Protein splicing

Last updated
The mechanism of protein splicing involving inteins. In this scheme, the N-extein is shown in red, the intein in black, and the C-extein in blue. X represents either an oxygen or sulfur atom. Intein mech.png
The mechanism of protein splicing involving inteins. In this scheme, the N-extein is shown in red, the intein in black, and the C-extein in blue. X represents either an oxygen or sulfur atom.

Protein splicing is an intramolecular reaction of a particular protein in which an internal protein segment (called an intein) is removed from a precursor protein with a ligation of C-terminal and N-terminal external proteins (called exteins) on both sides. The splicing junction of the precursor protein is mainly a cysteine or a serine, which are amino acids containing a nucleophilic side chain. The protein splicing reactions which are known now do not require exogenous cofactors or energy sources such as adenosine triphosphate (ATP) or guanosine triphosphate (GTP). Normally, splicing is associated only with pre-mRNA splicing. This precursor protein contains three segments—an N-extein followed by the intein followed by a C-extein. After splicing has taken place, the resulting protein contains the N-extein linked to the C-extein; this splicing product is also termed an extein.

Contents

History

The first intein was discovered in 1988 through sequence comparison between the Neurospora crassa [1] and carrot [2] vacuolar ATPase (without intein) and the homologous gene in yeast (with intein) that was first described as a putative calcium ion transporter. [3] In 1990 Hirata et al. [4] demonstrated that the extra sequence in the yeast gene was transcribed into mRNA and removed itself from the host protein only after translation. Since then, inteins have been found in all three domains of life (eukaryotes, bacteria, and archaea) and in viruses.

Protein splicing was unanticipated and its mechanisms were discovered by two groups (Anraku [5] and Stevens [6] ) in 1990. They both discovered a Saccharomyces cerevisiae VMA1 in a precursor of a vacuolar H+-ATPase enzyme. The amino acid sequence of the N- and C-termini corresponded to 70% DNA sequence of that of a vacuolar H+-ATPase from other organisms, while the amino acid sequence of the central position corresponded to 30% of the total DNA sequence of the yeast HO nuclease.

Many genes have unrelated intein-coding segments inserted at different positions. For these and other reasons, inteins (or more properly, the gene segments coding for inteins) are sometimes called selfish genetic elements , but it may be more accurate to call them parasitic. According to the gene centered view of evolution, most genes are "selfish" only insofar as to compete with other genes or alleles but usually they fulfill a function for the organisms, whereas "parasitic genetic elements", at least initially, do not make a positive contribution to the fitness of the organism. [7] [8]

As of December 2019, the UniProtKB database contains 188 entries manually annotated as inteins, ranging from just tens of amino acid residues to thousands. [9] The first intein was found encoded within the VMA gene of Saccharomyces cerevisiae. They were later found in fungi (ascomycetes, basidiomycetes, zygomycetes and chytrids) and in diverse proteins as well. A protein distantly related to known inteins containing protein, but closely related to metazoan hedgehog proteins, has been described to have the intein sequence from Glomeromycota. Many of the newly described inteins contain homing endonucleases and some of these are apparently active. [10] The abundance of intein in fungi indicates lateral transfer of intein-containing genes. While in eubacteria and archaea, there are 289 and 182 currently known inteins. Not surprisingly, most intein in eubacteria and archaea are found to be inserted into nucleic acid metabolic protein, like fungi. [10]

Inteins vary greatly, but many of the same intein-containing proteins are found in a number of species. For example, pre-mRNA processing factor 8 (Prp8) protein, instrumental in the spliceosome, has seven different intein insertion sites across eukaryotic species. [11] Intein-containing Prp8 is most commonly found in fungi, but is also seen in Amoebozoa, Chlorophyta, Capsaspora, and Choanoflagellida. Many mycobacteria contain inteins within DnaB (bacterial replicative helicase), RecA (bacterial DNA recombinase), and SufB (FeS cluster assembly protein). [12] [13] There is remarkable variety within the structure and number of DnaB inteins, both within the mycobacterium genus and beyond. Interestingly, intein-containing DnaB is also found in the chloroplasts of algae. [14] Intein-containing proteins found in archaea include RadA (RecA homolog), RFC, PolB, RNR. [15] Many of the same intein-containing proteins (or their homologs) are found in two or even all three domains of life. Inteins are also seen in the proteomes encoded by bacteriophages and eukaryotic viruses. Viruses may have been involved as vectors of intein distribution across the wide variety of intein containing organisms. [15]

Mechanism

The process for class 1 inteins begins with an N-O or N-S shift when the side chain of the first residue (a serine, threonine, or cysteine) of the intein portion of the precursor protein nucleophilically attacks the peptide bond of the residue immediately upstream (that is, the final residue of the N-extein) to form a linear ester (or thioester) intermediate. A transesterification occurs when the side chain of the first residue of the C-extein attacks the newly formed (thio)ester to free the N-terminal end of the intein. This forms a branched intermediate in which the N-extein and C-extein are attached, albeit not through a peptide bond. The last residue of the intein is always an asparagine (Asn), and the amide nitrogen atom of this side chain cleaves apart the peptide bond between the intein and the C-extein, resulting in a free intein segment with a terminal cyclic imide. Finally, the free amino group of the C-extein now attacks the (thio)ester linking the N- and C-exteins together. An O-N or S-N shift produces a peptide bond and the functional, ligated protein. [16]

Class 2 inteins have no nucleophilic first side chain, only an alanine. Instead, the reaction starts directly with a nucleophilic displacement, with the first residue of the C-extein atticking the peptide carboxyl on the final residue of the N-extein. The rest proceeds as usual, starting with Asn turning into a cyclic imide. [17]

Class 3 inteins have no nucleophilic first side chain, only an alanine, yet they have an internal noncontiguous "WCT" motif. The internal C (cysteine) residue attacks the peptide carboxyl on the final residue of the N-extein (nucleophilic displacement). Transesterification occurs when the first residue of the C-extein attacks the newly formed thioester. The rest proceeds as usual. [18]

The mechanism for the splicing effect is a naturally occurring analogy to the technique for chemically generating medium-sized proteins called native chemical ligation.

Intein

An intein is a segment of a protein that is able to excise itself and join the remaining portions (the exteins) with a peptide bond during protein splicing. [19] Inteins have also been called protein introns, by analogy with (RNA) introns.

Intein splicing occurs post-translationally in a self-catalytic process. Here, the extein is shown in red and the intein in blue. Image created with Biorender.com. Intein splicing dogma.png
Intein splicing occurs post-translationally in a self-catalytic process. Here, the extein is shown in red and the intein in blue. Image created with Biorender.com.

Naming conventions

The first part of an intein name is based on the scientific name of the organism in which it is found, and the second part is based on the name of the corresponding gene or extein. For example, the intein found in Thermoplasma acidophilum and associated with Vacuolar ATPase subunit A (VMA) is called "Tac VMA".

Normally, as in this example, just three letters suffice to specify the organism, but there are variations. For example, additional letters may be added to indicate a strain. If more than one intein is encoded in the corresponding gene, the inteins are given a numerical suffix starting from 5 to 3 or in order of their identification (for example, "Msm dnaB-1").

The segment of the gene that encodes the intein is usually given the same name as the intein, but to avoid confusion the name of the intein proper is usually capitalized (e.g., Pfu RIR1-1), whereas the name of the corresponding gene segment is italicized (e.g., Pfu rir1-1). A different disambiguating convention is to place a lowercase "i" after the source protein name, e.g. "Msm DnaBi1". [20]

Types of inteins

Inteins can be classified on many criteria.

Full and mini inteins

Inteins can contain a homing endonuclease gene (HEG) domain in addition to the splicing domains. This domain is responsible for the spread of the intein by cleaving DNA at an intein-free allele on the homologous chromosome, triggering the DNA double-stranded break repair (DSBR) system, which then repairs the break, thus copying the intein-coding DNA into a previously intein-free site. [17] The HEG domain is not necessary for intein splicing, and so it can be lost, forming a minimal, or mini, intein. Several studies have demonstrated the modular nature of inteins by adding or removing HEG domains and determining the activity of the new construct.[ citation needed ]

Split inteins

Sometimes, the intein of the precursor protein comes from two genes. In this case, the intein is said to be a split intein. For example, in cyanobacteria, DnaE, the catalytic subunit α of DNA polymerase III, is encoded by two separate genes, dnaE-n and dnaE-c. The dnaE-n product consists of an N-extein sequence followed by a 123-AA intein sequence, whereas the dnaE-c product consists of a 36-AA intein sequence followed by a C-extein sequence. [21]

Applications in biotechnology

Inteins are very efficient at protein splicing, and they have accordingly found an important role in biotechnology. There are more than 200 inteins identified to date; sizes range from 100800 AAs. Inteins have been engineered for particular applications such as protein semisynthesis [22] and the selective labeling of protein segments, which is useful for NMR studies of large proteins. [23]

Pharmaceutical inhibition of intein excision may be a useful tool for drug development; the protein that contains the intein will not carry out its normal function if the intein does not excise, since its structure will be disrupted.

It has been suggested that inteins could prove useful for achieving allotopic expression of certain highly hydrophobic proteins normally encoded by the mitochondrial genome, for example in gene therapy. [24] The hydrophobicity of these proteins is an obstacle to their import into mitochondria. Therefore, the insertion of a non-hydrophobic intein may allow this import to proceed. Excision of the intein after import would then restore the protein to wild-type.

Affinity tags have been widely used to purify recombinant proteins, as they allow the accumulation of recombinant protein with little impurities. However, the affinity tag must be removed by proteases in the final purification step. The extra proteolysis step raises the problems of protease specificity in removing affinity tags from recombinant protein, and the removal of the digestion product. This problem can be avoided by fusing an affinity tag to self-cleavable inteins in a controlled environment. The first generation of expression vectors of this kind used modified Saccharomyces cerevisiae VMA (Sce VMA) intein. Chong et al. [25] used a chitin binding domain (CBD) from Bacillus circulans as an affinity tag, and fused this tag with a modified Sce VMA intein. The modified intein undergoes a self-cleavage reaction at its N-terminal peptide linkage with 1,4-dithiothreitol (DTT), β-mercaptoethanol (β-ME), or cystine at low temperatures over a broad pH range. After expressing the recombinant protein, the cell homogenate is passed through the column containing chitin. This allows the CBD of the chimeric protein to bind to the column. Furthermore, when the temperature is lowered and the molecules described above pass through the column, the chimeric protein undergoes self-splicing and only the target protein is eluted. This novel technique eliminates the need for a proteolysis step, and modified Sce VMA stays in column attached to chitin through CBD. [25]

Recently inteins have been used to purify proteins based on self aggregating peptides. Elastin-like polypeptides (ELPs) are a useful tool in biotechnology. Fused with target protein, they tend to form aggregates inside the cells. [26] This eliminates the chromatographic step needed in protein purification. The ELP tags have been used in the fusion protein of intein, so that the aggregates can be isolated without chromatography (by centrifugation) and then intein and tag can be cleaved in controlled manner to release the target protein into solution. This protein isolation can be done using continuous media flow, yielding high amounts of protein, making this process more economically efficient than conventional methods. [26] Another group of researchers used smaller self aggregating tags to isolate target protein. Small amphipathic peptides 18A and ELK16 (figure 5) were used to form self cleaving aggregating protein. [27]

Applications in Antimicrobial Development

Over the last twenty years, there has been increasing interest in leveraging inteins for antimicrobial applications. [12] Intein splicing is found exclusively in unicellular organisms, with a particularly high abundance in pathogenic microorganisms. [28] Furthermore, inteins are commonly found within housekeeping proteins and/or proteins involved in the survival of the organism within a human host. Post-translational intein removal is necessary for the protein to properly fold and function. For example, Gaëlle Huet et al. demonstrated that in Mycobacterium tuberculosis , unspliced SufB prevents the formation of the SufBCD complex, a component of the SUF machinery. [29] As such, the inhibition of intein splicing may serve as a powerful platform for the development of antimicrobials.

Current research on intein splicing inhibitors has focused on developing antimycobacterials (M. tb. has three intein-containing proteins), as well as agents active against pathogenic fungi Cryptococcus and Aspergillus. [13] Cisplatin and similar platinum-containing compounds inhibit splicing of the M. tb. RecA intein through coordinating to catalytic residues. [30] Divalent cations, such as copper (II) and zinc (II) ions, function similarly to reversibly inhibit splicing. [12] However, neither of these methods are currently suitable for an effective and safe antibiotic. The fungal Prp8 intein is also inhibited by divalent cations and cisplatin through interfering with the catalytic Cys1 residue. [12] In 2021, Li et al. showed that small molecule inhibitors of Prp8 intein splicing were selective and effective at slowing the growth of C. neoformans and C. gattii, providing exciting evidence for the antimicrobial potential of intein splicing inhibitors. [31]

See also

Related Research Articles

<span class="mw-page-title-main">Central dogma of molecular biology</span> Explanation of the flow of genetic information within a biological system

The central dogma of molecular biology is an explanation of the flow of genetic information within a biological system. It is often stated as "DNA makes RNA, and RNA makes protein", although this is not its original meaning. It was first stated by Francis Crick in 1957, then published in 1958:

The Central Dogma. This states that once "information" has passed into protein it cannot get out again. In more detail, the transfer of information from nucleic acid to nucleic acid, or from nucleic acid to protein may be possible, but transfer from protein to protein, or from protein to nucleic acid is impossible. Information here means the precise determination of sequence, either of bases in the nucleic acid or of amino acid residues in the protein.

Marlene Belfort is an American biochemist known for her research on the factors that interrupt genes and proteins. She is a fellow of the American Academy of Arts and Sciences and has been admitted to the United States National Academy of Sciences.

<span class="mw-page-title-main">Fusion protein</span> Protein created by joining other proteins into a single polypeptide

Fusion proteins or chimeric (kī-ˈmir-ik) proteins are proteins created through the joining of two or more genes that originally coded for separate proteins. Translation of this fusion gene results in a single or multiple polypeptides with functional properties derived from each of the original proteins. Recombinant fusion proteins are created artificially by recombinant DNA technology for use in biological research or therapeutics. Chimeric or chimera usually designate hybrid proteins made of polypeptides having different functions or physico-chemical patterns. Chimeric mutant proteins occur naturally when a complex mutation, such as a chromosomal translocation, tandem duplication, or retrotransposition creates a novel coding sequence containing parts of the coding sequences from two different genes. Naturally occurring fusion proteins are commonly found in cancer cells, where they may function as oncoproteins. The bcr-abl fusion protein is a well-known example of an oncogenic fusion protein, and is considered to be the primary oncogenic driver of chronic myelogenous leukemia.

<span class="mw-page-title-main">Homing endonuclease</span>

The homing endonucleases are a collection of endonucleases encoded either as freestanding genes within introns, as fusions with host proteins, or as self-splicing inteins. They catalyze the hydrolysis of genomic DNA within the cells that synthesize them, but do so at very few, or even singular, locations. Repair of the hydrolyzed DNA by the host cell frequently results in the gene encoding the homing endonuclease having been copied into the cleavage site, hence the term 'homing' to describe the movement of these genes. Homing endonucleases can thereby transmit their genes horizontally within a host population, increasing their allele frequency at greater than Mendelian rates.

<span class="mw-page-title-main">ATP6AP1</span> Protein-coding gene in humans

The human gene ATP6AP1 encodes the S1 subunit of the enzyme V-type proton ATPase.

<span class="mw-page-title-main">ATP6V1C1</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit C 1 is an enzyme that in humans is encoded by the ATP6V1C1 gene.

<span class="mw-page-title-main">ATPase, H+ transporting, lysosomal V0 subunit a1</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase 116 kDa subunit a isoform 1 is an enzyme that in humans is encoded by the ATP6V0A1 gene.

<span class="mw-page-title-main">ATP6V1A</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase catalytic subunit A is an enzyme that in humans is encoded by the ATP6V1A gene.

<span class="mw-page-title-main">ATP6V0D1</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit d 1 is an enzyme that in humans is encoded by the ATP6V0D1 gene.

<span class="mw-page-title-main">ATP6V1H</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit H is an enzyme that in humans is encoded by the ATP6V1H gene.

<span class="mw-page-title-main">ATP6V1G1</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit G 1 is an enzyme that in humans is encoded by the ATP6V1G1 gene.

<span class="mw-page-title-main">ATP6V1D</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit D is an enzyme that in humans is encoded by the ATP6V1D gene.

<span class="mw-page-title-main">ATP6V1F</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit F is an enzyme that in humans is encoded by the ATP6V1F gene.

<span class="mw-page-title-main">ATP6V0B</span> Protein-coding gene in Homo sapiens

V-type proton ATPase 21 kDa proteolipid subunit is an enzyme that in humans is encoded by the ATP6V0B gene.

<span class="mw-page-title-main">ATP6V0E1</span> Protein-coding gene in the species Homo sapiens

V-type proton ATPase subunit e 1 is an enzyme that in humans is encoded by the ATP6V0E1 gene.

Split-intein circular ligation of peptides and proteins (SICLOPPS) is a biotechnology technique that permits the creation of cyclic peptides. These peptides are produced by ribosomal protein synthesis, followed by an intein-like event that splices the protein into a loop. By contrast with the nonribosomal peptide synthetases that produces some cyclic peptides like gramicidin S, SICLOPPS offers the advantage that the peptides' structure can be encoded by DNA in a simple manner according to the genetic code, but for this reason it imposes limitations on the types of amino acids incorporated that are comparable to those that apply to ordinary proteins. As implemented there is also some constraint on the peptide sequence of the cyclic sequence; for example, libraries may use the sequence SGXX..XXPL to increase the efficiency of circularization of the peptide. SICLOPPS is frequently used with a library of randomized DNA sequence that permits the simultaneous production and screening of large numbers of constructs at once, followed by the recovery of the DNA sequences responsible for the activity of the clone of interest.

<span class="mw-page-title-main">Prp8</span>

Prp8 refers to both the Prp8 protein and Prp8 gene. Prp8's name originates from its involvement in pre-mRNA processing. The Prp8 protein is a large, highly conserved, and unique protein that resides in the catalytic core of the spliceosome and has been found to have a central role in molecular rearrangements that occur there. Prp8 protein is a major central component of the catalytic core in the spliceosome, and the spliceosome is responsible for splicing of precursor mRNA that contains introns and exons. Unexpressed introns are removed by the spliceosome complex in order to create a more concise mRNA transcript. Splicing is just one of many different post-transcriptional modifications that mRNA must undergo before translation. Prp8 has also been hypothesized to be a cofactor in RNA catalysis.

Asparagine peptide lyase are one of the seven groups in which proteases, also termed proteolytic enzymes, peptidases, or proteinases, are classified according to their catalytic residue. The catalytic mechanism of the asparagine peptide lyases involves an asparagine residue acting as nucleophile to perform a nucleophilic elimination reaction, rather than hydrolysis, to catalyse the breaking of a peptide bond.

<span class="mw-page-title-main">Large tumor antigen</span>

The large tumor antigen is a protein encoded in the genomes of polyomaviruses, which are small double-stranded DNA viruses. LTag is expressed early in the infectious cycle and is essential for viral proliferation. Containing four well-conserved protein domains as well as several intrinsically disordered regions, LTag is a fairly large multifunctional protein; in most polyomaviruses, it ranges from around 600-800 amino acids in length. LTag has two primary functions, both related to replication of the viral genome: it unwinds the virus's DNA to prepare it for replication, and it interacts with proteins in the host cell to dysregulate the cell cycle so that the host's DNA replication machinery can be used to replicate the virus's genome. Some polyomavirus LTag proteins - most notably the well-studied SV40 large tumor antigen from the SV40 virus - are oncoproteins that can induce neoplastic transformation in the host cell.

References

  1. Bowman, EJ; Tenney, K; Bowman, BJ (Oct 1988). "Isolation of genes encoding the Neurospora vacuolar ATPase. Analysis of vma-1 encoding the 67-kDa subunit reveals homology to other ATPases". J Biol Chem. 263 (28): 13994–4001. doi: 10.1016/S0021-9258(18)68175-X . PMID   2971651.
  2. Zimniak, L; Dittrich, P; Gogarten, JP; Kibak, H; Taiz, L (Jul 1988). "The cDNA sequence of the 69-kDa subunit of the carrot vacuolar H+-ATPase. Homology to the beta-chain of F0F1-ATPases". J Biol Chem. 263 (19): 9102–12. doi: 10.1016/S0021-9258(19)76514-4 . PMID   2897965.
  3. Shih, CK; Wagner, R; Feinstein, S; Kanik-Ennulat, C; Neff, N (Aug 1988). "A dominant trifluoperazine resistance gene from Saccharomyces cerevisiae has homology with F0F1 ATP synthase and confers calcium-sensitive growth". Mol Cell Biol. 8 (8): 3094–103. doi:10.1128/mcb.8.8.3094. PMC   363536 . PMID   2905423.
  4. Hirata, R; Ohsumk, Y; Nakano, A; Kawasaki, H; Suzuki, K; Anraku, Y (Apr 1990). "Molecular structure of a gene, VMA1, encoding the catalytic subunit of H(+)-translocating adenosine triphosphatase from vacuolar membranes of Saccharomyces cerevisiae". J Biol Chem. 265 (12): 6726–33. doi: 10.1016/S0021-9258(19)39210-5 . PMID   2139027.
  5. Hirata R, Ohsumk Y, Nakano A, Kawasaki H, Suzuki K, Anraku Y (April 1990). "Molecular structure of a gene, VMA1, encoding the catalytic subunit of H(+)-translocating adenosine triphosphatase from vacuolar membranes of Saccharomyces cerevisiae". J. Biol. Chem. 265 (12): 6726–33. doi: 10.1016/S0021-9258(19)39210-5 . PMID   2139027.
  6. Kane PM, Yamashiro CT, Wolczyk DF, Neff N, Goebl M, Stevens TH (November 1990). "Protein splicing converts the yeast TFP1 gene product to the 69-kD subunit of the vacuolar H(+)-adenosine triphosphatase". Science. 250 (4981): 651–7. Bibcode:1990Sci...250..651K. doi:10.1126/science.2146742. PMID   2146742.
  7. Swithers, Kristen S.; Soucy, Shannon M.; Gogarten, J. Peter (2012). "The Role of Reticulate Evolution in Creating Innovation and Complexity". International Journal of Evolutionary Biology. 2012: 1–10. doi: 10.1155/2012/418964 . ISSN   2090-8032. PMC   3403396 . PMID   22844638.
  8. Dawkins, Richard (1976). The Selfish Gene . Oxford University Press.
  9. "UniProt: the universal protein knowledgebase". Nucleic Acids Research. 45 (D1): D158–D169. 2016-11-29. doi:10.1093/nar/gkw1099. ISSN   0305-1048. PMC   5210571 . PMID   27899622.
  10. 1 2 Perler, F. B. (2002). "InBase: the Intein Database". Nucleic Acids Research. 30 (1): 383–384. doi:10.1093/nar/30.1.383. ISSN   1362-4962. PMC   99080 . PMID   11752343.
  11. Green, Cathleen M.; Li, Zhong; Smith, Aaron D.; Novikova, Olga; Bacot-Davis, Valjean R.; Gao, Fengshan; Hu, Saiyang; Banavali, Nilesh K.; Thiele, Dennis J.; Li, Hongmin; Belfort, Marlene (2019-10-10). "Spliceosomal Prp8 intein at the crossroads of protein and RNA splicing". PLOS Biology. 17 (10): e3000104. doi: 10.1371/journal.pbio.3000104 . ISSN   1545-7885. PMC   6805012 . PMID   31600193.
  12. 1 2 3 4 Tharappel, Anil Mathew; Li, Zhong; Li, Hongmin (2022). "Inteins as Drug Targets and Therapeutic Tools". Frontiers in Molecular Biosciences. 9: 821146. doi: 10.3389/fmolb.2022.821146 . ISSN   2296-889X. PMC   8861304 . PMID   35211511.
  13. 1 2 Wall, Diana A.; Tarrant, Seanan P.; Wang, Chunyu; Mills, Kenneth V.; Lennon, Christopher W. (2021). "Intein Inhibitors as Novel Antimicrobials: Protein Splicing in Human Pathogens, Screening Methods, and Off-Target Considerations". Frontiers in Molecular Biosciences. 8: 752824. doi: 10.3389/fmolb.2021.752824 . ISSN   2296-889X. PMC   8529194 . PMID   34692773.
  14. Green, Cathleen M.; Novikova, Olga; Belfort, Marlene (2018-01-24). "The dynamic intein landscape of eukaryotes". Mobile DNA. 9 (1): 4. doi: 10.1186/s13100-018-0111-x . ISSN   1759-8753. PMC   5784728 . PMID   29416568.
  15. 1 2 Novikova, Olga; Topilina, Natalya; Belfort, Marlene (May 2014). "Enigmatic Distribution, Evolution, and Function of Inteins". Journal of Biological Chemistry. 289 (21): 14490–14497. doi: 10.1074/jbc.r114.548255 . ISSN   0021-9258. PMC   4031506 . PMID   24695741.
  16. Noren CJ, Wang J, Perler FB (2000). "Dissecting the chemistry of protein splicing and its applications". Angew Chem Int Ed Engl. 39 (3): 450–66. doi:10.1002/(sici)1521-3773(20000204)39:3<450::aid-anie450>3.3.co;2-6. PMID   10671234.
  17. 1 2 3 4 5 Nanda, A; Nasker, SS; Mehra, A; Panda, S; Nayak, S (16 December 2020). "Inteins in Science: Evolution to Application". Microorganisms. 8 (12): 2004. doi: 10.3390/microorganisms8122004 . PMC   7765530 . PMID   33339089.
  18. 1 2 3 Tori, K; Perler, FB (April 2011). "Expanding the definition of class 3 inteins and their proposed phage origin". Journal of Bacteriology. 193 (8): 2035–41. doi: 10.1128/JB.01407-10 . PMC   3133030 . PMID   21317331.
  19. Anraku, Y; Mizutani, R; Satow, Y (2005). "Protein splicing: its discovery and structural insight into novel chemical mechanisms". IUBMB Life. 57 (8): 563–74. doi: 10.1080/15216540500215499 . PMID   16118114.
  20. Kelley, Danielle S.; Lennon, Christopher W.; Li, Zhong; Miller, Michael R.; Banavali, Nilesh K.; Li, Hongmin; Belfort, Marlene (19 October 2018). "Mycobacterial DnaB helicase intein as oxidative stress sensor". Nature Communications. 9 (1): 4363. Bibcode:2018NatCo...9.4363K. doi: 10.1038/s41467-018-06554-x . PMC   6195587 . PMID   30341292.
  21. Wu, H.; Hu, Z.; Liu, X. Q. (1998). "Protein trans-splicing by a split intein encoded in a split DnaE gene of Synechocystis sp. PCC6803". Proceedings of the National Academy of Sciences of the United States of America. 95 (16): 9226–9231. Bibcode:1998PNAS...95.9226W. doi: 10.1073/pnas.95.16.9226 . PMC   21320 . PMID   9689062.
  22. Schwarzer D, Cole PA (2005). "Protein semisynthesis and expressed protein ligation: chasing a protein's tail". Curr Opin Chem Biol. 9 (6): 561–9. doi:10.1016/j.cbpa.2005.09.018. PMID   16226484.
  23. Muralidharan V, Muir TW (2006). "Protein ligation: an enabling technology for the biophysical analysis of proteins". Nat. Methods. 3 (6): 429–38. doi:10.1038/nmeth886. PMID   16721376. S2CID   12550693.
  24. de Grey, Aubrey D.N.J (2000). "Mitochondrial gene therapy: an arena for the biomedical use of inteins". Trends in Biotechnology. 18 (9): 394–399. doi:10.1016/S0167-7799(00)01476-1. ISSN   0167-7799. PMID   10942964.
  25. 1 2 Chong, Shaorong; Mersha, Fana B; Comb, Donald G; Scott, Melissa E; Landry, David; Vence, Luis M; Perler, Francine B; Benner, Jack; Kucera, Rebecca B; Hirvonen, Christine A; Pelletier, John J; Paulus, Henry; Xu, Ming-Qun (1997). "Single-column purification of free recombinant proteins using a self-cleavable affinity tag derived from a protein splicing element". Gene. 192 (2): 271–281. doi:10.1016/S0378-1119(97)00105-4. ISSN   0378-1119. PMID   9224900.
  26. 1 2 Fong, Baley A; Wood, David W (2010). "Expression and purification of ELP-intein-tagged target proteins in high cell density E. coli fermentation". Microbial Cell Factories. 9 (1): 77. doi: 10.1186/1475-2859-9-77 . ISSN   1475-2859. PMC   2978133 . PMID   20959011.
  27. Xing, Lei; Wu, Wei; Zhou, Bihong; Lin, Zhanglin (2011). "Streamlined protein expression and purification using cleavable self-aggregating tags". Microbial Cell Factories. 10 (1): 42. doi: 10.1186/1475-2859-10-42 . ISSN   1475-2859. PMC   3124420 . PMID   21631955.
  28. Shah, Neel H.; Muir, Tom W. (2013-12-24). "Inteins: nature's gift to protein chemists". Chemical Science. 5 (2): 446–461. doi:10.1039/C3SC52951G. ISSN   2041-6539. PMC   3949740 . PMID   24634716.
  29. Huet, Gaëlle; Castaing, Jean-Philippe; Fournier, Didier; Daffé, Mamadou; Saves, Isabelle (May 2006). "Protein Splicing of SufB Is Crucial for the Functionality of the Mycobacterium tuberculosis SUF Machinery". Journal of Bacteriology. 188 (9): 3412–3414. doi:10.1128/JB.188.9.3412-3414.2006. ISSN   0021-9193. PMC   1447444 . PMID   16621837.
  30. Chan, Hon; Pearson, C. Seth; Green, Cathleen M.; Li, Zhong; Zhang, Jing; Belfort, Georges; Shekhtman, Alex; Li, Hongmin; Belfort, Marlene (October 2016). "Exploring Intein Inhibition by Platinum Compounds as an Antimicrobial Strategy". Journal of Biological Chemistry. 291 (43): 22661–22670. doi: 10.1074/jbc.m116.747824 . ISSN   0021-9258. PMC   5077202 . PMID   27609519.
  31. Li, Zhong; Tharappel, Anil Mathew; Xu, Jimin; Lang, Yuekun; Green, Cathleen M.; Zhang, Jing; Lin, Qishan; Chaturvedi, Sudha; Zhou, Jia; Belfort, Marlene; Li, Hongmin (2021-01-12). "Small-molecule inhibitors for the Prp8 intein as antifungal agents". Proceedings of the National Academy of Sciences. 118 (2): e2008815118. Bibcode:2021PNAS..11808815L. doi: 10.1073/pnas.2008815118 . ISSN   0027-8424. PMC   7812778 . PMID   33397721.

Further reading