To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Group I catalytic intron

From Wikipedia, the free encyclopedia

Group I catalytic intron
Predicted secondary structure and sequence conservation of Group I catalytic intron
Identifiers
SymbolIntron_gpI
RfamRF00028
Other data
RNA typeIntron
Domain(s)Eukaryota; Bacteria; Viruses
GOGO:0000372
SOSO:0000587
PDB structuresPDBe

Group I introns are large self-splicing ribozymes. They catalyze their own excision from mRNA, tRNA and rRNA precursors in a wide range of organisms.[1][2][3] The core secondary structure consists of nine paired regions (P1-P9).[4] These fold to essentially two domains – the P4-P6 domain (formed from the stacking of P5, P4, P6 and P6a helices) and the P3-P9 domain (formed from the P8, P3, P7 and P9 helices).[2] The secondary structure mark-up for this family represents only this conserved core. Group I introns often have long open reading frames inserted in loop regions.

YouTube Encyclopedic

  • 1/5
    Views:
    18 591
    33 067
    138 969
    8 199
    7 232
  • Types I and II Self-Splicing (Autocatalytic) Introns in mRNA
  • Group I Intron Splicing (part 1)
  • Splicing
  • Group I Intron Splicing (part 2)
  • RNA- Splicing (Group-I and II introns)

Transcription

Catalysis

Splicing of group I introns is processed by two sequential transesterification reactions.[3] The exogenous guanosine or guanosine nucleotide (exoG) first docks onto the active G-binding site located in P7, and its 3'-OH is aligned to attack the phosphodiester bond at the 5' splice site located in P1, resulting in a free 3'-OH group at the upstream exon and the exoG being attached to the 5' end of the intron. Then the terminal G (omega G) of the intron swaps the exoG and occupies the G-binding site to organize the second ester-transfer reaction: the 3'-OH group of the upstream exon in P1 is aligned to attack the 3' splice site in P10, leading to the ligation of the adjacent upstream and downstream exons and release of the catalytic intron.

Two-metal-ion mechanism seen in protein polymerases and phosphatases was proposed to be used by group I and group II introns to process the phosphoryl transfer reactions,[5] which was unambiguously proven by a high-resolution structure of the Azoarcus group I intron in 2006.[6]

A 3D representation of the Group I catalytic intron. This view shows the active site in the crystal structure of the Tetrahymena ribozyme.[7]
A 3D representation of the Group I catalytic intron. This is the crystal structure of a phage Twort group I ribozyme-product complex.[8]
A 3D representation of the Group I catalytic intron. This is the structure of the Tetrahymena ribozyme with a base triple sandwich and metal ion at the active site.[9]

Intron folding

Since the early 1990s, scientists started to study how the group I intron achieves its native structure in vitro, and some mechanisms of RNA folding have been appreciated thus far.[10] It is agreed that the tertiary structure is folded after the formation of the secondary structure. During folding, RNA molecules are rapidly populated into different folding intermediates, the intermediates containing native interactions are further folded into the native structure through a fast folding pathway, while those containing non-native interactions are trapped in metastable or stable non-native conformations, and the process of conversion to the native structure occurs very slowly. It is evident that group I introns differing in the set of peripheral elements display different potentials in entering the fast folding pathway. Meanwhile, cooperative assembly of the tertiary structure is important for the folding of the native structure. Nevertheless, folding of group I introns in vitro encounters both thermodynamic and kinetic challenges. A few RNA binding proteins and chaperones have been shown to promote the folding of group I introns in vitro and in bacteria by stabilizing the native intermediates, and by destabilizing the non-native structures, respectively.

Distribution, phylogeny and mobility

Group I introns are distributed in bacteria, lower eukaryotes and higher plants. However, their occurrence in bacteria seems to be more sporadic than in lower eukaryotes, and they have become prevalent in higher plants. The genes that group I introns interrupt differ significantly: They interrupt rRNA, mRNA and tRNA genes in bacterial genomes, as well as in mitochondrial and chloroplast genomes of lower eukaryotes, but only invade rRNA genes in the nuclear genome of lower eukaryotes. In higher plants, these introns seem to be restricted to a few tRNA and mRNA genes of the chloroplasts and mitochondria.

Group I introns are also found inserted into genes of a wide variety of bacteriophages of Gram-positive bacteria.[11] However, their distribution in the phage of Gram-negative bacteria is mainly limited to the T4, T-even and T7-like bacteriophages.[11][12][13][14]

Both intron-early and intron-late theories have found evidences in explaining the origin of group I introns. Some group I introns encode homing endonuclease (HEG), which catalyzes intron mobility. It is proposed that HEGs move the intron from one location to another, from one organism to another and thus account for the wide spreading of the selfish group I introns. No biological role has been identified for group I introns thus far except for splicing of themselves from the precursor to prevent the death of the host that they live by. A small number of group I introns are also found to encode a class of proteins called maturases that facilitate the intron splicing.

See also

References

  1. ^ Nielsen H, Johansen SD (2009). "Group I introns: Moving in new directions". RNA Biol. 6 (4): 375–83. doi:10.4161/rna.6.4.9334. PMID 19667762. Retrieved 2010-07-15.
  2. ^ a b Cate JH, Gooding AR, Podell E, et al. (September 1996). "Crystal structure of a group I ribozyme domain: principles of RNA packing". Science. 273 (5282): 1678–85. Bibcode:1996Sci...273.1678C. doi:10.1126/science.273.5282.1678. PMID 8781224. S2CID 38185676.
  3. ^ a b Cech TR (1990). "Self-splicing of group I introns". Annu. Rev. Biochem. 59: 543–68. doi:10.1146/annurev.bi.59.070190.002551. PMID 2197983.
  4. ^ Woodson SA (June 2005). "Structure and assembly of group I introns". Curr. Opin. Struct. Biol. 15 (3): 324–30. doi:10.1016/j.sbi.2005.05.007. PMID 15922592.
  5. ^ Steitz, TA; Steitz JA (1993). "A general two-metal-ion mechanism for catalytic RNA". Proc Natl Acad Sci USA. 90 (14): 6498–6502. Bibcode:1993PNAS...90.6498S. doi:10.1073/pnas.90.14.6498. PMC 46959. PMID 8341661.
  6. ^ Stahley, MR; Strobel SA (2006). "RNA splicing: group I intron crystal structures reveal the basis of splice site selection and metal ion catalysis". Curr Opin Struct Biol. 16 (3): 319–326. doi:10.1016/j.sbi.2006.04.005. PMID 16697179.
  7. ^ Golden BL, Gooding AR, Podell ER, Cech TR (1998). "A preorganized active site in the crystal structure of the Tetrahymena ribozyme". Science. 282 (5387): 259–64. Bibcode:1998Sci...282..259G. doi:10.1126/science.282.5387.259. PMID 9841391.
  8. ^ Golden BL, Kim H, Chase E (2005). "Crystal structure of a phage Twort group I ribozyme-product complex". Nat Struct Mol Biol. 12 (1): 82–9. doi:10.1038/nsmb868. PMID 15580277. S2CID 33369317.
  9. ^ Guo F, Gooding AR, Cech TR (2004). "Structure of the Tetrahymena ribozyme: base triple sandwich and metal ion at the active site". Mol Cell. 16 (3): 351–62. doi:10.1016/j.molcel.2004.10.003. PMID 15525509.
  10. ^ Brion P, Westhof E (1997). "Hierarchy and dynamics of RNA folding". Annu Rev Biophys Biomol Struct. 26: 113–37. doi:10.1146/annurev.biophys.26.1.113. PMID 9241415.
  11. ^ a b Edgell DR, Belfort M, Shub DA (October 2000). "Barriers to intron promiscuity in bacteria". J. Bacteriol. 182 (19): 5281–9. doi:10.1128/jb.182.19.5281-5289.2000. PMC 110968. PMID 10986228.
  12. ^ Sandegren L, Sjöberg BM (May 2004). "Distribution, sequence homology, and homing of group I introns among T-even-like bacteriophages: evidence for recent transfer of old introns". J. Biol. Chem. 279 (21): 22218–27. doi:10.1074/jbc.M400929200. PMID 15026408.
  13. ^ Bonocora RP, Shub DA (December 2004). "A self-splicing group I intron in DNA polymerase genes of T7-like bacteriophages". J. Bacteriol. 186 (23): 8153–5. doi:10.1128/JB.186.23.8153-8155.2004. PMC 529087. PMID 15547290.
  14. ^ Lee CN, Lin JW, Weng SF, Tseng YH (December 2009). "Genomic characterization of the intron-containing T7-like phage phiL7 of Xanthomonas campestris". Appl. Environ. Microbiol. 75 (24): 7828–37. Bibcode:2009ApEnM..75.7828L. doi:10.1128/AEM.01214-09. PMC 2794104. PMID 19854925.

Further reading

External links

This page was last edited on 15 August 2023, at 05:26
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.