DNA methylation (“epigenetic identity”) but would require intricate control of restriction activity (reviewed in Mruk and Kobayashi 2014).

Control by the Cognate MTase

M·Ecl18kI and M·SsoII are two MTases that act as transcription factors and activate expression of their respective REase genes via binding to the regulatory site in the promoter region of these genes (Karyagina et al. 1997; Shilov et al. 1998; Fedotova et al. 2009). The amino-terminal region of M·Ecl18kI performs the regulatory function, but is also important for methylation activity. Loss of methylation activity per se does not prevent the MTase from performing its regulatory function and even increases its affinity to the regulatory site. However, the presence of the methylation domain is necessary for M·Ecl18kI to perform its regulatory function (Burenina et al. 2013).

PART B: TYPE I ENZYMES

Type I Families and Diversity

As discussed in Chapter 7, a single Type I common ancestor is likely, given the high sequence similarity of confirmed (biochemically analyzed) and putative enzymes and irrespective of the host within subclasses up to 80%–99%, between subclasses ∼20%–35%. This section is based on two reviews published in 2014 (Loenen et al. 2014a,b), and the reader is referred to these for more details and references. By 2013, ∼50% (1140/2145) of sequenced bacterial and archaeal genomes in REBASE carried one or more hsdR, hsdM, and hsdS genes and 40% appeared to have none, whereas the remainder had some but not all three genes or disrupted or scrambled genes (http://rebase.neb.com/rebase/rebase.html). On average, cells had two systems, although, for example, Desulfococcus oleovorans has eight systems. Type I enzymes could undergo specificity changes via TRD exchanges by homologous recombination, unequal crossing-over, or transposition (Chapter 6, Fig. 6). Domain shuffling is not limited to Type IA enzymes but can be observed between members of the same or different families (Loenen et al. 2014a). Within Type I families, HsdS subunits have the same organization, but between families they have different amino and carboxyl termini (circular permutations; see Loenen et al. 2014a for details). Circular permutation of HsdS of EcoAI indicated structural, but not necessarily functional equivalence, as different permutations resulted in an active R-M system, active in methylation only, or inactive, indicating that the HsdS termini interact with the HsdM and HsdR subunits (Janscak and Bickle 1998). Some bacteria have only one or two hsdR and hsdM genes but many hsdS genes (up to 22 in Mycoplasma sp.!) allowing multiple specificity changes providing protection against invaders (Sitaraman and Dybvig 1997; Dybvig et al. 1998; Loenen et al. 2014a). Shuffling of those 22 hsdS genes could easily result in more than 500 new specificities, “a defensive repertoire reminiscent of the immunoglobulins of higher organisms” (Loenen et al. 2014a). The advent of SMRT sequencing, which allows the localization of methylated bases, has led to a breakthrough in the determination of Type I recognition sites (Eid et al. 2009; Flusberg et al. 2010; Korlach et al. 2010; Clark et al. 2012; Korlach and Turner 2012), which may generate renewed interest in these “sophisticated molecular machines” (Murray 2000). SMRT sequencing not only led to an exponential increase in the number of known Type I recognition specificities (rising from approximately 40 biochemically characterized specificities in 2011 to more than 1100 by 2017) but also the discovery of Type I enzymes that produce m6A on one strand and m4C on the other (Morgan et al. 2016).

Single-Molecule Studies of EcoKI and EcoR124I

AFM and single-molecule studies, together with improved biochemical and biophysical methods, revealed new details about translocation by EcoKI via the motor domains that belong to superfamily 2 (SF2) (Neaves et al. 2009). Mutational analysis of the DEAD-box, RecA-domain-like, motifs of EcoR124I showed long-range effects of various mutations—for example, nuclease mutants could lower translocation and ATP usage rate, there could be a decrease in the off rate, and/or there could be slower restart and turnover (Sisakova et al. 2008a,b). Dimerization appeared to occur preferentially on two-site DNA, whereas DNA looping could occur in the absence of ATP hydrolysis. Would this be a common way to bring distant DNA regions together? Would this mean that SF2-dependent enzyme complexes in higher organisms also use such looping (which are involved in DNA repair, replication, recombination, chromosome remodeling, and RNA metabolism; for discussion, see, e.g., Tuteja and Tuteja 2004; Singleton et al. 2007; Fairman-Williams et al. 2010; Ramanathan and Agarwal 2011; Umate et al. 2011)?

Single-molecule studies using magnetic tweezers were designed to analyze single translocating molecules of EcoR124I in real time (Seidel et al. 2004, 2005, 2008; Stanley et al. 2006; Seidel and Dekker 2007). These experiments provided details on the rate of DNA translocation, as well as the processivity and ATP dependence of the HsdR motors. New facts emerged that may be of consequence for the studies on the eukaryotic SF2-dependent complexes mentioned above: (1) The two motors could work independently and the enzyme tracked along the helical pitch of the DNA on torsionally constrained molecules; (2) translocation could stop and restart by disassembly and reassembly, and (3) the HsdR subunits released the DNA roughly every 500 bp during this process (whereas the MTase remained attached to the recognition site); in other words, about four times over a distance of 2 kb. Concomitantly with this stop and restart process, the enzymes consumed vast amounts of ATP. A translocation block by collision with another HsdR or the presence of supercoiled DNA resulted in cleavage. The enzyme remained at the site but could be displaced by other proteins (e.g., E. coli RecBCD) (Bianco and Hurley 2005).

Type I Enzyme Atomic Structure

In the absence of crystals, the DNA recognition complex of EcoKI, the trimeric M·EcoKI (M2S1), had been modeled on 3D structures of other MTases (Chapter 7, Fig. 10). This model suggested a common origin of Type I and Type II MTases. Would this ancestral MTase combine with one or two HsdR molecules allowing translocation to the site of cleavage? Did translocation involve contacts with nonspecific DNA adjacent to the recognition site in a cleft in HsdR, which would close and reopen using ATP, Mg2+, and probably SAM, to fuel and control the conformational changes? Other questions remained to be answered: Would HsdR touch one strand or both strands of the dsDNA via backbone contacts, and what about step size, or the amount of DNA transported per physical step? And why no cleavage during the initial translocation? Was the translocation rate too high, or the catalytic PD region in the wrong conformation to contact the DNA? One thing seemed certain: Two REases were needed for dsDNA breaks—one for each strand. The answer to some of these questions came when finally the first structures of Type I enzymes appeared.

The Structure of the M·EcoKI (M2S1) Complex

The first crystal structures of HsdS subunits appeared in 2005 and 2010 (Calisto et al. 2005; Kim et al. 2005). The two TRDs are in inverted orientations, which makes the S subunit functionally symmetric (Fig. 8A, bottom right; see Loenen et al. 2014a for discussion). Each TRD consists of a globular DBD and an α-helical dimerization domain. The long α-helices (D1 and D2) encoded by the two conserved regions of the hsdS gene associate to form an antiparallel coiled-coil dimerization helix between the two variable HsdS specificity domains (S1 and S2) (Calisto et al. 2005; Kim et al. 2005). Amino acid side chains down their lengths interlock “like tines of a zipper” and form a hydrophobic core that holds the two helices together and separates the globular specificity domains by a fixed distance. S1 and S2 each recognize one-half of the recognition sequence. Each TRD also associates with one HsdM subunit to form an M2S trimer. Neither HsdS nor HsdM subunits bind to DNA alone, but the EcoKI trimer methylates both strands of the recognition sequence—the “top” strand of the 5′ half-sequence (Am6AC) and the “bottom” strand of the 3′ half-sequence (CGm6AC) of the bipartite AACNNNNNNGTCG—thus protecting the host DNA during replication (Fig. 8A).

FIGURE 8 

FIGURE 8 

FIGURE 8. (A) Model of the M·EcoKI MTase (PDB ID: 2Y7H). The S subunit is composed of two TRDs in inverted orientations. Each TRD comprises a globular DBD and an α-helical dimerization domain. The N-TRD (green) and C-TRD (orange) are specific for the two halves of the recognition sequence (AACNNNNNNGCAG). Zipper-like association of the helices separates the globular domains by a fixed distance and reverses the orientation of the C-TRD. Each TRD also associates with one M subunit (identical, but shown here in different shades of blue for clarity) to form an M2S trimer, which methylates both strands and protects the resident DNA during DNA replication. (B) Structure of the Type IA HsdS protein S-ORF132P (PDB ID: 1YF2). Structure of the Type IA HsdS protein S-MjaXI (PDB ID: 1YF2). The upper diagram shows the domain organization of the protein; arrows represent DBDs, and curly lines represent dimerization α-helices. The amino acid sequence of the protein is shown below, with the domains in corresponding colors. Below this are three views of the structure, from three perpendicular directions: “sideways,” “end-on,” and “above.” The panels on the left depict the protein; those on the right depict the protein with modeled DNA positioned approximately as it is bound. The DNA was taken from PDB ID: 2Y7H and transferred by structural alignment of the S subunits. (A,B, Reprinted from Loenen et al. 2014a.)

 

The structure and sequence of the S subunit from Methanocaldococcus jannaschii, S·MiaXI, is shown in the top part of Figure 8B. The recognition sequence of this protein is unknown. It is closely related to the Type IA family of EcoKI. Below this are three views of the structure from three perpendicular directions. The panels on the left show the protein on its own, and those on the right a model of the protein bound to DNA.

The structure of the trimeric M·EcoKI (M2S1) was resolved in 2009, thanks to the product of the 0.3 gene product of phage T7, T7 Ocr (Chapter 6), which proved to be a DNA mimic (see section Antagonists of Type I Action: Antirestriction, starting on page 230). Ocr was used to stabilize the otherwise labile MTase complex (Kennaway et al. 2009). Many different single M·EcoKI-Ocr complexes were imaged in the EM, allowing a reconstruction of the 3D complex to a resolution of 18 Å. A model of M·EcoKI is shown in Figure 8A, which depicts the location of the specificity domains of the S subunit in relation to the two M subunits (see Kennaway et al. 2009 for further details).

The Structure of the EcoKI and EcoR124I (R2M2S1) Complexes

The first data on the crystal structure of EcoR124I HsdR were published by Lapkouski et al. (Lapkouski et al. 2007, 2009). This suggested how the pentamer might be assembled and how the motors might translocate dsDNA (Lapkouski et al. 2009). The PD motif was found opposite the translocation domain, which would allow coupling of translocation to restriction. A model was proposed (Lapkouski et al. 2009) based on this HsdR structure, a DNA path across the subunit, and an early, incomplete model of the MTase core (Obarska et al. 2006). This model has much in common with the later model by Kennaway et al. (Kennaway et al. 2012) but differs in the orientation of the HsdR with respect to the MTase core and the path taken by the DNA (see below). Soon afterward, crystal data appeared on the amino-terminal fragment of a putative Type I enzyme from Vibrio sp. This contained three globular domains with an endonuclease core and the ATPase site close to the probable DNA-binding site for translocation. The authors suggested the involvement of a linker helix in the transition from DNA motor protein to nuclease (Uyen et al. 2008, 2009).

In 2012, years of efforts by David Dryden and coworkers finally paid off and the structure of the pentameric EcoKI and EcoR124I (R2M2S1) was elucidated by computer-assisted EM single-particle reconstructions (Kennaway et al. 2012). Single-particle analysis of negative stain EM images showed large differences between DNA-bound (Fig. 9A) and unbound EcoR124I enzymes (Fig. 9B), with their longest dimensions being ∼18 nm versus ∼22–26 nm, respectively. (The smaller particles were the R1M2S1 form and were analyzed separately as described later.) EcoR124I with DNA was in a closed conformation, whereas the enzyme alone was in an open form without DNA. Apparent twofold symmetry was visible in many image averages. Using these data, a 3D reconstruction was generated (Fig. 9A) of EcoR124I bound to a 30-bp dsDNA fragment with the unmethylated recognition site and the enzyme on its own (Fig. 9B; see Kennaway et al. 2012 for details).

FIGURE 8 

FIGURE 9. Gallery of Type I RM structures and conformations determined by EM and single-particle analysis. (A) EcoR124I + DNA (closed state) negative stain EM. (B) EcoR124I without DNA (open state) negative stain EM. (C) EcoKI + DNA negative stain EM. For each 3 × 3 panel, the top rows are image averages, the middle rows are their corresponding reprojections, and the bottom rows are 3D surface views of the 3D reconstruction (bars, 200 Å); on the right is a larger 3D surface perspective view. See Kennaway et al. 2012 for further details. (Reprinted from Kennaway et al. 2012, with permission from Cold Spring Harbor Laboratory Press.)

 

EcoR124I without DNA was highly extended and more flexible, which allowed a low-resolution (∼3.5-nm) 3D reconstruction of the enzyme (Fig. 9B). Most particles (∼80%) appeared to have their twofold axis roughly normal to the plane of the carbon film, but ∼5% were seen to be folded up into the closed state, indicating a dynamic equilibrium between states in the absence of cognate DNA. Some very thin connections between the domains were “likely pivot points for flexing to allow the enzyme to close up” (Kennaway et al. 2012). These data showed that the subunits strongly moved in a manner to allow entry of the DNA substrate (Fig. 9A,B). This change from an extended structure to a more compact form in the presence of DNA had been seen previously for the MTase (Kennaway et al. 2009; see Kennaway et al. 2012 for discussion).

Negatively stained particles of EcoKI with DNA bound (Fig. 9C) appeared smaller than EcoR124I with DNA (∼16 nm long) and appeared to be more rounded and variable than EcoR124I. The 3D reconstruction of EcoKI with a 75-bp fragment of dsDNA indicated a compact structure with many features similar to EcoR124I with DNA, including recognizable density for the five subunits in a matching arrangement, suggesting a common architecture for Type I enzymes. However, the EcoKI particles were compact and appeared to be identical with and without DNA, and not elongated, as seen for EcoR124I without DNA. The dynamic equilibrium between open and closed forms apparently favored the closed form for EcoKI under the conditions used for EM.

Scattering experiments were used to construct a model showing the location of HsdR and the MTase (Kennaway et al. 2012). In the elongated structure of EcoR124I, the two HsdR subunits were located toward the extreme ends on either side of the MTase core. Fortunately, a fraction of the enzyme existed as a tetramer (R1M2S1). In negative stain EM, the particles of EcoR124I with (Fig. 10A) or without (Fig. 10B) DNA were not 100% homogeneous, and further analysis showed a large missing region at the extremity of the smaller particles, which had to be the location of one of the HsdR subunits. The existence of tetrameric complexes with only one HsdR was consistent with previous biochemical data on EcoR124I (Janscak et al. 1998).

FIGURE 10 

FIGURE 10. 2D difference images from EM data show the position of the HsdR in the EcoR124I complex. (A) Difference imaging between image averages of large (left) and small (right) particles in the EcoR124I + DNA negative stain EM data set reveals a large “negative density” region (red contour at −2.5σ), consistent with a missing HsdR in the small particles. (B) Difference imaging of HsdR in the open state of EcoR124I (without DNA). Although the relative flexibility of the open complex gives rise to a less well-defined difference map, a region of negative density consistent with a missing HsdR is visible nevertheless (red contour). The deduced atomic structure of each EM particle is shown below the EM image. (Adapted from Kennaway et al. 2012, with permission from Cold Spring Harbor Laboratory Press.)

 

Although the DNA is not visible in these experiments, the authors were able to use T7 Ocr, which binds very tightly to the DNA-binding site of Type I enzymes (Atanasiu et al. 2002; Walkinshaw et al. 2002). EcoR124I-Ocr complexes adopted a closed conformation and further analysis revealed the position of a banana-shaped object running through the center of the enzyme but tilted at an angle relative to the long axis of the 3D map. This banana-like shape matched well with the structure of the Ocr protein (Walkinshaw et al. 2002). This orientation of Ocr in the EM map plus the structural models of the MTase core of EcoKI (Kennaway et al. 2009) and of EcoR124I (Kennaway et al. 2012) allowed only one possible orientation of the MTase with the dimerization helix of the S subunit exposed to the solvent. This was in agreement with previous observations that this region could accommodate small (Gubler and Bickle 1991) and large (Kannan et al. 1989) amino acid insertions, and even a fusion with green fluorescent protein (Chen et al. 2010), without loss of function. Moreover, limited proteolysis indicated preferential cleavage within the dimerization helix (Webb et al. 1995), and hence surface exposure of this region.

In these studies the carboxyl terminus of the R subunit of EcoR124I (aa residues 893–1038) was not visible, but could be modeled using known crystal structures (Kennaway et al. 2012). Using these crystal structures, the EM data, and scattering analyses, atomic models of complete R subunits for EcoKI and EcoR124I were constructed, backed up by the plethora of published biochemical data on these enzymes.

The authors proposed a model that fit the data and gave the location and directionality of the DNA motor domains by aligning these with those of the dsDNA-bound SWI2/SNF2 chromatin remodeling translocase from S. solfataricus (Lapkouski et al. 2009). The direction of DNA translocation of this translocase was known and imposed a similar directionality on each HsdR, and because these had to pull DNA in toward the MTase core of the Type I enzyme, the orientation of each HsdR relative to the MTase core became defined. Based on DNA footprinting experiments (Mernagh et al. 1998; Powell et al. 1998) and the known minimum length of 45 bp of DNA required for ATP hydrolysis (Roberts et al. 2011), the assumption was made that the DNA path between the DNA bound to the HsdR and the DNA bound to the core MTase could not be longer than ∼40 bp. This meant that the motor domains of the HsdR had to have their DNA-binding sites close to the DNA-binding site of the MTase core. Placement of the HsdR on either side of the MTase and interacting directly with DNA was further supported by the length of the structure of another DNA mimic protein, ArdA (Nekrasov et al. 2007; McMahon et al. 2009), which occupies the entire DNA-binding site on Type I enzymes. This then allowed the complete structures for the closed forms of EcoR124I and EcoKI to be constructed as shown in Figure 11.

FIGURE 11 

FIGURE 11 

FIGURE 11. Atomic models of EcoR124I + DNA, EcoR124I, and EcoKI + DNA docked into the EM map densities. (A) Two views of the EcoR124I + DNA model showing the MTase core closed around DNA (green; DNA bound to each HsdR is not shown for clarity). Adenine bases are flipped out into the active sites of each of the two HsdM (light and dark blue), induced by an ∼45° bend in the DNA. (Yellow) HsdS, (red) HsdR, with the β-sheets of the RecA-like motor domains (orange). (Gray) Residues missing from the crystal structures (the 44 and 152 carboxy-terminal residues of HsdM and HsdR, respectively) were modeled de novo. The carboxy-terminal regions of HsdM extend down to bind at the coiled coil of HsdS, and the HsdR carboxy-terminal domains fill some empty density next to the amino terminus of HsdM. (B) A model for EcoKI bound to DNA (colors as in A). HsdS and HsdM from the MTase structure (PDB ID: 2y2C) were docked in as a single rigid body; HsdR modeled on those from EcoR124I (PDB ID: 2w00) (see Supplemental Material of Kennaway et al. 2012) and placed in a position analogous to the EcoR124I model. (C) The model of EcoR124I in the open conformation (i.e., without DNA; colors as in A). Although the EM map is at a lower resolution, a full atomic model could be built, aided by the EcoR124I + DNA model, SANS data, and 2D difference imaging. HsdM and HsdR swing out as a unit away from HsdS. The predicted hinge regions in the carboxyl termini of the HsdM (gray) and their connections to HsdS are not well resolved. (Reprinted from Kennaway et al. 2012, with permission from Cold Spring Harbor Laboratory Press.)

 

Placement of the R subunit of EcoR124I forced two large kinks in the DNA to allow the DNA to thread through the MTase core (Fig. 11A). This kinked path shortens the through-space end-to-end distance of a duplex bound to the enzyme by ∼10 nm, in line with AFM measurements of complexes of EcoR124I on DNA that showed that binding of the enzyme shortened the length of a long linear DNA molecule by ∼11 nm (van Noort et al. 2004). AFM measurements of EcoKI bound to DNA also showed a pronounced kink (Walkinshaw et al. 2002; Neaves et al. 2009), and circular dichroism analysis of EcoR124I also indicated a large structural distortion to the DNA when bound (Taylor et al. 1994).

A fit of subunits into the EcoKI EM density map (Fig. 11B) corresponded closely to that of EcoR124I in the closed state. The thin protrusions at either side of the EM envelope for EcoKI could fit the long coiled-coil amino-terminal extensions of unknown function predicted in the R subunit of EcoKI but absent in EcoR124I (Fig. 11B). Significant sequence differences existed between the two enzymes, and this might account for other structural differences, although the overall architecture remained unchanged.

An optimal fit of subunits into the lower-resolution open EcoR124I map (Fig. 11C) was obtained by moving and rotating each HsdM–HsdR pair as a single rigid body away from HsdS. A relatively simple ∼90° rotation and an ∼80° twist around a pivot point near the carboxyl terminus of HsdM were sufficient to move between open and closed states. It had previously been shown that HsdR and HsdM can form a complex (Dryden et al. 1997), supporting movement of the two subunits as a rigid body. The carboxy-terminal residues of EcoKI HsdM were disordered in the crystal (PDB ID: 2ar0) (Kennaway et al. 2009) and were sensitive to proteolysis (Cooper and Dryden 1994) and could play the role of the flexible linker proposed. Proteolytic removal of this region inhibited assembly of the pentamer (Powell et al. 2003).

Taken together, the data showed how Type I enzymes assemble, bind, and distort DNA before the initiation of ATP-driven DNA translocation. Although these EM and small-angle scattering structures were of low resolution, the proposed atomic models are in agreement with the extensive data from biochemical, biophysical, and genetic studies (Murray 2000, 2002; Loenen 2003; Tock and Dryden 2005). These provided several further constraints on the subunit orientations and gave confidence in the atomic models shown in Figure 11. These make it clear that there is an equilibrium between open and closed forms of the Type I enzymes, with the equilibrium constant depending on the particular enzyme and the presence or absence of DNA (and presumably the cofactors SAM and ATP). EcoKI prefers to be closed whether DNA is present or not and must therefore transiently open up to allow DNA access to the MTase core. EcoR124I appears to prefer an open form in the absence of DNA but is closed with DNA bound.

It would appear possible for the Type I enzymes to reach the closed “initiation” complex with the S-shaped DNA path via different routes. The model shown in Figure 12A is based on the EcoR124I structure but David Dryden (pers. comm.) states that “it would also apply to other Type I enzymes if we could actually ever see the open conformation—which we did not for EcoKI but this does not mean that it does not exist transiently.” The open form can bind DNA nonspecifically using HsdR (left side of Fig. 12A) and diffuse along the DNA until the MTase core recognizes a target sequence or dissociates. The trigger for closing and formation of the initiation complex would be most likely the recognition of the target sequence by the MTase core. Alternatively, the closed form of the enzyme must open up transiently to allow DNA to enter the MTase core, followed by closing of the core around the DNA (right side of Fig. 12A) and diffusion of the enzyme on the DNA until it either recognizes its DNA target sequence or reopens and dissociates. Starting the process of target sequence location and recognition via this pathway means that the motor domains of the HsdR will have to rely on the inherent flexibility of DNA for them to grasp it and force it into the S-like shape shown in the initiation complex.

FIGURE 12 

FIGURE 12. Schematic of large-scale conformational change and initiation of DNA looping and translocation by EcoR124I. (A) Type I enzymes exist in a dynamic equilibrium between open and closed states (movement is shown by orange arrows, and pivot points in carboxy-terminal regions of HsdM are indicated by pink dots). DNA (green) binding to form encounter complexes can occur nonspecifically to the HsdR (red) or via the target sequence to the MTase core (HsdM is in light and dark blue, and HsdS is in yellow). Complete closure of the enzyme and bending of the DNA around the HsdR produces the initiation complex for DNA translocation. (B) The predicted complete path of the DNA (green dots) through the atomic model of EcoR124I with segments of bound DNA. This is the proposed initiation complex (from Fig. 10A). During active translocation, the DNA would then form expanding loops from each side (light-green dots for DNA, and the direction of translocation is shown by black arrows). The inset shows the initiation complex turned 90° to the main panel. (Reprinted from Kennaway et al. 2012, with permission from Cold Spring Harbor Laboratory Press.)

 

The introduction of sharp bends in the DNA would require considerable energy to be expended by the enzyme. This may come from the transition between open and closed forms of the RM enzyme, but it may also require the hydrolysis of ATP by the HsdR. The models suggest that once the enzyme has closed around DNA and the motor of an HsdR subunit has a good grip on a segment of DNA, further hydrolysis of ATP (required for translocation and cleavage, although not DNA binding) would push the segment bound to the motor toward the central MTase core, as indicated by large arrows in Figure 12A. Because the MTase core is also tightly bound to the DNA target sequence, DNA at the bend between the segments bound to the motor and to the MTase core would twist and perhaps even buckle, forming the small loop shown in Figure 12B. Formation of this highly strained loop is certain to be energetically unfavorable, in agreement with translocation measurements for Type I enzymes in which it appears that much ATP is used in abortive attempts to initiate translocation (Seidel et al. 2008). Once the loop has formed, further DNA translocation would occur as the motors pump DNA toward the MTase core. Single-molecule experiments make it clear that the motors can work independently (Seidel et al. 2004, 2008), perhaps explaining why early EM studies showed both single- and double-looped structures (Yuan et al. 1980; Endlich and Linn 1985). In light of the large changes occurring upon DNA binding, it is possible that the actively translocating enzyme undergoes further changes in structure (e.g., in the presence of ATP). One may speculate that this great flexibility would allow the enzyme to accommodate the stresses built up during the extensive DNA translocation periods observed for these molecular machines. It is noteworthy, in this respect, that a process of deassembly of the enzymes occurs after DNA cleavage, and some of the subunits—although not all and depending on the particular Type I enzyme—can be reused (Roberts et al. 2011; Simons and Szczelkun 2011).

As mentioned above, the earlier model of Lapkouski et al. (Lapkouski et al. 2009) differs from this new model in two respects: namely, the orientation of the HsdR with respect to the MTase core, and the path taken by the DNA. Previously (Lapkouski et al. 2009), the interface of HsdR with the MTase core was not defined when compared with the new models. More importantly, the DNA was proposed to bend across the motor domains of HsdR, so that it came near to the endonuclease domain in the same HsdR and could be cleaved. If this model was correct (Lapkouski et al. 2009), the partially assembled R1M2S1 form of EcoR124I should have been able to cleave DNA, which was, however, not the case (Janscak et al. 1998). The current model suggests that the endonuclease domain of one HsdR is in proximity to DNA translocated by the other HsdR (Fig. 12B). This would explain the absence of DNA cleavage by partially assembled R1M2S1 forms of EcoR124I, despite the fact that such an assembly translocates DNA effectively (Janscak et al. 1998; Seidel et al. 2004, 2008). Thus, the current models are a significant improvement on the previously published models (Davies et al. 1999; Lapkouski et al. 2009).

Last, the structural models presented can be compared with the structures of complex Type II REases enzymes in groups IIB and IIG (Roberts et al. 2003), which cleave at defined distances either on both sides (IIB) or only one side (IIG) of the target sequence. As mentioned above, these classes are Type I-like combined R-M systems, with an amino-terminal endonuclease PD domain directly fused to a γ-class MTase domain in a single protein, but without motor domains (Dryden 1999; Nakonieczna et al. 2009; Shen et al. 2011). Whereas Type IIB enzymes have a single HsdS-like subunit with two TRDs, the Type IIG enzymes BpuSI and MmeI have only one TRD. The Type IIB enzyme is effectively a dimeric Type IIG. Thus, a Type IIB enzyme is like a motorless Type I system, and a Type IIG system is like one-half of a motorless Type I RM enzyme. Figure 13 compares the relative locations of one endonuclease domain, one HsdM, and the HsdS from the closed form of EcoR124I with the structures of MmeI and BpuSI (Nakonieczna et al. 2009; Shen et al. 2011). It can be seen how fusion of the endonuclease domain from HsdR to the start of HsdM in EcoR124I would move it to the same location as observed in the Type IIG REases and lead to cleavage downstream from the target sequence. Thus, the proposed role of gene fusions in the evolution of different groups of Type II R-M systems (Mokrishcheva et al. 2011) can be extended to include the evolution of the Type I systems.

FIGURE 13 

FIGURE 13. Structural evolution of Type IIG enzymes from a Type I enzyme undergoing fusion of the carboxyl terminus of an endonuclease domain from HsdR via deletion of the motor domains, to the amino terminus of HsdM. (Left) Part of EcoR124I, with one endonuclease domain from HsdR (red), one HsdM (green is the amino-terminal domain, and blue is the MTase catalytic domain), and HsdS (yellow with two TRDs). DNA bound to the MTase core is shown, but DNA bound to HsdR is omitted for clarity. The dashed line shows how the end of the endonuclease domain could join with the amino terminus of HsdM to form a structure similar to the type IIG structures shown on the right. The catalytic motifs in the endonuclease domain and HsdM are shown in spacefill. (Middle) Model of MmeI with bound DNA with the same colors for the equivalent domains (Nakonieczna et al. 2009; coordinates from ftp://genesilico.pl/iamb/models/RM.MmeI). (Right) Crystal structure of BpuSI (PDB ID: 3s1s) with the same coloring of domains as in the other structures and with an inserted extra domain shown in gray (Shen et al. 2011). DNA is absent in this structure, and one can see that the endonuclease domain would be blocking the DNA-binding site on the TRD. Shen et al. (2011) proposed that the endonuclease domain would twist away to allow DNA sequence recognition. (Reprinted from Kennaway et al. 2012, with permission from Cold Spring Harbor Laboratory Press.)

 

Additional Roles for Type I Enzymes

In 1977, Werner Arber wondered whether REases might have additional functions in the cell (Arber 1977). The finding that a Type I enzyme could cleave a replication fork at its branch may indicate that the answer is yes (Ishikawa et al. 2009). This probably happens when the enzyme travels along the DNA and encounters a replication fork, halts, and cuts (Ishikawa et al. 2009). Possibly in line with this, the groups of Hirotada Mori in Nara and Chieko Wada in Kyoto identified protein–protein interactions in E. coli (Arifuzzaman et al. 2006) between the EcoKI subunits and some other proteins, in a comprehensive pull-down assay using a His-tagged library of ORFs. Some of these interactions may be just a matter of sticky proteins, but others could be of in vivo relevance—for example, ATP-dependent helicase HrpA, the replicative helicase DnaB, DNA polymerase III DnaE, or CTP synthase PyrG, which may point to potential fine-tuning of R-M activity with DNA replication and primary metabolism. Such cross talk between (endo)nucleases and primary metabolism may well be universal, although likely to be much more complex in eukaryotic systems. As such, the E. coli system remains a useful model system to study such complex processes.

In an interesting report, Marie Weiserova and colleagues used immunoblotting to show that HsdR is phosphorylated on threonine in vivo only when coproduced with the MTase subunits (HsdM and HsdS) (Cajthamlova et al. 2007). HsdR lacks this phosphorylation when introduced in the cell by itself. Is this as yet unexplained phosphorylation of EcoKI HsdR another way of restriction control or genome maintenance (e.g., involved in recruiting other enzymes to the DNA) (Cajthamlova et al. 2007)? It certainly warrants further investigation.

On a different track, the Type I ecoprrI system contains an additional gene, originally identified by the group of Tom Bickle (Tyndall et al. 1994; Kaufmann 2000). The protein encoded by prrC proved to be a member of the group of latent anticodon nucleases (ACNs), which are proteins that may be linked to stress responses, and have been called RNA-based innate immunity systems that distinguish self from non-self (Jain et al. 2011).

Most studies on Type I systems concerned enzymes from strains that could be cultured in the laboratory. However, with the advent of whole-genome sequencing, many Type I systems have been identified in benign as well as pathogenic bacteria, in which they may limit genetic exchange between species and contribute to genome and strain stability. Because of the presence in their genomes of sometimes large numbers and different types of REases, these enzymes have proven useful for strain typing of, for example, methicillin-resistant Staphylococcus aureus (MRSA) bacteria (Lindsay 2010). REases such as SauI can limit horizontal transfer through conjugation, phage transduction, or transposition of antibiotic-resistance genes or virulence factors, but they cannot completely prevent it (Waldron and Lindsay 2006; Veiga and Pinho 2009). Knowledge of the target sites of different SauI members in a wide range of S. aureus lineages will aid studies on these pathogens (Cooper et al. 2017).

In 2011, the Kobayashi group showed that within a gene, stretches of amino acids can move from one position to another (Furuta et al. 2011). The authors suggest that such lateral domain movements within genes may be a novel common route to generate new specificities. Finally, much attention has been paid to pathogens that use phase variation, as discussed in Part E.

Antagonists of Type I Action: Antirestriction

This section is a shorter version of the text by Loenen et al. (Loenen et al. 2014a). Note that this type of antirestriction is rather different from the ClpXP-mediated proteolysis mentioned before (Chapter 7; see also, e.g., Simons et al. 2014), or the lambda Ral protein (or the analogous prophage protein Lar), which alleviates restriction by changing EcoKI from a maintenance into a de novo MTase (Chapter 6; see also, e.g., Loenen 2003). Antirestriction (anti-R) and anti-restriction-modification (anti-RM) systems in phage, plasmids, and transposons enhance their survival in a new host. Much of the early seminal work on anti-R by the T-uneven phages T7 and T3 was carried out by the groups of Bill Studier (Studier 1975, 2013; Studier and Movva 1976; Dunn and Studier 1981; Mark and Studier 1981; Bandyopadhyay et al. 1985; Moffatt and Studier 1988) and Detlev Kruger (Kruger et al. 1977a,b,c, 1983). These phages inject a small part of their DNA carrying the 0.3 gene (which encodes Ocr; see next page). The Ocr protein is produced and inactivates EcoKI and EcoBI, before the remainder of their DNA enters the cell. The best-known “artful dodger” of host restriction is probably phage T4, which encodes multiple functions to escape host defenses, some of them useful to genetic engineers (e.g., polynucleotide kinase) (Kruger and Bickle 1983; Bickle and Kruger 1993; Miller et al. 2003; Rifat et al. 2008; Petrov et al. 2010). Other work on anti-RM was carried out by the groups of Tom Bickle and about 20 papers (mainly in Russian) by Belogurov and Zavil'gel'skii spanning nearly 30 years (reviewed in Kruger and Bickle 1983; Bickle and Kruger 1993; Zavil'gel'skii 2000; Thomas et al. 2003; Tock and Dryden 2005; Dryden 2006), whereas in the United Kingdom the Wilkins laboratory solved the riddle of control of anti-R of the self-transmissible IncI plasmid, thus identifying novel transcriptional regulation from a single-stranded promoter (see the next section; Althorpe et al. 1999; Bates et al. 1999; Wilkins 2002; Nasim et al. 2004). During evolution, new mechanisms and countermechanisms between anti-RM and RM appeared continuously (Kruger and Bickle 1983; Bickle and Kruger 1993; Zavil'gel'skii 2000; Wilkins 2002; Thomas et al. 2003; Putnam and Tainer 2005; Tock and Dryden 2005; Dryden 2006; Zavil'gel'skii and Rastorguev 2009). T7 Ocr proved to be a DNA mimic with a large negatively charged patch on its surface (Atanasiu et al. 2001; Walkinshaw et al. 2002); see the next section. In the wake of Ocr, the structures of several other anti-RM proteins have been elucidated: ArdA from Tn916 from Enterococcus faecalis (Serfiotis-Mitsa et al. 2008), ArdB from E. coli CFT073 (Oke et al. 2010), and KlcA, an ArdB homolog from plasmid pBP136 from Bordetella pertussis (Serfiotis-Mitsa et al. 2010); see the subsection The Structure of ArdA.

Whereas Ocr seems to be confined to phage (particularly T7 and its relatives), ArdA and ArdB proteins are usually encoded by conjugative plasmids and transposons (Gefter et al. 1966; Hausmann 1967; Chilley and Wilkins 1995; Chen et al. 2014). In T7, Ocr is synthesized for the first 2 min of infection, before entry of the remainder of the phage DNA (Gefter et al. 1966; Hausmann 1967, 1988; Hausmann and Messerschmid 1988a,b; Moffatt and Studier 1988; Garcia and Molineux 1996; Molineux 2001). This completely inhibits the resident Type I enzymes, and the rest of the T7 genome can safely enter the cell. Interestingly, Ocr could bind in vitro to E. coli RNA polymerase (Ratner 1974), leading to the possibility that Ocr has another role in the cell, like other “moonlighting” proteins (Mani et al. 2015; Jeffery 2016) as, for example, reported for SF2 proteins involved in the skin disease Xeroderma pigmentosum (Le May et al. 2010; Kuper et al. 2014; Compe and Egly 2016). These SF2s are involved in nucleotide excision repair, but also in transcription, leading to active demethylation of CpG islands (Le May et al. 2010).

In the case of ArdA of plasmid ColIb-P9, ssDNA (which is resistant to restriction) enters the cell and forms an unusual promoter via a dsDNA hairpin, which allows transcription of the ardA gene (Chilley and Wilkins 1995; Althorpe et al. 1999; Bates et al. 1999; Nasim et al. 2004). Production of ArdA or ArdB rapidly inhibits the resident Type I enzymes. This novel transcription method may well be more common and deserves further research. Importantly, genome sequencing projects indicate that ard genes are widespread and often accompanied by antibiotic-resistance genes. The impact of combined transfer of these genes on the rate of spread of resistance in bacterial populations will be obvious.

The Structure of Ocr

Ocr is a striking example of DNA mimicry by a protein (reviewed in Loenen et al. 2014a). It is a dimer of two monomers of 116 aa and shaped like a banana with a length of ∼7.5 nm and 2–2.5 nm thick (Dunn et al. 1981; Atanasiu et al. 2001; Blackstock et al. 2001; Walkinshaw et al. 2002; Zavil'gel'skii et al. 2009). In this way, it mimics the shape and surface charge of a section of B-form DNA (Fig. 14). Each monomer contains several α-helices, a long loop, and unstructured flexible amino and carboxyl termini, respectively. The thinness of the structure means that it has a minimal hydrophobic core with many aromatic amino acids, which may resemble the aromatic core of another DNA mimic, Qnr (Hegde et al. 2005). Despite the small core, Ocr is very stable to heat and chemical denaturation (Atanasiu et al. 2001). On the surface of each monomer are 34 negatively charged amino acids and only 6 positive amino acids (∼12 of each would be expected for a typical globular protein of this size), although not all of these are required for activity (Stephanou et al. 2009a,b; Kanwar et al. 2016). The negative surface charges are spaced at roughly the same separation as the phosphate groups on 24 bp of B-form DNA containing a bend in its center, which explains its affinity for Type I enzymes (Atanasiu et al. 2002; Su et al. 2005).

FIGURE 14 

FIGURE 14. Superimposition of two 12-bp B-DNA molecules on the T7 Ocr dimer. The phosphate groups of 12 bases in each DNA dodecamer overlap with 12 carboxyl groups on each Ocr monomer, thus mimicking the shape and surface charge of B-form DNA. Ocr is shown as a blue ribbon with amino (N) and carboxyl (C) termini indicated and the dimer interface shown as a red line. Phosphate groups are colored yellow (phosphorus) and purple (oxygen). The carboxyl groups are colored red (oxygen) and black (carbon). The sugar backbones of the DNA chains are colored in two shades of green with the base pairs omitted for clarity. Vectors for the DNA helical axes are drawn as black lines. (Reprinted from Atanasiu et al. 2002.)

 

The Structure of ArdA

ArdA is also a mimic of B-form DNA, like T7 Ocr (Fig. 15; McMahon et al. 2009). In the crystal structure, ArdA is a dimer but it can exist both as a dimer and as a higher multimer in solution (Serfiotis-Mitsa et al. 2008). ArdA from Tn916 from E. faecalis is 166 aa long with a very small dimer interface (like Ocr). The dimer is an elongated bent cylinder of ∼15 nm × 2 nm. The thinness of the structure again means that it has a minimal hydrophobic core, but unlike Ocr, ArdA is not very stable to denaturation (Serfiotis-Mitsa et al. 2008). The fold of ArdA is completely different from that of Ocr: Each ArdA monomer has three small, loosely packed domains, suggesting a flexible structure. The domain folds have been found in other protein structures with a mix of α-helices, β-strands, and loops. The surface of each monomer is covered with numerous carboxyl groups such that the dimer mimics ∼42 bp of bent B-form DNA. The flexibility of the structure may indicate that the ArdA protein can mold itself to the contorted S-shaped DNA-binding groove on the Type I enzyme, with different domains interacting with the different R, M, and S subunits (Kennaway et al. 2009, 2012).

FIGURE 15 

FIGURE 15. Protein inhibitors of Type I R–M enzymes. (Top to bottom) DNA model (hydrogen atoms omitted) from PDB: 2Y7H displayed on the same scale as the proteins for structural comparisons; T7 Ocr (PDB: 1S7Z and 2Y7C), ArdA (PDB: 2W82) from Tn916 of E. faecalis (Davies et al. 1999), and ArdB (PDB: 2WJ9) from a pathogenicity island of E. coli CFT073, respectively. All three proteins are homodimeric. Their subunits are identical, but are displayed here in different colors. (Reprinted from Loenen et al. 2014a.)

 

The Structure of ArdB

The structures of two members of the ArdB family have been solved by both crystallography (ArdB from a pathogenicity island in E. coli CFT073) and NMR spectroscopy (KlcA from B. pertussis plasmid pBP136 [Oke et al. 2010; Serfiotis-Mitsa et al. 2010]). The ArdB and KlcA amino acid sequences are close homologs with >30% sequence identity. Klc genes form part of the kor operon involved in a regulatory network of these promiscuous plasmids (Larsen and Figurski 1994). The two ArdB structures are clearly very different from those of Ocr and ArdA and are globular proteins with a novel fold. They are neither elongated nor possess significant charged patches, so they are unlikely to cause anti-R via DNA mimicry (see Loenen et al. 2014a for discussion).

Effect of Protein Inhibitors Ocr and Ard on Restriction and Modification

The effectiveness of Ocr, ArdA, and ArdB in inhibiting Type I R-M systems has been tested in vivo with the classical efficiency of plating test (e.o.p. test; see, e.g., Chapter 1) comparing the titer of phage on an r+ host carrying an anti-RM gene on a plasmid versus the strain lacking the plasmid (Walkinshaw et al. 2002; Serfiotis-Mitsa et al. 2008, 2010; Zavil'gel'skii and Rastorguev 2009; Zavil'gel'skii et al. 2009, 2011). Active anti-R enhances the number of recovered phages, which are then tested for modification by comparing the e.o.p. on the r+ versus r strain: anti-M activity leads to a lower e.o.p. on the former. A novel calibrated in vivo titration assay was designed for EcoKI by the group of Zavil'gel'skii (Zavil'gel'skii et al. 2009), which allows antirestriction proteins to be distinguished based on the quantitative differences seen at different expression levels of Type I enzymes ( Zavil'gel'skii and Rastorguev 2009; Zavil'gel'skii et al. 2009, 2011).

The plating assays show that Ocr blocks all Type I R-M systems (Walkinshaw et al. 2002). This is a direct consequence of the extremely strong binding of Ocr to the DNA-binding groove in the MTase core of the enzymes (Atanasiu et al. 2002). ArdA (Serfiotis-Mitsa et al. 2008) and ArdB/KlcA (Serfiotis-Mitsa et al. 2010) block restriction in all Type I families. ArdA discriminates between restriction and modification (Nekrasov et al. 2007). ArdA of plasmid R16 preferentially targets the restriction function of EcoKI (Thomas et al. 2003). This minimal anti-M effect is due to the binding of ArdA to the MTase core being of similar or weaker strength than DNA binding to the core. This weak binding is sufficient to prevent restriction but not modification. ArdA and ArdB differ in their propensity to block modification (Walkinshaw et al. 2002; Zavil'gel'skii and Rastorguev 2009; Zavil'gel'skii et al. 2009, 2011; Serfiotis-Mitsa et al. 2010). ArdB/KlcA wild variants all have strong restriction inhibition but weak effect on modification for four Type I classes (Serfiotis-Mitsa et al. 2010). ArdB also shows little or no anti-M effect in vivo and, in line with this, no interaction has been observed in vitro between ArdB and the MTase core. Furthermore, although ArdB causes anti-R in vivo, no effect could be demonstrated in vitro on restriction. Therefore, the mechanism of anti-R used by ArdB is indirect and requires further investigation. David Dryden states (Loenen et al. 2014a): “Our understanding of anti-RM is still in its infancy. Aside from the three systems described above, few others have been studied beyond their initial discovery. Given their synergistic role with restriction and modification in regulating horizontal gene transfer and the resistome (Wright 2010; Stern and Sorek 2011), this deficiency in our knowledge needs to be addressed.”

Type I Single Protein

Type ISP (Type I single protein) are similar to Type IIL REases (like, e.g., MmeI), but have an additional helicase-ATPase domain, which is essential for restriction (Smith et al. 2009a,b,c). The prototype ISP REases are two plasmid-encoded R-M systems in Lactococcus lactis, LlaGI and LlaBIII. LlaBIII is of commercial importance as protection against phage infections in milk fermentations. An α-helical coupler domain connects the Mrr-like PD domain–cum-SF2 helicase region to the m6A γ-MTase-TRD domain. With the exception of the TRD region, LlaBIII is >95% homologous to LlaGI (Sisakova et al. 2013). Structural and single-molecule studies indicated a conformational change after binding to the recognition site; the enzyme leaves the site and translocates DNA without looping, at ∼300 bp/sec at 25°C, consuming one to two ATP per base pair (Chand et al. 2015; Kulkarni et al. 2016). As the PD and helicase/ATPase domains are upstream of the direction of translocation, these results indicated that the enzyme could not simply dimerize via its nuclease domain like Type I enzymes (Chand et al. 2015; Kulkarni et al. 2016). Together with data from single cleavage sequence analysis, this led to the proposal that DNA cleavage occurred as a result of multiple nicks by colliding enzymes, roughly halfway between sites, with the nuclease domains distal (Chand et al. 2015). Recent experiments show that translocation activates the nuclease domains via distant interactions of the helicase or MTase-TRD, without requiring direct nuclease dimerization (van Aelst et al. 2015).

Sequence analyses of 552 Type ISP enzymes showed structurally well-conserved elements involved in target recognition of LlaGI and LlaBIII, although the primary sequences of the TRDs were not that well conserved (cf. Type II enzymes) (Kulkarni et al. 2016). This led to a partial consensus code for target recognition by this class of enzymes with specificity changes due to residues that contacted the bases as well as novel contacts (Kulkarni et al. 2016).

PART C: TYPE III ENZYMES

Introduction

As detailed in the previous chapters, research on the Type III REases was basically limited to the enzymes from phage P1 and plasmid P15 by the groups of Bickle, Rao, Kruger and Reuter, and later Szczelkun (McClelland 2004), whereas the structure of EcoP15I would be finally published by Aneel Aggarwal and coworkers (see the next section). The EcoP15I R-M complex acted as MTase in the absence of ATP, and as MTase or REase in the presence of ATP, depending on the methylation state of the recognition site. Important questions remained: Why was the MTase a dimer of two Mod subunits, as methylation occurred on only one strand of the recognition sequence? Did this play a role in the stability of the complex? There appeared to be quite convincing evidence for a translocation and cutting mechanism similar to that of the Type I enzymes, but why the differences with respect to the interaction with ATP and SAM, and the location of the cut site? Did cleavage (always) occur after DNA tracking with ATP as an energy source, and collision by two complexes? Did this involve one Res subunit or two Res subunits, and did this occur in one or both directions? Would collision result in a conformational change and cleavage 25–27 bp downstream from one of the two sites?

This part of the chapter is based on the historical perspective by Rao et al. (Rao et al. 2014), recent papers that addressed the role of ATP hydrolysis in long-distance communication between sites before cleavage could occur, and the different models based on 1D diffusion and/or 3D-DNA looping. Although it had been known that Type III enzymes needed two sites in a specific head-to-head orientation for cleavage, later evidence indicates that the sites can also be in a tail-to-tail configuration (van Aelst et al. 2010). Moreover, new data provided evidence for the trimeric nature of EcoP15I, EcoP1I, and PstII: one Res (and not two) and two Mod subunits (Butterer et al. 2014). The long-awaited structure appeared of the first Type III REase, which is also the first structure of a dimeric MTase, EcoP15I,which shed unexpected new light on the interactions of this Mod2 dimer with the Res subunit (Gupta et al. 2015); see the next page. Finally, whole-genome sequencing data indicated Type III R-M systems in many sequenced genomes, in which a role for these enzymes in “phase variation” is being unraveled with respect to pathogenicity and virulence of clinically relevant organisms, such as H. influenzae and biofilm formation. After its initial discovery in the 1970s by Andrzej Piekarowicz and Stuart Glover in H. influenzae (Glover and Piekarowicz 1972; Piekarowicz and Glover 1972; Piekarowicz 1974; Piekarowicz and Kalinowska 1974; Piekarowicz et al. 1974, 1975, 1976, 1981, 1986; Jablonska et al. 1975; Piekarowicz and Baj 1975; Kauc and Piekarowicz 1978; Piekarowicz and Brzezinski 1980; Brzezinski and Piekarowicz 1982; Piekarowicz 1982, 1984), phase variation has become an important phenomenon and has also been studied in, for example, Neisseria sp. (Piekarowicz et al. 1988; Kwiatek and Piekarowicz 2007; Adamczyk-Popławska et al. 2009; Kwiatek et al. 2010); see the video from Andrzej's talk at the aforementioned CSHL meeting in 2013 (Piekarowicz 2013), and in Part E.

The Structure of EcoP15I

Different types of experiments addressing the composition of the EcoP15I R-M complex and the mechanism of DNA translocation and looping resulted in different models and much controversy (Peakman and Szczelkun 2004; Raghavendra and Rao 2004; Reich et al. 2004; Crampton et al. 2007a,b; Moncke-Buchner et al. 2009; Ramanathan et al. 2009; van Aelst et al. 2010; Dryden et al. 2011; Szczelkun 2011; Wyszomirski et al. 2012; Schwarz et al. 2013; Rao et al. 2014). Was this related to differences in the composition of the complex (ResMod2 or Res2Mod2 [Rao et al. 2014]) or perhaps even multimeric complexes present in the preparation? There was evidence for translocation and DNA looping by EcoP15I, but was looping essential and/or could ATP drive 1D diffusion of the enzyme on the DNA? In the absence of crystals, Aneel Aggarwal and coworkers used small-angle X-ray scattering (SAXS) and analytical ultracentrifugation to analyze the structure of the EcoP15I R-M complex and the dimeric Mod2MTase (Gupta et al. 2012). Whereas the MTase was relatively compact, the R-M complex was an elongated crescent shape of ∼218 Å. Their data were in line with a model in which the MTase dimer was lodged between Res subunits. Did this mean that the Res subunits would come together and form a sliding clamp around the DNA in order to cut the DNA? Three years later (2015), the first crystal structures of the EcoP15I complex with DNA (and AMP) were published. These results came as a surprise and led to novel insights into the way in which the helicase and modification domains of EcoP15I interacted with DNA and each other (Gupta et al. 2015).

DNA Recognition by EcoP15I

Until this DNA-EcoP15I cocrystal structure appeared, by necessity the interpretation of data on dimeric MTase-DNA structures was limited to those obtained with monomeric MTases (all without DNA) (Gupta et al. 2015). The structure was a big surprise: One EcoP15I Mod subunit, ModA, turns out to be involved in specific recognition of the bases in the target site, whereas the other subunit, ModB, has the target adenine (CAGCAG) in its catalytic cleft, which is rotated 180° out of the DNA helix (Fig. 16).

FIGURE 16 

FIGURE 16. Overall structure of EcoP15I/DNA/AMP complex (PDB: 4ZCFI). (A) The domain arrangements of Mod and Res subunits. (B) An overall view of the Mod and Res subunits (Mod2Res1) bound to DNA and AMP. The two Mod protomers, ModA and ModB, are shown in cyan and blue, respectively, whereas the Res subunit is shown in magenta. The DNA is shown in gray, with the exception of the extrahelical adenine base (yellow). The AMP molecule is shown in yellow. ModA recognizes DNA through base-specific interaction from its TRD (TRDA) and interacts with Res through its MTase domain and CTD. The TRD of ModB (TRDB) does not enter the DNA major groove. CTD of ModA (CTDA) interacts with the Res subunit, whereas the CTD of ModB (CTDB) is exposed to solvent. ModA and ModB dimerize via their NTDs (NTDA/B) and central MTase domains (MTaseA/B). AMP binds in a cleft between the RecA1 and RecA2 motor domains of the Res subunit. The endonuclease domain that ensues the helical spacer is disordered and labeled in a dashed box. The proximity of TRDA of ModA and Pin domain of Res (interdomain distance B14 Å) is highlighted by a double-headed arrow. The intervening loops in the structure that are not modeled because of weak density are represented by colored dashes. (Reprinted from Gupta et al. 2015.)

 

In contrast to γ-class MTases (Type I HsdM or Type II M·TaqI) or α-class MTases (Type II M·FokI, EcoDam), in which the TRD is adjacent to the active site cleft, EcoP15I Mod belongs to the β-class MTases, where the TRD lies far off this cleft (Gupta et al. 2015). The authors suggested that, by extension, a similar division of labor (ModA subunit for recognition, ModB subunit for modification) may be used by other, mainly dimeric, β-class MTases (e.g., M·RsrI) (Thomas and Gumport 2006) and even also apply to other DNA or RNA m6A-MTases in other organisms, including mammalian cells (Gupta et al. 2015). Such RNA methylation is very common in both nucleus and cytoplasm and, for example, is implicated in RNA metabolism (transcription/splicing) and stem cell development (possibly involving the β-class MTases METTL3/METTL14) (Gupta et al. 2015). Is this division of labor universal? Is it active in all kingdoms, exemplified, for example, by the plant de novo MTase, DRM2 (domains rearranged MTase 2, which methylates only one DNA strand, like EcoP15I) or SPOUT RNA MTases (in which the RNA binds in a cleft between two monomers, and the target base is in the catalytic pocket of one monomer) (Gupta et al. 2015)?

Restriction by EcoP15I

During the early 1990s, helicases had been considered ATP-dependent DNA and RNA unwinding enzymes, but this view was subsequently challenged by data on Type I and other enzymes indicating translocation without strand separation. It became clear that the specificity of helicases or translocases for different substrates was dictated by additional regions in between the RecA motor domains and/or the amino- or carboxy-terminal flanking regions (Singleton et al. 2007). For example, true helicases contain a wedge-like domain between the RecA domains to disrupt the base pair for unwinding (Singleton et al. 2007). During unwinding or translocation, the motors consume ATP with every step, but why did some enzymes consume very little ATP while traveling long distances on the DNA? The answer came from single-molecule fluorescent microscopy studies. These indicated that ATP hydrolysis of EcoP15I bound to its target site did not result in DNA translocation: The energy generated induced a conformational change that resulted in long-range diffusion of the enzyme on the DNA (Schwarz et al. 2013). Such ATP-triggered change of state, which allows sliding, was named “molecular switching” of the enzyme, which could happen on DNA as well as RNA (Szczelkun et al. 2010; Schwarz et al. 2013; Szczelkun 2013), but could probably also cause other events such as protein–protein interactions, for example, for the clamp loader (Kelch 2016). Subsequent kinetic studies support this notion of two distinct ATPase phases, a rapid consumption of ∼10 ATP inducing a conformational change and a slower phase related to the rate of dissociation of the enzyme from the recognition site (Toth et al. 2015).

The novel EcoP15I structure revealed three new substructures, two within the RecA1 segment and one between RecA1 and RecA2: a loop, a β-hairpin-like “Q-arm,” and a novel “Pin” domain (Fig. 16; Gupta et al. 2015). These are scattered through the RecA-like domain. Two are in the amino-terminal half and the Pin domain is an insertion that divides RecA1 half (motifs I-III) from RecA2 (motifs IV–VI) (see Supplementary Fig. 5 of Gupta et al. 2015 and Fig. 1a in Mackeldanz et al. 2013). The motor domains bound dsDNA and facilitated DNA sliding via this specialized Pin domain. The Pin domain adopted a tertiary structure that extended toward the ModA TRD subunit and interacted with the translocating strand of the DNA duplex (Gupta et al. 2015). The DNA was severely distorted from B form at two sites along its axis (i.e., where the adenine was ejected from the recognition site and near the ModA-Res interface). The first distortion was due to intrusion of ModA into the DNA major groove, whereas at the ModA-Res junction the DNA was bent ∼24° toward the minor groove, in the direction of the ModA TRD and Res Pin domain. The result was a reduction in the distance between these domains to <14 Å that “may facilitate an interaction between the two domains when EcoP15I assumes a diffusive or sliding state on DNA” (Gupta et al. 2015). Therefore, the EcoP15I motor domain interacted predominantly with the translocating strand. The semiclosed configuration of the EcoP15I motor domain differed from the more open structure of another translocase, S. solfataricus SF2 translocase, which might indicate that EcoP15I was in an intermediate state, following ATP hydrolysis but before AMP dissociation (see Gupta et al. 2015 for further discussion).

With this structure containing AMP in hand, could one predict what might happen in the presence of ATP? How might the enzyme slide? According to Aneel Aggarwal and coworkers (Gupta et al. 2015), the ModA TRD might move from the DNA major groove to the Pin domain and in this way adopt a “nonspecific” conformation that would result in DNA sliding. Such a movement of ∼40° would be possible because of a flexible linker between the TRD and MTase domain, which would prevent the Pin domain reaching the DNA. This structural model is in agreement with single-molecule studies, which suggest that the ResMod2 complex moves along the DNA (like Type ISP, but unlike Type I, in which the M2S complex remains bound to the recognition site). This trimeric complex would slide along the DNA until it collides with another bound complex, which would make it cleavage competent (Schwarz et al. 2013).

PART D: TYPE IV ENZYMES

Introduction

Like the Type IIM REases, Type IV modification-dependent REases (MDEs) recognize a variety of DNA modifications at cytosine or adenine bases (http://rebase.neb.com/rebase/rebase.html; Carlson et al. 1994; Roberts et al. 2003). This section is based on two recent reviews (Loenen and Raleigh 2014; Weigele and Raleigh 2016), and the reader is referred to these for more details and references. Over the years many papers were published reporting phages protecting their DNA against a wide range of Type I, II, and III REases by base modification such as methylation. This did not protect them against Type IV enzymes that preferentially or exclusively attack modified Cs and As. In the laboratory strain E. coli K12, McrA and McrBC recognize hm5C DNA and m5C, but the pathogenic E. coli CT596 also carries the gmrS and gmrD (glucose-modified hm5C restriction) genes allowing restriction of ghm5C DNA in, for example, wild-type T4 DNA (Bair and Black 2007). Interestingly, this activity in turn could be inhibited by T4 IPI*, a small protein encoded by this champion dodger of bacterial defense systems (Bair et al. 2007; Rifat et al. 2008).

Another Type IV REase, Mrr (modified DNA rejection and restriction), was identified in E. coli K12, because it caused cloning trouble by recognizing both m5C and m6A (Heitman and Model 1987; Waite-Rees et al. 1991). In addition to m5C, m6A, and m4C, many other modifications exist in both prokaryotes and eukaryotes with probable roles in defense and stress situations, in part via regulation of replication and transcription (Freitag and Selker 2005; Lobner-Olesen et al. 2005; Zhou et al. 2005; Borst and Sabatini 2008; Kaminska and Bujnicki 2008; Low and Casadesus 2008; Iyer et al. 2009; Kriaucionis and Heintz 2009; Ou et al. 2009; Tahiliani et al. 2009; Xu et al. 2009; Prohaska et al. 2010; Wang et al. 2011; Loenen and Raleigh 2014). The Type IV REases are hard to identify as they lack a cognate MTase, and therefore one cannot use Pacific Biosciences (PacBio) SMRT sequencing to find the recognition sites. Their discovery relies on a genetic system and suitable phages or plasmids as challengers. Cell extracts to digest DNA in vitro and analyze on gel are of little use because of severe degradation of the DNA (Eid et al. 2009; Korlach and Turner 2012; Roberts et al. 2015).

It is not known whether all Type IV enzymes flip the methylated base out of the helix, but that may be common. Figure 17 shows the structure of McrBC compared to other base-flipping proteins.

FIGURE 17 

FIGURE 17. McrB-N in comparison to other base-flipping proteins. (A) SRA domains SUVH5 (3Q0C) and UHRF1 (2ZKD) use loops extending from a crescent formed from two beta sheets to flip C or m5C from undeformed B-form DNA into a pocket (top row), whereas McrB-N (3SSC; bottom row) uses loops from one beta sheet to distort the DNA and flip the base. It resembles the human alkyladenine glycosylase (1BNK; bottom row) in bending the DNA toward the major groove, while flipping the base via the minor groove (Sukackaite et al. 2012). (B) The SRA-like hemimethylated m5C recognition domains. A ribbon model of the amino-terminal domain of the MspJI structure (4F0Q and 4F0P; top) compared with the SRA domain of URHF1 (PDB: 3FDE; bottom). The crescent shape formed by interacting beta sheets and helices αB and αC are the conserved features of the SRA domain highlighted here. Loops on the concave side of UHRF1 participate in flipping the base, and similar loops presumably do so for MspJI. Two of these vary in length among family members and may play roles in sequence context specificity (Horton et al. 2012). (Reprinted from Loenen and Raleigh 2014; A, originally adapted from Sukackaite et al. 2012; B, originally adapted from Horton et al. 2012.)

 

Fusions of DNA Binding and Cleavage Domains

All characterized Type IV proteins are fusions of various cleavage and DNA recognition domains. By 2014, six groups of enzymes had been identified that recognize modified DNA with low sequence selectivity, with the prototypes McrA, McrBC, Mrr, SauUSI, MspJI, PvuRts1I, and GmrSD (Loenen and Raleigh 2014). Five more groups have various fusions of the DBD and the cleavage domain, whereas they may be ATP- or GTP-dependent for recognition or cleavage (Weigele and Raleigh 2016). As PvuRts1I is qualitatively similar to MspJI family enzymes, and in turn to DpnI (i.e., typical IIM REases), it is debatable whether some enzymes should be classified as Type IIM or Type IV: The key feature discriminating IIM and IV enzymes is cleavage position, although this position is known (and fixed) for IIM enzymes (DpnI, PvuRts1I family, MspJI family), the cleavage position for the Type IV enzymes is either variable (McrBC) or unknown (McrA, Mrr).

McrA

McrA (recognition sequence YCGR) of E. coli K12 has been extensively analyzed by bioinformatics and mutagenesis. McrA has an amino-terminal DBD, and a carboxy-terminal HNH domain required for restriction in vivo (Bujnicki et al. 2000; Anton and Raleigh 2004). It recognizes m5C and hm5C (but not ghm5C), with a preference for C or T at the 5′ position, (Y>R)m5CGR; in vitro it binds m5CpG DNA, but does not restrict (Mulligan and Dunn 2008; Mulligan et al. 2010; Loenen and Raleigh 2014; Loenen et al. 2014b). Interestingly, a mutation in the DBD enabled in vivo discrimination between m5C (still recognized) and hm5C (not recognized) (Anton and Raleigh 2004). A protein with a similar HNH nuclease domain, but different DBD, in Streptomyces coelicolor A3, ScoA3McrA, cuts DNA modified by the E. coli Dcm protein (Cm5CWGG) or phosphorothioate (PT)-modified sites (or both) at a variable distance, and cleavage depends on Mn2+ or Co2+ (Gonzalez-Ceron et al. 2009; Xu et al. 2009; Liu et al. 2010).

McrBC

McrBC of E. coli K12 has been well characterized (Raleigh 1992; Sutherland et al. 1992; Gast et al. 1997; Pieper et al. 1997, 2002; Stewart and Raleigh 1998; Stewart et al. 2000; Panne et al. 2001; Pieper and Pingoud 2002; Sukackaite et al. 2012; for details, see Loenen and Raleigh 2014; Weigele and Raleigh 2016). The mcrB gene also encodes an additional protein, McrB(S), which starts at an internal translation initiation codon (thus lacking the first 161 aa) and appears to have a regulatory function (Beary et al. 1997). Figure 18 shows a model for the assembly of the McrBC complex (Loenen and Raleigh 2014). McrBC is a GTP-dependent heteroheptamer in which McrC (with the PD nuclease motif) binds a complex of McrB with GTP and DNA. McrBC makes a double-strand cut near one Rm5C site but requires the cooperation of two sites or a translocation block. The sites may be on different daughters across a fork. These are separated by 30-3000 bp, may be on either strand, and minor cleavage clusters are found ∼40, ∼50, and ∼60 nt from the m5C (Pieper et al. 2002).

FIGURE 18 

FIGURE 18. McrBC assembly model (Loenen and Raleigh 2014). Two proteins are expressed from mcrB in vivo. Both the complete protein (McrB-L) and a small one missing the amino terminus (McrB-S; top row) bind GTP, forming high-order multimers detected by gel filtration (second row). When visualized by scanning transmission EM, these appear as heptameric rings with a central channel. Rings of McrB-L in top views show projections that may correspond to the amino-terminal DBD (red segment). Both forms can then associate with McrC, judged again by gel filtration. McrB-L: GTP can bind to its specific substrate (RmC) in the absence of McrC (third row); in its presence, the substrate is cleaved (fourth row). GTP hydrolysis is required for cleavage (arrow): A supershifted binding complex forms in the presence of GTP-γ-S, but no cleavage occurs. Translocation accompanies GTP hydrolysis; dsDNA cleavage requires collaboration between two complexes, or a translocation block. The path of the DNA in the figure is arbitrary, as is the conformation of McrC. (Modified from Bourniquel and Bickle 2002, with permission from Elsevier Masson SAS.)

 

Mrr

Mrr of E. coli K12 recognizes m6A and m5C (with uncertain specificity), which prevented cloning of some R-M systems (e.g., PstI [CTGm6AG], HhaII [Gm6ANTC], and StyLTI [CAGm6AG]) and other genes (Heitman and Model 1987; Kelleher and Raleigh 1991; Waite-Rees et al. 1991; Loenen and Raleigh 2014; Weigele and Raleigh 2016). Mrr contains a predicted variant of the PD motif in the carboxy-terminal domain, with a presumed amino-terminal winged-helix DBD, like the MspJI family (Loenen and Raleigh 2014; Weigele and Raleigh 2016).

SauUSI

The (d)ATP-dependent SauUSI (SCNGS) of S. aureus recognizes Sm5CNGS (S = C or G) (Loenen and Raleigh 2014). It has a PLD nuclease domain (like BfiI and other REases), a carboxy-terminal DBD, a helicase/translocase domain in between, and a cleavage domain (Xu et al. 2011) with a reaction mechanism resembling that of topoisomerases and transposases (Interthal et al. 2001; Sasnauskas et al. 2003, 2007), but not quite the same: PLD nucleases employ a covalent protein–DNA intermediate, but, unlike topoisomerases and some (e.g., serine) transposases, the covalent linkage is made by a histidine rather than serine or tyrosine.

PvuRts1I

PvuRts1I from P. vulgaris Rts1 restricted glucosylated T-even phages in vivo (Janosi et al. 1994; Loenen and Raleigh 2014). It recognizes (g)mC(N11-13/N9-10)G—that is, it cuts 11–13 nt downstream from a modified C, which is 20–23 nt before a G (hence, not very specific). It has approximately 20 active homologs, including AbaSI form Acinetobacter baumanii, which recognize m5C slightly and hm5C and ghm5C better, with differing preferences and weak and variable selectivity for the sequence surrounding the modified base (Szwagierczak et al. 2011; Borgaro and Zhu 2013). These enzymes require two modified sites for dsDNA cleavage ∼22 nt apart, with incisions ∼11–13 nt 3′ to the modified base on the one strand and 9–10 nt 3′ on the other. Bioinformatics, mutational evidence, and crystal structures indicate a carboxy-terminal SRA-like (SET and ring finger–associated) DBD, and an amino-terminal cleavage domain that is a divergent member of the PD family (Kazrani et al. 2014; Shao et al. 2014; Weigele and Raleigh 2016). The structure of AbaSI confirmed the results with PvuRts1I, with an amino-terminal nuclease domain resembling that of VSR, whereas the carboxy-terminal domain resembles SRA family members (Horton et al. 2014a; Weigele and Raleigh 2016).

GmrSD

The aforementioned GmrSD enzyme, encoded by the gmrS and gmrD genes from a pathogenic E. coli strain, also exists as a single fusion protein with similar characteristics (Bair et al. 2007; He et al. 2015; Machnicka et al. 2015). In vitro, the enzyme prefers ghm5C DNA to unmodified m5C. GmrD is the nuclease, whereas the GmrS domain has a ParB/Srx fold, present in conjugative plasmids (He et al. 2015; Machnicka et al. 2015; see Weigele and Raleigh 2016 for details). The enzyme has a strong preference for UTP over GTP and CTP (Loenen and Raleigh 2014; Weigele and Raleigh 2016).

MspJI

MspJI from Mycobacterium sp. JLS, was the first member of a family of Type IV enzymes, which cleave with a four-base 5′ extension 12 nt from the m5C, and 16–17 nt on the opposite strand (Bujnicki and Rychlewski 2001; Zheng et al. 2010; Cohen-Karni et al. 2011; Horton et al. 2012, 2014b,c). Members recognize m5C and hm5C, but not ghm5C. MspJI has a SRA-like amino-terminal domain, which flips m5C out of the helix, like other SRA-like proteins. Swapping experiments with short protein loops contacting nearby bases, which varied among three members, significantly reduced selectivity for sequences flanking the modified base (Sasnauskas et al. 2015; Weigele and Raleigh 2016).

The endonuclease domain cuts a different DNA strand than the DNA bound by that polypeptide. This is reminiscent of M·EcoP15I in which the DBD of one subunit binds the recognition sequence, whereas the catalytic domains of, in this case, two other subunits, effect dsDNA cleavage. MspJI (and the isoschizomer AspBHI [Horton et al. 2014b,c]) can be thought of as Type IIS enzymes (albeit enzymes that bind only when the recognition site is modified) and their tetrameric structure and domain cooperation could be typical of many “normal” Type IIS enzymes, as well as the Type IIB, IIC, and IIG enzymes. The convention is to think of these acting as transient homodimers, but they might well act as tetramers (and even as multimers of tetramers, e.g., BcgI) instead.

PART E: PHASE VARIATION

Introduction

Bacterial pathogens not only infect the host, but also try to maintain themselves (“colonize”) by hiding from the immune system and/or confusing it. One way of doing this is by hypermutation at simple sequence repeats or homopolymeric tracks located within the reading frame or in the promoter of a subset of genes (Moxon et al. 2006). This is due to polymerase slippage (called slipped-strand mispairing [SSM]), which could be an important evolutionary strategy (Levinson and Gutman 1987a,b). Genetic variation in the population of pathogenic bacteria via SSM would result in two or (many) more different phenotypes that allow the strain to evade the host immune system (Robertson and Meyer 1992). SSM changes the number of repeats or bases, which switches promoters “on” or “off” (e.g., by changing the distance between the -35 and -10 regions), causes frameshifts in coding regions, and/or dictates alternative usage of multiple translation initiation codons in different reading frames, thus altering or abolishing DNA recognition (see, e.g., van Ham et al. 1993; van Belkum et al. 1998; De Bolle et al. 2000; van der Woude and Baumler 2004; Srikhanta et al. 2005, 2010; Moxon et al. 2006; van der Woude 2006; Dixon et al. 2007).

The groups of Richard Moxon, Andrzej Piekarowicz, and Michael Jennings have studied repeat variation that created this so-called “phase variation” in three pathogens, H. influenzae, H. pylori, and Neisseria sp., revealing altered expression of up to 80 genes, including genes important for iron uptake, DNA repair, electron transport, amino acid transport, and growth (reviewed in Srikhanta et al. 2010) but also for MTases. In the latter cases, the MTases can function as an on–off switch for multiple genes that allow the pathogen to combat host immunity (De Bolle et al. 2000; Seib et al. 2002; Srikhanta et al. 2005, 2010; Casadesus and Low 2006; Moxon et al. 2006; Wion and Casadesus 2006; Marinus and Casadesus 2009). Various other clinical isolates also contain MTases with repeats of variable length, which are either lone MTases or associated with Type II systems (Kong et al. 2000; Xu et al. 2000a,b; Lin et al. 2001; Vitkute et al. 2001; Skoglund et al. 2007). These genetic systems have been called “phasevarions” or “phase-variable regulons,” and appear to be a common strategy to randomly switch between distinct cell types and create phenotypic heterogeneity in the bacterial pathogenic population (Weiser et al. 1990; van Ham et al. 1993; Hallet 2001; Srikhanta et al. 2010). Such methylation-driven alternative gene expression can be very complex because of the presence of multiple phase-variable genes (see Srikhanta et al. 2010 for further details).

Taken together, different types of variation may alter gene expression: (1) reversible changes (i.e., alternative states of the same genes result in expression or not, or expression of different genes [e.g., flagella or tail fibers], via inversion of promoters or amino termini, or slipped mispairing); (2) diversity selected changes, in which a rare genotype is favored because of changing circumstances (involving outside forces); and (3) replacement or addition variation, in which a gene is replaced by another gene or is added extra.

Variable Type II systems

Around the turn of the century, two dozen potential R-M systems were identified in two completely sequenced H. pylori strains, 26695 and J99, amounting for >4% of the total genome (i.e., much more than in other sequenced bacterial genomes) (Kong et al. 2000; Lin et al. 2001). Although nearly 90% of the R-M genes were present in both strains, <30% of the Type II R-M systems were functional in both strains with different sets active in each strain. An interesting observation was that all strain-specific R-M genes were active, whereas most shared genes were inactive. Did this indicate that these active strain-specific genes had been acquired recently via horizontal transfer from other bacteria? And did these pathogenic strains constantly acquire new R-M systems and inactivate and delete the old ones (Lin et al. 2001)? Were these multiple R-M systems a “primitive bacterial immune system, by alternatively turning on/off a subset of numerous R-M systems” (Kong et al. 2000)? Support for this idea came from other H. pylori strains, in which the R-M systems also proved to be highly diverse (Xu et al. 2000a). In addition, several REases had novel specificities: Hpy178III (TCNNGA), Hpy99I (GGWCG), and Hpy188I (TCNGA) (Xu et al. 2000a). The latter system was absent in the 26695 and J99 strains, whereas the GC content implied that the Hpy188I system had been recently introduced into the H. pylori genome (Xu et al. 2000b).

In 2016, phase variation of a Type IIG enzyme was reported in Campylobacter jejuni NCTC11168 (Anjum et al. 2016). This IIG protein methylates adenine in CCCGA and CCTGA sequences. Using both inhibition of restriction and PacBio-derived methylome analyses of mutants and phase variants, the cj0031c allele in this strain was demonstrated to alter site-specific methylation patterns and gene expression, which “may indirectly change adaptive traits” (Anjum et al. 2016).

Phase-Variable Type III systems

Phase-variable Type III mod genes have been identified in H. influenzae, H. pylori, and Neisseria sp., which contain tandem repeats that may be homopolymeric, or repeat tracks of 2, 3, 4, or 5 nt (Srikhanta et al. 2010). The mod-like gene of H. influenzae Rd has 40 AGTC repeats within its ORF. This mod gene was found in 21 out of 23 other H. influenzae strains, and in 13 of those the locus contained repeats of variable length (De Bolle et al. 2000). These repeats comprised a hypervariable region in the central region of the mod gene of 22 nontypeable H. influenzae strains, whereas the res gene was conserved (Bayliss et al. 2006). Moreover, similar mod genes with similar hypervariable regions were identified in pathogenic Neisseria sp., suggesting horizontal transfer of these genes between different species. This high phase variability of these MTases would not only protect against phage infections but might “also have implications for other fitness attributes of these bacterial species” (Bayliss et al. 2006). An example of phase variation that allows typing of these so-called untypeable H. influenzae (“NTHi”) strains is shown in Figure 19, redesigned from Figure 3 in Fox et al. 2007. Variation of the number of repeats generates ON/OFF switches. Switching within a clonal population results in subpopulations with and without RM activity. In Haemophilus and other taxa, this results in variable expression of distant genes as well, presumably regulated by the presence of modification in regulatory regions (e.g., Tan et al. 2016; see, for a review, Sanchez-Romero et al. 2015). The variable regulation at distant sites does not require activity of the restriction function (Fox et al. 2007). Also observed in this figure is the presence of variable segments (TRDs) that give distinct recognition specificities. These variations are generated by horizontal transfer between strains, species, and even genera (Bayliss et al. 2006). Further studies indicated that in H. influenzae Rd, phase variation of modA was calculated to occur at a frequency of 4 × 10−6 mutations/division/repeat unit for off-to-on switches and 7 × 10−6 mutations/division/repeat unit for on-to-off switches (De Bolle et al. 2000). The restriction phenotype of a Type III system in N. gonorrhoeae, NgoAXP, switched randomly because of a change in the number of pentanucleotides (CCAAC/G) present at the 5′ end of the coding region of the ngoAXPmod gene (Adamczyk-Popławska et al. 2009). The mod gene in another N. gonorrhoeae strain, FA1090, was linked to biofilm formation, adhesion to human cells and epithelial cell invasion, and hence to pathogenicity and systemic infection (Kwiatek et al. 2015). Various serotypes of Pasteurella haemolytica have CACAG repeats within the 5′ end of a Type III R-M system, repeats which could change in length upon serial subculture and most likely also occurred as a result of DNA SSM (Ryan and Lo 1999).

FIGURE 19 

FIGURE 19. Phase variation. A representation of phase-variable methyltransferase genes present in different strains of nontypeable H. influenzae (NTHi). The prototypical strain of each NTHi strain in which each allele/arrangement is present is shown on the left side. Each modA gene is represented as a white arrow, with the DNA recognition domain (DRD) represented by a colored box. Downstream from each modA is the cognate restriction endonuclease gene, res, with the locations of the ATP-binding motif (TGxGKT), the ATP-hydrolysis motif (DEAH), and the endonuclease motif, PD · · · (D/E)XK, indicated above. Phase-variable modA genes are represented by the five most common modA alleles found in NTHi isolates from patients with middle ear infection (otitis media [OM]). These alleles—modA2, 4, 5, 9, and 10—all contain a simple-sequence repeat (SSR) tract; in this case the sequence AGCC repeats n times. The SSR tract is represented by a gray box in these alleles, with the number of AGCC repeats and the expression status of this number of repeats shown underneath each allele. For example, modA2 in NTHi strain 723 contains 16 AGCC repeats, which leads to expression (ON) of the gene. NTHi strain R539 contains a deletion of the entire mod-res region. NTHi strain 162 contains the modA7 allele that does not contain SSRs and therefore is not phase variable. ModA alleles that are not phase-variable contain the 12-nt sequence 5′-TCAGATAGTCAG-3′ in place of a SSR. In all cases, the genes flanking the mod-res locus are conserved: upstream is the gene rnhB (represented by the blue arrow to the left of each modA gene); downstream is the gene pcaC (represented by the yellow arrow to the right of each res gene). (Courtesy of John Atack.)

 

Phase-Variable Type I systems

Phase variation also occurs in Type I systems. HindI of H. influenzae has a (GACGA)4 repeat, and changes in the number of pentanucleotides (which is influenced by Dam methylation) within the coding sequence of hsdM were linked to protection against phage (Zaleski et al. 2005). In Mycoplasma pulmonis, two hsd loci each contain two hsdS genes with complex, site-specific DNA inversion systems (Dybvig et al. 1998). This generates a complete family of related hsdS genes with extensive sequence variations that recognize different DNA sequences, suggestive of additional roles in genome maintenance (Dybvig et al. 1998).

In the Type IC NgoAV system from N. gonorrhoeae, the length of tandem repeats of four amino acids were involved in the generation of a truncated or full-length HsdS protein, and only the long protein could complement other Type IC systems (Adamczyk-Popławska et al. 2011). Similar tetra-amino acid repeats (either TAEL, LEAT, SEAL, or TSEL) were identified in other Type IC systems in distantly related bacteria (Adamczyk-Popławska et al. 2003). Was this a common special characteristic of Type IC systems (Adamczyk-Popławska et al. 2003), and is there a tale to tell? Do these data shed light on the origin of the switch between the EcoR124I and EcoR124II systems mentioned in Chapter 6 (Fig. 6)? These proteins recognize GAA (N6) RTGC and GAA (N7) RTGC, respectively, and the presence of either six or seven bases between the two specific DNA regions was shown to relate to the presence of either two or three TAEL repeats in the respective HsdS subunits (Price et al. 1989). Price et al. (1989) suggested that the switch between the two specificities could be due to unequal crossing-over in these repeats (Price et al. 1989). Is this a coincidence or are these repeats perhaps the end result of prolonged SSM? Were these TAEL repeats once much longer, and were they slowly eliminated during continued growth under laboratory conditions, because they were no longer needed?

Such an assumption could fit in with LacZ fusion experiments by Bolle et al. (2000) with respect to the above-mentioned mod-like gene of H. influenzae Rd (with 40 AGTC repeats within its ORF). These authors fused a lacZ reporter to a chromosomal copy of mod downstream from the repeats, which resulted in high-level phase variation. Changing the number of repeats changed mutation rates. Phase variation occurred at a high frequency in strains with the wild-type number of repeats. Rates increased linearly with tract length over the range 17–38 repeat units. The majority of tract alterations were insertions or deletions of one repeat unit with a 2:1 bias toward contractions of the tract (De Bolle et al. 2000). This could be interpreted to mean that the shorter the track, the higher the chance that the track would become shorter faster. As under laboratory conditions the pressure is absent to protect against either phage or host immunity, the bacteria with shorter tracts would have a favorable advantage over bacteria with longer tracts. That advantage would eventually be lost once the number of repeats became very small, and the length of the spacer between the two recognition domains could no longer be decreased without loss of the DNA recognition function. The end result would be two active enzymes, EcoR124I and EcoR124II, with only two or three TAEL repeats left.

Monika Adamczyk-Popławska et al. (Adamczyk-Popławska et al. 2011) analyzed a poly(G) tract in the hsdS(NgoAV1) gene in N. gonorrhoeae. Deletion of 1 nt in this tract with seven guanines led to a frameshift at the 3′ end of the hsdS(NgoAV1) gene and fusion to a second downstream hsdS gene, hsdS(NgoAV2) (Adamczyk-Popławska et al. 2011). This resulted in a longer HsdS protein with two TRDs, rather than the original truncated HsdS protein with a single TRD derived from hsdS(NgoAV1) (Adamczyk-Popławska et al. 2011). Such a contraction of the poly(G) tract that caused this frameshift might well occur in vivo, as the authors found a minor subpopulation of cells that appeared to have only six guanines. Thus, it could be argued that the strain could switch this Type I system “on” (two TRDs = protection against phage and/or other foreign DNA) or “off” (one TRD and the possibility of DNA exchange via horizontal transfer).

FINAL THOUGHTS

It has proven very difficult to generate mutant or hybrid restriction enzymes with long DNA specificity sites that would serve as good tools for gene therapy. The new RNA-based method of genome editing using the CRISPR–Cas9 system may present a better alternative, but also has its drawbacks. To answer the question of how foolproof the CRISPR system is, Bull and Malik (2017) state in their discussion of a recent paper by Champer et al. (2017) that “the easy targeting of CRISPR, the very property that has led to its current popularity, may also be its downfall as a practical means to control populations or suppress disease transmission.” This would fit in with the research described in this book, which proves that organisms go to incredible lengths to keep genome alterations under tight control to avoid chromosomal instability and allow only a low level of heterogeneity.

The year of the 5th NEB meeting in Bristol in 2004 (Fig. 20) was also the year of the publication of the first and only book totally dedicated to restriction enzymes (edited by Alfred Pingoud [Pingoud 2004]). It reflects the growing realization of the importance of restriction-modification systems in “cells, not eppendorfs” (King and Murray 1994), and their role outside the laboratory in communities with benign and pathogenic bacteria and archaea.

FIGURE 20 

FIGURE 20. The NEB meetings between 1988 and 2015. (Top left) The first four NEB meetings on restriction and modification enzymes. (Top right) The 5th NEB meeting on Restriction/Modification in Bristol (2004) was organized by Mark Sczcelkun, Bernard Connolly, and David Dryden. (Bottom left) The 6th NEB meeting on DNA Restriction and Modification in Bremen (2010) was organized by Albert Jeltsch, Alfred Pingoud, and Wolfgang Wende. (Bottom right) The 7th NEB meeting on DNA Restriction and Modification in Gdansk (2015) was organized by Iwona Mruk, Geoffrey Wilson, and Richard Morgan. (Courtesy of Rich Roberts.)

 

The large current number of restriction enzymes and the 50-odd structures indicate overlap between types and subtypes in different ways. The enzymes may share a common ancestor with three separate domains for DNA recognition, restriction, and methylation, whereas Type I and III enzymes also contain the ATP-dependent SF2-related molecular motor domains. The cooperation of restriction enzymes between sites with or without looping and with or without collision/stalling may be more the rule than the exception. Different catalytic nuclease domains, mainly with the PD · · · (D/E)XK motif, but also HNH, GIY-YIG, or PLD domains, seem to have been “mixed and matched” during the course of evolution. The structures indicate careful control of positioning of the nuclease domain and large conformational changes before cleavage can occur, plus a variety of other control systems to avoid rampant nuclease activity, which would result in genome instability. Such other control may inhibit synthesis of the restriction enzymes in the cell at the level of transcription at the operon promoter or via C proteins, at the level of translation and/or posttranslation, and finally via DNA mimics employed by various plasmids and phages that inhibit Type I enzymes. The structure of the EcoP15I Type III enzyme indicates a division of labor of the two modification subunits—one for DNA recognition and the other for methylation, which may be a more general mechanism and may also apply to EcoKI (see below). EcoP15I hydrolyses approximately 30 ATP molecules in two steps (a fast consumption of approximately 10 ATP molecules followed by a slower consumption of approximately 20 ATP, which switches the enzyme into another rather distinct structural state that can diffuse on DNA over long distances (Schwarz et al. 2013). This resembles the situation with EcoKI, in which ATP also acts as allosteric effector (Burckhardt et al. 1981). The big difference with EcoP15I is that the EcoKI methyltransferase complex remains bound to the recognition site, whereas the HsdR subunits translocate DNA and hydrolyze ATP, and the whole EcoP15I complex slides away from the site (thus consuming less ATP). Both the ATP-triggered thermal diffusion and ATP-dependent initiation of translocation are important observations for studies into the much more complex eukaryotic methyltransferases and helicases.

One of the most important applications of restriction enzymes with concomitant impact on science and society has been the development of the DNA fingerprinting technique by Alec Jeffreys. Alec started his “Zoo blots” in the 1970s next door to the author of this book in Leicester, where he continued his work on restriction fragment length polymorphisms (RFLPs), which research could fill a book by itself (Fig. 21; Jeffreys 2006).

FIGURE 21 

FIGURE 21. The first DNA fingerprint developed by Alec Jeffreys in the Genetics Department in Leicester, September 10, 1984. (Courtesy of Alec Jeffreys.)

 

A final note about the intriguing story of EcoKI, hemimethylation, and cofactor SAM, a subject close to the author's heart (Loenen and Murray 1986; Loenen 2003, 2006, 2010, 2017, 2018): As already suggested in 1981 (Burckhardt et al. 1981), EcoKI has two types of binding sites for SAM—one for DNA recognition (the effector site) and the other for methylation (the methyl donor site). How does this relate to the ability of EcoKI to switch from a maintenance methyltransferase (with a strong preference for methylation of hemimethylated DNA) to a de novo methyltransferase? This ability is a property of Type IA enzymes (and not Type IC enzymes such as EcoR124I) that has been observed in the presence of the lambda Ral protein (Zabeau et al. 1980; Loenen and Murray 1986). How can a small protein like Ral cause such an important switch? Are there similar proteins to be discovered in eukaryotic systems? And what about the de novo methylation EcoKI mutants (e.g., the HsdM L113Q mutant made in Noreen Murray's laboratory) (Kelleher et al. 1991)? In contrast to wild-type EcoKI, which has two indistinguishable high-affinity SAM-binding sites, mutants like L113Q have only one high-affinity site, whereas the second site has a low affinity for SAM (Winter 1997). Is it the asymmetry of the complex with DNA that controls the switch between methylation and restriction? How the ability to bind only one SAM molecule properly turns L113Q into a de novo enzyme, without apparently affecting the methylation reaction as such, requires further investigation, as the ability to distinguish between unmethylated and hemimethylated DNA is of fundamental importance to cellular activities.

REFERENCES

Abrosimova LA, Monakhova MV, Migur AY, Wolfgang W, Pingoud A, Kubareva EA, Oretskaya TS. 2013. Thermo-switchable activity of the restriction endonuclease SsoII achieved by site-directed enzyme modification. IUBMB Life 65: 1012–1016. 10.1002/iub.1222

Adamczyk-Popławska M, Kondrzycka A, Urbanek K, Piekarowicz A. 2003. Tetra-amino-acid tandem repeats are involved in HsdS complementation in type IC restriction-modification systems. Microbiology 149: 3311–3319. 10.1099/mic.0.26497-0

Adamczyk-Popławska M, Lower M, Piekarowicz A. 2009. Characterization of the NgoAXP: phase-variable type III restriction-modification system in Neisseria gonorrhoeae. FEMS Microbiol Lett 300: 25–35. 10.1111/j.1574-6968.2009.01760.x

Adamczyk-Popławska M, Lower M, Piekarowicz A. 2011. Deletion of one nucleotide within the homonucleotide tract present in the hsdS gene alters the DNA sequence specificity of type I restriction-modification system NgoAV. J Bacteroiol 193: 6750–6759. 10.1128/JB.05672-11

Aiken C, Milarski-Brown K., Gumport R.I. 1986. RsrI and EcoRI are two different restriction endonucleases that recognise the same DNA sequence. Fed Proc 45: 1914.

Althorpe NJ, Chilley PM, Thomas AT, Brammar WJ, Wilkins BM. 1999. Transient transcriptional activation of the Incl1 plasmid anti-restriction gene (ardA) and SOS inhibition gene (psiB) early in conjugating recipient bacteria. Mol Microbiol 31: 133–142. 10.1046/j.1365-2958.1999.01153.x

Anjum A, Brathwaite KJ, Aidley J, Connerton PL, Cummings NJ, Parkhill J, Connerton I, Bayliss CD. 2016. Phase variation of a Type IIG restriction-modification enzyme alters site-specific methylation patterns and gene expression in Campylobacter jejuni strain NCTC11168. Nucleic Acids Res 44: 4581–4594. 10.1093/nar/gkw019

Anton BP, Raleigh EA. 2004. Transposon-mediated linker insertion scanning mutagenesis of the Escherichia coli McrA endonuclease. J Bacteriol 186: 5699–5707. 10.1128/JB.186.17.5699-5707.2004

Anton BP, Heiter DF, Benner JS, Hess EJ, Greenough L, Moran LS, Slatko BE, Brooks JE. 1997. Cloning and characterization of the BglII restriction-modification system reveals a possible evolutionary footprint. Gene 187: 19–27. 10.1016/S0378-1119(96)00638-5

Arber W. 1977. What is the function of restriction enzymes? Trends Biochem Sci 2: N176–N178. 10.1016/0968-0004(77)90062-7

Arifuzzaman M, Maeda M, Itoh A, Nishikata K, Takita C, Saito R, Ara T, Nakahigashi K, Huang HC, Hirai A, et al. 2006. Large-scale identification of protein–protein interaction of Escherichia coli K-12. Genome Res 16: 686–691. 10.1101/gr.4527806

Armalyte E, Bujnicki JM, Giedriene J, Gasiunas G, Kosinski J, Lubys A. 2005. Mva1269I: a monomeric type IIS restriction endonuclease from Micrococcus varians with two EcoRI- and FokI-like catalytic domains. J Biol Chem 280: 41584–41594. 10.1074/jbc.M506775200

Atanasiu C, Byron O, McMiken H, Sturrock SS, Dryden DT. 2001. Characterisation of the structure of ocr, the gene 0.3 protein of bacteriophage T7. Nucleic Acids Res 29: 3059–3068. 10.1093/nar/29.14.3059

Atanasiu C, Su TJ, Sturrock SS, Dryden DT. 2002. Interaction of the ocr gene 0.3 protein of bacteriophage T7 with EcoKI restriction/modification enzyme. Nucleic Acids Res 30: 3936–3944. 10.1093/nar/gkf518

Bair CL, Black LW. 2007. A type IV modification dependent restriction nuclease that targets glucosylated hydroxymethyl cytosine modified DNAs. J Mol Biol 366: 768–778. 10.1016/j.jmb.2006.11.051

Bair CL, Rifat D, Black LW. 2007. Exclusion of glucosyl-hydroxymethylcytosine DNA containing bacteriophages is overcome by the injected protein inhibitor IPI*. J Mol Biol 366: 779–789. 10.1016/j.jmb.2006.11.049

Baker O, Tsurkan S, Fu J, Klink B, Rump A, Obst M, Kranz A, Schrock E, Anastassiadis K, Stewart AF. 2017. The contribution of homology arms to nuclease-assisted genome engineering. Nucleic Acids Res 45: 8105–8115. 10.1093/nar/gkx497

Ball N, Streeter SD, Kneale GG, McGeehan JE. 2009. Structure of the restriction-modification controller protein C.Esp1396I. Acta Crystallogr, Sect D: Biol Crystallogr 65: 900–905. 10.1107/S0907444909020514

Ball NJ, McGeehan JE, Streeter SD, Thresh SJ, Kneale GG. 2012. The structural basis of differential DNA sequence recognition by restriction-modification controller proteins. Nucleic Acids Res 40: 10532–10542. 10.1093/nar/gks718

Bandyopadhyay PK, Studier FW, Hamilton DL, Yuan R. 1985. Inhibition of the type I restriction-modification enzymes EcoB and EcoK by the gene 0.3 protein of bacteriophage T7. J Mol Biol 182: 567–578. 10.1016/0022-2836(85)90242-6

Bart A, Dankert J, van der Ende A. 1999. Operator sequences for the regulatory proteins of restriction modification systems. Mol Microbiol 31: 1277–1278. 10.1046/j.1365-2958.1999.01253.x

Bates S, Roscoe RA, Althorpe NJ, Brammar WJ, Wilkins BM. 1999. Expression of leading region genes on IncI1 plasmid ColIb-P9: genetic evidence for single-stranded DNA transcription. Microbiology 145 (Pt 10): 2655–2662. 10.1099/00221287-145-10-2655

Baumann K. 2014. Epigenetics: Enhancers under TET control. Nat Rev Mol Cell Biol 15: 699. 10.1038/nrm3901

Bayliss CD, Callaghan MJ, Moxon ER. 2006. High allelic diversity in the methyltransferase gene of a phase variable type III restriction-modification system has implications for the fitness of Haemophilus influenzae. Nucleic Acids Res 34: 4046–4059. 10.1093/nar/gkl568

Beary TP, Braymer HD, Achberger EC. 1997. Evidence of participation of McrB(S) in McrBC restriction in Escherichia coli K-12. J Bacteriol 179: 7768–7775. 10.1128/jb.179.24.7768-7775.1997

Bellamy SR, Milsom SE, Scott DJ, Daniels LE, Wilson GG, Halford SE. 2005. Cleavage of individual DNA strands by the different subunits of the heterodimeric restriction endonuclease BbvCI. J Mol Biol 348: 641–653. 10.1016/j.jmb.2005.02.035

Bellamy SR, Milsom SE, Kovacheva YS, Sessions RB, Halford SE. 2007. A switch in the mechanism of communication between the two DNA-binding sites in the SfiI restriction endonuclease. J Mol Biol 373: 1169–1183. 10.1016/j.jmb.2007.08.030

Bellamy SR, Mina P, Retter SE, Halford SE. 2008. Fidelity of DNA sequence recognition by the SfiI restriction endonuclease is determined by communications between its two DNA-binding sites. J Mol Biol 384: 557–563. 10.1016/j.jmb.2008.09.057

Bellamy SR, Kovacheva YS, Zulkipli IH, Halford SE. 2009. Differences between Ca2+ and Mg2+ in DNA binding and release by the SfiI restriction endonuclease: implications for DNA looping. Nucleic Acids Res 37: 5443–5453. 10.1093/nar/gkp569

Bhakta MS, Henry IM, Ousterout DG, Das KT, Lockwood SH, Meckler JF, Wallen MC, Zykovich A, Yu Y, Leo H, et al. 2013. Highly active zinc-finger nucleases by extended modular assembly. Genome Res 23: 530–538. 10.1101/gr.143693.112

Bianco PR, Hurley EM. 2005. The type I restriction endonuclease EcoR124I, couples ATP hydrolysis to bidirectional DNA translocation. J Mol Biol 352: 837–859. 10.1016/j.jmb.2005.07.055

Bibikova M, Golic M, Golic KG, Carroll D. 2002. Targeted chromosomal cleavage and mutagenesis in Drosophila using zinc-finger nucleases. Genetics 161: 1169–1175.

Bickle TA, Kruger DH. 1993. Biology of DNA restriction. Microbiol Rev 57: 434–450.

Biebricher A, Wende W, Escude C, Pingoud A, Desbiolles P. 2009. Tracking of single quantum dot labeled EcoRV sliding along DNA manipulated by double optical tweezers. Biophys J 96: L50–L52. 10.1016/j.bpj.2009.01.035

Bitinaite J, Schildkraut I. 2002. Self-generated DNA termini relax the specificity of SgrAI restriction endonuclease. Proc Natl Acad Sci 99: 1164–1169. 10.1073/pnas.022346799

Bitinaite J, Wah DA, Aggarwal AK, Schildkraut I. 1998. FokI dimerization is required for DNA cleavage. Proc Natl Acad Sci 95: 10570–10575. 10.1073/pnas.95.18.10570

Blackstock JJ, Egelhaaf SU, Atanasiu C, Dryden DT, Poon WC. 2001. Shape of Ocr, the gene 0.3 protein of bacteriophage T7: modeling based on light scattering experiments. Biochemistry 40: 9944–9949. 10.1021/bi010587+

Boch J, Scholze H, Schornack S, Landgraf A, Hahn S, Kay S, Lahaye T, Nickstadt A, Bonas U. 2009. Breaking the code of DNA binding specificity of TAL-type III effectors. Science 326: 1509–1512. 10.1126/science.1178811

Bochtler M, Szczepanowski RH, Tamulaitis G, Gražulis S, Czapinska H, Manakova E, Šikšnys V. 2006. Nucleotide flips determine the specificity of the Ecl18kI restriction endonuclease. EMBO J 25: 2219–2229. 10.1038/sj.emboj.7601096

Bogdanova E, Djordjevic M, Papapanagiotou I, Heyduk T, Kneale G, Severinov K. 2008. Transcription regulation of the type II restriction-modification system AhdI. Nucleic Acids Res 36: 1429–1442. 10.1093/nar/gkm1116

Bogdanova E, Zakharova M, Streeter S, Taylor J, Heyduk T, Kneale G, Severinov K. 2009. Transcription regulation of restriction-modification system Esp1396I. Nucleic Acids Res 37: 3354–3366. 10.1093/nar/gkp210

Bogdanove AJ, Bohm A, Mille JC, Morgan RD, Stoddard BL. 2018. Engineering altered protein–DNA recognition specificity. Nucleic Acids Res 46: 4845–4871. 10.1093/nar/gky289

Bonnet I, Biebricher A, Porte PL, Loverdo C, Benichou O, Voituriez R, Escude C, Wende W, Pingoud A, Desbiolles P. 2008. Sliding and jumping of single EcoRV restriction enzymes on non-cognate DNA. Nucleic Acids Res 36: 4118–4127. 10.1093/nar/gkn376

Borgaro JG, Zhu Z. 2013. Characterization of the 5-hydroxymethylcytosine-specific DNA restriction endonucleases. Nucleic Acids Res 41: 4198–4206. 10.1093/nar/gkt102

Borst P, Sabatini R. 2008. Base J: discovery, biosynthesis, and possible functions. Ann Rev Microbiol 62: 235–251. 10.1146/annurev.micro.62.081307.162750

Bourniquel AA, Bickle TA. 2002. Complex restriction enzymes: NTP-driven molecular motors. Biochimie 84: 1047–1059. 10.1016/S0300-9084(02)00020-2

Boyd AC, Charles IG, Keyte JW, Brammar WJ. 1986. Isolation and computer-aided characterization of MmeI, a type II restriction endonuclease from Methylophilus methylotrophus. Nucleic Acids Res 14: 5255–5274. 10.1093/nar/14.13.5255

Bozic D, Gražulis S, Šikšnys V, Huber R. 1996. Crystal structure of Citrobacter freundii restriction endonuclease Cfr10I at 2.15 A resolution. J Mol Biol 255: 176–186. 10.1006/jmbi.1996.0015

Brooks JE, Benner JS, Heiter DF, Silber KR, Sznyter LA, Jager-Quinton T, Moran LS, Slatko BE, Wilson GG, Nwankwo DO. 1989. Cloning the BamHI restriction modification system. Nucleic Acids Res 17: 979–997. 10.1093/nar/17.3.979

Brzezinski R, Piekarowicz A. 1982. Steps in the reaction mechanism of the Haemophilus influenzae Rf restriction endonuclease. J Mol Biol 154: 615–627. 10.1016/S0022-2836(82)80018-1

Bujnicki JM. 2004. Molecular phylogenetics of restriction endonucleases. In Restriction endonucleases (ed. Pingoud A), pp. 63–93. Springer, Berlin.

Bujnicki JM, Rychlewski L. 2001. Grouping together highly diverged PD-(D/E)XK nucleases and identification of novel superfamily members using structure-guided alignment of sequence profiles. J Mol Microbiol Biotechnol 3: 69–72.

Bujnicki JM, Radlinska M, Rychlewski L. 2000. Atomic model of the 5-methylcytosine-specific restriction enzyme McrA reveals an atypical zinc finger and structural similarity to ββαMe endonucleases. Mol Microbiol 37: 1280–1281. 10.1046/j.1365-2958.2000.02010.x

Bujnicki JM, Radlinska M, Rychlewski L. 2001. Polyphyletic evolution of type II restriction enzymes revisited: two independent sources of second-hand folds revealed. Trends Biochem Sci 26: 9–11. 10.1016/S0968-0004(00)01690-X

Bull JJ, Malik HS. 2017. The gene drive bubble: new realities. PLoS Genet 13: e1006850. 10.1371/journal.pgen.1006850

Burckhardt J, Weisemann J, Yuan R. 1981. Characterization of the DNA methylase activity of the restriction enzyme from Escherichia coli K. J Biol Chem 256: 4024–4032.

Burenina OY, Fedotova EA, Ryazanova AY, Protsenko AS, Zakharova MV, Karyagina AS, Solonin AS, Oretskaya TS, Kubareva EA. 2013. Peculiarities of the regulation of gene expression in the Ecl18kI restriction-modification system. Acta Nat 5: 70–80.

Butterer A, Pernstich C, Smith RM, Sobott F, Szczelkun MD, Toth J. 2014. Type III restriction endonucleases are heterotrimeric: comprising one helicase-nuclease subunit and a dimeric methyltransferase that binds only one specific DNA. Nucleic Acids Res 42: 5139–5150. 10.1093/nar/gku122

Cajthamlova K, Sisakova E, Weiser J, Weiserova M. 2007. Phosphorylation of Type IA restriction-modification complex enzyme EcoKI on the HsdR subunit. FEMS Microbiol Lett 270: 171–177. 10.1111/j.1574-6968.2007.00663.x

Calisto BM, Pich OQ, Pinol J, Fita I, Querol E, Carpena X. 2005. Crystal structure of a putative type I restriction-modification S subunit from Mycoplasma genitalium. J Mol Biol 351: 749–762. 10.1016/j.jmb.2005.06.050

Callahan SJ, Morgan RD, Jain R, Townson SA, Wilson GG, Roberts RJ, Aggarwal AK. 2011. Crystallization and preliminary crystallographic analysis of the type IIL restriction enzyme MmeI in complex with DNA. Acta Crystallogr, Sect F: Struct Biol Cryst Commun 67: 1262–1265. 10.1107/S1744309111028041

Callahan SJ, Luyten YA, Gupta YK, Wilson GG, Roberts RJ, Morgan RD, Aggarwal AK. 2016. Structure of Type IIL restriction-modification enzyme MmeI in complex with DNA has implications for engineering new specificities. PLoS Biol 14: e1002442. 10.1371/journal.pbio.1002442

Callow P, Sukhodub A, Taylor JE, Kneale GG. 2007. Shape and subunit organisation of the DNA methyltransferase M.AhdI by small-angle neutron scattering. J Mol Biol 369: 177–185. 10.1016/j.jmb.2007.03.012

Capoluongo E, Giglio A, Leonetti F, Belardi M, Giannetti A, Caprilli F, Ameglio F. 2000. DNA heterogeneity of Staphylococcus aureus strains evaluated by SmaI and SgrAI pulsed-field gel electrophoresis in patients with impetigo. Res Microbiol 151: 53–61. 10.1016/S0923-2508(00)00127-3

Carlson K, Raleigh EA, Hattman S. 1994. In Molecular biology of bacteriophage T4 (ed. Karam JD, et al.). American Society for Microbiology, Washington, DC.

Carroll D. 2011a. Genome engineering with zinc-finger nucleases. Genetics 188: 773–782. 10.1534/genetics.111.131433

Carroll D. 2011b. Zinc-finger nucleases: a panoramic view. Curr Gene Ther 11: 2–10. 10.2174/156652311794520076

Carroll D. 2014. Genome engineering with targetable nucleases. Ann Rev Biochem 83: 409–439. 10.1146/annurev-biochem-060713-035418

Carroll D, Beumer KJ. 2014. Genome engineering with TALENs and ZFNs: repair pathways and donor design. Methods 69: 137–141. 10.1016/j.ymeth.2014.03.026

Casadesus J, Low D. 2006. Epigenetic gene regulation in the bacterial world. Microbiol Mol Biol Rev 70: 830–856. 10.1128/MMBR.00016-06

Catto LE, Ganguly S, Milsom SE, Welsh AJ, Halford SE. 2006. Protein assembly and DNA looping by the FokI restriction endonuclease. Nucleic Acids Res 34: 1711–1720. 10.1093/nar/gkl076

Catto LE, Bellamy SR, Retter SE, Halford SE. 2008. Dynamics and consequences of DNA looping by the FokI restriction endonuclease. Nucleic Acids Res 36: 2073–2081. 10.1093/nar/gkn051

Cesnaviciene E, Mitkaite G, Stankevicius K, Janulaitis A, Lubys A. 2003. Esp1396I restriction-modification system: structural organization and mode of regulation. Nucleic Acids Res 31: 743–749. 10.1093/nar/gkg135

Champer J, Reeves R, Oh SY, Liu C, Liu J, Clark AG, Messer PW. 2017. Novel CRISPR/Cas9 gene drive constructs reveal insights into mechanisms of resistance allele formation and drive efficiency in genetically diverse populations. PLoS Genet 13: e1006796. 10.1371/journal.pgen.1006796

Chan SH, Stoddard BL, Xu SY. 2011. Natural and engineered nicking endonucleases—from cleavage mechanism to engineering of strand-specificity. Nucleic Acids Res 39: 1–18. 10.1093/nar/gkq742

Chand MK, Nirwan N, Diffin FM, van Aelst K, Kulkarni M, Pernstich C, Szczelkun MD, Saikrishnan K. 2015. Translocation-coupled DNA cleavage by the Type ISP restriction-modification enzymes. Nat Chem Biol 11: 870–877. 10.1038/nchembio.1926

Chandrasegaran S, Smith J. 1999. Chimeric restriction enzymes: what is next? Biol Chem 380: 841–848. 10.1515/BC.1999.103

Chandrashekaran S, Saravanan M, Radha DR, Nagaraja V. 2004. Ca2+-mediated site-specific DNA cleavage and suppression of promiscuous activity of KpnI restriction endonuclease. J Biol Chem 279: 49736–49740. 10.1074/jbc.M409483200

Chen K, Roberts GA, Stephanou AS, Cooper LP, White JH, Dryden DT. 2010. Fusion of GFP to the M·EcoKI DNA methyltransferase produces a new probe of Type I DNA restriction and modification enzymes. Biochem Biophys Res Commun 398: 254–259. 10.1016/j.bbrc.2010.06.069

Chen K, Reuter M, Sanghvi B, Roberts GA, Cooper LP, Tilling M, Blakely GW, Dryden DT. 2014. ArdA proteins from different mobile genetic elements can bind to the EcoKI Type I DNA methyltransferase of E. coli K12. Biochim Biophys Acta 1844: 505–511. 10.1016/j.bbapap.2013.12.008

Cheng X, Roberts RJ. 2001. AdoMet-dependent methylation, DNA methyltransferases and base flipping. Nucleic Acids Res 29: 3784–3795. 10.1093/nar/29.18.3784

Cheng X, Balendiran K, Schildkraut I, Anderson JE. 1994. Structure of PvuII endonuclease with cognate DNA. EMBO J 13: 3927–3935. 10.1002/j.1460-2075.1994.tb06708.x

Chilley PM, Wilkins BM. 1995. Distribution of the ardA family of antirestriction genes on conjugative plasmids. Microbiology 141 (Pt 9): 2157–2164. 10.1099/13500872-141-9-2157

Chuluunbaatar T, Ivanenko-Johnston T, Fuxreiter M, Meleshko R, Rasko T, Simon I, Heitman J, Kiss A. 2007. An EcoRI-RsrI chimeric restriction endonuclease retains parental sequence specificity. Biochim Biophys Acta 1774: 583–594. 10.1016/j.bbapap.2007.02.011

Clark TA, Murray IA, Morgan RD, Kislyuk AO, Spittle KE, Boitano M, Fomenkov A, Roberts RJ, Korlach J. 2012. Characterization of DNA methyltransferase specificities using single-molecule, real-time DNA sequencing. Nucleic Acids Res 40: e29. 10.1093/nar/gkr1146

Cohen-Karni D, Xu D, Apone L, Fomenkov A, Sun Z, Davis PJ, Kinney SR, Yamada-Mabuchi M, Xu SY, Davis T, et al. 2011. The MspJI family of modification-dependent restriction endonucleases for epigenetic studies. Proc Natl Acad Sci 108: 11040–11045. 10.1073/pnas.1018448108

Compe E, Egly JM. 2016. Nucleotide excision repair and transcriptional regulation: TFIIH and beyond. Ann Rev Biochem 85: 265–290. 10.1146/annurev-biochem-060815-014857

Cooper LP, Dryden DT. 1994. The domains of a type I DNA methyltransferase. Interactions and role in recognition of DNA methylation. J Mol Biol 236: 1011–1021. 10.1016/0022-2836(94)90008-6

Cooper LP, Roberts GA, White JH, Luyten YA, Bower EKM, Morgan RD, Roberts RJ, Lindsay JA, Dryden DTF. 2017. DNA target recognition domains in the Type I restriction and modification systems of Staphylococcus aureus. Nucleic Acids Res 45: 3395–3406. 10.1093/nar/gkx067

Crampton N, Roes S, Dryden DT, Rao DN, Edwardson JM, Henderson RM. 2007a. DNA looping and translocation provide an optimal cleavage mechanism for the type III restriction enzymes. EMBO J 26: 3815–3825. 10.1038/sj.emboj.7601807

Crampton N, Yokokawa M, Dryden DT, Edwardson JM, Rao DN, Takeyasu K, Yoshimura SH, Henderson RM. 2007b. Fast-scan atomic force microscopy reveals that the type III restriction enzyme EcoP15I is capable of DNA translocation and looping. Proc Natl Acad Sci 104: 12755–12760. 10.1073/pnas.0700483104

Csefalvay E, Lapkouski M, Guzanova A, Csefalvay L, Baikova T, Shevelev I, Bialevich V, Shamayeva K, Janscak P, Kuta Smatanova I, et al. 2015. Functional coupling of duplex translocation to DNA cleavage in a type I restriction enzyme. PloS One 10: e0128700. doi:10.1371/journal.pone.0128700. 10.1371/journal.pone.0128700

Cymerman IA, Obarska A, Skowronek KJ, Lubys A, Bujnicki JM. 2006. Identification of a new subfamily of HNH nucleases and experimental characterization of a representative member, HphI restriction endonuclease. Proteins 65: 867–876. 10.1002/prot.21156

Daniels LE, Wood KM, Scott DJ, Halford SE. 2003. Subunit assembly for DNA cleavage by restriction endonuclease SgrAI. J Mol Biol 327: 579–591. 10.1016/S0022-2836(03)00143-8

Davies GP, Martin I, Sturrock SS, Cronshaw A, Murray NE, Dryden DT. 1999. On the structure and operation of type I DNA restriction enzymes. J Mol Biol 290: 565–579. 10.1006/jmbi.1999.2908

De Bolle X, Bayliss CD, Field D, van de Ven T, Saunders NJ, Hood DW, Moxon ER. 2000. The length of a tetranucleotide repeat tract in Haemophilus influenzae determines the phase variation rate of a gene with homology to type III DNA methyltransferases. Mol Microbiol 35: 211–222. 10.1046/j.1365-2958.2000.01701.x

Dedkov VS, Degtyarev SK. 1998. Actinobacillus and Streptococcus: producers of isoschizomers of the restriction endonucleases R.HphI, R.SauI, R.NheI, R.MboI and R.SwaI. Biol Chem 379: 573–574.

Deibert M, Gražulis S, Sasnauskas G, Šikšnys V, Huber R. 2000. Structure of the tetrameric restriction endonuclease NgoMIV in complex with cleaved DNA. Nat Struct Biol 7: 792–799. 10.1038/79032

Deng D, Yan C, Pan X, Mahfouz M, Wang J, Zhu JK, Shi Y, Yan N. 2012. Structural basis for sequence-specific recognition of DNA by TAL effectors. Science 335: 720–723. 10.1126/science.1215670

Denjmukhametov MM, Brevnov MG, Zakharova MV, Repyk AV, Solonin AS, Petrauskene OV, Gromova ES. 1998. The Ecl18kI restriction-modification system: cloning, expression, properties of the purified enzymes. FEBS Lett 433: 233–236. 10.1016/S0014-5793(98)00921-1

Dixon K, Bayliss CD, Makepeace K, Moxon ER, Hood DW. 2007. Identification of the functional initiation codons of a phase-variable gene of Haemophilus influenzae, lic2A, with the potential for differential expression. J Bacteriol 189: 511–521. 10.1128/JB.00815-06

Dreyer AK, Hoffmann D, Lachmann N, Ackermann M, Steinemann D, Timm B, Siler U, Reichenbach J, Grez M, Moritz T, et al. 2015. TALEN-mediated functional correction of X-linked chronic granulomatous disease in patient-derived induced pluripotent stem cells. Biomaterials 69: 191–200. 10.1016/j.biomaterials.2015.07.057

Dryden DTF. 1999. Bacterial methyltransferases. In S-adenosylmethionine-dependent methyltransferases: structures and functions (ed. Cheng X, Blumenthal RM), pp. 283–340. World Scientific, Singapore.

Dryden DT. 2006. DNA mimicry by proteins and the control of enzymatic activity on DNA. Trends Biotechnol 24: 378–382. 10.1016/j.tibtech.2006.06.004

Dryden DT, Cooper LP, Thorpe PH, Byron O. 1997. The in vitro assembly of the EcoKI type I DNA restriction/modification enzyme and its in vivo implications. Biochemistry 36: 1065–1076. 10.1021/bi9619435

Dryden DT, Edwardson JM, Henderson RM. 2011. DNA translocation by type III restriction enzymes: a comparison of current models of their operation derived from ensemble and single-molecule measurements. Nucleic Acids Res 39: 4525–4531. 10.1093/nar/gkq1285

Dunn DB, Smith JD. 1955a. The occurrence of 6-methylaminopurine in microbial deoxyribonucleic acids. Biochem J 60: pxvii.

Dunn DB, Smith JD. 1955b. Occurrence of a new base in the deoxyribonucleic acid of a strain of Bacterium coli. Nature 175: 336–337. 10.1038/175336a0

Dunn JJ, Studier FW. 1981. Nucleotide sequence from the genetic left end of bacteriophage T7 DNA to the beginning of gene 4. J Mol Biol 148: 303–330. 10.1016/0022-2836(81)90178-9

Dunn JJ, Elzinga M, Mark KK, Studier FW. 1981. Amino acid sequence of the gene 0.3 protein of bacteriophage T7 and nucleotide sequence of its mRNA. J Biol Chem 256: 2579–2585.

Dunten PW, Little EJ, Gregory MT, Manohar VM, Dalton M, Hough D, Bitinaite J, Horton NC. 2008. The structure of SgrAI bound to DNA; recognition of an 8 base pair target. Nucleic Acids Res 36: 5405–5416. 10.1093/nar/gkn510

Dunten PW, Little EJ, Horton NC. 2009. The restriction enzyme SgrAI: structure solution via combination of poor MIRAS and MR phases. Acta Crystallogr, Sect D: Biol Crystallogr 65: 393–398. 10.1107/S0907444909003266

Durai S, Mani M, Kandavelou K, Wu J, Porteus MH, Chandrasegaran S. 2005. Zinc finger nucleases: custom-designed molecular scissors for genome engineering of plant and mammalian cells. Nucleic Acids Res 33: 5978–5990. 10.1093/nar/gki912

Dybvig K, Sitaraman R, French CT. 1998. A family of phase-variable restriction enzymes with differing specificities generated by high-frequency gene rearrangements. Proc Natl Acad Sci 95: 13923–13928. 10.1073/pnas.95.23.13923

Ehrlich M, Gama-Sosa MA, Carreira LH, Ljungdahl LG, Kuo KC, Gehrke CW. 1985. DNA methylation in thermophilic bacteria: N4-methylcytosine, 5-methylcytosine, and N6-methyladenine. Nucleic Acids Res 13: 1399–1412. 10.1093/nar/13.4.1399

Ehrlich M, Wilson GG, Kuo KC, Gehrke CW. 1987. N4-methylcytosine as a minor base in bacterial DNA. J Bacteriol 169: 939–943. 10.1128/jb.169.3.939-943.1987

Eid J, Fehr A, Gray J, Luong K, Lyle J, Otto G, Peluso P, Rank D, Baybayan P, Bettman B, et al. 2009. Real-time DNA sequencing from single polymerase molecules. Science 323: 133–138. 10.1126/science.1162986

Embleton ML, Šikšnys V, Halford SE. 2001. DNA cleavage reactions by type II restriction enzymes that require two copies of their recognition sites. J Mol Biol 311: 503–514. 10.1006/jmbi.2001.4892

Embleton ML, Vologodskii AV, Halford SE. 2004. Dynamics of DNA loop capture by the SfiI restriction endonuclease on supercoiled and relaxed DNA. J Mol Biol 339: 53–66. 10.1016/j.jmb.2004.03.046

Endlich B, Linn S. 1985. The DNA restriction endonuclease of Escherichia coli B. I. Studies of the DNA translocation and the ATPase activities. J Biol Chem 260: 5720–5728.

Fairman-Williams ME, Guenther UP, Jankowsky E. 2010. SF1 and SF2 helicases: family matters. Curr Opin Struct Biol 20: 313–324. 10.1016/j.sbi.2010.03.011

Fedotova EA, Protsenko AS, Zakharova MV, Lavrova NV, Alekseevsky AV, Oretskaya TS, Karyagina AS, Solonin AS, Kubareva EA. 2009. SsoII-like DNA-methyltransferase Ecl18kI: interaction between regulatory and methylating functions. Biochem Biokhim 74: 85–91. 10.1134/S0006297909010131

Flusberg BA, Webster DR, Lee JH, Travers KJ, Olivares EC, Clark TA, Korlach J, Turner SW. 2010. Direct detection of DNA methylation during single-molecule, real-time sequencing. Nat Meth 7: 461–465. 10.1038/nmeth.1459

Fox KL, Dowideit SJ, Erwin AL, Srikhanta YN, Smith AL, Jennings MP. 2007. Haemophilus influenzae phasevarions have evolved from type III DNA restriction systems into epigenetic regulators of gene expression. Nucleic Acids Res 35: 5242–5252. 10.1093/nar/gkm571

Freitag M, Selker EU. 2005. Controlling DNA methylation: many roads to one modification. Curr Opin Genet Dev 15: 191–199. 10.1016/j.gde.2005.02.003

Friedhoff P, Franke I, Meiss G, Wende W, Krause KL, Pingoud A. 1999. A similar active site for non-specific and specific endonucleases. Nat Struct Biol 6: 112–113. 10.1038/5796

Friedrich T, Fatemi M, Gowhar H, Leismann O, Jeltsch A. 2000. Specificity of DNA binding and methylation by the M.FokI DNA methyltransferase. Biochim Biophys Acta 1480: 145–159. 10.1016/S0167-4838(00)00065-0

Furuta Y, Kawai M, Uchiyama I, Kobayashi I. 2011. Domain movement within a gene: a novel evolutionary mechanism for protein diversification. PloS One 6: e18819. 10.1371/journal.pone.0018819

Gabriel R, Lombardo A, Arens A, Miller JC, Genovese P, Kaeppel C, Nowrouzi A, Bartholomae CC, Wang J, Friedman G, et al. 2011. An unbiased genome-wide analysis of zinc-finger nuclease specificity. Nat Biotechnol 29: 816–823. 10.1038/nbt.1948

Gabsalilow L, Schierling B, Friedhoff P, Pingoud A, Wende W. 2013. Site- and strand-specific nicking of DNA by fusion proteins derived from MutH and I-SceI or TALE repeats. Nucleic Acids Res 41: e83. doi:10.1093/nar/gkt1080. 10.1093/nar/gkt080

Galburt EA, Chevalier B, Tang W, Jurica MS, Flick KE, Monnat RJ Jr, Stoddard BL. 1999. A novel endonuclease mechanism directly visualized for I-PpoI. Nat Struct Biol 6: 1096–1099. 10.1038/70027

Galetto R, Duchateau P, Paques F. 2009. Targeted approaches for gene therapy and the emergence of engineered meganucleases. Expert Opin Biol Ther 9: 1289–1303. 10.1517/14712590903213669

Garcia LR, Molineux IJ. 1996. Transcription-independent DNA translocation of bacteriophage T7 DNA into Escherichia coli. J Bacteriol 178: 6921–6929. 10.1128/jb.178.23.6921-6929.1996

Gasiunas G, Sasnauskas G, Tamulaitis G, Urbanke C, Razaniene D, Šikšnys V. 2008. Tetrameric restriction enzymes: expansion to the GIY-YIG nuclease family. Nucleic Acids Res 36: 938–949. 10.1093/nar/gkm1090

Gast FU, Brinkmann T, Pieper U, Kruger T, Noyer-Weidner M, Pingoud A. 1997. The recognition of methylated DNA by the GTP-dependent restriction endonuclease McrBC resides in the N-terminal domain of McrB. Biol Chem 378: 975–982. 10.1515/bchm.1997.378.9.975

Gefter M, Hausmann R, Gold M, Hurwitz J. 1966. The enzymatic methylation of ribonucleic acid and deoxyribonucleic acid. X. Bacteriophage T3-induced S-adenosylmethionine cleavage. J Biol Chem 241: 1995–2006.

Gemmen GJ, Millin R, Smith DE. 2006. Tension-dependent DNA cleavage by restriction endonucleases: two-site enzymes are “switched off” at low force. Proc Natl Acad Sci 103: 11555–11560. 10.1073/pnas.0604463103

Gilmore JL, Suzuki Y, Tamulaitis G, Šikšnys V, Takeyasu K, Lyubchenko YL. 2009. Single-molecule dynamics of the DNA-EcoRII protein complexes revealed with high-speed atomic force microscopy. Biochemistry 48: 10492–10498. 10.1021/bi9010368

Glover SW, Piekarowicz A. 1972. Host specificity of DNA in Haemophilus influenzae: restriction and modification in strain Rd. Biochem Biophys Res Commun 46: 1610–1617. 10.1016/0006-291X(72)90793-0

Golovenko D, Manakova E, Tamulaitienė G, Gražulis S, Šikšnys V. 2009. Structural mechanisms for the 5′-CCWGG sequence recognition by the N- and C-terminal domains of EcoRII. Nucleic Acids Res 37: 6613–6624. 10.1093/nar/gkp699

Golovenko D, Manakova E, Zakrys L, Zaremba M, Sasnauskas G, Gražulis S, Šikšnys V. 2014. Structural insight into the specificity of the B3 DNA-binding domains provided by the co-crystal structure of the C-terminal fragment of BfiI restriction enzyme. Nucleic Acids Res 42: 4113–4122. 10.1093/nar/gkt1368

González-Cerón G, Miranda-Olivares OJ, Servin-González L. 2009. Characterization of the methyl-specific restriction system of Streptomyces coelicolor A3(2) and of the role played by laterally acquired nucleases. FEMS Microbiol Lett 301: 35–43. 10.1111/j.1574-6968.2009.01790.x

Goszczynski B, McGhee JD. 1991. Resolution of sequencing ambiguities: a universal FokI adapter permits Maxam-Gilbert re-sequencing of single-stranded phagemid DNA. Gene 104: 71–74. 10.1016/0378-1119(91)90466-O

Gowers DM, Halford SE. 2003. Protein motion from non-specific to specific DNA by three-dimensional routes aided by supercoiling. EMBO J 22: 1410–1418. 10.1093/emboj/cdg125

Gowers DM, Bellamy SR, Halford SE. 2004. One recognition sequence, seven restriction enzymes, five reaction mechanisms. Nucleic Acids Res 32: 3469–3479. 10.1093/nar/gkh685

Gowers DM, Wilson GG, Halford SE. 2005. Measurement of the contributions of 1D and 3D pathways to the translocation of a protein along DNA. Proc Natl Acad Sci 102: 15883–15888. 10.1073/pnas.0505378102

Gražulis S, Deibert M, Rimseliene R, Skirgaila R, Sasnauskas G, Lagunavicius A, Repin V, Urbanke C, Huber R, Šikšnys V. 2002. Crystal structure of the Bse634I restriction endonuclease: comparison of two enzymes recognizing the same DNA sequence. Nucleic Acids Res 30: 876–885. 10.1093/nar/30.4.876

Gražulis S, Manakova E, Roessle M, Bochtler M, Tamulaitienė G, Huber R, Šikšnys V. 2005. Structure of the metal-independent restriction enzyme BfiI reveals fusion of a specific DNA-binding domain with a nonspecific nuclease. Proc Natl Acad Sci 102: 15797–15802. 10.1073/pnas.0507949102

Gubler M, Bickle TA. 1991. Increased protein flexibility leads to promiscuous protein–DNA interactions in type IC restriction-modification systems. EMBO J 10: 951–957. 10.1002/j.1460-2075.1991.tb08029.x

Guilinger JP, Pattanayak V, Reyon D, Tsai SQ, Sander JD, Joung JK, Liu DR. 2014a. Broad specificity profiling of TALENs results in engineered nucleases with improved DNA-cleavage specificity. Nat Methods 11: 429–435. 10.1038/nmeth.2845

Guilinger JP, Thompson DB, Liu DR. 2014b. Fusion of catalytically inactive Cas9 to FokI nuclease improves the specificity of genome modification. Nat Biotechnol 32: 577–582. 10.1038/nbt.2909

Guo J, Gaj T, Barbas CF 3rd. 2010. Directed evolution of an enhanced and highly efficient FokI cleavage domain for zinc finger nucleases. J Mol Biol 400: 96–107. 10.1016/j.jmb.2010.04.060

Gupta R, Nagarajan A, Wajapeyee N. 2010. Advances in genome-wide DNA methylation analysis. BioTechniques 49: ii–ixi.

Gupta YK, Yang L, Chan SH, Samuelson JC, Xu SY, Aggarwal AK. 2012. Structural insights into the assembly and shape of Type III restriction-modification (R-M) EcoP15I complex by small-angle X-ray scattering. J Mol Biol 420: 261–268. 10.1016/j.jmb.2012.04.026

Gupta YK, Chan SH, Xu SY, Aggarwal AK. 2015. Structural basis of asymmetric DNA methylation and ATP-triggered long-range diffusion by EcoP15I. Nat Commun 6: 7363. doi: 10.1038/ncomms8363. 10.1038/ncomms8363

Halford SE. 2001. Hopping, jumping and looping by restriction enzymes. Biochem Soc Trans 29: 363–374. 10.1042/bst0290363

Halford SE. 2013. http://library.cshl.edu/Meetings/restriction-enzymes/Halford.php.

Halford SE, Marko JF. 2004. How do site-specific DNA-binding proteins find their targets? Nucleic Acids Res 32: 3040–3052. 10.1093/nar/gkh624

Halford SE, Bilcock DT, Stanford NP, Williams SA, Milsom SE, Gormley NA, Watson MA, Bath AJ, Embleton ML, Gowers DM, et al. 1999. Restriction endonuclease reactions requiring two recognition sites. Biochem Soc Trans 27: 696–699. 10.1042/bst0270696

Halford SE, Welsh AJ, Szczelkun MD. 2004. Enzyme-mediated DNA looping. Ann Rev Biophys Biomol Struct 33: 1–24. 10.1146/annurev.biophys.33.110502.132711

Halford SE, Catto LE, Pernstich C, Rusling DA, Sanders KL. 2011. The reaction mechanism of FokI excludes the possibility of targeting zinc finger nucleases to unique DNA sites. Biochem Soc Trans 39: 584–588. 10.1042/BST0390584

Hallet B. 2001. Playing Dr. Jekyll and Mr. Hyde: combined mechanisms of phase variation in bacteria. Curr Opin Microbiol 4: 570–581. 10.1016/S1369-5274(00)00253-8

Handel EM, Cathomen T. 2011. Zinc-finger nuclease based genome surgery: it's all about specificity. Curr Gene Ther 11: 28–37. 10.2174/156652311794520120

Handel EM, Gellhaus K, Khan K, Bednarski C, Cornu TI, Muller-Lerch F, Kotin RM, Heilbronn R, Cathomen T. 2012. Versatile and efficient genome editing in human cells by combining zinc-finger nucleases with adeno-associated viral vectors. Hum Gene Ther 23: 321–329. 10.1089/hum.2011.140

Hasan N, Kim SC, Podhajska AJ, Szybalski W. 1986. A novel multistep method for generating precise unidirectional deletions using BspMI, a class-IIS restriction enzyme. Gene 50: 55–62. 10.1016/0378-1119(86)90309-4

Hausmann R. 1967. Synthesis of an S-adenosylmethionine-cleaving enzyme in T3-infected Escherichia coli and its disturbance by co-infection with enzymatically incompetent bacteriophage. J Virol 1: 57–63. 10.1099/0022-1317-1-1-57

Hausmann R, Messerschmid M. 1988. Inhibition of gene expression of T7-related phages by prophage P1. Mol Gen Genet 212: 543–547. 10.1007/BF00330862

Hausmann R, Messerschmid M. 1988. The T7 group. In The bacteriophages (ed. Calendar R), Vol.1, 259–290. Plenum, New York.

He X, Hull V, Thomas JA, Fu X, Gidwani S, Gupta YK, Black LW, Xu SY. 2015. Expression and purification of a single-chain Type IV restriction enzyme Eco94GmrSD and determination of its substrate preference. Sci Rep 5: 9747. doi:9710.1038/srep09747. 10.1038/srep09747

Hegde SS, Vetting MW, Roderick SL, Mitchenall LA, Maxwell A, Takiff HE, Blanchard JS. 2005. A fluoroquinolone resistance protein from Mycobacterium tuberculosis that mimics DNA. Science 308: 1480–1483. 10.1126/science.1110699

Heidmann S, Seifert W, Kessler C, Domdey H. 1989. Cloning, characterization and heterologous expression of the SmaI restriction-modification system. Nucleic Acids Res 17: 9783–9796. 10.1093/nar/17.23.9783

Heiter DF, Lunnen KD, Wilson GG. 2005. Site-specific DNA-nicking mutants of the heterodimeric restriction endonuclease R·BbvCI. J Mol Biol 348: 631–640. 10.1016/j.jmb.2005.02.034

Heiter DF, Lunnen KD, Morgan RD, Wilson GG. 2015. When less is more: AspCNI and other REases that cleave poorly at high concentrations. In 7th NEB meeting on restriction and modification, Gdansk: Poster P20.

Heitman J, Model P. 1987. Site-specific methylases induce the SOS DNA repair response in Escherichia coli. J Bacteriol 169: 3243–3250. 10.1128/jb.169.7.3243-3250.1987

Heitman J, Zinder ND, Model P. 1989. Repair of the Escherichia coli chromosome after in vivo scission by the EcoRI endonuclease. Proc Natl Acad Sci 86: 2281–2285. 10.1073/pnas.86.7.2281

Hendel A, Fine EJ, Bao G, Porteus MH. 2015. Quantifying on- and off-target genome editing. Trends Biotechnol 33: 132–140. 10.1016/j.tibtech.2014.12.001

Hendrickson PG, Cairns BR. 2016. Tet proteins enhance the developmental hourglass. Nat Genet 48: 345–347. 10.1038/ng.3533

Hershey AD, Dixon J, Chase M. 1953. Nucleic acid economy in bacteria infected with bacteriophage T2. I. Purine and pyrimidine composition. J Gen Physiol 36: 777–789. 10.1085/jgp.36.6.777

Hien le T, Zatsepin TS, Schierling B, Volkov EM, Wende W, Pingoud A, Kubareva EA, Oretskaya TS. 2011. Restriction endonuclease SsoII with photoregulated activity—a “molecular gate” approach. Bioconjug Chem 22: 1366–1373. 10.1021/bc200063m

Higgins LS, Besnier C, Kong H. 2001. The nicking endonuclease N·BstNBI is closely related to type IIs restriction endonucleases MlyI and PleI. Nucleic Acids Res 29: 2492–2501. 10.1093/nar/29.12.2492

Hingorani-Varma K, Bitinaite J. 2003. Kinetic analysis of the coordinated interaction of SgrAI restriction endonuclease with different DNA targets. J Biol Chem 278: 40392–40399. 10.1074/jbc.M304603200

Hirsch JA, Wah DA, Dorner LF, Schildkraut I, Aggarwal AK. 1997. Crystallization and preliminary X-ray analysis of restriction endonuclease FokI bound to DNA. FEBS Lett 403: 136–138. 10.1016/S0014-5793(97)00039-2

Hockemeyer D, Wang H, Kiani S, Lai CS, Gao Q, Cassady JP, Cost GJ, Zhang L, Santiago Y, Miller JC, et al. 2011. Genetic engineering of human pluripotent cells using TALE nucleases. Nat Biotechnol 29: 731–734. 10.1038/nbt.1927

Horton NC. 2015. Allosteric control of DNA cleavage rate and DNA sequence specificity via run-on oligomerization. In 7th NEB meeting on restriction and modification (Gdansk), Talk T2.

Horton JR, Blumenthal RM, Cheng X. 2004. Restriction endonucleases: structure of the conserved catalytic core and the role of metal ions in DNA cleavage. In Restriction endonucleases (ed. Pingoud A), pp. 361–392. Springer, Berlin.

Horton JR, Zhang X, Maunus R, Yang Z, Wilson GG, Roberts RJ, Cheng X. 2006. DNA nicking by HinP1I endonuclease: bending, base flipping and minor groove expansion. Nucleic Acids Res 34: 939–948. 10.1093/nar/gkj484

Horton JR, Mabuchi MY, Cohen-Karni D, Zhang X, Griggs RM, Samaranayake M, Roberts RJ, Zheng Y, Cheng X. 2012. Structure and cleavage activity of the tetrameric MspJI DNA modification-dependent restriction endonuclease. Nucleic Acids Res 40: 9763–9773. 10.1093/nar/gks719

Horton JR, Borgaro JG, Griggs RM, Quimby A, Guan S, Zhang X, Wilson GG, Zheng Y, Zhu Z, Cheng X. 2014a. Structure of 5-hydroxymethylcytosine-specific restriction enzyme, AbaSI, in complex with DNA. Nucleic Acids Res 42: 7947–7959. 10.1093/nar/gku497

Horton JR, Nugent RL, Li A, Mabuchi MY, Fomenkov A, Cohen-Karni D, Griggs RM, Zhang X, Wilson GG, Zheng Y, et al. 2014b. Structure and mutagenesis of the DNA modification-dependent restriction endonuclease AspBHI. Sci Rep 4: 4246. doi:10.1038/srep04246. 10.1038/srep04246

Horton JR, Wang H, Mabuchi MY, Zhang X, Roberts RJ, Zheng Y, Wilson GG, Cheng X. 2014c. Modification-dependent restriction endonuclease, MspJI, flips 5-methylcytosine out of the DNA helix. Nucleic Acids Res 42: 12092–12101. 10.1093/nar/gku871

Huai Q, Colandene JD, Chen Y, Luo F, Zhao Y, Topal MD, Ke H. 2000. Crystal structure of NaeI-an evolutionary bridge between DNA endonuclease and topoisomerase. EMBO J 19: 3110–3118. 10.1093/emboj/19.12.3110

Ibryashkina EM, Zakharova MV, Baskunov VB, Bogdanova ES, Nagornykh MO, Den'mukhamedov MM, Melnik BS, Kolinski A, Gront D, Feder M, et al. 2007. Type II restriction endonuclease R.Eco29kI is a member of the GIY-YIG nuclease superfamily. BMC Struct Biol 7: 48. 10.1186/1472-6807-7-48

Imanishi M, Negi S, Sugiura Y. 2010. Non-FokI-based zinc finger nucleases. Meth Mol Biol 649: 337–349. 10.1007/978-1-60761-753-2_21

Interthal H, Pouliot JJ, Champoux JJ. 2001. The tyrosyl-DNA phosphodiesterase Tdp1 is a member of the phospholipase D superfamily. Proc Natl Acad Sci 98: 12009–12014. 10.1073/pnas.211429198

Ishikawa K, Watanabe M, Kuroita T, Uchiyama I, Bujnicki JM, Kawakami B, Tanokura M, Kobayashi I. 2005. Discovery of a novel restriction endonuclease by genome comparison and application of a wheat-germ-based cell-free translation assay: PabI (5′-GTA/C) from the hyperthermophilic archaeon Pyrococcus abyssi. Nucleic Acids Res 33: e112. doi:110.1093/nar/gni1113. 10.1093/nar/gni113

Ishikawa K, Handa N, Kobayashi I. 2009. Cleavage of a model DNA replication fork by a Type I restriction endonuclease. Nucleic Acids Res 37: 3531–3544. 10.1093/nar/gkp214

Ives CL, Nathan PD, Brooks JE. 1992. Regulation of the BamHI restriction-modification system by a small intergenic open reading frame, bamHIC, in both Escherichia coli and Bacillus subtilis. J Bacteriol 174: 7194–7201. 10.1128/jb.174.22.7194-7201.1992

Iyer LM, Tahiliani M, Rao A, Aravind L. 2009. Prediction of novel families of enzymes involved in oxidative and other complex modifications of bases in nucleic acids. Cell Cycle 8: 1698–1710. 10.4161/cc.8.11.8580

Jablonska E, Kauc L, Piekarowicz A. 1975. An Haemophilus influenzae mutant which inhibits the growth of HP1c1 phage. Mol Gen Genet 139: 157–166. 10.1007/BF00264695

Jain R, Poulos MG, Gros J, Chakravarty AK, Shuman S. 2011. Substrate specificity and mutational analysis of Kluyveromyces lactis γ-toxin, a eukaryal tRNA anticodon nuclease. RNA 17: 1336–1343. 10.1261/rna.2722711

Janosi L, Yonemitsu H, Hong H, Kaji A. 1994. Molecular cloning and expression of a novel hydroxymethylcytosine-specific restriction enzyme (PvuRts1I) modulated by glucosylation of DNA. J Mol Biol 242: 45–61. 10.1006/jmbi.1994.1556

Janscak P, Bickle TA. 1998. The DNA recognition subunit of the type IB restriction-modification enzyme EcoAI tolerates circular permutions of its polypeptide chain. J Mol Biol 284: 937–948. 10.1006/jmbi.1998.2250

Janscak P, Dryden DT, Firman K. 1998. Analysis of the subunit assembly of the type IC restriction-modification enzyme EcoR124I. Nucleic Acids Res 26: 4439–4445. 10.1093/nar/26.19.4439

Janulaitis A, Klimasauskas S, Petrusyte M, Butkus V. 1983. Cytosine modification in DNA by BcnI methylase yields N4-methylcytosine. FEBS Lett 161: 131–134. 10.1016/0014-5793(83)80745-5

Janulaitis A, Petrusyte M, Maneliene Z, Klimasauskas S, Butkus V. 1992a. Purification and properties of the Eco57I restriction endonuclease and methylase—prototypes of a new class (type IV). Nucleic Acids Res 20: 6043–6049. 10.1093/nar/20.22.6043

Janulaitis A, Vaisvila R, Timinskas A, Klimasauskas S, Butkus V. 1992b. Cloning and sequence analysis of the genes coding for Eco57I type IV restriction-modification enzymes. Nucleic Acids Res 20: 6051–6056. 10.1093/nar/20.22.6051

Jeffery CJ. 2016. Protein species and moonlighting proteins: Very small changes in a protein's covalent structure can change its biochemical function. J Proteomics 134: 19–24. 10.1016/j.jprot.2015.10.003

Jeffreys AJ. 2006. See, e.g., https://www.knaw.nl/en/awards/laureates/dr-h-p-heinekenprijs-voor-biochemie-en-biofysica/alec-j-jeffreys-1950-groot-brittannia; https://www.ncbi.nlm.nih.gov/pubmed/20011117 (interview with Jane Gitschier 2009); https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3831583/pdf/2041-2223-4-21.pdf (2013).

Jeltsch A, Wenz C, Wende W, Selent U, Pingoud A. 1996. Engineering novel restriction endonucleases: principles and applications. Trends Biotechnol 14: 235–238. 10.1016/0167-7799(96)10030-5

Jeschke J, Collignon E, Fuks F. 2016. Portraits of TET-mediated DNA hydroxymethylation in cancer. Curr Opin Genet Dev 36: 16–26. 10.1016/j.gde.2016.01.004

Ji M, Tan L, Jen-Jacobson L, Saxena S. 2014. Insights into copper coordination in the EcoRI-DNA complex by ESR spectroscopy. Mol Phys 112: 3173–3182. 10.1080/00268976.2014.934313

Johnson TB, Coghill RD. 1925. Researches on pyrimidines. C111. The discovery of S-methyl-cytosine in tuberculinic acid, the nucleic acid of the tubercle bacillus. J Am Chem Soc 47(11): 2838–2844. 10.1021/ja01688a030

Joung JK, Sander JD. 2013. TALENs: a widely applicable technology for targeted genome editing. Nat Rev Mol Cell Biol 14: 49–55. 10.1038/nrm3486

Jurenaite-Urbanaviciene S, Serksnaite J, Kriukiene E, Giedriene J, Venclovas C, Lubys A. 2007. Generation of DNA cleavage specificities of type II restriction endonucleases by reassortment of target recognition domains. Proc Natl Acad Sci 104: 10358–10363. 10.1073/pnas.0610365104

Jurica MS, Stoddard BL. 1999. Homing endonucleases: structure, function and evolution. Cell Mol Life Sci 55: 1304–1326. 10.1007/s000180050372

Kachalova GS, Rogulin EA, Yunusova AK, Artyukh RI, Perevyazova TA, Matvienko NI, Zheleznaya LA, Bartunik HD. 2008. Structural analysis of the heterodimeric type IIS restriction endonuclease R.BspD6I acting as a complex between a monomeric site-specific nickase and a catalytic subunit. J Mol Biol 384: 489–502. 10.1016/j.jmb.2008.09.033

Kaczorowski T, Skowron P, Podhajska AJ. 1989. Purification and characterization of the FokI restriction endonuclease. Gene 80: 209–216. 10.1016/0378-1119(89)90285-0

Kaminska KH, Bujnicki JM. 2008. Bacteriophage Mu Mom protein responsible for DNA modification is a new member of the acyltransferase superfamily. Cell Cycle 7: 120–121. 10.4161/cc.7.1.5158

Kaminska KH, Kawai M, Boniecki M, Kobayashi I, Bujnicki JM. 2008. Type II restriction endonuclease R.Hpy188I belongs to the GIY-YIG nuclease superfamily, but exhibits an unusual active site. BMC Struct Biol 8: 48. doi:10.1186/1472-6807-1188-1148. 10.1186/1472-6807-8-48

Kannan P, Cowan GM, Daniel AS, Gann AA, Murray NE. 1989. Conservation of organization in the specificity polypeptides of two families of type I restriction enzymes. J Mol Biol 209: 335–344. 10.1016/0022-2836(89)90001-6

Kanwar N, Roberts GA, Cooper LP, Stephanou AS, Dryden DT. 2016. The evolutionary pathway from a biologically inactive polypeptide sequence to a folded, active structural mimic of DNA. Nucleic Acids Res 44: 4289–4303. 10.1093/nar/gkw234

Karyagina A, Shilov I, Tashlitskii V, Khodoun M, Vasil'ev S, Lau PC, Nikolskaya I. 1997. Specific binding of sso II DNA methyltransferase to its promoter region provides the regulation of Sso II restriction-modification gene expression. Nucleic Acids Res 25: 2114–2120. 10.1093/nar/25.11.2114

Kauc L, Piekarowicz A. 1978. Purification and properties of a new restriction endonuclease from Haemophilus influenzae Rf. Eur J Biochem/FEBS 92: 417–426. 10.1111/j.1432-1033.1978.tb12762.x

Kaufmann G. 2000. Anticodon nucleases. Trends Biochem Sci 25: 70–74. 10.1016/S0968-0004(99)01525-X

Kaus-Drobek M, Czapinska H, Sokolowska M, Tamulaitis G, Szczepanowski RH, Urbanke C, Šikšnys V, Bochtler M. 2007. Restriction endonuclease MvaI is a monomer that recognizes its target sequence asymmetrically. Nucleic Acids Res 35: 2035–2046. 10.1093/nar/gkm064

Kazrani AA, Kowalska M, Czapinska H, Bochtler M. 2014. Crystal structure of the 5hmC specific endonuclease PvuRts1I. Nucleic Acids Res 42: 5929–5936. 10.1093/nar/gku186

Kelch BA. 2016. Review: The lord of the rings: Structure and mechanism of the sliding clamp loader. Biopolymers 105: 532–546. 10.1002/bip.22827

Kelleher JE, Raleigh EA. 1991. A novel activity in Escherichia coli K-12 that directs restriction of DNA modified at CG dinucleotides. J Bacteriol 173: 5220–5223. 10.1128/jb.173.16.5220-5223.1991

Kelleher JE, Daniel AS, Murray NE. 1991. Mutations that confer de novo activity upon a maintenance methyltransferase. J Mol Biol 221: 431–440. 10.1016/0022-2836(91)80064-2

Kennaway CK, Obarska-Kosinska A, White JH, Tuszynska I, Cooper LP, Bujnicki JM, Trinick J, Dryden DT. 2009. The structure of M.EcoKI Type I DNA methyltransferase with a DNA mimic antirestriction protein. Nucleic Acids Res 37: 762–770. 10.1093/nar/gkn988

Kennaway CK, Taylor JE, Song CF, Potrzebowski W, Nicholson W, White JH, Swiderska A, Obarska-Kosinska A, Callow P, Cooper LP, et al. 2012. Structure and operation of the DNA-translocating type I DNA restriction enzymes. Genes Dev 26: 92–104. 10.1101/gad.179085.111

Khan F, Furuta Y, Kawai M, Kaminska KH, Ishikawa K, Bujnicki JM, Kobayashi I. 2010. A putative mobile genetic element carrying a novel type IIF restriction-modification system (PluTI). Nucleic Acids Res 38: 3019–3030. 10.1093/nar/gkp1221

Kim SC, Podhajska AJ, Szybalski W. 1988. Cleaving DNA at any predetermined site with adapter-primers and class-IIS restriction enzymes. Science 240: 504–506. 10.1126/science.2833816

Kim YC, Grable JC, Love R, Greene PJ, Rosenberg JM. 1990. Refinement of Eco RI endonuclease crystal structure: a revised protein chain tracing. Science 249: 1307–1309. 10.1126/science.2399465

Kim YG, Li L, Chandrasegaran S. 1994. Insertion and deletion mutants of FokI restriction endonuclease. J Biol Chem 269: 31978–31982.

Kim SC, Skowron PM, Szybalski W. 1996a. Structural requirements for FokI-DNA interaction and oligodeoxyribonucleotide-instructed cleavage. J Mol Biol 258: 638–649. 10.1006/jmbi.1996.0275

Kim YG, Cha J, Chandrasegaran S. 1996b. Hybrid restriction enzymes: zinc finger fusions to FokI cleavage domain. Proc Natl Acad Sci 93: 1156–1160. 10.1073/pnas.93.3.1156

Kim YG, Shi Y, Berg JM, Chandrasegaran S. 1997. Site-specific cleavage of DNA–RNA hybrids by zinc finger/FokI cleavage domain fusions. Gene 203: 43–49. 10.1016/S0378-1119(97)00489-7

Kim YG, Smith J, Durgesha M, Chandrasegaran S. 1998. Chimeric restriction enzyme: Gal4 fusion to FokI cleavage domain. Biol Chem 379: 489–495. 10.1515/bchm.1998.379.4-5.489

Kim JS, DeGiovanni A, Jancarik J, Adams PD, Yokota H, Kim R, Kim SH. 2005. Crystal structure of DNA sequence specificity subunit of a type I restriction-modification enzyme and its functional implications. Proc Natl Acad Sci 102: 3248–3253. 10.1073/pnas.0409851102

King G, Murray NE. 1994. Restriction enzymes in cells, not eppendorfs. Trends Microbiol 2: 465–469. 10.1016/0966-842X(94)90649-1

Kingston IJ, Gormley NA, Halford SE. 2003. DNA supercoiling enables the type IIS restriction enzyme BspMI to recognise the relative orientation of two DNA sequences. Nucleic Acids Res 31: 5221–5228. 10.1093/nar/gkg743

Kita K, Kotani H, Hiraoka N, Nakamura T, Yonaha K. 1989a. Overproduction and crystallization of FokI restriction endonuclease. Nucleic Acids Res 17: 8741–8753. 10.1093/nar/17.21.8741

Kita K, Kotani H, Sugisaki H, Takanami M. 1989b. The FokI restriction-modification system. I. Organization and nucleotide sequences of the restriction and modification genes. J Biol Chem 264: 5751–5756.

Kita K, Kotani H, Ohta H, Yanase H, Kato N. 1992a. StsI, a new FokI isoschizomer from Streptococcus sanguis 54, cleaves 5′ GGATG(N)10/14 3′. Nucleic Acids Res 20: 618. PMCID: 310441. 10.1093/nar/20.3.618

Kita K, Suisha M, Kotani H, Yanase H, Kato N. 1992b. Cloning and sequence analysis of the StsI restriction-modification gene: presence of homology to FokI restriction-modification enzymes. Nucleic Acids Res 20: 4167–4172. 10.1093/nar/20.16.4167

Kleinstiver BP, Berube-Janzen W, Fernandes AD, Edgell DR. 2011. Divalent metal ion differentially regulates the sequential nicking reactions of the GIY-YIG homing endonuclease I-BmoI. PloS One 6: e23804. doi:10.1371/journal.pone.0023804. 10.1371/journal.pone.0023804

Kleinstiver BP, Wolfs JM, Edgell DR. 2013. The monomeric GIY-YIG homing endonuclease I-BmoI uses a molecular anchor and a flexible tether to sequentially nick DNA. Nucleic Acids Res 41: 5413–5427. 10.1093/nar/gkt186

Klug A. 2010a. The discovery of zinc fingers and their applications in gene regulation and genome manipulation. Annu Rev Biochem 79: 213–231. 10.1146/annurev-biochem-010909-095056

Klug A. 2010b. The discovery of zinc fingers and their development for practical applications in gene regulation and genome manipulation. Q Rev Biophys 43: 1–21. 10.1017/S0033583510000089

Knowle D, Lintner RE, Touma YM, Blumenthal RM. 2005. Nature of the promoter activated by C·PvuII, an unusual regulatory protein conserved among restriction-modification systems. J Bacteriol 187: 488–497. 10.1128/JB.187.2.488-497.2005

Kojima KK, Kobayashi I. 2015. Transmission of the PabI family of restriction DNA glycosylase genes: mobility and long-term inheritance. BMC Genomics 16: 817. doi:810.1186/s12864-12015-12021-12863. 10.1186/s12864-015-2021-3

Kong H. 1998. “Characterization of a new R-M system, the BcgI system of Bacillus coagulans” PhD thesis, Boston University.

Kong H, Smith CL. 1997. Substrate DNA and cofactor regulate the activities of a multi-functional restriction-modification enzyme, BcgI. Nucleic Acids Res 25: 3687–3692. 10.1093/nar/25.18.3687

Kong H, Smith CL. 1998. Does BcgI, a unique restriction endonuclease, require two recognition sites for cleavage? Biol Chem 379: 605–609.

Kong H, Morgan RD, Maunus RE, Schildkraut I. 1993. A unique restriction endonuclease, BcgI, from Bacillus coagulans. Nucleic Acids Res 21: 987–991. 10.1093/nar/21.4.987

Kong H, Roemer SE, Waite-Rees PA, Benner JS, Wilson GG, Nwankwo DO. 1994. Characterization of BcgI, a new kind of restriction-modification system. J Biol Chem 269: 683–690.

Kong H, Lin LF, Porter N, Stickel S, Byrd D, Posfai J, Roberts RJ. 2000. Functional analysis of putative restriction-modification system genes in the Helicobacter pylori J99 genome. Nucleic Acids Res 28: 3216–3223. 10.1093/nar/28.17.3216

Korlach J, Turner SW. 2012. Going beyond five bases in DNA sequencing. Curr Opin Struct Biol 22: 251–261. 10.1016/j.sbi.2012.04.002

Korlach J, Bjornson KP, Chaudhuri BP, Cicero RL, Flusberg BA, Gray JJ, Holden D, Saxena R, Wegener J, Turner SW. 2010. Real-time DNA sequencing from single polymerase molecules. Methods Enzymol 472: 431–455. 10.1016/S0076-6879(10)72001-2

Kostiuk G, Dikic J, Sasnauskas G, Seidel R, Šikšnys V. 2015. Single-molecule analysis of the monomeric REase BcnI. In 7th NEB meeting on restriction and modification (Gdansk): Poster P30.

Kostiuk G, Sasnauskas G, Tamulaitienė G, Šikšnys V. 2011. Degenerate sequence recognition by the monomeric restriction enzyme: single mutation converts BcnI into a strand-specific nicking endonuclease. Nucleic Acids Res 39: 3744–3753. 10.1093/nar/gkq1351

Kostiuk G, Dikic J, Schwarz FW, Sasnauskas G, Seidel R, Šikšnys V. 2017. The dynamics of the monomeric restriction endonuclease BcnI during its interaction with DNA. Nucleic Acids Res 45: 5968–5979. 10.1093/nar/gkx294

Kostrewa D, Winkler FK. 1995. Mg2+ binding to the active site of EcoRV endonuclease: a crystallographic study of complexes with substrate and product DNA at 2 A resolution. Biochemistry 34: 683–696. 10.1021/bi00002a036

Kovall RA, Matthews BW. 1999. Type II restriction endonucleases: structural, functional and evolutionary relationships. Curr Opin Chem Biol 3: 578–583. 10.1016/S1367-5931(99)00012-5

Kriaucionis S, Heintz N. 2009. The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 324: 929–930. 10.1126/science.1169786

Kriukiene E. 2006. Domain organization and metal ion requirement of the Type IIS restriction endonuclease MnlI. FEBS Lett 580: 6115–6122. 10.1016/j.febslet.2006.09.075

Kriukiene E, Lubiene J, Lagunavicius A, Lubys A. 2005. MnlI—The member of H-N-H subtype of Type IIS restriction endonucleases. Biochim Biophys Acta 1751: 194–204. 10.1016/j.bbapap.2005.06.006

Kruger DH, Bickle TA. 1983. Bacteriophage survival: multiple mechanisms for avoiding the deoxyribonucleic acid restriction systems of their hosts. Microbiol Rev 47: 345–360.

Kruger DH, Reuter M. 2005. Reliable detection of DNA cytosine methylation at CpNpG sites using the engineered restriction enzyme EcoRII-C. BioTechniques 38: 855–856. 10.2144/05386BM01

Kruger DH, Hansen S, Schroeder C, Presber W. 1977a. Host-dependent modification of bacteriophage T7 and SAMase-negative T3 derivatives affecting their adsorption ability. Mol Gen Genet 153: 107–110. 10.1007/BF01036002

Kruger DH, Presber W, Hansen S, Rosenthal HA. 1977b. Different restriction of bacteriophages T3 and T7 by P1-lysogenic cells and the role of the T3-coded SAMase. Z Allg Mikrobiol 17: 581–591. 10.1002/jobm.3630170802

Kruger DH, Schroeder C, Hansen S, Rosenthal HA. 1977c. Active protection by bacteriophages T3 and T7 against E. coli B- and K-specific restriction of their DNA. Mol General Genet 153: 99–106. 10.1007/BF01036001

Kruger DH, Reuter M, Schroeder C, Glatman LI, Chernin LS. 1983. Restriction of bacteriophage T3 and T7 ocr+ strains by the type II restriction endonuclease EcoRV. Mol Gen Genet 190: 349–351. 10.1007/BF00330663

Kruger DH, Barcak GJ, Reuter M, Smith HO. 1988. EcoRII can be activated to cleave refractory DNA recognition sites. Nucleic Acids Res 16: 3997–4008. 10.1093/nar/16.9.3997

Kubareva EA, Petrauskene OV, Karyagina AS, Tashlitsky VN, Nikolskaya II, Gromova ES. 1992. Cleavage of synthetic substrates containing non-nucleotide inserts by restriction endonucleases. Change in the cleavage specificity of endonuclease SsoII. Nucleic Acids Res 20: 4533–4538. 10.1093/nar/20.17.4533

Kubareva EA, Thole H, Karyagina AS, Oretskaya TS, Pingoud A, Pingoud V. 2000. Identification of a base-specific contact between the restriction endonuclease SsoII and its recognition sequence by photocross-linking. Nucleic Acids Res 28: 1085–1091. 10.1093/nar/28.5.1085

Kulkarni M, Nirwan N, van Aelst K, Szczelkun MD, Saikrishnan K. 2016. Structural insights into DNA sequence recognition by Type ISP restriction-modification enzymes. Nucleic Acids Res 44: 4396–4408. 10.1093/nar/gkw154

Kuper J, Braun C, Elias A, Michels G, Sauer F, Schmitt DR, Poterszman A, Egly JM, Kisker C. 2014. In TFIIH, XPD helicase is exclusively devoted to DNA repair. PLoS Biol 12: e1001954. doi:10.1371/journal.pbio.1001954. 10.1371/journal.pbio.1001954

Kurpiewski MR, Engler LE, Wozniak LA, Kobylanska A, Koziolkiewicz M, Stec WJ, Jen-Jacobson L. 2004. Mechanisms of coupling between DNA recognition specificity and catalysis in EcoRI endonuclease. Structure 12: 1775–1788. 10.1016/j.str.2004.07.016

Kwiatek A, Luczkiewicz M, Bandyra K, Stein DC, Piekarowicz A. 2010. Neisseria gonorrhoeae FA1090 carries genes encoding two classes of Vsr endonucleases. J Bacteriol 192: 3951–3960. 10.1128/JB.00098-10

Kwiatek A, Mrozek A, Bacal P, Piekarowicz A, Adamczyk-Popławska M. 2015. Type III methyltransferase M.NgoAX from Neisseria gonorrhoeae FA1090 regulates biofilm formation and interactions with human cells. Frontiers Microbiol 6: 1426. doi:1410.3389/fmicb.2015.01426. 10.3389/fmicb.2015.01426

Lacks S, Greenberg B. 1975. A deoxyribonuclease of Diplococcus pneumoniae specific for methylated DNA. J Biol Chem 250: 4060–4066.

Lacks S, Greenberg B. 1977. Complementary specificity of restriction endonucleases of Diplococcus pneumoniae with respect to DNA methylation. J Mol Biol 114: 153–168. 10.1016/0022-2836(77)90289-3

Lagunavicius A, Sasnauskas G, Halford SE, Šikšnys V. 2003. The metal-independent type IIs restriction enzyme BfiI is a dimer that binds two DNA sites but has only one catalytic centre. J Mol Biol 326: 1051–1064. 10.1016/S0022-2836(03)00020-2

Lambert AR, Sussman D, Shen B, Maunus R, Nix J, Samuelson J, Xu SY, Stoddard BL. 2008. Structures of the rare-cutting restriction endonuclease NotI reveal a unique metal binding fold involved in DNA binding. Structure 16: 558–569. 10.1016/j.str.2008.01.017

Landry D, Looney MC, Feehery GR, Slatko BE, Jack WE, Schildkraut I, Wilson GG. 1989. M·FokI methylates adenine in both strands of its asymmetric recognition sequence. Gene 77: 1–10. 10.1016/0378-1119(89)90353-3

Lanio T, Jeltsch A, Pingoud A. 2000. On the possibilities and limitations of rational protein design to expand the specificity of restriction enzymes: a case study employing EcoRV as the target. Protein Eng 13: 275–281. 10.1093/protein/13.4.275

Lapkouski M, Panjikar S, Kuta Smatanova I, Csefalvay E. 2007. Purification, crystallization and preliminary X-ray analysis of the HsdR subunit of the EcoR124I endonuclease from Escherichia coli. Acta Crystallogr, Sect F: Struct Biol Cryst Commun 63: 582–585. 10.1107/S174430910702622X

Lapkouski M, Panjikar S, Janscak P, Smatanova IK, Carey J, Ettrich R, Csefalvay E. 2009. Structure of the motor subunit of type I restriction-modification complex EcoR124I. Nat Struct Mol Biol 16: 94–95. 10.1038/nsmb.1523

Larsen MH, Figurski DH. 1994. Structure, expression, and regulation of the kilC operon of promiscuous IncPα plasmids. J Bacteriol 176: 5022–5032. 10.1128/jb.176.16.5022-5032.1994

Laue F, Ankenbauer W, Schmitz GG, Kessler C. 1990. The selective inhibitory effect of netropsin on relaxation of sequence specificity of restriction endonuclease SgrAI recognizing the octanucleotide sequence 5′-CR decreases CCGGYG-3′. Nucleic Acids Res 18: 3421. 10.1093/nar/18.11.3421

Laurens N, Bellamy SR, Harms AF, Kovacheva YS, Halford SE, Wuite GJ. 2009. Dissecting protein-induced DNA looping dynamics in real time. Nucleic Acids Res 37: 5454–5464. 10.1093/nar/gkp570

Laurens N, Rusling DA, Pernstich C, Brouwer I, Halford SE, Wuite GJ. 2012. DNA looping by FokI: the impact of twisting and bending rigidity on protein-induced looping dynamics. Nucleic Acids Res 40: 4988–4997. 10.1093/nar/gks184

Le May N, Egly JM, Coin F. 2010. True lies: the double life of the nucleotide excision repair factors in transcription and DNA repair. J Nucleic Acids 2010. doi:10.4061/2010/616342. 10.4061/2010/616342

Leismann O, Roth M, Friedrich T, Wende W, Jeltsch A. 1998. The Flavobacterium okeanokoites adenine-N6-specific DNA-methyltransferase M·FokI is a tandem enzyme of two independent domains with very different kinetic properties. Eur J Biochem/FEBS 251: 899–906. 10.1046/j.1432-1327.1998.2510899.x

Lepikhov K, Tchernov A, Zheleznaja L, Matvienko N, Walter J, Trautner TA. 2001. Characterization of the type IV restriction modification system BspLU11III from Bacillus sp. LU11. Nucleic Acids Res 29: 4691–4698. 10.1093/nar/29.22.4691

Levinson G, Gutman GA. 1987a. High frequencies of short frameshifts in poly-CA/TG tandem repeats borne by bacteriophage M13 in Escherichia coli K-12. Nucleic Acids Res 15: 5323–5338. 10.1093/nar/15.13.5323

Levinson G, Gutman GA. 1987b. Slipped-strand mispairing: a major mechanism for DNA sequence evolution. Mol Biol Evol 4: 203–221.

Li L, Chandrasegaran S. 1993. Alteration of the cleavage distance of Fok I restriction endonuclease by insertion mutagenesis. Proc Natl Acad Sci 90: 2764–2768. 10.1073/pnas.90.7.2764

Li L, Wu LP, Chandrasegaran S. 1992. Functional domains in Fok I restriction endonuclease. Proc Natl Acad Sci 89: 4275–4279. 10.1073/pnas.89.10.4275

Li L, Wu LP, Clarke R, Chandrasegaran S. 1993. C-terminal deletion mutants of the FokI restriction endonuclease. Gene 133: 79–84. 10.1016/0378-1119(93)90227-T

Li T, Huang S, Jiang WZ, Wright D, Spalding MH, Weeks DP, Yang B. 2011. TAL nucleases (TALNs): hybrid proteins composed of TAL effectors and FokI DNA-cleavage domain. Nucleic Acids Res 39: 359–372. 10.1093/nar/gkq704

Liang J, Blumenthal RM. 2013. Naturally-occurring, dually-functional fusions between restriction endonucleases and regulatory proteins. BMC Evol Biol 13: 218. 10.1186/1471-2148-13-218

Lin LF, Posfai J, Roberts RJ, Kong H. 2001. Comparative genomics of the restriction-modification systems in Helicobacter pylori. Proc Natl Acad Sci 98: 2740–2745. 10.1073/pnas.051612298

Lindsay JA. 2010. Genomic variation and evolution of Staphylococcus aureus. Int J Med Microbiol 300: 98–103. 10.1016/j.ijmm.2009.08.013

Little EJ, Dunten PW, Bitinaite J, Horton NC. 2011. New clues in the allosteric activation of DNA cleavage by SgrAI: structures of SgrAI bound to cleaved primary-site DNA and uncleaved secondary-site DNA. Acta Crystallogr, Sect D: Biol Crystallogr 67: 67–74. 10.1107/S0907444910047785

Liu Y, Kobayashi I. 2007. Negative regulation of the EcoRI restriction enzyme gene is associated with intragenic reverse promoters. J Bacteriol 189: 6928–6935. 10.1128/JB.00127-07

Liu Y, Ichige A, Kobayashi I. 2007. Regulation of the EcoRI restriction-modification system: Identification of ecoRIM gene promoters and their upstream negative regulators in the ecoRIR gene. Gene 400: 140–149. 10.1016/j.gene.2007.06.006

Liu G, Ou HY, Wang T, Li L, Tan H, Zhou X, Rajakumar K, Deng Z, He X. 2010. Cleavage of phosphorothioated DNA and methylated DNA by the type IV restriction endonuclease ScoMcrA. PLoS Genet 6: e1001253. doi:10.1371/journal.pgen.1001253. 10.1371/journal.pgen.1001253

Lobner-Olesen A, Skovgaard O, Marinus MG. 2005. Dam methylation: coordinating cellular processes. Curr Opin Microbiol 8: 154–160. 10.1016/j.mib.2005.02.009

Loenen WAM. 2003. Tracking EcoKI and DNA fifty years on: a golden story full of surprises. Nucleic Acids Res 31: 7059–7069. 10.1093/nar/gkg944

Loenen WAM. 2006. S-adenosylmethionine: jack of all trades and master of everything? Biochem Soc Trans 34: 330–333. 10.1042/BST0340330

Loenen WAM. 2010. S-adenosylmethionine: simple agent of methylation and secret to aging and metabolism? In Epigenetics of aging (ed. Tollefsbol TO), pp 107–131. Springer, Berlin.

Loenen WAM. 2017. S-adenosylmethionine: a promising avenue in anti-aging medicine? In Anti-aging drugs: from basic research to clinical practice (ed. Vaiserman, Alexander M), RSC Drug Discovery Series, Vol. 57, pp. 435–473. Royal Society of Chemistry, Cambridge.

Loenen WAM. 2018. S-adenosylmethionine metabolism and aging. In Epigenetics of aging and longevity (ed. Moskalev A, Vaiserman A), Ch. 3, pp. 59–93. Elsevier, New York.

Loenen WA, Murray NE. 1986. Modification enhancement by the restriction alleviation protein (Ral) of bacteriophage λ. J Mol Biol 190: 11–22. 10.1016/0022-2836(86)90071-9

Loenen WA, Raleigh EA. 2014. The other face of restriction: modification-dependent enzymes. Nucleic Acids Res 42: 56–69. 10.1093/nar/gkt747

Loenen WA, Dryden DT, Raleigh EA, Wilson GG. 2014a. Type I restriction enzymes and their relatives. Nucleic Acids Res 42: 20–44. 10.1093/nar/gkt847

Loenen WA, Dryden DT, Raleigh EA, Wilson GG, Murray NE. 2014b. Highlights of the DNA cutters: a short history of the restriction enzymes. Nucleic Acids Res 42: 3–19. 10.1093/nar/gkt990

Looney MC, Moran LS, Jack WE, Feehery GR, Benner JS, Slatko BE, Wilson GG. 1989. Nucleotide sequence of the FokI restriction-modification system: separate strand-specificity domains in the methyltransferase. Gene 80: 193–208. 10.1016/0378-1119(89)90284-9

Low DA, Casadesus J. 2008. Clocks and switches: bacterial gene regulation by DNA adenine methylation. Curr Opin Microbiol 11: 106–112. 10.1016/j.mib.2008.02.012

Lu X, Zhao BS, He C. 2015. TET family proteins: oxidation activity, interacting molecules, and functions in diseases. Chem Rev 115: 2225–2239. 10.1021/cr500470n

Lukacs CM, Kucera R, Schildkraut I, Aggarwal AK. 2000. Understanding the immutability of restriction enzymes: crystal structure of BglII and its DNA substrate at 1.5 Å resolution. Nat Struct Biol 7: 134–140. 10.1038/72405

Luria SE, Human ML. 1952. A nonhereditary, host-induced variation of bacterial viruses. J Bacteriol 64: 557–569.

Lyumkis D, Talley H, Stewart A, Shah S, Park CK, Tama F, Potter CS, Carragher B, Horton NC. 2013. Allosteric regulation of DNA cleavage and sequence-specificity through run-on oligomerization. Structure 21: 1848–1858. 10.1016/j.str.2013.08.012

Ma L, Chen K, Clarke DJ, Nortcliffe CP, Wilson GG, Edwardson JM, Morton AJ, Jones AC, Dryden DT. 2013a. Restriction endonuclease TseI cleaves A:A and T:T mismatches in CAG and CTG repeats. Nucleic Acids Res 41: 4999–5009. 10.1093/nar/gkt176

Ma X, Shah S, Zhou M, Park CK, Wysocki VH, Horton NC. 2013b. Structural analysis of activated SgrAI-DNA oligomers using ion mobility mass spectrometry. Biochemistry 52: 4373–4381. 10.1021/bi3013214

Machnicka MA, Kaminska KH, Dunin-Horkawicz S, Bujnicki JM. 2015. Phylogenomics and sequence-structure-function relationships in the GmrSD family of Type IV restriction enzymes. BMC Bioinf 16: 336. doi:310.1186/s12859-12015-10773-z. 10.1186/s12859-015-0773-z

Mackeldanz P, Alves J, Moncke-Buchner E, Wyszomirski KH, Kruger DH, Reuter M. 2013. Functional consequences of mutating conserved SF2 helicase motifs in the Type III restriction endonuclease EcoP15I translocase domain. Biochimie 95: 817–823. 10.1016/j.biochi.2012.11.014

Mak AN, Lambert AR, Stoddard BL. 2010. Folding, DNA recognition, and function of GIY-YIG endonucleases: crystal structures of R.Eco29kI. Structure 18: 1321–1331. 10.1016/j.str.2010.07.006

Mak AN, Bradley P, Cernadas RA, Bogdanove AJ, Stoddard BL. 2012. The crystal structure of TAL effector PthXo1 bound to its DNA target. Science 335: 716–719. 10.1126/science.1216211

Mak AN, Bradley P, Bogdanove AJ, Stoddard BL. 2013. TAL effectors: function, structure, engineering and applications. Curr Opin Struct Biol 23: 93–99. 10.1016/j.sbi.2012.11.001

Manakova E, Gražulis S, Zaremba M, Tamulaitienė G, Golovenko D, Šikšnys V. 2012. Structural mechanisms of the degenerate sequence recognition by Bse634I restriction endonuclease. Nucleic Acids Res 40: 6741–6751. 10.1093/nar/gks300

Manakova E, Tamulaitienė G, Tamulaitis G, Mikutenaite M, Šikšnys V. 2015. Crystallographic and functional studies of restriction enzymes recognizing 5′-CCGG tetranucleotide core: Kpn2I and PfoI. In 7th NEB meeting on restriction and modification (Gdansk): Poster P1.

Mandecki W, Bolling TJ. 1988. FokI method of gene synthesis. Gene 68: 101–107. 10.1016/0378-1119(88)90603-8

Mani M, Chen C, Amblee V, Liu H, Mathur T, Zwicke G, Zabad S, Patel B, Thakkar J, Jeffery CJ. 2015. MoonProt: a database for proteins that are known to moonlight. Nucleic Acids Res 43: D277–D282. 10.1093/nar/gku954

Marinus MG, Casadesus J. 2009. Roles of DNA adenine methylation in host-pathogen interactions: mismatch repair, transcriptional regulation, and more. FEMS Microbiol Rev 33: 488–503. 10.1111/j.1574-6976.2008.00159.x

Mark KK, Studier FW. 1981. Purification of the gene 0.3 protein of bacteriophage T7, an inhibitor of the DNA restriction system of Escherichia coli. J Biol Chem 256: 2573–2578.

Marks P, McGeehan J, Wilson G, Errington N, Kneale G. 2003. Purification and characterisation of a novel DNA methyltransferase, M.AhdI. Nucleic Acids Res 31: 2803–2810. 10.1093/nar/gkg399

Marshall JJ, Halford SE. 2010. The type IIB restriction endonucleases. Biochem Soc Trans 38: 410–416. 10.1042/BST0890410

Marshall JJ, Gowers DM, Halford SE. 2007. Restriction endonucleases that bridge and excise two recognition sites from DNA. J Mol Biol 367: 419–431. 10.1016/j.jmb.2006.12.070

Marshall JJ, Smith RM, Ganguly S, Halford SE. 2011. Concerted action at eight phosphodiester bonds by the BcgI restriction endonuclease. Nucleic Acids Res 39: 7630–7640. 10.1093/nar/gkr453

McClarin JA, Frederick CA, Wang BC, Greene P, Boyer HW, Grable J, Rosenberg JM. 1986. Structure of the DNA-Eco RI endonuclease recognition complex at 3 A resolution. Science 234: 1526–1541. 10.1126/science.3024321

McClelland SE, Szczelkun MD. 2004. The Type I and III restriction endonucleases: structural elements in molecular motors that process DNA. In Restriction endonucleases (ed. Pingoud A), pp. 111–135. Springer, Berlin.

McGeehan JE, Streeter S, Cooper JB, Mohammed F, Fox GC, Kneale GG. 2004. Crystallization and preliminary X-ray analysis of the controller protein C.AhdI from Aeromonas hydrophilia. Acta Crystallogr, Sect D: Biol Crystallogr 60: 323–325. 10.1107/S0907444903026143

McGeehan JE, Streeter SD, Papapanagiotou I, Fox GC, Kneale GG. 2005. High-resolution crystal structure of the restriction-modification controller protein C.AhdI from Aeromonas hydrophila. J Mol Biol 346: 689–701. 10.1016/j.jmb.2004.12.025

McGeehan JE, Papapanagiotou I, Streeter SD, Kneale GG. 2006. Cooperative binding of the C.AhdI controller protein to the C/R promoter and its role in endonuclease gene expression. J Mol Biol 358: 523–531. 10.1016/j.jmb.2006.02.003

McGeehan JE, Streeter SD, Thresh SJ, Ball N, Ravelli RB, Kneale GG. 2008. Structural analysis of the genetic switch that regulates the expression of restriction-modification genes. Nucleic Acids Res 36: 4778–4787. 10.1093/nar/gkn448

McGeehan JE, Ball NJ, Streeter SD, Thresh SJ, Kneale GG. 2012. Recognition of dual symmetry by the controller protein C.Esp1396I based on the structure of the transcriptional activation complex. Nucleic Acids Res 40: 4158–4167. 10.1093/nar/gkr1250

McMahon SA, Roberts GA, Johnson KA, Cooper LP, Liu H, White JH, Carter LG, Sanghvi B, Oke M, Walkinshaw MD, et al. 2009. Extensive DNA mimicry by the ArdA anti-restriction protein and its role in the spread of antibiotic resistance. Nucleic Acids Res 37: 4887–4897. 10.1093/nar/gkp478

Mernagh DR, Janscak P, Firman K, Kneale GG. 1998. Protein-protein and protein-DNA interactions in the type I restriction endonuclease R.EcoR124I. Biol Chem 379: 497–503. 10.1515/bchm.1998.379.4-5.497

Mierzejewska K, Siwek W, Czapinska H, Kaus-Drobek M, Radlinska M, Skowronek K, Bujnicki JM, Dadlez M, Bochtler M. 2014. Structural basis of the methylation specificity of R.DpnI. Nucleic Acids Res 42: 8745–8754. 10.1093/nar/gku546

Miller J, McLachlan AD, Klug A. 1985. Repetitive zinc-binding domains in the protein transcription factor IIIA from Xenopus oocytes. EMBO J 4: 1609–1614. 10.1002/j.1460-2075.1985.tb03825.x

Miller ES, Kutter E, Mosig G, Arisaka F, Kunisawa T, Ruger W. 2003. Bacteriophage T4 genome. Microbiol Mol Biol Rev 67: 86–156. 10.1128/MMBR.67.1.86-156.2003

Miller JC, Holmes MC, Wang J, Guschin DY, Lee YL, Rupniewski I, Beausejour CM, Waite AJ, Wang NS, Kim KA, et al. 2007. An improved zinc-finger nuclease architecture for highly specific genome editing. Nat Biotechnol 25: 778–785. 10.1038/nbt1319

Miller JC, Tan S, Qiao G, Barlow KA, Wang J, Xia DF, Meng X, Paschon DE, Leung E, Hinkley SJ, et al. 2011. A TALE nuclease architecture for efficient genome editing. Nat Biotechnol 29: 143–148. 10.1038/nbt.1755

Mino T, Aoyama Y, Sera T. 2009. Efficient double-stranded DNA cleavage by artificial zinc-finger nucleases composed of one zinc-finger protein and a single-chain FokI dimer. J Biotechnol 140: 156–161. 10.1016/j.jbiotec.2009.02.004

Mino T, Mori T, Aoyama Y, Sera T. 2014. Inhibition of DNA replication of human papillomavirus by using zinc finger-single-chain FokI dimer hybrid. Mol Biotechnol 56: 731–737. 10.1007/s12033-014-9751-3

Miyazono K, Watanabe M, Kosinski J, Ishikawa K, Kamo M, Sawasaki T, Nagata K, Bujnicki JM, Endo Y, Tanokura M, et al. 2007. Novel protein fold discovered in the PabI family of restriction enzymes. Nucleic Acids Res 35: 1908–1918. 10.1093/nar/gkm091

Miyazono K, Furuta Y, Watanabe-Matsui M, Miyakawa T, Ito T, Kobayashi I, Tanokura M. 2014. A sequence-specific DNA glycosylase mediates restriction-modification in Pyrococcus abyssi. Nat Commun 5: 3178. 10.1038/ncomms4178

Moffatt BA, Studier FW. 1988. Entry of bacteriophage T7 DNA into the cell and escape from host restriction. J Bacteriol 170: 2095–2105. 10.1128/jb.170.5.2095-2105.1988

Mokrishcheva ML, Solonin AS, Nikitin DV. 2011. Fused eco29kIR- and M genes coding for a fully functional hybrid polypeptide as a model of molecular evolution of restriction-modification systems. BMC Evol Biol 11: 35. doi:10.1186/1471-2148-1111-1135. 10.1186/1471-2148-11-35

Molineux IJ. 2001. No syringes please, ejection of phage T7 DNA from the virion is enzyme driven. Mol Microbiol 40: 1–8. 10.1046/j.1365-2958.2001.02357.x

Moncke-Buchner E, Reich S, Mucke M, Reuter M, Messer W, Wanker EE, Kruger DH. 2002. Counting CAG repeats in the Huntington's disease gene by restriction endonuclease EcoP15I cleavage. Nucleic Acids Res 30: e83. PMID: 12177311. 10.1093/nar/gnf082

Moncke-Buchner E, Rothenberg M, Reich S, Wagenfuhr K, Matsumura H, Terauchi R, Kruger DH, Reuter M. 2009. Functional characterization and modulation of the DNA cleavage efficiency of type III restriction endonuclease EcoP15I in its interaction with two sites in the DNA target. J M Biol 387: 1309–1319. 10.1016/j.jmb.2009.02.047

Morgan RD, Luyten YA. 2009. Rational engineering of type II restriction endonuclease DNA binding and cleavage specificity. Nucleic Acids Res 37: 5222–5233. 10.1093/nar/gkp535

Morgan RD, Calvet C, Demeter M, Agra R, Kong H. 2000. Characterization of the specific DNA nicking activity of restriction endonuclease N.BstNBI. Biol Chem 381: 1123–1125. 10.1515/BC.2000.137

Morgan RD, Bhatia TK, Lovasco L, Davis TB. 2008. MmeI: a minimal Type II restriction-modification system that only modifies one DNA strand for host protection. Nucleic Acids Res 36: 6558–6570. 10.1093/nar/gkn711

Morgan RD, Dwinell EA, Bhatia TK, Lang EM, Luyten YA. 2009. The MmeI family: type II restriction-modification enzymes that employ single-strand modification for host protection. Nucleic Acids Res 37: 5208–5221. 10.1093/nar/gkp534

Morgan RD, Luyten YA, Johnson SA, Clough EM, Clark TA, Roberts RJ. 2016. Novel m4C modification in type I restriction-modification systems. Nucleic Acids Res 44: 9413–9425.

Mori T, Kagatsume I, Shinomiya K, Aoyama Y, Sera T. 2009. Sandwiched zinc-finger nucleases harboring a single-chain FokI dimer as a DNA-cleavage domain. Biochem Biophys Res Commun 390: 694–697. 10.1016/j.bbrc.2009.10.030

Moscou MJ, Bogdanove AJ. 2009. A simple cipher governs DNA recognition by TAL effectors. Science 326: 1501. 10.1126/science.1178817

Moxon R, Bayliss C, Hood D. 2006. Bacterial contingency loci: the role of simple sequence DNA repeats in bacterial adaptation. Ann Rev Genet 40: 307–333. 10.1146/annurev.genet.40.110405.090442

Mruk I, Kobayashi I. 2014. To be or not to be: regulation of restriction-modification systems and other toxin-antitoxin systems. Nucleic Acids Res 42: 70–86. 10.1093/nar/gkt711

Mruk I, Liu Y, Ge L, Kobayashi I. 2011. Antisense RNA associated with biological regulation of a restriction-modification system. Nucleic Acids Res 39: 5622–5632. 10.1093/nar/gkr166

Mulligan EA, Dunn JJ. 2008. Cloning, purification and initial characterization of E. coli McrA, a putative 5-methylcytosine-specific nuclease. Protein Expression Purif 62: 98–103. 10.1016/j.pep.2008.06.016

Mulligan EA, Hatchwell E, McCorkle SR, Dunn JJ. 2010. Differential binding of Escherichia coli McrA protein to DNA sequences that contain the dinucleotide m5CpG. Nucleic Acids Res 38: 1997–2005. 10.1093/nar/gkp1120

Murray NE. 2000. Type I restriction systems: sophisticated molecular machines (a legacy of Bertani and Weigle). Microbiol Mol Biol Rev 64: 412–434. 10.1128/MMBR.64.2.412-434.2000

Murray NE. 2002. 2001 Fred Griffith review lecture. Immigration control of DNA in bacteria: self versus non-self. Microbiology 148: 3–20. 10.1099/00221287-148-1-3

Murray IA, Stickel SK, Roberts RJ. 2010. Sequence-specific cleavage of RNA by Type II restriction enzymes. Nucleic Acids Res 38: 8257–8268. 10.1093/nar/gkq702

Mussolino C, Alzubi J, Fine EJ, Morbitzer R, Cradick TJ, Lahaye T, Bao G, Cathomen T. 2014. TALENs facilitate targeted genome editing in human cells with high specificity and low cytotoxicity. Nucleic Acids Res 42: 6762–6773. 10.1093/nar/gku305

Nakayama Y, Kobayashi I. 1998. Restriction-modification gene complexes as selfish gene entities: roles of a regulatory system in their establishment, maintenance, and apoptotic mutual exclusion. Proc Natl Acad Sci 95: 6442–6447. 10.1073/pnas.95.11.6442

Nakonieczna J, Zmijewski JW, Banecki B, Podhajska AJ. 2007. Binding of MmeI restriction–modification enzyme to its specific recognition sequence is stimulated by S-adenosyl-l-methionine. Mol Biotechnol 37: 127–135. 10.1007/s12033-007-0034-0

Nakonieczna J, Kaczorowski T, Obarska-Kosinska A, Bujnicki JM. 2009. Functional analysis of MmeI from methanol utilizer Methylophilus methylotrophus, a subtype IIC restriction-modification enzyme related to type I enzymes. Appl Environ Microbiol 75: 212–223. 10.1128/AEM.01322-08

Nasim MT, Eperon IC, Wilkins BM, Brammar WJ. 2004. The activity of a single-stranded promoter of plasmid ColIb-P9 depends on its secondary structure. Mol Microbiol 53: 405–417. 10.1111/j.1365-2958.2004.04114.x

Neaves KJ, Cooper LP, White JH, Carnally SM, Dryden DT, Edwardson JM, Henderson RM. 2009. Atomic force microscopy of the EcoKI Type I DNA restriction enzyme bound to DNA shows enzyme dimerization and DNA looping. Nucleic Acids Res 37: 2053–2063. 10.1093/nar/gkp042

Nekrasov SV, Agafonova OV, Belogurova NG, Delver EP, Belogurov AA. 2007. Plasmid-encoded antirestriction protein ArdA can discriminate between type I methyltransferase and complete restriction-modification system. J Mol Biol 365: 284–297. 10.1016/j.jmb.2006.09.087

Newman M, Strzelecka T, Dorner LF, Schildkraut I, Aggarwal AK. 1994. Structure of restriction endonuclease BamHI and its relationship to EcoRI. Nature 368: 660–664. 10.1038/368660a0

Niv MY, Ripoll DR, Vila JA, Liwo A, Vanamee ES, Aggarwal AK, Weinstein H, Scheraga HA. 2007. Topology of Type II REases revisited; structural classes and the common conserved core. Nucleic Acids Res 35: 2227–2237. 10.1093/nar/gkm045

Noyer-Weidner M, Diaz R, Reiners L. 1986. Cytosine-specific DNA modification interferes with plasmid establishment in Escherichia coli K12: involvement of rglB. Mol Gen Genet 205: 469–475. 10.1007/BF00338084

Nwankwo D, Wilson G. 1987. Cloning of two type II methylase genes that recognise asymmetric nucleotide sequences: FokI and HgaI. Mol Gen Genet 209: 570–574. 10.1007/BF00331164

Obarska A, Blundell A, Feder M, Vejsadova S, Sisakova E, Weiserova M, Bujnicki JM, Firman K. 2006. Structural model for the multisubunit Type IC restriction-modification DNA methyltransferase M.EcoR124I in complex with DNA. Nucleic Acids Res 34: 1992–2005. 10.1093/nar/gkl132

Oke M, Carter LG, Johnson KA, Liu H, McMahon SA, Yan X, Kerou M, Weikart ND, Kadi N, Sheikh MA, et al. 2010. The Scottish Structural Proteomics Facility: targets, methods and outputs. J Struct Funct Genomics 11: 167–180. 10.1007/s10969-010-9090-y

Orlowski J, Bujnicki JM. 2008. Structural and evolutionary classification of Type II restriction enzymes based on theoretical and experimental analyses. Nucleic Acids Res 36: 3552–3569. 10.1093/nar/gkn175

Ou HY, He X, Shao Y, Tai C, Rajakumar K, Deng Z. 2009. dndDB: a database focused on phosphorothioation of the DNA backbone. PloS One 4: e5132. doi:10.1371/journal.pone.0005132. 10.1371/journal.pone.0005132

Panne D, Muller SA, Wirtz S, Engel A, Bickle TA. 2001. The McrBC restriction endonuclease assembles into a ring structure in the presence of G nucleotides. EMBO J 20: 3210–3217. 10.1093/emboj/20.12.3210

Papapanagiotou I, Streeter SD, Cary PD, Kneale GG. 2007. DNA structural deformations in the interaction of the controller protein C.AhdI with its operator sequence. Nucleic Acids Res 35: 2643–2650. 10.1093/nar/gkm129

Park CK, Stiteler AP, Shah S, Ghare MI, Bitinaite J, Horton NC. 2010. Activation of DNA cleavage by oligomerization of DNA-bound SgrAI. Biochemistry 49: 8818–8830. 10.1021/bi100557v

Pastor WA, Aravind L, Rao A. 2013. TETonic shift: biological roles of TET proteins in DNA demethylation and transcription. Nat Rev Mol Cell Biol 14: 341–356. 10.1038/nrm3589

Pattanayak V, Ramirez CL, Joung JK, Liu DR. 2011. Revealing off-target cleavage specificities of zinc-finger nucleases by in vitro selection. Nat Methods 8: 765–770. 10.1038/nmeth.1670

Peakman LJ, Szczelkun MD. 2004. DNA communications by Type III restriction endonucleases—confirmation of 1D translocation over 3D looping. Nucleic Acids Res 32: 4166–4174. 10.1093/nar/gkh762

Penn NW, Suwalski R, O'Riley C, Bojanowski K, Yura R. 1972. The presence of 5-hydroxymethylcytosine in animal deoxyribonucleic acid. Biochem J 126: 781–790. 10.1042/bj1260781

Perez-Pinera P, Ousterout DG, Brown MT, Gersbach CA. 2012a. Gene targeting to the ROSA26 locus directed by engineered zinc finger nucleases. Nucleic Acids Res 40: 3741–3752. 10.1093/nar/gkr1214

Perez-Pinera P, Ousterout DG, Gersbach CA. 2012b. Advances in targeted genome editing. Curr Opin Chem Biol 16: 268–277. 10.1016/j.cbpa.2012.06.007

Pernstich C, Halford SE. 2012. Illuminating the reaction pathway of the FokI restriction endonuclease by fluorescence resonance energy transfer. Nucleic Acids Res 40: 1203–1213. 10.1093/nar/gkr809

Pertzev AV, Kravetz AN, Mayorov SG, Zakharova MV, Solonin AS. 1997. Isolation of a strain overproducing endonuclease Eco29kI: purification and characterization of the homogeneous enzyme. Biochem Biokhim 62: 732–741.

Petrov VM, Ratnayaka S, Nolan JM, Miller ES, Karam JD. 2010. Genomes of the T4-related bacteriophages as windows on microbial genome evolution. Virol J 7: 292. doi:10.1186/1743-422X-7-292. 10.1186/1743-422X-7-292

Piekarowicz A. 1974. The influence of methionine deprivation on restriction properties of Haemophilus influenzae Rd and Ra strains. Acta Microbiol Pol, Ser A 6: 71–74.

Piekarowicz A. 1982. HineI is an isoschizomer of HinfIII restriction endonuclease. J Mol Biol 157: 373–381. 10.1016/0022-2836(82)90240-6

Piekarowicz A. 1984. Preferential cleavage by restriction endonuclease HinfIII. Acta Biochim Pol 31: 453–464.

Piekarowicz A. 2013. REases and DNA MTases in H. influenzae and N. gonorrhoeae. Talk at CSHL meeting on the history of Restriction and Modification.

Piekarowicz A, Baj J. 1975. Host specificity of DNA in haemophilus influenzae: The physiological and genetical bases of instability of restriction and modification of DNA in strain Rd. Acta Microbiol Pol, Ser A 8: 119–130.

Piekarowicz A, Brzezinski R. 1980. Cleavage and methylation of DNA by the restriction endonuclease HinfIII isolated from Haemophilus influenzae Rf. J Mol Biol 144: 415–429. 10.1016/0022-2836(80)90329-0

Piekarowicz A, Glover SW. 1972. Host specificity of DNA in Haemophilus influenzae: the two restriction and modification systems in strain Ra. Mol Gen Genet 116: 11–25. 10.1007/BF00334255

Piekarowicz A, Kalinowska J. 1974. Host specificity of DNA in Haemophilus influenzae: similarity between host-specificity types of Haemophilus influenzae Re and Rf. J Gen Microbiol 81: 405–411. 10.1099/00221287-81-2-405

Piekarowicz A, Kauc L, Glover SW. 1974. Host specificity of DNA in Haemophilus influenzae: the restriction and modification systems in strains Rb and Rf. J Gen Microbiol 81: 391–403. 10.1099/00221287-81-2-391

Piekarowicz A, Brzezinski R, Kauc L. 1975. Host specificity of DNA in Haemophilus influenzae: the in vivo action of the restriction endonucleases on phage and bacterial DNA. Acta Microbiol Pol, Ser A 7: 51–65.

Piekarowicz A, Brzezinski R, Kauc L. 1976. Host specificity of DNA in Haemophilus influenzae: DNA restriction enzyme from H. influenzae Rf232. Acta Microbiol Pol 25: 307–312.

Piekarowicz A, Bickle TA, Shepherd JC, Ineichen K. 1981. The DNA sequence recognised by the HinfIII restriction endonuclease. J Mol Biol 146: 167–172. 10.1016/0022-2836(81)90372-7

Piekarowicz A, Brzezinski R, Smorawinska M, Kauc L, Skowronek K, Lenarczyk M, Golembiowska M, Siwinska M. 1986. Major spontaneous genomic rearrangements in Haemophilus influenzae S2 and HP1c1 bacteriophages. Gene 49: 111–118. 10.1016/0378-1119(86)90390-2

Piekarowicz A, Yuan R, Stein DC. 1988. Identification of a new restriction endonuclease, R.NgoBI, from Neisseria gonorrhoeae. Nucleic Acids Res 16: 9868. PMID: 3141904. 10.1093/nar/16.20.9868

Pieper U, Pingoud A. 2002. A mutational analysis of the PD…D/EXK motif suggests that McrC harbors the catalytic center for DNA cleavage by the GTP-dependent restriction enzyme McrBC from Escherichia coli. Biochemistry 41: 5236–5244. 10.1021/bi0156862

Pieper U, Brinkmann T, Kruger T, Noyer-Weidner M, Pingoud A. 1997. Characterization of the interaction between the restriction endonuclease McrBC from E. coli and its cofactor GTP. J Mol Biol 272: 190–199. 10.1006/jmbi.1997.1228

Pieper U, Groll DH, Wunsch S, Gast FU, Speck C, Mucke N, Pingoud A. 2002. The GTP-dependent restriction enzyme McrBC from Escherichia coli forms high-molecular mass complexes with DNA and produces a cleavage pattern with a characteristic 10-base pair repeat. Biochemistry 41: 5245–5254. 10.1021/bi015687u

Pingoud A. 2004. Restriction endonucleases. Springer, Berlin.

Pingoud A, Silva GH. 2007. Precision genome surgery. Nat Biotechnol 25: 743–744. 10.1038/nbt0707-743

Pingoud V, Kubareva E, Stengel G, Friedhoff P, Bujnicki JM, Urbanke C, Sudina A, Pingoud A. 2002. Evolutionary relationship between different subgroups of restriction endonucleases. J Biol Chem 277: 14306–14314. 10.1074/jbc.M111625200

Pingoud V, Conzelmann C, Kinzebach S, Sudina A, Metelev V, Kubareva E, Bujnicki JM, Lurz R, Luder G, Xu SY, et al. 2003. PspGI, a type II restriction endonuclease from the extreme thermophile Pyrococcus sp.: structural and functional studies to investigate an evolutionary relationship with several mesophilic restriction enzymes. J Mol Biol 329: 913–929. 10.1016/S0022-2836(03)00523-0

Pingoud A, Fuxreiter M, Pingoud V, Wende W. 2005a. Type II restriction endonucleases: structure and mechanism. Cell Mol Life Sci 62: 685–707. 10.1007/s00018-004-4513-1

Pingoud V, Geyer H, Geyer R, Kubareva E, Bujnicki JM, Pingoud A. 2005b. Identification of base-specific contacts in protein-DNA complexes by photocrosslinking and mass spectrometry: a case study using the restriction endonuclease SsoII. Mol BioSyst 1: 135–141. 10.1039/B503091A

Pingoud V, Sudina A, Geyer H, Bujnicki JM, Lurz R, Luder G, Morgan R, Kubareva E, Pingoud A. 2005c. Specificity changes in the evolution of type II restriction endonucleases: a biochemical and bioinformatic analysis of restriction enzymes that recognize unrelated sequences. J Biol Chem 280: 4289–4298. 10.1074/jbc.M409020200

Pingoud A, Wilson GG, Wende W. 2014. Type II restriction endonucleases—a historical perspective and more. Nucleic Acids Res 42: 7489–7527. 10.1093/nar/gku447

Powell LM, Dryden DT, Murray NE. 1998. Sequence-specific DNA binding by EcoKI, a type IA DNA restriction enzyme. J Mol Biol 283: 963–976. 10.1006/jmbi.1998.2143

Powell LM, Lejeune E, Hussain FS, Cronshaw AD, Kelly SM, Price NC, Dryden DT. 2003. Assembly of EcoKI DNA methyltransferase requires the C-terminal region of the HsdM modification subunit. Biophys Chem 103: 129–137. 10.1016/S0301-4622(02)00251-X

Price C, Lingner J, Bickle TA, Firman K, Glover SW. 1989. Basis for changes in DNA recognition by the EcoR124 and EcoR124/3 type I DNA restriction and modification enzymes. J Mol Biol 205: 115–125. 10.1016/0022-2836(89)90369-0

Prohaska SJ, Stadler PF, Krakauer DC. 2010. Innovation in gene regulation: the case of chromatin computation. J Theor Biol 265: 27–44. 10.1016/j.jtbi.2010.03.011

Protsenko A, Zakharova M, Nagornykh M, Solonin A, Severinov K. 2009. Transcription regulation of restriction-modification system Ecl18kI. Nucleic Acids Res 37: 5322–5330. 10.1093/nar/gkp579

Putnam CD, Tainer JA. 2005. Protein mimicry of DNA and pathway regulation. DNA Repair (Amst) 4: 1410–1420. 10.1016/j.dnarep.2005.08.007

Raghavendra NK, Rao DN. 2004. Unidirectional translocation from recognition site and a necessary interaction with DNA end for cleavage by Type III restriction enzyme. Nucleic Acids Res 32: 5703–5711. 10.1093/nar/gkh899

Raleigh EA. 1992. Organization and function of the mcrBC genes of Escherichia coli K-12. Mol Microbiol 6: 1079–1086. 10.1111/j.1365-2958.1992.tb01546.x

Raleigh EA, Wilson G. 1986. Escherichia coli K-12 restricts DNA containing 5-methylcytosine. Proc Natl Acad Sci 83: 9070–9074. 10.1073/pnas.83.23.9070

Ramalingam S, Kandavelou K, Rajenderan R, Chandrasegaran S. 2011. Creating designed zinc-finger nucleases with minimal cytotoxicity. J Mol Biol 405: 630–641. 10.1016/j.jmb.2010.10.043

Ramalingam S, London V, Kandavelou K, Cebotaru L, Guggino W, Civin C, Chandrasegaran S. 2013. Generation and genetic engineering of human induced pluripotent stem cells using designed zinc finger nucleases. Stem Cells Dev 22: 595–610. 10.1089/scd.2012.0245

Ramalingam S, Annaluru N, Kandavelou K, Chandrasegaran S. 2014. TALEN-mediated generation and genetic correction of disease-specific human induced pluripotent stem cells. Curr Gene Ther 14: 461–472. 10.2174/1566523214666140918101725

Ramanathan A, Agarwal PK. 2011. Evolutionarily conserved linkage between enzyme fold, flexibility, and catalysis. PLoS Biol 9: e1001193. doi:10.1371/journal.pbio.1001193. 10.1371/journal.pbio.1001193

Ramanathan SP, van Aelst K, Sears A, Peakman LJ, Diffin FM, Szczelkun MD, Seidel R. 2009. Type III restriction enzymes communicate in 1D without looping between their target sites. Proc Natl Acad Sci 106: 1748–1753. 10.1073/pnas.0807193106

Rao DN, Dryden DT, Bheemanaik S. 2014. Type III restriction-modification enzymes: a historical perspective. Nucleic Acids Res 42: 45–55. 10.1093/nar/gkt616

Rasko T, Der A, Klement E, Slaska-Kiss K, Posfai E, Medzihradszky KF, Marshak DR, Roberts RJ, Kiss A. 2010. BspRI restriction endonuclease: cloning, expression in Escherichia coli and sequential cleavage mechanism. Nucleic Acids Res 38: 7155–7166. 10.1093/nar/gkq567

Ratner D. 1974. The interaction bacterial and phage proteins with immobilized Escherichia coli RNA polymerase. J Mol Biol 88: 373–383. 10.1016/0022-2836(74)90488-4

Reich S, Gossl I, Reuter M, Rabe JP, Kruger DH. 2004. Scanning force microscopy of DNA translocation by the Type III restriction enzyme EcoP15I. J Mol Biol 341: 337–343. 10.1016/j.jmb.2004.06.031

Reuter M, Mucke M, Kruger DH. 2004. Structure and function of Type IIE restriction endonucleases—or: from a plasmid that restricts phage replication to a new molecular DNA recognition mechanism. In Restriction endonucleases (ed. Pingoud A), pp. 261–295. Springer, Berlin.

Revel HR, Luria SE. 1970. DNA-glucosylation in T-even phage: genetic determination and role in phagehost interaction. Ann Rev Genet 4: 177–192. 10.1146/annurev.ge.04.120170.001141

Rezulak M, Borsuk I, Mruk I. 2016. Natural C-independent expression of restriction endonuclease in a C protein-associated restriction-modification system. Nucleic Acids Res 44: 2646–2660. 10.1093/nar/gkv1331

Rifat D, Wright NT, Varney KM, Weber DJ, Black LW. 2008. Restriction endonuclease inhibitor IPI* of bacteriophage T4: a novel structure for a dedicated target. J Mol Biol 375: 720–734. 10.1016/j.jmb.2007.10.064

Rimseliene R, Vaisvila R, Janulaitis A. 1995. The eco72IC gene specifies a trans-acting factor which influences expression of both DNA methyltransferase and endonuclease from the Eco72I restriction-modification system. Gene 157: 217–219. 10.1016/0378-1119(94)00794-S

Rimseliene R, Maneliene Z, Lubys A, Janulaitis A. 2003. Engineering of restriction endonucleases: using methylation activity of the bifunctional endonuclease Eco57I to select the mutant with a novel sequence specificity. J Mol Biol 327: 383–391. 10.1016/S0022-2836(03)00142-6

Roberts RJ, Cheng X. 1998. Base flipping. Ann Rev Biochem 67: 181–198. 10.1146/annurev.biochem.67.1.181

Roberts RJ, Halford SE. 1993. Type II restriction enzymes. In Nucleases, 2nd ed. (ed. Linn SM, Lloyd RS, Roberts RJ), pp. 35–88. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.

Roberts RJ, Belfort M, Bestor T, Bhagwat AS, Bickle TA, Bitinaite J, Blumenthal RM, Degtyarev S, Dryden DT, Dybvig K, et al. 2003. A nomenclature for restriction enzymes, DNA methyltransferases, homing endonucleases and their genes. Nucleic Acids Res 31: 1805–1812. 10.1093/nar/gkg274

Roberts RJ, Vincze T, Posfai J, Macelis D. 2010. REBASE—a database for DNA restriction and modification: enzymes, genes and genomes. Nucleic Acids Res 38: D234–D236. 10.1093/nar/gkp874

Roberts GA, Cooper LP, White JH, Su TJ, Zipprich JT, Geary P, Kennedy C, Dryden DT. 2011. An investigation of the structural requirements for ATP hydrolysis and DNA cleavage by the EcoKI Type I DNA restriction and modification enzyme. Nucleic Acids Res 39: 7667–7676. 10.1093/nar/gkr480

Roberts RJ, Vincze T, Pósfai J, Macelis D. 2015. REBASE—a database for DNA restriction and modification: enzymes, genes and genomes. Nucleic Acids Res 43: D298–D299. 10.1093/nar/gku1046

Robertson BD, Meyer TF. 1992. Genetic variation in pathogenic bacteria. Trends Genet 8: 422–427. 10.1016/0168-9525(92)90325-X

Rogers JM, Barrera LA, Reyon D, Sander JD, Kellis M, Joung JK, Bulyk ML. 2015. Context influences on TALE-DNA binding revealed by quantitative profiling. Nat Commun 6: 7440. doi:10.1038/ncomms8440. 10.1038/ncomms8440

Rohs R, Jin X, West SM, Joshi R, Honig B, Mann RS. 2010. Origins of specificity in protein-DNA recognition. Ann Rev Biochem 79: 233–269. 10.1146/annurev-biochem-060408-091030

Rusling DA, Laurens N, Pernstich C, Wuite GJ, Halford SE. 2012. DNA looping by FokI: the impact of synapse geometry on loop topology at varied site orientations. Nucleic Acids Res 40: 4977–4987. 10.1093/nar/gks183

Rutkauskas D, Petkelyte M, Naujalis P, Sasnauskas G, Tamulaitis G, Zaremba M, Šikšnys V. 2014. Restriction enzyme Ecl18kI-induced DNA looping dynamics by single-molecule FRET. J Phys Chem B 118: 8575–8582. 10.1021/jp504546v

Ryan KA, Lo RY. 1999. Characterization of a CACAG pentanucleotide repeat in Pasteurella haemolytica and its possible role in modulation of a novel type III restriction-modification system. Nucleic Acids Res 27: 1505–1511. 10.1093/nar/27.6.1505

Sanchez-Romero MA, Cota I, Casadesus J. 2015. DNA methylation in bacteria: from the methyl group to the methylome. Curr Opin Microbiol 25: 9–16. 10.1016/j.mib.2015.03.004

Sanders KL, Catto LE, Bellamy SR, Halford SE. 2009. Targeting individual subunits of the FokI restriction endonuclease to specific DNA strands. Nucleic Acids Res 37: 2105–2115. 10.1093/nar/gkp046

Sapienza PJ, Dela Torre CA, McCoy WHt, Jana SV, Jen-Jacobson L. 2005. Thermodynamic and kinetic basis for the relaxed DNA sequence specificity of “promiscuous” mutant EcoRI endonucleases. J Mol Biol 348: 307–324. 10.1016/j.jmb.2005.02.051

Sapienza PJ, Rosenberg JM, Jen-Jacobson L. 2007. Structural and thermodynamic basis for enhanced DNA binding by a promiscuous mutant EcoRI endonuclease. Structure 15: 1368–1382. 10.1016/j.str.2007.09.014

Sapienza PJ, Niu T, Kurpiewski MR, Grigorescu A, Jen-Jacobson L. 2014. Thermodynamic and structural basis for relaxation of specificity in protein-DNA recognition. J Mol Biol 426: 84–104. 10.1016/j.jmb.2013.09.005

Sapranauskas R, Sasnauskas G, Lagunavicius A, Vilkaitis G, Lubys A, Šikšnys V. 2000. Novel subtype of type IIs restriction enzymes. BfiI endonuclease exhibits similarities to the EDTA-resistant nuclease Nuc of Salmonella typhimurium. J Biol Chem 275: 30878–30885. 10.1074/jbc.M003350200

Saravanan M, Bujnicki JM, Cymerman IA, Rao DN, Nagaraja V. 2004. Type II restriction endonuclease R.KpnI is a member of the HNH nuclease superfamily. Nucleic Acids Res 32: 6129–6135. 10.1093/nar/gkh951

Saravanan M, Vasu K, Ghosh S, Nagaraja V. 2007a. Dual role for Zn2+ in maintaining structural integrity and inducing DNA sequence specificity in a promiscuous endonuclease. J Biol Chem 282: 32320–32326. 10.1074/jbc.M705927200

Saravanan M, Vasu K, Kanakaraj R, Rao DN, Nagaraja V. 2007b. R.KpnI, an HNH superfamily REase, exhibits differential discrimination at non-canonical sequences in the presence of Ca2+ and Mg2+. Nucleic Acids Res 35: 2777–2786. 10.1093/nar/gkm114

Sarrade-Loucheur A, Xu SY, Chan SH. 2013. The role of the methyltransferase domain of bifunctional restriction enzyme RM.BpuSI in cleavage activity. PloS One 8: e80967. doi:10.1371/journal.pone.0080967. 10.1371/journal.pone.0080967

Sasnauskas G, Halford SE, Šikšnys V. 2003. How the BfiI restriction enzyme uses one active site to cut two DNA strands. Proc Natl Acad Sci 100: 6410–6415. 10.1073/pnas.1131003100

Sasnauskas G, Connolly BA, Halford SE, Šikšnys V. 2007. Site-specific DNA transesterification catalyzed by a restriction enzyme. Proc Natl Acad Sci 104: 2115–2120. 10.1073/pnas.0608689104

Sasnauskas G, Connolly BA, Halford SE, Šikšnys V. 2008. Template-directed addition of nucleosides to DNA by the BfiI restriction enzyme. Nucleic Acids Res 36: 3969–3977. 10.1093/nar/gkn343

Sasnauskas G, Zakrys L, Zaremba M, Cosstick R, Gaynor JW, Halford SE, Šikšnys V. 2010. A novel mechanism for the scission of double-stranded DNA: BfiI cuts both 3′–5′ and 5′–3′ strands by rotating a single active site. Nucleic Acids Res 38: 2399–2410. 10.1093/nar/gkp1194

Sasnauskas G, Kostiuk G, Tamulaitis G, Šikšnys V. 2011. Target site cleavage by the monomeric restriction enzyme BcnI requires translocation to a random DNA sequence and a switch in enzyme orientation. Nucleic Acids Res 39: 8844–8856. 10.1093/nar/gkr588

Sasnauskas G, Tamulaitienė G, Tamulaitis G, Calyseva J, Laime M, Šikšnys V. 2015a. Biochemical and structural characterization of a monomeric Type IIE REase. In 7th NEB meeting on restriction and modification (Gdansk). Short talk S1.

Sasnauskas G, Zagorskaite E, Kauneckaite K, Tamulaitienė G, Šikšnys V. 2015b. Structure-guided sequence specificity engineering of the modification-dependent restriction endonuclease LpnPI. Nucleic Acids Res 43: 6144–6155. 10.1093/nar/gkv548

Sasnauskas G, Tamulaitienė G, Tamulaitis G, Calyseva J, Laime M, Rimseliene R, Lubys A, Šikšnys V. 2017. UbaLAI is a monomeric Type IIE restriction enzyme. Nucleic Acids Res 45: 9583–9594. 10.1093/nar/gkx634

Sawaya MR, Zhu Z, Mersha F, Chan SH, Dabur R, Xu SY, Balendiran GK. 2005. Crystal structure of the restriction-modification system control element C.Bcll and mapping of its binding site. Structure 13: 1837–1847. 10.1016/j.str.2005.08.017

Schierling B, Noel AJ, Wende W, Hien le T, Volkov E, Kubareva E, Oretskaya T, Kokkinidis M, Rompp A, Spengler B, et al. 2010. Controlling the enzymatic activity of a restriction enzyme by light. Proc Natl Acad Sci 107: 1361–1366. 10.1073/pnas.0909444107

Schierling B, Dannemann N, Gabsalilow L, Wende W, Cathomen T, Pingoud A. 2012. A novel zinc-finger nuclease platform with a sequence-specific cleavage module. Nucleic Acids Res 40: 2623–2638. 10.1093/nar/gkr1112

Schwarz FW, Toth J, van Aelst K, Cui G, Clausing S, Szczelkun MD, Seidel R. 2013. The helicase-like domains of type III restriction enzymes trigger long-range diffusion along DNA. Science 340: 353–356. 10.1126/science.1231122

Schwarz FW, van Aelst K, Toth J, Seidel R, Szczelkun MD. 2011. DNA cleavage site selection by Type III restriction enzymes provides evidence for head-on protein collisions following 1D bidirectional motion. Nucleic Acids Res 39: 8042–8051. 10.1093/nar/gkr502

Sears LE, Zhou B, Aliotta JM, Morgan RD, Kong H. 1996. BaeI, another unusual BcgI-like restriction endonuclease. Nucleic Acids Res 24: 3590–3592. 10.1093/nar/24.18.3590

Seib KL, Peak IR, Jennings MP. 2002. Phase variable restriction-modification systems in Moraxella catarrhalis. FEMS Immunol Med Microbiol 32: 159–165.

Seidel R, Dekker C. 2007. Single-molecule studies of nucleic acid motors. Curr Opin Struct Biol 17: 80–86. 10.1016/j.sbi.2006.12.003

Seidel R, van Noort J, van der Scheer C, Bloom JG, Dekker NH, Dutta CF, Blundell A, Robinson T, Firman K, Dekker C. 2004. Real-time observation of DNA translocation by the type I restriction modification enzyme EcoR124I. Nat Struct Mol Biol 11: 838–843. 10.1038/nsmb816

Seidel R, Bloom JG, van Noort J, Dutta CF, Dekker NH, Firman K, Szczelkun MD, Dekker C. 2005. Dynamics of initiation, termination and reinitiation of DNA translocation by the motor protein EcoR124I. EMBO J 24: 4188–4197. 10.1038/sj.emboj.7600881

Seidel R, Bloom JG, Dekker C, Szczelkun MD. 2008. Motor step size and ATP coupling efficiency of the dsDNA translocase EcoR124I. EMBO J 27: 1388–1398. 10.1038/emboj.2008.69

Semenova E, Minakhin L, Bogdanova E, Nagornykh M, Vasilov A, Heyduk T, Solonin A, Zakharova M, Severinov K. 2005. Transcription regulation of the EcoRV restriction-modification system. Nucleic Acids Res 33: 6942–6951. 10.1093/nar/gki998

Serfiotis-Mitsa D, Roberts GA, Cooper LP, White JH, Nutley M, Cooper A, Blakely GW, Dryden DT. 2008. The Orf18 gene product from conjugative transposon Tn916 is an ArdA antirestriction protein that inhibits type I DNA restriction-modification systems. J Mol Biol 383: 970–981. 10.1016/j.jmb.2008.06.005

Serfiotis-Mitsa D, Herbert AP, Roberts GA, Soares DC, White JH, Blakely GW, Uhrin D, Dryden DT. 2010. The structure of the KlcA and ArdB proteins reveals a novel fold and antirestriction activity against Type I DNA restriction systems in vivo but not in vitro. Nucleic Acids Res 38: 1723–1737. 10.1093/nar/gkp1144

Shao C, Wang C, Zang J. 2014. Structural basis for the substrate selectivity of PvuRts1I, a 5-hydroxymethylcytosine DNA restriction endonuclease. Acta Crystallogr, Sect D: Biol Crystallogr 70: 2477–2486. 10.1107/S139900471401606X

Shen BW, Heiter DF, Chan SH, Wang H, Xu SY, Morgan RD, Wilson GG, Stoddard BL. 2010. Unusual target site disruption by the rare-cutting HNH restriction endonuclease PacI. Structure 18: 734–743. 10.1016/j.str.2010.03.009

Shen BW, Xu D, Chan SH, Zheng Y, Zhu Z, Xu SY, Stoddard BL. 2011. Characterization and crystal structure of the type IIG restriction endonuclease RM.BpuSI. Nucleic Acids Res 39: 8223–8236. 10.1093/nar/gkr543

Shen BW, Heiter DF, Lunnen KD, Wilson GG, Stoddard BL. 2015. Crystal structure of a 8-bp REase, SwaI. In 7th NEB meeting on restriction and modification (Gdansk): Poster P3.

Shen BW, Heiter DF, Lunnen KD, Wilson GG, Stoddard BL. 2017. DNA recognition by the SwaI restriction endonuclease involves unusual distortion of an 8 base pair A:T-rich target. Nucleic Acids Res 45: 1516–1528. 10.1093/nar/gkw1200

Shilov I, Tashlitsky V, Khodoun M, Vasil'ev S, Alekseev Y, Kuzubov A, Kubareva E, Karyagina A. 1998. DNA-methyltransferase SsoII interaction with own promoter region binding site. Nucleic Acids Res 26: 2659–2664. 10.1093/nar/26.11.2659

Shlyakhtenko LS, Gilmore J, Portillo A, Tamulaitis G, Šikšnys V, Lyubchenko YL. 2007. Direct visualization of the EcoRII-DNA triple synaptic complex by atomic force microscopy. Biochemistry 46: 11128–11136. 10.1021/bi701123u

Simoncsits A, Tjornhammar ML, Rasko T, Kiss A, Pongor S. 2001. Covalent joining of the subunits of a homodimeric type II restriction endonuclease: single-chain PvuII endonuclease. J Mol Biol 309: 89–97. 10.1006/jmbi.2001.4651

Simons M, Szczelkun MD. 2011. Recycling of protein subunits during DNA translocation and cleavage by Type I restriction-modification enzymes. Nucleic Acids Res 39: 7656–7666. 10.1093/nar/gkr479

Simons M, Diffin FM, Szczelkun MD. 2014. ClpXP protease targets long-lived DNA translocation states of a helicase-like motor to cause restriction alleviation. Nucleic Acids Res 42: 12082–12091. 10.1093/nar/gku851

Singleton MR, Dillingham MS, Wigley DB. 2007. Structure and mechanism of helicases and nucleic acid translocases. Ann Rev Biochem 76: 23–50. 10.1146/annurev.biochem.76.052305.115300

Sisakova E, Stanley LK, Weiserova M, Szczelkun MD. 2008a. A RecB-family nuclease motif in the Type I restriction endonuclease EcoR124I. Nucleic Acids Res 36: 3939–3949. 10.1093/nar/gkn333

Sisakova E, Weiserova M, Dekker C, Seidel R, Szczelkun MD. 2008b. The interrelationship of helicase and nuclease domains during DNA translocation by the molecular motor EcoR124I. J Mol Biol 384: 1273–1286. 10.1016/j.jmb.2008.10.017

Sisakova E, van Aelst K, Diffin FM, Szczelkun MD. 2013. The Type ISP restriction-modification enzymes LlaBIII and LlaGI use a translocation-collision mechanism to cleave non-specific DNA distant from their recognition sites. Nucleic Acids Res 41: 1071–1080. 10.1093/nar/gks1209

Šikšnys V, Gražulis S, Huber R. 2004. Structure and function of the tetrameric restriction enzymes. In Restriction endonucleases (ed. Pingoud A), pp. 237–259. Springer, Berlin.

Šikšnys V, Skirgaila R, Sasnauskas G, Urbanke C, Cherny D, Grazulis S, Huber R. 1999. The Cfr10I restriction enzyme is functional as a tetramer. J Mol Biol 291: 1105–1118. 10.1006/jmbi.1999.2977

Sitaraman R, Dybvig K. 1997. The hsd loci of Mycoplasma pulmonis: organization, rearrangements and expression of genes. Mol Microbioly 26: 109–120. 10.1046/j.1365-2958.1997.5571938.x

Siwek W, Czapinska H, Bochtler M, Bujnicki JM, Skowronek K. 2012. Crystal structure and mechanism of action of the N6-methyladenine-dependent type IIM restriction endonuclease R.DpnI. Nucleic Acids Res 40: 7563–7572. 10.1093/nar/gks428

Skirgaila R, Šikšnys V. 1998. Ca2+-ions stimulate DNA binding specificity of Cfr10I restriction enzyme. Biol Chem 379: 595–598.

Skirgaila R, Grazulis S, Bozic D, Huber R, Šikšnys V. 1998. Structure-based redesign of the catalytic/metal binding site of Cfr10I restriction endonuclease reveals importance of spatial rather than sequence conservation of active centre residues. J Mol Biol 279: 473–481. 10.1006/jmbi.1998.1803

Skoglund A, Bjorkholm B, Nilsson C, Andersson AF, Jernberg C, Schirwitz K, Enroth C, Krabbe M, Engstrand L. 2007. Functional analysis of the M.HpyAIV DNA methyltransferase of Helicobacter pylori. J Bacteriol 189: 8914–8921. 10.1128/JB.00108-07

Skowron P, Kaczorowski T, Tucholski J, Podhajska AJ. 1993. Atypical DNA-binding properties of class-IIS restriction endonucleases: evidence for recognition of the cognate sequence by a FokI monomer. Gene 125: 1–10. 10.1016/0378-1119(93)90738-O

Skowron PM, Harasimowicz R, Rutkowska SM. 1996. GCN4 eukaryotic transcription factor/FokI endonuclease-mediated ‘Achilles’ heel cleavage': quantitative study of protein-DNA interaction. Gene 170: 1–8. 10.1016/0378-1119(95)00857-8

Skowron PM, Majewski J, Zylicz-Stachula A, Rutkowska SM, Jaworowska I, Harasimowicz-Slowinska RI. 2003. A new Thermus sp. class-IIS enzyme sub-family: isolation of a ‘twin’ endonuclease TspDTI with a novel specificity 5′-ATGAA-3′, related to TspGWI, TaqII and Tth111II. Nucleic Acids Res 31: e74. PMC167652. 10.1093/nar/gng074

Smith RM, Diffin FM, Savery NJ, Josephsen J, Szczelkun MD. 2009a. DNA cleavage and methylation specificity of the single polypeptide restriction-modification enzyme LlaGI. Nucleic Acids Res 37: 7206–7218. 10.1093/nar/gkp790

Smith RM, Josephsen J, Szczelkun MD. 2009b. An Mrr-family nuclease motif in the single polypeptide restriction-modification enzyme LlaGI. Nucleic Acids Res 37: 7231–7238. 10.1093/nar/gkp795

Smith RM, Josephsen J, Szczelkun MD. 2009c. The single polypeptide restriction-modification enzyme LlaGI is a self-contained molecular motor that translocates DNA loops. Nucleic Acids Res 37: 7219–7230. 10.1093/nar/gkp794

Smith RM, Jacklin AJ, Marshall JJ, Sobott F, Halford SE. 2013a. Organization of the BcgI restriction-modification protein for the transfer of one methyl group to DNA. Nucleic Acids Res 41: 405–417. 10.1093/nar/gks1000

Smith RM, Marshall JJ, Jacklin AJ, Retter SE, Halford SE, Sobott F. 2013b. Organization of the BcgI restriction-modification protein for the cleavage of eight phosphodiester bonds in DNA. Nucleic Acids Res 41: 391–404. 10.1093/nar/gks1023

Smith RM, Pernstich C, Halford SE. 2014. TstI, a Type II restriction-modification protein with DNA recognition, cleavage and methylation functions in a single polypeptide. Nucleic Acids Res 42: 5809–5822. 10.1093/nar/gku187

Sohail A, Ives CL, Brooks JE. 1995. Purification and characterization of C.BamHI, a regulator of the BamHI restriction-modification system. Gene 157: 227–228. 10.1016/0378-1119(94)00698-R

Sokolowska M, Kaus-Drobek M, Czapinska H, Tamulaitis G, Šikšnys V, Bochtler M. 2007a. Restriction endonucleases that resemble a component of the bacterial DNA repair machinery. Cell Mol Life Sci 64: 2351–2357. 10.1007/s00018-007-7124-9

Sokolowska M, Kaus-Drobek M, Czapinska H, Tamulaitis G, Szczepanowski RH, Urbanke C, Šikšnys V, Bochtler M. 2007b. Monomeric restriction endonuclease BcnI in the apo form and in an asymmetric complex with target DNA. J Mol Biol 369: 722–734. 10.1016/j.jmb.2007.03.018

Sokolowska M, Czapinska H, Bochtler M. 2009. Crystal structure of the ββα-Me type II restriction endonuclease Hpy99I with target DNA. Nucleic Acids Res 37: 3799–3810. 10.1093/nar/gkp228

Sokolowska M, Czapinska H, Bochtler M. 2011. Hpy188I–DNA pre- and post-cleavage complexes—snapshots of the GIY-YIG nuclease mediated catalysis. Nucleic Acids Res 39: 1554–1564. 10.1093/nar/gkq821

Srikhanta YN, Maguire TL, Stacey KJ, Grimmond SM, Jennings MP. 2005. The phasevarion: a genetic system controlling coordinated, random switching of expression of multiple genes. Proc Natl Acad Sci 102: 5547–5551. 10.1073/pnas.0501169102

Srikhanta YN, Fox KL, Jennings MP. 2010. The phasevarion: phase variation of type III DNA methyltransferases controls coordinated switching in multiple genes. Nat Rev Microbiol 8: 196–206. 10.1038/nrmicro2283

Stahl F, Wende W, Jeltsch A, Pingoud A. 1996. Introduction of asymmetry in the naturally symmetric restriction endonuclease EcoRV to investigate intersubunit communication in the homodimeric protein. Proc Natl Acad Sci 93: 6175–6180. 10.1073/pnas.93.12.6175

Stanley LK, Seidel R, van der Scheer C, Dekker NH, Szczelkun MD, Dekker C. 2006. When a helicase is not a helicase: dsDNA tracking by the motor protein EcoR124I. EMBO J 25: 2230–2239. 10.1038/sj.emboj.7601104

Steczkiewicz K, Muszewska A, Knizewski L, Rychlewski L, Ginalski K. 2012. Sequence, structure and functional diversity of PD-(D/E)XK phosphodiesterase superfamily. Nucleic Acids Res 40: 7016–7045. 10.1093/nar/gks382

Stephanou AS, Roberts GA, Cooper LP, Clarke DJ, Thomson AR, MacKay CL, Nutley M, Cooper A, Dryden DT. 2009a. Dissection of the DNA mimicry of the bacteriophage T7 Ocr protein using chemical modification. J Mol Biol 391: 565–576. 10.1016/j.jmb.2009.06.020

Stephanou AS, Roberts GA, Tock MR, Pritchard EH, Turkington R, Nutley M, Cooper A, Dryden DT. 2009b. A mutational analysis of DNA mimicry by ocr, the gene 0.3 antirestriction protein of bacteriophage T7. Biochem Biophys Res Commun 378: 129–132. 10.1016/j.bbrc.2008.11.014

Stephenson FH, Ballard BT, Boyer HW, Rosenberg JM, Greene PJ. 1989. Comparison of the nucleotide and amino acid sequences of the RsrI and EcoRI restriction endonucleases. Gene 85: 1–13. 10.1016/0378-1119(89)90458-7

Stern A, Sorek R. 2011. The phage-host arms race: shaping the evolution of microbes. Bioessays 33: 43–51. 10.1002/bies.201000071

Stewart FJ, Raleigh EA. 1998. Dependence of McrBC cleavage on distance between recognition elements. Biol Chem 379: 611–616.

Stewart FJ, Panne D, Bickle TA, Raleigh EA. 2000. Methyl-specific DNA binding by McrBC, a modification-dependent restriction enzyme. J Mol Biol 298: 611–622. 10.1006/jmbi.2000.3697

Stoddard BL. 2005. Homing endonuclease structure and function. Q. Rev Biophys 38: 49–95. 10.1017/S0033583505004063

Stower H. 2014. Epigenetics: reprogramming with TET. Nat Rev Genet 15: 66. doi:10.1038/nrg3659.

Streeter SD, Papapanagiotou I, McGeehan JE, Kneale GG. 2004. DNA footprinting and biophysical characterization of the controller protein C.AhdI suggests the basis of a genetic switch. Nucleic Acids Res 32: 6445–6453. 10.1093/nar/gkh975

Studier FW. 1975. Gene 0.3 of bacteriophage T7 acts to overcome the DNA restriction system of the host. J Mol Biol 94: 283–295. 10.1016/0022-2836(75)90083-2

Studier FW. 2013. Phage T7 is neither modified nor restircted by EcoKI or EcoBI, which led to the discovery of the T7 0.3 gene encoding Ocr, as well as the collision model for Type I enzymes. http://library.cshl.edu/Meetings/restriction-enzymes/v-Studier.php.

Studier FW, Movva NR. 1976. SAMase gene of bacteriophage T3 is responsible for overcoming host restriction. J Virol 19: 136–145.

Su TJ, Tock MR, Egelhaaf SU, Poon WC, Dryden DT. 2005. DNA bending by M.EcoKI methyltransferase is coupled to nucleotide flipping. Nucleic Acids Res 33: 3235–3244. 10.1093/nar/gki618

Sudina AE, Zatsepin TS, Pingoud V, Pingoud A, Oretskaya TS, Kubareva EA. 2005. Affinity modification of the restriction endonuclease SsoII by 2′-aldehyde-containing double stranded DNAs. Biochem Biokhim 70: 941–947. 10.1007/s10541-005-0206-0

Sugisaki H, Kanazawa S. 1981. New restriction endonucleases from Flavobacterium okeanokoites (FokI) and Micrococcus luteus (MluI). Gene 16: 73–78. 10.1016/0378-1119(81)90062-7

Sugisaki H, Kita K, Takanami M. 1989. The FokI restriction-modification system. II. Presence of two domains in FokI methylase responsible for modification of different DNA strands. J Biol Chem 264: 5757–5761.

Sukackaite R, Gražulis S, Tamulaitis G, Šikšnys V. 2012. The recognition domain of the methyl-specific endonuclease McrBC flips out 5-methylcytosine. Nucleic Acids Res 40: 7552–7562. 10.1093/nar/gks332

Sutherland E, Coe L, Raleigh EA. 1992. McrBC: a multisubunit GTP-dependent restriction endonuclease. J Mol Biol 225: 327–348. 10.1016/0022-2836(92)90925-A

Szczelkun MD. 2011. Translocation, switching and gating: potential roles for ATP in long-range communication on DNA by Type III restriction endonucleases. Biochem Soc Trans 39: 589–594. 10.1042/BST0390589

Szczelkun MD. 2013. Roles for helicases as ATP-dependent molecular switches. Adv Exp Med Biol 767: 225–244. 10.1007/978-1-4614-5037-5_11

Szczelkun MD, Friedhoff P, Seidel R. 2010. Maintaining a sense of direction during long-range communication on DNA. Biochem Soc Trans 38: 404–409. 10.1042/BST0380404

Szczepanowski RH, Carpenter MA, Czapinska H, Zaremba M, Tamulaitis G, Šikšnys V, Bhagwat AS, Bochtler M. 2008. Central base pair flipping and discrimination by PspGI. Nucleic Acids Res 36: 6109–6117. 10.1093/nar/gkn622

Szczepek M, Brondani V, Buchel J, Serrano L, Segal DJ, Cathomen T. 2007. Structure-based redesign of the dimerization interface reduces the toxicity of zinc-finger nucleases. Nat Biotechnol 25: 786–793. 10.1038/nbt1317

Szczepek M, Mackeldanz P, Moncke-Buchner E, Alves J, Kruger DH, Reuter M. 2009. Molecular analysis of restriction endonuclease EcoRII from Escherichia coli reveals precise regulation of its enzymatic activity by autoinhibition. Mol Microbiol 72: 1011–1021. 10.1111/j.1365-2958.2009.06702.x

Szwagierczak A, Brachmann A, Schmidt CS, Bultmann S, Leonhardt H, Spada F. 2011. Characterization of PvuRts1I endonuclease as a tool to investigate genomic 5-hydroxymethylcytosine. Nucleic Acids Res 39: 5149–5156. 10.1093/nar/gkr118

Szybalski W, Kim SC, Hasan N, Podhajska AJ. 1991. Class-IIS restriction enzymes—a review. Gene 100: 13–26. 10.1016/0378-1119(91)90345-C

Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, Brudno Y, Agarwal S, Iyer LM, Liu DR, Aravind L, et al. 2009. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324: 930–935. 10.1126/science.1170116

Tamulaitienė G, Silanskas A, Grazulis S, Zaremba M, Siksnys V. 2014. Crystal structure of the R-protein of the multisubunit ATP-dependent restriction endonuclease NgoAVII. Nucleic Acids Res 42: 14022–14030. 10.1093/nar/gku1237

Tamulaitienė G, Jovaisaite V, Tamulaitis G, Songailiene I, Manakova E, Zaremba M, Gražulis S, Xu SY, Šikšnys V. 2017. Restriction endonuclease AgeI is a monomer which dimerizes to cleave DNA. Nucleic Acids Res 45: 3547–3558.

Tamulaitis G, Mucke M, Šikšnys V. 2006a. Biochemical and mutational analysis of EcoRII functional domains reveals evolutionary links between restriction enzymes. FEBS Lett 580: 1665–1671. 10.1016/j.febslet.2006.02.010

Tamulaitis G, Sasnauskas G, Mucke M, Šikšnys V. 2006b. Simultaneous binding of three recognition sites is necessary for a concerted plasmid DNA cleavage by EcoRII restriction endonuclease. J Mol Biol 358: 406–419. 10.1016/j.jmb.2006.02.024

Tamulaitis G, Zaremba M, Szczepanowski RH, Bochtler M, Šikšnys V. 2007. Nucleotide flipping by restriction enzymes analyzed by 2-aminopurine steady-state fluorescence. Nucleic Acids Res 35: 4792–4799. 10.1093/nar/gkm513

Tamulaitis G, Zaremba M, Szczepanowski RH, Bochtler M, Šikšnys V. 2008. How PspGI, catalytic domain of EcoRII and Ecl18kI acquire specificities for different DNA targets. Nucleic Acids Res 36: 6101–6108. 10.1093/nar/gkn621

Tamulaitis G, Rutkauskas M, Zaremba M, Gražulis S, Tamulaitiene G, Šikšnys V. 2015. Functional significance of protein assemblies predicted by the crystal structure of the restriction endonuclease BsaWI. Nucleic Acids Res 43: 8100–8110. 10.1093/nar/gkv768

Tan A, Atack JM, Jennings MP, Seib KL. 2016. The capricious nature of bacterial pathogens: phasevarions and vaccine development. Front Immunol 7: 586. PMC5149525. 10.3389/fimmu.2016.00586

Tao T, Blumenthal RM. 1992. Sequence and characterization of pvuIIR, the PvuII endonuclease gene, and of pvuIIC, its regulatory gene. J Bacteriol 174: 3395–3398. 10.1128/jb.174.10.3395-3398.1992

Tao T, Bourne JC, Blumenthal RM. 1991. A family of regulatory genes associated with type II restriction-modification systems. J Bacteriol 173: 1367–1375. 10.1128/jb.173.4.1367-1375.1991

Tautz N, Kaluza K, Frey B, Jarsch M, Schmitz GG, Kessler C. 1990. SgrAI, a novel class-II restriction endonuclease from Streptomyces griseus recognizing the octanucleotide sequence 5′-CR/CCGGYG-3′[corrected]. Nucleic Acids Res 18: 3087. PMCID: 330872. 10.1093/nar/18.10.3087

Taylor IA, Davis KG, Watts D, Kneale GG. 1994. DNA-binding induces a major structural transition in a type I methyltransferase. EMBO J 13: 5772–5778. 10.1002/j.1460-2075.1994.tb06915.x

Thanisch K, Schneider K, Morbitzer R, Solovei I, Lahaye T, Bultmann S, Leonhardt H. 2014. Targeting and tracing of specific DNA sequences with dTALEs in living cells. Nucleic Acids Res 42: e38. doi:10.1093/nar/gkt1348. 10.1093/nar/gkt1348

The 7th NEB Meeting on DNA Restriction and Modification, August 24–29, 2015. Uniwersytet Gdanski ulica, Kladki 24, Gdańsk, Poland.

Thomas CB, Gumport RI. 2006. Dimerization of the bacterial RsrI N6-adenine DNA methyltransferase. Nucleic Acids Res 34: 806–815. 10.1093/nar/gkj486

Thomas AT, Brammar WJ, Wilkins BM. 2003. Plasmid R16 ArdA protein preferentially targets restriction activity of the type I restriction-modification system EcoKI. J Bacteriol 185: 2022–2025. 10.1128/JB.185.6.2022-2025.2003

Tock MR, Dryden DT. 2005. The biology of restriction and anti-restriction. Curr Opin Microbiol 8: 466–472. 10.1016/j.mib.2005.06.003

Too PH, Zhu Z, Chan SH, Xu SY. 2010. Engineering Nt.BtsCI and Nb.BtsCI nicking enzymes and applications in generating long overhangs. Nucleic Acids Res 38: 1294–1303. 10.1093/nar/gkp1092

Toth J, Bollins J, Szczelkun MD. 2015. Re-evaluating the kinetics of ATP hydrolysis during initiation of DNA sliding by Type III restriction enzymes. Nucleic Acids Res 43: 10870–10881. 10.1093/nar/gkv1154

Townson SA, Samuelson JC, Vanamee ES, Edwards TA, Escalante CR, Xu SY, Aggarwal AK. 2004. Crystal structure of BstYI at 1.85A resolution: a thermophilic restriction endonuclease with overlapping specificities to BamHI and BglII. J Mol Biol 338: 725–733. 10.1016/j.jmb.2004.02.074

Townson SA, Samuelson JC, Xu SY, Aggarwal AK. 2005. Implications for switching restriction enzyme specificities from the structure of BstYI bound to a BglII DNA sequence. Structure 13: 791–801. 10.1016/j.str.2005.02.018

Tucholski J, Skowron PM, Podhajska AJ. 1995. MmeI, a class-IIS restriction endonuclease: purification and characterization. Gene 157: 87–92. 10.1016/0378-1119(94)00787-S

Tucholski J, Zmijewski JW, Podhajska AJ. 1998. Two intertwined methylation activities of the MmeI restriction-modification class-IIS system from Methylophilus methylotrophus. Gene 223: 293–302. 10.1016/S0378-1119(98)00450-8

Tuteja N, Tuteja R. 2004. Prokaryotic and eukaryotic DNA helicases. Essential molecular motor proteins for cellular machinery. Eur J Biochem/FEBS 271: 1835–1848. 10.1111/j.1432-1033.2004.04093.x

Tyndall C, Meister J, Bickle TA. 1994. The Escherichia coli prr region encodes a functional type IC DNA restriction system closely integrated with an anticodon nuclease gene. J Mol Biol 237: 266–274. 10.1006/jmbi.1994.1230

Umate P, Tuteja N, Tuteja R. 2011. Genome-wide comprehensive analysis of human helicases. Commun Integr Biol 4: 118–137. 10.4161/cib.13844

Urnov FD, Miller JC, Lee YL, Beausejour CM, Rock JM, Augustus S, Jamieson AC, Porteus MH, Gregory PD, Holmes MC. 2005. Highly efficient endogenous human gene correction using designed zinc-finger nucleases. Nature 435: 646–651. 10.1038/nature03556

Urnov FD, Rebar EJ, Holmes MC, Zhang HS, Gregory PD. 2010. Genome editing with engineered zinc finger nucleases. Nat Rev Genet 11: 636–646. 10.1038/nrg2842

Uyen NT, Nishi K, Park SY, Choi JW, Lee HJ, Kim JS. 2008. Crystallization and preliminary X-ray diffraction analysis of the HsdR subunit of a putative type I restriction enzyme from Vibrio vulnificus YJ016. Acta Crystallographica, Sect F: Struct Biol Cryst Commun 64: 926–928. 10.1107/S1744309108027516

Uyen NT, Park SY, Choi JW, Lee HJ, Nishi K, Kim JS. 2009. The fragment structure of a putative HsdR subunit of a type I restriction enzyme from Vibrio vulnificus YJ016: implications for DNA restriction and translocation activity. Nucleic Acids Res 37: 6960–6969. 10.1093/nar/gkp603

van Aelst K, Toth J, Ramanathan SP, Schwarz FW, Seidel R, Szczelkun MD. 2010. Type III restriction enzymes cleave DNA by long-range interaction between sites in both head-to-head and tail-to-tail inverted repeat. Proc Natl Acad Sci 107: 9123–9128. 10.1073/pnas.1001637107

van Aelst K, Saikrishnan K, Szczelkun MD. 2015. Mapping DNA cleavage by the Type ISP restriction-modification enzymes following long-range communication between DNA sites in different orientations. Nucleic Acids Res 43: 10430–10443.

van Belkum A, Scherer S, van Alphen L, Verbrugh H. 1998. Short-sequence DNA repeats in prokaryotic genomes. Microbiol Mol Biol Rev 62: 275–293.

van den Broek B, Vanzi F, Normanno D, Pavone FS, Wuite GJ. 2006. Real-time observation of DNA looping dynamics of Type IIE restriction enzymes NaeI and NarI. Nucleic Acids Res 34: 167–174. 10.1093/nar/gkj432

van der Woude MW. 2006. Re-examining the role and random nature of phase variation. FEMS Microbiol Lett 254: 190–197. 10.1111/j.1574-6968.2005.00038.x

van der Woude MW, Baumler AJ. 2004. Phase and antigenic variation in bacteria. Clin Microbiol Rev 17: 581–611. 10.1128/CMR.17.3.581-611.2004

van Ham SM, van Alphen L, Mooi FR, van Putten JP. 1993. Phase variation of H. influenzae fimbriae: transcriptional control of two divergent genes through a variable combined promoter region. Cell 73: 1187–1196. 10.1016/0092-8674(93)90647-9

van Noort J, van der Heijden T, Dutta CF, Firman K, Dekker C. 2004. Initiation of translocation by Type I restriction-modification enzymes is associated with a short DNA extrusion. Nucleic Acids Res 32: 6540–6547. 10.1093/nar/gkh999

Vanamee ES, Santagata S, Aggarwal AK. 2001. FokI requires two specific DNA sites for cleavage. J Mol Biol 309: 69–78. 10.1006/jmbi.2001.4635

Vanamee ES, Hsieh P, Zhu Z, Yates D, Garman E, Xu S, Aggarwal AK. 2003. Glucocorticoid receptor-like Zn(Cys)4 motifs in BslI restriction endonuclease. J Mol Biol 334: 595–603. 10.1016/j.jmb.2003.09.043

Vanamee ES, Viadiu H, Kucera R, Dorner L, Picone S, Schildkraut I, Aggarwal AK. 2005. A view of consecutive binding events from structures of tetrameric endonuclease SfiI bound to DNA. EMBO J 24: 4198–4208. 10.1038/sj.emboj.7600880

Vanamee ES, Berriman J, Aggarwal AK. 2007. An EM view of the FokI synaptic complex by single particle analysis. J Mol Biol 370: 207–212. 10.1016/j.jmb.2007.04.066

Vanamee ES, Viadiu H, Chan SH, Ummat A, Hartline AM, Xu SY, Aggarwal AK. 2011. Asymmetric DNA recognition by the OkrAI endonuclease, an isoschizomer of BamHI. Nucleic Acids Res 39: 712–719. 10.1093/nar/gkq779

VanderVeen LA, Harris TM, Jen-Jacobson L, Marnett LJ. 2008. Formation of DNA-protein cross-links between γ-hydroxypropanodeoxyguanosine and EcoRI. Chem Res Toxicol 21: 1733–1738. 10.1021/tx800092g

Vasu K, Nagaraja V. 2013. Diverse functions of restriction-modification systems in addition to cellular defense. Microbiol Mol Biol Rev 77: 53–72. 10.1128/MMBR.00044-12

Vasu K, Nagamalleswari E, Nagaraja V. 2012. Promiscuous restriction is a cellular defense strategy that confers fitness advantage to bacteria. Proc Natl Acad Sci 109: E1287–E1293. 10.1073/pnas.1119226109

Vasu K, Nagamalleswari E, Zahran M, Imhof P, Xu SY, Zhu Z, Chan SH, Nagaraja V. 2013. Increasing cleavage specificity and activity of restriction endonuclease KpnI. Nucleic Acids Res 41: 9812–9824. 10.1093/nar/gkt734

Veiga H, Pinho MG. 2009. Inactivation of the SauI type I restriction-modification system is not sufficient to generate Staphylococcus aureus strains capable of efficiently accepting foreign DNA. Appl Environ Microbiol 75: 3034–3038. 10.1128/AEM.01862-08

Venclovas C, Timinskas A, Šikšnys V. 1994. Five-stranded β-sheet sandwiched with two α-helices: a structural link between restriction endonucleases EcoRI and EcoRV. Proteins 20: 279–282. 10.1002/prot.340200308

Veron N, Peters AH. 2011. Epigenetics: Tet proteins in the limelight. Nature 473: 293–294. 10.1038/473293a

Viadiu H, Vanamee ES, Jacobson EM, Schildkraut I, Aggarwal AK. 2003. Crystallization of restriction endonuclease SfiI in complex with DNA. Acta Crystallogr, Sect D: Biol Crystallogr 59: 1493–1495. 10.1107/S0907444903011910

Vitkute J, Stankevicius K, Tamulaitienė G, Maneliene Z, Timinskas A, Berg DE, Janulaitis A. 2001. Specificities of eleven different DNA methyltransferases of Helicobacter pylori strain 26695. J Bacteriol 183: 443–450. 10.1128/JB.183.2.443-450.2001

Wah DA, Hirsch JA, Dorner LF, Schildkraut I, Aggarwal AK. 1997. Structure of the multimodular endonuclease FokI bound to DNA. Nature 388: 97–100. 10.1038/40446

Wah DA, Bitinaite J, Schildkraut I, Aggarwal AK. 1998. Structure of FokI has implications for DNA cleavage. Proc Natl Acad Sci 95: 10564–10569. 10.1073/pnas.95.18.10564

Waite-Rees PA, Keating CJ, Moran LS, Slatko BE, Hornstra LJ, Benner JS. 1991. Characterization and expression of the Escherichia coli Mrr restriction system. J Bacteriol 173: 5207–5219. 10.1128/jb.173.16.5207-5219.1991

Waldron DE, Lindsay JA. 2006. Sau1: a novel lineage-specific type I restriction-modification system that blocks horizontal gene transfer into Staphylococcus aureus and between S. aureus isolates of different lineages. J Bacteriol 188: 5578–5585. 10.1128/JB.00418-06

Walkinshaw MD, Taylor P, Sturrock SS, Atanasiu C, Berge T, Henderson RM, Edwardson JM, Dryden DT. 2002. Structure of Ocr from bacteriophage T7, a protein that mimics B-form DNA. Mol Cell 9: 187–194. 10.1016/S1097-2765(02)00435-5

Wang L, Chen S, Vergin KL, Giovannoni SJ, Chan SW, DeMott MS, Taghizadeh K, Cordero OX, Cutler M, Timberlake S, et al. 2011a. DNA phosphorothioation is widespread and quantized in bacterial genomes. Proc Natl Acad Sci 108: 2963–2968. 10.1073/pnas.1017261108

Wang H, Guan S, Quimby A, Cohen-Karni D, Pradhan S, Wilson G, Roberts RJ, Zhu Z, Zheng Y. 2011b. Comparative characterization of the PvuRts1I family of restriction enzymes and their application in mapping genomic 5-hydroxymethylcytosine. Nucleic Acids Res 39: 9294–9305. 10.1093/nar/gkr607

Ward DF, Murray NE. 1979. Convergent transcription in bacteriophage λ: interference with gene expression. J Mol Biol 133: 249–266. 10.1016/0022-2836(79)90533-3

Warren RA. 1980. Modified bases in bacteriophage DNAs. Annu Rev Microbiol 34: 137–158. 10.1146/annurev.mi.34.100180.001033

Watanabe M, Yuzawa H, Handa N, Kobayashi I. 2006. Hyperthermophilic DNA methyltransferase M.PabI from the archaeon Pyrococcus abyssi. Appl Environ Microbiol 72: 5367–5375. 10.1128/AEM.00433-06

Waugh DS, Sauer RT. 1993. Single amino acid substitutions uncouple the DNA binding and strand scission activities of Fok I endonuclease. Proc Natl Acad Sci 90: 9596–9600. 10.1073/pnas.90.20.9596

Waugh DS, Sauer RT. 1994. A novel class of FokI restriction endonuclease mutants that cleave hemi-methylated substrates. J Biol Chem 269: 12298–12303.

Webb B, Sali A. 2016. Comparative protein structure modeling using MODELLER. In Current protocols in protein science Vol. 86 pp. 2.9.1–2.9.37. John Wiley & Sons, Hoboken, NJ.

Webb M, Taylor IA, Firman K, Kneale GG. 1995. Probing the domain structure of the type IC DNA methyltransferase M.EcoR124I by limited proteolysis. J Mol Biol 250: 181–190. 10.1006/jmbi.1995.0369

Weigele P, Raleigh EA. 2016. Biosynthesis and function of modified bases in bacteria and their viruses. Chem Rev. doi:10.1021/acs.chemrev.1026b00114.

Weiser JN, Williams A, Moxon ER. 1990. Phase-variable lipopolysaccharide structures enhance the invasive capacity of Haemophilus influenzae. Infect Immun 58: 3455–3457.

Welsh AJ, Halford SE, Scott DJ. 2004. Analysis of Type II restriction endonucleases that interact with two recognition sites. In Restriction endonucleases (ed. Pingoud A), pp. 297–317. Springer, Berlin.

Wende W, Stahl F, Pingoud A. 1996. The production and characterization of artificial heterodimers of the restriction endonuclease EcoRV. Biol Chem 377: 625–632.

Wentzell LM, Nobbs TJ, Halford SE. 1995. The SfiI restriction endonuclease makes a four-strand DNA break at two copies of its recognition sequence. J Mol Biol 248: 581–595. 10.1006/jmbi.1995.0244

Wilkins BM. 2002. Plasmid promiscuity: meeting the challenge of DNA immigration control. Environ Microbiol 4: 495–500. 10.1046/j.1462-2920.2002.00332.x

Williams K, Savageau MA, Blumenthal RM. 2013. A bistable hysteretic switch in an activator-repressor regulated restriction-modification system. Nucleic Acids Res 41: 6045–6057. 10.1093/nar/gkt324

Wilson GG, Murray NE. 1991. Restriction and modification systems. Annu Rev Genet 25: 585–627. 10.1146/annurev.ge.25.120191.003101

Winkler FK, Banner DW, Oefner C, Tsernoglou D, Brown RS, Heathman SP, Bryan RK, Martin PD, Petratos K, Wilson KS. 1993. The crystal structure of EcoRV endonuclease and of its complexes with cognate and non-cognate DNA fragments. EMBO J 12: 1781–1795. 10.1002/j.1460-2075.1993.tb05826.x

Winter M. 1997. “Investigation of de novo methylation activity in mutants of the EcoKI methyltransferase.” PhD thesis, University of Edinburgh, Edinburgh.

Wion D, Casadesus J. 2006. N6-methyl-adenine: an epigenetic signal for DNA-protein interactions. Nat Rev Microbiol 4: 183–192. 10.1038/nrmicro1350

Wolfes H, Alves J, Fliess A, Geiger R, Pingoud A. 1986. Site directed mutagenesis experiments suggest that Glu 111, Glu 144 and Arg 145 are essential for endonucleolytic activity of EcoRI. Nucleic Acids Res 14: 9063–9080. 10.1093/nar/14.22.9063

Wood KM, Daniels LE, Halford SE. 2005. Long-range communications between DNA sites by the dimeric restriction endonuclease SgrAI. J Mol Biol 350: 240–253. 10.1016/j.jmb.2005.04.053

Wright GD. 2010. Antibiotic resistance in the environment: a link to the clinic? Curr Opin Microbiol 13: 589–594. 10.1016/j.mib.2010.08.005

Wu J, Kandavelou K, Chandrasegaran S. 2007. Custom-designed zinc finger nucleases: what is next? Cell Mol Life Scis 64: 2933–2944. 10.1007/s00018-007-7206-8

Wyatt GR, Cohen SS. 1953. The bases of the nucleic acids of some bacterial and animal viruses: the occurrence of 5-hydroxymethylcytosine. Biochem J 55: 774–782. 10.1042/bj0550774

Wyszomirski KH, Curth U, Alves J, Mackeldanz P, Moncke-Buchner E, Schutkowski M, Kruger DH, Reuter M. 2012. Type III restriction endonuclease EcoP15I is a heterotrimeric complex containing one Res subunit with several DNA-binding regions and ATPase activity. Nucleic Acids Res 40: 3610–3622. 10.1093/nar/gkr1239

Xiao Y, Jung C, Marx AD, Winkler I, Wyman C, Lebbink JH, Friedhoff P, Cristovao M. 2011. Generation of DNA nanocircles containing mismatched bases. BioTechniques 51: p259–262, 264–255.

Xu Q, Morgan RD, Roberts RJ, Blaser MJ. 2000a. Identification of type II restriction and modification systems in Helicobacter pylori reveals their substantial diversity among strains. Proc Natl Acad Sci 97: 9671–9676. 10.1073/pnas.97.17.9671

Xu Q, Stickel S, Roberts RJ, Blaser MJ, Morgan RD. 2000b. Purification of the novel endonuclease, Hpy188I, and cloning of its restriction-modification genes reveal evidence of its horizontal transfer to the Helicobacter pylori genome. J Biol Chem 275: 17086–17093. 10.1074/jbc.M910303199

Xu SY, Zhu Z, Zhang P, Chan SH, Samuelson JC, Xiao J, Ingalls D, Wilson GG. 2007. Discovery of natural nicking endonucleases Nb.BsrDI and Nb.BtsI and engineering of top-strand nicking variants from BsrDI and BtsI. Nucleic Acids Res 35: 4608–4618. 10.1093/nar/gkm481

Xu T, Liang J, Chen S, Wang L, He X, You D, Wang Z, Li A, Xu Z, Zhou X, et al. 2009. DNA phosphorothioation in Streptomyces lividans: mutational analysis of the dnd locus. BMC Microbiol 9: 41 doi:10.1186/1471-2180-1189-1141. 10.1186/1471-2180-9-41

Xu SY, Corvaglia AR, Chan SH, Zheng Y, Linder P. 2011. A type IV modification-dependent restriction enzyme SauUSI from Staphylococcus aureus subsp. aureus USA300. Nucleic Acids Res 39: 5597–5610. 10.1093/nar/gkr098

Xu S-y, Gidwani S, Heiter D. 2015. Rearranging the subdomains in the BsaXI S subunit to generate a new specificity. In 7th NEB meeting on restriction and modification (Gdansk): Poster P34.

Xu SY, Klein P, Degtyarev S, Roberts RJ. 2016. Expression and purification of the modification-dependent restriction enzyme BisI and its homologous enzymes. Sci Rep 6: 28579. doi:10.1038/srep28579. 10.1038/srep28579

Yang Z, Horton JR, Maunus R, Wilson GG, Roberts RJ, Cheng X. 2005. Structure of HinP1I endonuclease reveals a striking similarity to the monomeric restriction enzyme MspI. Nucleic Acids Res 33: 1892–1901. 10.1093/nar/gki337

Yanik M, Alzubi J, Lahaye T, Cathomen T, Pingoud A, Wende W. 2013. TALE-PvuII fusion proteins—novel tools for gene targeting. PloS One 8: e82539. doi:10.1371/journal.pone.0082539. 10.1371/journal.pone.0082539

Yonezawa A, Sugiura Y. 1994. DNA binding mode of class-IIS restriction endonuclease FokI revealed by DNA footprinting analysis. Biochim Biophys Acta 1219: 369–379. 10.1016/0167-4781(94)90061-2

Yuan R, Hamilton DL, Burckhardt J. 1980. DNA translocation by the restriction enzyme from E. coli K. Cell 20: 237–244. 10.1016/0092-8674(80)90251-2

Yunusova AK, Rogulin EA, Artyukh RI, Zheleznaya LA, Matvienko NI. 2006. Nickase and a protein encoded by an open reading frame downstream from the nickase BspD6I gene form a restriction endonuclease complex. Biochem Biokhim 71: 815–820. 10.1134/S0006297906070157

Zabeau M, Friedman S, Van Montagu M, Schell J. 1980. The ral gene of phage λ. I. Identification of a non-essential gene that modulates restriction and modification in E. coli. Mol Gen Genet 179: 63–73. 10.1007/BF00268447

Zagorskaite E, Sasnauskas G. 2014. Chemical display of pyrimidine bases flipped out by modification-dependent restriction endonucleases of MspJI and PvuRts1I families. PloS One 9: e114580. doi:10.1371/journal.pone.0114580. 10.1371/journal.pone.0114580

Zaleski P, Wojciechowski M, Piekarowicz A. 2005. The role of Dam methylation in phase variation of Haemophilus influenzae genes involved in defence against phage infection. Microbiology 151: 3361–3369. 10.1099/mic.0.28184-0

Zaremba M, Urbanke C, Halford SE, Šikšnys V. 2004. Generation of the BfiI restriction endonuclease from the fusion of a DNA recognition domain to a non-specific nuclease from the phospholipase D superfamily. J Mol Biol 336: 81–92. 10.1016/j.jmb.2003.12.012

Zaremba M, Sasnauskas G, Urbanke C, Šikšnys V. 2005. Conversion of the tetrameric restriction endonuclease Bse634I into a dimer: oligomeric structure-stability-function correlations. J Mol Biol 348: 459–478. 10.1016/j.jmb.2005.02.037

Zaremba M, Sasnauskas G, Urbanke C, Šikšnys V. 2006. Allosteric communication network in the tetrameric restriction endonuclease Bse634I. J Mol Biol 363: 800–812. 10.1016/j.jmb.2006.08.050

Zaremba M, Owsicka A, Tamulaitis G, Sasnauskas G, Shlyakhtenko LS, Lushnikov AY, Lyubchenko YL, Laurens N, van den Broek B, Wuite GJ, et al. 2010. DNA synapsis through transient tetramerization triggers cleavage by Ecl18kI restriction enzyme. Nucleic Acids Res 38: 7142–7154. 10.1093/nar/gkq560

Zaremba M, Sasnauskas G, Šikšnys V. 2012. The link between restriction endonuclease fidelity and oligomeric state: a study with Bse634I. FEBS Lett 586: 3324–3329. 10.1016/j.febslet.2012.07.009

Zaremba M, Toliusis P, Grigaitis R, Manakova E, Silanskas A, Tamulaitienė G, Szczelkun MD, Šikšnys V. 2014. DNA cleavage by CgII and NgoAVII requires interaction between N- and R-proteins and extensive nucleotide hydrolysis. Nucleic Acids Res 42: 13887–13896. 10.1093/nar/gku1236

Zaremba M, Toliusis P, Silanskas A, Manakova E, Szczelkun M, Šikšnys V. 2015. DNA cleavage by CgII requires assembly of a heterotetramer of R- and H proteins and extensive ATP hydrolysis. 7th NEB meeting on restriction and modification (Gdansk): Short talk S8.

Zavil'gel'skii GB. 2000. [Antirestriction]. Mol Biol 34: 854–862.

Zavil'gel'skii GB, Rastorguev SM. 2009. [Antirestriction proteins ardA and Ocr as effective inhibitors of the type I restriction-modification enzymes]. Mol Biol 43: 264–273.

Zavil'gel'skii GB, Kotova V, Rastorguev SM. 2009. [Antirestriction and antimodification activities of the T7 Ocr protein: effect of mutations in interface]. Mol Biol 43: 103–110.

Zavil'gel'skii GV, Kotova V, Rastorguev SM. 2011. [Antimodification activity of the ArdA and Ocr proteins]. Genetika 47: 159–167.

Zhang M, Huang J, Deng M, Weng X, Ma H, Zhou X. 2009. Sensitive and visual detection of adenosine by a rationally designed FokI-based biosensing strategy. Chemistry Asian J 4: 1420–1423. 10.1002/asia.200900200

Zheleznaya LA, Kainov DE, Yunusova AK, Matvienko NI. 2003. Regulatory C protein of the EcoRV modification-restriction system. Biochem Biokhim 68: 105–110. 10.1023/A:1022105804578

Zheng Y, Cohen-Karni D, Xu D, Chin HG, Wilson G, Pradhan S, Roberts RJ. 2010. A unique family of Mrr-like modification-dependent restriction endonucleases. Nucleic Acids Res 38: 5527–5534. 10.1093/nar/gkq327

Zhou EX, Reuter M, Meehan EJ, Chen L. 2002. A new crystal form of restriction endonuclease EcoRII that diffracts to 2.8 Å resolution. Acta Crystallogr, Sect D: Biol Crystallogr 58: 1343–1345. 10.1107/S0907444902009125

Zhou XE, Wang Y, Reuter M, Mackeldanz P, Kruger DH, Meehan EJ, Chen L. 2003. A single mutation of restriction endonuclease EcoRII led to a new crystal form that diffracts to 2.1 Å resolution. Acta Crystallogr, Sect D: Biol Crystallogr 59: 910–912. 10.1107/S0907444903004086

Zhou XE, Wang Y, Reuter M, Mucke M, Kruger DH, Meehan EJ, Chen L. 2004. Crystal structure of type IIE restriction endonuclease EcoRII reveals an autoinhibition mechanism by a novel effector-binding fold. J Mol Biol 335: 307–319. 10.1016/j.jmb.2003.10.030

Zhou X, He X, Liang J, Li A, Xu T, Kieser T, Helmann JD, Deng Z. 2005. A novel DNA modification by sulphur. Mol Microbiol 57: 1428–1438. 10.1111/j.1365-2958.2005.04764.x

Zylicz-Stachula A, Bujnicki JM, Skowron PM. 2009. Cloning and analysis of a bifunctional methyltransferase/restriction endonuclease TspGWI, the prototype of a Thermus sp. enzyme family. BMC Mol Biol 10: 52. doi:10.1186/1471-2199-1110-1152. 10.1186/1471-2199-10-52

Zylicz-Stachula A, Jezewska-Frackowiak J, Skowron PM. 2014. Cofactor analogue-induced chemical reactivation of endonuclease activity in a DNA cleavage/methylation deficient TspGWI N(4)(7)(3)A variant in the NPPY motif. Mol Biol Rep 41: 2313–2323. 10.1007/s11033-014-3085-x

WWW RESOURCES

http://library.cshl.edu/Meetings/restriction-enzymes/v-GeoffWilson.php Wilson G. 2013. The cloning efforts at NEB.

http://library.cshl.edu/Meetings/restriction-enzymes/v-Janulaitis.php Janulaitis A. 2013. Science and politics: three phases of commercialization at Fermentas.

http://library.cshl.edu/Meetings/restriction-enzymes/v-Lubys.php Lubys A. 2013. The cloning efforts at Fermentas.

http://library.cshl.edu/Meetings/restriction-enzymes/v-Roberts.php Roberts R. 2013. Many more REases at CSHL, the start of REBASE and more recent work.

http://library.cshl.edu/Meetings/restriction-enzymes/v-Studier.php Studier FW. 2013. Phage T7 is neither modified nor restricted by EcoKI or EcoBI, which led to the discovery of the T7 0.3 gene encoding Ocr, as well as to the collision model for Type I enzymes.

http://rebase.neb.com/cgi-bin/cryyearbar 2017. REBASE crystals. Restriction enzyme structures per year.

http://rebase.neb.com/rebase/rebase.html REBASE. The restriction enzyme database.

http://scop.mrc-lmb.cam.ac.uk SCOP. Structural classification of proteins.

https://www.knaw.nl/en/awards/laureates/dr-h-p-heinekenprijs-voor-biochemie-en-biofysica/alec-j-jeffreys-1950-groot-brittannia Koninklijke Nederlandse Academie van Wetenschappen site (Royal Dutch Academy of “Arts and Sciences”), which is hosting an Alec J. Jeffreys biography.

https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3831583/pdf/2041-2223-4-21.pdf

https://www.ncbi.nlm.nih.gov/pubmed/20011117 Gitschier J. 2009. DNA fingerprints: an interview with professor Sir Alec Jeffreys.

 

APPENDIX 1: Table of selected REases studied in recent years

Type II name and restriction site Sub-typea Detailsb Reference(s)c

AhdI
GACNNN↓NNGTC

H P

Hybrid IIP with tetrameric M&S MTase, missing link Type I and II?

Marks et al. 2003; Pingoud et al. 2014

AspCNI
GCSGC

P

PLD REase cleaves poorly at high concentrations.

Heiter et al. 2015

BbvCI
CCTCAGC (−5/−2)

A T

Two catalytic sites from different subunits, each cleaving own strand; useful as nicking enzyme.

Bellamy et al. 2005; Heiter et al. 2005

BcgI
(10/12) CGA(N6)TGC (12/10)

B G H S

Cleaves four bonds in concerted action.

Kong and Smith 1998; Marshall et al. 2007, 2011; Marshall and Halford 2010; Halford 2013; Smith et al. 2013a,b; Pingoud et al. 2014

BcnI
CC↓SGG

P

Monomer localizes target site by 1D and 3D diffusion, nicks one DNA strand, turns 180°, and cleaves the second strand; the switch in orientation proceeds without dissociation into bulk solution (Sasnauskas et al. 2011); design of nicking endonucleases (Kostiuk et al. 2011).

Janulaitis et al. 1983; Sokolowska et al. 2007a; Kostiuk et al. 2011; Sasnauskas et al. 2011; Kostiuk et al. 2015

BfiI
ACTGGG (5/4)

S

PLD. The first non-PD-(D/E)XK restriction enzyme identified (Sapranauskas et al. 2000). Well-studied; carboxy-terminal DBD resembles that of B3-like plant transcription factors; cleaves one strand at a time via covalent intermediate; catalyzes both DNA hydrolysis and transesterification reactions (Sasnauskas et al. 2007).

Sapranauskas et al. 2000; Lagunavicius et al. 2003; Sasnauskas et al. 2003, 2007, 2008, 2010; Zaremba et al. 2004; Gražulis et al. 2005; Marshall and Halford 2010; Golovenko et al. 2014; Pingoud et al. 2014

BisI
Gm5CNGC

M

Gm5CNGC; relatives widespread requiring different numbers of m5C.

Xu et al. 2016

BpuSI
GGGAC (10/14)

G S

RM·BpuSI has FokI-like carboxy-terminal domain; accompanied by two MTases.

Niv et al. 2007; Shen et al. 2011; Sarrade-Loucheur et al. 2013; Pingoud et al. 2014

Bse634I
R↓CCGGY

F P

Tetramer. Dimeric mutant still has same fidelity as tetramer. Member of CCGG family.

Gražulis et al. 2002; Zaremba et al. 2005, 2006, 2012; Manakova et al. 2012

BspD6I
GACTC (4/6)

S

N·BspD6I is a site-specific nickase in a heterodimeric complex with a catalytic subunit.

Kachalova et al. 2008

BspMI
ACCTGC (4/8)

S

Detection relative orientation of two target sites.

Kingston et al. 2003

BspRI
GG↓CC

P

Monomeric REase cuts dsDNA recognition site in two independent binding events.

Rasko et al. 2010

BstYI
R/GATCY

P

Thermophilic REase with overlapping specificity to BamHI and BglII.

Townson et al. 2004, 2005

BsaXI
(9/12) ACNNNNNCTCC (10/7)

B

Domain swapping and circular permutation of TRD subdomains (or deletion) → active protein with altered specificity/poor protein yields or inactivity.

Xu et al. 2015

Cfr10I
R↓CCGGY

F

Tetramer. Importance of spatial rather than sequence conservation of aa at active center. Member of CCGG family.

Bozic et al. 1996; Skirgaila and Šikšnys 1998; Skirgaila et al. 1998; Šikšnys et al. 1999

CglI
GCSGC

P

Heterotetramer of R + H with extensive ATP hydrolysis. R protein is similar to BfiI.

Zaremba et al. 2014, 2015

DpnI
Gm6A↓TC

M P

Methylation-dependent; cuts as monomer, one strand at a time; amino-terminal PD domain and carboxy-terminal winged helix (wH) allosteric activator domain; both domains bind methylated DNA with sequence specificity.

Lacks and Greenberg 1975; Siwek et al. 2012; Mierzejewska et al. 2014; Pingoud et al. 2014

Ecl18kI
↓CCNGG

P

The first REase which flips out nucleotides from its target site to accommodate interrupted CCGG sequences in the conserved active site. Transient tetramerization triggers cleavage. Member of CCGG family.

Denjmukhametov et al. 1998; Bochtler et al. 2006; Tamulaitis et al. 2008; Fedotova et al. 2009; Protsenko et al. 2009; Zaremba et al. 2010; Burenina et al. 2013; Rutkauskas et al. 2014

Eco29kI
CCGC↓GG

P

GIY-YIG structure.

Mak et al. 2010; Pertzev et al. 1997; Ibryashkina et al. 2007; Orlowski and Bujnicki 2008; Mokrishcheva et al. 2011

Eco57I
CTGAAG (16/14)

E G S

Accompanied by one MTase (cuts 1½ turn away). Sequence specificity was altered by the methylation activity–based selection technique.

Janulaitis et al. 1992a,b; Rimseliene et al. 2003; Pingoud et al. 2014

EcoRI
G↓AATTC

P

The first structure of a restriction enzyme; one of the most extensively studied “classical” restriction enzyme. Studies on inhibition by Cu2+ ions; relaxed specificity and structure of mutants that cleave EcoRI star sites; role of flanking sequences; regulation ecoRIRM operon.

McClarin et al. 1986; Kim et al. 1990; Kurpiewski et al. 2004; Sapienza et al. 2005, 2007; Liu and Kobayashi 2007; Liu et al. 2007; VanderVeen et al. 2008; Ji et al. 2014; Sapienza et al. 2014;

EcoRII
↓CCWGG

E P

For the first time it was shown that REase can be activated to cleave refractory DNA recognition sites (Kruger et al. 1988).The only demonstrated restriction enzyme which interacts with three recognition sites to effectively cleave one DNA site (Tamulaitis et al. 2006b). The first case of autoinhibition, a mechanism described for many transcription factors and signal transducing proteins (Zhou et al. 2004). Member of CCGG family.

Kruger et al. 1988; Zhou et al. 2002, 2003, 2004; Reuter et al. 2004; Kruger and Reuter 2005; Tamulaitis et al. 2006a,b, 2008; Shlyakhtenko et al. 2007; Gilmore et al. 2009; Golovenko et al. 2009; Szczepek et al. 2009

EcoRV
GAT↓ATC

P

One of the most extensively studied “classical” restriction enzymes. Single molecule studies; tracking of single quantum dot labeled EcoRV sliding along DNA manipulated by double optical tweezers indicated that during sliding, EcoRV stays in close contact with the DNA.

Winkler et al. 1993; Kostrewa and Winkler 1995; Bonnet et al. 2008; Biebricher et al. 2009

FokI
GGATG (9/13)

S

Early, best-known IIS; two MTases fused in single protein; crystals indicate catalytic domain hidden behind DNA-binding domain; DNA-cleavage domain used for engineering purposes.

Sugisaki and Kanazawa 1981; Miller et al. 1985; Nwankwo and Wilson 1987; Mandecki and Bolling 1988; Kaczorowski et al. 1989; Kita et al. 1989a,b; Landry et al. 1989; Looney et al. 1989; Sugisaki et al. 1989; Goszczynski and McGhee 1991; Szybalski et al. 1991; Li et al. 1992, 1993; Li and Chandrasegaran 1993; Skowron et al. 1993, 1996; Waugh and Sauer 1993; Kim et al. 1994, 1996a,b, 1997, 1998; Waugh and Sauer 1994; Yonezawa and Sugiura 1994; Hirsch et al. 1997; Wah et al. 1997, 1998; Bitinaite et al. 1998; Leismann et al. 1998; Chandrasegaran and Smith 1999; Friedrich et al. 2000; Vanamee et al. 2001; Bibikova et al. 2002; Urnov et al. 2005; Catto et al. 2006, 2008; Gemmen et al. 2006; Bellamy et al. 2007; Miller et al. 2007; Szczepek et al. 2007; Vanamee et al. 2007; Mino et al. 2009, 2014; Mori et al. 2009; Sanders et al. 2009; Zhang et al. 2009; Guo et al. 2010; Imanishi et al. 2010; Klug 2010a,b; Carroll 2011a,b; Gabriel et al. 2011; Halford et al. 2011; Handel and Cathomen 2011; Li et al. 2011; Pattanayak et al. 2011; Ramalingam et al. 2011; Handel et al. 2012; Laurens et al. 2012; Pernstich and Halford 2012; Rusling et al. 2012; Bhakta et al. 2013; Ramalingam et al. 2013; Guilinger et al. 2014a; Pingoud et al. 2014

HinP1I
G↓CGC

P

Monomeric REase with structural (but no sequence) similarity to MspI; back-to-back dimer with two active sites and two DNA duplexes bound on the outer surfaces of the dimer facing away from each other. The function of the base flipping is unclear and it seems to be part of great DNA distortions.

Yang et al. 2005; Horton et al. 2006

HphI
GGTGA (8/7)

S

HNH REase

Cymerman et al. 2006

Hpy188I
TCN↓GA

P

GIY-YIG structure with highly specific recognition structure, in contrast with HEases or DNA repair enzymes.

Xu et al. 2000b; Kaminska et al. 2008; Orlowski and Bujnicki 2008; Sokolowska et al. 2011

KpnI
GGTAC↓C

P

The first member of a HNH family REase-mediated death benefits population; computer model, but no crystal structure available.

Chandrashekaran et al. 2004; Saravanan et al. 2004; 2007a,b; Gupta et al. 2010; Vasu et al. 2012, 2013; Vasu and Nagaraja 2013

MnlI
CCTC (7/6)

S

Member of Type IIS with the HNH-type active site within carboxy-terminal domain, DNA recognition amino-terminal domain.

Kriukiene et al. 2005; Kriukiene 2006

NaeI
GCC↓GGC

E

Needs to interact with two copies of the recognition sequence for efficient cleavage of one. Real-time observation of DNA looping dynamics of NaeI and NarI compared.

Huai et al. 2000; van den Broek et al. 2006

NarI
GG↓CGCC

E

Real-time observation of DNA looping dynamics of NaeI and NarI compared.

van den Broek et al. 2006

Mva1269I
GAATGC (1/−1)

S

Monomeric REase with two EcoRI-like and FokI-like catalytic domains; design of nicking endonucleases (Armalyte et al. 2005).

Armalyte et al. 2005; Pingoud et al. 2014

MmeI
TCCRAC (20/18)

G S

No separate MTase (cuts two turns away, crystal); changes in the S domain alter recognition site for R and M (like Type I enzymes); altered specificities could be predicted.

Boyd et al. 1986; Tucholski et al. 1995, 1998; Nakonieczna et al. 2007, 2009; Morgan et al. 2008, 2009; Morgan and Luyten 2009; Callahan et al. 2011, 2016; Pingoud et al. 2014

NotI
GC↓GGCCGC

P

Well-known 8-bp cutter; structure reveals unique metal binding fold (also in other putative endonucleases) with an iron atom within Cys4 motif

Lambert et al. 2008

PabI
GTA↓C

P

“half-pipe”; cuts GTA/C; no true REase, as it flips all four purines out of the helix; DNA adenine glycosylase excises adenines.

Miyazono et al. 2007, 2014; Ishikawa et al. 2005; Watanabe et al. 2006; Pingoud et al. 2014; Kojima and Kobayashi 2015

PacI
TTAAT↓TAA

P

HNH REase; 8–base pair rare-cutting homodimer, each subunit with two Zn2+-bound motifs surrounding a beta-beta-alpha-metal catalytic site.

Shen et al. 2010

PspGI
↓CCWGG

P

Use nucleotide flipping as a part of its DNA recognition mechanism. Member of CCGG family.

Pingoud et al. 2003; Szczepanowski et al. 2008; Tamulaitis et al. 2008

SfiI
GGCCNNNN↓NGGCC

F P

The first characterized homotetrameric enzyme: Structures reveal two different binding states of SfiI: one with both DNA-binding sites fully occupied and the other with fully and partially occupied sites.

Wentzell et al. 1995; Viadiu et al. 2003; Embleton et al. 2004; Vanamee et al. 2005; Bellamy et al. 2007, 2008, 2009; Laurens et al. 2009

SgrAI
CR↓CCGGYG

F P

Filaments by cryoEM; Member of CCGG family; preferentially cleaves concertedly at two sites; assembles into homotetramers, then other molecules join to generate helical structures with one DNA-bound homodimer after another.

Laue et al. 1990; Tautz et al. 1990; Capoluongo et al. 2000; Bitinaite and Schildkraut 2002; Daniels et al. 2003; Hingorani-Varma and Bitinaite 2003; Šikšnys et al. 2004; Wood et al. 2005; Dunten et al. 2008, 2009; Park et al. 2010; Little et al. 2011; Lyumkis et al. 2013; Ma et al. 2013b; Horton 2015

SsoII
↓CCNGG

P

Member of CCGG family; tested for applications such as genome surgery.

Kubareva et al. 1992, 2000; Pingoud et al. 2005, 2014; Sudina et al. 2005; Bochtler et al. 2006; Pingoud and Silva 2007; Fedotova et al. 2009; Schierling et al. 2010; Hien le et al. 2011; Abrosimova et al. 2013

SwaI
ATTT↓AAAT

P

8–base pair cutter of AT-rich site.

Dedkov and Degtyarev 1998; Shen et al. 2015

TseI
G↓CWGC

P

Cuts A:A and T:T mismatch in CAG and CTG repeats (useful for typing in Huntington's disease).

Ma et al. 2013a

TspGWI
ACGGA (11/9)

G S

Monomeric bifunctional R-M with independent REase and MTase activities that can be uncoupled; resembles Type I, but lacks translocation domain (“half” Type I).

Skowron et al. 2003; Zylicz-Stachula et al. 2009, 2014

TstI
(8/13) CACNNNNNNTCC (12/7)

B

DNA recognition, cleavage, and methylation in one polypeptide.

Smith et al. 2014

Other Types d

Structure.

McrB-N
Rm5C(N30-35/)-(N30-3000)-Rm5C

IV

Structure recognition domain McrBC; flips out the modified cytosine.

Sukackaite et al. 2012

PvuRts1I
mC(N11-13/N9-10)G

IV

Amino-terminal, atypical PD-(D/E)XK REase domain and carboxy-terminal SRA domain with potential pocket for a flipped hm5C or ghm5C. Epigenome studies (Szwagierczak et al. 2011; Wang et al. 2011). This one and other family members could also be classified as Type IIM (like DpnI). Also a Type IIS. Unlike Type IV (McrBC), PvuRts1I family enzymes cut at a fixed distance from the recognition site.

Janosi et al. 1994; Szwagierczak et al. 2011; Wang et al. 2011b; Kazrani et al. 2014; Shao et al. 2014; Zagorskaite and Sasnauskas 2014

EcoKI
(AACNNNNNNTGC)
&
EcoR124I
(GAANNNNNNNRTCG)

I

Structure based on EM analysis.

Kennaway et al. 2012

I

Structure HsdR.

Csefalvay et al. 2015

EcoP15I
(CAGCAG)

III

Structure.

Gupta et al. 2015

MspJI (and other family members, e.g., LpnPI, AspBHI, …)

IIM , IIS (methylation-dependent)

This is a large family of methylation-dependent (Type IIM and IIS) restriction enzymes. Structures of MspJI (apo- [Horton et al. 2012] and DNA bound, with a flipped base in the SRA domain (Horton et al. 2014c); structures of apo-LpnPI (Sasnauskas et al. 2015b), apo-AspBHI (Horton et al. 2014b) are also available. Base flipping in solution (Zagorskaite and Sasnauskas 2014). Applications in genome methylation studies (Zheng et al. 2010; Cohen-Karni et al. 2011).

Zheng et al. 2010; Cohen-Karni et al. 2011; Horton et al. 2012, 2014b,c; Zagorskaite and Sasnauskas 2014; Sasnauskas et al. 2015b

Update September 2017; courtesy of Vilnius group.

a Subtypes as listed on REBASE, with some adaptations by Pingoud et al. (2014) to indicate the overlap between different subtypes.

b The REases that were the first of their (sub)type (adapted from Horton 2015) are indicated in bold: For example, BcnI was the first Type II enzyme to be identified as a monomer (Sokolowska et al. 2007).

c References include older papers to the respective REases, where suitable: For example, SgrAI was the first REase to be shown to form filaments by cryoEM a few years ago (Lyumkis et al. 2013), but the enzyme was already first reported in 1990.

d See Parts B, C, and D within the chapter (on Type I, III, and IV, respectively) for details and other references.

Get the full book:
PDF / ePub / mobi

Buy the Print Book