Psychology Wiki
Register
Advertisement

Assessment | Biopsychology | Comparative | Cognitive | Developmental | Language | Individual differences | Personality | Philosophy | Social |
Methods | Statistics | Clinical | Educational | Industrial | Professional items | World psychology |

Biological: Behavioural genetics · Evolutionary psychology · Neuroanatomy · Neurochemistry · Neuroendocrinology · Neuroscience · Psychoneuroimmunology · Physiological Psychology · Psychopharmacology (Index, Outline)


File:Pre-mRNA-1ysv.png-tubes.png

A hairpin loop from a pre-mRNA. Highlighted are the bases (light green) and backbone (sky blue).

Ribonucleic acid (RNA) is a biologically important type of molecule that consists of a long chain of nucleotide units. Each nucleotide consists of a nitrogenous base, a ribose sugar, and a phosphate. RNA is very similar to DNA, but differs in a few important structural details: in the cell, RNA is usually single-stranded, while DNA is usually double-stranded; RNA nucleotides contain ribose while DNA contains deoxyribose (a type of ribose that lacks one oxygen atom); and RNA has the base uracil rather than thymine that is present in DNA.

RNA is transcribed from DNA by enzymes called RNA polymerases and is generally further processed by other enzymes. RNA is central to protein synthesis. Here, a type of RNA called messenger RNA carries information from DNA to structures called ribosomes. These ribosomes are made from proteins and ribosomal RNAs, which come together to form a molecular machine that can read messenger RNAs and translate the information they carry into proteins. There are many RNAs with other roles – in particular regulating which genes are expressed, but also as the genomes of most viruses.

Structure[]

File:Piwi-siRNA-basepairing.png

Watson-Crick base pairs in a siRNA (hydrogen atoms are not shown)

Each nucleotide in RNA contains a ribose sugar, with carbons numbered 1' through 5'. A base is attached to the 1' position, generally adenine (A), cytosine (C), guanine (G) or uracil (U). Adenine and guanine are purines, cytosine and uracil are pyrimidines. A phosphate group is attached to the 3' position of one ribose and the 5' position of the next. The phosphate groups have a negative charge each at physiological pH, making RNA a charged molecule (polyanion). The bases may form hydrogen bonds between cytosine and guanine, between adenine and uracil and between guanine and uracil.[1] However other interactions are possible, such as a group of adenine bases binding to each other in a bulge,[2] or the GNRA tetraloop that has a guanine–adenine base-pair.[1]

File:RNA chemical structure.GIF

Chemical structure of RNA

An important structural feature of RNA that distinguishes it from DNA is the presence of a hydroxyl group at the 2' position of the ribose sugar. The presence of this functional group causes the helix to adopt the A-form geometry rather than the B-form most commonly observed in DNA.[3] This results in a very deep and narrow major groove and a shallow and wide minor groove.[4] A second consequence of the presence of the 2'-hydroxyl group is that in conformationally flexible regions of an RNA molecule (that is, not involved in formation of a double helix), it can chemically attack the adjacent phosphodiester bond to cleave the backbone.[5]

File:Ciliate telomerase RNA.JPG

Secondary structure of a telomerase RNA.

RNA is transcribed with only four bases (adenine, cytosine, guanine and uracil),[6] but there are numerous modified bases and sugars in mature RNAs. Pseudouridine (Ψ), in which the linkage between uracil and ribose is changed from a C–N bond to a C–C bond, and ribothymidine (T), are found in various places (most notably in the TΨC loop of tRNA).[7] Another notable modified base is hypoxanthine, a deaminated adenine base whose nucleoside is called inosine (I). Inosine plays a key role in the wobble hypothesis of the genetic code.[8] There are nearly 100 other naturally occurring modified nucleosides,[9] of which pseudouridine and nucleosides with 2'-O-methylribose are the most common.[10] The specific roles of many of these modifications in RNA are not fully understood. However, it is notable that in ribosomal RNA, many of the post-transcriptional modifications occur in highly functional regions, such as the peptidyl transferase center and the subunit interface, implying that they are important for normal function.[11]

The functional form of single stranded RNA molecules, just like proteins, frequently requires a specific tertiary structure. The scaffold for this structure is provided by secondary structural elements which are hydrogen bonds within the molecule. This leads to several recognizable "domains" of secondary structure like hairpin loops, bulges and internal loops.[12] Since RNA is charged, metal ions such as Mg2+ are needed to stabilise many secondary and tertiary structures.[13]

Comparison with DNA[]

RNA and DNA are both nucleic acids, but differ in three main ways. First, unlike DNA which is double-stranded, RNA is a single-stranded molecule in most of its biological roles and has a much shorter chain of nucleotides. Second, while DNA contains deoxyribose, RNA contains ribose (there is no hydroxyl group attached to the pentose ring in the 2' position in DNA). These hydroxyl groups make RNA less stable than DNA because it is more prone to hydrolysis. Third, the complementary base to adenine is not thymine, as it is in DNA, but rather uracil, which is an unmethylated form of thymine.[14]

Ribosome 50s

The 50S ribosomal subunit. RNA is in orange, protein in blue. The active site is in the middle (red).

Like DNA, most biologically active RNAs, including mRNA, tRNA, rRNA, snRNAs and other non-coding RNAs, contain self-complementary sequences that allow parts of the RNA to fold and pair with itself to form double helices. Structural analysis of these RNAs have revealed that they are highly structured. Unlike DNA, their structures do not consist of long double helices but rather collections of short helices packed together into structures akin to proteins. In this fashion, RNAs can achieve chemical catalysis, like enzymes.[15] For instance, determination of the structure of the ribosome—an enzyme that catalyzes peptide bond formation—revealed that its active site is composed entirely of RNA.[16]

Synthesis[]

Synthesis of RNA is usually catalyzed by an enzyme—RNA polymerase—using DNA as a template, a process known as transcription. Initiation of transcription begins with the binding of the enzyme to a promoter sequence in the DNA (usually found "upstream" of a gene). The DNA double helix is unwound by the helicase activity of the enzyme. The enzyme then progresses along the template strand in the 3’ to 5’ direction, synthesizing a complementary RNA molecule with elongation occurring in the 5’ to 3’ direction. The DNA sequence also dictates where termination of RNA synthesis will occur.[17]

RNAs are often modified by enzymes after transcription. For example, a poly(A) tail and a 5' cap are added to eukaryotic pre-mRNA and introns are removed by the spliceosome.

There are also a number of RNA-dependent RNA polymerases that use RNA as their template for synthesis of a new strand of RNA. For instance, a number of RNA viruses (such as poliovirus) use this type of enzyme to replicate their genetic material.[18] Also, RNA-dependent RNA polymerase is part of the RNA interference pathway in many organisms.[19]

Types of RNA[]

See also: List of RNAs

Overview[]

File:Full length hammerhead ribozyme.png

Structure of a hammerhead ribozyme, a ribozyme that cuts RNA

Messenger RNA (mRNA) is the RNA that carries information from DNA to the ribosome, the sites of protein synthesis (translation) in the cell. The coding sequence of the mRNA determines the amino acid sequence in the protein that is produced.[20] Many RNAs do not code for protein however (about 97% of the transcriptial output is non-protein-coding in eukaryotes [21][22][23][24]).

These so-called non-coding RNAs ("ncRNA") can be encoded by their own genes (RNA genes), but can also derive from mRNA introns.[25] The most prominent examples of non-coding RNAs are transfer RNA (tRNA) and ribosomal RNA (rRNA), both of which are involved in the process of translation.[14] There are also non-coding RNAs involved in gene regulation, RNA processing and other roles. Certain RNAs are able to catalyse chemical reactions such as cutting and ligating other RNA molecules,[26] and the catalysis of peptide bond formation in the ribosome;[16] these are known as ribozymes.

In translation[]

Messenger RNA (mRNA) carries information about a protein sequence to the ribosomes, the protein synthesis factories in the cell. It is coded so that every three nucleotides (a codon) correspond to one amino acid. In eukaryotic cells, once precursor mRNA (pre-mRNA) has been transcribed from DNA, it is processed to mature mRNA. This removes its introns—non-coding sections of the pre-mRNA. The mRNA is then exported from the nucleus to the cytoplasm, where it is bound to ribosomes and translated into its corresponding protein form with the help of tRNA. In prokaryotic cells, which do not have nucleus and cytoplasm compartments, mRNA can bind to ribosomes while it is being transcribed from DNA. After a certain amount of time the message degrades into its component nucleotides with the assistance of ribonucleases.[20]

Transfer RNA (tRNA) is a small RNA chain of about 80 nucleotides that transfers a specific amino acid to a growing polypeptide chain at the ribosomal site of protein synthesis during translation. It has sites for amino acid attachment and an anticodon region for codon recognition that binds to a specific sequence on the messenger RNA chain through hydrogen bonding.[25]

Ribosomal RNA (rRNA) is the catalytic component of the ribosomes. Eukaryotic ribosomes contain four different rRNA molecules: 18S, 5.8S, 28S and 5S rRNA. Three of the rRNA molecules are synthesized in the nucleolus, and one is synthesized elsewhere. In the cytoplasm, ribosomal RNA and protein combine to form a nucleoprotein called a ribosome. The ribosome binds mRNA and carries out protein synthesis. Several ribosomes may be attached to a single mRNA at any time.[20] rRNA is extremely abundant and makes up 80% of the 10 mg/ml RNA found in a typical eukaryotic cytoplasm.[27]

Transfer-messenger RNA (tmRNA) is found in many bacteria and plastids. It tags proteins encoded by mRNAs that lack stop codons for degradation and prevents the ribosome from stalling.[28]

Regulatory RNAs[]

Several types of RNA can downregulate gene expression by being complementary to a part of an mRNA or a gene's DNA. MicroRNAs (miRNA; 21-22 nt) are found in eukaryotes and act through RNA interference (RNAi), where an effector complex of miRNA and enzymes can break down mRNA which the miRNA is complementary to, block the mRNA from being translated, or accelerate its degradation.[29][30] While small interfering RNAs (siRNA; 20-25 nt) are often produced by breakdown of viral RNA, there are also endogenous sources of siRNAs.[31][32] siRNAs act through RNA interference in a fashion similar to miRNAs. Some miRNAs and siRNAs can cause genes they target to be methylated, thereby decreasing or increasing transcription of those genes.[33][34][35] Animals have Piwi-interacting RNAs (piRNA; 29-30 nt) which are active in germline cells and are thought to be a defense against transposons and play a role in gametogenesis.[36][37] All prokaryotes have CRISPR RNAs, a regulatory system analogous to RNA interference.[38] Antisense RNAs are widespread; most downregulate a gene, but a few are activators of transcription.[39] One way antisense RNA can act is by binding to an mRNA, forming double-stranded RNA that is enzymatically degraded.[40] There are many long noncoding RNAs that regulate genes in eukaryotes,[41] one such RNA is Xist which coats one X chromosome in female mammals and inactivates it.[42]

An mRNA may contain regulatory elements itself, such as riboswitches, in the 5' untranslated region or 3' untranslated region; these cis-regulatory elements regulate the activity of that mRNA.[43] The untranslated regions can also contain elements that regulate other genes.[44]

In RNA processing[]

File:Uridine to pseudouridine.GIF

Uridine to pseudouridine is a common RNA modification.

Many RNAs are involved in modifying other RNAs. Introns are spliced out of pre-mRNA by spliceosomes, which contain several small nuclear RNAs (snRNA),[14] or the introns can be ribozymes that are spliced by themselves.[45] RNA can also be altered by having its nucleotides modified to other nucleotides than A, C, G and U. In eukaryotes, modifications of RNA nucleotides are generally directed by small nucleolar RNAs (snoRNA; 60-300 nt),[25] found in the nucleolus and cajal bodies. snoRNAs associate with enzymes and guide them to a spot on an RNA by basepairing to that RNA. These enzymes then perform the nucleotide modification. rRNAs and tRNAs are extensively modified, but snRNAs and mRNAs can also be the target of base modification.[46][47]

RNA genomes[]

Like DNA, RNA can carry genetic information. RNA viruses have genomes composed of RNA, and a variety of proteins encoded by that genome. The viral genome is replicated by some of those proteins, while other proteins protect the genome as the virus particle moves to a new host cell. Viroids are another group of pathogens, but they consist only of RNA, do not encode any protein and are replicated by a host plant cell's polymerase.[48]

In reverse transcription[]

Reverse transcribing viruses replicate their genomes by reverse transcribing DNA copies from their RNA; these DNA copies are then transcribed to new RNA. Retrotransposons also spread by copying DNA and RNA from one another,[49] and telomerase contains an RNA that is used as template for building the ends of eukaryotic chromosomes.[50]

Double-stranded RNA[]

Double-stranded RNA (dsRNA) is RNA with two complementary strands, similar to the DNA found in all cells. dsRNA forms the genetic material of some viruses (double-stranded RNA viruses). Double-stranded RNA such as viral RNA or siRNA can trigger RNA interference in eukaryotes, as well as interferon response in vertebrates.[51][52][53]

Discovery[]

Nucleic acids were discovered in 1868 by Friedrich Miescher, who called the material 'nuclein' since it was found in the nucleus.[54] It was later discovered that prokaryotic cells, which do not have a nucleus, also contain nucleic acids. The role of RNA in protein synthesis was suspected already in 1939.[55] Severo Ochoa won the 1959 Nobel Prize in Medicine after he discovered how RNA is synthesized.[56] The sequence of the 77 nucleotides of a yeast tRNA was found by Robert W. Holley in 1965,[57] winning Holley the 1968 Nobel Prize in Medicine. In 1967, Carl Woese realized RNA can be catalytic and proposed that the earliest forms of life relied on RNA both to carry genetic information and to catalyze biochemical reactions—an RNA world.[58][59] In 1976, Walter Fiers and his team determined the first complete nucleotide sequence of an RNA virus genome, that of bacteriophage MS2.[60] In 1990 it was found in petunia that introduced genes can silence similar genes of the plant's own, now known to be a result of RNA interference.[61][62] At about the same time, 22 nt long RNAs, now called microRNAs, were found to have a role in the development of C. elegans.[63] The discovery of gene regulatory RNAs has led to attempts to develop drugs made of RNA, such as siRNA, to silence genes.[64]

See also[]

Template:MolBioGeneExp

  • Genetics
  • Molecular biology
  • Oligonucleotide synthesis
  • Quantification of nucleic acids
  • RNA Ontology Consortium
  • Sequence profiling tool
  • RNA extraction
  • RNA structure prediction

References[]

  1. 1.0 1.1 Lee JC, Gutell RR (2004). Diversity of base-pair conformations and their occurrence in rRNA structure and RNA structural motifs. J. Mol. Biol. 344 (5): 1225–49.
  2. Barciszewski J, Frederic B, Clark C (1999). RNA biochemistry and biotechnology, 73–87, Springer.
  3. Salazar M, Fedoroff OY, Miller JM, Ribeiro NS, Reid BR (1992). The DNA strand in DNAoRNA hybrid duplexes is neither B-form nor A-form in solution. Biochemistry 32 (16): 4207–15.
  4. Hermann T, Patel DJ (2000). RNA bulges as architectural and recognition motifs. Structure 8 (3): R47–R54.
  5. Mikkola S, Nurmi K, Yousefi-Salakdeh E, Strömberg R, Lönnberg H (1999). The mechanism of the metal ion promoted cleavage of RNA phosphodiester bonds involves a general acid catalysis by the metal aquo ion on the departure of the leaving group. Perkin transactions 2: 1619–26.
  6. Jankowski JAZ, Polak JM (1996). Clinical gene analysis and manipulation: Tools, techniques and troubleshooting, 14, Cambridge University Press.
  7. Yu Q, Morrow CD (2001). Identification of critical elements in the tRNA acceptor stem and TΨC loop necessary for human immunodeficiency virus type 1 infectivity. J Virol. 75 (10): 4902–6.
  8. Elliott MS, Trewyn RW (1983). Inosine biosynthesis in transfer RNA by an enzymatic insertion of hypoxanthine. J. Biol. Chem. 259 (4): 2407–10.
  9. Söll D, RajBhandary U (1995). TRNA: Structure, biosynthesis, and function, 165, ASM Press.
  10. Kiss T (2001). Small nucleolar RNA-guided post-transcriptional modification of cellular RNAs. The EMBO Journal 20: 3617–22.
  11. King TH, Liu B, McCully RR, Fournier MJ (2002). Ribosome structure and activity are altered in cells lacking snoRNPs that form pseudouridines in the peptidyl transferase center. Molecular Cell 11 (2): 425–35.
  12. Mathews DH, Disney MD, Childs JL, Schroeder SJ, Zuker M, Turner DH (2004). Incorporating chemical modification constraints into a dynamic programming algorithm for prediction of RNA secondary structure. Proc. Natl. Acad. Sci. USA 101 (19): 7287–92.
  13. Tan ZJ, Chen SJ (2008). Salt dependence of nucleic acid hairpin stability. Biophys. J. 95 (2): 738–52.
  14. 14.0 14.1 14.2 Berg JM, Tymoczko JL, Stryer L (2002). Biochemistry, 5th, 118–19, 781–808, WH Freeman and Company.
  15. Higgs PG (2000). RNA secondary structure: physical and computational aspects. Quarterly Reviews of Biophysics 33: 199–253.
  16. 16.0 16.1 Nissen P, Hansen J, Ban N, Moore PB, Steitz TA (2000). The structural basis of ribosome activity in peptide bond synthesis. Science 289 (5481): 920–30.
  17. Nudler E, Gottesman ME (2002). Transcription termination and anti-termination in E. coli. Genes to Cells 7: 755–68.
  18. Jeffrey L Hansen, Alexander M Long, Steve C Schultz (1997). Structure of the RNA-dependent RNA polymerase of poliovirus. Structure 5 (8): 1109–22.
  19. Ahlquist P (2002). RNA-Dependent RNA Polymerases, Viruses, and RNA Silencing. Science 296 (5571): 1270–73.
  20. 20.0 20.1 20.2 Cooper GC, Hausman RE (2004). The Cell: A Molecular Approach, 3rd, 261–76, 297, 339–44, Sinauer.
  21. Mattick JS, Gagen MJ (1 September 2001). The evolution of controlled multitasked gene networks: the role of introns and other noncoding RNAs in the development of complex organisms. Mol. Biol. Evol. 18 (9): 1611–30.
  22. Mattick, J.S. (2001). “Noncoding RNAs: the architects of eukaryotic complexity”. EMBO Reports, 2(11), 986-991. [1]
  23. Mattick, J.S. (2003). “Challenging the dogma: The hidden layer of non-protein-coding RNAs on complex organisms” Bioessays. 25, 930-939. [2]
  24. Mattick, J.S. (2004). “The hidden genetic program of complex organisms” Scientific American. 291(4), 30-37. [3]
  25. 25.0 25.1 25.2 Wirta W (2006). Mining the transcriptome – methods and applications, Stockholm: School of Biotechnology, Royal Institute of Technology.
  26. Rossi JJ (2004). Ribozyme diagnostics comes of age. Chemistry & Biology 11 (7): 894–95.
  27. Kampers T, Friedhoff P, Biernat J, Mandelkow E-M, Mandelkow E (1996). RNA stimulates aggregation of microtubule-associated protein tau into Alzheimer-like paired helical filaments. FEBS Letters 399 (3): 104D.
  28. Gueneau de Novoa P, Williams KP (2004). The tmRNA website: reductive evolution of tmRNA in plastids and other endosymbionts. Nucleic Acids Res. 32 (Database issue): D104–8.
  29. Wu L, Belasco JG (January 2008). Let me count the ways: mechanisms of gene regulation by miRNAs and siRNAs. Mol. Cell 29 (1): 1–7.
  30. Matzke MA, Matzke AJM (2004). Planting the seeds of a new paradigm. PLoS Biology 2 (5): e133.
  31. Vazquez F, Vaucheret H, Rajagopalan R, Lepers C, Gasciolli V, Mallory AC, Hilbert J, Bartel DP, Crété P (2004). Endogenous trans-acting siRNAs regulate the accumulation of Arabidopsis mRNAs. Molecular Cell 16 (1): 69–79.
  32. Watanabe T, Totoki Y, Toyoda A, et al. (May 2008). Endogenous siRNAs from naturally formed dsRNAs regulate transcripts in mouse oocytes. Nature 453 (7194): 539–43.
  33. Sontheimer EJ, Carthew RW (July 2005). Silence from within: endogenous siRNAs and miRNAs. Cell 122 (1): 9–12.
  34. Doran G (2007). RNAi – Is one suffix sufficient?. Journal of RNAi and Gene Silencing 3 (1): 217–19.
  35. Pushparaj PN, Aarthi JJ, Kumar SD, Manikandan J (2008). RNAi and RNAa - The Yin and Yang of RNAome. Bioinformation 2 (6): 235–7.
  36. Horwich MD, Li C Matranga C, Vagin V, Farley G, Wang P, Zamore PD (2007). The Drosophila RNA methyltransferase, DmHen1, modifies germline piRNAs and single-stranded siRNAs in RISC. Current Biology 17: 1265–72.
  37. Girard A, Sachidanandam R, Hannon GJ, Carmell MA (2006). A germline-specific class of small RNAs binds mammalian Piwi proteins. Nature 442: 199–202.
  38. Makarova KS, Grishin NV, Shabalina SA, Wolf YI, Koonin EV (2006). A putative RNA-interference-based immune system in prokaryotes: computational analysis of the predicted enzymatic machinery, functional analogies with eukaryotic RNAi, and hypothetical mechanisms of action. Biol. Direct 1: 7.
  39. Wagner EG, Altuvia S, Romby P (2002). Antisense RNAs in bacteria and their genetic elements. Adv Genet. 46: 361–98.
  40. Gilbert SF (2003). Developmental Biology, 7th, 101–3, Sinauer.
  41. Amaral PP, Mattick JS (October 2008). Noncoding RNA in development. Mammalian genome : official journal of the International Mammalian Genome Society 19 (7-8): 454.
  42. Heard E, Mongelard F, Arnaud D, Chureau C, Vourc'h C, Avner P (1999). Human XIST yeast artificial chromosome transgenes show partial X inactivation center function in mouse embryonic stem cells. Proc. Natl. Acad. Sci. USA 96 (12): 6841–46.
  43. Batey RT (2006). Structures of regulatory elements in mRNAs. Curr. Opin. Struct. Biol. 16 (3): 299–306.
  44. Scotto L, Assoian RK (June 1993). A GC-rich domain with bifunctional effects on mRNA and protein levels: implications for control of transforming growth factor beta 1 expression. Mol. Cell. Biol. 13 (6): 3588–97.
  45. Steitz TA, Steitz JA (1993). A general two-metal-ion mechanism for catalytic RNA. Proc. Natl. Acad. Sci. U.S.A. 90 (14): 6498–502.
  46. Xie J, Zhang M, Zhou T, Hua X, Tang L, Wu W (2007). Sno/scaRNAbase: a curated database for small nucleolar RNAs and cajal body-specific RNAs. Nucleic Acids Res. 35 (Database issue): D183–7.
  47. Omer AD, Ziesche S, Decatur WA, Fournier MJ, Dennis PP (2003). RNA-modifying machines in archaea. Molecular Microbiology 48 (3): 617–29.
  48. Daròs JA, Elena SF, Flores R (2006). Viroids: an Ariadne's thread into the RNA labyrinth. EMBO Rep. 7 (6): 593–8.
  49. Kalendar R, Vicient CM, Peleg O, Anamthawat-Jonsson K, Bolshoy A, Schulman AH (2004). Large retrotransposon derivatives: abundant, conserved but nonautonomous retroelements of barley and related genomes. Genetics 166 (3): D339.
  50. Podlevsky JD, Bley CJ, Omana RV, Qi X, Chen JJ (2008). The telomerase database. Nucleic Acids Res. 36 (Database issue): D339–43.
  51. Blevins T et al. (2006). Four plant Dicers mediate viral small RNA biogenesis and DNA virus induced silencing. Nucleic Acids Res 34 (21): 6233–46.
  52. Jana S, Chakraborty C, Nandi S, Deb JK (2004). RNA interference: potential therapeutic targets. Appl. Microbiol. Biotechnol. 65 (6): 649–57.
  53. Schultz U, Kaspers B, Staeheli P (2004). The interferon system of non-mammalian vertebrates. Dev. Comp. Immunol. 28 (5): 499–508.
  54. Dahm R (2005). Friedrich Miescher and the discovery of DNA. Developmental Biology 278 (2): 274–88.
  55. Caspersson T, Schultz J (1939). Pentose nucleotides in the cytoplasm of growing tissues. Nature 143: 602–3.
  56. Ochoa S (1959). Enzymatic synthesis of ribonucleic acid. Nobel Lecture.
  57. Holley RW et al. (1965). Structure of a ribonucleic acid. Science 147 (1664): 1462–65.
  58. Siebert S (2006). Common sequence structure properties and stable regions in RNA secondary structures. Dissertation, Albert-Ludwigs-Universität, Freiburg im Breisgau.
  59. Szathmáry E (1999). The origin of the genetic code: amino acids as cofactors in an RNA world. Trends Genet. 15 (6): 223–9.
  60. Fiers W et al. (1976). Complete nucleotide-sequence of bacteriophage MS2-RNA: primary and secondary structure of replicase gene. Nature 260 (5551): 500–7.
  61. Napoli C, Lemieux C, Jorgensen R (1990). Introduction of a chimeric chalcone synthase gene into petunia results in reversible co-suppression of homologous genes in trans. Plant Cell 2 (4): 279–89.
  62. Dafny-Yelin M, Chung SM, Frankman EL, Tzfira T (December 2007). pSAT RNA interference vectors: a modular series for multiple gene down-regulation in plants. Plant Physiol. 145 (4): 1272–81.
  63. Ruvkun G (2001). Glimpses of a tiny RNA world. Science 294 (5543): 797–99.
  64. Fichou Y, Férec C (2006). The potential of oligonucleotides for therapeutic applications. Trends in Biotechnology 24 (12): 563–70.

External links[]

Commons logo
Wikimedia Commons has media related to:

Template:RNA-footer


This page uses Creative Commons Licensed content from Wikipedia (view authors).
Advertisement