Polyadenylation

This is a good article. Click here for more information.
Source: Wikipedia, the free encyclopedia.
(Redirected from
Poly(A) tail
)

Typical structure of a mature eukaryotic mRNA

Polyadenylation is the addition of a poly(A) tail to an RNA transcript, typically a messenger RNA (mRNA). The poly(A) tail consists of multiple adenosine monophosphates; in other words, it is a stretch of RNA that has only adenine bases. In eukaryotes, polyadenylation is part of the process that produces mature mRNA for translation. In many bacteria, the poly(A) tail promotes degradation of the mRNA. It, therefore, forms part of the larger process of gene expression.

The process of polyadenylation begins as the

terminates. The 3′-most segment of the newly made pre-mRNA is first cleaved off by a set of proteins; these proteins then synthesize the poly(A) tail at the RNA's 3′ end. In some genes these proteins add a poly(A) tail at one of several possible sites. Therefore, polyadenylation can produce more than one transcript from a single gene (alternative polyadenylation), similar to alternative splicing.[1]

The poly(A) tail is important for the nuclear export, translation and stability of mRNA. The tail is shortened over time, and, when it is short enough, the mRNA is enzymatically degraded.[2] However, in a few cell types, mRNAs with short poly(A) tails are stored for later activation by re-polyadenylation in the cytosol.[3] In contrast, when polyadenylation occurs in bacteria, it promotes RNA degradation.[4] This is also sometimes the case for eukaryotic non-coding RNAs.[5][6]

mRNA molecules in both prokaryotes and eukaryotes have polyadenylated 3′-ends, with the prokaryotic poly(A) tails generally shorter and fewer mRNA molecules polyadenylated.[7]

Background on RNA

Chemical structure of RNA. The sequence of bases differs between RNA molecules.

RNAs are a type of large biological molecules, whose individual building blocks are called nucleotides. The name poly(A) tail (for polyadenylic acid tail)

transcribed) from a DNA template. By convention, RNA sequences are written in a 5′ to 3′ direction. The 5′ end is the part of the RNA molecule that is transcribed first, and the 3′ end is transcribed last. The 3′ end is also where the poly(A) tail is found on polyadenylated RNAs.[1][9]

Messenger RNA (mRNA) is RNA that has a coding region that acts as a template for protein synthesis (translation). The rest of the mRNA, the untranslated regions, tune how active the mRNA is.[10] There are also many RNAs that are not translated, called non-coding RNAs. Like the untranslated regions, many of these non-coding RNAs have regulatory roles.[11]

Nuclear polyadenylation

Function

In nuclear polyadenylation, a poly(A) tail is added to an RNA at the end of transcription. On mRNAs, the poly(A) tail protects the mRNA molecule from enzymatic degradation in the cytoplasm and aids in transcription termination, export of the mRNA from the nucleus, and translation.[2] Almost all eukaryotic mRNAs are polyadenylated,[12] with the exception of animal replication-dependent histone mRNAs.[13] These are the only mRNAs in eukaryotes that lack a poly(A) tail, ending instead in a stem-loop structure followed by a purine-rich sequence, termed histone downstream element, that directs where the RNA is cut so that the 3′ end of the histone mRNA is formed.[14]

Many eukaryotic non-coding RNAs are always polyadenylated at the end of transcription. There are small RNAs where the poly(A) tail is seen only in intermediary forms and not in the mature RNA as the ends are removed during processing, the notable ones being

X chromosome inactivation – a poly(A) tail is part of the mature RNA.[17]

Mechanism

Proteins involved:[12][18]

CPSF: cleavage/polyadenylation specificity factor

CstF
: cleavage stimulation factor
PAP: polyadenylate polymerase
PABII: polyadenylate binding protein 2
CFI: cleavage factor I
CFII: cleavage factor II

The

CPSF exist.[18][20] Two other proteins add specificity to the binding to an RNA: CstF and CFI. CstF binds to a GU-rich region further downstream of CPSF's site.[21] CFI recognises a third site on the RNA (a set of UGUAA sequences in mammals[22][23][24]) and can recruit CPSF even if the AAUAAA sequence is missing.[25][26] The polyadenylation signal – the sequence motif recognised by the RNA cleavage complex – varies between groups of eukaryotes. Most human polyadenylation sites contain the AAUAAA sequence,[21] but this sequence is less common in plants and fungi.[27]

The RNA is typically cleaved before transcription termination, as CstF also binds to RNA polymerase II.[28] Through a poorly understood mechanism (as of 2002), it signals for RNA polymerase II to slip off of the transcript.[29] Cleavage also involves the protein CFII, though it is unknown how.[30] The cleavage site associated with a polyadenylation signal can vary up to some 50 nucleotides.[31]

When the RNA is cleaved, polyadenylation starts, catalysed by polyadenylate polymerase.

Polyadenylate polymerase builds the poly(A) tail by adding adenosine monophosphate units from adenosine triphosphate to the RNA, cleaving off pyrophosphate.[32] Another protein, PAB2, binds to the new, short poly(A) tail and increases the affinity of polyadenylate polymerase for the RNA. When the poly(A) tail is approximately 250 nucleotides long the enzyme can no longer bind to CPSF and polyadenylation stops, thus determining the length of the poly(A) tail.[33][34] CPSF is in contact with RNA polymerase II, allowing it to signal the polymerase to terminate transcription.[35][36] When RNA polymerase II reaches a "termination sequence" (⁵'TTTATT3' on the DNA template and ⁵'AAUAAA3' on the primary transcript), the end of transcription is signaled.[37] The polyadenylation machinery is also physically linked to the spliceosome, a complex that removes introns from RNAs.[26]

Downstream effects

The poly(A) tail acts as the binding site for

40S ribosomal subunit.[41] However, a poly(A) tail is not required for the translation of all mRNAs.[42] Further, poly(A) tailing (oligo-adenylation) can determine the fate of RNA molecules that are usually not poly(A)-tailed (such as (small) non-coding (sn)RNAs etc.) and thereby induce their RNA decay.[43]

Deadenylation

In eukaryotic

In animals, poly(A) ribonuclease (

budding yeast and human cells, most notably the CCR4-Not complex.[49]

Cytoplasmic polyadenylation

There is polyadenylation in the cytosol of some animal cell types, namely in the

embryogenesis and in post-synaptic sites of nerve cells. This lengthens the poly(A) tail of an mRNA with a shortened poly(A) tail, so that the mRNA will be translated.[44][50] These shortened poly(A) tails are often less than 20 nucleotides, and are lengthened to around 80–150 nucleotides.[3]

In the early mouse embryo, cytoplasmic polyadenylation of maternal RNAs from the egg cell allows the cell to survive and grow even though transcription does not start until the middle of the 2-cell stage (4-cell stage in human).[51][52] In the brain, cytoplasmic polyadenylation is active during learning and could play a role in long-term potentiation, which is the strengthening of the signal transmission from a nerve cell to another in response to nerve impulses and is important for learning and memory formation.[3][53]

Cytoplasmic polyadenylation requires the RNA-binding proteins CPSF and CPEB, and can involve other RNA-binding proteins like Pumilio.[54] Depending on the cell type, the polymerase can be the same type of polyadenylate polymerase (PAP) that is used in the nuclear process, or the cytoplasmic polymerase GLD-2.[55]

Results of using different polyadenylation sites on the same gene

Alternative polyadenylation

Many protein-coding genes have more than one polyadenylation site, so a gene can code for several mRNAs that differ in their

3′ end.[27][56][57] The 3’ region of a transcript contains many polyadenylation signals (PAS). When more proximal (closer towards 5’ end) PAS sites are utilized, this shortens the length of the 3’ untranslated region (3' UTR) of a transcript.[58] Studies in both humans and flies have shown tissue specific APA. With neuronal tissues preferring distal PAS usage, leading to longer 3’ UTRs and testis tissues preferring proximal PAS leading to shorter 3’ UTRs.[59][60] Studies have shown there is a correlation between a gene's conservation level and its tendency to do alternative polyadenylation, with highly conserved genes exhibiting more APA. Similarly, highly expressed genes follow this same pattern.[61] Ribo-sequencing data (sequencing of only mRNAs inside ribosomes) has shown that mRNA isoforms with shorter 3’ UTRs are more likely to be translated.[58]

Since alternative polyadenylation changes the length of the 3' UTR,[62] it can also change which binding sites are available for microRNAs in the 3′ UTR.[19][63] MicroRNAs tend to repress translation and promote degradation of the mRNAs they bind to, although there are examples of microRNAs that stabilise transcripts.[64][65] Alternative polyadenylation can also shorten the coding region, thus making the mRNA code for a different protein,[66][67] but this is much less common than just shortening the 3′ untranslated region.[27]

The choice of poly(A) site can be influenced by extracellular stimuli and depends on the expression of the proteins that take part in polyadenylation.

TNF-α. These mRNAs then have longer half-lives and produce more of these proteins.[68] RNA-binding proteins other than those in the polyadenylation machinery can also affect whether a polyadenylation site is used,[70][71][72][73] as can DNA methylation near the polyadenylation signal.[74] In addition, numerous other components involved in transcription, splicing or other mechanisms regulating RNA biology can affect APA.[75]

Tagging for degradation in eukaryotes

For many

snoRNA, polyadenylation is a way of marking the RNA for degradation, at least in yeast.[76] This polyadenylation is done in the nucleus by the TRAMP complex, which maintains a tail that is around 4 nucleotides long to the 3′ end.[77][78] The RNA is then degraded by the exosome.[79] Poly(A) tails have also been found on human rRNA fragments, both the form of homopolymeric (A only) and heterpolymeric (mostly A) tails.[80]

In prokaryotes and organelles

Polyadenylation in bacteria helps polynucleotide phosphorylase degrade past secondary structure

In many bacteria, both mRNAs and non-coding RNAs can be polyadenylated. This poly(A) tail promotes degradation by the

secondary structure would otherwise block the 3′ end. Successive rounds of polyadenylation and degradation of the 3′ end by polynucleotide phosphorylase allows the degradosome to overcome these secondary structures. The poly(A) tail can also recruit RNases that cut the RNA in two.[81] These bacterial poly(A) tails are about 30 nucleotides long.[82]

In as different groups as animals and

trypanosomes, the mitochondria contain both stabilising and destabilising poly(A) tails. Destabilising polyadenylation targets both mRNA and noncoding RNAs. The poly(A) tails are 43 nucleotides long on average. The stabilising ones start at the stop codon, and without them the stop codon (UAA) is not complete as the genome only encodes the U or UA part. Plant mitochondria have only destabilising polyadenylation. Mitochondrial polyadenylation has never been observed in either budding or fission yeast.[83][84]

While many bacteria and mitochondria have polyadenylate polymerases, they also have another type of polyadenylation, performed by polynucleotide phosphorylase itself. This enzyme is found in bacteria,[85] mitochondria,[86] plastids[87] and as a constituent of the archaeal exosome (in those archaea that have an exosome).[88] It can synthesise a 3′ extension where the vast majority of the bases are adenines. Like in bacteria, polyadenylation by polynucleotide phosphorylase promotes degradation of the RNA in plastids[89] and likely also archaea.[83]

Evolution

Although polyadenylation is seen in almost all organisms, it is not universal.

last universal common ancestor of all living organisms, it is presumed, had some form of polyadenylation system.[82] A few organisms do not polyadenylate mRNA, which implies that they have lost their polyadenylation machineries during evolution. Although no examples of eukaryotes that lack polyadenylation are known, mRNAs from the bacterium Mycoplasma gallisepticum and the salt-tolerant archaean Haloferax volcanii lack this modification.[91][92]

The most ancient polyadenylating enzyme is polynucleotide phosphorylase. This enzyme is part of both the bacterial degradosome and the archaeal exosome,[93] two closely related complexes that recycle RNA into nucleotides. This enzyme degrades RNA by attacking the bond between the 3′-most nucleotides with a phosphate, breaking off a diphosphate nucleotide. This reaction is reversible, and so the enzyme can also extend RNA with more nucleotides. The heteropolymeric tail added by polynucleotide phosphorylase is very rich in adenine. The choice of adenine is most likely the result of higher ADP concentrations than other nucleotides as a result of using ATP as an energy currency, making it more likely to be incorporated in this tail in early lifeforms. It has been suggested that the involvement of adenine-rich tails in RNA degradation prompted the later evolution of polyadenylate polymerases (the enzymes that produce poly(A) tails with no other nucleotides in them).[94]

Polyadenylate polymerases are not as ancient. They have separately evolved in both bacteria and eukaryotes from CCA-adding enzyme, which is the enzyme that completes the 3′ ends of tRNAs. Its catalytic domain is homologous to that of other polymerases.[79] It is presumed that the horizontal transfer of bacterial CCA-adding enzyme to eukaryotes allowed the archaeal-like CCA-adding enzyme to switch function to a poly(A) polymerase.[82] Some lineages, like archaea and cyanobacteria, never evolved a polyadenylate polymerase.[94]

Polyadenylate tails are observed in several

HIV-1 and Poliovirus, inhibit the cell's poly-A binding protein (PABPC1) in order to emphasize their own genes' expression over the host cell's.[99]

History

Poly(A)polymerase was first identified in 1960 as an enzymatic activity in extracts made from cell nuclei that could polymerise ATP, but not ADP, into polyadenine.[100][101] Although identified in many types of cells, this activity had no known function until 1971, when poly(A) sequences were found in mRNAs.[102][103] The only function of these sequences was thought at first to be protection of the 3′ end of the RNA from nucleases, but later the specific roles of polyadenylation in nuclear export and translation were identified. The polymerases responsible for polyadenylation were first purified and characterized in the 1960s and 1970s, but the large number of accessory proteins that control this process were discovered only in the early 1990s.[102]

See also

References

  1. ^
    S2CID 478260
    .
  2. ^ .
  3. ^ .
  4. .
  5. .
  6. .
  7. ^ .
  8. .
  9. ]
  10. .
  11. .
  12. ^ .
  13. ^ .
  14. .
  15. .
  16. .
  17. .
  18. ^ .
  19. ^ .
  20. .
  21. ^ .
  22. .
  23. .
  24. .
  25. .
  26. ^ .
  27. ^ .
  28. .
  29. ^ Molecular Biology of the Cell, Chapter 6, "From DNA to RNA". 4th edition. Alberts B, Johnson A, Lewis J, et al. New York: Garland Science; 2002.
  30. S2CID 34840144
    .
  31. .
  32. .
  33. .
  34. .
  35. .
  36. .
  37. .
  38. .
  39. ^ .
  40. .
  41. .
  42. .
  43. .
  44. ^ .
  45. .
  46. .
  47. .
  48. .
  49. .
  50. .
  51. .
  52. .
  53. .
  54. .
  55. .
  56. .
  57. .
  58. ^ .
  59. .
  60. .
  61. .
  62. .
  63. .
  64. .
  65. .
  66. .
  67. .
  68. ^ .
  69. .
  70. .
  71. .
  72. .
  73. .
  74. .
  75. .
  76. .
  77. .
  78. .
  79. ^ .
  80. .
  81. .
  82. ^ .
  83. ^ .
  84. .
  85. .
  86. .
  87. .
  88. .
  89. .
  90. .
  91. .
  92. .
  93. .
  94. ^ .
  95. .
  96. .
  97. .
  98. .
  99. ^ "Inhibition of host poly(A)-binding protein by virus ~ ViralZone". viralzone.expasy.org.
  100. .
  101. .
  102. ^ .
  103. .

Further reading

External links