ABSTRACT
The RmInt1 group II intron is an efficient self-splicing mobile retroelement that catalyzes its own excision as lariat, linear and circular molecules. In vivo, the RmInt1 lariat and the reverse transcriptase (IEP) it encodes form a ribonucleoprotein particle (RNP) that recognizes the DNA target for site-specific full intron insertion via a two-step reverse splicing reaction. RNPs containing linear group II intron RNA are generally thought to be unable to complete the reverse splicing reaction. Here, we show that reconstituted in vitro RNPs containing linear RmInt1 ΔORF RNA can mediate the cleavage of single-stranded DNA substrates in a very precise manner with the attachment of the intron RNA to the 3´exon as the first step of a reverse splicing reaction. Notably, we also observe molecules in which the 5´exon is linked to the RmInt1 RNA, suggesting the completion of the reverse splicing reaction, albeit rather low and inefficiently. That process depends on DNA target recognition and can be successful completed by RmInt1 RNPs with linear RNA displaying 5´ modifications.
KEYWORDS: Group II introns, linear RNA, reverse splicing, Sinorhizobium meliloti, RmInt1
Introduction
Group II introns are autocatalytic RNAs that are spliced out of a precursor transcript (pre-RNA) in a two-step mechanism resembling the excision of spliceosomal introns [1,2]. This mechanism involves two sequential transesterification reactions: the 2´-OH group of the bulged adenosine residue in domain VI first attacks the phosphate group at the 5´ splice site, generating the branched intron structure and exposing the 3´-OH group of the 5´ exon. This activated group then reacts with the phosphate group at the 3´ splice site, joining the exons and releasing the intron lariat RNA [3]. For some introns, the first step in the splicing process is hydrolysis, yielding a free linear intron RNA rather than the lariat form [4–8]. Moreover, an excision mechanism generating RNA circles has been observed for some group II introns. This mechanism involves the release of the 3´ exon and a further attack on the first intron residue by a reactive OH group on the terminal nucleotide [9–13].
Some group II introns also act as mobile retroelements, integrating into new genomic locations by a reverse splicing process [14–19]. The functional unit for mobility is a ribonucleoprotein particle (RNP) consisting of the lariat intron RNA and an intron-encoded protein (IEP). Target recognition mostly involves base-pairing interactions between the exon-binding sites (EBSs) in domain I of the intron RNA and the corresponding sequences at the target site (intron-binding sites, IBSs). The intron RNA reverse splices into the top strand and is then reverse-transcribed by the IEP [20]. Host proteins are then responsible for completing the insertion process [21–25]. This reaction has been extensively demonstrated for RNPs formed with lariat intron RNA, but linear group II intron RNAs are generally thought to be unable to complete the reverse splicing reaction [26,27].
The Sinorhizobium meliloti RmInt1 group II intron is an efficient, but unusual mobile retroelement [28] that self-splices in vitro to generate lariat, linear and circular intron molecules together with a wide variety of truncated variants of all of them and aberrant intermediates [10,29,30], but is excised in vivo to form intron lariat and circle RNAs [10]. RmInt1 encodes an IEP without a DNA endonuclease domain (En−) [31], but with a C-terminal extension involved in intron function [32]. RmInt1 displays a strong bias for reverse splicing into the lagging strand template during replication, and it has been suggested that this intron recognizes its target principally as a single-stranded (ss)DNA [33]. We recently showed that RmInt1 lariat RNPs reconstituted from self-spliced intron precursor and a maltose binding protein (MBP)-IEP fusion protein could reverse-splice into DNA substrates in vitro [34].
Here, we used the RmInt1 linear intron RNA and the MBP-IEP fusion protein to reconstitute linear RmInt1 RNPs, and to determine their biochemical activities and requirements for DNA cleavage and reverse splicing. We show, for the first time, that RNPs containing linear intron RNA can complete the precise joining of the RNA molecule to both exons of a DNA substrate in vitro. We also show that the efficiency and fidelity of this reaction are compromised by modifications/insertions at the 5´ or 3´ end of the intron sequence or by the proper recognition of the DNA target sequence by the intron RNA. These findings could have important implications for the spread and mobility mechanisms of group II introns that excises as linear molecules in vivo.
Results
DNA cleavage and reverse splicing of RNPs containing linear RmInt1 RNA
We previously showed high levels of in vitro cleavage and partial reverse splicing into ssDNA substrates for RNPs containing the RmInt1 lariat, whereas full reverse splicing was barely detectable [34]. To test the ability of RNPs reconstituted in vitro with the linear RmInt1 (ΔORF) RNA to perform these activities, we carried out DNA cleavage assays with 5´ and 3´ end-32P-labeled 70 nt ssDNA substrates encompassing the RmInt1 target site. The products obtained were separated by electrophoresis in a denaturing polyacrylamide gel (Figure 1A). We found that the RNPs containing linear intron RNA cleaved the DNA substrate as efficiently as RNPs containing lariat RNA (lanes 3–4 versus lanes 5–6). Notably, low-mobility bands were detected at the top of the gel regardless of the end-labeled substrate used. Both types of RNPs yielded mostly partial reverse splicing products (Figure 1A, top of the gel, lanes 4 and 6), but the RNPs containing linear intron RNA also generated products corresponding to intron RNA molecules attached to the 5´ exon, similar to the RNPs containing lariat intron RNA (5´ end-labeled substrate, lanes 3 and 5). As expected, DNA cleavage/reverse splicing activity was not detected in the absence of the RNP complex (-RNPs, lanes 1 and 2) or for RNPs reconstituted with a catalytically inactive linear intron RNA (Linear dV mutant RNPs, lanes 7 and 8). To confirm the nature of the low-mobility bands, we subjected DNA cleavage reaction mixtures to RNase digestion before gel electrophoresis (Figure 1B). The DNA cleavage bands remained visible at the bottom of the gel, while the bands migrating with a low mobility, possibly corresponding to the linear intron RNA attached to the 5´ exon (lane 3) or to the 3´ exon (lane 4), completely disappeared on RNase treatment (lanes 1 and 2). Similar results were obtained by subjecting reaction mixtures to alkaline digestion at 70°C for 10 minutes (data not shown). These results confirm that the bands at the top of the gel are RNA-DNA hybrid molecules susceptible to RNA degradation.
Figure 1.

Biochemical activities of RNPs reconstituted with linear RmInt1 intron RNA. (A) Comparison of the DNA cleavage activities of linear and lariat intron RNPs with DNA substrates radioactively labeled at their 5´ (5´*) and 3´ (3´*) ends. (B) Verification of hybrid RNA/DNA molecules by RNase digestion. The bands are identified on the right and their sizes are indicated on the left. 5´ DNA exon, black box; 3´ DNA exon, white box; RNA, thin line. Detailed panels correspond to the same area of the gel overexposed.
Time-courses of DNA cleavage/reverse splicing of RNPs containing linear intron RNA (Figure 2) revealed that almost immediately after the mixing of the reaction components, products derived from DNA cleavage and low migrating bands were detected with both 5´ end- and 3´ end-labeled substrates. We detected 3´ end-labeled low-mobility products barely the two minutes after the addition of RNPs to the reaction tube, and the amounts of these products reached the saturation point 30 minutes into the reaction (left panel). However, even though the 5´ end-labeled product resulting from DNA cleavage also appeared two minutes after initiation of the reaction, the band at the top of the gel corresponding to a product in which the DNA substrate is attached to the intron RNA was of a much weaker intensity than that for the partial reverse splicing product (see the upper plot for quantification). Similar behavior was observed in assays with RNPs reconstituted with lariat RNA (Figure S1). A strong accumulation of free exons was observed after two hours of reaction (Figs. 1A and 2), whereas the levels of low-mobility products stabilized after about 30 minutes (Figure 2, lanes 9–10, both panels). Spliced exon reopening (SER) has been described for other group II introns and is usually observed during forward-splicing assays, consistent with the hydrolytic cleavage of the joined exons at the splice site catalyzed by group II introns, but without intron insertion [4,35]. The accumulation of free 32P-labeled 3´ exon may therefore indicate the occurrence of a parallel reaction contributing to exon reopening. Taken together, our data indicate that the first step of RmInt1 linear intron reverse splicing is rapid, specific and precise, but that the completion of intron integration is rate-limiting.
Figure 2.

Time-course of the reaction between the 5´-triphosphorylated linear intron and the DNA substrates labeled at the 3´end (left) or 5´end (right gel). Samples were taken at 0, 2, 5, 10, 15, 20, 30, 60, 120, and 300 minutes. The progress of the reaction for band quantification was plotted over time, with reactions performed with 3´end-labeled DNA substrates shown in gray and reactions carried out with 5´end-32P-labeled DNA substrates shown in black. The inset graph in the reverse splicing plot focus closely on the reverse splicing levels of RNPs reconstituted with linear RNA on 5´-labeled DNA substrates. For product identification, see the caption to Figure 1; *, labeled molecules.
Identification of low migrating intron products generated with RNPs containing linear RmInt1 RNA
To test whether the linear intron was linked to the 5´and 3´exons, the low-mobility products of the reaction of linear intron RNPs with labeled ssDNA substrates were gel-purified and sequenced after RT-PCR (Figure 3). Different combinations of primer pairs were used to detect the putative fully reverse-spliced product (primer 1 and 4) and to identify sequences at the 5´and 3´- integration junctions from the linear intron RNA (5´ junction, primers 1/2; and 3´ junction, primers 3/4). PCR products were cloned and sequenced as individual clones. The sequences of low-mobility products from different independent reactions were determined (Figure 3). Amplification with a primer pair binding to the 3´ intron-exon junction mostly generated a product corresponding to the 3´ end of the RmInt1-ΔORF intron attached to the 3´ exon. By contrast, an analysis of the 5´-integration junction revealed that most of the products corresponded to the 5´end of the intron attached to the 5´exon, but that smaller products were also generated, consistent with an intron truncated at nucleotide position 31, close to a loop region in subdomain IA. The analysis of the sequence at that position did not shed light about the appearance of this product because we could not identify any sequence resembling the intron EBSs. Finally, the sequencing of products amplified from a cDNA obtained with random oligonucleotides and primers binding to both exons identified the full-length RmInt1 ΔORF intron, inserted into the target site. Thus, our results indicate that RNPs containing the linear RmInt1 RNA can promote the full reverse splicing of the intron into ssDNA substrates at the correct insertion site in vitro. This is the first time, to our knowledge, that full reverse splicing has been detected in vitro for group II intron RNPs reconstituted with linear intron RNA.
Figure 3.

Identification of the reverse splicing products obtained with RNPs reconstituted with linear RmInt1 intron RNA. (A) The sites of primer binding for PCR amplification are shown. 5´ DNA exon, black box; 3´ DNA exon, white box; RNA, thin line. (B) A representative gel of the amplification products corresponding to the 5´ junction, 3´ junction and the full reverse splicing molecule is presented. The template for the negative control (−) was the −35/ΔORF/+75 RNA precursor transcript. The template for the positive control (+) was the cDNA obtained from a run-off transcript of the precursor intron containing −35/+75 exon sequences. Lanes −/+ RT identify the templates derived from the −/+ reverse transcription of the low-mobility bands extracted from polyacrylamide gels. (C) Sequences of the 5´ and 3´ integration junction resulting from the reverse splicing of linear RNPs, obtained by RT-PCR. Numbers on the right side correspond to the number of individual clones presenting the same sequence for the 5´ junction (upper sequences) and for the 3´ junction (lower sequences). The sequences derived from the full reverse splicing product (Full Rv Sp) are excluded in these surveys. The position numbers at the beginning and end of the intron sequences are indicated as for the ΔORF intron sequence. The expected sequence is shown in bold and the underlined sequences represent the intron binding sites (IBS1, 2 and 3). The experiment was performed at least four times, with different RNPs and substrate preparations.
DNA cleavage and full insertion of linear RmInt1 RNA depend on DNA target recognition
The RmInt1 RNP complex containing the lariat intron recognizes the DNA target site via both its RNA and protein components [36,37]. In particular, the RNA component base-pairs via its EBS sequences to nucleotide residues in the 5´ and 3´ exons (mostly at positions −13 to +1 relative to the intron insertion site), whereas the IEP interacts with distal positions in the 5´ exon (T −15 nucleotide). We assessed the relevance of target-site recognition for the efficiency of DNA cleavage and full reverse splicing in vitro for RmInt1 linear RNA, by incubating RNPs reconstituted with linear RmInt1 introns with 5´ end-labeled ssDNA substrates modified at different positions in the DNA target site (Figure 4). The substitutions made were chosen on the basis of retrohoming results obtained in vivo [36]: in the 5´ exon, the sequence was independently modified at T-15 (−15G), T-12 (−12G), and G-2 (−2C); and in the 3´ exon, the sequence was modified at C + 1 (+1A) and G + 4 (+4T). All mutant target DNAs affected DNA cleavage and/or full reverse splicing efficiency (Figure 4). The modifications to the 3´ exon slightly increased the DNA cleavage activity of the linear RNA RNPs, but did not proportionally increase the amounts of linear intron RNA attached to the 5´exon (lanes 11 and 12). Moreover, the intron RNA yielded a double cleavage band when incubated with the +1A mutant DNA substrate (lane 11), probably due to the misrecognition by EBS3 (G). Nucleotide modifications to the 5´ exon had different effects on DNA cleavage and full reverse splicing activities. The −2C substitution affecting IBS1-EBS1 pairing increased DNA cleavage and full intron insertion into the target site (lane 10). By contrast, nucleotide changes in the IBS2 (−12G) or distal 5´ exon (−15G) regions decreased DNA cleavage. However, only modifications affecting the distal region of the 5´ exon, potentially recognized by the IEP, resulted in lower levels of the low migrating product (lane 8). As expected, neither dV-mutant intron RNA (lanes 1 to 6) nor RNPs mixed with a substrate with an alien sequence (data not shown) generated any specific DNA cleavage-derived products. Similar results were obtained for DNA cleavage with RNPs containing lariat RNA, although the reverse splicing rates might indicate that the RNPs containing lariat RNA discriminate better the different mutated DNA substrates than those containing linear RNA (Figure S2, compare lanes 2 and 8, or 3 and 9). Then, the insertion of the linear intron into the target DNA therefore seems to be limited mostly by recognition of the distal 5´exon region enabling EBS-IBS pairing to occur, as with RNPs containing lariat intron RNA. Nevertheless, the lariat conformation in the RNP complexes seems to constrict the recognition the DNA substrate for reverse splicing.
Figure 4.

RNP activity is dependent on DNA target recognition. DNA cleavage and reverse splicing products observed when 5´ end-labeled point-mutated DNA substrates react with wild-type or domain V-inactive linear intron RNP complexes. The detailed upper panel corresponds to the same area of the gel overexposed. Quantification graphs show the mean values for three independent assays with different RNP preparations. Error bars correspond to the standard errors.
Linear RmInt1 RNA modified at the 5´ and 3´ ends can still insert into the target DNA site
We assessed whether modifications at the 5´ and 3´ ends of the linear intron RNA influence its insertion into the DNA target. RNPs prepared with various linear RmInt1 RNAs modified at their boundaries were incubated with 70 nt 5´ and 3´end-32P-labeled ssDNA substrates (Figure 5). Modifications to the 5´ end of the intron had little, if any, effect on intron activity (lanes 3 to 8), whereas changes at the 3´ end affected intron functionality (lanes 9 to 16). DNA cleavage and reverse splicing levels were similar to those of the wild-type RmInt1 RNA when intron transcripts extended by 5 nucleotides at the 5´-end were used for RNP reconstitution, indicating that intron insertion may involve the incorporation of extra residues at the 5´ end of the intron (compare lanes 1 and 2 with 3 and 4, or 7 and 8), which was further confirmed by amplification and sequencing the corresponding low-mobility products. In addition, we found that the first position (G) of the intron RNA seemed to be dispensable for DNA cleavage and intron insertion in vitro (lanes 5 and 6), but a C-to-A point mutation (C740A) at the 3´ end of the intron RNA decreased intron functionality (lanes 15 and 16), whereas the same substitution at the penultimate position in the intron (C739A) had no effect on intron activity (lanes 13 and 14). The presence of these modifications in the linear intron molecules inserted between the two DNA exons was checked by sequencing. Finally, the addition of 5 nt at the 3´end of the intron greatly decreased reverse splicing reactions regardless of whether the last nucleotide was an A or the natural C (lanes 9 to 12). Together, these data suggest that the linear RmInt1 RNA can be modified at its 5´end with no significant effect on the activity of the intron RNP.
Figure 5.

Modifications to the ends of the intron RNA used for reconstitution alter RNP activity. (A) Diagram showing the wild-type intron RNA and the intron RNAs with modifications to the 5´ end and the 3´ end used for the preparation of RNPs. Sequence insertions and substitutions are shown in bold. (B) DNA cleavage and reverse splicing products observed when substrates radioactively labeled at their 5´ end (5´*) or 3´ end (3´*) were incubated with the different linear intron RNP preparations. The lower panel corresponds to the bottom part of the gel, whereas the upper panel shows the top of the gel but after a longer exposure. For product identification, see the legend to Figure 1.
Discussion
The reconstitution of RNPs in vitro has been a cornerstone of the development of potential biotechnological applications for group II introns since the late 1990s [34,38–45]. We have analyzed here the DNA cleavage and reverse splicing activities of RNPs reconstituted with linear RmInt1 RNA and an MBP-IEP fusion protein and show that they can catalyze the attachment of the intron RNA to both exons of single-stranded DNA substrates in vitro, unlike other described group II introns [43,46–48].
RNPs containing linear RmInt1 RNA catalyzed the covalent binding of the intron RNA to DNA exons but also the hydrolytic cleavage of DNA substrates
In vitro endonuclease reactions on double-stranded DNA substrates have shown that Ll.ltrB group II intron RNPs generate a double-strand break leading either to complete intron insertion if the lariat RNA is used for reconstitution [38,39,43], or to partial insertion of the intron into the top strand if linear RNA is used for RNP reconstitution [43]. Other linear group II introns lacking a protein partner can bind to the 3´ exon (DNA or RNA) but cannot reverse the first step of forward splicing [46,47], even if a potential leaving group is placed at the 5´ end [48]. Like these introns, the linear RmInt1 RNA attaches to a precise position in vitro, in the 3´ exon, due to the reversal of the second step of the splicing reaction. Furthermore, RmInt1 RNPs containing linear RNA catalyze the binding of the linear ribozyme to the last residue of the 5´ exon. As we were able to sequence the low-mobility bands isolated from our DNA cleavage reactions, it therefore seems likely that the 5´ exon is linked to the 5´triphosphate linear intron molecules by a 3´-5´phosphodiester bond releasing pyrophosphate. Nevertheless, preliminary results indicate that the attachment of the 5´ exon to a 5´ monophosphate linear molecule is also produced with similar efficiency (data not shown). Previous studies on other group II introns have yielded conflicting results. Experiments using a linear ai5γ intron modified with a suitable leaving group (a triphosphate or an m7G-cap) at the 5´ end generated no products corresponding to reversal of the first step in splicing [48], whereas a linear bI1 intron extended 12 nt at its 5´end and attached to the 3´exon seemed to use the breaking energy of an α-β bond in a triphosphate group for transesterification reconstituting a 5´ exon-12 nt-extended intron RNA-3´ exon molecule [47]. The dephosphorylated version of this 12 nt-extended bI1 intron also supported transfer of the 5´ exon, albeit less efficiently and generating molecules truncated at the 5´ exon-intron junction [47]. Future research should determine the mechanism by which the linear RmInt1 RNA is bound to the 5´exon.
It is plausible that molecules of 5´ exon covalently bound to the linear RmInt1 RNA but with no 3´ exon could be produced by the reaction of a hydroxyl group at the end of free 5´ exon generated by SER with the 5´ end of the triphosphorylated linear intron [47]. However, reverse splicing assays including only the 5´ exon and RmInt1 linear RNA RNPs have shown that these products did not react (Figure S4). Furthermore, we detected full intron insertion at the target site, demonstrating the occurrence of complete reverse splicing.
The differences in band intensity between experiments performed with 5´ and 3´ end-labeled substrates suggest that the second step in the reverse splicing of linear RmInt1 intron RNA may be rate-limiting. This is not particularly surprising given the low levels of full reverse splicing observed for RmInt1 lariat RNPs with ssDNA substrates [34] or for the lariat form of other group II introns in vitro with their own RNA substrates [46–48]. Indeed, the reverse branching reaction is generally considered to be inefficient in wild-type introns due to structural constraints on the catalytic core due to the change in conformation occurring between the two steps of reverse splicing [41,49–53]. On the other hand, our time-course experiments showed a rapid first-phase reaction, followed by a slow second-phase reaction. These observations are similar to the kinetic findings for the reverse splicing of other lariat (full reverse splicing) and linear (partial reverse splicing) group II introns in vitro. In these situations, the biphasic kinetics can be explained by three concomitant reactions: reverse splicing, forward splicing and SER [48,51,54]. We observed the accumulation of free exons over long reaction times. We therefore assumed that SER occurred and that this irreversible reaction competed for the substrate. The observed slower rate of hydrolytic cleavage of the substrate (SER) by linear RmInt1 RNPs than by other group II introns [48] may be related to the presence of the IEP, which not only facilitates correct intron RNA folding, but may also restrict the access of water or metal-coordinated hydroxyl groups to the catalytic core.
The RmInt1 IEP may control the rate of reverse splicing of the RmInt1 RNA
For group II introns, insertion site selection has been shown to be associated not only with the precise structure of the catalytic core, but mostly, to the correct recognition of the IBS-EBS helical structure [55,56]. As expected, DNA cleavage and reverse splicing requirements in vitro of linear RmInt1 RNPs are similar to those described for in vivo retrohoming of the RmInt1 intron [36,57]. However, conditions in vitro seem to be more permissive because single mutations of the DNA target that completely abolish retrohoming in vivo do not abolish reverse splicing activity in vitro. The residues directly involved in the base pairing with the intron RNA seem to be important for cleavage of the DNA substrate. However, the mutant target −15G (5´exon distal region) did not only produces a ladder of aberrant cleavage products, but also it allows even lower reverse splicing levels than the other mutated DNA substrates. It has been suggested that this residue is recognized by the RmInt1 IEP [36], which would suggest that the rate of reverse splicing is also dependent on the correct positioning of the protein component of the RNP complex. Indeed, an analysis of the cryo-EM structure of the L. lactis L1.LtrB group II intron in complex with its IEP suggested a direct role of the thumb domain of the LtrA protein in the stabilization of EBS-IBS interactions [58]. Moreover, it has been considered that the RNPs bind to its DNA target nonspecifically, scan the DNA, and eventually a rate-limiting conformational event of the complex occurs at suitable sequences, when the specific recognition of the DNA target by the protein component take place [41]. We cannot rule out that the architecture of the linear intron and/or the large MBP-IEP fusion protein could impair the structural reorganization of the RNP complex to recognize specifically the DNA target, generating an abundance of unproductive reaction complexes for integration. Thus, it is plausible that the forward splicing reactions are favored in our in vitro assays, resulting in a very low detection of reverse splicing products.
Modifications of nucleotides at RNA-intron boundaries may alter catalytic core architecture
The nucleotides at the intron boundaries of group II introns are strongly conserved, with consensus sequences of GUGYG at the 5′ end of the intron and AY (where Y is C or U) at the 3′ end [59], although the RmInt1 intron has two C residues at its 3´ end. The effect of these terminal residues has been studied in detail only for forward splicing efficiency. Lariat formation seems to require a non-Watson–Crick interaction between the first and penultimate nucleotides of the intron [60,61]. It is therefore feasible that the absence of the G + 1 nucleotide or the replacement of the penultimate nucleotide of the linear RmInt1 ΔORF RNA had little effect on reverse splicing efficiency. Furthermore, the last residue of the intron (γ′) is engaged in a Watson–Crick interaction with a purine (γ) residue in the segment separating domain 2 from domain 3 (J2/3), which is required for the positioning of the 3′ splice site for exon ligation [62–64] and stabilization of the catalytic core [52,58]. Despite discrepancies concerning the location of the γ position in structural studies, either in [52,58] or outside the active center [53,65], the γ´ residue seems to be located in the catalytic core. It therefore seems possible that substitution of the 3´terminal nucleotide of the intron decreases the activity of the linear intron RNP.
Finally, our detection of 5´ exon molecules linked to linear intron RNAs carrying a five-nucleotide extension at their 5´ end is a remarkable finding. Such a situation has been described for 10 organelle group II introns lacking the bulging A of domain VI, in which the end of the IBS1 sequence and the GUGYG consensus sequence are separated by one to 33 intervening nucleotides [8]. However, the catalytic core is subject to structural constraints [52,53,58,65], and the presence of additional nucleotides at intron boundaries (5´and 3´ends), and, consequently, in the active site of the ribozyme, may distort the conformation of the reactive groups, decreasing reverse splicing activity.
Potential biological relevance
Natural group II introns excising exclusively as linear molecules appear to have lost their mobility [6,8]. Interestingly, some bacterial introns generating only trace amounts of branched products during in vitro splicing are present as multiple copies in their host genomes [66–68]. Moreover, Lambowitz and coworkers demonstrated the retrohoming of a linear group II intron (Ll.ltrB) in eukaryotes by partial reverse splicing followed by reverse transcription of the intron RNA. The cDNA generated was then attached to the 5´ exon through a variant of the non-homologous end-joining (NHEJ) mechanism, a process dependent on host enzymes (Lig4 and PolQ) that frequently leads to imprecise intron insertion, with sequence losses and gains at the 5´ intron-exon junction [43–45]. Our observations that the full reverse splicing reaction of linear RmInt1 RNA occurs in vitro, even though is low and inefficient, open the possibility that some group II introns which excises as linear molecules in vivo might spread and impact the evolution of the host bacterial genome.
Materials and methods
In vitro transcription for linear RmInt1 RNA preparations
A 740 nt RmInt1-ΔORF linear intron was obtained by run-off transcription with the T7 RNA polymerase from a PCR product containing the T7 promoter at the 5´ end. The template for in vitro transcription was amplified from pKGEMA4 [69] with the Phusion DNA polymerase (Thermo Scientific) and the sense primer LMS94 and the antisense primer Oligo1, an oligonucleotide 2´-OMe-modified at the last two nucleotides at the 5´ end to prevent the addition of untemplated nucleotides to the 3´end of the transcripts [70,71] (see Table S1). We performed DNA cleavage assays on 5´-labeled 70-mer ssDNA substrates, with RNPs reconstituted from homogeneous intron transcripts with identical 3´ ends (Li 3´Homo) or with additional nucleotides at the 3´-end (Li 3´Hetero) (Figure S4), obtaining similar levels of DNA cleavage activity for both types of reconstituted RNPs, although reverse splicing levels were slightly higher when homogeneous intron RNA molecules with identical 3´-termini were used to prepare the RmInt1 RNPs [70,71]. The RmInt1 intron RNA precursor was synthesized as described by Molina-Sánchez et al. (2016) [34]. For modifications introduced into the 5´end of the RNA, the Oligo1 primer was used as the 3´ primer; conversely, if the RNA was modified at its 3´end, the corresponding primer in 3´ was LMS94. Transcripts were generated with 15 µg of the PCR product and 70 units of T7 RNA polymerase in 1 ml reaction buffer containing 1× transcription buffer [10× transcription buffer: 150 mM MgCl2, 400 mM Tris–HCl, pH 7.5, 20 mM spermidine, and 50 mM DTT], 0.96 mM NTPs, 10 mM DTT and 400 units of RNaseOUT (Invitrogen). RNA products were precipitated after 5 h of incubation at 37°C, treated with DNase I and gel-purified before use [30,34]. RNA molecules of the correct size were excised from the gel and eluted by simple diffusion during overnight incubation in elution buffer (0.3 M NaCl, 10 mM MOPS pH 6, 1 mM EDTA). RNA was precipitated, dried and stored in 10 mM MOPS pH 6 and 1 mM EDTA.
RmInt1 IEP expression and purification
The MBP-IEP fusion protein was produced and purified essentially as described by Molina-Sánchez et al. (2016) [34]. Freshly transformed E. coli Rosetta-gami (DE3):pLysS cells harboring pMALFlagIEP were cultured in 50 ml of LB medium supplemented with 0.2% glucose, 100 µg·ml−1 ampicillin and 50 µg·ml−1 chloramphenicol for 4 h at 37°C with vigorous shaking, until the OD600 reached 0.5 units. They were then induced with 0.3 mM IPTG at 20°C for 16–22 h. Cells were lysed by freeze-thaw cycles followed by sonication. Cleared lysates were passed through chromatography columns (BioRad) containing amylose beads (New England Biolabs). MBP-IEP was eluted by adding 10 mM maltose (Sigma-Aldrich), dialyzed and concentrated with YM-30 centrifugal filters (Amicron Ultra, Millipore). Protein concentrations were determined by the Bradford method, with the Bio-Rad protein assay reagent and BSA as the standard (Bio-Rad). Protein preparations were analyzed by electrophoresis in Coomassie blue-stained 0.1% SDS-10% polyacrylamide gels and by immunoblot analysis with antibodies against the Flag epitope (Sigma-Aldrich).
Reconstitution of RNP particles in vitro
RmInt1-ΔORF RNP particles were reconstituted from in vitro-synthesized intron RNA and purified MBP-IEP, as described by Molina-Sánchez et al. (2016) [34]. Linear intron RNA (2.5 µM) in 40 mM Tris-HCl pH 7.5 was denatured by heating at 90°C for 1 minute and then renatured by successive incubations at 50°C for 1 minute and slow cooling at room temperature, before the addition of 500 mM NH4Cl, 5 mM MgCl2 and 5 µM purified MBP-IEP protein. The RNA/protein mixture was transferred to a water bath at 30°C for 1 h, and the reconstituted RNPs were then kept on ice until use.
Substrate labeling and DNA cleavage assays
DNA cleavage assays were carried out principally on a 70-mer single-stranded DNA oligonucleotide substrate containing the intron insertion site at position 35 (see Table S2). The substrate was labelled at its 5´ end with [γ-32P]ATP (6,000 Ci/mmol; Perkin-Elmer) and T4 polynucleotide kinase (New England Biolabs), or at its 3´ end with terminal transferase (New England Biolabs) and cordycepin 5´-triphosphate [α-32P] 3´-deoxyadenosine (6,000 Ci/mmol; Perkin Elmer). The sequences of the other mutated 70-mer single-stranded DNA substrates used for the experiment described in Figure 4 are included in Table S2.
Reconstituted RNP particles (2.5 µl; ~900 nM) were incubated with the substrate (300,000 cpm) at 37°C for 2 h in reaction buffer containing 50 mM Tris-HCl pH 7.5, 10 mM KCl, 25 mM MgCl2 and 5 mM DTT. For time-course experiments, samples were taken at the indicated times and placed on ice. The reaction mixtures were cleaned up by extraction with phenol-chloroform-isoamyl alcohol (25:24:1) followed by ethanol precipitation. Products were analyzed by electrophoresis in denaturing 7 M urea-6% (w/v) polyacrylamide gels and quantified with Quantity One software (Bio-Rad).
Identification of low-migrating products
Low-migrating products were gel-purified and eluted by simple diffusion overnight in elution buffer (0.3 M NaCl, 10 mM MOPS pH 6, 1 mM EDTA). The DNA/RNA hybrid molecules were precipitated, dried and stored in RNase-free water. cDNA was synthesized with 2 µg N6 random oligonucleotides or 10 pmol S70ds/DN (complementary to the 3´ end of the substrate) and 200 units SuperScript II reverse transcriptase (Invitrogen) in the presence of 1.5 mM dNTPs, 15 units RNase OUT (Invitrogen), 50 mM Tris-HCl pH 8.3, 75 mM KCl, 2 mM MgCl2 and 20 mM DTT for 2 h at 42°C. Integration junctions were identified by PCR with 10 pmol of each primer and 2 units Taq polymerase. For the 5´ exon-intron junction, S70ds/UP (complementary to the 5´ end of the substrate) and R1 (complementary to positions 348–365 of intron domain I) were used, whereas for the intron-3´ exon junction, H3 (located at position 520–539 of the intron domain IV) was used with oligonucleotide +9 (binding between positions 9 and 28 in the 3´ exon). The PCR conditions were as follows: denaturation at 94°C for 3 min, followed by 25 cycles of 20 s at 94°C, 20 s at 60°C, 30 s at 72°C, and a final extension at 72°C for 10 min. The full reverse splicing product was amplified with S70ds/UP and S70ds/DN, by increasing the extension time to 60 s and the number of cycles to 35. The RT-PCR products were visualized on agarose gels and inserted into the pGEM-T easy vector (Promega) for the sequencing of individual molecules. The primer sequences are indicated in Table S3.
Funding Statement
This work was supported by the Ministerio de Ciencia e Innovación [BIO2017-8244-P]; Ministerio de Ciencia e Innovación [BIO2014-51953-P].
Acknowledgments
We thank Jose María del Arco and Ascensión Martos for technical assistance. This work was supported by research grants [BIO2014-51953-P] and [BIO2017-8244-P] from the Plan Nacional de I+D+i, biotechnology program of the Spanish Ministerio de Ciencia, Innovación y Universidades including European Regional Development Funds (ERDF).
Disclosure statement
The authors have no competing interests to declare.
Supplementary material
Supplemental data for this article can be accessed here.
References
- [1].Keating KS, Toor N, Perlman PS, et al. A structural analysis of the group II intron active site and implications for the spliceosome. RNA. 2010;16(1):1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [2].Lambowitz AM, Belfort M.. Mobile bacterial group II introns at the crux of eukaryotic evolution. Microbiol Spectr. 2015;3(1):MDNA3-0050-2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [3].Peebles CL, Perlman PS, Mecklenburg KL, et al. A self-splicing RNA excises an intron lariat. Cell. 1986;44(2):213–223. [DOI] [PubMed] [Google Scholar]
- [4].Jarrell KA, Peebles CL, Dietrich RC, et al. Group II intron self-splicing. Alternative reaction conditions yield novel products. J Biol Chem. 1988;263(7):3432–3439. [PubMed] [Google Scholar]
- [5].van der Veen R, Kwakman JH, Grivell LA. Mutations at the lariat acceptor site allow self-splicing of a group II intron without lariat formation. Embo J. 1987;6(12):3827–3831. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [6].Vogel J, Börner T. Lariat formation and a hydrolytic pathway in plant chloroplast group II intron splicing. Embo J. 2002;21(14):3794–3803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Li CF, Costa M, Michel F. Linking the branchpoint helix to a newly found receptor allows lariat formation by a group II intron. Embo J. 2011;30(15):3040–3051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Li CF, Costa M, Bassi G, et al. Recurrent insertion of 5′-terminal nucleotides and loss of the branchpoint motif in lineages of group II introns inserted in mitochondrial preribosomal RNAs. RNA. 2011;17(7):1321–1335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [9].Murray HL, Mikheeva S, Coljee VW, et al. Excision of group II introns as circles. Mol Cell. 2001;8(1):201–211. [DOI] [PubMed] [Google Scholar]
- [10].Molina-Sánchez MD, Martinez-Abarca F, Toro N. Excision of the Sinorhizobium meliloti group II intron RmInt1 as circles in vivo. J Biol Chem. 2006;281(39):28737–28744. [DOI] [PubMed] [Google Scholar]
- [11].Li-Pook-Than J, Bonen L. Multiple physical forms of excised group II intron RNAs in wheat mitochondria. Nucleic Acids Res. 2006;34(9):2782–2790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [12].Monat C, Quiroga C, Laroche-Johnston F, et al. The Ll.LtrB intron from Lactococcus lactis excises as circles in vivo: insights into the group II intron circularization pathway. RNA. 2015;21(7):1286–1293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Monat C, Cousineau B. Circularization pathway of a bacterial group II intron. Nucleic Acids Res. 2016;44(4):1845–1853. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [14].Augustin S, Müller MW, Schweyen RJ. Reverse self-splicing of group II intron RNAs in vitro. Nature. 1990;343(6256):383–386. [DOI] [PubMed] [Google Scholar]
- [15].Yang J, Zimmerly S, Perlman PS, et al. Efficient integration of an intron RNA into double-stranded DNA by reverse splicing. Nature. 1996;381(6580):332–335. [DOI] [PubMed] [Google Scholar]
- [16].Cousineau B, Smith D, Lawrence-Cavanagh S, et al. Retrohoming of a bacterial group II intron: mobility via complete reverse splicing, independent of homologous DNA recombination. Cell. 1998;94(4):451–462. [DOI] [PubMed] [Google Scholar]
- [17].Dickson L, Huang HR, Liu L, et al. Retrotransposition of a yeast group II intron occurs by reverse splicing directly into ectopic DNA sites. Proc Natl Acad Sci USA. 2001;98(23):13207–13212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [18].Eskes R, Yang J, Lambowitz AM, et al. Mobility of yeast mitochondrial group II introns: engineering a new site specificity and retrohoming via full reverse splicing. Cell. 1997;88(6):865–874. [DOI] [PubMed] [Google Scholar]
- [19].Ichiyanagi K, Beauregard A, Lawrence S, et al. Retrotransposition of the Ll.LtrB group II intron proceeds predominantly via reverse splicing into DNA targets. Mol Microbiol. 2002;46(5):1259–1272. [DOI] [PubMed] [Google Scholar]
- [20].Lambowitz AM, Zimmerly S. Group II introns: mobile ribozymes that invade DNA. Cold Spring Harb Perspect Biol. 2011;3(8):a003616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [21].Smith D, Zhong J, Matsuura M, et al. Recruitment of host functions suggests a repair pathway for late steps in group II intron retrohoming. Genes Dev. 2005;19(20):2477–2487. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [22].Coros CJ, Piazza CL, Chalamcharla VR, et al. A mutant screen reveals RNase E as a silencer of group II intron retromobility in Escherichia coli. RNA. 2008;14(12):2634–2644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [23].Coros CJ, Piazza CL, Chalamcharla VR, et al. Global regulators orchestrate group II intron retromobility. Mol Cell. 2009;34(2):250–256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [24].Yao J, Truong DM, Lambowitz AM. Genetic and biochemical assays reveal a key role for replication restart proteins in group II intron retrohoming. PLoS Genet. 2013;9(4):e1003469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Nisa-Martínez R, Molina-Sánchez MD, Toro N. Host factors influencing the retrohoming pathway of group II intron RmInt1, which has an intron-encoded protein naturally devoid of endonuclease activity. PLoS One. 2016;11(9):e0162275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [26].Mohr S, Matsuura M, Perlman PS, et al. A DEAD-box protein alone promotes group II intron splicing and reverse splicing by acting as an RNA chaperone. Proc Natl Acad Sci USA. 2006;103(10):3569–3574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Gordon PM, Fong R, Piccirilli JA. A second divalent metal ion in the group II intron reaction center. Chem Biol. 2007;14(6):607–612. [DOI] [PubMed] [Google Scholar]
- [28].Martínez-Abarca F, García-Rodríguez FM, Toro N. Homing of a bacterial group II intron with an intron-encoded protein lacking a recognizable endonuclease domain. Mol Microbiol. 2000;35(6):1405–1412. [DOI] [PubMed] [Google Scholar]
- [29].Costa M, Michel F, Molina-Sánchez MD, et al. An alternative intron-exon pairing scheme implied by unexpected in vitro activities of group II intron RmInt1 from Sinorhizobium meliloti. Biochimie. 2006;88(6):711–717. [DOI] [PubMed] [Google Scholar]
- [30].Chillón I, Molina-Sánchez MD, Fedorova O, et al. In vitro characterization of the splicing efficiency and fidelity of the RmInt1 group II intron as a means of controlling the dispersion of its host mobile element. RNA. 2014;20(12):2000–2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [31].Muñoz-Adelantado E, San Filippo J, Martínez-Abarca F, et al. Mobility of the Sinorhizobium meliloti group II intron RmInt1 occurs by reverse splicing into DNA, but requires an unknown reverse transcriptase priming mechanism. J Mol Biol. 2003;327(5):931–943. [DOI] [PubMed] [Google Scholar]
- [32].Molina-Sánchez MD, Martínez-Abarca F, Toro N. Structural features in the C-terminal region of the Sinorhizobium meliloti RmInt1 group II intron-encoded protein contribute to its maturase and intron DNA-insertion function. FEBS J. 2010;277(1):244–254. [DOI] [PubMed] [Google Scholar]
- [33].Martínez-Abarca F, Barrientos-Durán A, Fernández-López M, et al. The RmInt1 group II intron has two different retrohoming pathways for mobility using predominantly the nascent lagging strand at DNA replication forks for priming. Nucleic Acids Res. 2004;32(9):2880–2888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [34].Molina-Sánchez MD, García-Rodríguez FM, Toro N. Functionality of in vitro reconstituted group II intron RmInt1-derived ribonucleoprotein particles. Front Mol Biosci. 2016;3:58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Peebles CL, Benatan EJ, Jarrell KA, et al. Group II intron self-splicing: development of alternative reaction conditions and identification of a predicted intermediate. Cold Spring Harb Symp Quant Biol. 1987;52:223–232. [DOI] [PubMed] [Google Scholar]
- [36].Jiménez-Zurdo JI, García-Rodríguez FM, Barrientos-Durán A, et al. DNA target site requirements for homing in vivo of a bacterial group II intron encoding a protein lacking the DNA endonuclease domain. J Mol Biol. 2003;326(2):413–423. [DOI] [PubMed] [Google Scholar]
- [37].Martínez-Rodríguez L, García-Rodríguez FM, Molina-Sánchez MD, et al. Insights into the strategies used by related group II introns to adapt successfully for the colonisation of a bacterial genome. RNA Biol. 2014;11(8):1061–1071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Matsuura M, Saldanha R, Ma H, et al. A bacterial group II intron encoding reverse transcriptase, maturase, and DNA endonuclease activities: biochemical demonstration of maturase activity and insertion of new genetic information within the intron. Genes Dev. 1997;11(21):2910–2924. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Saldanha R, Chen B, Wank H, et al. RNA and protein catalysis in group II intron splicing and mobility reactions using purified components. Biochemistry. 1999;38(28):9069–9083. [DOI] [PubMed] [Google Scholar]
- [40].Singh NN, Lambowitz AM. Interaction of a group II intron ribonucleoprotein endonuclease with its DNA target site investigated by DNA footprinting and modification interference. J Mol Biol. 2001;309(2):361–386. [DOI] [PubMed] [Google Scholar]
- [41].Aizawa Y, Xiang Q, Lambowitz AM, et al. The pathway for DNA recognition and RNA integration by a group II intron retrotransposon. Mol Cell. 2003;11(3):795–805. [DOI] [PubMed] [Google Scholar]
- [42].Noah JW, Park S, Whitt JT, et al. Atomic force microscopy reveals DNA bending during group II intron ribonucleoprotein particle integration into double-stranded DNA. Biochemistry. 2006;45(41):12424–12435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [43].Mastroianni M, Watanabe K, White TB, et al. Group II intron-based gene targeting reactions in eukaryotes. PLoS One. 2008;3(9):e3121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [44].Zhuang F, Mastroianni M, White TB, et al. Linear group II intron RNAs can retrohome in eukaryotes and may use nonhomologous end-joining for cDNA ligation. Proc Natl Acad Sci USA. 2009;106(43):18189–18194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].White TB, Lambowitz AM. The retrohoming of linear group II intron RNAs in Drosophila melanogaster occurs by both DNA ligase 4-dependent and -independent mechanisms. PLoS Genet. 2012;8(2):e1002534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [46].Mörl M, Schmelzer C. Integration of group II intron bI1 into a foreign RNA by reversal of the self-splicing reaction in vitro. Cell. 1990;60(4):629–636. [DOI] [PubMed] [Google Scholar]
- [47].Mörl M, Niemer I, Schmelzer C. New reactions catalyzed by a group II intron ribozyme with RNA and DNA substrates. Cell. 1992;70(5):803–810. [DOI] [PubMed] [Google Scholar]
- [48].Roitzsch M, Pyle AM. The linear form of a group II intron catalyzes efficient autocatalytic reverse splicing, establishing a potential for mobility. RNA. 2009;15(3):473–482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Chanfreau G, Jacquier A. An RNA conformational change between the two chemical steps of group II self-splicing. Embo J. 1996;15(13):3466–3476. [PMC free article] [PubMed] [Google Scholar]
- [50].Nolte A, Chanfreau G, Jacquier A. Influence of substrate structure on in vitro ribozyme activity of a group II intron. RNA. 1998;4(6):694–708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [51].Chin K, Pyle AM. Branch-point attack in group II introns is a highly reversible transesterification, providing a potential proofreading mechanism for 5‘-splice site selection. RNA. 1995;1(4):391–406. [PMC free article] [PubMed] [Google Scholar]
- [52].Costa M, Walbott H, Monachello D, et al. Crystal structures of a group II intron lariat primed for reverse splicing. Science. 2016;354(6316):aaf9258. [DOI] [PubMed] [Google Scholar]
- [53].Marcia M, Pyle AM. Visualizing group II intron catalysis through the stages of splicing. Cell. 2012;151(3):497–507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [54].Dème E, Nolte A, Jacquier A. Unexpected metal ion requirements specific for catalysis of the branching reaction in a group II intron. Biochemistry. 1999;38(10):3157–3167. [DOI] [PubMed] [Google Scholar]
- [55].Jacquier A, Jacquesson-Breuleux N. Splice site selection and role of the lariat in a group II intron. J Mol Biol. 1991;219(3):415–428. [DOI] [PubMed] [Google Scholar]
- [56].Su LJ, Qin PZ, Michels WJ, et al. Guiding ribozyme cleavage through motif recognition: the mechanism of cleavage site selection by a group II intron ribozyme. J Mol Biol. 2001;306(4):655–668. [DOI] [PubMed] [Google Scholar]
- [57].García-Rodríguez FM, Hernández-Gutiérrez T, Díaz-Prado V, et al. Use of the computer-retargeted group II intron RmInt1 of Sinorhizobium meliloti for gene targeting. RNA Biol. 2014;11(4):391–401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [58].Qu G, Kaushal PS, Wang J, et al. Structure of a group II intron in complex with its reverse transcriptase. Nat Struct Mol Biol. 2016;23(6):549–557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Michel F, Umesono K, Ozeki H. Comparative and functional anatomy of group II catalytic introns–a review. Gene. 1989;82(1):5–30. [DOI] [PubMed] [Google Scholar]
- [60].Chanfreau G, Jacquier A. Interaction of intronic boundaries is required for the second splicing step efficiency of a group II intron. Embo J. 1993;12(13):5173–5180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [61].Peebles CL, Belcher SM, Zhang M, et al. Mutation of the conserved first nucleotide of a group II intron from yeast mitochondrial DNA reduces the rate but allows accurate splicing. J Biol Chem. 1993;268(16):11929–11938. [PubMed] [Google Scholar]
- [62].Jacquier A, Michel F. Base-pairing interactions involving the 5‘ and 3‘-terminal nucleotides of group II self-splicing introns. J Mol Biol. 1990;213(3):437–447. [DOI] [PubMed] [Google Scholar]
- [63].Mikheeva S, Murray HL, Zhou H, et al. Deletion of a conserved dinucleotide inhibits the second step of group II intron splicing. RNA. 2000;6(11):1509–1515. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Chillón I, Martínez-Abarca F, Toro N. Splicing of the Sinorhizobium meliloti RmInt1 group II intron provides evidence of retroelement behavior. Nucleic Acids Res. 2011;39(3):1095–1104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [65].Marcia M, Somarowthu S, Pyle AM. Now on display: a gallery of group II intron structures at different stages of catalysis. Mob DNA. 2013;4(1):14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Nagy V, Pirakitikulr N, Zhou KI, et al. Predicted group II intron lineages E and F comprise catalytically active ribozymes. RNA. 2013;19(9):1266–1278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Granlund M, Michel F, Norgren M. Mutually exclusive distribution of IS1548 and GBSi1, an active group II intron identified in human isolates of group B streptococci. J Bacteriol. 2001;183(8):2560–2569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [68].Toor N, Robart AR, Christianson J, et al. Self-splicing of a group IIC intron: 5‘ exon recognition and alternative 5‘ splicing events implicate the stem-loop motif of a transcriptional terminator. Nucleic Acids Res. 2006;34(22):6461–6471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [69].Nisa-Martínez R, Jiménez-Zurdo JI, Martínez-Abarca F, et al. Dispersion of the RmInt1 group II intron in the Sinorhizobium meliloti genome upon acquisition by conjugative transfer. Nucleic Acids Res. 2007;35(1):214–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Kao C, Zheng M, Rüdisser S. A simple and efficient method to reduce nontemplated nucleotide addition at the 3 terminus of RNAs transcribed by T7 RNA polymerase. RNA. 1999;5(9):1268–1272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Kao C, Rüdisser S, Zheng M. A simple and efficient method to transcribe RNAs with reduced 3‘ heterogeneity. Methods. 2001;23(3):201–205. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
