Download Genetic code ambiguity: an unexpected source of proteome

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Interactome wikipedia , lookup

Point mutation wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Peptide synthesis wikipedia , lookup

Personalized medicine wikipedia , lookup

Expression vector wikipedia , lookup

Protein–protein interaction wikipedia , lookup

Messenger RNA wikipedia , lookup

Metabolism wikipedia , lookup

Protein wikipedia , lookup

Genetic engineering wikipedia , lookup

Two-hybrid screening wikipedia , lookup

Proteolysis wikipedia , lookup

Epitranscriptome wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Biochemistry wikipedia , lookup

Amino acid synthesis wikipedia , lookup

Transfer RNA wikipedia , lookup

Biosynthesis wikipedia , lookup

Genetic code wikipedia , lookup

Transcript
COMICR-698; NO OF PAGES 7
Available online at www.sciencedirect.com
Genetic code ambiguity: an unexpected source of proteome
innovation and phenotypic diversity
Gabriela R Moura, Laura C Carreto and Manuel AS Santos
Translation of the genome into the proteome is a highly
accurate biological process. However, the molecular
mechanisms involved in protein synthesis are not error free and
downstream protein quality control systems are needed to
counteract the negative effects of translational errors
(mistranslation) on proteome and cell homeostasis. This plus
human and mice diseases caused by translational error
generalized the idea that codon ambiguity is detrimental to life.
Here we depart from this classical view of deleterious
translational error and highlight how codon ambiguity can play
important roles in the evolution of novel proteins. We also
explain how tRNA mischarging can be relevant for the synthesis
of functional proteomes, how codon ambiguity generates
phenotypic and genetic diversity and how advantageous
phenotypes can be selected, fixed, and inherited. A brief
introduction to the molecular nature of translational error is
provided; however, detailed information on the mechanistic
aspects of mistranslation or comprehensive literature reviews
of this topic should be obtained elsewhere.
Address
Department of Biology and CESAM, University of Aveiro, 3810-193
Aveiro, Portugal
Corresponding author: Santos, Manuel AS ([email protected])
Current Opinion in Microbiology 2009, 12:1–7
This review comes from a themed issue on
Growth and development: eukaryotes
Edited by Judith Berman
1369-5274/$ – see front matter
# 2009 Elsevier Ltd. All rights reserved.
DOI 10.1016/j.mib.2009.09.004
Introduction
Genome translation errors can arise during tRNA charging by aminoacyl-tRNA synthetases (aaRSs) and during
mRNA decoding by the ribosome. Aminoacylation errors
are mainly caused either by failure of the aaRSs to
differentiate between amino acids with similar chemical
properties or by the incorrect recognition of tRNAs. Such
errors are minimized by aaRSs editing mechanisms,
which discard incorrectly bound amino acids, and by
highly specific tRNA–aaRS interaction networks [1].
At the ribosome level, mRNA decoding is affected by
four main types of errors: first, missense errors cause
www.sciencedirect.com
incorrect amino acid incorporation into growing polypeptide chains and result in the synthesis of mutant proteins;
second, nonsense errors cause readthrough of stop codons
and produce proteins with extended C-termini; third,
frameshifting errors alter the mRNA reading frame normally to the 1 or to the +1 frames, producing out-offrame truncated proteins; and fourth, processivity errors
terminate translation prematurely and also produce truncated proteins (Figure 1) [2].
On average, the frequency of translational errors is in the
order of 104 [1]. These are regarded as physiological
errors and quality control systems minimize the toxic
effects of mistranslated proteins. However, the fidelity
of protein synthesis is influenced by intracellular and
extracellular factors and fluctuates over time and along
mRNA reading frames. Translation errors increase sharply under amino acid starvation and are affected by codon
usage, codon context, mRNA structure, aminoacyl-tRNA
concentration, and tRNA modification, which in turn are
influenced by cellular metabolism [3,4–7]. Also, in yeast
and probably in all eukaryotes, translation termination
fidelity is controlled by the [PSI+] factor that is a prion
form of the eukaryotic translation termination factor 3
(eRF3) [8,9]. Environmental conditions that promote
[PSI+] formation increase suppression of stop codons
and synthesis of proteins with extended C-termini [10].
Therefore, translational error is dynamic and can be
regulated. We discuss below how natural selection
explored mistranslation to expand the genetic code and
to create novel functional classes of proteins. We also
discuss the evolutionary implications of recent studies
showing that codon ambiguity generates phenotypic
variability and genetic diversity. For simplicity, throughout the text, translational error, mistranslation, genetic
code ambiguity, and codon ambiguity have identical
meaning.
tRNA mischarging as a source of proteome
innovation
The evolutionary relevance of genetic code ambiguity is
unequivocally demonstrated by the natural reassignments
of UGA and UAG stop codons to selenocysteine (Sec) and
pyrrolysine (Pyl), respectively, and by the widespread
tRNA-dependent synthesis of asparagine (Asn), glutamine (Gln), and cysteine (Cys) [11,12,13], as discussed
below.
Selenocysteine (Sec) replaces cysteine (Cys) in the active
site of a group of redox proteins generically known as
Current Opinion in Microbiology 2009, 12:1–7
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004
COMICR-698; NO OF PAGES 7
2 Growth and development: eukaryotes
Figure 1
Sources of genetic code ambiguity during genome translation into the proteome. Codons may be misassigned to amino acids due to tRNA
mischarging by aminoacyl-tRNA synthetases (aaRSs) or due to codon misreading during mRNA decoding in the ribosome. Mischarging of tRNAs may
be caused by the failure of aaRSs to recognize their cognate tRNAs or by the activation of incorrectly bound amino acids. To minimize mischarging,
some aaRSs have editing mechanisms that discard chemically similar amino acids from their active sites. In some organisms, tRNA mischarging is
necessary for the synthesis of selenocysteine, asparagine, glutamine, and cysteine. Codon decoding errors are mainly caused by misreading of sense
codons (missense errors) and nonsense codons (nonsense errors) by near cognate or noncognate tRNAs. These errors result in the synthesis of
mutant proteins. Other mRNA decoding errors are caused by loss of the reading frame (frameshifting) or premature ribosome drop off from the mRNA
(processivity errors). These errors result in premature translation termination and synthesis of truncated polypeptides.
selenoproteins. Sec is translationally inserted at specific
UGA codons which are reprogramed by a complex translational machinery called the selenosome [11,14]. Initial
charging of a unique tRNASec with serine (Ser) by a seryltRNA synthetase (SerRS) produces a mischarged tRNA
species (Ser-tRNASec) whose Ser moiety is then converted to selenocysteyl-tRNASec [Sec-tRNASec] by a
selenocysteine synthase (SelA), using selenophosphate
as substrate [12]. The existence of the mischarged SertRNASec suggests, therefore, that Ser can be misincorporated at Sec-UGA codons, creating codon ambiguity.
Expression of the Eubacterium acidaminophilum peroxiredoxin PrxU gene in Escherichia coli results in the misincorporation of selenocysteine, tryptophan, and serine at
the UGA-Sec codons [14], confirming that hypothesis.
In the ciliate Euplotes crassus cysteine and selenocysteine
are naturally inserted at UGA codons [15]. This ciliate
genome encodes two tRNAs with cognate anticodons for
the UGA codon, namely the tRNAUCASec and the tRNAUCys
. Its thioredoxin reductase genes (eTR1 and eTR2)
CA
contain seven UGAs, the first six direct insertion of Cys
while the last one is used to insert Sec. Cys is incorporated
by the tRNAUCACys because of failure of the E. crassus
eRF1 to recognize the UGA stop codon and Sec is
Current Opinion in Microbiology 2009, 12:1–7
incorporated by the tRNAUCASec through recoding of
the UGA by a SECIS element present in the 30 -UTR
of eTR1 and eTR2. Cys and Sec incorporation at those
UGAs are necessary to produce active eTR1 and eTR2
[15], thus demonstrating that such ambiguity is functional.
Pyrrolysine (Pyl) provides additional evidence for positive
roles of genetic code ambiguity. Pyl is cotranslationally
inserted into the active center of methyltransferases of
Methanosarcineace species, Desulfitobacterium hafniense, and
in the glutless worm Olavius algarvensis [16]. Pyl is charged
on a dedicated nonsense suppressor tRNA (tRNAPyl) by a
pyrrolysyl-tRNA synthetase (PylRS) that catalyzes the
formation of Pyl-tRNAPyl [12]. Surprisingly, the PylRS
has broad amino acid specificity and binds various nonnatural lysine derivatives, namely Ne-D-prolyl-L-lysine, Necyclopentyloxycarbonyl-L-lysine (Cyc), Ne-(tert-butoxycarbonyl (t-Boc))-L-lysine (Boc-lysine), Ne-t-Boc-L-6-amino2-hydroxyhexanoic acid (Boc-LysOH) [17,18]. These
amino acids are mischarged onto the tRNAPyl by PylRS
and can be inserted into the genetic code of E. coli, yeast, or
mammalian cells using the orthogonal PylRS.tRNAPyl pair
[18,19,20]. Furthermore, the Methanosarcina barkeri
Fusaro tRNAPyl can be charged with serine in vitro by
www.sciencedirect.com
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004
COMICR-698; NO OF PAGES 7
Mistranslation generates phenotypic diversity Moura, Carreto and Santos 3
one of its SerRS enzymes and in vivo by the E. coli SerRS
[19], that is, it can exist as Pyl-tRNAPyl or Ser-tRNAPyl.
This broad amino acid specificity of the PylRS and ancestral competition of the tRNAPyl with release factor 1 (RF1)
for UAG codons strongly suggest that the ancestral UAGPyl was ambiguous and that its recoding evolved to minimize such ambiguity.
Other cases of natural tRNA ambiguity are illustrated by
tRNA-dependent asparagine (Asn), glutamine (Gln), and
cysteine (Cys) biosynthesis in various organisms where
AsnRS, GlnRS, and CysRS do not exist (reviewed in
[13]). In most bacteria and archaea Asn is synthesized
directly on a tRNAAsn which is initially charged with Asp
by a nondiscriminating AspRS (Asp-tRNAAsn). In all
known archaea, most bacteria and chloroplasts, Gln is
synthesized on a tRNAGln that is mischarged with Glu by
a nondiscriminating GluRS (Glu-tRNAGln) [21,22].
Amido transferases (Glu-AdT and Asp-AdT) carry out
the conversion of Asp and Glu into Asn and Gln and also
phosphorylate the mischarged tRNAs producing two
additional mischarged intermediate tRNA species,
namely g-phosphoryl-Glu-tRNAGln (P-Glu-tRNAGln)
(P-Asp-tRNAAsn)
and
b-phosphoryl-Asp-tRNAAsn
[23,24]. Similarly, in most methanogenic archaea,
cysteine (Cys) is synthesized on a tRNACys that is initially
charged with O-phosphoserine (Sep) by the enzyme Ophosphoseryl-tRNA synthase (SepRS) (Sep-tRNACys). A
Sep-tRNA:Cys-tRNA synthase (SepCysS) then transforms Sep-tRNACys into Cys-tRNACys [25,26,27,28].
Considering that mischarged tRNA can participate in
mRNA translation [29,30,31], it is reasonable to assume
that Asn, Gln, and Cys were also incorporated into the
genetic code of the ancestor of modern organisms through
ambiguous translation of the respective codons. It is not
yet clear why Asn, Gln, and Cys are synthesized directly
on mischarged tRNAs in these organisms, but these
alternative biosynthetic pathways may integrate amino
acid metabolism, protein synthesis, and cellular signaling
pathways providing new layers of cellular control [13].
Codon ambiguity as a generator of phenotypic
diversity
In several species of the genus Candida thousands of
leucine CUG codons have been reassigned to serine using
a novel serine tRNA (tRNACAGSer) [30,32]. These CUG
codons remain ambiguous in extant Candida species due
to charging of the tRNACAGSer by both the SerRS and the
LeuRS [33]. This leads to the formation of a correctly
charged Ser-tRNACAGSer and a mischarged Leu-tRNASer
which participate in translation. Under laboratory
CAG
conditions, serine and leucine are incorporated into the
Figure 2
The consequences of genetic code ambiguity. Codon ambiguity results in the synthesis of mutant proteins, which misfold and are targeted for
degradation by the ubiquitin–proteasome quality control system. Alternatively, such proteins aggregate and accumulate in the cell forming toxic
protein aggregates. Some mutant proteins may still fold with the help of molecular chaperones and may retain their functions or may acquire new
functions. Accumulation of aggregated and soluble mutant proteins and overloading of protein quality control systems with misfolded proteins activate
the stress response and reprogram gene expression, leading to exposure of hidden phenotypic variation. Mutant DNA polymerases and DNA repair
enzymes create hypermutagenic clones with high adaptation potential. Under these genetic code ambiguity conditions, cell viability, and homeostasis
are guarantied by wild type proteins that represent the major component of the proteome.
www.sciencedirect.com
Current Opinion in Microbiology 2009, 12:1–7
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004
COMICR-698; NO OF PAGES 7
4 Growth and development: eukaryotes
Figure 3
Model for fixation and inheritance of phenotypes associated with genetic code ambiguity. Ambiguous codon decoding expands the proteome and
generates new phenotypes creating phenotypic diversity. Selection of advantageous phenotypes creates positive feedback pressure that maintains
ambiguous codon decoding. This ultimately leads to synthesis of mutant DNA polymerases and DNA repair enzymes and emergence of
hypermutagenic clones and to an exponential increase in genome mutational load. This accelerates the fixation of advantageous phenotypes and their
transmission to the progeny.
Candida albicans proteome with efficiencies of 97% and
3%, respectively, but such CUG ambiguity increases up
to 5% under stress. C. albicans tolerates up to 28% of
leucine misincorporation without visible effects on
growth rate, suggesting that CUG ambiguity may fluctuate between 3% and 28%, depending on environmental
conditions. Remarkably, CUG ambiguity increases
secretion of lipases and proteinases and cell adhesion
and spawns a wide variety of colony morphologies
[34,35]. In S. cerevisiae, similar CUG ambiguity does
not generate such phenotypic and morphological diversity; however, it increases tolerance to various environmental stressors and permits growth in the presence of
lethal doses of cycloheximide, arsenite, cadmium, and salt
[36]. Most interestingly, the phenotypic diversity
spawned by CUG ambiguity in C. albicans is similar to
that spawned in S. cerevisiae by the [PSI+] prion [37]. The
[PSI+] prion reduces decoding efficiency of the translation termination machinery (eRF1–eRF3 complex)
leading to readthrough of UAG, UAA, and UGA stop
codons and to synthesis of proteins with extended Ctermini. Accumulation of such aberrant proteins spawns a
wide variety of growth and morphological phenotypes
with high selective potential [10]. [PSI+] also controls the
biosynthesis of polyamines by regulating the expression
of ornithine decarboxilase antizyme through a +1 frameshift at a shifty-UGA site [38]. Therefore, [PSI+]
mediated stop codon ambiguity in S. cerevisiae and
CUG ambiguity in C. albicans are used as sources of
phenotypic variation and for novel regulatory mechanisms (Figure 2).
Generation of genetic diversity and
inheritance of phenotypic traits
An important question related to the phenotypic diversity
that arises from genetic code ambiguity is how can it be
Current Opinion in Microbiology 2009, 12:1–7
selected and inherited? Clearly, advantageous phenotypes exposed by codon ambiguity can only be evolutionarily relevant if transmitted to the progeny. There are
two fundamental difficulties here. Firstly, the phenotypic
diversity generated by codon ambiguity is diverse and
somewhat stochastic. Secondly, in general, codon ambiguity decreases fitness and should be eliminated by
negative selection. The discovery in E. coli and other
bacteria of a hypermutagenesis phenotype associated
with codon ambiguity, the ‘translational stress mutagenesis phenotype’ (TSM), provides a fascinating solution for
this apparent genetic paradox [39–43]. A mutant tRNAGly
that misincorporates glycine at aspartate (GAC/U)
codons, various mutant tRNAAla that misread noncognate
codons, ambiguous ribosomes, and the mistranslation
drug streptomycin all induce spontaneous hypermutagenesis in E. coli via synthesis of mutant DNA polymerase III
and RecABC/RuvABC complexes [39,40,41]. This
increased genome mutation rate generates genetic diversity and, therefore, provides an indirect mechanism for
rapid fixation of the advantageous phenotypes linked to
genetic code ambiguity (Figure 3).
Conclusions
The view that codon ambiguity is an aberration of nature
is clearly a poor interpretation of the biology of this old
phenomenon. There is no doubt that above a certain
mistranslation threshold the proteome is disrupted, cell
fitness decreases, and cell death increases [31,44,45]. But,
codon ambiguity can increase proteome diversity by
creating statistical populations of proteins and can generate genetic and phenotypic diversity, which can be
explored by natural selection for developmental, metabolic, and regulatory innovation. It is now clear that
bacteria and eukaryotes are far more resistant to mistranslation than anticipated and that mischarged tRNAs can
www.sciencedirect.com
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004
COMICR-698; NO OF PAGES 7
Mistranslation generates phenotypic diversity Moura, Carreto and Santos 5
Table 1
Some known cases of functional genetic code ambiguity
Function
Organisms
Reference
Natural ambiguity
Ser-tRNA Sec
Asp-tRNA Asn
P-Asp-tRNA Asn
Glu-tRNA Gln
P-Glu-tRNA Gln
Sep-tRNA Cys
Leu-tRNA Ser
PSI prion
Selenocysteine incorporation
Asp synthesis
Asp synthesis
Gln synthesis
Gln synthesis
Cys synthesis
Leu misincorporation
Readthrough of stop codons
Most prokaryotes and eukaryotes
Most bacteria, archaea, and mitochondria
Most bacteria, archaea, and mitochondria
Most bacteria, archaea, and mitochondria
Most bacteria, archaea, and mitochondria
Most bacteria, archaea, and mitochondria
Various Candida species
Yeast and other eukaryotes
[11,12]
[13]
[13]
[13]
[13]
[13]
[32,33]
[37]
Artificial ambiguity
aa*-tRNA–aaRS
Orthogonal pairs
Cys-tRNAPro
Ser-tRNAThr
X*-tRNA Pyl
Genetic code expansion
Mistranslation
Mistranslation
Genetic code expansion
Escherichia coli, yeast, and mammalian cells
E. coli
E. coli
E. coli
[50]
[29]
[29]
[19]
Ambiguity at the tRNA charging level is not always translated into proteins due to translational buffering systems, namely the capacity of the
elongation factor Tu (EF-Tu) to negatively discriminate mischarged tRNAs. However, naturally mischarged tRNAs should have participated in
translation before such buffering systems appeared. This is corroborated by lack of such buffering systems in E. coli, yeast, and mammalian cells
engineered for the incorporation of both artificial and natural amino acids through orthogonal tRNA–aaRS pairs [49,50]. X* indicates artificial
derivatives of lysine.
participate in mRNA translation (Table 1) [29,46].
The high cellular tolerance to mistranslation opens the
possibility for the evolution of proteomic and phenotypic
novelty, which can be fixed rapidly through hypermutagenesis. In other words, genetic code ambiguity is an
epigenetic evolution evolvability system much similar to
[PSI+] or Hsp90 [47,48,49].
5.
Parker J, Precup J: Mistranslation during phenylalanine
starvation. Mol Gen Genet 1986, 204:70-74.
6.
Precup J, Parker J: Missense misreading of asparagine codons
as a function of codon identity and context. J Biol Chem 1987,
262:11351-11355.
7.
Precup J, Ulrich AK, Roopnarine O, Parker J: Context specific
misreading of phenylalanine codons. Mol Gen Genet 1989,
218:397-401.
8.
Stansfield I, Jones KM, Kushnirov VV, Dagkesamanskaya AR,
Poznyakovski AI, Paushkin SV, Nierras CR, Cox BS, TerAvanesyan MD, Tuite MF: The products of the SUP45 (eRF1) and
SUP35 genes interact to mediate translation termination in
Saccharomyces cerevisiae. EMBO J 1995, 14:4365-4373.
9.
Song H, Mugnier P, Das AK, Webb HM, Evans DR, Tuite MF,
Hemmings BA, Barford D: The crystal structure of human
eukaryotic release factor eRF1 — mechanism of stop
codon recognition and peptidyl-tRNA hydrolysis. Cell 2000,
100:311-321.
Acknowledgements
We are thankful to the Portuguese Foundation for Science and Technology
and to the Human Frontier Science Program for funding our work through
projects PTDC/BIA-BCM/64745 and RGP0045/2005-C, respectively. We
are also grateful to Dieter Söll and Judith Berman for reading and
commenting on the manuscript before publication.
References and recommended reading
Papers of particular interest, published within the period of review,
have been highlighted as:
of special interest
of outstanding interest
1. Ling J, Reynolds N, Ibba M: Aminoacyl-tRNA synthesis and
translational quality control. Annu Rev Microbiol 2009, 63:61-78.
Comprehensive review of the editing mechanisms of aminoacyl-tRNA
synthetases. The authors also discuss the biological implications of
translational error.
2.
Ogle JM, Ramakrishnan V: Structural insights into translational
fidelity. Annu Rev Biochem 2005, 74:129-177.
3.
Kramer EB, Farabaugh PJ: The frequency of translational
misreading errors in E. coli is largely determined by tRNA
competition. RNA 2007, 13:87-96.
The authors developed a highly sensitive luminescent methodology to
study translational error in vivo in E. coli. They demonstrate that average
decoding error in vivo is 3.4 104 per codon and that it varies widely
between codons. A main source of error is the competition between
tRNAs at the ribosome A-site.
4.
Stahl G, Salem SN, Chen L, Zhao B, Farabaugh PJ: Translational
accuracy during exponential, postdiauxic, and stationary
growth phases in Saccharomyces cerevisiae. Eukaryot Cell
2004, 3:331-338.
www.sciencedirect.com
10. Eaglestone SS, Cox BS, Tuite MF: Translation termination
efficiency can be regulated in Saccharomyces cerevisiae by
environmental stress through a prion-mediated mechanism.
EMBO J 1999, 18:1974-1981.
11. Allmang C, Wurth L, Krol A: The selenium to selenoprotein
pathway in eukaryotes: more molecular partners than
anticipated. Biochim Biophys Acta 2009 doi: 10.1016/
j.bbagen.2009.03.003.
Review of the eukaryotic selenosome machinery. The authors review the
selenocysteine (Sec) biosynthesis pathway, in particular the enzymes that
convert serine into phosphoserine and selenocysteine, and also the
translation factors that are involved in Sec insertion at UGA codons.
12. Ambrogelly A, Palioura S, Soll D: Natural expansion of the
genetic code. Nat Chem Biol 2007, 3:29-35.
Recent review of the genetic code alterations discovered in bacteria,
fungi, and in the mitochondria of many eukaryotes. The expansions of the
genetic code to incorporate selenocysteine and pyrrolysine are discussed as well as the molecular mechanisms of codon reassignment.
13. Sheppard K, Yuan J, Hohn MJ, Jester B, Devine KM, Soll D: From
one amino acid to another: tRNA-dependent amino acid
biosynthesis. Nucleic Acids Res 2008, 36:1813-1825.
The pathways of tRNA-dependent asparagine, glutamine, and cysteine
biosynthesis in bacteria and mitochondria are described in detail in this
review.
Current Opinion in Microbiology 2009, 12:1–7
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004
COMICR-698; NO OF PAGES 7
6 Growth and development: eukaryotes
14. Gursinsky T, Grobe D, Schierhorn A, Jager J, Andreesen JR,
Sohling B: Factors and selenocysteine insertion sequence
requirements for the synthesis of selenoproteins from a Grampositive anaerobe in Escherichia coli. Appl Environ Microbiol
2008, 74:1385-1393.
15. Turanov AA, Lobanov AV, Fomenko DE, Morrison HG, Sogin ML,
Klobutcher LA, Hatfield DL, Gladyshev VN: Genetic code
supports targeted insertion of two amino acids by one codon.
Science 2009, 323:259-261.
This study demonstrates that cysteine (Cys) and selenocysteine (Sec) are
inserted at UGA codons in the ciliate Euplotes crassus and that such
codon ambiguity is necessary to produce fully functional proteins. There
are specialized tRNAs for Cys and Sec insertion at UGAs. These tRNAs
discriminate UGAs with the help of selenosome components involved in
recoding of UGA for Sec insertion.
16. Zhang Y, Gladyshev VN: High content of proteins containing
21st and 22nd amino acids, selenocysteine and pyrrolysine, in
a symbiotic deltaproteobacterium of gutless worm Olavius
algarvensis. Nucleic Acids Res 2007, 35:4952-4963.
17. Polycarpo CR, Herring S, Berube A, Wood JL, Soll D,
Ambrogelly A: Pyrrolysine analogues as substrates
for pyrrolysyl-tRNA synthetase. FEBS Lett 2006,
580:6695-6700.
18. Kobayashi T, Yanagisawa T, Sakamoto K, Yokoyama S:
Recognition of non-alpha-amino substrates by pyrrolysyltRNA synthetase. J Mol Biol 2009, 385:1352-1360.
The binding activity of the Methanosarcina mazei PylRS for non-a-amino
derivatives of Boc-lysine is studied. The authors demonstrate that the
PylRs has broad amino acid specificity and is able to bind and activate
various non-a-amino acids. They also show that these amino acids can
be successfully incorporated into an E. coli reporter protein.
19. Gundllapalli S, Ambrogelly A, Umehara T, Li D, Polycarpo C, Soll D:
Misacylation of pyrrolysine tRNA in vitro and in vivo. FEBS Lett
2008, 582:3353-3358.
Methanosarcina barkeri Fusaro tRNAPyl can be misacylated with serine in
vitro and in vivo in E. coli and serine can be misincorporated into E. coli
DHFR reporter protein.
20. Yanagisawa T, Ishii R, Fukunaga R, Kobayashi T, Sakamoto K,
Yokoyama S: Multistep engineering of pyrrolysyl-tRNA
synthetase to genetically encode N(epsilon)-(oazidobenzyloxycarbonyl) lysine for site-specific protein
modification. Chem Biol 2008, 15:1187-1197.
21. Curnow AW, Ibba M, Soll D: tRNA-dependent asparagine
formation. Nature 1996, 382:589-590.
22. Jahn D, Chen MW, Soll D: Purification and functional
characterization of glutamate-1-semialdehyde
aminotransferase from Chlamydomonas reinhardtii. J Biol
Chem 1991, 266:161-167.
23. Wilcox M: Gamma-glutamyl phosphate attached to glutaminespecific tRNA. A precursor of glutaminyl-tRNA in Bacillus
subtilis. Eur J Biochem 1969, 11:405-412.
High-resolution structure of the complex SepRS-tRNACys–O-phosphoserine. The authors describe the mechanism of O-phosphoserine binding
and activation by the SepRS and demonstrate that the C-terminal domain
of the enzyme recognizes the anticodon of the tRNACys.
29. Ruan B, Palioura S, Sabina J, Marvin-Guy L, Kochhar S,
LaRossa RA, Soll D: Quality control despite mistranslation
caused by an ambiguous genetic code. Proc Natl Acad Sci U S A
2008, 105:16502-16507.
Amino acid misincorporation into various E. coli reporter proteins was
tested using misacylating aminoacyl-tRNA synthetases, namely ProRS,
NN-ThrRS, ND-GluRS, and ND-AspRS. The study demonstrates that
mischarged tRNAs circumvent EF-Tu discrimination and can participate
in protein synthesis. Remarkably, E. coli cells were able to tolerate up to
12% of Asp misincorporation at Asn codons by a mischarged AsptRNAAsn. It is also demonstrated that mistranslating E. coli requires the
heat-shock response and proteases for survival.
30. Santos MAS, Tuite MF: The CUG codon is decoded in vivo as
serine and not leucine in Candida albicans. Nucleic Acids Res
1995, 23:1481-1486.
31. Santos MAS, Perreau VM, Tuite MF: Transfer RNA structural
change is a key element in the reassignment of the CUG codon
in Candida albicans. EMBO J 1996, 15:5060-5068.
32. Butler G, Rasmussen MD, Lin MF, Santos MA, Sakthikumar S,
Munro CA, Rheinbay E, Grabherr M, Forche A, Reedy JL et al.:
Evolution of pathogenicity and sexual reproduction in eight
Candida genomes. Nature 2009, 459:657-662.
33. Suzuki T, Ueda T, Watanabe K: The ‘polysemous’ codon — a
codon with multiple amino acid assignment caused by dual
specificity of tRNA identity. EMBO J 1997, 16:1122-1134.
34. Gomes AC, Miranda I, Silva RM, Moura GR, Thomas B,
Akoulitchev A, Santos MA: A genetic code alteration generates
a proteome of high diversity in the human pathogen Candida
albicans. Genome Biol 2007, 8:R206.
Misincorporation of serine at leucine CUG codons, mediated by a mischarged Leu-tRNACAGSer is quantified by direct mass spectrometry. The
study demonstrates that CUG codons are naturally ambiguous in C.
albicans and that environmental stress increases such ambiguity. C.
albicans tolerates 28% of CUG ambiguity with minimal decrease in
growth rate and such ambiguity generates phenotypic diversity. The
study also shows that CUG ambiguity expands the size of the C. albicans
proteome exponentially.
35. Miranda I, Rocha R, Santos MC, Mateus DD, Moura GR, Carreto L,
Santos MAS: A genetic code alteration is a phenotype diversity
generator in the human pathogen Candida albicans. PLoS ONE
2007, 2:e996.
Extensive description of phenotypic variability induced by CUG ambiguity
in the human pathogen C. albicans. CUG ambiguity generates extensive
morphological variation, increases flocculation, hypha formation, and
augments secretion of lipases and proteinases. The study also demonstrates that such ambiguity increases expression of C. albicans morphogenesis specific genes.
24. Wilcox M: Gamma-phosphoryl ester of glu-tRNA-GLN as an
intermediate in Bacillus subtilis glutaminyl-tRNA synthesis.
Cold Spring Harb Symp Quant Biol 1969, 34:521-528.
36. Santos MAS, Cheesman C, Costa V, Moradas-Ferreira P,
Tuite MF: Selective advantages created by codon ambiguity
allowed for the evolution of an alternative genetic code in
Candida spp.. Mol Microbiol 1999, 31:937-947.
25. Borup B, Ferry JG: Cysteine biosynthesis in the Archaea:
Methanosarcina thermophila utilizes O-acetylserine
sulfhydrylase. FEMS Microbiol Lett 2000, 189:205-210.
37. True HL, Lindquist SL: A yeast prion provides a mechanism for
genetic variation and phenotypic diversity. Nature 2000,
407:477-483.
26. Borup B, Ferry JG: O-Acetylserine sulfhydrylase from
Methanosarcina thermophila. J Bacteriol 2000, 182:45-50.
38. Namy O, Galopier A, Martini C, Matsufuji S, Fabret C, Rousset JP:
Epigenetic control of polyamines by the prion [PSI+]. Nat Cell
Biol 2008, 10:1069-1075.
The yeast [PSI+] prion regulates the expression of ornithine decarboxilase
antizyme, which is a negative regulator of cellular polyamines, by increasing the frequency of frameshifting at a UGA-shifty site. Polyamines are
involved in many cellular processes and variation in their cellular concentration induces phenotypic variation.
27. Fukunaga R, Yokoyama S: Structural insights into the
second step of RNA-dependent cysteine biosynthesis in
archaea: crystal structure of Sep-tRNA:Cys-tRNA
synthase from Archaeoglobus fulgidus. J Mol Biol 2007,
370:128-141.
Crystal structure of the SepCysS from Archaeoglobus fulgidus at 2.4 Å
resolution. The SepCysS is closely related to the pyridoxal 50 -phosphate
(PLP)-dependent cysteine desulfurases. Binding of the substrate PLP to
the active site and a model of the Sep-tRNACys–SepCysS complex are
also described.
28. Fukunaga R, Yokoyama S: Structural insights into the first step
of RNA-dependent cysteine biosynthesis in archaea. Nat Struct
Mol Biol 2007, 14:272-279.
Current Opinion in Microbiology 2009, 12:1–7
39. Al Mamun AA, Rahman MS, Humayun MZ: Escherichia coli cells
bearing mutA, a mutant glyV tRNA gene, express a recAdependent error-prone DNA replication activity. Mol Microbiol
1999, 33:732-740.
40. Al Mamun AA, Marians KJ, Humayun MZ: DNA polymerase III
from Escherichia coli cells expressing mutA mistranslator
tRNA is error-prone. J Biol Chem 2002, 277:46319-46327.
www.sciencedirect.com
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004
COMICR-698; NO OF PAGES 7
Mistranslation generates phenotypic diversity Moura, Carreto and Santos 7
41. Balashov S, Humayun MZ: Mistranslation induced by
streptomycin provokes a RecABC/RuvABC-dependent
mutator phenotype in Escherichia coli cells. J Mol Biol 2002,
315:513-527.
42. Balashov S, Humayun MZ: Escherichia coli cells bearing a
ribosomal ambiguity mutation in rpsD have a mutator
phenotype that correlates with increased mistranslation.
J Bacteriol 2003, 185:5015-5018.
43. Dorazi R, Lingutla JJ, Humayun MZ: Expression of mutant
alanine tRNAs increases spontaneous mutagenesis in
Escherichia coli. Mol Microbiol 2002, 44:131-141.
44. Nangle LA, Crecy-Lagard V, Doring V, Schimmel VP: Genetic
code ambiguity. Cell viability related to severity of editing
defects in mutant tRNA synthetases. J Biol Chem 2002,
277:45729-45733.
45. Nangle LA, Motta CM, Schimmel P: Global effects of
mistranslation from an editing defect in mammalian cells.
Chem Biol 2006, 13:1091-1100.
46. Silva RM, Paredes JA, Moura GR, Manadas B, Lima-Costa T,
Rocha R, Miranda I, Gomes AC, Koerkamp MJ, Perrot M et al.:
Critical roles for a genetic code alteration in the evolution of
the genus Candida. EMBO J 2007, 26:4555-4565.
Ambiguous decoding of leucine CUG codons as serine in S. cerevisiae
blocks sexual reproduction and induces important genomic alterations,
www.sciencedirect.com
namely increased ploidy and aneuploidy. It also reprograms gene expression, activates the stress response, and generates selective advantages
under environmental stress conditions.
47. Rutherford SL: Between genotype and phenotype: protein
chaperones and evolvability. Nat Rev Genet 2003, 4:263-274.
48. Tyedmers J, Madariaga ML, Lindquist S: Prion switching in
response to environmental stress. PLoS Biol 2008, 6:e294.
In yeast, the prion [PSI+] exposes an array of hidden phenotypes by
increasing mistranslation of the three stop codons (UAG, UGA, and
UAA). However, the frequency of appearance of the [PSI+] prion is very
low and many wild type strains do not have it, questioning the evolutionary
relevance of this epigenetic evolvability system. The authors show that the
frequency of [PSI+] induction increases as much as 60-fold when yeast is
exposed to environmental stressors, namely hydrogen peroxide and salt.
49. Griswold CK, Masel J: Complex adaptations can drive the
evolution of the capacitor [PSI], even with realistic rates of
yeast sex. PLoS Genet 2009, 5:e1000517.
50. Wang Q, Parrish AR, Wang L: Expanding the genetic code for
biological studies. Chem. Biol. 2009, 16:323-336.
Review of the non-natural amino acids that have already been inserted
into the genetic code of E. coli, yeast and mammalian cells, using genetic
engineering methodologies. The authors highlight how incorporation of
artificial amino acids into the genetic code can be used to produce
proteins with novel properties.
Current Opinion in Microbiology 2009, 12:1–7
Please cite this article in press as: Moura GR, et al. Genetic code ambiguity: an unexpected source of proteome innovation and phenotypic diversity, Curr Opin Microbiol (2009), doi:10.1016/
j.mib.2009.09.004