Site-specific recombination

Last updated

Site-specific recombination, also known as conservative site-specific recombination, is a type of genetic recombination in which DNA strand exchange takes place between segments possessing at least a certain degree of sequence homology. [1] [2] [3] Enzymes known as site-specific recombinases (SSRs) perform rearrangements of DNA segments by recognizing and binding to short, specific DNA sequences (sites), at which they cleave the DNA backbone, exchange the two DNA helices involved, and rejoin the DNA strands. In some cases the presence of a recombinase enzyme and the recombination sites is sufficient for the reaction to proceed; in other systems a number of accessory proteins and/or accessory sites are required. Many different genome modification strategies, among these recombinase-mediated cassette exchange (RMCE), an advanced approach for the targeted introduction of transcription units into predetermined genomic loci, rely on SSRs.

Contents

Site-specific recombination systems are highly specific, fast, and efficient, even when faced with complex eukaryotic genomes. [4] They are employed naturally in a variety of cellular processes, including bacterial genome replication, differentiation and pathogenesis, and movement of mobile genetic elements. [5] For the same reasons, they present a potential basis for the development of genetic engineering tools. [6]

Recombination sites are typically between 30 and 200 nucleotides in length and consist of two motifs with a partial inverted-repeat symmetry, to which the recombinase binds, and which flank a central crossover sequence at which the recombination takes place. The pairs of sites between which the recombination occurs are usually identical, but there are exceptions (e.g. attP and attB of λ integrase). [7]

Classification: tyrosine- vs. serine- recombinases

Fig. 1. Tyr-Recombinases: Details of the crossover step.

Top: Traditional view including strand-exchange followed by branch-migration (proofreading). The mechanism occurs in the framework of a synaptic complex (1) including both DNA sites in parallel orientation. While branch-migration explains the specific homology requirements and the reversibility of the process in a straightforward manner, it cannot be reconciled with the motions recombinase subunits have to undergo in three dimensions.

Bottom: Current view. Two simultaneous strand-swaps, each depending on the complementarity of three successive bases at (or close to) the edges of the 8-bp spacer (dashed lines indicate base-pairing). Didactic complications arise from the fact that, in this model, the synaptic complex must accommodate both substrates in an anti-parallel orientation.

This synaptic complex (1) arises from the association of two individual recombinase subunits ("protomers"; gray ovals) with the respective target site. Its formation depends on inter-protomer contacts and DNA bending, which in turn define the subunits (green) with an active role during the first crossover reaction. Both representations illustrate only one half of the respective pathway. These parts are separated by a Holliday junction/isomerization step before the product (3) can be released. STswap.png
Fig. 1. Tyr-Recombinases: Details of the crossover step.

Top: Traditional view including strand-exchange followed by branch-migration (proofreading). The mechanism occurs in the framework of a synaptic complex (1) including both DNA sites in parallel orientation. While branch-migration explains the specific homology requirements and the reversibility of the process in a straightforward manner, it cannot be reconciled with the motions recombinase subunits have to undergo in three dimensions.

Bottom: Current view. Two simultaneous strand-swaps, each depending on the complementarity of three successive bases at (or close to) the edges of the 8-bp spacer (dashed lines indicate base-pairing). Didactic complications arise from the fact that, in this model, the synaptic complex must accommodate both substrates in an anti-parallel orientation.

This synaptic complex (1) arises from the association of two individual recombinase subunits ("protomers"; gray ovals) with the respective target site. Its formation depends on inter-protomer contacts and DNA bending, which in turn define the subunits (green) with an active role during the first crossover reaction. Both representations illustrate only one half of the respective pathway. These parts are separated by a Holliday junction/isomerization step before the product (3) can be released.
Fig. 2. Ser-Recombinases: The (essentially irreversible) subunit-rotation pathway.

Contrary to Tyr-recombinases, the four participating DNA strands are cut in synchrony at points staggered by only 2 bp (leaving little room for proofreading). Subunit-rotation (180deg) permits the exchange of strands while covalently linked to the protein partner. The intermediate exposure of double-strand breaks bears risks of triggering illegitimate recombination and thereby secondary reactions.

Here, the synaptic complex arises from the association of pre-formed recombinase dimers with the respective target sites (CTD/NTD, C-/N-terminal domain). Like for Tyr-recombinases, each site contains two arms, each accommodating one protomer. As both arms are structured slightly differently, the subunits become conformationally tuned and thereby prepared for their respective role in the recombination cycle. Contrary to members of the Tyr-class the recombination pathway converts two different substrate sites (attP and attB) to site-hybrids (attL and attR). This explains the irreversible nature of this particular recombination pathway, which can only be overcome by auxiliary "recombination directionality factors" (RDFs). SUrot.png
Fig. 2. Ser-Recombinases: The (essentially irreversible) subunit-rotation pathway.

Contrary to Tyr-recombinases, the four participating DNA strands are cut in synchrony at points staggered by only 2 bp (leaving little room for proofreading). Subunit-rotation (180°) permits the exchange of strands while covalently linked to the protein partner. The intermediate exposure of double-strand breaks bears risks of triggering illegitimate recombination and thereby secondary reactions.

Here, the synaptic complex arises from the association of pre-formed recombinase dimers with the respective target sites (CTD/NTD, C-/N-terminal domain). Like for Tyr-recombinases, each site contains two arms, each accommodating one protomer. As both arms are structured slightly differently, the subunits become conformationally tuned and thereby prepared for their respective role in the recombination cycle. Contrary to members of the Tyr-class the recombination pathway converts two different substrate sites (attP and attB) to site-hybrids (attL and attR). This explains the irreversible nature of this particular recombination pathway, which can only be overcome by auxiliary "recombination directionality factors" (RDFs).

Based on amino acid sequence homologies and mechanistic relatedness, most site-specific recombinases are grouped into one of two families: the tyrosine (Tyr) recombinase family or serine (Ser) recombinase family. The names stem from the conserved nucleophilic amino acid residue present in each class of recombinase which is used to attack the DNA and which becomes covalently linked to it during strand exchange. The earliest identified members of the serine recombinase family were known as resolvases or DNA invertases, while the founding member of the tyrosine recombinases, lambda phage integrase (using attP/B recognition sites), differs from the now well-known enzymes such as Cre (from the P1 phage) and FLP (from the yeast Saccharomyces cerevisiae ). Famous serine recombinases include enzymes such as gamma-delta resolvase (from the Tn1000 transposon), Tn3 resolvase (from the Tn3 transposon), and φC31 integrase (from the φC31 phage). [8]

Although the individual members of the two recombinase families can perform reactions with the same practical outcomes, the families are unrelated to each other, having different protein structures and reaction mechanisms. Unlike tyrosine recombinases, serine recombinases are highly modular, as was first hinted by biochemical studies [9] and later shown by crystallographic structures. [10] [11] Knowledge of these protein structures could prove useful when attempting to re-engineer recombinase proteins as tools for genetic manipulation.

Mechanism

Recombination between two DNA sites begins by the recognition and binding of these sites – one site on each of two separate double-stranded DNA molecules, or at least two distant segments of the same molecule – by the recombinase enzyme. This is followed by synapsis, i.e. bringing the sites together to form the synaptic complex. It is within this synaptic complex that the strand exchange takes place, as the DNA is cleaved and rejoined by controlled transesterification reactions. During strand exchange, each double-stranded DNA molecule is cut at a fixed point within the crossover region of the recognition site, releasing a deoxyribose hydroxyl group, while the recombinase enzyme forms a transient covalent bond to a DNA backbone phosphate. This phosphodiester bond between the hydroxyl group of the nucleophilic serine or tyrosine residue conserves the energy that was expended in cleaving the DNA. Energy stored in this bond is subsequently used for the rejoining of the DNA to the corresponding deoxyribose hydroxyl group on the other DNA molecule. The entire reaction therefore proceeds without the need for external energy-rich cofactors such as ATP.

Although the basic chemical reaction is the same for both tyrosine and serine recombinases, there are some differences between them. [12] Tyrosine recombinases, such as Cre or FLP, cleave one DNA strand at a time at points that are staggered by 6–8bp, linking the 3' end of the strand to the hydroxyl group of the tyrosine nucleophile (Fig. 1). [13] Strand exchange then proceeds via a crossed strand intermediate analogous to the Holliday junction in which only one pair of strands has been exchanged. [14] [15]

The mechanism and control of serine recombinases is much less well understood. This group of enzymes was only discovered in the mid-1990s and is still relatively small. The now classical members gamma-delta and Tn3 resolvase, but also new additions like φC31-, Bxb1-, and R4 integrases, cut all four DNA strands simultaneously at points that are staggered by 2 bp (Fig. 2). [16] During cleavage, a protein–DNA bond is formed via a transesterification reaction, in which a phosphodiester bond is replaced by a phosphoserine bond between a 5' phosphate at the cleavage site and the hydroxyl group of the conserved serine residue (S10 in resolvase). [17] [18]

Fig. 3A. Reversible insertion and excision by the Cre-lox system. Cre-lox insertion excision.svg
Fig. 3A. Reversible insertion and excision by the Cre-lox system.
Fig. 3B. Inversion by the Cre-lox system. Cre-lox inversion.svg
Fig. 3B. Inversion by the Cre-lox system.

It is still not entirely clear how the strand exchange occurs after the DNA has been cleaved. However, it has been shown that the strands are exchanged while covalently linked to the protein, with a resulting net rotation of 180°. [19] [20] The most quoted (but not the only) model accounting for these facts is the "subunit rotation model" (Fig. 2). [12] [21] Independent of the model, DNA duplexes are situated outside of the protein complex, and large movement of the protein is needed to achieve the strand exchange. In this case the recombination sites are slightly asymmetric, which allows the enzyme to tell apart the left and right ends of the site. When generating products, left ends are always joined to the right ends of their partner sites, and vice versa. This causes different recombination hybrid sites to be reconstituted in the recombination products. Joining of left ends to left or right to right is avoided due to the asymmetric "overlap" sequence between the staggered points of top and bottom strand exchange, which is in stark contrast to the mechanism employed by tyrosine recombinases. [12]

The reaction catalysed by Cre-recombinase, for instance, may lead to excision of the DNA segment flanked by the two sites (Fig. 3A), but may also lead to integration or inversion of the orientation of the flanked DNA segment (Fig. 3B). What the outcome of the reaction will be is dictated mainly by the relative locations and orientations of the sites that are to be recombined, but also by the innate specificity of the site-specific system in question. Excisions and inversions occur if the recombination takes place between two sites that are found on the same molecule (intramolecular recombination), and if the sites are in the same (direct repeat) or in an opposite orientation (inverted repeat), respectively. Insertions, on the other hand, take place if the recombination occurs on sites that are situated on two different DNA molecules (intermolecular recombination), provided that at least one of these molecules is circular. Most site-specific systems are highly specialised, catalysing only one of these different types of reaction, and have evolved to ignore the sites that are in the "wrong" orientation.

See also

Related Research Articles

<span class="mw-page-title-main">Chymotrypsin</span> Digestive enzyme

Chymotrypsin (EC 3.4.21.1, chymotrypsins A and B, alpha-chymar ophth, avazyme, chymar, chymotest, enzeon, quimar, quimotrase, alpha-chymar, alpha-chymotrypsin A, alpha-chymotrypsin) is a digestive enzyme component of pancreatic juice acting in the duodenum, where it performs proteolysis, the breakdown of proteins and polypeptides. Chymotrypsin preferentially cleaves peptide amide bonds where the side chain of the amino acid N-terminal to the scissile amide bond (the P1 position) is a large hydrophobic amino acid (tyrosine, tryptophan, and phenylalanine). These amino acids contain an aromatic ring in their side chain that fits into a hydrophobic pocket (the S1 position) of the enzyme. It is activated in the presence of trypsin. The hydrophobic and shape complementarity between the peptide substrate P1 side chain and the enzyme S1 binding cavity accounts for the substrate specificity of this enzyme. Chymotrypsin also hydrolyzes other amide bonds in peptides at slower rates, particularly those containing leucine at the P1 position.

<span class="mw-page-title-main">Protein primary structure</span> Linear sequence of amino acids in a peptide or protein

Protein primary structure is the linear sequence of amino acids in a peptide or protein. By convention, the primary structure of a protein is reported starting from the amino-terminal (N) end to the carboxyl-terminal (C) end. Protein biosynthesis is most commonly performed by ribosomes in cells. Peptides can also be synthesized in the laboratory. Protein primary structures can be directly sequenced, or inferred from DNA sequences.

<span class="mw-page-title-main">Integrase</span> Class of enzymes

Retroviral integrase (IN) is an enzyme produced by a retrovirus that integrates its genetic information into that of the host cell it infects. Retroviral INs are not to be confused with phage integrases (recombinases) used in biotechnology, such as λ phage integrase, as discussed in site-specific recombination.

<span class="mw-page-title-main">RecBCD</span> Family of protein complexes in bacteria

Exodeoxyribonuclease V is an enzyme of E. coli that initiates recombinational repair from potentially lethal double strand breaks in DNA which may result from ionizing radiation, replication errors, endonucleases, oxidative damage, and a host of other factors. The RecBCD enzyme is both a helicase that unwinds, or separates the strands of DNA, and a nuclease that makes single-stranded nicks in DNA. It catalyses exonucleolytic cleavage in either 5′- to 3′- or 3′- to 5′-direction to yield 5′-phosphooligonucleotides.

<span class="mw-page-title-main">Serine protease</span> Class of enzymes

Serine proteases are enzymes that cleave peptide bonds in proteins. Serine serves as the nucleophilic amino acid at the (enzyme's) active site. They are found ubiquitously in both eukaryotes and prokaryotes. Serine proteases fall into two broad categories based on their structure: chymotrypsin-like (trypsin-like) or subtilisin-like.

A transposase is any of a class of enzymes capable of binding to the end of a transposon and catalysing its movement to another part of a genome, typically by a cut-and-paste mechanism or a replicative mechanism, in a process known as transposition. The word "transposase" was first coined by the individuals who cloned the enzyme required for transposition of the Tn3 transposon. The existence of transposons was postulated in the late 1940s by Barbara McClintock, who was studying the inheritance of maize, but the actual molecular basis for transposition was described by later groups. McClintock discovered that some segments of chromosomes changed their position, jumping between different loci or from one chromosome to another. The repositioning of these transposons allowed other genes for pigment to be expressed. Transposition in maize causes changes in color; however, in other organisms, such as bacteria, it can cause antibiotic resistance. Transposition is also important in creating genetic diversity within species and generating adaptability to changing living conditions.

In biochemistry, dephosphorylation is the removal of a phosphate (PO43−) group from an organic compound by hydrolysis. It is a reversible post-translational modification. Dephosphorylation and its counterpart, phosphorylation, activate and deactivate enzymes by detaching or attaching phosphoric esters and anhydrides. A notable occurrence of dephosphorylation is the conversion of ATP to ADP and inorganic phosphate.

Cre-Lox recombination is a site-specific recombinase technology, used to carry out deletions, insertions, translocations and inversions at specific sites in the DNA of cells. It allows the DNA modification to be targeted to a specific cell type or be triggered by a specific external stimulus. It is implemented both in eukaryotic and prokaryotic systems. The Cre-lox recombination system has been particularly useful to help neuroscientists to study the brain in which complex cell types and neural circuits come together to generate cognition and behaviors. NIH Blueprint for Neuroscience Research has created several hundreds of Cre driver mouse lines which are currently used by the worldwide neuroscience community.

Site-specific recombinase technologies are genome engineering tools that depend on recombinase enzymes to replace targeted sections of DNA.

<span class="mw-page-title-main">Cre recombinase</span> Genetic recombination enzyme

Cre recombinase is a tyrosine recombinase enzyme derived from the P1 bacteriophage. The enzyme uses a topoisomerase I-like mechanism to carry out site specific recombination events. The enzyme is a member of the integrase family of site specific recombinase and it is known to catalyse the site specific recombination event between two DNA recognition sites. This 34 base pair (bp) loxP recognition site consists of two 13 bp palindromic sequences which flank an 8bp spacer region. The products of Cre-mediated recombination at loxP sites are dependent upon the location and relative orientation of the loxP sites. Two separate DNA species both containing loxP sites can undergo fusion as the result of Cre mediated recombination. DNA sequences found between two loxP sites are said to be "floxed". In this case the products of Cre mediated recombination depends upon the orientation of the loxP sites. DNA found between two loxP sites oriented in the same direction will be excised as a circular loop of DNA whilst intervening DNA between two loxP sites that are opposingly orientated will be inverted. The enzyme requires no additional cofactors or accessory proteins for its function.

<span class="mw-page-title-main">Holliday junction</span> Branched nucleic acid structure

A Holliday junction is a branched nucleic acid structure that contains four double-stranded arms joined. These arms may adopt one of several conformations depending on buffer salt concentrations and the sequence of nucleobases closest to the junction. The structure is named after Robin Holliday, the molecular biologist who proposed its existence in 1964.

Recombinases are genetic recombination enzymes.

DNA footprinting is a method of investigating the sequence specificity of DNA-binding proteins in vitro. This technique can be used to study protein-DNA interactions both outside and within cells.

<span class="mw-page-title-main">FLP-FRT recombination</span>

In genetics, Flp-FRT recombination is a site-directed recombination technology, increasingly used to manipulate an organism's DNA under controlled conditions in vivo. It is analogous to Cre-lox recombination but involves the recombination of sequences between short flippase recognition target (FRT) sites by the recombinase flippase (Flp) derived from the 2 µ plasmid of baker's yeast Saccharomyces cerevisiae.

Protein metabolism denotes the various biochemical processes responsible for the synthesis of proteins and amino acids (anabolism), and the breakdown of proteins by catabolism.

Hin recombinase is a 21kD protein composed of 198 amino acids that is found in the bacteria Salmonella. Hin belongs to the serine recombinase family (B2) of DNA invertases in which it relies on the active site serine to initiate DNA cleavage and recombination. The related protein, gamma-delta resolvase shares high similarity to Hin, of which much structural work has been done, including structures bound to DNA and reaction intermediates. Hin functions to invert a 900 base pair (bp) DNA segment within the salmonella genome that contains a promoter for downstream flagellar genes, fljA and fljB. Inversion of the intervening DNA alternates the direction of the promoter and thereby alternates expression of the flagellar genes. This is advantageous to the bacterium as a means of escape from the host immune response.

P1 is a temperate bacteriophage that infects Escherichia coli and some other bacteria. When undergoing a lysogenic cycle the phage genome exists as a plasmid in the bacterium unlike other phages that integrate into the host DNA. P1 has an icosahedral head containing the DNA attached to a contractile tail with six tail fibers. The P1 phage has gained research interest because it can be used to transfer DNA from one bacterial cell to another in a process known as transduction. As it replicates during its lytic cycle it captures fragments of the host chromosome. If the resulting viral particles are used to infect a different host the captured DNA fragments can be integrated into the new host's genome. This method of in vivo genetic engineering was widely used for many years and is still used today, though to a lesser extent. P1 can also be used to create the P1-derived artificial chromosome cloning vector which can carry relatively large fragments of DNA. P1 encodes a site-specific recombinase, Cre, that is widely used to carry out cell-specific or time-specific DNA recombination by flanking the target DNA with loxP sites.

The Tn3 transposon is a 4957 base pair mobile genetic element, found in prokaryotes. It encodes three proteins:

<span class="mw-page-title-main">ADP-ribosylation</span> Addition of one or more ADP-ribose moieties to a protein.

ADP-ribosylation is the addition of one or more ADP-ribose moieties to a protein. It is a reversible post-translational modification that is involved in many cellular processes, including cell signaling, DNA repair, gene regulation and apoptosis. Improper ADP-ribosylation has been implicated in some forms of cancer. It is also the basis for the toxicity of bacterial compounds such as cholera toxin, diphtheria toxin, and others.

<span class="mw-page-title-main">HIV integration</span>

AIDS is caused by the human immunodeficiency virus (HIV). Individuals with HIV have what is referred to as a "HIV infection". When infected semen, vaginal secretions, or blood come in contact with the mucous membranes or broken skin of an uninfected person, HIV may be transferred to the uninfected person, causing another infection. Additionally, HIV can also be passed from infected pregnant women to their uninfected baby during pregnancy and/or delivery, or via breastfeeding. As a result of HIV infection, a portion of these individuals will progress and go on to develop clinically significant AIDS.

References

  1. Bode, J; Schlake, T; asadasasada Iber, M; Schuebeler, D; Seibler, J; Snezhkov, E; Nikolaev, L (2000). "The transgeneticist's toolbox: novel methods for the targeted modification of eukaryotic genomes". Biol. Chem. 381 (9–10): 801–813. doi:10.1515/BC.2000.103. PMID   11076013. S2CID   36479502.
  2. Kolb, A.F. (2002). "Genome Engineering Using Site-Specific Recombinases". Cloning & Stem Cells. 4 (1): 65–80. doi:10.1089/153623002753632066. PMID   12006158.
  3. Coates, C.J.; Kaminski, JM; Summers, JB; Segal, DJ; Miller, AD; Kolb, AF (2005). "Site-directed genome modification: derivatives of BAL-modifying enzymes as targeting tools" (PDF). Trends in Biotechnology. 23 (8): 407–19. doi:10.1016/j.tibtech.2005.06.009. PMID   15993503. Archived from the original (PDF) on 2006-08-29.
  4. Sauer, B. (1998). "Inducible Gene Targeting in Mice Using the Cre/loxSystem" (PDF). Methods. 14 (4): 381–92. doi:10.1006/meth.1998.0593. PMID   9608509. Archived from the original (PDF) on 2011-06-11.
  5. Nash, H. A. (1996). Site-specific recombination: integration, excision, resolution, and inversion of defined DNA segments. Escherichia coli and Salmonella: cellular and molecular biology, 2, pp. 2363–2376.
  6. Akopian, A.; Stark, W.M. (2005). Site-Specific DNA Recombinases as Instruments for Genomic Surgery. Advances in Genetics. Vol. 55. pp. 1–23. doi:10.1016/S0065-2660(05)55001-6. ISBN   978-0-12-017655-7. PMID   16291210.
  7. Landy, A. (1989). "Dynamic, Structural, and Regulatory Aspects of lambda Site-Specific Recombination". Annual Review of Biochemistry. 58 (1): 913–41. doi:10.1146/annurev.bi.58.070189.004405. PMID   2528323.
  8. Stark, W.M.; Boocock, M.R. (1995). "Topological selectivity in site-specific recombination". Mobile Genetic Elements. Oxford University Press. pp. 101–129.
  9. Abdel-Meguid, S.S.; Grindley, N.D.; Templeton, N.S.; Steitz, T.A. (April 1984). "Cleavage of the site-specific recombination protein gamma delta resolvase: the smaller of two fragments binds DNA specifically". Proc. Natl. Acad. Sci. U.S.A. 81 (7): 2001–5. Bibcode:1984PNAS...81.2001A. doi: 10.1073/pnas.81.7.2001 . PMC   345424 . PMID   6326096.
  10. Yang, W.; Steitz, T.A. (1995). "Crystal structure of the site-specific recombinase gamma delta resolvase complexed with a 34 bp cleavage site". Cell. 82 (2): 193–207. doi: 10.1016/0092-8674(95)90307-0 . PMID   7628011. S2CID   15849525.
  11. Li, W.; Kamtekar, S; Xiong, Y; Sarkis, GJ; Grindley, ND; Steitz, TA (2005). "Structure of a Synaptic Gamma Delta Resolvase Tetramer Covalently Linked to Two Cleaved DNAs". Science. 309 (5738): 1210–5. Bibcode:2005Sci...309.1210L. doi:10.1126/science.1112064. PMID   15994378. S2CID   84409916.
  12. 1 2 3 Turan, S.; Bode, J. (2011). "Review: Site-specific recombinases: from tag-and-target- to tag-and-exchange-based genomic modifications". FASEB J. 25 (12): 4088–4107. doi: 10.1096/fj.11-186940 . PMID   21891781. S2CID   7075677.
  13. Van Duyne, G.D. (2002). "A structural view of tyrosine recombinase site-specific recombination". Mobile DNA II. ASM Press. pp. 93–117.
  14. Holliday, R. (1964). "A mechanism for gene conversion in fungi" (PDF). Genetics Research. 5 (2): 282–304. doi: 10.1017/S0016672300001233 .
  15. Grainge, I.; Jayaram, M. (1999). "The integrase family of recombinases: organization and function of the active site". Molecular Microbiology. 33 (3): 449–56. doi: 10.1046/j.1365-2958.1999.01493.x . PMID   10577069.
  16. Stark, W.M.; Boocock, M.R.; Sherratt, DJ (1992). "Catalysis by site-specific recombinases". Trends in Genetics. 8 (12): 432–9. doi:10.1016/0168-9525(92)90327-Z. PMID   1337225.
  17. Reed, R.R.; Grindley, N.D. (1981). "Transposon-mediated site-specific recombination in vitro: DNA cleavage and protein-DNA linkage at the recombination site". Cell. 25 (3): 721–8. doi:10.1016/0092-8674(81)90179-3. PMID   6269756. S2CID   28410571.
  18. Reed, R.R.; Moser, C.D. (1984). "Resolvase-mediated recombination intermediates contain a serine residue covalently linked to DNA". Cold Spring Harb Symp Quant Biol. 49: 245–9. doi:10.1101/sqb.1984.049.01.028. PMID   6099239.
  19. Stark, M.W.; Sherratt, DJ; Boocock, MR (1989). "Site-specific recombination by Tn 3 resolvase: topological changes in the forward and reverse reactions". Cell. 58 (4): 779–90. doi:10.1016/0092-8674(89)90111-6. PMID   2548736. S2CID   46508016.
  20. Stark, W.M.; Boocock, M.R. (1994). "The Linkage Change of a Knotting Reaction Catalysed by Tn3 Resolvase". Journal of Molecular Biology. 239 (1): 25–36. doi:10.1006/jmbi.1994.1348. PMID   8196046.
  21. Sarkis, G.J; Murley, LL; Leschziner, AE; Boocock, MR; Stark, WM; Grindley, ND (2001). "A model for the gamma-delta resolvase synaptic complex". Molecular Cell. 8 (3): 623–31. doi: 10.1016/S1097-2765(01)00334-3 . PMID   11583624.