key: cord-0012199-y49slc0u authors: Akhverdyan, Valerii Z.; Gak, Evgueni R.; Tokmakova, Irina L.; Stoynova, Nataliya V.; Yomantas, Yurgis A. V.; Mashko, Sergey V. title: Application of the bacteriophage Mu-driven system for the integration/amplification of target genes in the chromosomes of engineered Gram-negative bacteria—mini review date: 2011-06-23 journal: Appl Microbiol Biotechnol DOI: 10.1007/s00253-011-3416-y sha: 05b9d83503f1e8ee4ffc687860e458a1f86df8dd doc_id: 12199 cord_uid: y49slc0u The advantages of phage Mu transposition-based systems for the chromosomal editing of plasmid-less strains are reviewed. The cis and trans requirements for Mu phage-mediated transposition, which include the L/R ends of the Mu DNA, the transposition factors MuA and MuB, and the cis/trans functioning of the E element as an enhancer, are presented. Mini-Mu(LR)/(LER) units are Mu derivatives that lack most of the Mu genes but contain the L/R ends or a properly arranged E element in cis to the L/R ends. The dual-component system, which consists of an integrative plasmid with a mini-Mu and an easily eliminated helper plasmid encoding inducible transposition factors, is described in detail as a tool for the integration/amplification of recombinant DNAs. This chromosomal editing method is based on replicative transposition through the formation of a cointegrate that can be resolved in a recombination-dependent manner. (E-plus)- or (E-minus)-helpers that differ in the presence of the trans-acting E element are used to achieve the proper mini-Mu transposition intensity. The systems that have been developed for the construction of stably maintained mini-Mu multi-integrant strains of Escherichia coli and Methylophilus methylotrophus are described. A novel integration/amplification/fixation strategy is proposed for consecutive independent replicative transpositions of different mini-Mu(LER) units with “excisable” E elements in methylotrophic cells. The relevance and practical significance of constructing plasmid-less recombinant bacterial strains for use in applied microbiology and biotechnology are increasing due to the potential for genetic instability to reduce the number of active recombinant alleles in plasmids (Friehs 2004) and restrictions on the application of plasmids in large-scale industries of the First World (European Council Directives 1990 , 1998 . In vivo chromosomal editing methods, which are primarily based on homologous and/or site-specific recombination of DNA as well as on transposition mechanisms, have been used to engineer plasmid-less bacteria (Balbás and Gosset 2001) . The increase in genomic copy number of native or previously modified target genes is an important tool for chromosomal editing and the construction of stably maintained bacterial genomes. The development of new approaches for the integration (Minaeva et al. 2008; Rivero-Müller et al. 2007 ) and the increase in genomic copy number (Tyo et al. 2009 ) of recombinant DNA fragments remains relevant even for Escherichia coli and closely related Gram-negative bacteria for which recombineeringbased technologies (Court et al. 2002; Sawitzke et al. 2007; Sharan et al. 2009 ) have already been developed to finely modulate the expression of chromosomal genes (De Mey et al. 2010; Katashkina et al. 2005; Meynial-Salles et al. 2005) and sometimes conditionally increase (Doroshenko et al. 2010b) or silence (Krylov et al. 2010 ) the level of their transcription. The methods of integration/amplification are especially important for bacteria for which the genetic tools for the chromosomal editing are not as comprehensive and diversified as those available for E. coli. The integration of target genes in a bacterial chromosome followed by the increase in genomic copy number can be efficiently achieved using a phage Mu-driven transposition system originally developed for E. coli and initially characterized more than 20 years ago (Castilho et al. 1984; Chaconas et al. 1981a, b) . Since the mid-1980s, several Mu derivatives, mini-Mu(s), have been constructed and extensively used for applications in classical in vivo transposition, including insertional mutagenesis, gene fusion, and mapping techniques, as well as for gene cloning and DNA sequencing strategies (Groisman 1991; Groisman and Casadaban 1986, 1987) . The highly efficient Mu-based technology of the in vitro DNA transposition has been developed as a convenient tool for the functional analysis of genes, genomes, and proteins because of the accurate nature reaction and the low stringency for the target preference (Haapa et al. 1999; Haapa-Paananen et al. 2002; Savilahti et al. 1995; Savilahti and Mizuuchi 1996; Turakainen et al. 2009 ). The construction of recombinant plasmid-less L-threonine-overproducing E. coli strains may have been the first application of mini-Mu as a vector for the integration/amplification of target pathway genes for metabolic engineering (Kurahashi et al. 1990 ; Kurahashi and Takinami 1991) . Comprehensive investigations, primarily in vitro, of the Mu-driven system have provided extensive insights into the molecular aspects of the transposition mechanism. These works have been recently summarized in excellent research papers and overviews (Abdelhakim et al. 2008; Au et al. 2006; Chaconas and Harshey 2002; Choi and Harshey 2010; Gueguen et al. 2005; Harshey and Jayaram 2006; Nakai et al. 2001; Rice and Baker 2001) . However, in vivo Mu-driven systems may be underutilized in applied microbiology and biotechnology. The main aim of this review is to highlight, especially for the new generation of investigators, the potential application of Mu-driven systems as a powerful genetic tool for the integration/ amplification of target genes and the construction of plasmid-less engineered bacterial strains. Known as the most efficient transposon, Mu is a temperate bacteriophage that is capable of growth on many enteric bacteria, including E. coli K-12. The Mu phage undergoes two alternative transposition pathways ( Fig. 1 ) at different stages of its life cycle (Howe 1997; Paolozzi and Ghelardini 2006; Symonds et al. 1987 ). The Mu-driven "integrative" (or "conservative" or "nonreplicative") transposition, which is also designated as a "simple insertion" (Akroyd and Symonds 1983; Harshey 1984; Liebart et al. 1982) , and the "replicative" transposition (Chaconas et al. 1981b are temporally separated. During the Mu phage infection step, various flanking sequences from the previous host are lost during the integrative transposition of the linear Mu DNA into a random site of the bacterial chromosome (Au et al. 2006; Bukhari and Zipser 1972) . Replicative transposition through the formation of a "cointegrate" structure is obligatory for replication and for the production of approximately 100 phages during lytic growth (Chaconas et al. 1981b; Chaconas and Harshey 2002; Craigie and Mizuuchi 1985) . As a result of replicative transposition, new copies of the Mu DNA occur at many sites in the bacterial genome, frequently within several kilobase pairs of one another. The most obvious factor that affects the random integration of Mu DNA into bacterial genes is their efficient transcription, which has a clearly negative impact on transposition (Manna et al. 2004) . Both Mu-mediated transposition pathways are catalyzed by a high-order protein-DNA complex called a transpososome (reviewed in Gueguen et al. 2005; Harshey and Jayaram 2006) . The Mu DNA sites and proteins involved in transpososome assembly and function are presented in Figs. 2 and 3, respectively (for detailed references, see Chaconas and Harshey 2002; Harshey and Jayaram 2006) . The core of the Mu transpososome is composed of two Mu end DNA segments (L and R ends) that are synapsed by stably bound multiple subunits of the transposase MuA, which catalyzes the specific DNA cleavage and joining required for transposition (Craigie et al. 1984; Kuo et al. 1991; Lavoie et al. 1991) . The L/R ends each contain three MuA-binding sites (L1-L3 and R1-R3, correspondingly) with different spacings (Fig. 2a ) (Craigie et al. 1984) . MuA binds as a monomer to each L/R site, introducing an 80°to 90°bend (Kuo et al. 1991) . The function of the cis-or trans-encoded MuA protein is essential for the transposition in vivo of the mini-Mu that is presented on supercoiled DNA and consists of the L/R sites (Patterson et al. 1986 ). The six binding sites are not equally important for transposome assembly. Indeed, genetic experiments show the unimportance of the R3 site; a deletion of the L3 site results only in a 10-fold reduction in transposition in vivo (Groenen et al. 1985) . It is shown in vitro that a MuAmediated stable synaptic complex with only three (L1, R1, R2) sites could be formed in which the donor DNA strands are nicked and fully competent in the subsequent strandtransfer step of transposition (Kuo et al. 1991) . The linear mini-Mu DNA flanked by the (R1R2) sites in an inverted orientation is efficiently used for in vitro MuA-mediated assemble the transpososome followed by integrative transposition in vitro (Haapa et al. 1999; Savilahti et al. 1995) or in vivo in genomes of different organisms (Lamberg et al. 2002; Paatero et al. 2008; Pajunen et al. 2005) . In E. coli, Mu transpososome assembly is facilitated by two host-encoded DNA bending proteins-HU and a sequence-specific integration host factor (IHF) (Swinger and Rice 2004) . HU binds in vivo to the spacer between L1 and L2 , introducing a bend and presumably drawing the bound MuA protomers together (Gueguen et al. 2005) . The binding site for IHF is in an E element that is located approximately 1 kb from the L end in the native Mu DNA. The enhancer element, E [earlier named as the internal activation sequence or IAS Mizuuchi and Mizuuchi 1989) ], stimulates transposition more than 100-fold in vitro and in vivo (Castilho et al. 1984; Leung et al. 1989; . The E element partially overlaps with the O1-O3 region (Krause and Higgins 1986), which was initially identified as the operator to which the Mu c-repressor binds to silence the transcription of early phage functions from the P e promoter (Fig. 2a) . The E element is composed of two clusters of MuA-binding sequences separated by a binding site for IHF (Mizuuchi and Mizuuchi 1989; ). The site-specific IHF-mediated bending of DNA at the E element is presumed to assist MuA-mediated end-enhancer interactions (Harshey and Jayaram 2006) . IHF is required in vitro for the efficient transposition of the mini-Mu carrying the E element when the superhelical density (σ) of the donor DNA is low , i.e., σ decreases from −0.05 (a character value for naked bacterial superhelical DNA) to a level of σ~−0.025 [this density is typical for the "restrained" protein-bound DNA molecule in vivo (Dillon and Dorman 2010; Pettijohn 1996; Pettijohn and Pfenninger 1980) ]. (Daniell et al. 1973; Bukhari and Taylor 1975) ] to a bacterial cell. The ends of the infecting DNA are bound by the MuN protein to form a noncovalent circular structure (Gloor and Chaconas 1986; Harshey and Bukhari 1983) . In an integrative transposition that requires the MuA protein, the flanking sequences are lost and the Mu DNA is randomly inserted into the bacterial genome. After integration, approximately 1-10% of the phages become lysogens (Howe and Bade 1975) . The majority of the phages immediately enter the lytic cycle (which is also the ultimate fate of the induced lysogens) during which MuA, assisted by the activities of several factors, catalyzes multiple replicative transpositions that eventually produce a burst size of approximately 100 new phages. Adapted from Choi and Harshey (2010) and Sokolsky and Baker (2003) The in cis orientation of the E element in mini-Mu with respect to the L/R ends is critical to its function, while its distance from the ends does not seem to be significant. The E element can stimulate transposition if it is present in trans in an unlinked DNA molecule (Surette and Chaconas 1992) . The presence of the E element is important for correct transpososome assembly ( Fig. 3 ) (Allison and Chaconas 1992; Lavoie and Chaconas 1995; Watson and Chaconas 1996) , and it remains associated with this complex throughout the transposition process (Pathania et al. 2002 (Pathania et al. , 2003 Yin et al. 2007) . MuA is a 663-aa residue protein that can be divided into three domains by partial proteolysis; each domain can be functionally or structurally divided into subdomains (Fig. 2b , and reviewed in Rice and Baker 2001) . The N-terminal domain of the MuA protein contacts the sequences of the L/ R ends and the E element through separate regions . The central domain is involved in the catalysis of the transposition reactions. The C-terminal domain is responsible for the interaction with the host protein, ClpX (see below), and with the auxiliary transposition factor, the MuB protein. (Wu and Chaconas 1995) . The minimal region required for DNA binding, cleavage, and strand transfer in vitro, including subdomains Iβ, Iγ, the catalytic core (domain II), and the N-terminal portion of domain III (IIIα), is indicated according to Krementsova et al. (1998) . Adapted from Baker and Luo (1994) , Harshey (2010), and Haapa-Paananen (2002) During the multistep reaction, the MuA present in the transpososome complex catalyzes the specific DNA singlestranded cleavages at the terminal CA dinucleotides of the L1 and R1 sites, using water as the nucleophile in such a way that 3′-OH groups of the transposable element are exposed (Mizuuchi 1984) . These 3′ ends serve as nucleophiles for the subsequent joining or strand-transfer step, attacking phosphodiester bonds spaced 5 bp apart on target DNA in a one-step transesterification reaction (Mizuuchi and Adzuma 1991) . The resulting θ-like DNA structure is a common intermediate for both transposition pathways (Chaconas and Harshey 2002; Harshey and Jayaram 2006) . The host ClpX unfoldase ultimately destabilizes the transpososome, which facilitates the recruitment of the host-dependent DNA replication/repair machinery to finalize the "nick-joinreplicative" or "nick-join-repair" transposition pathways ( Fig. 4) (Abdelhakim et al. 2008; Nakai et al. 2001) . In both cases, the strand transfer at staggered positions in target DNA strands will cause a duplication of 5 bp of the target DNA flanking the transposed Mu-based element. Probably, ClpX participates, as well, in stimulating the nuclease activity of the C-terminal domain of the transposase MuA (Wu and Chaconas 1995) in a highly regulated reaction that removes the attached host DNA after the incoming Mu genome has inserted into a bacterial chromosome according to the integrative transposition (Choi and Harshey 2010) . The structures of mini-Mus can differ significantly. Some mini-Mus carry all of the genetic elements essential for transposition, replication, and packaging. Other derivatives possess only the Mu ends (later designated as mini-Mu(LR) units) and can be complemented in trans to perform Muspecific functions in E. coli (Chaconas et al. 1981a; Harshey 1983; Patterson et al. 1986 ). Mini-Mu(cts) chromosomal vectors, which lack the genes required for Fig. 4 The two outcomes of Mu-driven transposition. The common θ intermediate can be resolved differently by the DNA repair/replication host-dependent machinery through integrative or replicative transposition. Transposition of the mini-Mu(LR) unit from the integrative plasmid (I) into the bacterial chromosome (BC) results in a "simple insertion" in which BC gains a copy of the unit or a "cointegrate" in which I and BC fuse and two copies of the mini-Mu unit border the junction of the fused replicons as direct repeats. The cointegrate can subsequently be resolved by homologous recombination between two mini-Mu units. Adapted from Abalakina et al. (2008a) and Choi and Harshey (2010) 3′-OH groups now function in the joining or strand-transfer step catalyzed by the complex and ClpX-mediated exchanges MuA with a set of host proteins in preparation for repair or replication to finalize the integrative or replicative transposition pathways, respectively. For details, see Au et al. (2006) , Harshey (2010), and Jayaram (2006) making phage particles, have been used for simultaneous thermo-induced replicative transposition and heterologous protein synthesis in E. coli (Weinberg et al. 1993) . However, the cis-encoded transposition factors in the mini-Mu(cts) constructs can cause instability in engineered strains even under noninduced conditions ). Stabilization has been achieved through the use of dual-component systems in which the genes for the MuA and MuB transposition factors are encoded in trans on an unlinked/nontransposed DNA molecule and eliminated after mini-Mu transposition. Different variants of this system have been developed Castilho et al. 1984; Chaconas et al. 1981a; Groenen et al. 1985; Patterson et al. 1986 ). One of the most popular systems includes an "integrative" plasmid that consists of a mini-Mu(LR) unit as the first component and a compatible "helper" plasmid carrying the inducible MuA and MuB genes as the second component. Usually, the helper plasmid possesses an unstable replicon and can be easily eliminated from the cells. The transposed genes in a set of mini-Mu(LR) units are flanked by Rho-independent transcription terminators in a directly repeated orientation (Abalakina et al. 2008a, b; Gulevich et al. 2009 ). The upstream terminator prevents readthrough transcription of the randomly integrated mini-Mu unit from the chromosomal promoter. In turn, the downstream terminator interrupts internal readthrough transcription of the mini-Mu that could interfere with transpososome assembly in vivo (Patterson et al. 1986 ). In the mini-Mu derivatives that could be assigned as mini-Mu(LER) units, an E element properly arranged between the L and R ends positively influences transposition (Leung et al. 1989 ). However, intense expression of the MuA and MuB genes, even when located in trans, leads to host cell lethality because of the overly efficient replicative transposition of the mini-Mu(LER) units Lee 2002) . Although artificial mini-Mu(LER) units are actively used in fundamental investigations of transpososome assembly and function in vitro and in vivo (Lee 2002) , they have not been widely applied in stably engineered E. coli strains. In contrast, in trans E element has been successfully used to significantly increase the transposition efficiency of the mini-Mu(LR) unit. A comparison of the E-carrier ((Eplus)-helper 1 ) and the helper that did not contain the E element ((E-minus)-helper 2 ) revealed a nearly two-order magnitude of increase in the transposition efficiency of the mini-Mu(LR) units when the (E-plus)-helper was used. To facilitate the selection of the transposed mini-Mu, an antibiotic resistance marker (Ant R ) can be included. 3 Markers that are flanked by the sequences essential for site-specific recombination, λattL/R (Peredelchuk and Bennett 1997) , loxP (Arakawa et al. 2001), or FRT (Datsenko and Wanner 2000) , are usually used. These markers can be excised in vivo from the mini-Mu units by the corresponding recombinase to obtain the marker-less recombinant strain (Abalakina et al. 2008a; Gulevich et al. 2009; Wei et al. 2010) . Several protocols have been developed for mini-Mu(LR) unit transposition into the E. coli chromosome with the help of the dual-component Mu-driven system. The best results have been obtained when an (E-plus)-helper plasmid is first transferred into the recipient strain and is stably maintained under noninducing conditions. Then, an integrative plasmid is transferred into the helper-carrier strain; the efficacy of this step mainly determines the total efficiency of the detected mini-Mu(LR) unit transposition. Therefore, different strategies that are based on the presence of a selective marker in the mini-Mu(LR) unit are employed. For a marker-carrier mini-Mu, replication of the integrative plasmid is not necessary in the recipient cell. So, the transposase can be induced during plasmid entry followed by direct selection of the clones-integrants. Usually, transposition occurs in the majority of cells that receive the integrative plasmid ). If the marker-less mini-Mu(LR) unit is used, the integrative plasmid has to be selectively transformed into the helper-carrier recipient. The transposase is then expressed in the cells that contain both autonomously replicating components of the Mu-driven system. Finally, the obtained clones-integrants are cured of the unstable plasmid(s). Usually, the transposition of the mini-Mu(LR) units can be detected in at least 10% of the induced cells ). Both strategies can lead to high-efficiency formation of clones possessing one or more (up to five to ten) copies of the mini-Mu(LR) unit in the chromosome ). The copy number of the integrated mini-Mu can be increased further by reintroducing and expressing the (Eplus)-helper plasmid in the corresponding plasmid-less strain-integrant. The (E-minus)-helper plasmid has a significantly decreased but still detectable ability to facilitate intrachromosomal replicative transposition of mini-Mu(LR) units in E. coli (Gak et al., in preparation). A dual-component Mu-driven system has been frequently used for E. coli chromosome editing and for constructing different bacterial strains for basic research and metabolically engineered amino acid production (Table 1) . In most cases, the expression level of the Muintegrated genes increased proportionally as the copy number in the chromosome increased. As a rule, the amplification of a biosynthetic operon leads to the increased accumulation of the corresponding amino acid. The amino acid production levels of the plasmid-less strains obtained were comparable to those of their plasmid-carrier recombinant analogs, and the former were significantly more stable during nonselective cultivation Savrasova et al. 2007) . The transposition of the mini-Mu(LR) unit into the bacterial chromosome mainly occurs through a "nick-joinreplicative" mechanism with the formation of a cointegrate in which the integrative plasmid and the chromosome fuse and two copies of the mini-Mu unit border the junction of the fused replicons as direct repeats (Fig. 4) (Abalakina et al. 2008a; Patterson et al. 1986; . The cointegrate is subsequently resolved by a reciprocal recombination event between the two mini-Mu units. The chromosomal structure of the final strain that contains one copy of the mini-Mu unit is the same as if integrative transposition of this unit had occurred. The sites of mini-Mu insertion in the E. coli chromosome could be precisely determined (Wei et al. 2010; Zimenkov et al. 2004 ) by inverse polymerase chain reaction (Ochman et al. 1988) . Several copies of the mini-Mu(LR) units can be easily detected in the bacterial genome after using the dualcomponent Mu-driven system. It could not be excluded that these multiple copies are due to independent transpositions from several copies of the nonreplicated integrative plasmid that were initially transformed into one cell or from several copies of autonomously replicated integrative plasmid. However, especially when the nonreplicating integrative plasmid was used, it seems more likely that the amplification is achieved due to the mini-Mu intrachromosomal replicative transposition. Certainly, the genome of the final stable plasmid-less multi-integrant strain may contain transposition-mediated chromosomal rearrangements, including inversions and/or deletions of bacterial DNA fragments (Watson et al. 2004 ). Interest in the use of methylotrophic bacteria for applied microbiology and biotechnology has increased in recent years (Schrader et al. 2009 ). Several genomes of methylotrophs have been sequenced (Chistoserdova et al. 2007; Vuilleumier et al. 2009 ). Significant progress has been made in elucidating the metabolism of these bacteria , and the number of tools available for genetic and metabolic engineering has expanded greatly (Bélanger et al. 2004; Choi et al. 2006; Lidstrom 2001, 2004) . Strategies to produce fine and bulk chemicals using methylotrophs have been described previously (Bourque et al. 1995; Fitzgerald and Lidstrom 2003; Motoyama et al. 1993 Motoyama et al. , 2001 . In particular, the obligate methylotroph Methylophilus methylotrophus AS1, which was extensively studied with respect to the industrial-scale production of single cell protein from methanol in the 1970s (Anthony 1982; Vasey and Powell 1984) , was recently metabolically engineered for the biosynthesis of L-lysine (Gunji and Yasueda 2006; Tsujimoto et al. 2006 ) and L-phenylalanine ) and for the efficient secretion of recombinant proteins (Itaya et al. 2008) . The adaptation of mini-Mu transposition for chromosomal editing in M. methylotrophus AS1 (Abalakina et al. 2008a, b) was a significant development. Mu-driven integration of the E. coli transporter gene into the methylotrophic genome was the basis for the development of an efficient method of constructing auxotrophic mutants with the following exploiting the developed system for metabolic engineering of L-phenylalanine overproduction (Iomantas and Abalakina 2002; Tokmakova et al. 2008 Tokmakova et al. , 2010 . The adaptation of the Mu-driven system to M. methylotrophus AS1 included several stages (Abalakina et al. 2008a ). Initially, a DNA fragment containing the genes from E. coli plasmid, pMH10 ), for the thermo-induced transposition factors was cloned into vectors containing the broad host range replicons of the IncQ (Chistoserdov and Tsygankov 1986) or IncPα groups (Ditta et al. 1985; Pansegrau et al. 1994) . These new (Eplus)-helpers (Fig. 5a) were transformed into methylotrophic cells (by mobilization or electroporation) and maintained under strictly selective conditions without expressing the Mu transposition factors. All of the helpers could be easily eliminated from M. methylotrophus AS1 by aerobically culturing the cells in liquid medium without antibiotics (Abalakina et al. 2008a) . The constructed integrative plasmids could not autonomously replicate in methylotrophic cells. Because it was unclear if it was possible for Mu-driven transposition to occur, the Mob RP4 + element (Simon et al. 1983 (Simon et al. , 1984 was included in the initial integrative plasmids to ensure their highly efficient transfer into M. methylotrophus by mobilization (Abalakina et al. 2008a) . The next generation of integrative plasmids (Fig. 5b , see experimental details in Tokmakova 2010) did not carry the Mob + element and could be autonomously maintained only in the pir + E. coli strain used to propagate γ-replicon originating from R6K (Bowers et al. 2007 ). These integrative plasmids were introduced into the methylotrophic recipient cells by electroporation. The Km R gene, flanked by FRT sites, was used as the "excisable" selective marker in the constructed mini-Mu(LR) units (Abalakina et al. 2008a) . The mini-Mu(LR) unit can be detected in the M. methylotrophus AS1 chromosome after the transfer of the integrative plasmid into the helper-carrier cells in the presence of the partially induced (37°C) transposition factors. Curing the selected integrants of the helper plasmid and Flp-mediated elimination of the marker results in a plasmid-less marker-less recombinant strain that is ready for the next round of mini-Mu transposition (Abalakina et al. 2008a) . The efficiencies (10 −2 -10 −3 ) of plasmid mobilization and Mu-driven integration are correlated. This means that, under the developed conditions, a mini-Mu(LR) unit is transposed into the bacterial chromosome of each cell of the recipient strain that received the integrative plasmid. As in E. coli, the transposition of the mini-Mu(LR) unit into the M. methylotrophus AS1 chromosome occurs through "nick-join-replicative" formation of the cointegrate, followed by its recombination-mediated resolution (Abalakina et al. 2008a) . These factors suggested that the mini-Mu(LR) units might be amplified in M. methylotrophus with the assistance of the same set of host proteins (Au et al. 2006; Gueguen et al. 2005; North and Nakai 2005) . However, the MuA-and MuB-mediated intrachromosomal duplication of the mini-Mu(LR) unit was detected only with a low (10 −4 ) frequency (Abalakina et al. 2008a) . The efficiency of the Mu-driven increase in genomic copy number in M. methylotrophus AS1 was increased 100fold by exploiting the mini-Mu(LER) unit (Fig. 5b) . When transferred as part of the integrative plasmid into M. methylotrophus, these units, like earlier used mini-Mu (LR), could be integrated into the bacterial chromosome with the formation of a cointegrate; the transposition efficiency of this process was equally high whether the helper was (E-plus) or (E-minus). On the contrary, the increase in genomic copy number efficacy of the mini-Mu (LER) or mini-Mu(LR) units was significantly dependent on the type of helper used ( Table 2 ). The detected capacities for increase in genomic copy number for the mini-Mu (LER) and mini-Mu(LR) units differed by approximately four orders of magnitude in the presence of the expressed (E-minus)-helper. These results served as the basis for a novel integration/ amplification/fixation strategy that was developed for the M. methylotrophus AS1 Mu-driven system. The E element in the mini-Mu(LER) unit was substituted with its excisable analog, which was bracketed by loxP-like sites [lox66 and lox71 (Albert et al. 1995) ] that could serve as the target for irreversible excision by phage P1 Cre recombinase (Abremski and Hoess 1984) . Thus, the mini-Mu(L[ ex E]R) unit (Fig. 5c ) could be integrated and amplified in the M. methylotrophus AS1 genome according to the standard procedure (using (Eplus)-or (E-minus)-helpers), followed by Cre-mediated excision of the E element. The "residual" copies of the mini-Mu(LR) units in the chromosome could not be further efficiently amplified in the presence of the (E-minus)-helper plasmid. This allows the integration and independent amplification of another mini-Mu(L[ ex E]R) unit into the M. methylotrophus AS1 genome without changing the number and location of the previously integrated genes (Fig. 6 ) . There are still many fundamental questions concerning the differences in the efficiencies of the Mu-driven increase in genomic copy number in E. coli and M. methylotrophus. The answers likely depend on the precise nature of the respective transpososome assembly processes. These pro- The increase in genomic copy number efficiency was evaluated as the part of multi-integrants obtained after induced transposition factors expression in the single-integrant cells selected on the media with increased concentrations of streptomycin (2 mg/ml of Sm) used as a selective marker according to Abalakina et al. (2008a) cesses could depend, in particular, on the intracellular concentrations of the expressed transposition factors and the DNA-bending host proteins that facilitate transposition at the "restrained" host DNAs. We have reviewed the adaptation of the dual-component Mu-driven system for chromosomal editing and the construction of plasmid-less marker-less strains of several Gram-negative bacteria due to the integration/amplification of target genes by the "nick-join-replicative" pathway in vivo. Certainly, other methods could be used for the chromosomal amplification of the target genes as well. The same purposes could be achieved through the single insertion of a DNA cassette with multiple copies of the target gene (Choi et al. 2006) . The marker-containing single insertions in different specific (Haldimann and Wanner 2001; Minaeva et al. 2008) or random (De Lorenzo and Timmis 1994; Peredelchuk and Bennett 1997; Wei et al. 2010 ) sites could be combined in one strain by general P1mediated transduction, followed by marker curing. These alternative methods, each of which possesses its own specific advantages, are significantly inferior to the Mudriven amplification approach with respect to ease-of-use and the speed with which a target can be achieved, especially if selection for the best variants of the multiintegrant is possible. Up today, in vitro-assembled Mu transpososome complexes have been efficiently applied for the in vivo random insertion of recombinant DNAs into different bacterial genomes by the "nick-join-repair" pathway (Haapa et al. 1999; Laasik et al. 2005; Lamberg et al. 2002; Lanckriet et al. 2009; Pajunen et al. 2005; Savilahti et al. 1995; Savilahti and Mizuuchi 1996; Tu Quoc et al. 2007; Wei et al. 2010; Wu et al. 2009 ). Moreover, efficient Mu transpososomebased integration has been verified even in yeast, mouse, and human genomes (Paatero et al. 2008; Turakainen et al. 2009 ). Many more host proteins likely participate in the process of Mu-driven replicative transposition than in the simple insertion of mini-Mu into the host chromosome (Au et al. 2006; North and Nakai 2005) . It is possible that only integrative, but not replicative, transposition occurs in this broad range of host organisms, even when efficient intracellular expression of Mu transposition factors is provided. However, to broaden the range of hosts in which the Mu-driven system can be used for the integration/ amplification of target genes, its adaptation should be attempted, especially by using mini-Mu(LR)/(LER) units in combination with different (E-plus)/(E-minus)-helpers expressed in different genetic backgrounds. The combined Mu-driven system could be proposed when the in vitro-assembled MuA-mediated transpososome Then, the E element is excised by Cre-mediated site-specific recombi-nation. The "truncated" mini-Mu(LR) units could be further amplified with very low frequency by (E-minus)-helper, i.e., their positions in the chromosome could be considered as fixated. So, the obtained multiintegrant could be served as the recipient for integration and independent amplification of new Mu(L[ ex E]R) units results single-copy integrants due to the integrative transposition, followed by the increase in genomic copy number of the integrated mini-Mu in vivo by the replicative transposition in the presence of the helper plasmid expressing MuA and MuB factors. Phage Mu-driven twoplasmid system for integration of recombinant DNA in the Methylophilus methylotrphus genome Exploiting of the Mu-phage transposition system for recombinant DNA integration in the Methylophilus methylotrophus chromosome Unique contacts direct high-priority recognition of the tetrameric Mu transposase-DNA complex by the AAA+ unfoldase ClpX Bacteriophage P1 site-specific recombination. Purification and properties of the Cre recombinase protein Development of the mini-Mu system providing an effective integration and amplification of the genetic material in the Evidence for a conservative pathway of transposition of bacteriophage Mu Site-specific integration of DNA into wild-type and mutant lox sites placed in the plant genome Role of the A protein-binding sites in the in vitro transposition of Mu DNA. II. A complex circuit of interactions involving the Mu-ends and the transpositional enhancer The commercial exploitation of methylotrophs Mutant loxP vectors for selective marker recycle and conditional knock-outs Chromosomal integration mechanism of infecting Mu virion DNA Identification of residues in the Mu transposase essential for catalysis Chromosomal editing in Escherichia coli. Vectors for DNA integration and excision Production of heterologous protein by Methylobacterium extorquens in high cell density fermentation High cell density production of poly-beta-hydrobutyrate (PHB) from methanol by Methylobacterium extorquens: production of high-molecularmass PHB Mechanism of origin activation by monomers of R6K-encoded π protein Influence of insertions on packaging of host sequences covalently linked to bacteriophage Mu DNA Random insertion of Mu-1 DNAwithin a single gene Plasmid insertion mutagenesis and lac gene fusion with mini-Mu bacteriophage transposons Transposition of phage Mu DNA In vitro and in vivo manipulations of bacteriophage Mu DNA: cloning of Mu ends and construction of mini-Mu's carrying selectable markers Predominant end-products of prophage Mu DNA transposition during the lytic cycle are replicon fusions DNA transposition: jumping gene machine, some assembly required Broad host range vectors derived from an RSF1010::Tn1 plasmid Genome of Methylobacillus flagellatus, molecular basis for obligate methylotrophy, and polyphyletic origin of methylotrophy The expanding world of methylotrophic metabolism DNA repair by the cryptic endonuclease activity of Mu transposase Multicopy integration and expression of heterologous genes in Methylobacterium extorquens ATCC 55366 Genetic engineering using homologous recombination Mechanism of transposition of bacteriophage Mu: structure of a transposition intermediate Site-specific recognition of the bacteriophage Mu ends by the Mu A protein Heteroduplex structures of bacteriophage Mu DNA One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products Analysis and construction of stable phenotypes in gram-negative bacteria with Tn5-and Tn10-derived minitransposons Promoter knock-in: a novel rational method for the fine tuning of genes Bacterial nucleoid-associated proteins, nucleoid structure and gene expression Plasmid related to the broad host range vector, pRK290, useful for gene cloning and for monitoring gene expression YddG from Escherichia coli promotes export of aromatic amino acids Construction of an L-phenylalanine-producing tyrosine-prototrophic Escherichia coli strain using tyrA ssrA-like tagged alleles Pho regulon promoter-mediated transcription of the key pathway gene aroG Fbr improves the performance of an L-phenylalanine-producing Escherichia coli strain DNAprotein complexes during attachment-site synapsis in Mu DNA transposition Transposon Mu as a genetic tool for stabilized amplification of genetic information Stabilized amplification in Gram negative bacteria with Mu phage and its application in L-threonine production The yeaS (leuE) gene of Escherichia coli encodes an exporter of leucine, and the Lrp protein regulates its expression Novel mutants of Erwinia carotovora subsp. carotovora defective in the production of plant cell wall degrading enzymes generated by Mu transpososome-mediated insertion mutagenesis Efficient insertion mutagenesis strategy for bacterial genomes involving electroporation of in vitro-assembled DNA transposition complexes of bacteriophage Mu Generation of single-copy transposon insertions in Clostridium perfringens by electroporation of phage Mu DNA transposition complexes Transposition of phage Mu DNA Structural aspects of a higher order nucleoprotein complex: induction of an altered DNA structure at the Mu-host junction of the Mu type 1 transpososome Anatomy of a flexer-DNA complex inside a higher-order transposition intermediate University of Texas at Austin Leung PC, Teplow DB, Harshey R (1989) Interaction of distinct domains in Mu transposase with Mu DNA ends and an internal transpositional enhancer Conservative integration of bacteriophage Mu DNA into pBR322 plasmid Methods of making amino acids using E. coli transformed with csc genes Microarray analysis of transposition targets in Escherichia coli: the impact of transcription Development of improved versatile broad-host-range vectors for use in methylotrophs and other gram-negative bacteria Development of an insertional expression vector system for Methylobacterium extorquens AM1 and generation of null mutants lacking mtdA and/or fch Fluorescent proteins from nonbioluminescent Anthozoa species New tool for metabolic pathway engineering in Escherichia coli: one step method to modulate expression of chromosomal genes Dual in/out strategy for genes integration into bacterial chromosome: a novel approach to step-by-step construction of plasmid-less marker-less recombinant E. coli strains with predesigned genome structure Mechanism of transposition of bacteriophage Mu: polarity of the strand transfer reaction at the initiation of transposition Inversion of the phosphate chirality at the target site of Mu DNA strand transfer: evidence for a one-step transesterification mechanism Efficient Mu transposition requires interaction of transposase with a DNA sequence at the Mu operator: implication for regulation Assembly of phage Mu transpososomes: cooperative transitions assisted by protein and DNA scaffolds Bacteriophage Mu genome sequence: analysis and comparison with Mu-like prophage in Haemophilus, Neisseria and Deinococcus Amino acid production from methanol by Methylobacillus glycogens mutants: isolation of L-glutamic acid hyper-producing mutants from M. glucogenes strains, and derivation of Lthreonine-and L-lysine-producing mutants from them Overproduction of L-lysine from methanol by Methylobacillus glycogens derivatives carrying a plasmid with a mutated dapA gene Handoff from recombinase to replisome: insights from transposition Host factors that promote transpososome disassembly and the PriA-PriC pathway for restart primosome assembly Search for integrationdeficient mutants of bacteriophage Mu Genetic applications of an inverse polymerase chain reaction Bacteriophage Mu integration in yeast and mammalian genomes Generation of transposon insertion mutant libraries for Gram-positive bacteria by electroporation of phage Mu DNA transposition complexes Complete nucleotide sequence of Birmingham IncP alpha plasmids. Compilation and comparative analysis The bacteriophage Mu. In: Calendar R (ed) The bacteriophages Path of DNA within the Mu transpososome: transposase interacts bridging two Mu ends and the enhancer trap five DNA supercoils A unique right endenhancer complex precedes synapsis of Mu ends: the enhancer is sequestered within the transpososome throughout transposition Transposition studies of mini-Mu plasmids constructed from the chemically synthesized ends of bacteriophage Mu A method for construction of E. coli strains with multiple DNA insertions in the chromosome The nucleoid. In: Neidhardt FC (ed) Escherichia coli and Salmonella: cellular and molecular biology Supercoils in prokaryotic DNA restrained in vivo Comparative architecture of transposase and integrase complexes Assisted large fragment insertion by red/ET-recombination (ALFIRE)-an alternative and enhanced method for large fragment recombineering Differential role of the MuB protein in phage Mu integration vs. replication: mechanistic insights into two transposition pathways Mu transpositional recombination: donor DNA cleavage and strand transfer in trans by the Mu transposase The phage Mu transpososome core: DNA requirements for assembly and function Method for preparing L-amino acids by fermentation of mixture of glucose and pentoses. Patent RU2273666 Creation of mini-Mu system without selective markers for integration of genes into Use of the valine biosynthetic pathway to convert glucose into isobutanol Recombineering: in vivo genetic engineering in E. coli, S. enterica, and beyond Methanol-based industrial biotechnology: current status and future perspectives of methylotrophic bacteria Recombineering: a homologous recombination-based method of genetic engineering A broad host range mobilization system for in vivo genetic engineering: transposon mutagenesis in gram negative bacteria High frequency mobilization of gram-negative bacterial replicons by the in vitro constructed Tn5-mob transposon Design and study on characteristics of auto-and smoothly regulated genetic element O 3 /P lacUV5 /O lac →lacI DNA gyrase requirements distinguish the alternate pathways of Mu transposition A protein factor which reduces the negative supercoiling requirement in the Mu DNA strand transfer reaction is Escherichia coli integration host factor The Mu transpositional enhancer can function in trans: requirement of the enhancer for synapsis but not strand cleavage Action at a distance in Mu DNA transposition: an enhancer-like element is the site of action of supercoiling relief activity by integration host factor (IHF) IHF and HU: flexible architects of bent DNA Aerobic catabolism of threonine in Escherichia coli strain with feedback resistant biosynthetic threonine deaminase Development of chromosomal editing methods for metabolic engineering of an obligate methylotrophic bacteria Methylophilus methylotrophus AS1 Method for imparting auxotrophicity on L-amino acid to Methylophilus genus bacteria, and method of producing L-amino acid L-lysine biosynthetic pathway of Methylophilus methylotrophus and construction of an L-lysine producer Isolation and characterization of biofilm formation-defective mutants of Staphylococcus aureus Transposition-based method for the rapid generation of gene-targeting vectors to produce Cre/Flp-modifiable conditional knock-out mice Stabilized gene duplication enables long-term selection-free heterologous pathway expression Single-cell protein Methylobacterium genome sequences: a reference blueprint to investigate microbial metabolism of C1 compounds from natural and industrial sources Three-site synapsis during Mu DNA transposition: a critical intermediate preceding engagement of the active site Molecular biology of the gene, 5th edn A mini-Mu transposon-based method for multiple DNA fragment integration into bacterial genomes A chromosomal expression vector for Escherichia coli based on the bacteriophage Mu A novel DNA binding and nuclease activity in domain III of Mu transposase: evidence for a catalytic region involved in donor cleavage Mu transposition complex mutagenesis in Lactococcus lactisidentification of genes affecting nisin production Interactions of phage Mu enhancer and termini that specify the assembly of a topologically unique interwrapped transpososome Aromatic amino acid auxotrophs constructed by recombinant marker exchange in Methylophilus methylotrophus AS1 cells expressing the aroP encoded transporter of Escherichia coli The regions in E. coli chromosome preferable for phage Mu-driven system of gene integration Acknowledgements This paper is devoted to the memory of Prof. Yurii I. Kozlov (1944 who initiated the investigation and application of the Mu-driven integration/amplification systems at AGRI. The authors are extremely thankful to all of the AGRI members who participated in these investigations. In particular, we thank Drs. E.G. Abalakina Open Access This article is distributed under the terms of the Creative Commons Attribution Noncommercial License which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.