DNA and RNA codon tables

Last updated
Aminoacids table.svg
The standard RNA codon table organized in a wheel

A codon table can be used to translate a genetic code into a sequence of amino acids. [1] [2] The standard genetic code is traditionally represented as an RNA codon table, because when proteins are made in a cell by ribosomes, it is messenger RNA (mRNA) that directs protein synthesis. [2] [3] The mRNA sequence is determined by the sequence of genomic DNA. [4] In this context, the standard genetic code is referred to as translation table 1. [3] It can also be represented in a DNA codon table. The DNA codons in such tables occur on the sense DNA strand and are arranged in a 5-to-3 direction. Different tables with alternate codons are used depending on the source of the genetic code, such as from a cell nucleus, mitochondrion, plastid, or hydrogenosome. [5]

Contents

There are 64 different codons in the genetic code and the below tables; most specify an amino acid. [6] Three sequences, UAG, UGA, and UAA, known as stop codons, [note 1] do not code for an amino acid but instead signal the release of the nascent polypeptide from the ribosome. [7] In the standard code, the sequence AUG—read as methionine—can serve as a start codon and, along with sequences such as an initiation factor, initiates translation. [3] [8] [9] In rare instances, start codons in the standard code may also include GUG or UUG; these codons normally represent valine and leucine, respectively, but as start codons they are translated as methionine or formylmethionine. [3] [9]

The first table—the standard table—can be used to translate nucleotide triplets into the corresponding amino acid or appropriate signal if it is a start or stop codon. The second table, appropriately called the inverse, does the opposite: it can be used to deduce a possible triplet code if the amino acid is known. As multiple codons can code for the same amino acid, the International Union of Pure and Applied Chemistry's (IUPAC) nucleic acid notation is given in some instances.

Translation table 1

Standard RNA codon table

Amino-acid biochemical propertiesNonpolar (np)Polar (p)Basic (b)Acidic (a)Termination: stop codon *Initiation: possible start codon ⇒
Standard genetic code [1] [10]
1st
base
2nd base3rd
base
UCAG
U UUU(Phe/F) Phenylalanine (np)UCU(Ser/S) Serine (p)UAU(Tyr/Y) Tyrosine (p)UGU(Cys/C) Cysteine (p)U
UUCUCCUACUGCC
UUA(Leu/L) Leucine (np)UCAUAA Stop (Ochre) * [note 2] UGA Stop (Opal) * [note 2] A
UUG ⇒UCGUAG Stop (Amber) * [note 2] UGG(Trp/W) Tryptophan (np)G
C CUUCCU(Pro/P) Proline (np)CAU(His/H) Histidine (b)CGU(Arg/R) Arginine (b)U
CUCCCCCACCGCC
CUACCACAA(Gln/Q) Glutamine (p)CGAA
CUGCCGCAGCGGG
A AUU(Ile/I) Isoleucine (np)ACU(Thr/T) Threonine (p)AAU(Asn/N) Asparagine (p)AGU(Ser/S) Serine (p)U
AUCACCAACAGCC
AUAACAAAA(Lys/K) Lysine (b)AGA(Arg/R) Arginine (b)A
AUG ⇒(Met/M) Methionine (np)ACGAAGAGGG
G GUU(Val/V) Valine (np)GCU(Ala/A) Alanine (np)GAU(Asp/D) Aspartic acid (a)GGU(Gly/G) Glycine (np)U
GUCGCCGACGGCC
GUAGCAGAA(Glu/E) Glutamic acid (a)GGAA
GUG ⇒GCGGAGGGGG

As shown in the above table, NCBI table 1 includes the less-canonical start codons GUG and UUG. [3]

Inverse RNA codon table

Inverse table for the standard genetic code (compressed using IUPAC notation) [13]
Amino acidRNA codonsCompressedAmino acidRNA codonsCompressed
Ala, AGCU, GCC, GCA, GCGGCNIle, IAUU, AUC, AUAAUH
Arg, RCGU, CGC, CGA, CGG; AGA, AGGCGN, AGR; or
CGY, MGR
Leu, LCUU, CUC, CUA, CUG; UUA, UUGCUN, UUR; or
CUY, YUR
Asn, NAAU, AACAAYLys, KAAA, AAGAAR
Asp, DGAU, GACGAYMet, MAUG
Asn or Asp, BAAU, AAC; GAU, GACRAYPhe, FUUU, UUCUUY
Cys, CUGU, UGCUGYPro, PCCU, CCC, CCA, CCGCCN
Gln, QCAA, CAGCARSer, SUCU, UCC, UCA, UCG; AGU, AGCUCN, AGY
Glu, EGAA, GAGGARThr, TACU, ACC, ACA, ACGACN
Gln or Glu, ZCAA, CAG; GAA, GAGSARTrp, WUGG
Gly, GGGU, GGC, GGA, GGGGGNTyr, YUAU, UACUAY
His, HCAU, CACCAYVal, VGUU, GUC, GUA, GUGGUN
STARTAUG, CUG, UUGHUGSTOPUAA, UGA, UAGURA, UAG; or
UGA, UAR

Standard DNA codon table

Amino-acid biochemical propertiesNonpolar (np)Polar (p)Basic (b)Acidic (a)Termination: stop codon *Initiation: possible start codon ⇒
Standard genetic code [14] [note 3]
1st
base
2nd base3rd
base
TCAG
T TTT(Phe/F) Phenylalanine (np)TCT(Ser/S) Serine (p)TAT(Tyr/Y) Tyrosine (p)TGT(Cys/C) Cysteine (p)T
TTCTCCTACTGCC
TTA(Leu/L) Leucine (np)TCATAA Stop (Ochre) * [note 2] TGA Stop (Opal) * [note 2] A
TTG ⇒TCGTAG Stop (Amber) * [note 2] TGG(Trp/W) Tryptophan (np)G
C CTTCCT(Pro/P) Proline (np)CAT(His/H) Histidine (b)CGT(Arg/R) Arginine (b)T
CTCCCCCACCGCC
CTACCACAA(Gln/Q) Glutamine (p)CGAA
CTGCCGCAGCGGG
A ATT(Ile/I) Isoleucine (np)ACT(Thr/T) Threonine (p)AAT(Asn/N) Asparagine (p)AGT(Ser/S) Serine (p)T
ATCACCAACAGCC
ATAACAAAA(Lys/K) Lysine (b)AGA(Arg/R) Arginine (b)A
ATG ⇒(Met/M) Methionine (np)ACGAAGAGGG
G GTT(Val/V) Valine (np)GCT(Ala/A) Alanine (np)GAT(Asp/D) Aspartic acid (a)GGT(Gly/G) Glycine (np)T
GTCGCCGACGGCC
GTAGCAGAA(Glu/E) Glutamic acid (a)GGAA
GTG ⇒GCGGAGGGGG

Inverse DNA codon table

Inverse table for the standard genetic code (compressed using IUPAC notation) [13]
Amino acidDNA codonsCompressedAmino acidDNA codonsCompressed
Ala, AGCT, GCC, GCA, GCGGCNIle, IATT, ATC, ATAATH
Arg, RCGT, CGC, CGA, CGG; AGA, AGGCGN, AGR; or
CGY, MGR
Leu, LCTT, CTC, CTA, CTG; TTA, TTGCTN, TTR; or
CTY, YTR
Asn, NAAT, AACAAYLys, KAAA, AAGAAR
Asp, DGAT, GACGAYMet, MATG
Asn or Asp, BAAT, AAC; GAT, GACRAYPhe, FTTT, TTCTTY
Cys, CTGT, TGCTGYPro, PCCT, CCC, CCA, CCGCCN
Gln, QCAA, CAGCARSer, STCT, TCC, TCA, TCG; AGT, AGCTCN, AGY
Glu, EGAA, GAGGARThr, TACT, ACC, ACA, ACGACN
Gln or Glu, ZCAA, CAG; GAA, GAGSARTrp, WTGG
Gly, GGGT, GGC, GGA, GGGGGNTyr, YTAT, TACTAY
His, HCAT, CACCAYVal, VGTT, GTC, GTA, GTGGTN
STARTATG, CTG, UTGHTGSTOPTAA, TGA, TAGTRA, TAR

Alternative codons in other translation tables

The genetic code was once believed to be universal: [16] a codon would code for the same amino acid regardless of the organism or source. However, it is now agreed that the genetic code evolves, [17] resulting in discrepancies in how a codon is translated depending on the genetic source. [16] [17] For example, in 1981, it was discovered that the use of codons AUA, UGA, AGA and AGG by the coding system in mammalian mitochondria differed from the universal code. [16] Stop codons can also be affected: in ciliated protozoa, the universal stop codons UAA and UAG code for glutamine. [17] [note 4] The following table displays these alternative codons.

Amino-acid biochemical propertiesNonpolar (np)Polar (p)Basic (b)Acidic (a)Termination: stop codon *
Comparison between codon translations with alternative and standard genetic codes [3]
CodeTranslation
table
DNA codon involvedRNA codon involvedTranslation
with this code
Standard translationNotes
Standard 1Includes translation table 8 (plant chloroplasts).
Vertebrate mitochondrial 2AGAAGA Stop *Arg (R) (b)
AGGAGG Stop *Arg (R) (b)
ATAAUAMet (M) (np)Ile (I) (np)
TGAUGATrp (W) (np) Stop *
Yeast mitochondrial 3ATAAUAMet (M) (np)Ile (I) (np)
CTTCUUThr (T) (p)Leu (L) (np)
CTCCUCThr (T) (p)Leu (L) (np)
CTACUAThr (T) (p)Leu (L) (np)
CTGCUGThr (T) (p)Leu (L) (np)
TGAUGATrp (W) (np) Stop *
CGACGAabsentArg (R) (b)
CGCCGCabsentArg (R) (b)
Mold, protozoan, and coelenterate mitochondrial + Mycoplasma / Spiroplasma 4TGAUGATrp (W) (np) Stop *Includes the translation table 7 (kinetoplasts).
Invertebrate mitochondrial 5AGAAGASer (S) (p)Arg (R) (b)
AGGAGGSer (S) (p)Arg (R) (b)
ATAAUAMet (M) (np)Ile (I) (np)
TGAUGATrp (W) (np) Stop *
Ciliate, dasycladacean and Hexamita nuclear 6TAAUAAGln (Q) (p) Stop *
TAGUAGGln (Q) (p) Stop *
Echinoderm and flatworm mitochondrial 9AAAAAAAsn (N) (p)Lys (K) (b)
AGAAGASer (S) (p)Arg (R) (b)
AGGAGGSer (S) (p)Arg (R) (b)
TGAUGATrp (W) (np) Stop *
Euplotid nuclear 10TGAUGACys (C) (p) Stop *
Bacterial, archaeal and plant plastid 11See translation table 1.
Alternative yeast nuclear 12CTGCUGSer (S) (p)Leu (L) (np)
Ascidian mitochondrial 13AGAAGAGly (G) (np)Arg (R) (b)
AGGAGGGly (G) (np)Arg (R) (b)
ATAAUAMet (M) (np)Ile (I) (np)
TGAUGATrp (W) (np) Stop *
Alternative flatworm mitochondrial 14AAAAAAAsn (N) (p)Lys (K) (b)
AGAAGASer (S) (p)Arg (R) (b)
AGGAGGSer (S) (p)Arg (R) (b)
TAAUAATyr (Y) (p) Stop *
TGAUGATrp (W) (np) Stop *
Blepharisma nuclear 15TAGUAGGln (Q) (p) Stop *As of Nov. 18, 2016: absent from the NCBI update. Similar to translation table 6.
Chlorophycean mitochondrial 16TAGUAGLeu (L) (np) Stop *
Trematode mitochondrial 21TGAUGATrp (W) (np) Stop *
ATAAUAMet (M) (np)Ile (I) (np)
AGAAGASer (S)Arg (R) (b)
AGGAGGSer (S) (p)Arg (R) (b)
AAAAAAAsn (N) (p)Lys (K) (b)
Scenedesmus obliquus mitochondrial 22TCAUCA Stop *Ser (S) (p)
TAGUAGLeu (L) (np) Stop *
Thraustochytrium mitochondrial 23TTAUUA Stop *Leu (L) (np)Similar to translation table 11.
Pterobranchia mitochondrial 24AGAAGASer (S) (p)Arg (R) (b)
AGGAGGLys (K) (b)Arg (R) (b)
TGAUGATrp (W) (np) Stop *
Candidate division SR1 and Gracilibacteria 25TGAUGAGly (G) (np) Stop *
Pachysolen tannophilus nuclear 26CTGCUGAla (A) (np)Leu (L) (np)
Karyorelict nuclear 27TAAUAAGln (Q) (p) Stop *
TAGUAGGln (Q) (p) Stop *
TGUGA Stop *orTrp (W) (np) Stop *
Condylostoma nuclear 28TAAUAA Stop *orGln (Q) (p) Stop *
TAGUAG Stop *orGln (Q) (p) Stop *
TGAUGA Stop *orTrp (W) (np) Stop *
Mesodinium nuclear 29TAAUAATyr (Y) (p) Stop *
TAGUAGTyr (Y) (p) Stop *
Peritrich nuclear 30TAUAAGlu (E) (a) Stop *
TAGUAGGlu (E) (a) Stop *
Blastocrithidia nuclear 31TAAUAA Stop *orGlu (E) (a) Stop *
TAGUAG Stop *orGlu (E) (a) Stop *
TGAUGATrp (W) (np) Stop *
Cephalodiscidae mitochondrial code 33AGAAGASer (S) (p)Arg (R) (b)Similar to translation table 24.
AGGAGGLys (K) (b)Arg (R) (b)
TAAUAATyr (Y) (p) Stop *
TGAUGATrp (W) (np) Stop *

See also

Notes

  1. Each stop codon has a specific name: UAG is amber, UGA is opal or umber, and UAA is ochre. [7] In DNA, these stop codons are TAG, TGA, and TAA, respectively.
  2. 1 2 3 4 5 6 The historical basis for designating the stop codons as amber, ochre and opal is described in the autobiography of Sydney Brenner [11] and in a historical article by Bob Edgar. [12]
  3. The major difference between DNA and RNA is that thymine (T) is only found in the former. In RNA, it is replaced with uracil (U). [15] This is the only difference between the standard RNA codon table and the standard DNA codon table.
  4. Euplotes octacarinatus is an exception. [17]

Related Research Articles

<span class="mw-page-title-main">Genetic code</span> Rules by which information encoded within genetic material is translated into proteins

The genetic code is the set of rules used by living cells to translate information encoded within genetic material into proteins. Translation is accomplished by the ribosome, which links proteinogenic amino acids in an order specified by messenger RNA (mRNA), using transfer RNA (tRNA) molecules to carry amino acids and to read the mRNA three nucleotides at a time. The genetic code is highly similar among all organisms and can be expressed in a simple table with 64 entries.

<span class="mw-page-title-main">Protein biosynthesis</span> Assembly of proteins inside biological cells

Protein biosynthesis is a core biological process, occurring inside cells, balancing the loss of cellular proteins through the production of new proteins. Proteins perform a number of critical functions as enzymes, structural proteins or hormones. Protein synthesis is a very similar process for both prokaryotes and eukaryotes but there are some distinct differences.

<span class="mw-page-title-main">Stop codon</span> Codon that marks the end of a protein-coding sequence

In molecular biology, a stop codon is a codon that signals the termination of the translation process of the current protein. Most codons in messenger RNA correspond to the addition of an amino acid to a growing polypeptide chain, which may ultimately become a protein; stop codons signal the termination of this process by binding release factors, which cause the ribosomal subunits to disassociate, releasing the amino acid chain.

<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.

<span class="mw-page-title-main">Translation (biology)</span> Cellular process of protein synthesis

In biology, translation is the process in living cells in which proteins are produced using RNA molecules as templates. The generated protein is a sequence of amino acids. This sequence is determined by the sequence of nucleotides in the RNA. The nucleotides are considered three at a time. Each such triple results in addition of one specific amino acid to the protein being generated. The matching from nucleotide triple to amino acid is called the genetic code. The translation is performed by a large complex of functional RNA and proteins called ribosomes. The entire process is called gene expression.

<span class="mw-page-title-main">Proteinogenic amino acid</span> Amino acid that is incorporated biosynthetically into proteins during translation

Proteinogenic amino acids are amino acids that are incorporated biosynthetically into proteins during translation. The word "proteinogenic" means "protein creating". Throughout known life, there are 22 genetically encoded (proteinogenic) amino acids, 20 in the standard genetic code and an additional 2 that can be incorporated by special translation mechanisms.

<span class="mw-page-title-main">Reading frame</span> Division of RNA/DNA sequences into sets of triplets which correspond to amino acids

In molecular biology, a reading frame is a way of dividing the sequence of nucleotides in a nucleic acid molecule into a set of consecutive, non-overlapping triplets. Where these triplets equate to amino acids or stop signals during translation, they are called codons.

<span class="mw-page-title-main">Wobble base pair</span> RNA base pair that does not follow Watson-Crick base pair rules

A wobble base pair is a pairing between two nucleotides in RNA molecules that does not follow Watson-Crick base pair rules. The four main wobble base pairs are guanine-uracil (G-U), hypoxanthine-uracil (I-U), hypoxanthine-adenine (I-A), and hypoxanthine-cytosine (I-C). In order to maintain consistency of nucleic acid nomenclature, "I" is used for hypoxanthine because hypoxanthine is the nucleobase of inosine; nomenclature otherwise follows the names of nucleobases and their corresponding nucleosides. The thermodynamic stability of a wobble base pair is comparable to that of a Watson-Crick base pair. Wobble base pairs are fundamental in RNA secondary structure and are critical for the proper translation of the genetic code.

<span class="mw-page-title-main">Frameshift mutation</span> Mutation that shifts codon alignment

A frameshift mutation is a genetic mutation caused by indels of a number of nucleotides in a DNA sequence that is not divisible by three. Due to the triplet nature of gene expression by codons, the insertion or deletion can change the reading frame, resulting in a completely different translation from the original. The earlier in the sequence the deletion or insertion occurs, the more altered the protein. A frameshift mutation is not the same as a single-nucleotide polymorphism in which a nucleotide is replaced, rather than inserted or deleted. A frameshift mutation will in general cause the reading of the codons after the mutation to code for different amino acids. The frameshift mutation will also alter the first stop codon encountered in the sequence. The polypeptide being created could be abnormally short or abnormally long, and will most likely not be functional.

<span class="mw-page-title-main">Nirenberg and Matthaei experiment</span>

The Nirenberg and Matthaei experiment was a scientific experiment performed in May 1961 by Marshall W. Nirenberg and his post-doctoral fellow, J. Heinrich Matthaei, at the National Institutes of Health (NIH). The experiment deciphered the first of the 64 triplet codons in the genetic code by using nucleic acid homopolymers to translate specific amino acids.

<span class="mw-page-title-main">Nirenberg and Leder experiment</span>

The Nirenberg and Leder experiment was a scientific experiment performed in 1964 by Marshall W. Nirenberg and Philip Leder. The experiment elucidated the triplet nature of the genetic code and allowed the remaining ambiguous codons in the genetic code to be deciphered.

In genetics, a nonsense mutation is a point mutation in a sequence of DNA that results in a nonsense codon, or a premature stop codon in the transcribed mRNA, and leads to a truncated, incomplete, and possibly nonfunctional protein product. Nonsense mutations are not always harmful; the functional effect of a nonsense mutation depends on many aspects, such as the location of the stop codon within the coding DNA. For example, the effect of a nonsense mutation depends on the proximity of the nonsense mutation to the original stop codon, and the degree to which functional subdomains of the protein are affected. As nonsense mutations leads to premature termination of polypeptide chains; they are also called chain termination mutations.

In molecular biology, reading frames are defined as spans of DNA sequence between the start and stop codons. Usually, this is considered within a studied region of a prokaryotic DNA sequence, where only one of the six possible reading frames will be "open". Such an ORF may contain a start codon and by definition cannot extend beyond a stop codon. That start codon indicates where translation may start. The transcription termination site is located after the ORF, beyond the translation stop codon. If transcription were to cease before the stop codon, an incomplete protein would be made during translation.

<span class="mw-page-title-main">Silent mutation</span> DNA mutation with no observable effect on an organisms phenotype

Silent mutations are mutations in DNA that do not have an observable effect on the organism's phenotype. They are a specific type of neutral mutation. The phrase silent mutation is often used interchangeably with the phrase synonymous mutation; however, synonymous mutations are not always silent, nor vice versa. Synonymous mutations can affect transcription, splicing, mRNA transport, and translation, any of which could alter phenotype, rendering the synonymous mutation non-silent. The substrate specificity of the tRNA to the rare codon can affect the timing of translation, and in turn the co-translational folding of the protein. This is reflected in the codon usage bias that is observed in many species. Mutations that cause the altered codon to produce an amino acid with similar functionality are often classified as silent; if the properties of the amino acid are conserved, this mutation does not usually significantly affect protein function.

Xenobiology (XB) is a subfield of synthetic biology, the study of synthesizing and manipulating biological devices and systems. The name "xenobiology" derives from the Greek word xenos, which means "stranger, alien". Xenobiology is a form of biology that is not (yet) familiar to science and is not found in nature. In practice, it describes novel biological systems and biochemistries that differ from the canonical DNA–RNA-20 amino acid system. For example, instead of DNA or RNA, XB explores nucleic acid analogues, termed xeno nucleic acid (XNA) as information carriers. It also focuses on an expanded genetic code and the incorporation of non-proteinogenic amino acids, or “xeno amino acids” into proteins.

The Shine–Dalgarno (SD) sequence is a ribosomal binding site in bacterial and archaeal messenger RNA, generally located around 8 bases upstream of the start codon AUG. The RNA sequence helps recruit the ribosome to the messenger RNA (mRNA) to initiate protein synthesis by aligning the ribosome with the start codon. Once recruited, tRNA may add amino acids in sequence as dictated by the codons, moving downstream from the translational start site.

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

A synonymous substitution is the evolutionary substitution of one base for another in an exon of a gene coding for a protein, such that the produced amino acid sequence is not modified. This is possible because the genetic code is "degenerate", meaning that some amino acids are coded for by more than one three-base-pair codon; since some of the codons for a given amino acid differ by just one base pair from others coding for the same amino acid, a mutation that replaces the "normal" base by one of the alternatives will result in incorporation of the same amino acid into the growing polypeptide chain when the gene is translated. Synonymous substitutions and mutations affecting noncoding DNA are often considered silent mutations; however, it is not always the case that the mutation is silent.

<span class="mw-page-title-main">Start codon</span> First codon of a messenger RNA translated by a ribosome

The start codon is the first codon of a messenger RNA (mRNA) transcript translated by a ribosome. The start codon always codes for methionine in eukaryotes and archaea and a N-formylmethionine (fMet) in bacteria, mitochondria and plastids.

A nonsense suppressor is a factor which can inhibit the effect of the nonsense mutation. Nonsense suppressors can be generally divided into two classes: a) a mutated tRNA which can bind with a termination codon on mRNA; b) a mutation on ribosomes decreasing the effect of a termination codon. It is believed that nonsense suppressors keep a low concentration in the cell and do not disrupt normal translation most of the time. In addition, many genes do not have only one termination codon, and cells commonly use ochre codons as the termination signal, whose nonsense suppressors are usually inefficient.

<span class="mw-page-title-main">Expanded genetic code</span> Modified genetic code

An expanded genetic code is an artificially modified genetic code in which one or more specific codons have been re-allocated to encode an amino acid that is not among the 22 common naturally-encoded proteinogenic amino acids.

References

  1. 1 2 "Amino Acid Translation Table". Oregon State University. Archived from the original on 29 May 2020. Retrieved 2 December 2020.
  2. 1 2 Bartee, Lisa; Brook, Jack. MHCC Biology 112: Biology for Health Professions. Open Oregon. p. 42. Archived from the original on 6 December 2020. Retrieved 6 December 2020.
  3. 1 2 3 4 5 6 Elzanowski A, Ostell J (7 January 2019). "The Genetic Codes". National Center for Biotechnology Information. Archived from the original on 5 October 2020. Retrieved 21 February 2019.
  4. "RNA Functions". Scitable. Nature Education. Archived from the original on 18 October 2008. Retrieved 5 January 2021.
  5. "The Genetic Codes". National Center for Biotechnology Information. Archived from the original on 13 May 2011. Retrieved 2 December 2020.
  6. "Codon". National Human Genome Research Institute. Archived from the original on 22 October 2020. Retrieved 10 October 2020.
  7. 1 2 Maloy S. (29 November 2003). "How nonsense mutations got their names". Microbial Genetics Course. San Diego State University. Archived from the original on 23 September 2020. Retrieved 10 October 2020.
  8. Hinnebusch AG (2011). "Molecular Mechanism of Scanning and Start Codon Selection in Eukaryotes". Microbiology and Molecular Biology Reviews. 75 (3): 434–467. doi: 10.1128/MMBR.00008-11 . PMC   3165540 . PMID   21885680.
  9. 1 2 Touriol C, Bornes S, Bonnal S, Audigier S, Prats H, Prats AC, Vagner S (2003). "Generation of protein isoform diversity by alternative initiation of translation at non-AUG codons". Biology of the Cell. 95 (3–4): 169–78. doi: 10.1016/S0248-4900(03)00033-9 . PMID   12867081.
  10. "The Information in DNA Determines Cellular Function via Translation". Scitable. Nature Education. Archived from the original on 23 September 2017. Retrieved 5 December 2020.
  11. Brenner, Sydney; Wolpert, Lewis (2001). A Life in Science. Biomed Central Limited. pp. 101–104. ISBN   9780954027803.
  12. Edgar B (2004). "The genome of bacteriophage T4: an archeological dig". Genetics. 168 (2): 575–82. doi:10.1093/genetics/168.2.575. PMC   1448817 . PMID   15514035. see pages 580–581
  13. 1 2 IUPAC—IUB Commission on Biochemical Nomenclature. "Abbreviations and Symbols for Nucleic Acids, Polynucleotides and Their Constituents" (PDF). International Union of Pure and Applied Chemistry. Retrieved 5 December 2020.
  14. "What does DNA do?". Your Genome. Welcome Genome Campus. Archived from the original on 29 November 2020. Retrieved 12 January 2021.
  15. "Genes". DNA, Genetics, and Evolution. Boston University. Archived from the original on 28 April 2020. Retrieved 10 December 2020.
  16. 1 2 3 Osawa, A (November 1993). "Evolutionary changes in the genetic code". Comparative Biochemistry and Physiology. 106 (2): 489–94. doi:10.1016/0305-0491(93)90122-l. PMID   8281749.
  17. 1 2 3 4 Osawa S, Jukes TH, Watanabe K, Muto A (March 1992). "Recent evidence for evolution of the genetic code". Microbiological Reviews. 56 (1): 229–64. doi:10.1128/MR.56.1.229-264.1992. PMC   372862 . PMID   1579111.

Further reading