Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Protein Trans-Splicing of an Atypical Split Intein Showing Structural Flexibility and Cross-Reactivity

  • Huiling Song,

    Affiliations Institute of Biological Sciences and Biotechnology, Donghua University, Shanghai, People’s Republic of China, Department of Biochemistry and Molecular Biology, Dalhousie University, Halifax, Nova Scotia, Canada

  • Qing Meng ,

    mengqing@dhu.edu.cn (QM); paul.liu@dal.ca (XQL)

    Affiliation Institute of Biological Sciences and Biotechnology, Donghua University, Shanghai, People’s Republic of China

  • Xiang-Qin Liu

    mengqing@dhu.edu.cn (QM); paul.liu@dal.ca (XQL)

    Affiliation Department of Biochemistry and Molecular Biology, Dalhousie University, Halifax, Nova Scotia, Canada

Abstract

Inteins catalyze a protein splicing reaction to excise the intein from a precursor protein and join the flanking sequences (exteins) with a peptide bond. In a split intein, the intein fragments (IN and IC) can reassemble non-covalently to catalyze a trans-splicing reaction that joins the exteins from separate polypeptides. An atypical split intein having a very small IN and a large IC is particularly useful for joining synthetic peptides with recombinant proteins, which can be a generally useful method of introducing site-specific chemical labeling or modifications into proteins. However, a large IC derived from an Ssp DnaX intein was found recently to undergo spontaneous C-cleavage, which raised questions regarding its structure-function and ability to trans-splice. Here, we show that this IC could undergo trans-splicing in the presence of IN, and the trans-splicing activity completely suppressed the C-cleavage activity. We also found that this IC could trans-splice with small IN sequences derived from two other inteins, showing a cross-reactivity of this atypical split intein. Furthermore, we found that this IC could trans-splice even when the IN sequence was embedded in a nearly complete intein sequence, suggesting that the small IN could project out of the central pocket of the intein to become accessible to the IC. Overall, these findings uncovered a new atypical split intein that can be valuable for peptide-protein trans-splicing, and they also revealed an interesting structural flexibility and cross-reactivity at the active site of this intein.

Introduction

Inteins are internal protein sequences that can catalyze a protein splicing reaction to excise themselves from a precursor protein and at the same time join the flanking sequences (N- and C-exteins) with a peptide bond [1], [2]. The catalytic mechanism of protein splicing typically consists of four steps [2], [3]: 1) an N-S or N-O acyl rearrangement at the upstream splicing junction replaces the peptide bond with an ester bond; 2) a transesterification reaction transfers the N-extein from the upstream splice junction to the downstream splice junction; 3) an asparagine cyclization at the C-terminus of the intein breaks the peptide bond, which separates the intein from the exteins; and 4) a S-N or O-N acyl rearrangement forms a peptide bond joining the N- and C-exteins. Inteins can sometimes catalyze protein cleavage at their N- or C-termini when the splicing mechanism is disrupted [3]. For example, when step 1 of the splicing mechanism is blocked or missing, step 3 may still occur and break the peptide bond at the C-terminus of the intein, in what is termed C-cleavage. Whereas different inteins show low levels of similarity at the amino acid sequence level, the crystal structures of different inteins (the protein splicing domain) appear very similar [4], [5], [6], [7]. The splicing domains of inteins are typically shaped like a flattened disk consisting of ∼12 β-strands, with the N- and C-terminal parts of the intein folded into a centrally located catalytic pocket.

In a split intein, the intein sequence is broken into two separate fragments, and these intein fragments (IN and IC) can reassemble non-covalently to catalyze a protein trans-splicing reaction [8], [9]. This reaction takes place between two separate polypeptides, with one polypeptide consisting of an N-extein fused to the IN and the other polypeptide consisting of the IC fused to a C-extein. During trans-splicing, the IN and IC are excised, while the N- and C-exteins are concomitantly joined with a peptide bond. Protein-protein trans-splicing using split inteins has proven to be a useful technology with a wide range of applications. Examples include production of cytotoxic proteins that cannot be expressed in a single piece [10], [11], segmental isotope labeling of proteins for NMR studies [12], and a gene therapy procedure using split genes [13]. For peptide-protein trans-splicing, a non-canonical split intein (Ssp DnaB S1) consisting of a small IN and a large IC has been engineered [14]. This atypical split intein is particularly suitable for splicing synthetic peptides onto the N-terminus of recombinant proteins, because the extremely small IN (11 aa long) can be more readily produced together with a small N-extein through chemical synthesis. Because the chemically synthesized N-extein can potentially contain any desired chemical group, this atypical split intein has been used successfully in adding fluorescent labels to the N-terminus of recombinant proteins in vitro [15] and on the surface of live mammalian cells [16].

Further development of the intein-based method of protein site-specific modification or labeling is of great interest for protein research and engineering. For example, fluorescent or isotope labels can be useful for studying cellular location and trafficking of proteins, and chemical modifications (e.g. unnatural amino acids) can aid studies of protein's structure-function relationship. Standard chemical methods often produce mixed populations of the modified protein, because such methods usually target certain amino acid side chains (thiols, carboxyls, amines) that may exist at multiple locations in the protein [17]. Other methods have been developed for site-specific protein modifications with limited success. Certain polypeptide tags have been used in recombinant proteins to attach a chemical group through an enzymatic reaction (e.g. [18]), but the tag remains incorporated in the modified protein and may interfere with protein function. Specially engineered tRNA charging systems have been used to add unnatural amino acids to proteins during translation [19], but it is difficult or impossible to engineer special tRNA charging systems for every desired unnatural amino acid or chemical modification. Intein-based protein-peptide trans-splicing is a newer and potentially more useful method for site-specific protein modifications [20], [21], [22], because it does not leave a large tag in the modified protein and may be used generally with any chemical moieties on the chemically synthesized extein peptide. To further develop this intein-based method, it is important to find new and atypical split inteins for the peptide-protein trans-splicing, because different inteins may exhibit different splicing efficiencies with different exteins [23], [24], [25], [26]. In a recent such study, however, a large C-intein (IC) derived from the Ssp DnaX intein (a natural intein in DnaX protein of Synechocystis sp. PCC6803) was found to undergo spontaneous C-cleavage [27], which is unlike the similarly constructed IC from the Ssp DnaB intein (a natural intein in DnaB protein of Synechocystis sp. PCC6803). This C-cleavage activity was unexpected, because the large IC without IN was thought to have a structural hole in the catalytic pocket of the intein, based on the predicted structure of the intein. This surprise finding raised interesting questions regarding the structure and function of this intein, namely whether it can still catalyze protein trans-splicing in the presence of the IN and, if so, whether its catalytic pocket has unusual structural flexibilities that are not apparent from intein crystal structures.

In this study, we initially found that the unusual IC of the Ssp DnaX intein could undergo protein trans-splicing with the small IN on another protein, and this trans-splicing activity completely predominated over the C-cleavage activity. The IC could also trans-splice with small IN sequences derived from other different inteins, which revealed for the first time a cross-reactivity of atypical split inteins. We also found that the small IN could be replaced functionally by a nearly complete intein containing the IN. These findings not only generated a new and second atypical split intein suitable for trans-splicing peptides onto the N-terminus of proteins, they also have interesting implications for the structure-function of this atypical intein and perhaps also other inteins. We suggest that the centrally located catalytic pocket of the intein might undergo reversible transitions between an open state for the trans-splicing function and a closed state for the C-cleavage function, and this structural flexibility might permit the IN part of the intein to swing out of the central pocket of the intein.

Results

First we determined whether the 139-aa C-intein (IC) of the Ssp DnaX intein could undergo protein trans-splicing when the missing 11-aa N-terminal part (N-intein or IN) was provided in trans, because the IC alone had been found to undergo spontaneous C-cleavage [27]. As illustrated in Figure 1A, a maltose binding protein (M) and a thioredoxin protein (T) were used as the N-extein and C-extein, respectively, so that a trans-splicing reaction would join these two exteins to form the splicing product MT. As seen in Figure 1B, the splicing product MT was produced both in vivo when the two precursor proteins (MIN and ICT) were co-expressed in E. coli and in vitro when the purified precursor proteins were incubated together in a test tube.

thumbnail
Figure 1. Detection of trans-splicing and C-cleavage activities.

A. Schematic illustration of the C-cleavage and trans-splicing reactions. Precursor protein ICT is a fusion protein consisting of the 139-aa C-intein (IC) of the Ssp DnaX split intein fused to a thioredoxin protein (T). Precursor protein MIN is a fusion protein consisting of a maltose binding protein (M) and the 11-aa N-intein (IN) of the Ssp DnaX split intein. B. Experimental analysis of the reactions. For analysis in vivo, MIN and ICT proteins were co-expressed in E. coli cells. Total cellular proteins before (NI) and after (I) the IPTG-induced expression were resolved by SDS-PAGE, and protein bands were visualized either by staining (Coomassie stained) or by Western blotting using an anti-thioredoxin (Anti-T) antibody. For analysis in vitro, the MIN and ICT proteins were separately produced and purified. These two proteins were then mixed and incubated at room temperature for 20 hours. Reaction products were analyzed and visualized by staining or Western blotting as above. Positions are indicated for the precursor proteins (MIN and ICT), the splicing product (MT), and the C-cleavage products (IC and T). Size markers (Marker) are shown on the left. C. Effects of reaction times and temperatures. The in vitro reactions were carried out for the specified length of times and at the specified temperatures. Reaction products were analyzed by Western blotting as above.

https://doi.org/10.1371/journal.pone.0045355.g001

Interestingly, no C-cleavage activity was detected under these conditions, as indicated by an absence of the cleavage product T. The precursor and product proteins were identified by their predicted sizes and specific recognition of an anti-thioredoxin (anti-T) antibody through Western blotting. For the in vivo analysis in E. coli cells, only the splicing product MT protein was detected using anti-T antibody, indicating that the precursor protein ICT had trans-spliced completely to form the MT protein. For the in vitro analysis, the purified precursor protein MIN was added in excess to the precursor protein ICT to drive the trans-splicing reaction to greater completion. In producing the ICT protein alone in E. coli, a significant amount of spontaneous C-cleavage occurred as reported previously [27]. In the purification of ICT using an affinity tag (hexahistidine) contained in the IC, the cleavage product IC was co-purified with the remaining ICT in the purified sample, while the cleavage product T lacked the hexahistidine tag and was absent in the purified sample. The purified ICT protein did not show new C-cleavage during subsequent co-incubation with the MIN protein for trans-splicing, as indicated by the absence of any new formation of the C-cleavage products IC and T. Under the in vitro conditions used, approximately 85% of the ICT protein was trans-spliced to form the MT protein after 20 hours of incubation at room temperature (Figure 1B). We also tested shorter reaction times and different temperatures (Figure 1C). The efficiency of trans-splicing was nearly identical at four tested temperatures (4, 25, and 37°C, and on ice) after 24 hours of reaction. With a shorter reaction time of 15 minutes, the efficiency of trans-splicing was a little lower at 4°C and significantly lower on ice.

We then investigated whether the trans-splicing reaction could still occur when the small IN is embedded in a near complete intein. As illustrated in Figure 2A, the intein fragment INL was designed to contain the N-terminal 144-aa sequence of the 150-aa Ssp DnaX intein that lacked the C-terminal 6 aa of the intein, to prevent possible self-cleavage or cis-splicing. This INL was found to trans-splice efficiently with IC in vitro at three different temperatures (4, 25, and 37°C), where M and T were the exteins (Figure 2B). Because the 144-aa INL consists of the small (11-aa) IN plus other parts of the intein, we asked whether the other parts of the intein also participated in the trans-splicing reaction. To answer this question, a double mutation (TXXH to AXXA) was introduced in the Block B motif of the intein, because this conserved intein sequence motif is outside the 11-aa IN and known to be functionally important in inteins [25]. As seen in Figure 2C, mutating the Block B motif of IC (resulting in ICm) destroyed its ability to trans-splice with IN, as expected. In contrast, mutating the Block B motif of INL (resulting in INLm) did not affect its ability to trans-splice with IC, indicating that the Block B motif in INL did not participate in the reaction. We also tested mutated C-intein (ICm) in a combination with the non-mutated version of INL, and found that latter could compensate for the mutated Block B motif in the former for the trans-splicing reaction. When the Block B motif was mutated in both INL (resulting in INLm) and IC (resulting in ICm), the trans-splicing reaction was abolished, as expected.

thumbnail
Figure 2. Effects of IN size and Block B mutation on trans-splicing.

A. Schematic illustration of experimental designs. INL is the N-terminal 144-aa sequence of the Ssp DnaX intein. INLm and ICm are the same as INL and IC, respectively, except that the conserved Block B sequence of the intein was mutated from TXXH to AXXA. Others are the same as in Figure 1A. B. Trans-splicing of ICT with MINL. A mixture of the two precursor proteins was incubated at the specified temperatures for 20 hours to allow reaction, with the protein bands visualized by Western blotting using anti-T antibody. C. Trans-splicing of different precursor protein pairs listed in panel A. Each pair of precursor proteins was incubated together at room temperature for 20 hours, with the protein bands visualized by Western blotting using anti-T antibody. Names of only intein parts of the precursor proteins are specified on the top of the panel.

https://doi.org/10.1371/journal.pone.0045355.g002

To further explore the structural flexibility and versatility of this atypical split intein, we asked whether the large IC could trans-splice with small IN sequences derived from other inteins. The 12-aa INRB was derived from the N-terminus of the Rma DnaB intein (a natural intein in DnaB protein of Rhodothermus marinus) [28] that is highly similar to the Ssp DnaB intein from which the first atypical split intein was derived [13]. The 12-aa INSG was derived from the N-terminus of the Ssp GyrB intein (a natural intein in GyrB protein of Synechocystis sp. PCC6803) [29]. As shown in Figure 3A, the INRB sequence is 41% identical (58% similar) to the IN sequence, and the INSG sequence is 50% identical (75% similar) to the IN sequence. Under in vivo conditions in E. coli, both INRB and INSG trans-spliced efficiently with IC, as indicated by the accumulation of splicing product MT but not precursor protein ICT (Figure 3B). Under in vitro conditions using purified precursor proteins, INRB trans-spliced with IC, but INSG did not. For INRB, the in vitro trans-splicing reaction did not go to completion, with ∼60% of the precursor protein ICT remaining. This was not due to a lesser amount of INRB, because Coomassie-stained gel pictures showed an excess amount of the precursor protein MINRB in the reaction (Figure 3B). This may indicate an inefficient use of INRB under the in vitro conditions used, although an efficient use of INRB was seen in E. coli cells. With both INRB and INSG, the precursor protein ICT underwent a small amount of C-cleavage, as indicated by the accumulation of a small amount of the C-cleavage product T.

thumbnail
Figure 3. Cross-reactivity of IC with IN from different inteins.

A. Comparison of the amino acid sequences of IN from Ssp DnaX intein (IN), Rma DnaB intein (INRB), and Ssp GyrB intein (INSG). Identical and similar residues are marked with a | and a :, respectively. A gap (represented with a -) is introduced in the IN sequence to maximize the sequence alignment. B. Analysis of trans-splicing reactions between IC and INRB (or INSG) as specified on top. For in vivo analysis, MINRB (or MINSG) and ICT proteins were co-expressed in E. coli cells, and total cellular proteins were analyzed by Western blotting using an anti-thioredoxin (Anti-T) antibody. For in vitro analysis, purified MINRB (or MINSG) and ICT proteins were co-incubated at room temperature for 20 hours, and the reaction products were analyzed by Western blotting as above. Positions are marked for the precursor ICT, splicing product MT, and C-cleavage product T. Size markers are shown on the left.

https://doi.org/10.1371/journal.pone.0045355.g003

Discussion

The atypical Ssp DnaX intein is found, for the first time, to be capable of protein trans-splicing, despite the fact that its large C-intein (IC) had been known to undergo spontaneous C-cleavage. This finding has interesting implications on the structure-function of inteins’ active site. Previously the IC part of this intein was found to undergo spontaneous C-cleavage in the absence of IN [27], which was quite unexpected and unlike other inteins. The highly conserved crystal structures of inteins predict that the N- and C-terminal parts of an intein are located in a central catalytic pocket [5], [6], [7], as illustrated by a computer modeling of the Ssp DnaX intein shown in Figure 4A. The 11-aa IN sequence forms two small β-strands named β1 and β2, with β1 being buried deep inside the intein structure. Without IN, the IC structure has been predicted to have a structural void (hole) in its catalytic pocket [27], as illustrated in Figure 4C. A similar prediction has also been made for IC of the Ssp DnaB intein [30], where the hole was thought to be a docking place for the IN to trigger a C-cleavage reaction. To explain why the IC of Ssp DnaX intein (but not of Ssp DnaB intein) could undergo spontaneous C-cleavage in the absence of IN, it was suggested that the predicted hole of this IC might be sealed or compensated to allow the spontaneous C-cleavage [27]. Our findings in this study show that this IC can also perform protein trans-splicing with IN, indicating that the predicted hole of IC is open at least some of the time, in order for the IN to dock for trans-splicing. We suggest that the hole of IC may exist in two equilibrium states: an open state (hole is open) allowing the docking of IN for trans-splicing, and a closed state (hole is closed) allowing spontaneous C-cleavage without IN. Our findings also indicate that the open state predominates, because the trans-splicing reaction completely suppressed the C-cleavage reaction when IN was present. Our suggestion is consistent with an earlier study of the Ssp DnaB intein, where the horseshoe-like structure of the large IC was suggested to open up and clamp onto the small IN [30].

thumbnail
Figure 4. Structural modeling of the intein and its parts.

Computer-based modeling of the Ssp DnaX intein (A) and its parts (B, C, D) used the VMD program (http://www.ks.uiuc.edu/Research/vmd/) and the crystal structure of the Ssp DnaB mini-intein [7]. The 11-aa IN part of the intein is shown in red. The C-terminal 6-aa part of the intein is shown in blue.

https://doi.org/10.1371/journal.pone.0045355.g004

The above suggestion is further supported by our finding that the trans-splicing reaction could occur when the small IN was replaced with INL, where INL is a near complete intein containing the IN. The IN part of INL must have participated in the trans-splicing reaction, whereas the remaining part of INL (at least the conserved Block B motif) apparently did not participate in the reaction, because a mutation in the Block B motif of INL did not prevent trans-splicing. To participate in the trans-splicing reaction, the IN part of INL needs to move out from its buried position in INL (Figure 4D), before it can dock into the catalytic pocket (open hole) of IC for trans-splicing. This suggests that the INL structure is also able to open up in order for the IN part to ‘swing out’ to an exposed position. Consistent with this suggestion, a mutated C-intein (ICm) trans-spliced with INL, suggesting that the ICm structure could open up to allow its C-terminal part to “swing out” to an exposed position for participation in the trans-splicing reaction. Furthermore, we discovered cross-reactivity between IC of the atypical Ssp DnaX split intein and the small IN from two other inteins, which is the first finding of cross-reactivity for such atypical split inteins. Considering that the IN of the other inteins is only 40–50% identical to the native IN in amino acid sequence, it is interesting that the non-native IN can correctly dock into the structural hole of the IC and catalyze the trans-splicing reaction. Overall, these findings revealed an interesting structural flexibility at or near the catalytic pocket of inteins, which can have significant implications on future engineering of split inteins for peptide-protein or protein-protein trans-splicing for various applications.

The trans-splicing function of this atypical Ssp DnaX split intein also makes a significant addition to the intein-based toolbox for general uses. Previously only the Ssp DnaB intein has been engineered into such an atypical split intein, and was named the S1 split intein [14]. Unlike other forms of split inteins, the S1 split intein has an extremely small IN and is therefore particularly useful for splicing synthetic peptides onto the N-terminus of target proteins. The synthetic peptide can easily accommodate the 11-aa IN plus a small N-extein to be spliced onto the N-terminus of a target protein, with the target protein being a fusion protein containing the IC. This peptide-protein trans-splicing is useful for site-specific labeling or modifications of proteins, because the synthetic N-extein may be engineered to carry a variety of chemical moieties, including fluorescent groups, modified or unnatural amino acids, and drug molecules, as long as the chemical moiety does not block trans-splicing. Finding and understanding new S1 split inteins, as we have done in this study, is important for wide uses of this peptide-protein trans-splicing method, because different inteins have been known to splice differently when used on different target proteins [25], [31]. It is impressive that this new S1 split intein could perform the trans-splicing reaction at temperatures ranging from ∼1°C (on ice) to 37°C, although the reaction speed was somewhat lower at 1–4°C temperatures. This temperature tolerance may be due to the fact that this intein was derived from a natural intein found in a cyanobacterium (Syenochocystis sp. PCC6803) that lives under a wide range of environmental temperatures. This robust nature of the S1 split intein can be an advantage in practical applications where one may need to achieve trans-splicing under low temperatures. Our finding of cross-reactivity between the IC of the atypical Ssp DnaX split intein and the small IN from two other inteins also has interesting implications. On the one hand, it permits different choices for the IN for doing peptide-protein trans-splicing and suggests that the IN sequence may tolerate many sequence changes, which can be useful information for designing and producing synthetic peptides containing IN. On the other hand, two atypical split inteins may not be used together in a mixed system to achieve labeling or modification of two different target proteins in a protein-specific manner.

Materials and Methods

Plasmid Construction

Plasmid pMSX-S1 for in vivo experiments was constructed as described previously [14], in which the two open reading frames expressing the MIN and ICT proteins were separated by a spacer sequence. A restriction enzyme cutting site Afl II was introduced at the split site. To construct plasmid pMSX-S1N expressing the MIN protein alone, a DNA fragment between Afl II and Hind III sites was deleted from plasmid pMSX-S1. Plasmid pMSX-S1C expressing the ICT protein alone was from [27]. Plasmid pMSX-S1NL expressing the MINL protein was constructed by replacing the IN coding sequence in pMSX-S1N with the first 141 codons of the Ssp DnaX intein [32], which used standard recombinant DNA methods including PCR, DNA cutting(XhoI- Afl II), and ligation. Plasmid pMRB-S1N expressing the MINRB protein and plasmid pMSG-S1N expressing the MINSG protein were constructed by replacing the IN coding sequence in pMSX-S1N with the first 12 codons of the Rma DnaB intein [28] and Ssp GyrB intein [29], respectively. To construct pMRBSX-S1, pMSGSX-S1, the IN coding sequence between XhoI and Afl II in pMSX-S1was replaced with INRB or INSG sequence respectively. To construct plasmid pmMSX-S1C (or pmMSX-S1NL) expressing the ICmT (or MINLm) protein, site-directed mutations were introduced into plasmid pMSX-S1C (or pMSX-S1NL), using a standard method of inverse PCR.

Protein Expression, Purification and in vitro Reactions

Plasmids pMSX-S1, pMSX-S1N, pMSX-S1NL, pMRB-S1N, pMSG-S1N and pmMSX-S1NL were each transformed into Escherichia coli DH5α cells, while plasmids pMSX-S1C and pmMSX-S1C were each transformed into Escherichia coli BL21(DE3) cells, all using a standard E. coli transformation protocol. The transformed E. coli cells were grown in 50 mL of Luria Broth (LB) medium at 37°C to mid-logarithmic phase (OD600 of ∼0.6) and induced by 0.8 mM IPTG to express, at room temperature overnight, the plasmid-encoded protein(s) of interest. To analyze total cellular proteins, cells were harvested by centrifugation and solubilized directly in SDS loading buffer. For protein purification, harvested cells were lysed using a French Press (14,000 PSI), and the cell lysate was centrifuged to remove any insoluble materials. To purify the MIN, MINL, and MINLm proteins containing the maltose binding protein, amylose resin was used according to the manufacturer’s instructions (New England Biolabs). To purify the ICT and ICmT proteins containing a hexahistidine tag, Ni-NTA resin (QIAGEN) was used according to the manufacturer’s instructions. For in vitro trans-splicing or cleavage reactions, the specified precursor proteins were mixed and incubated under specified conditions, with 1 mM DTT added to all in vitro reactions. Western blotting used an anti-thioredoxin (anti-T) antibody (Invitrogen) and the Enhanced Chem-luminescence detection kit (GE Healthcare), all according to the manufacturer’s instructions.

Computational Simulations of Intein Structure

A simulated three-dimensional structure of Ssp DanX mini-intein was obtained by using the fully automated homology-modeling pipeline SWISS-MODEL [33], [34]. The intein amino acid sequence [27] was uploaded to the Automatic Modeling Workspace (http://swissmodel.expasy.org/workspace), the crystal structure of Ssp DnaB mini-intein was selected automatically as a closest template by homology, and structural models of Ssp DanX mini-intein were generated and presented in NewCartoon style.

Acknowledgments

We thank Dr. Xingmei Qi for providing the plasmid pMSX-S1C and Dr. David Spencer for a critical reading of this manuscript.

Author Contributions

Conceived and designed the experiments: XQL. Performed the experiments: HS. Analyzed the data: HS QM XQL. Contributed reagents/materials/analysis tools: QM XQL. Wrote the paper: HS XQL.

References

  1. 1. Perler FB, Davis EO, Dean GE, Gimble FS, Jack WE, et al. (1994) Protein splicing elements: inteins and exteins–a definition of terms and recommended nomenclature. Nucleic Acids Res 22: 1125–1127.
  2. 2. Perler FB (2005) Protein splicing mechanisms and applications. IUBMB Life 57: 469–476.
  3. 3. Xu MQ, Perler FB (1996) The mechanism of protein splicing and its modulation by mutation. Embo J 15: 5146–5153.
  4. 4. Klabunde T, Sharma S, Telenti A, Jacobs WR Jr, Sacchettini JC (1998) Crystal structure of GyrA intein from Mycobacterium xenopi reveals structural basis of protein splicing. Nat Struct Biol 5: 31–36.
  5. 5. Ichiyanagi K, Ishino Y, Ariyoshi M, Komori K, Morikawa K (2000) Crystal structure of an archaeal intein-encoded homing endonuclease PI-PfuI. J Mol Biol 300: 889–901.
  6. 6. Hall TM, Porter JA, Young KE, Koonin EV, Beachy PA, et al. (1997) Crystal structure of a Hedgehog autoprocessing domain: homology between Hedgehog and self-splicing proteins. Cell 91: 85–97.
  7. 7. Ding Y, Xu MQ, Ghosh I, Chen X, Ferrandon S, et al. (2003) Crystal structure of a mini-intein reveals a conserved catalytic module involved in side chain cyclization of asparagine during protein splicing. J Biol Chem 278: 39133–39142.
  8. 8. Wu H, Hu Z, Liu XQ (1998) Protein trans-splicing by a split intein encoded in a split DnaE gene of Synechocystis sp. PCC6803. Proc Natl Acad Sci U S A 95: 9226–9231.
  9. 9. Evans TC Jr, Martin D, Kolly R, Panne D, Sun L, et al. (2000) Protein trans-splicing and cyclization by a naturally split intein from the dnaE gene of Synechocystis species PCC6803. J Biol Chem 275: 9091–9094.
  10. 10. Evans TC Jr, Benner J, Xu MQ (1998) Semisynthesis of cytotoxic proteins using a modified protein splicing element. Protein Sci 7: 2256–2264.
  11. 11. Wu W, Wood DW, Belfort G, Derbyshire V, Belfort M (2002) Intein-mediated purification of cytotoxic endonuclease I-TevI by insertional inactivation and pH-controllable splicing. Nucleic Acids Res 30: 4864–4871.
  12. 12. Toshio Yamazaki TO, Natsuko Oda, Yoshimasa Kyogoku, Koichi Uegaki, Nobutoshi Ito, Yoshizumi Ishino, Haruki Nakamura (1998) Segmental isotope labeling for protein NMR using peptide splicing. Journal of the American Chemical Society 120: 2.
  13. 13. Li J, Sun W, Wang B, Xiao X, Liu XQ (2008) Protein trans-splicing as a means for viral vector-mediated in vivo gene therapy. Hum Gene Ther 19: 958–964.
  14. 14. Sun W, Yang J, Liu XQ (2004) Synthetic two-piece and three-piece split inteins for protein trans-splicing. J Biol Chem 279: 35281–35286.
  15. 15. Ludwig C, Pfeiff M, Linne U, Mootz HD (2006) Ligation of a synthetic peptide to the N terminus of a recombinant protein using semisynthetic protein trans-splicing. Angew Chem Int Ed Engl 45: 5218–5221.
  16. 16. Ando T, Tsukiji S, Tanaka T, Nagamune T (2007) Construction of a small-molecule-integrated semisynthetic split intein for in vivo protein ligation. Chem Commun (Camb): 4995–4997.
  17. 17. Gilmore JM, Scheck RA, Esser-Kahn AP, Joshi NS, Francis MB (2006) N-terminal protein modification through a biomimetic transamination reaction. Angew Chem Int Ed Engl 45: 5307–5311.
  18. 18. Lin CW, Ting AY (2006) Transglutaminase-catalyzed site-specific conjugation of small-molecule probes to proteins in vitro and on the surface of living cells. J Am Chem Soc 128: 4542–4543.
  19. 19. Xie J, Schultz PG (2006) A chemical toolkit for proteins–an expanded genetic code. Nat Rev Mol Cell Biol 7: 775–782.
  20. 20. Muir TW (2003) Semisynthesis of proteins by expressed protein ligation. Annu Rev Biochem 72: 249–289.
  21. 21. Tan LP, Yao SQ (2005) Intein-mediated, in vitro and in vivo protein modifications with small molecules. Protein Pept Lett 12: 769–775.
  22. 22. Kalia J, Raines RT (2006) Reactivity of intein thioesters: appending a functional group to a protein. Chembiochem 7: 1375–1383.
  23. 23. Amitai G, Callahan BP, Stanger MJ, Belfort G, Belfort M (2009) Modulation of intein activity by its neighboring extein substrates. Proc Natl Acad Sci U S A 106: 11005–11010.
  24. 24. Iwai H, Zuger S, Jin J, Tam PH (2006) Highly efficient protein trans-splicing by a naturally split DnaE intein from Nostoc punctiforme. FEBS Lett 580: 1853–1858.
  25. 25. Noren CJ, Wang J, Perler FB (2000) Dissecting the Chemistry of Protein Splicing and Its Applications. Angew Chem Int Ed Engl 39: 450–466.
  26. 26. Appleby-Tagoe JH, Thiel IV, Wang Y, Wang Y, Mootz HD, et al. (2011) Highly efficient and more general cis- and trans-splicing inteins through sequential directed evolution. J Biol Chem 286: 34440–34447.
  27. 27. Qi X, Meng Q, Liu XQ (2011) Spontaneous C-cleavage of a mini-intein without its conserved N-terminal motif A. FEBS Lett. 585: 2513–2518.
  28. 28. Liu XQ, Hu Z (1997) A DnaB intein in Rhodothermus marinus: indication of recent intein homing across remotely related organisms. Proc Natl Acad Sci U S A 94: 7851–7856.
  29. 29. Gorbalenya AE (1998) Non-canonical inteins. Nucleic Acids Res 26: 1741–1748.
  30. 30. Volkmann G, Sun W, Liu XQ (2009) Controllable protein cleavages through intein fragment complementation. Protein Sci 18: 2393–2402.
  31. 31. Paulus H (2001) Inteins as enzymes. Bioorg Chem 29: 119–129.
  32. 32. Liu XQ, Hu Z (1997) Identification and characterization of a cyanobacterial DnaX intein. FEBS Lett 408: 311–314.
  33. 33. Kiefer F, Arnold K, Kunzli M, Bordoli L, Schwede T (2009) The SWISS-MODEL Repository and associated resources. Nucleic Acids Res 37: D387–392.
  34. 34. Kopp J, Schwede T (2004) The SWISS-MODEL Repository of annotated three-dimensional protein structure homology models. Nucleic Acids Res 32: D230–234.