Jump to content

Genetic code

This is a good article. Click here for more information.
From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 155.69.125.199 (talk) at 13:56, 13 October 2017. The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

A series of codons in part of a messenger RNA (mRNA) molecule. Each codon consists of three nucleotides, usually corresponding to a single amino acid. The nucleotides are abbreviated with the letters A, U, G and C. This is mRNA, which uses U (uracil). DNA uses T (thymine) instead. This mRNA molecule will instruct a ribosome to synthesize a protein according to this code.

The genetic code is the set of rules by which information encoded within genetic material (DNA or mRNA sequences) is translated into proteins by living cells. Translation is accomplished by the ribosome, which links amino acids in an order specified by 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.[1]

The code defines how sequences of nucleotide triplets, called codons, specify which amino acid will be added next during protein synthesis. With some exceptions,[2] a three-nucleotide codon in a nucleic acid sequence specifies a single amino acid. The vast majority of genes are encoded with a single scheme (see the RNA codon table). That scheme is often referred to as the canonical or standard genetic code, or simply the genetic code, though variant codes (such as in human mitochondria) exist.

While the "genetic code" determines a protein's amino acid sequence, other genomic regions determine when and where these proteins are produced according to various "gene regulatory codes".

History

The genetic code

Efforts to understand how proteins are encoded began after DNA's structure was discovered in 1953. George Gamow postulated that sets of three bases must be employed to encode the 20 standard amino acids used by living cells to build proteins, which would allow a maximum of 43 = 64 amino acids.[3]

Codons

The Crick, Brenner et al. experiment first demonstrated that codons consist of three DNA bases. Marshall Nirenberg and Heinrich J. Matthaei were the first to reveal the nature of a codon in 1961.

They used a cell-free system to translate a poly-uracil RNA sequence (i.e., UUUUU...) and discovered that the polypeptide that they had synthesized consisted of only the amino acid phenylalanine.[4] They thereby deduced that the codon UUU specified the amino acid phenylalanine.

This was followed by experiments in Severo Ochoa's laboratory that demonstrated that the poly-adenine RNA sequence (AAAAA...) coded for the polypeptide poly-lysine[5] and that the poly-cytosine RNA sequence (CCCCC...) coded for the polypeptide poly-proline.[6] Therefore, the codon AAA specified the amino acid lysine, and the codon CCC specified the amino acid proline. Using various copolymers most of the remaining codons were then determined.

Subsequent work by Har Gobind Khorana identified the rest of the genetic code. Shortly thereafter, Robert W. Holley determined the structure of transfer RNA (tRNA), the adapter molecule that facilitates the process of translating RNA into protein. This work was based upon Ochoa's earlier studies, yielding the latter the Nobel Prize in Physiology or Medicine in 1959 for work on the enzymology of RNA synthesis.[7]

Extending this work, Nirenberg and Philip Leder revealed the code's triplet nature and deciphered its codons. In these experiments, various combinations of mRNA were passed through a filter that contained ribosomes, the components of cells that translate RNA into protein. Unique triplets promoted the binding of specific tRNAs to the ribosome. Leder and Nirenberg were able to determine the sequences of 54 out of 64 codons in their experiments.[8] Khorana, Holley and Nirenberg received the 1968 Nobel for their work.[9]

The three stop codons were named by discoverers Richard Epstein and Charles Steinberg. "Amber" was named after their friend Harris Bernstein, whose last name means "amber" in German.[10] The other two stop codons were named "ochre" and "opal" in order to keep the "color names" theme.

Expanded genetic codes (synthetic biology)

Since 2001, 40 non-natural amino acids have been added into protein by creating a unique codon (recoding) and a corresponding transfer-RNA:aminoacyl – tRNA-synthetase pair to encode it with diverse physicochemical and biological properties in order to be used as a tool to exploring protein structure and function or to create novel or enhanced proteins.[11] [12]

H. Murakami and M. Sisido extended some codons to have four and five bases. Steven A. Benner constructed a functional 65th (in vivo) codon.[13]

In 2015 N. Budisa, D. Söll and co-workers reported the full substitution of all 20,899 tryptophan residues (UGG codons) with unnatural thienopyrrole-alanine in the genetic code of the bacterium Escherichia coli.[14]

In 2016 the first stable semisynthetic organism was created. It was a (single cell) bacterium with two synthetic bases (called X and Y). The bases survived cell division.[15][16]

In 2017 a mouse engineered with an extended genetic code that can produce proteins with unnatural amino acids was reported.[17]

Features

Reading frame

A codon is defined by the initial nucleotide from which translation starts and sets the frame for a run of successive triplets, which is known as an "open reading frame" (ORF). For example, the string GGGAAACCC, if read from the first position, contains the codons GGG, AAA, and CCC; if read from the second position, it contains the codons GGA and AAC; and if read from the third position, GAA and ACC. Every sequence can, thus, be read in its 5' → 3' direction in three reading frames, each producing a possibly distinct amino acid sequence (in the given example, Gly-Lys-Pro, Gly-Asn, or Glu-Thr, respectively). DNA is double-stranded defining six possible reading frames, three in the forward orientation on one strand and three reverse on the opposite strand.[18]: 330  Protein-coding frames are defined by a start codon, usually the first AUG codon in the sequence.

In eukaryotes, ORFs in exons are often interrupted by introns.

Start/stop codons

Translation starts with a chain-initiation codon or start codon. The start codon alone is not sufficient to begin the process. Nearby sequences such as the Shine-Dalgarno sequence in E. coli and initiation factors are also required to start translation. The most common start codon is AUG, which is read as methionine or, in bacteria, as formylmethionine. Alternative start codons depending on the organism include "GUG" or "UUG"; these codons normally represent valine and leucine, respectively, but as start codons they are translated as methionine or formylmethionine.[19]

The three stop codons have names: UAG is amber, UGA is opal (sometimes also called umber), and UAA is ochre. Stop codons are also called "termination" or "nonsense" codons. They signal release of the nascent polypeptide from the ribosome because no cognate tRNA has anticodons complementary to these stop signals, allowing a release factor to bind to the ribosome instead.[20]

Effect of mutations

Examples of notable mutations that can occur in humans.[21]

During the process of DNA replication, errors occasionally occur in the polymerization of the second strand. These errors, mutations, can affect an organism's phenotype, especially if they occur within the protein coding sequence of a gene. Error rates are typically 1 error in every 10–100 million bases—due to the "proofreading" ability of DNA polymerases.[22][23]

Missense mutations and nonsense mutations are examples of point mutations that can cause genetic diseases such as sickle-cell disease and thalassemia respectively.[24][25][26] Clinically important missense mutations generally change the properties of the coded amino acid residue among basic, acidic, polar or non-polar states, whereas nonsense mutations result in a stop codon.[18]: 266

Mutations that disrupt the reading frame sequence by indels (insertions or deletions) of a non-multiple of 3 nucleotide bases are known as frameshift mutations. These mutations usually result in a completely different translation from the original, and likely cause a stop codon to be read, which truncates the protein.[27] These mutations may impair the protein's function and are thus rare in in vivo protein-coding sequences. One reason inheritance of frameshift mutations is rare is that, if the protein being translated is essential for growth under the selective pressures the organism faces, absence of a functional protein may cause death before the organism becomes viable.[28] Frameshift mutations may result in severe genetic diseases such as Tay-Sachs disease.[29]

Although most mutations that change protein sequences are harmful or neutral, some mutations have benefits.[30] These mutations may enable the mutant organism to withstand particular environmental stresses better than wild type organisms, or reproduce more quickly. In these cases a mutation will tend to become more common in a population through natural selection.[31] Viruses that use RNA as their genetic material have rapid mutation rates,[32] which can be an advantage, since these viruses thereby evolve rapidly, and thus evade the immune system defensive responses.[33] In large populations of asexually reproducing organisms, for example, E. coli, multiple beneficial mutations may co-occur. This phenomenon is called clonal interference and causes competition among the mutations.[34]

Degeneracy

Degeneracy is the redundancy of the genetic code. This term was given by Bernfield and Nirenberg. The genetic code has redundancy but no ambiguity (see the codon tables below for the full correlation). For example, although codons GAA and GAG both specify glutamic acid (redundancy), neither specifies another amino acid (no ambiguity). The codons encoding one amino acid may differ in any of their three positions. For example, the amino acid leucine is specified by YUR or CUN (UUA, UUG, CUU, CUC, CUA, or CUG) codons (difference in the first or third position indicated using IUPAC notation), while the amino acid serine is specified by UCN or AGY (UCA, UCG, UCC, UCU, AGU, or AGC) codons (difference in the first, second, or third position).[35]: 102–117 : 521–522 A practical consequence of redundancy is that errors in the third position of the triplet codon cause only a silent mutation or an error that would not affect the protein because the hydrophilicity or hydrophobicity is maintained by equivalent substitution of amino acids; for example, a codon of NUN (where N = any nucleotide) tends to code for hydrophobic amino acids. NCN yields amino acid residues that are small in size and moderate in hydropathy; NAN encodes average size hydrophilic residues. The genetic code is so well-structured for hydropathy that a mathematical analysis (Singular Value Decomposition) of 12 variables (4 nucleotides x 3 positions) yields a remarkable correlation (C = 0.95) for predicting the hydropathy of the encoded amino acid directly from the triplet nucleotide sequence, without translation.[36][37] Note in the table, below, eight amino acids are not affected at all by mutations at the third position of the codon, whereas in the figure above, a mutation at the second position is likely to cause a radical change in the physicochemical properties of the encoded amino acid.

Grouping of codons by amino acid residue molar volume and hydropathy. A more detailed version is available.

Codon usage bias

The frequency of codons, also known as codon usage bias, can vary from species to species with functional implications for the control of translation. The following codon usage table is for the human genome.[38]

Human genome codon frequency table
Human genome codon frequency
Codon AA[A] Fraction[B] Freq [C] Number[D] Codon AA Fraction Freq Number Codon AA Fraction Freq Number Codon AA Fraction Freq Number
UUU F 0.46 17.6 714,298 UCU S 0.19 15.2 618,711 UAU Y 0.44 12.2 495,699 UGU C 0.46 10.6 430,311
UUC F 0.54 20.3 824,692 UCC S 0.22 17.7 718,892 UAC Y 0.56 15.3 622,407 UGC C 0.54 12.6 513,028
UUA L 0.08 7.7 311,881 UCA S 0.15 12.2 496,448 UAA * 0.30 1.0 40,285 UGA * 0.47 1.6 63,237
UUG L 0.13 12.9 525,688 UCG S 0.05 4.4 179,419 UAG * 0.24 0.8 32,109 UGG W 1.00 13.2 535,595
CUU L 0.13 13.2 536,515 CCU P 0.29 17.5 713,233 CAU H 0.42 10.9 441,711 CGU R 0.08 4.5 184,609
CUC L 0.20 19.6 796,638 CCC P 0.32 19.8 804,620 CAC H 0.58 15.1 613,713 CGC R 0.18 10.4 423,516
CUA L 0.07 7.2 290,751 CCA P 0.28 16.9 688,038 CAA Q 0.27 12.3 501,911 CGA R 0.11 6.2 250,760
CUG L 0.40 39.6 1,611,801 CCG P 0.11 6.9 281,570 CAG Q 0.73 34.2 1,391,973 CGG R 0.20 11.4 464,485
AUU I 0.36 16.0 650,473 ACU T 0.25 13.1 533,609 AAU N 0.47 17.0 689,701 AGU S 0.15 12.1 493,429
AUC I 0.47 20.8 846,466 ACC T 0.36 18.9 768,147 AAC N 0.53 19.1 776,603 AGC S 0.24 19.5 791,383
AUA I 0.17 7.5 304,565 ACA T 0.28 15.1 614,523 AAA K 0.43 24.4 993,621 AGA R 0.21 12.2 494,682
AUG M 1.00 22.0 896,005 ACG T 0.11 6.1 246,105 AAG K 0.57 31.9 1,295,568 AGG R 0.21 12.0 486,463
GUU V 0.18 11.0 448,607 GCU A 0.27 18.4 750,096 GAU D 0.46 21.8 885,429 GGU G 0.16 10.8 437,126
GUC V 0.24 14.5 588,138 GCC A 0.40 27.7 1,127,679 GAC D 0.54 25.1 1,020,595 GGC G 0.34 22.2 903,565
GUA V 0.12 7.1 287,712 GCA A 0.23 15.8 643,471 GAA E 0.42 29.0 1,177,632 GGA G 0.25 16.5 669,873
GUG V 0.46 28.1 1,143,534 GCG A 0.11 7.4 299,495 GAG E 0.58 39.6 1,609,975 GGG G 0.25 16.5 669,768

A Amino acid. B Fraction of each codon among all those specifying a given amino acid. C Frequency among 40,662,582 codons of 93,487 coding sequences. D Number.

Standard codon tables

RNA codon table

Amino-acid biochemical properties Nonpolar Polar Basic Acidic Termination: stop codon
Standard genetic code (NCBI table 1)[39]
1st
base
2nd base 3rd
base
U C A G
U UUU (Phe/F) Phenylalanine UCU (Ser/S) Serine UAU (Tyr/Y) Tyrosine UGU (Cys/C) Cysteine U
UUC UCC UAC UGC C
UUA (Leu/L) Leucine UCA UAA Stop (Ochre)[B] UGA Stop (Opal)[B] A
UUG[A] UCG UAG Stop (Amber)[B] UGG (Trp/W) Tryptophan G
C CUU CCU (Pro/P) Proline CAU (His/H) Histidine CGU (Arg/R) Arginine U
CUC CCC CAC CGC C
CUA CCA CAA (Gln/Q) Glutamine CGA A
CUG CCG CAG CGG G
A AUU (Ile/I) Isoleucine ACU (Thr/T) Threonine AAU (Asn/N) Asparagine AGU (Ser/S) Serine U
AUC ACC AAC AGC C
AUA ACA AAA (Lys/K) Lysine AGA (Arg/R) Arginine A
AUG[A] (Met/M) Methionine ACG AAG AGG G
G GUU (Val/V) Valine GCU (Ala/A) Alanine GAU (Asp/D) Aspartic acid GGU (Gly/G) Glycine U
GUC GCC GAC GGC C
GUA GCA GAA (Glu/E) Glutamic acid GGA A
GUG[A] GCG GAG GGG G
A Possible start codons in NCBI table 1. AUG is most common.[40] The two other start codons listed by table 1 (GUG and UUG) are rare in eukaryotes.[41] Prokaryotes have less strigent start codon requirements; they are described by NCBI table 11.
B ^ ^ ^ The historical basis for designating the stop codons as amber, ochre and opal is described in an autobiography by Sydney Brenner[42] and in a historical article by Bob Edgar.[43]
Inverse table for the standard genetic code (compressed using IUPAC notation)
Amino acid DNA codons Compressed Amino acid DNA codons Compressed
Ala, A GCU, GCC, GCA, GCG GCN Ile, I AUU, AUC, AUA AUH
Arg, R CGU, CGC, CGA, CGG; AGA, AGG CGN, AGR; or
CGY, MGR
Leu, L CUU, CUC, CUA, CUG; UUA, UUG CUN, UUR; or
CUY, YUR
Asn, N AAU, AAC AAY Lys, K AAA, AAG AAR
Asp, D GAU, GAC GAY Met, M AUG
Asn or Asp, B AAU, AAC; GAU, GAC RAY Phe, F UUU, UUC UUY
Cys, C UGU, UGC UGY Pro, P CCU, CCC, CCA, CCG CCN
Gln, Q CAA, CAG CAR Ser, S UCU, UCC, UCA, UCG; AGU, AGC UCN, AGY
Glu, E GAA, GAG GAR Thr, T ACU, ACC, ACA, ACG ACN
Gln or Glu, Z CAA, CAG; GAA, GAG SAR Trp, W UGG
Gly, G GGU, GGC, GGA, GGG GGN Tyr, Y UAU, UAC UAY
His, H CAU, CAC CAY Val, V GUU, GUC, GUA, GUG GUN
START AUG, CUG, UUG HUG STOP UAA, UGA, UAG URA, UAR

DNA codon table

The DNA codon table is essentially identical to that for RNA, but with U replaced by T.

Alternative genetic codes

Non-standard amino acids

In some proteins, non-standard amino acids are substituted for standard stop codons, depending on associated signal sequences in the messenger RNA. For example, UGA can code for selenocysteine and UAG can code for pyrrolysine. Selenocysteine became to be seen as the 21st amino acid, and pyrrolysine as the 22nd.[44] Unlike selenocysteine, pyrrolysine-encoded UAG is translated with the participation of a dedicated aminoacyl-tRNA synthetase.[45] Both selenocysteine and pyrrolysine may be present in the same organism.[46] Although the genetic code is normally fixed in an organism, the achaeal prokaryote Acetohalobium arabaticum can expand its genetic code from 20 to 21 amino acids (by including pyrrolysine) under different conditions of growth.[47]

Variations

Genetic code logo of the Globobulimina pseudospinescens mitochondrial genome. The logo shows the 64 codons from left to right, predicted alternatives in red (relative to the standard genetic code). Red line: stop codons. The height of each amino acid in the stack shows how often it is aligned to the codon in homologous protein domains. The stack height indicates the support for the prediction.

Variations on the standard code were predicted in the 1970s.[48] The first was discovered in 1979, by researchers studying human mitochondrial genes.[49] Many slight variants were discovered thereafter,[44] including various alternative mitochondrial codes.[50] Small variants such as translation of the codon UGA as tryptophan in Mycoplasma species, and translation of CUG as a serine rather than leucine in yeasts of the "CTG clade" (such as Candida albicans).[51][52][53] Because viruses must use the same genetic code as their hosts, modifications to the standard genetic code could interfere with viral protein synthesis or functioning.[54] However, viruses such as totiviruses adapted to the host's genetic code modification.[55] In bacteria and archaea, GUG and UUG are common start codons. In rare cases, certain proteins may use alternative start codons.[44] Surprisingly, variations in the interpretation of the genetic code exist also in human nuclear-encoded genes: In 2016, researchers studying the translation of malate dehydrogenase found that in about 4% of the mRNAs encoding this enzyme the stop codon is naturally used to encode the amino acids tryptophan and arginine.[56] This type of recoding is induced by a high-readthrough stop codon context[57] and it is referred to as functional translational readthrough.[58]

Variant genetic codes used by an organism can be inferred by identifying highly conserved genes encoded in that genome, and comparing its codon usage to the amino acids in homologous proteins of other organisms. For example, the program FACIL[59] infers a genetic code by searching which amino acids in homologous protein domains are most often aligned to every codon. The resulting amino acid probabilities for each codon are displayed in a genetic code logo, that also shows the support for a stop codon.

Despite these differences, all known naturally occurring codes are very similar. The coding mechanism is the same for all organisms: three-base codons, tRNA, ribosomes, single direction reading and translating single codons into single amino acids.

List of alternative codons

List of alternative codons
Amino acids biochemical properties nonpolar polar basic acidic Termination: stop codon
Comparison between codon translations with alternative and standard genetic codes
Code Translation
table
DNA codon involved RNA codon involved Translation
with this code
Standard translation Notes
Standard 1 Includes translation table 8 (plant chloroplasts).
Vertebrate mitochondrial 2 AGA AGA Ter (*) Arg (R)
AGG AGG Ter (*) Arg (R)
ATA AUA Met (M) Ile (I)
TGA UGA Trp (W) Ter (*)
Yeast mitochondrial 3 ATA AUA Met (M) Ile (I)
CTT CUU Thr (T) Leu (L)
CTC CUC Thr (T) Leu (L)
CTA CUA Thr (T) Leu (L)
CTG CUG Thr (T) Leu (L)
TGA UGA Trp (W) Ter (*)
CGA CGA absent Arg (R)
CGC CGC absent Arg (R)
Mold, protozoan, and coelenterate mitochondrial + Mycoplasma / Spiroplasma 4 TGA UGA Trp (W) Ter (*) Includes the translation table 7 (kinetoplasts).
Invertebrate mitochondrial 5 AGA AGA Ser (S) Arg (R)
AGG AGG Ser (S) Arg (R)
ATA AUA Met (M) Ile (I)
TGA UGA Trp (W) Ter (*)
Ciliate, dasycladacean and Hexamita nuclear 6 TAA UAA Gln (Q) Ter (*)
TAG UAG Gln (Q) Ter (*)
Echinoderm and flatworm mitochondrial 9 AAA AAA Asn (N) Lys (K)
AGA AGA Ser (S) Arg (R)
AGG AGG Ser (S) Arg (R)
TGA UGA Trp (W) Ter (*)
Euplotid nuclear 10 TGA UGA Cys (C) Ter (*)
Bacterial, archaeal and plant plastid 11 See translation table 1.
Alternative yeast nuclear 12 CTG CUG Ser (S) Leu (L)
Ascidian mitochondrial 13 AGA AGA Gly (G) Arg (R)
AGG AGG Gly (G) Arg (R)
ATA AUA Met (M) Ile (I)
TGA UGA Trp (W) Ter (*)
Alternative flatworm mitochondrial 14 AAA AAA Asn (N) Lys (K)
AGA AGA Ser (S) Arg (R)
AGG AGG Ser (S) Arg (R)
TAA UAA Tyr (Y) Ter (*)
TGA UGA Trp (W) Ter (*)
Blepharisma nuclear 15 TAG UAG Gln (Q) Ter (*) As of Nov. 18, 2016: absent from the NCBI update.
Chlorophycean mitochondrial 16 TAG UAG Leu (L) Ter (*)
Trematode mitochondrial 21 TGA UGA Trp (W) Ter (*)
ATA AUA Met (M) Ile (I)
AGA AGA Ser (S) Arg (R)
AGG AGG Ser (S) Arg (R)
AAA AAA Asn (N) Lys (K)
Scenedesmus obliquus mitochondrial 22 TCA UCA Ter (*) Ser (S)
TAG UAG Leu (L) Ter (*)
Thraustochytrium mitochondrial 23 TTA UUA Ter (*) Leu (L) Similar to translation table 11.
Pterobranchia mitochondrial 24 AGA AGA Ser (S) Arg (R)
AGG AGG Lys (K) Arg (R)
TGA UGA Trp (W) Ter (*)
Candidate division SR1 and Gracilibacteria 25 TGA UGA Gly (G) Ter (*)
Pachysolen tannophilus nuclear 26 CTG CUG Ala (A) Leu (L)
Karyorelict nuclear 27 TAA UAA Gln (Q) Ter (*)
TAG UAG Gln (Q) Ter (*)
TGA UGA Ter (*) or Trp (W) Ter (*)
Condylostoma nuclear 28 TAA UAA Ter (*) or Gln (Q) Ter (*)
TAG UAG Ter (*) or Gln (Q) Ter (*)
TGA UGA Ter (*) or Trp (W) Ter (*)
Mesodinium nuclear 29 TAA UAA Tyr (Y) Ter (*)
TAG UAG Tyr (Y) Ter (*)
Peritrich nuclear 30 TAA UAA Glu (E) Ter (*)
TAG UAG Glu (E) Ter (*)
Blastocrithidia nuclear 31 TAA UAA Ter (*) or Gln (Q) Ter (*)
TAG UAG Ter (*) or Gln (Q) Ter (*)
TGA UGA Trp (W) Ter (*)

Origin

The genetic code is a key part of the story of life. The main hypothesis for life's origin is the RNA world hypothesis. Any model for the emergence of genetic code is intimately related to a model of the transfer from ribozymes (RNA enzymes) to proteins as the principal enzymes in cells. In line with the RNA world hypothesis, transfer RNA molecules appear to have evolved before modern aminoacyl-tRNA synthetases, so the latter cannot be part of the explanation of its patterns.[60]

A hypothetical randomly evolved genetic code further motivates a biochemical or evolutionary model for its origin. If amino acids were randomly assigned to triplet codons, there would be 1.5 × 1084 possible genetic codes.[61]: 163 This number is found by calculating the number of ways that 21 items (20 amino acids plus one stop) can be placed in 64 bins, wherein each item is used at least once.[62] However, the distribution of codon assignments in the genetic code is nonrandom.[63] In particular, the genetic code clusters certain amino acid assignments.

Amino acids that share the same biosynthetic pathway tend to have the same first base in their codons. This could be an evolutionary relic of an early, simpler genetic code with fewer amino acids that later evolved to code a larger set of amino acids.[64] It could also reflect steric and chemical properties that had another effect on the codon during its evolution. Amino acids with similar physical properties also tend to have similar codons,[65][66] reducing the problems caused by point mutations and mistranslations.[63]

Given the non-random genetic triplet coding scheme, a tenable hypothesis for the origin of genetic code could address multiple aspects of the codon table, such as absence of codons for D-amino acids, secondary codon patterns for some amino acids, confinement of synonymous positions to third position, the small set of only 20 amino acids (instead of a number approaching 64), and the relation of stop codon patterns to amino acid coding patterns.[67]

Three main hypotheses address the origin of the genetic code. Many models belong to one of them or to a hybrid:[68]

  • Random freeze: the genetic code was randomly created. For example, early tRNA-like ribozymes may have had different affinities for amino acids, with codons emerging from another part of the ribozyme that exhibited random variability. Once enough peptides were coded for, any major random change in the genetic code would have been lethal; hence it became "frozen".[69]
  • Stereochemical affinity: the genetic code is a result of a high affinity between each amino acid and its codon or anti-codon; the latter option implies that pre-tRNA molecules matched their corresponding amino acids by this affinity. Later during evolution, this matching was gradually replaced with matching by aminoacyl-tRNA synthetases.[67][70][71]
  • Optimality: the genetic code continued to evolve after its initial creation, so that the current code maximizes some fitness function, usually some kind of error minimization.[67][68]

Hypotheses have addressed a variety of scenarios:[72]

  • Chemical principles govern specific RNA interaction with amino acids. Experiments with aptamers showed that some amino acids have a selective chemical affinity for their codons.[73] Experiments showed that of 8 amino acids tested, 6 show some RNA triplet-amino acid association.[61][71]
  • Biosynthetic expansion. The genetic code grew from a simpler earlier code through a process of "biosynthetic expansion". Primordial life "discovered" new amino acids (for example, as by-products of metabolism) and later incorporated some of these into the machinery of genetic coding.[74] Although much circumstantial evidence has been found to suggest that fewer amino acid types were used in the past,[75] precise and detailed hypotheses about which amino acids entered the code in what order are controversial.[76][77]
  • Natural selection has led to codon assignments of the genetic code that minimize the effects of mutations.[78] A recent hypothesis[79] suggests that the triplet code was derived from codes that used longer than triplet codons (such as quadruplet codons). Longer than triplet decoding would greater codon redundancy and would be more error resistant. This feature could allow accurate decoding absent complex translational machinery such as the ribosome, such as before cells began making ribosomes.
  • Information channels: Information-theoretic approaches model the process of translating the genetic code into corresponding amino acids as an error-prone information channel.[80] The inherent noise (that is, the error) in the channel poses the organism with a fundamental question: how can a genetic code be constructed to withstand noise[81] while accurately and efficiently translating information? These "rate-distortion" models[82] suggest that the genetic code originated as a result of the interplay of the three conflicting evolutionary forces: the needs for diverse amino-acids,[83] for error-tolerance[78] and for minimal resource cost. The code emerges at a transition when the mapping of codons to amino-acids becomes nonrandom. The code's emergence is governed by the topology defined by the probable errors and is related to the map coloring problem.[84]
  • Game theory: Models based on signaling games combine elements of game theory, natural selection and information channels. Such models have been used to suggest that the first polypeptides were likely short and had non-enzymatic function. Game theoretic models suggested that the organization of RNA strings into cells may have been necessary to prevent "deceptive" use of the genetic code, i.e. preventing the ancient equivalent of viruses from overwhelming the RNA world.[85]
  • Stop codons: Codons for translational stops are also an interesting aspect to the problem of the origin of the genetic code. As an example for addressing stop codon evolution, it has been suggested that the stop codons are such that they are most likely to terminate translation early in the case of a frame shift error.[86] In contrast, some stereochemical molecular models explain the origin of stop codons as "unassignable".[67]

References

  1. ^ Shu, Jian-Jun (2017). "A new integrated symmetrical table for genetic codes". BioSystems. 151: 21–26. doi:10.1016/j.biosystems.2016.11.004.
  2. ^ Turanov AA, Lobanov AV, Fomenko DE, Morrison HG, Sogin ML, Klobutcher LA, Hatfield DL, Gladyshev VN (January 2009). "Genetic code supports targeted insertion of two amino acids by one codon". Science. 323 (5911): 259–61. doi:10.1126/science.1164748. PMC 3088105. PMID 19131629.
  3. ^ Crick, Francis (10 July 1990). "Chapter 8: The genetic code". What Mad Pursuit: A Personal View of Scientific Discovery. Basic Books. pp. 89–101. ISBN 0-465-09138-5.
  4. ^ Nirenberg MW, Matthaei JH (October 1961). "The dependence of cell-free protein synthesis in E. coli upon naturally occurring or synthetic polyribonucleotides". Proceedings of the National Academy of Sciences of the United States of America. 47 (10): 1588–602. Bibcode:1961PNAS...47.1588N. doi:10.1073/pnas.47.10.1588. PMC 223178. PMID 14479932.
  5. ^ Gardner RS, Wahba AJ, Basilio C, Miller RS, Lengyel P, Speyer JF (December 1962). "Synthetic polynucleotides and the amino acid code. VII". Proceedings of the National Academy of Sciences of the United States of America. 48 (12): 2087–94. Bibcode:1962PNAS...48.2087G. doi:10.1073/pnas.48.12.2087. PMC 221128. PMID 13946552.
  6. ^ Wahba AJ, Gardner RS, Basilio C, Miller RS, Speyer JF, Lengyel P (January 1963). "Synthetic polynucleotides and the amino acid code. VIII". Proceedings of the National Academy of Sciences of the United States of America. 49 (1): 116–22. Bibcode:1963PNAS...49..116W. doi:10.1073/pnas.49.1.116. PMC 300638. PMID 13998282.
  7. ^ "The Nobel Prize in Physiology or Medicine 1959" (Press release). The Royal Swedish Academy of Science. 1959. Retrieved 27 February 2010. The Nobel Prize in Physiology or Medicine 1959 was awarded jointly to Severo Ochoa and Arthur Kornberg 'for their discovery of the mechanisms in the biological synthesis of ribonucleic acid and deoxyribonucleic acid'.
  8. ^ Nirenberg M, Leder P, Bernfield M, Brimacombe R, Trupin J, Rottman F, O'Neal C (May 1965). "RNA codewords and protein synthesis, VII. On the general nature of the RNA code". Proceedings of the National Academy of Sciences of the United States of America. 53 (5): 1161–8. Bibcode:1965PNAS...53.1161N. doi:10.1073/pnas.53.5.1161. PMC 301388. PMID 5330357.
  9. ^ "The Nobel Prize in Physiology or Medicine 1968" (Press release). The Royal Swedish Academy of Science. 1968. Retrieved 27 February 2010. The Nobel Prize in Physiology or Medicine 1968 was awarded jointly to Robert W. Holley, Har Gobind Khorana and Marshall W. Nirenberg 'for their interpretation of the genetic code and its function in protein synthesis'.
  10. ^ Edgar B (October 2004). "The genome of bacteriophage T4: an archeological dig". Genetics. 168 (2): 575–82. PMC 1448817. PMID 15514035.
  11. ^ Xie J, Schultz PG (December 2005). "Adding amino acids to the genetic repertoire". Current Opinion in Chemical Biology. 9 (6): 548–54. doi:10.1016/j.cbpa.2005.10.011. PMID 16260173.
  12. ^ Wang Q, Parrish AR, Wang L (March 2009). "Expanding the genetic code for biological studies". Chemistry & Biology. 16 (3): 323–36. doi:10.1016/j.chembiol.2009.03.001. PMC 2696486. PMID 19318213.
  13. ^ Simon, Matthew (7 January 2005). Emergent Computation: Emphasizing Bioinformatics. Springer Science & Business Media. pp. 105–106. ISBN 978-0-387-22046-8. {{cite book}}: Unknown parameter |name-list-format= ignored (|name-list-style= suggested) (help)
  14. ^ Hoesl, M. G.; Oehm, S.; Durkin, P.; Darmon, E.; Peil, L.; Aerni, H.-R.; Rappsilber, J.; Rinehart, J.; Leach, D.; Söll, D.; Budisa, N. (2015). "Chemical evolution of a bacterial proteome". Angewandte Chemie International Edition. 54: 10030–10034. doi:10.1002/anie.201502868. PMC 4782924. NIHMSID: NIHMS711205
  15. ^ "First stable semisynthetic organism created | KurzweilAI". www.kurzweilai.net. 3 February 2017. Retrieved 9 February 2017.
  16. ^ Zhang Y, Lamb BM, Feldman AW, Zhou AX, Lavergne T, Li L, Romesberg FE (February 2017). "A semisynthetic organism engineered for the stable expansion of the genetic alphabet". Proceedings of the National Academy of Sciences of the United States of America. 114 (6): 1317–1322. doi:10.1073/pnas.1616443114. PMC 5307467. PMID 28115716.
  17. ^ Han S, Yang A, Lee S, Lee HW, Park CB, Park HS (February 2017). "Expanding the genetic code of Mus musculus". Nature Communications. 8: 14568. doi:10.1038/ncomms14568. PMC 5321798. PMID 28220771.
  18. ^ a b King, Robert C.; Mulligan, Pamela; Stansfield, William (10 January 2013). A Dictionary of Genetics. OUP USA. p. 608. ISBN 978-0-19-976644-4. {{cite book}}: Unknown parameter |name-list-format= ignored (|name-list-style= suggested) (help)
  19. ^ 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 / Under the Auspices of the European Cell Biology Organization. 95 (3–4): 169–78. doi:10.1016/S0248-4900(03)00033-9. PMID 12867081.
  20. ^ Maloy S (29 November 2003). "How nonsense mutations got their names". Microbial Genetics Course. San Diego State University. Retrieved 10 March 2010.
  21. ^ References for the image are found in Wikimedia Commons page at: Commons:File:Notable mutations.svg#References.
  22. ^ Griffiths, Anthony J. F.; Miller, Jeffrey H.; Suzuki, David T.; Lewontin, Richard C.; et al., eds. (2000). "Spontaneous mutations". An Introduction to Genetic Analysis (7th ed.). New York: W. H. Freeman. ISBN 0-7167-3520-2. {{cite book}}: External link in |chapterurl= (help); Unknown parameter |chapterurl= ignored (|chapter-url= suggested) (help); Unknown parameter |name-list-format= ignored (|name-list-style= suggested) (help)
  23. ^ Freisinger E, Grollman AP, Miller H, Kisker C (April 2004). "Lesion (in)tolerance reveals insights into DNA replication fidelity". The EMBO Journal. 23 (7): 1494–505. doi:10.1038/sj.emboj.7600158. PMC 391067. PMID 15057282.
  24. ^ (Boillée 2006, p. 39)
  25. ^ Chang JC, Kan YW (June 1979). "beta 0 thalassemia, a nonsense mutation in man". Proceedings of the National Academy of Sciences of the United States of America. 76 (6): 2886–9. Bibcode:1979PNAS...76.2886C. doi:10.1073/pnas.76.6.2886. PMC 383714. PMID 88735.
  26. ^ Boillée S, Vande Velde C, Cleveland DW (October 2006). "ALS: a disease of motor neurons and their nonneuronal neighbors". Neuron. 52 (1): 39–59. doi:10.1016/j.neuron.2006.09.018. PMID 17015226.
  27. ^ Isbrandt D, Hopwood JJ, von Figura K, Peters C (1996). "Two novel frameshift mutations causing premature stop codons in a patient with the severe form of Maroteaux-Lamy syndrome". Human Mutation. 7 (4): 361–3. doi:10.1002/(SICI)1098-1004(1996)7:4<361::AID-HUMU12>3.0.CO;2-0. PMID 8723688.
  28. ^ Crow JF (1993). "How much do we know about spontaneous human mutation rates?". Environmental and Molecular Mutagenesis. 21 (2): 122–9. doi:10.1002/em.2850210205. PMID 8444142.
  29. ^ Lewis, Ricki (2005). Human Genetics: Concepts and Applications (6th ed.). Boston, Mass: McGraw Hill. pp. 227–228. ISBN 0-07-111156-5. {{cite book}}: Unknown parameter |name-list-format= ignored (|name-list-style= suggested) (help)
  30. ^ Sawyer SA, Parsch J, Zhang Z, Hartl DL (April 2007). "Prevalence of positive selection among nearly neutral amino acid replacements in Drosophila". Proceedings of the National Academy of Sciences of the United States of America. 104 (16): 6504–10. Bibcode:2007PNAS..104.6504S. doi:10.1073/pnas.0701572104. PMC 1871816. PMID 17409186.
  31. ^ Bridges KR (2002). "Malaria and the Red Cell". Harvard. Archived from the original on 27 November 2011. {{cite journal}}: Unknown parameter |deadurl= ignored (|url-status= suggested) (help)
  32. ^ Drake JW, Holland JJ (November 1999). "Mutation rates among RNA viruses". Proceedings of the National Academy of Sciences of the United States of America. 96 (24): 13910–3. Bibcode:1999PNAS...9613910D. doi:10.1073/pnas.96.24.13910. PMC 24164. PMID 10570172.
  33. ^ Holland J, Spindler K, Horodyski F, Grabau E, Nichol S, VandePol S (March 1982). "Rapid evolution of RNA genomes". Science. 215 (4540): 1577–85. Bibcode:1982Sci...215.1577H. doi:10.1126/science.7041255. PMID 7041255.
  34. ^ de Visser JA, Rozen DE (April 2006). "Clonal interference and the periodic selection of new beneficial mutations in Escherichia coli". Genetics. 172 (4): 2093–100. doi:10.1534/genetics.105.052373. PMC 1456385. PMID 16489229.
  35. ^ Watson, James D. (2008). Molecular Biology of the Gene. Pearson/Benjamin Cummings. ISBN 978-0-8053-9592-1.
  36. ^ Michel-Beyerle, Maria Elisabeth (1990). Reaction centers of photosynthetic bacteria: Feldafing-II-Meeting. Springer-Verlag. ISBN 978-3-540-53420-4.
  37. ^ Füllen G, Youvan DC (1994). "Genetic Algorithms and Recursive Ensemble Mutagenesis in Protein Engineering". Complexity International 1.
  38. ^ "Codon usage table".
  39. ^ 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.
  40. ^ Nakamoto T (March 2009). "Evolution and the universality of the mechanism of initiation of protein synthesis". Gene. 432 (1–2): 1–6. doi:10.1016/j.gene.2008.11.001. PMID 19056476.
  41. ^ Asano, K (2014). "Why is start codon selection so precise in eukaryotes?". Translation (Austin, Tex.). 2 (1): e28387. doi:10.4161/trla.28387. PMID 26779403.
  42. ^ Brenner S. A Life in Science (2001) Published by Biomed Central Limited ISBN 0-9540278-0-9 see pages 101-104
  43. ^ Edgar B (2004). "The genome of bacteriophage T4: an archeological dig". Genetics. 168 (2): 575–82. PMC 1448817. PMID 15514035. see pages 580-581
  44. ^ a b c Elzanowski A, Ostell J (7 April 2008). "The Genetic Codes". National Center for Biotechnology Information (NCBI). Retrieved 10 March 2010.
  45. ^ Krzycki JA (December 2005). "The direct genetic encoding of pyrrolysine". Current Opinion in Microbiology. 8 (6): 706–12. doi:10.1016/j.mib.2005.10.009. PMID 16256420.
  46. ^ Zhang Y, Baranov PV, Atkins JF, Gladyshev VN (May 2005). "Pyrrolysine and selenocysteine use dissimilar decoding strategies". The Journal of Biological Chemistry. 280 (21): 20740–51. doi:10.1074/jbc.M501458200. PMID 15788401.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  47. ^ Prat L, Heinemann IU, Aerni HR, Rinehart J, O'Donoghue P, Söll D (December 2012). "Carbon source-dependent expansion of the genetic code in bacteria". Proceedings of the National Academy of Sciences of the United States of America. 109 (51): 21070–5. Bibcode:2012PNAS..10921070P. doi:10.1073/pnas.1218613110. PMC 3529041. PMID 23185002.
  48. ^ Crick FH, Orgel LE (1973). "Directed panspermia". Icarus. 19 (3): 341–6, 344. Bibcode:1973Icar...19..341C. doi:10.1016/0019-1035(73)90110-3. It is a little surprising that organisms with somewhat different codes do not coexist. (Further discussion)
  49. ^ Barrell BG, Bankier AT, Drouin J (1979). "A different genetic code in human mitochondria". Nature. 282 (5735): 189–194. doi:10.1038/282189a0. PMID 226894. ([1])
  50. ^ Jukes TH, Osawa S (December 1990). "The genetic code in mitochondria and chloroplasts". Experientia. 46 (11–12): 1117–26. doi:10.1007/BF01936921. PMID 2253709.
  51. ^ Fitzpatrick DA, Logue ME, Stajich JE, Butler G (1 January 2006). "A fungal phylogeny based on 42 complete genomes derived from supertree and combined gene analysis". BMC Evolutionary Biology. 6: 99. doi:10.1186/1471-2148-6-99. PMC 1679813. PMID 17121679.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  52. ^ Santos MA, Tuite MF (May 1995). "The CUG codon is decoded in vivo as serine and not leucine in Candida albicans". Nucleic Acids Research. 23 (9): 1481–6. doi:10.1093/nar/23.9.1481. PMC 306886. PMID 7784200.
  53. ^ Butler G, Rasmussen MD, Lin MF, et al. (June 2009). "Evolution of pathogenicity and sexual reproduction in eight Candida genomes". Nature. 459 (7247): 657–62. Bibcode:2009Natur.459..657B. doi:10.1038/nature08064. PMC 2834264. PMID 19465905.
  54. ^ Witzany G (2016). "Crucial steps to life: From chemical reactions to code using agents". Biosystems. 140: 49–57. doi:10.1016/j.biosystems.2015.12.007. PMID 26723230.
  55. ^ Taylor DJ, Ballinger MJ, Bowman SM, Bruenn JA (2013). "Virus-host co-evolution under a modified nuclear genetic code". PeerJ. 1: e50. doi:10.7717/peerj.50. PMC 3628385. PMID 23638388.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  56. ^ Hofhuis J, Schueren F, Nötzel C, Lingner T, Gärtner J, Jahn O, Thoms S (2016). "The functional readthrough extension of malate dehydrogenase reveals a modification of the genetic code". Open Biol. 6 (11): 160246. doi:10.1098/rsob.160246. PMC 5133446. PMID 27881739.
  57. ^ Schueren F, Lingner T, George R, Hofhuis J, Gartner J, Thoms S (2014). "Peroxisomal lactate dehydrogenase is generated by translational readthrough in mammals". eLife. 3: e03640. doi:10.7554/eLife.03640. PMC 4359377. PMID 25247702.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  58. ^ F. Schueren und S. Thoms (2016). "Functional Translational Readthrough: A Systems Biology Perspective". PLOS Genetics. 12 (8): e1006196. doi:10.1371/journal.pgen.1006196. PMC 4973966. PMID 27490485.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  59. ^ Dutilh BE, Jurgelenaite R, Szklarczyk R, van Hijum SA, Harhangi HR, Schmid M, de Wild B, Françoijs KJ, Stunnenberg HG, Strous M, Jetten MS, Op den Camp HJ, Huynen MA (July 2011). "FACIL: Fast and Accurate Genetic Code Inference and Logo". Bioinformatics. 27 (14): 1929–33. doi:10.1093/bioinformatics/btr316. PMC 3129529. PMID 21653513.
  60. ^ Ribas de Pouplana L, Turner RJ, Steer BA, Schimmel P (September 1998). "Genetic code origins: tRNAs older than their synthetases?". Proceedings of the National Academy of Sciences of the United States of America. 95 (19): 11295–300. Bibcode:1998PNAS...9511295D. doi:10.1073/pnas.95.19.11295. PMC 21636. PMID 9736730.
  61. ^ a b Yarus, Michael (2010). Life from an RNA World: The Ancestor Within. Harvard University Press. ISBN 978-0-674-05075-4.
  62. ^ "Mathematica function for # possible arrangements of items in bins? - Online Technical Discussion Groups—Wolfram Community". community.wolfram.com. Retrieved 3 February 2017.
  63. ^ a b Freeland SJ, Hurst LD (September 1998). "The genetic code is one in a million". Journal of Molecular Evolution. 47 (3): 238–48. doi:10.1007/PL00006381. PMID 9732450.
  64. ^ Taylor FJ, Coates D (1989). "The code within the codons". Bio Systems. 22 (3): 177–87. doi:10.1016/0303-2647(89)90059-2. PMID 2650752.
  65. ^ Di Giulio M (October 1989). "The extension reached by the minimization of the polarity distances during the evolution of the genetic code". Journal of Molecular Evolution. 29 (4): 288–93. doi:10.1007/BF02103616. PMID 2514270.
  66. ^ Wong JT (February 1980). "Role of minimization of chemical distances between amino acids in the evolution of the genetic code". Proceedings of the National Academy of Sciences of the United States of America. 77 (2): 1083–6. Bibcode:1980PNAS...77.1083W. doi:10.1073/pnas.77.2.1083. PMC 348428. PMID 6928661.
  67. ^ a b c d Erives A (August 2011). "A model of proto-anti-codon RNA enzymes requiring L-amino acid homochirality". Journal of Molecular Evolution. 73 (1–2): 10–22. doi:10.1007/s00239-011-9453-4. PMC 3223571. PMID 21779963.
  68. ^ a b Freeland SJ, Knight RD, Landweber LF, Hurst LD (April 2000). "Early fixation of an optimal genetic code". Molecular Biology and Evolution. 17 (4): 511–18. doi:10.1093/oxfordjournals.molbev.a026331. PMID 10742043.
  69. ^ Crick FH (December 1968). "The origin of the genetic code". Journal of Molecular Evolution. 38 (3): 367–79. doi:10.1016/0022-2836(68)90392-6. PMID 4887876.
  70. ^ Hopfield JJ (1978). "Origin of the genetic code: a testable hypothesis based on tRNA structure, sequence, and kinetic proofreading". PNAS. 75 (9): 4334–4338. doi:10.1073/pnas.75.9.4334. PMC 336109. PMID 279919.
  71. ^ a b Yarus M, Widmann JJ, Knight R (November 2009). "RNA-amino acid binding: a stereochemical era for the genetic code". Journal of Molecular Evolution. 69 (5): 406–29. doi:10.1007/s00239-009-9270-1. PMID 19795157.
  72. ^ Knight RD, Freeland SJ, Landweber LF (June 1999). "Selection, history and chemistry: the three faces of the genetic code". Trends in Biochemical Sciences. 24 (6): 241–7. doi:10.1016/S0968-0004(99)01392-4. PMID 10366854.
  73. ^ Knight RD, Landweber LF (September 1998). "Rhyme or reason: RNA-arginine interactions and the genetic code". Chemistry & Biology. 5 (9): R215–20. doi:10.1016/S1074-5521(98)90001-1. PMID 9751648.
  74. ^ Sengupta S, Higgs PG (2015). "Pathways of genetic code evolution in ancient and modern organisms". Journal of Molecular Evolution. 80 (5–6): 229–243. doi:10.1007/s00239-015-9686-8. PMID 26054480.
  75. ^ Brooks DJ, Fresco JR, Lesk AM, Singh M (October 2002). "Evolution of amino acid frequencies in proteins over deep time: inferred order of introduction of amino acids into the genetic code". Molecular Biology and Evolution. 19 (10): 1645–55. doi:10.1093/oxfordjournals.molbev.a003988. PMID 12270892.
  76. ^ Amirnovin R (May 1997). "An analysis of the metabolic theory of the origin of the genetic code". Journal of Molecular Evolution. 44 (5): 473–6. doi:10.1007/PL00006170. PMID 9115171.
  77. ^ Ronneberg TA, Landweber LF, Freeland SJ (December 2000). "Testing a biosynthetic theory of the genetic code: fact or artifact?". Proceedings of the National Academy of Sciences of the United States of America. 97 (25): 13690–5. Bibcode:2000PNAS...9713690R. doi:10.1073/pnas.250403097. PMC 17637. PMID 11087835.
  78. ^ a b Freeland SJ, Wu T, Keulmann N (October 2003). "The case for an error minimizing standard genetic code". Origins of Life and Evolution of the Biosphere. 33 (4–5): 457–77. doi:10.1023/A:1025771327614. PMID 14604186.
  79. ^ Baranov PV, Venin M, Provan G (2009). Gemmell NJ (ed.). "Codon size reduction as the origin of the triplet genetic code". PLOS ONE. 4 (5): e5708. Bibcode:2009PLoSO...4.5708B. doi:10.1371/journal.pone.0005708. PMC 2682656. PMID 19479032. {{cite journal}}: Unknown parameter |name-list-format= ignored (|name-list-style= suggested) (help)CS1 maint: unflagged free DOI (link)
  80. ^ Tlusty T (November 2007). "A model for the emergence of the genetic code as a transition in a noisy information channel". Journal of Theoretical Biology. 249 (2): 331–42. doi:10.1016/j.jtbi.2007.07.029. PMID 17826800.
  81. ^ Sonneborn TM (1965). Bryson V, Vogel H (eds.). Evolving genes and proteins. New York: Academic Press. pp. 377–397.
  82. ^ Tlusty T (February 2008). "Rate-distortion scenario for the emergence and evolution of noisy molecular codes". Physical Review Letters. 100 (4): 048101. arXiv:1007.4149. Bibcode:2008PhRvL.100d8101T. doi:10.1103/PhysRevLett.100.048101. PMID 18352335.
  83. ^ Sella G, Ardell DH (September 2006). "The coevolution of genes and genetic codes: Crick's frozen accident revisited". Journal of Molecular Evolution. 63 (3): 297–313. doi:10.1007/s00239-004-0176-7. PMID 16838217.
  84. ^ Tlusty T (September 2010). "A colorful origin for the genetic code: information theory, statistical mechanics and the emergence of molecular codes". Physics of Life Reviews. 7 (3): 362–76. arXiv:1007.3906. Bibcode:2010PhLRv...7..362T. doi:10.1016/j.plrev.2010.06.002. PMID 20558115.
  85. ^ Jee J, Sundstrom A, Massey SE, Mishra B (November 2013). "What can information-asymmetric games tell us about the context of Crick's 'frozen accident'?". Journal of the Royal Society, Interface / the Royal Society. 10 (88): 20130614. doi:10.1098/rsif.2013.0614. PMC 3785830. PMID 23985735.
  86. ^ Itzkovitz S, Alon U (2007). "The genetic code is nearly optimal for allowing additional information within protein-coding sequences". Genome Research. 17 (4): 405–412. doi:10.1101/gr.5987307. PMC 1832087. PMID 17293451.

Further reading