ArticlePDF Available

The gene of an archaeal α-L-fucosidase is expressed by translational frameshifting

Authors:

Abstract and Figures

The standard rules of genetic translational decoding are altered in specific genes by different events that are globally termed recoding. In Archaea recoding has been unequivocally determined so far only for termination codon readthrough events. We study here the mechanism of expression of a gene encoding for a α-l-fucosidase from the archaeon Sulfolobus solfataricus (fucA1), which is split in two open reading frames separated by a −1 frameshifting. The expression in Escherichia coli of the wild-type split gene led to the production by frameshifting of full-length polypeptides with an efficiency of 5%. Mutations in the regulatory site where the shift takes place demonstrate that the expression in vivo occurs in a programmed way. Further, we identify a full-length product of fucA1 in S.solfataricus extracts, which translate this gene in vitro by following programmed −1 frameshifting. This is the first experimental demonstration that this kind of recoding is present in Archaea.
Content may be subject to copyright.
The gene of an archaeal a-L-fucosidase is expressed
by translational frameshifting
Beatrice Cobucci-Ponzano
1
, Fiorella Conte
1
, Dario Benelli
2
, Paola Londei
2
,
Angela Flagiello
3
, Maria Monti
3
, Piero Pucci
3
, Mose
`Rossi
1,4
and Marco Moracci
1,
*
1
Institute of Protein Biochemistry—Consiglio Nazionale delle Ricerche, Via P. Castellino 111, 80131 Naples,
Italy,
2
Dipartimento di Biochimica Medica e Biologia Medica (DIBIM), Universita
`di Bari—Piazzale Giulio Cesare,
70124 Bari, Italy,
3
CEINGE Biotecnologie Avanzate s.c.a r.l., Dipartimento di Chimica Organica e Biochimica,
Universita
`di Napoli Federico II, Complesso Universitario di Monte S. Angelo, via Cinthia 4, 80126 Napoli, Italy and
4
Dipartimento di Biologia Strutturale e Funzionale, Universita
`di Napoli ‘Federico II’, Complesso Universitario di Monte
S. Angelo, Via Cinthia 4, 80126 Naples, Italy
Received April 19, 2006; Revised July 21, 2006; Accepted July 22, 2006
ABSTRACT
The standard rules of genetic translational decoding
are altered in specific genes by different events
that are globally termed recoding. In Archaea reco-
ding has been unequivocally determined so far
only for termination codon readthrough events. We
study here the mechanism of expression of a gene
encoding for a a-L-fucosidase from the archaeon
Sulfolobus solfataricus (fucA1), which is split in two
open reading frames separated by a 1 frame-
shifting. The expression in Escherichia coli of
the wild-type split gene led to the production by
frameshifting of full-length polypeptides with an
efficiency of 5%. Mutations in the regulatory site
where the shift takes place demonstrate that the
expression in vivo occurs in a programmed way.
Further, we identify a full-length product of fucA1
in S.solfataricus extracts, which translate this gene
in vitro by following programmed 1 frameshifting.
This is the first experimental demonstration that
this kind of recoding is present in Archaea.
INTRODUCTION
Translation is optimally accurate and the correspondence
between the nucleotide and the protein sequences are often
considered as an immutable dogma. However, the genetic
code is not quite universal: in certain organelles and in a
small number of organisms the meaning of different codons
has been reassigned and all the mRNAs are decoded
accordingly. More surprisingly, the standard rules of genetic
decoding are altered in specific genes by different events that
are globally termed recoding (1). In all cases, translational
recoding occurs in competition with normal decoding, with
a proportion of the ribosomes not obeying to the ‘universal’
rules. Translational recoding has been identified in both
prokaryotes and eukaryotes. It has crucial roles in the regula-
tion of gene expression and includes stop codon readthrough,
ribosome hopping and ±1 programmed frameshifting [for
reviews see (2–4)].
In stop codon readthrough a stop codon is decoded by a
tRNA carrying an unusual amino acid rather than a transla-
tional release factor. Specific stimulatory elements down-
stream to the stop codon regulate this process (5). Hopping,
in which the ribosome stops translation in a particular site
of the mRNA and re-start few nucleotides downstream, is a
rare event and it has been studied in detail only in the bacte-
riophage T4 (6). In programmed frameshifting, ribosomes are
induced to shift to an alternative, overlapping reading frame
1nt3
0-wards (+1 frameshifting) or 50-wards (1 frameshifting)
of the mRNA. This process is regulated and its frequency
varies in different genes. The ±1 programmed frameshifting
has been studied extensively in viruses, retrotransposons and
insertion elements for which many cases are documented
(7–9). Instead, this phenomenon is by far less common in
cellular genes. A single case of programmed +1 frameshift-
ing is known in prokaryotes (10,11) while in eukaryotes,
including humans, several genes regulated by this recoding
event have been described previously [(4) and references
therein]. Compared to +1 frameshifting, 1 frameshifting
is less widespread with only two examples in prokaryotes
(12–14) and few others in eukaryotes (15–17).
The programmed 1 frameshifting is triggered by several
elements in the mRNA. The slippery sequence, showing the
X-XXY-YYZ motif, in which X can be any base, Y is usually
A or U, and Z is any base but G, has the function of favouring
the tRNA misalignment and it is the site where the shift takes
place (3,18). Frameshifting could be further stimulated by
other elements flanking the slippery sequence: a codon for
a low-abundance tRNA, a stop codon, a Shine–Dalgarno
sequence and an mRNA secondary structure. It has been
*To whom correspondence should be addressed. Tel: +39 081 6132271; Fax: +39 081 6132277; Email: m.moracci@ibp.cnr.it
2006 The Author(s).
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/
by-nc/2.0/uk/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.
4258–4268 Nucleic Acids Research, 2006, Vol. 34, No. 15 Published online 18 August 2006
doi:10.1093/nar/gkl574
reported that these elements, alone or in combination,
enhance frameshifting by pausing the translating ribosome
on the slippery sequence (4,18).
Noticeably, known cases of recoding in Archaea [recently
reviewed in (19)] are limited to termination codon
readthrough events that regulate the incorporation of the
21st and 22nd amino acids selenocysteine and pyrrolysine,
respectively (20–23).
No archaeal genes regulated by translational programmed
frameshifting and ribosome hopping have been identified
experimentally so far; therefore, if compared with the others
domains of life, the study of translational recoding in Archaea
is still at its dawn.
We showed that the a-L-fucosidase gene from the crenar-
chaeon Sulfolobus solfataricus is putatively expressed by pro-
grammed 1 frameshifting (24). This gene, named fucA1,is
organized in the open reading frames (ORFs) SSO11867 and
SSO3060 of 81 and 426 amino acids, respectively, which are
separated by a 1 frameshifting in a 40 base overlap
(Figure 1A). We have reported previously that the region of
overlap between the two ORFs had the characteristic features
of the genes expressed by programmed 1 frameshifting
including a slippery heptanucleotide A-AAA-AAT (codons
are shown in the zero frame) flanked by a putative stem–
loop and the rare codons CAC (Figure 1A) resembling the
prokaryotic stem–loops/hairpins and the Shine–Dalgarno-
like sites (24). We showed that the frameshifting, obtained
by mutating by site-directed mutagenesis the fucA1 gene
exactly in the position predicted from the slippery site,
produced a full-length gene, named fucA1
A
, encoding for a
polypeptide of 495 amino acids (Figure 1B). This mutant
gene expressed in Escherichia coli a fully functional
a-L-fucosidase, named Ssa-fuc, which was thermophilic,
thermostable and had an unusual nonameric structure
(24,25). More recently, we determined the reaction mecha-
nism and the function of the residues of the active site of
the mutant enzyme (26,27).
The functionality of the product of the mutant gene
fucA1
A
does not provide direct experimental evidence that
programmed 1 frameshifting occurs in vivo and in
S.solfataricus. To address these issues, we report here the
study of the expression of the wild-type split gene fucA1
and of its mutants in the slippery sequence. We demonstrate
here that fucA1 is expressed by programmed 1 frameshift-
ing in both E.coli and S.solfataricus. This is the first experi-
mental demonstration that this kind of recoding is present in
the Archaea domain of life. The relevance of programmed 1
frameshifting in Archaea is also discussed.
Figure 1. The a-fucosidase gene. (A) Region of overlap in the wild-type split fucA1 gene. The N-terminal SSO11867 ORF is in the zero frame, the C-terminal
SSO3060 ORF, for which only a fragment is shown, is in the 1 frame. The slippery heptameric sequence is underlined; the rare codons are boxed and the arrows
indicate the stems of the putative mRNA secondary structure. The amino acids involved in the programmed 1 frameshifting and the first codon translated after
this event in the 1 frame are shadowed. (B) Fragment of the full-length mutant fucA1
A
gene. The small arrows indicate the mutated nucleotides.
Nucleic Acids Research, 2006, Vol. 34, No. 15 4259
MATERIALS AND METHODS
Analysis of the a-fucosidase expression
S.solfataricus cells were grown, and cell extracts obtained, as
described previously (24,28).
The expression in the E.coli strain BL21(RB791) of
the wild-type gene fucA1 and of the mutant genes fucA1
A
[previously named FrameFuc in (24)], fucA1
B
,fucA1
sm
and
fucA1
tm
as fusions of glutathione S-transferase (GST) and
the purification of the recombinant proteins were performed
as reported previously (23). The nomenclature used in this
paper for the different a-fucosidase genes is listed in Table 1.
For the western blot studies, equal amounts of E.coli
cultures expressing the wild-type and mutant fucA1 genes,
normalized for the OD
600
, were resuspended in SDS–PAGE
loading buffer containing 0.03 M Tris–HCl buffer, pH 6.8,
3% SDS (w/v), 6.7% glycerol (w/v), 6.7% 2-mercaptoethanol
(w/v) and 0.002% blue bromophenol (w/v). The samples were
incubated at 100C for 5 min (unless otherwise indicated)
and were directly loaded on to the gel. Western blot analyses
were performed by blotting SDS–PAGEs of the concen-
trations indicated on Hybond-P polyvinylidenfluorid filters
(Amersham Biosciences, Uppsala, Sweden); polyclonal
anti-Ssa-fuc antibodies from rabbit (PRIMM, Milan, Italy)
and anti-GST antibodies (Amersham Biosciences) were
diluted 1:5000 and 1:40 000, respectively. The filters were
washed and incubated with the ImmunoPure anti-rabbit IgG
antibody conjugated with the horseradish peroxidase (HRP)
from Pierce Biotechnology (Rockford, IL, USA). Filters
were developed with the ECL-plus Western Blotting Detec-
tion system (Amersham Biosciences) by following the manu-
facturer’s indications. The molecular weight markers used in
the western blot analyses were the ECL streptavidin–HRP
conjugate (Amersham Biosciences).
The protein concentration of the samples was measured
with the method of Bradford (29) and the amounts of
sample loaded on to the SDS–PAGEs are those indicated.
The quantification of the bands identified by western blot
was performed by using the program Quantity One 4.4.0 in a
ChemiDoc EQ System (Bio-Rad, Hercules, CA, USA) with
the volume analysis tool. The frameshifting efficiency was cal-
culated as the ratio of the intensity of the bands of the frame-
shifted product/frameshifted product +termination product.
The mutants in the slippery sequence of the wild-type gene
fucA1 were prepared by site-directed mutagenesis from the
vector pGEX-11867/3060, described previously (24,27).
The synthetic oligonucleotides used (PRIMM) were the
following: FucA1sm-rev, 50-TTTAGGTGATATTGGTGTT-
CTGGTCTATCT-30; FucA1sm-fwd, 50-GAACACCAATAT-
CACCTAAAGAATTCGGCCCA-30; FucA1tm-rev, 50-AGG-
TGATATTGGTGTTCTGGTCTATCTGGC-30; FucA1tm-fwd,
50-CCAGAACACCAATATCACCTCAAGAACTCGGCCCA-
GT-30, where the mismatched nucleotides in the mutagenic
primers are underlined. Direct sequencing identified the plas-
mids containing the desired mutations and the mutant genes,
named fucA1
sm
and fucA1
tm
, were completely re-sequenced.
Expression and characterization of Ssa-fuc
B
The mutant Ssa-fuc
B
was prepared by site-directed muta-
genesis from the vector pGEX-11867/3060, by using the
same site-directed mutagenesis kit described above. The syn-
thetic oligonucleotides used were FucA1sm-rev (described
above) and the following mutagenic oligonucleotide: Fuc-B,
50-GAACACCAATATCACCTAAAGAAGTTCGGCCC-
AGT-30, where the mismatched nucleotides are underlined.
Direct sequencing identified the plasmid containing the desired
mutations and the mutant gene, named fucA1
B
, was completely
re-sequenced. The enzymatic characterization of Ssa-fuc
B
was
performed as described previously (24,27).
Mass spectrometry experiments
Samples of the proteins expressed in E.coli from the wild-
type gene fucA1 and the mutants fucA1
A
and fucA1
sm
, purified
as described, were fractionated on an SDS–PAGE. Protein
bands were excised from the gel, washed in 50 mM ammo-
nium bicarbonate, pH 8.0, in 50% acetonitrile, reduced with
10 mM DTT at 56C for 45 min and alkylated with 55 mM
iodoacetamide for 30 min at room temperature in the dark.
The gel pieces were washed several times with the buffer,
resuspended in 50 mM ammonium bicarbonate and incubated
with 100 ng of trypsin for 2 h at 4C and overnight at
37C. The supernatant containing peptides was analysed by
MALDIMS on an Applied Biosystem Voyager DE-PRO
mass spectrometer using a-cyano-4-hydroxycynnamic acid
as matrix. Mass calibration was performed by using the
standard mixture provided by manufacturer.
Liquid chromatography online tandem mass spectrometry
(LCMSMS) analyses were performed on a Q-TOF hybrid
mass spectrometer (Micromass, Waters, Milford, MA,
USA) coupled with a CapLC capillary chromatographic sys-
tem (Waters). Peptide ions were selected in the collision cell
and fragmented. Analysis of the daughter ion spectra led to
the reconstruction of peptide sequences.
Experiments of translation in vitro
Genomic DNA from S.solfataricus P2 strain was prepared as
described previously (24). A DNA fragment of 1538 nt con-
taining the complete fucA1 gene, was prepared by PCR, by
using the following synthetic oligonucleotides (Genenco,
Florence, Italy): FucA1-fwd, 50-CTGGAGGCGCGCTAA-
TACGACTCACTATAGGTCAGTTAAATGTCACAAAA-
TTCT-30; FucA1-rev, 50-GACTTGGCGCGCCTATCTAT-
AATCTAGGATAACCCTTAT-30, in which the sequence
corresponding to the genome of S.solfataricus is underlined.
In the FucA1-fwd primer, the sequence of the promoter of
the T7 RNA polymerase is in boldface and the sequence of
the BssHII site is shown in italics. The PCR amplification
was performed as described previously (24) and the ampli-
fication products were cloned in the BssHII site of the plas-
mid pBluescript II KS+.ThefucA1 gene was completely
re-sequenced to check if undesired mutations were introduced
by PCR and the recombinant vector obtained, named pBlu-
FucA1, was used for translation in vitro experiments.
The plasmids expressing the mutant genes fucA1
A
,fucA1
sm
and fucA1
tm
for experiments of translation in vitro were pre-
pared by substituting the KpnI–NcoI wild-type fragment,
containing the slippery site, with those isolated from the
mutants. To check that the resulting plasmids had the correct
sequence, the mutant genes were completely re-sequenced.
4260 Nucleic Acids Research, 2006, Vol. 34, No. 15
The mRNAs encoding wild-type fucA1 and its various
mutants were obtained by in vitro run-off transcription.
About 2 mg of each plasmid was linearized with BssHII
and incubated with 50 U of T7 RNA polymerase for 1 h
30 min at 37C. The transcription mixtures were then treated
with 10 U of DNAseI (RNAse free) for 30 min. The tran-
scribed RNAs were recovered by extracting the samples
twice with phenol (pH 4.7) and once with phenol/chloroform
1:1 followed by precipitation with ethanol. The mRNAs
were resuspended in DEPC-treated H
2
O at the approximate
concentration of 0.6 pmol/ml.
In vitro translation assays were performed essentially as
described by Condo
`et al. (28). The samples (25 ml final
volume) contained 5 mlofS.solfataricus cell extract,
10 mM KCl, 20 mM Tris–HCl, pH 7.0, 20 mM Mg acetate,
3 mM ATP, 1 mM GTP, 5 mg of bulk S.solfataricus tRNA,
2mlof[
35
S]methionine (1200 Ci/mmol at 10 mCi/ml) and
10 pmol of each mRNA. The mixtures were incubated at
70C for 45 min. After this time, the synthesized proteins
were resolved by electrophoresis 12.5% acrylamide–SDS
gels and revealed by autoradiography of the dried gels on
an Instant Imager apparatus.
Transcriptional analysis of fucA1
Cells of S.solfataricus, strain P2, were grown in minimal
salts culture media supplemented with yeast extract (0.1%),
casamino acids (0.1%), plus glucose (0.1%) (YGM) or
sucrose (0.1%) (YSM). The extraction of total RNA was
performed as reported previously (24). Total RNA was
extensively digested with DNAse (Ambion, Austin, TX,
USA) and the absence of DNA was assessed by the lack
of PCR amplification with each sets of primers described
below. The RT–PCR experiments were performed as
reported previously (24) by using the primers described pre-
viously that allowed the amplification of a region of 833 nt
(positions 1–833, in which the A of the first ATG codon is
numbered as one) overlapping the ORFs SSO11867 and
SSO3060 (24).
For real-time PCR experiments total cDNA was obtained
using the kit Quantitect RT (Qiagen GmbH, Hilden,
Germany) from 500 ng of the same preparation of RNA
described above. cDNA was then amplified in a Bio-Rad
LightCycler using the DyNAmo HS Syber Green qPCR Kit
(Finnzymes Oy, Espoo, Finland). Synthetic oligonucleotides
(PRIMM) used for the amplification of a region at the 30of
the ORF SSO3060 were as follows: 50-Real: 50-TAAATGGC-
GAAGCGATTTTC-30;3
0-Real: 50-ATATGCCTTTGTCGC-
GGATA-30for the gene fucA1.5
0-GAATGGGGGTGATA-
CTGTCG-30and 50-TTTACAGCCGGGACTACAGG-30for
the 16S rRNA gene.
For each amplification of the fucA1 gene was used 2500-
fold more cDNA than that used for the amplification of the
16S rRNA. Controls with no template cDNA were always
included. PCR conditions were 15 min at 95C for initial
denaturation, followed by 40 cycles of 10 s at 95C, 25 s at
56C and 35 s at 72C, and a final step of 10 min at 72C.
Product purity was controlled by melting point analysis of
setpoints with 0.5C temperature increase from 72 to 95C.
PCR products were analysed on 2% agarose gels and visual-
ized by ethidium bromide staining.
The expression values of fucA1 gene were normalized to
the values determined for the 16S rRNA gene. Absolute
expression levels were calculated as fucA1/16S ratio in
YSM and YGM cells, respectively. Relative mRNA expres-
sion levels (YSM/YGM ratio) were calculated as ( fucA1/
16S ratio in YGM cells)/(fucA1/16S ratio in YSM). Each
cDNA was used in triplicate for each amplification.
RESULTS
Expression of fucA1 in E.coli
The wild-type fucA1 gene, expressed in E.coli as a GST-fused
protein, produced trace amounts of a-fucosidase activity
(2.3 ·10
2
units mg
1
after removal of GST), suggesting
that a programmed 1 frameshifting may occur in E.coli
(24). The enzyme was then purified by using the GST puri-
fication system and analysed by SDS–PAGE revealing a
major protein band (Figure 2A). The sample and control
bands were excised from the gel, digested in situ with trypsin
and directly analysed by matrix-assisted laser desorption/
ionization mass spectrometry (MALDIMS). As shown in
Figure 2B and C, both spectra revealed the occurrence of
an identical mass signal at m/z 1244.6 corresponding to a
peptide (Peptide A) encompassing the overlapping region of
the two ORFs. This result was confirmed by liquid chro-
matography online tandem mass spectrometry (LCMSMS)
analysis of the peptide mixtures. The fragmentation spectra
of the two signals showed the common sequence Asn-Phe-
Gly-Pro-Val-Thr-Asp-Phe-Gly-Tyr-Lys in which the amino
acid from the ORF SSO11867 is underlined. These results
unequivocally demonstrate that the protein containing the
Peptide A is produced in E.coli by a frameshifting event
that occurred exactly within the slippery heptamer predicted
from the analysis of the DNA sequence in the region of over-
lap between the ORFs SSO11867 and SSO3060 (Figure 1A).
Remarkably, the MALDIMS analysis of the products of
the wild-type fucA1 gene revealed the presence of a second
Peptide B at m/z 1258.6 that is absent in the spectra of the
Ssa-fuc control protein (Figure 2B and C). The sequence of
Peptide B obtained by LCMSMS (Figure 2D) was Lys-Phe-
Gly-Pro-Val-Thr-Asp-Phe-Gly-Tyr-Lys. This sequence dif-
fers only by one amino acid from Peptide A demonstrating
that the interrupted gene fucA1 expresses in E.coli two full-
length proteins originated by different 1 frameshifting
events. Polypeptide A results from a shift in a site A and it
is identical to Ssa-fuc prepared by site-directed mutagenesis
(24), suggesting that the expression occurred with the simul-
taneous P- and A-site slippage. Instead, polypeptide B, named
Ssa-fuc
B
, is generated by frameshifting in a second site B as
the result of a single P-site slippage (Figure 2E).
To measure the global efficiency of frameshifting in the
two sites of the wild-type gene fucA1 we analysed the total
extracts of E.coli by western blot using anti-GST antibodies
(Figure 2F). Two bands with marked different electrophoretic
mobility were observed: the polypeptide of 78.7 ± 1.1 kDa
migrated like GST-Ssa-fuc fusion and was identified as origi-
nated from frameshifting in either site A or B of fucA1. The
protein of 38.1 ± 1.2 kDa, which is not expressed by the mut-
ant gene fucA1
A
(not shown), had an electrophoretic mobility
compatible with GST fused to the polypeptide encoded by the
Nucleic Acids Research, 2006, Vol. 34, No. 15 4261
ORF SSO11867 solely (27 and 9.6 kDa, respectively). This
polypeptide originated from the translational termination of
the ribosome at the OCH codon of the fucA1 N-terminal
ORF (Figure 1A). The calculated ratio of frameshifting to
the termination products was 5%.
Preparation and characterization of Ssa-fuc
B
To test if the full-length a-fucosidase produced by the 1
frameshifting event in site B (Ssa-fuc
B
), resulting from the
single P-site slippage has different properties from Ssa-fuc,
whose sequence arises from the simultaneous P- and A- site
slippage, we prepared the enzyme by site-directed muta-
genesis. The slippery sequence in fucA1 A-AAA-AAT was
mutated in A-AAG-AAG-T where mutations are underlined.
The new mutant gene was named fucA1
B
. The first G, produc-
ing the conservative mutation AAA!AAG, was made to
disrupt the slippery sequence and hence reducing the shifting
efficiency. The second G was inserted to produce the
frameshifting that results in the amino acid sequence of
Peptide B. Therefore, the sequence of the two full-length
mutant genes fucA1
A
and fucA1
B
differs only in the region
of the slippery sequence: A-AAG-AAT-TTC-GGC and
A-AAG-AAG-TTC-GGC, respectively (the mutations are
underlined, the nucleotides in boldface were originally in
the 1 frame) (Table 1).
The recombinant Ssa-fuc
B
was purified up to 95%
(Materials and Methods). Gel filtration chromatography demon-
strated that in native conditions Ssa-fuc
B
had the same non-
americ structure of Ssa-fuc with an identical molecular weight
of 508 kDa (data not shown). In addition, Ssa-fuc
B
had the same
high substrate selectivity of Ssa-fuc. The two enzymes have
high affinity for 4-nitrophenyl-a-L-fucoside (4NP-Fuc) sub-
strate at 65C; the K
M
is identical within the experimental
error (0.0287 ± 0.005 mM) while the k
cat
of Ssa-fuc
B
(137 ±
5.7 s
1
)is48%ofthatofSsa-fuc (287 ± 11 s
1
). In addition,
4-nitrophenyl-a-L-arabinoside, -rhamnoside, 4-nitrophenyl-a-
D-glucoside, -xyloside, -galactoside and -mannoside were not
substrates of Ssa-fuc
B
as shown previously for Ssa-fuc (24).
This suggests that the different amino acid sequence did not
Figure 2. Analysis of the expression of fucA1 in E.coli.(A) Coomassie stained 7% SDS–PAGE showing (line 1) the recombinant Ssa-fuc (3 mg) and (line 2) the
purified products of the wild-type split fucA1 gene. The protein concentration of the latter sample could not be quantified because of the scarcity of
the purification yields. MALDIMS of the purified products of the wild-type fucA1 gene and of Ssa-fuc are shown in (B) and (C), respectively. Peptide A and B
are indicated. (D) LCMSMS analysis of peptide B. (E) The proposed frameshifting sites in the fucA1 gene. The open and the closed arrows indicate the shifting
sites A and B, respectively (for details see text). The sequence of the peptides A and B are also indicated. (F) Western blot of E.coli cellular extracts expressing
the wild-type fucA1 gene. The blot was probed with anti-GST antibodies. The pre-stained molecular weight markers were b-galactosidase (175 000), paramyosin
(83 000), glutamic dehydrogenase (62 000), aldolase (47 500) and triosephosphate isomerase (32 500).
Table 1. Nomenclature and characteristics of the a-fucosidase genes
Gene name Status Name of the
recombinant protein
Slippery heptamer
a
fucA1 wild type 1 frameshifted A-AAA-AAT
fucA1
A
mutant Full-length Ssa-fuc A-AAG-AAT
fucA1
B
mutant Full-length Ssa-fuc
B
A-AAG-AAG
fucA1
sm
mutant 1 frameshifted A-AAG-AAT
fucA1
tm
mutant 1 frameshifted C-AAG-AAC
a
Nucleotides modified by substitution and insertion mutations are underlined
and in boldface, respectively.
4262 Nucleic Acids Research, 2006, Vol. 34, No. 15
significantly affect the active site. Both enzymes showed an
identical profile of specific activity versus temperature with an
optimal temperature higher than 95C (data not shown). The
heat stability and the pH dependence of Ssa-fuc and Ssa-fuc
B
are reported in Figure 3. At 80C, the optimal growth
temperature of S.solfataricus, the half-life of Ssa-fuc
B
is
45 min, almost 4-fold lower than that of Ssa-fuc (Figure 3A).
The two enzymes showed different behaviour at pH <6.0 at
which Ssa-fuc
B
is only barely active and stable (Figure 3B);
however, the two enzymes showed similar values of specific
activity at pHs above 6.0, which is close to the intracellular pH
of S.solfataricus (30).
Characterization of the slippery sequence of
fucA1 in E.coli
The experimental data reported above indicate that the pre-
dicted slippery heptanucleotide in the region of overlap
between the ORFs SSO11867 and SSO3060 of the wild-
type gene fucA1 could regulate in cis the frameshifting events
observed in E.coli. To test this hypothesis, we mutated the
sequence A-AAA-AAT into A-AAG-AAT and C-AAG-
AAC (mutations are underlined) obtaining the fucA1 single
mutant ( fucA1
sm
) and triple mutant ( fucA1
tm
) genes, respec-
tively. It is worth noting that the mutations disrupt the
slippery sequence, but they maintain the 1 frameshift
between the two ORFs (Table 1).
Surprisingly, the expression of fucA1
sm
in E.coli produced
a full-length polypeptide that, after purification by affinity
chromatography and removal of the GST protein, showed
the same electrophoretic migration of Ssa-fuc and Ssa-fuc
B
(Figure 4A). This protein was then characterized by mass
spectrometry analyses following in situ tryptic digestion.
Interestingly, the MALDI spectra revealed the presence of a
single peptide encompassing the overlapping region between
the two ORFs with a mass value of 1259.7 Da (peptide C;
Figure 4B). The sequence of peptide C, determined from
the fragmentation spectra obtained by LCMSMS analysis, was
Glu-Phe-Gly-Pro-Val-Thr-Asp-Phe-Gly-Tyr-Lys (Figure 4C).
Remarkably, apart from the Glu residue, this sequence is
identical to that of peptide B produced from fucA1, indicat-
ing that in the mutant gene fucA1
sm
only one of the two
frameshifting events observed in the wild-type fucA1 gene
had occurred. The presence of a Glu instead of Lys was
not unexpected. The mutation A-AAA-AAT!A-AAG-
AAT in fucA1
sm
was conservative in the zero frame of the
ORF SSO11867 (AAA!AAG, both encoding Lys), but it
produced the mutation AAA!GAA (Lys!Glu) in the 1
frame of the ORF SSO3060.
It is worth noting that the frameshifting efficiency of the
gene fucA1
sm
, calculated by western blot as described
above, was 2-folds higher (10%) if compared to fucA1 (5%)
(Figure 4D). This indicates that the mutation cancelled the
frameshifting site A and, in the same time, enhanced the
frameshifting efficiency of site B.
In contrast, the triple mutant fucA1
tm
produced in E.coli
only the low molecular weight band resulting from transla-
tional termination (Figure 4D). No full-length protein could
be detected in western blots probed with either anti-GST
(Figure 4D) or anti-Ssa-fuc antibodies (Figure 4E). These
data show that the disruption of the heptameric slippery
sequence completely abolished the frameshifting in E.coli
confirming that this sequence has a direct role in controlling
the frameshifting in vivo.
Expression of fucA1 in S.solfataricus
To test whether fucA1 is expressed in S.solfataricus we anal-
ysed the extracts of cells grown on yeast extract, sucrose and
casaminoacids medium (YSM). Accurate assays showed that
S.solfataricus extracts contained 3.4 ·10
4
units mg
1
of
a-fucosidase activity. These very low amounts hampered
the purification of the enzyme. The extracts of S.solfataricus
cells grown on YSM revealed by western blot a band of a
molecular mass >97 kDa and no signals were detected with
the pre-immune serum confirming the specificity of the
anti-Ssa-fuc antibodies (Figure 5A). The different molecular
mass may result from post-translational modifications
occurred in the archaeon or from the incomplete denaturation
of a protein complex. In particular, the latter event is not
unusual among enzymes from hyperthermophilic archaea
(31,32). To test which hypotheses were appropriate, cellular
extracts of S.solfataricus were analysed by western blot
extending the incubation at 100C to 2 h. Interestingly, this
Figure 3. Comparison of the stability and pH dependence of Ssa-fuc and
Ssa-fuc
B
.(A) Thermal stability of Ssa-fuc (open circles) and Ssa-fuc
B
(closed circles) at 80C. (B) pH dependence of Ssa-fuc (open circles) and
Ssa-fuc
B
(closed circles) at 65C.
Nucleic Acids Research, 2006, Vol. 34, No. 15 4263
treatment shifted the high-molecular mass band to 67.6 ± 1.2
kDa (Figure 5B and C), which still differs from that of the
recombinant Ssa-fuc, 58.9 ± 1.2 kDa, leaving the question
on the origin of this difference unsolved. To try to shed
some light we immunoprecipitated extracts of S.solfataricus
with anti-Ssa-fuc antibodies and we analysed the major
protein band by MALDIMS. Unfortunately, we could not
observe any peptide compatible with the fucosidase because
the heavy IgG chain co-migrated with the band of the
expected molecular weight (data not shown).
To test if the scarce amounts of the a-fucosidase in
S.solfataricus extracts was the result of reduced expression
at transcriptional level, we performed a northern blot analysis
of total RNA extracted from cells grown either on YSM or
YGM media. We could not observe any signal by using
probes matching the 30of the ORF SSO3060 (data not
shown). These results suggest that fucA1 produced a rare tran-
script; therefore, we analysed the level of mRNA by RT–PCR
and by real-time PCR. A band corresponding to the region of
overlap between the ORFs SSO11867 and SSO3060 was
observed in the RNA extracted from cells grown on YSM
Figure 5. Analysis of the expression of the a-fucosidase in S.solfataricus.
(A) Western blot analysis of recombinant Ssa-fuc (lanes 1, 2, 5 and 6,
0.14 mg) and of extracts of S.solfataricus cells grown on YSM (lanes 3, 4, 7
and 8, 153 mg). Samples in lanes 1, 3, 5 and 7 were not denaturated
before loading. The left panel shows the blot probed with anti-Ssa-fuc
antibodies; the right panel was probed with the pre-immune serum diluted
1:5 000. (B) Western blot analysis: recombinant Ssa-fuc (lanes 1, 2 and 3,
0.5 mg) incubated at 100C for 5 min, 1 h and 2 h, respectively; extracts
of S.solfataricus cells (lanes 4, 5 and 6, 1 mg) incubated at 100C for 5 min,
1 h and 2 h, respectively. (C) Western blot analysis of recombinant
Ssa-fuc (lane 1, 0.1 mg) incubated at 100C for 5 min and of extracts of
S.solfataricus cells (lane 2, 1 mg) incubated at 100C for 2 h, respectively.
The molecular weight markers were: phosphorylase b (97 000), albumin
(66 000), ovalbumin (45 000), carbonic anhydrase (31 000) and trypsin
inhibitor (20 100).
Figure 4. Analysis of the expression in E.coli of the mutants in the slippery sequence. (A) Coomassie stained 7% SDS–PAGE showing (arrow) the purified
recombinant Ssa-fuc
B
(1.2 mg), the product of the gene fucA1
sm
(2 mg), and Ssa-fuc (4 mg). The bands with faster electrophoretic mobility result from the
proteolytic cleavage of the full-length protein (25). (B) Partial MALDIMS spectrum of the tryptic digest from mutant fucA1
sm
expressed in E.coli. The mass
signal corresponding to peptide C encompassing the overlapping region is indicated. (C) LCMSMS analysis of peptide C. The amino acid sequence inferred from
fragmentation spectra is reported. (D) Western blot of E.coli cellular extracts expressing fucA1
A
, the wild-type fucA1,fucA1
sm
and fucA1
tm
genes (Materials and
Methods). The blot was probed with anti-GST antibodies. (E) Western blot of partially purified protein samples expressed in E.coli fused to GST from wild-type
and mutant fucA1 genes. Cellular extracts were loaded on GST–Sepharose matrix. After washing, equal amounts of slurries (30 ml of 300 ml) were denaturated
and loaded on a 8% SDS–PAGE. Extracts of E.coli cells expressing the parental plasmid pGEX-2TK were used as the negative control (pGEX). The blot was
probed with anti-Ssa-fuc antibodies.
4264 Nucleic Acids Research, 2006, Vol. 34, No. 15
and YGM media, demonstrating that under these conditions
the two ORFs were co-transcribed (Figure 6A).
The experiments of real-time PCR shown in Figure 6B
demonstrated that rRNA16S was amplified after 17 cycles
while the amplification of fucA1 mRNA was observed after
38 cycles, despite the fact that we used 2500-fold more
cDNA for the amplification of fucA1. This indicates that
the gene fucA1 is transcribed at very low level. No significant
differences in the fucA1 mRNA level were observed in cells
grown in YSM or YGM media. This is further confirmed by
the analysis by western blot of the extracts of the same cells
of S.solfataricus used to prepare the total RNAs, which
revealed equal amounts of a-fucosidase in the two extracts
(Figure 6C). Therefore, the low a-fucosidase activity
observed under the conditions tested is the result of the
poor transcription of the fucA1 gene.
Analysis of the expression of fucA1 in S.solfataricus by
in vitro translation
To determine whether, and with what efficiency, the 1
frameshifting could be performed by S.solfataricus ribo-
somes, mRNAs obtained by in vitro transcription of the
cloned wild-type fucA1 gene and the mutants thereof were
used to program an in vitro translation system prepared as
described by Condo
`et al. (28). To this aim, a promoter of
T7 polymerase was inserted ahead of the gene of interest to
obtain RNA transcripts endowed with the short 50-
untranslated region of 9 nt observed for the natural fucA1
mRNA (24). Autoradiography of an SDS–PAGE of the trans-
lation products (Figure 7) revealed that the wild-type fucA1
transcript produced a tiny but clear band whose molecular
weight corresponded to that of the full-length Ssa-fuc
obtained by site-directed mutagenesis (24); the latter was
translated quite efficiently in the cell-free system in spite of
being encoded by a quasi-leaderless mRNA. Judging from
the relative intensity of the signals given by the translation
products of the wild-type fucA1 and the full-length mutant
fucA1
A
, the efficiency of the 1 frameshifting in the homo-
logous system was 10%. No signals corresponding to the
polypeptides expected from the separated ORFs SSO11867
and SSO3060 (9.6 and 46.5 kDa, respectively) were
observed. However, it should be noted that the product of
SSO11867, even if synthesized, is too small to be detected
in the gel system employed for this experiment. The larger
product of ORF SSO3060, on the other hand, is certainly
absent. These data unequivocally demonstrate that the
ribosomes of S.solfataricus can decode the split fucA1 gene
by programmed 1 frameshifting with considerable effici-
ency producing a full-length polypeptide from the two
ORFs SSO11867 and SSO3060.
Remarkably, under the same conditions at which fucA1
drives the expression of the full-length protein, we could
not observe any product from the fucA1
sm
and fucA1
tm
con-
structs. These data demonstrate that the integrity of the hep-
tanucleotide is essential for the expression of the fucA1 gene
in S.solfataricus, thus further confirming that the gene is
decoded by programmed 1 frameshifting in this organism.
In addition, the lack of expression of fucA1
sm
by translation
in vitro in S.solfataricus contrasts with the efficient
expression of this mutant in E.coli, indicating that the two
Figure 6. Analysis of the expression of fucA1 in different media. (A) Agarose
gel showing the products of RT–PCR encompassing the ORFs SSO11867 and
SSO3060 by using total cellular RNA extracted from cells grown in YSM
(lanes 1 and 2) and YGM (lanes 3 and 4); lanes 1 and 3, control (amplification
of total RNA supplemented with Taq and without reverse transcriptase
enzyme); lane 2 and 4, fucA1.(B) Comparison of the fucA1 mRNA levels in
YSM and YGM by real-time PCR. The inset shows the corresponding
products found in the real-time PCR visualized by ethidium bromide staining.
(C) Western blot of S.solfataricus extracts of cells grown in YSM (lane 2,
80 mg) and YGM (lane 3, 80 mg). Lane 1 recombinant Ssa-fuc (0.2 mg).
Nucleic Acids Research, 2006, Vol. 34, No. 15 4265
organisms recognize different sequences regulating the
translational frameshifting.
DISCUSSION
The identification of genes whose expression is regulated by
recoding events is often serendipitous. In the framework
of our studies on glycosidases from hyperthermophiles, we
identified in the genome of the archaeon S.solfataricus
a split gene encoding a putative a-fucosidase, which could
be expressed through programmed 1 frameshifting (24).
We tackled this issue by studying the expression of fucA1
in S.solfataricus and in E.coli to overcome the problems con-
nected to the scarcity of expression of the a-fucosidase gene
and to the manipulation of hyperthermophiles. As already
reported by others, in fact, it is a common strategy to study
recoding events from different organisms in E.coli (23,33).
The expression in E.coli of the wild-type split gene fucA1
led to the production by frameshifting of two full-length
polypeptides with an efficiency of 5%. This is a value higher
than that observed in other genes expressed by translational
frameshifting in a heterologous system such as the proteins
gpG and gpGT (0.3–3.5%) (33).
The gene fucA1 is expressed in S.solfataricus at very low
level under the conditions tested. In particular, the transcrip-
tional analysis of the gene revealed that it is expressed at very
low level in both YSM and YGM media. Similarly, no differ-
ences in the two media could be found by western blot probed
with anti-Ssa-fuc antibodies, indicating that the low expres-
sion of the enzyme in S.solfataricus is the result of scarce
transcription rather than suppressed translation.
Western blots allowed us to identify a specific band
8.7 kDa heavier than that of the recombinant Ssa-fuc and
experiments of translation in vitro showed that the wild-type
gene expresses a full-length polypeptide exhibiting the same
molecular mass of the recombinant protein. This demon-
strates that the translational machinery of S.solfataricus is
fully competent to perform programmed frameshifting. It
seems likely that the observed discrepancy in molecular
mass might arise from post-translational modifications that
cannot be produced by the translation in vitro. Further experi-
ments are required to characterize the a-L-fucosidase identi-
fied in S.solfataricus.
MALDIMS and LCMSMS analyses of the products in
E.coli of the wild-type split gene fucA1 demonstrated that
two independent frameshifting events occurred in vivo in
the proposed slippery site. In particular, the sequences
obtained by LCMSMS demonstrate that peptide A results
from a simultaneous backward slippage of both the P- and
the A-site tRNAs (Figure 8A). Instead, the sequence of
peptide B is the result of the re-positioning on the 1
frame of only the P-site tRNA; in fact, the next incorporated
amino acid is specified by the codon in the new frame
(Figure 8B). Therefore, the expression by 1 frameshifting
of the wild-type gene fucA1 in E.coli follows the models
proposed for ribosomal frameshifting (34). We confirmed
the significance of the slippery heptanucleotide in promoting
the programmed frameshifting in vivo by mutating the
putative regulatory sequence. The triple mutant fucA1
tm
gave no full-length products; presumably, the mutations in
both the P- and in the A-site of the slippery sequence dramati-
cally reduced the efficiency of the 1 frameshifting as
observed previously in metazoans (35). This result confirms
that the intact slippery sequence in the wild-type gene
fucA1 is absolutely necessary for its expression in E.coli.In
contrast, surprisingly, the single mutant fucA1
sm
showed an
even increased frequency of frameshifting (10%) if compared
to the wild-type and produced only one polypeptide by
shifting specifically in site B. We explained this result obser-
ving that the mutation in the P-site of the slippery sequence
A-AAA-AAT!A-AAG-AAT created a novel slippery
sequence A-AAG identical to that controlling the expression
by programmed 1 frameshifting of a transposase gene in
E.coli (36). Therefore, apparently, the single mutation inacti-
vated the simultaneous P- and A-site tRNA re-positioning
and, in the same time, fostered the shifting efficiency of
the tRNA in the P-site. It is worth noting that, instead, in
S.solfataricus, only the simultaneous slippage is effective
(Figure 8B) and even the single mutation in the slippery
sequence of fucA1
sm
completely annulled the expression of
the gene. This indicates that this sequence is essential in
the archaeon and that programmed frameshifting in
S.solfataricus and E.coli exploits different mechanisms.
Furthermore, since the only difference between the enzymes
produced by the frameshifting sites A and B, Ssa-fuc and
Ssa-fuc
B
, respectively, is the stability at 80C, which is the
S.solfataricus physiological temperature, the functionality of
Ssa-fuc
B
in the archaeon appears questionable.
The reason why fucA1 is regulated by programmed 1
frameshifting is not known. However, the physiological
significance of programmed frameshifting has been assigned
to a minority of the cellular genes while for most of them it
is still uncertain [see (4) and reference therein; (16)]. This
mechanism of recoding is exploited to set the ratio of two
polypeptides such as the tand gsubunits of the DNA
polymerase III holoenzyme in E.coli (12). Alternatively,
programmed frameshifting balances the expression of a
protein, as the bacterial translational release factor 2 and
the eukaryotic ornithine decarboxylase antizyme [see (4)
and (18) and references therein]. In the case of fucA1, the
polypeptide encoded by the smaller ORF SSO11867 could
never be detected by western blots analyses. In addition,
the modelling of Ssa-fuc on the high-resolution crystal
structure of the a-L-fucosidase from Thermotoga maritima
(25,37) showed that the fucA1 N-terminal polypeptide is
not an independent domain. Moreover, we have shown
recently that SSO11867 includes essential catalytic residues
(27), excluding the possibility that a functional a-fucosidase
can be obtained from the ORF SSO3060 alone. There-
fore, several lines of evidence allow us to exclude that
Figure 7. In vitro translation. Of each sample 15 ml was loaded on 12.5%
acrylamide–SDS gel and the newly synthesized proteins were revealed by
autoradiography. Lane 1, no mRNA added; lane 2, fucA1
sm
; lane 3, wild-type
fucA1; lane 4, full-length fucA1
A
; lane 5, fucA1
tm
.
4266 Nucleic Acids Research, 2006, Vol. 34, No. 15
programmed 1 frameshifting is used to set the ratio of two
polypeptides of the a-fucosidase from S.solfataricus. More
probably, this translational mechanism might be required to
control the expression level of fucA1.
Noticeably, this is the only fucosidase gene expressed by
programmed 1 frameshifting. Among carbohydrate active
enzymes, the only example of expression through this recod-
ing mechanism is that reported for a gene encoding for a
a(1,2)-fucosyltransferase from Helicobacter pylori that is
interrupted by a 1 frameshifting (38). In this case, the
expression by programmed frameshifting would lead to a
functional enzyme synthesizing components of the surface
lipopolysaccharides to evade the human immune defensive
system. It is hard to parallel this model to fucA1. Neverthe-
less, the monosaccharide fucose is involved in a variety of
biological functions (39). Therefore, the a-L-fucosidase
might play a role in the metabolism of fucosylated oligosac-
charides; experiments are currently in progress to knockout
the wild-type fucA1 gene and to insert constitutive functional
mutants of this gene in S.solfataricus.
FucA1 is the only archaeal a-L-fucosidase gene identified
so far; hence, it is probably the result of a horizontal gene
transfer event in S.solfataricus. However, since there are no
a-fucosidases genes regulated by programmed frameshifting
in Bacteria and Eukarya, it is tempting to speculate that
this sophisticated mechanism of translational regulation pre-
existed in S.solfataricus and it was applied to the fucosidase
gene for physiological reasons. The identification of other
genes interrupted by 1 frameshifts in S.solfataricus would
open the possibility that they are regulated by programmed
1 frameshifting. Recently, the computational analysis of
prokaryotic genomes revealed that seven Archaea harbour
interrupted coding sequences, but S.solfataricus is not
included in this study (40). A computational analysis on sev-
eral archaeal genomes revealed that 34 interrupted genes are
present in the genome of S.solfataricus, 11 of these genes
are composed by two ORFs separated by 1 frameshifting
and could be expressed by recoding (B. Cobucci-Ponzano,
M. Rossi and M. Moracci, manuscript in preparation).
We have experimentally shown here, for the first time, that
programmed 1 frameshifting is present in the Archaea
domain. This finding is the missing piece in the puzzle of
the phylogenetic distribution of programmed frameshifting
demonstrating that this mechanism is universally conserved.
ACKNOWLEDGEMENTS
We are grateful to Maria Carmina Ferrara for technical
assistance in the real-time PCR experiments and to Maria
Ciaramella and Massimo Di Giulio for useful discussion. We
thank the TIGEM-IGB DNA sequencing core for the sequenc-
ing of the clones. The IBP-CNR belongs to the Centro
Regionale di Competenza in Applicazioni Tecnologico-
Industriali di Biomolecole e Biosistemi. P.P., M. R. and
M. M. were supported by MIUR project ‘Folding di proteine:
l’altra meta
`del codice genetico’ RBAU015B47. This
work was partially supported by the ASI project MoMa n.
1/014/06/0. Funding to pay the Open Access publication
charges for this article was provided by MIUR project
RBAU015B47.
Conflict of interest statement. None declared.
REFERENCES
1. Gesteland,R.F., Weiss,R.B. and Atkins,J.F. (1992) Recoding:
reprogrammed genetic decoding. Science,257, 1640–1641.
2. Gesteland,R.F. and Atkins,J.F. (1996) Recoding: dynamic
reprogramming of translation. Annu. Rev. Biochem.,65, 741–768.
3. Farabaugh,P.J. (1996) Programmed translational frameshifting.
Annu. Rev. Genet.,30, 507–528.
Figure 8. Proposed mechanisms of programmed 1 frameshifting of fucA1.(A) Simultaneous P- and A-site slippage; (B) P-site slippage. For the sake of clarity,
the amino acids bound to the tRNA and in the peptides identified by LCMSMS are reported with the one- and the three-letters codes, respectively. The amino
acids bound to the tRNAs shown in the A-site are highlighted in boldface and italics.
Nucleic Acids Research, 2006, Vol. 34, No. 15 4267
4. Namy,O., Rousset,J.P., Napthine,S. and Brierley,I. (2004)
Reprogrammed genetic decoding in cellular gene expression. Mol. Cell,
13, 157–168.
5. Bertram,G., Innes,S., Minella,O., Richardson,J. and Stansfield,I. (2001)
Endless possibilities: translation termination and stop codon
recognition. Microbiology,147, 255–269.
6. Herr,A.J., Wills,N.M., Nelson,C.C., Gesteland,R.F. and Atkins,J.F.
(2004) Factors that influence selection of coding resumption sites in
translational bypassing: minimal conventional peptidyl-tRNA: mRNA
pairing can suffice. J. Biol. Chem.,279, 11081–11087.
7. Engelberg-Kulka,H. and Schoulaker-Schwarz,R. (1994) Regulatory
implications of translational frameshifting in cellular gene expression.
Mol. Microbiol.,1, 3–8.
8. Sung,D. and Kang,H. (2003) Prokaryotic and eukaryotic translational
machineries respond differently to the frameshifting RNA signal from
plant or animal virus. Virus Res.,92, 165–170.
9. Dos Ramos,F., Carrasco,M., Doyle,T. and Brierley,I. (2004)
Programmed 1 ribosomal frameshifting in the SARS coronavirus.
Biochem. Soc. Trans.,32, 1081–1083.
10. Craigen,W.J. and Caskey,C.T. (1986) Expression of peptide chain
release factor 2 requires high-efficiency frameshift. Nature,322,
273–275.
11. Baranov,P.V., Gesteland,R.F. and Atkins,J.F. (2002) Release factor
2 frameshifting sites in different bacteria. EMBO Rep.,4, 373–377.
12. Tsuchihashi,Z. and Kornberg,A. (1990) Translational frameshifting
generates the gamma subunit of DNA polymerase III holoenzyme.
Proc. Natl Acad. Sci. USA,87, 2516–2520.
13. Tsuchihashi,Z. and Brown,P.O. (1992) Sequence requirements for
efficient translational frameshifting in the Escherichia coli dnaX gene
and the role of an unstable interaction between tRNA(Lys) and an AAG
lysine codon. Genes Dev.,6, 511–519.
14. Mejlhede,N., Atkins,J.F. and Neuhard,J. (1999) Ribosomal 1
frameshifting during decoding of Bacillus subtilis cdd occurs at the
sequence CGA AAG. J. Bacteriol.,181, 2930–2937.
15. Shigemoto,K., Brennan,J., Walls,E., Watson,C.J., Stott,D., Rigby,P.W.
and Reith,A.D. (2001) Identification and characterisation of a
developmentally regulated mammalian gene that utilises 1
programmed ribosomal frameshifting. Nucleic Acids Res.,29,
4079–4088.
16. Wills,N.M., Moore,B., Hammer,A., Gesteland,R.F. and Atkins,J.F.
(2006) A functional 1 ribosomal frameshift signal in the human
paraneoplastic Ma3 gene. J. Biol. Chem.,281, 7082–7088.
17. Hammell,A.B., Taylor,R.C., Peltz,S.W. and Dinman,J.D. (1999)
Identification of putative programmed 1 ribosomal frameshift signals
in large DNA databases. Genome Res.,9, 417–427.
18. Baranov,P.V., Gesteland,R.F. and Atkins,J.F. (2002) Recoding:
translational bifurcations in gene expression. Gene,286, 187–201.
19. Cobucci-Ponzano,B., Rossi,M. and Moracci,M. (2005) Recoding in
archaea. Mol. Microbiol.,55, 339–348.
20. Wilting,R., Schorling,S., Persson,B.C. and Bock,A. (1997)
Selenoprotein synthesis in archaea: identification of an mRNA element
of Methanococcus jannaschii probably directing selenocysteine
insertion. J. Mol. Biol.,266, 637–641.
21. Hao,B., Gong,W., Ferguson,T.K., James,C.M., Krzycki,J.A. and
Chan,M.K. (2002) A new UAG-encoded residue in the structure of
a methanogen methyltransferase. Science,296, 1462–1466.
22. Polycarpo,C., Ambrogelly,A., Ruan,B., Tumbula-Hansen,D.,
Ataide,S.F., Ishitani,R., Yokoyama,S., Nureki,O., Ibba,M. and So
¨ll,D.
(2003) Activation of the pyrrolysine suppressor tRNA requires
formation of a ternary complex with class I and class II lysyl-tRNA
synthetases. Mol. Cell,12, 287–294.
23. Blight,S.K., Larue,R.C., Mahapatra,A., Longstaff,D.G., Chang,E.,
Zhao,G., Kang,P.T., Green-Church,K.B., Chan,M.K. and Krzycki,J.A.
(2004) Direct charging of tRNA(CUA) with pyrrolysine in vitro and
in vivo.Nature,431, 333–335.
24. Cobucci-Ponzano,B., Trincone,A., Giordano,A., Rossi,M. and
Moracci,M. (2003) Identification of an archaeal alpha-L-fucosidase
encoded by an interrupted gene. Production of a functional enzyme by
mutations mimicking programmed 1 frameshifting. J. Biol. Chem.,
278, 14622–14631.
25. Rosano,C., Zuccotti,S., Cobucci-Ponzano,B., Mazzone,M., Rossi,M.,
Moracci,M., Petoukhov,M.V., Svergun,D.I. and Bolognesi,M. (2004)
Structural characterization of the nonameric assembly of an Archaeal
alpha-L-fucosidase by synchrotron small angle X-ray scattering.
Biochem. Biophys. Res. Commun.,320, 176–182.
26. Cobucci-Ponzano,B., Trincone,A., Giordano,A., Rossi,M. and
Moracci,M. (2003) Identification of the catalytic nucleophile of the
family 29 alpha-L-fucosidase from Sulfolobus solfataricus via chemical
rescue of an inactive mutant. Biochemistry,42, 9525–9531.
27. Cobucci-Ponzano,B., Mazzone,M., Rossi,M. and Moracci,M. (2005)
Probing the catalytically essential residues of the alpha-L-fucosidase
from the hyperthermophilic archaeon Sulfolobus solfataricus.
Biochemistry,44, 6331–6342.
28. Condo
`,I., Ciammaruconi,A., Benelli,D., Ruggero,D. and Londei,P.
(1999) Cis-acting signals controlling translational initiation in the
thermophilic archaeon Sulfolobus solfataricus.Mol. Microbiol.,34,
377–384.
29. Bradford,M.M. (1976) A rapid and sensitive method for the
quantitation of microgram quantities of protein utilizing the principle of
protein-dye binding. Anal. Biochem.,72, 248–254.
30. Lu
¨bben,M. and Scha
¨fer,G. (1989) Chemiosmotic energy conversion of
the archaebacterial thermoacidophile Sulfolobus acidocaldarius:
oxidative phosphorylation and the presence of an F0-related
N, N0-dicyclohexylcarbodiimide-binding proteolipid. J. Bacteriol.,
171, 6106–6116.
31. Moracci,M., La Volpe,A., Pulitzer,J.F., Rossi,M. and Ciaramella,M.
(1992) Expression of the thermostable beta-galactosidase gene from the
archaebacterium Sulfolobus solfataricus in Saccharomyces cerevisiae
and characterization of a new inducible promoter for heterologous
expression. J. Bacteriol.,174, 873–882.
32. Tachibana,Y., Kuramura,A., Shirasaka,N., Suzuki,Y., Yamamoto,T.,
Fujiwara,S., Takagi,M. and Imanaka,T. (1999) Purification and
characterization of an extremely thermostable cyclomaltodextrin
glucanotransferase from a newly isolated hyperthermophilic archaeon,
aThermococcus sp. Appl. Environ. Microbiol.,65, 1991–1997.
33. Xu,J., Hendrix,R.W. and Duda,R.L. (2004) Conserved translational
frameshift in dsDNA bacteriophage tail assembly genes. Mol. Cell,
16, 11–21.
34. Baranov,P.V., Gesteland,R.F. and Atkins,J.F. (2004) P-site tRNA is
a crucial initiator of ribosomal frameshifting. RNA,10, 221–230.
35. Plant,E.P. and Dinman,J.D. (2006) Comparative study of the effects of
heptameric slippery site composition on 1 frameshifting among
different eukaryotic systems. RNA,12, 666–673.
36. Sekine,Y., Eisaki,N. and Ohtsubo,E. (1994) Translational control in
production of transposase and in transposition of insertion sequence
IS3. J. Mol. Biol.,235, 1406–1420.
37. Sulzenbacher,G., Bignon,C., Nishimura,T., Tarling,C.A., Withers,S.G.,
Henrissat,B. and Bourne,Y. (2004) Crystal structure of Thermotoga
maritima alpha-L-fucosidase. Insights into the catalytic mechanism and
the molecular basis for fucosidosis. J. Biol. Chem.,279, 13119–13128.
38. Wang,G., Rasko,D.A., Sherburne,R. and Taylor,D.E. (1999) Molecular
genetic basis for the variable expression of Lewis Y antigen in
Helicobacter pylori: analysis of the alpha (1,2) fucosyltransferase gene.
Mol. Microbiol.,31, 1265–1274.
39. Staudacher,E., Altmann,F., Wilson,I.B. and Marz,L. (1999) Fucose in
N-glycans: from plant to man. Biochim. Biophys. Acta,1473, 216–236.
40. Perrodou,E., Deshayes,C., Muller,J., Schaeffer,C., Van Dorsselaer,A.,
Ripp,R., Poch,O., Reyrat,J.M. and Lecompte,O. (2006) ICDS database:
interrupted CoDing sequences in prokaryotic genomes. Nucleic Acids
Res.,34, 338–343.
4268 Nucleic Acids Research, 2006, Vol. 34, No. 15
... These events regulate protein expression at the translational level, and their mechanisms are well known and characterized in viruses, bacteria, and eukaryotes. In archaea, it was demonstrated and studied that translational recoding regulates the decoding of selenocysteine and pyrrolysine, the 21st and the 22nd amino acids, and one case of programmed −1 frameshifting has been reported so far in Saccharolobus solfataricus P2 [94,[96][97][98][99]. The pyl genes have been discovered in the genomes of bacteria and archaea, and they are typically grouped together with other genes involved in the methylamine metabolism and methylamine methyltransferases [90]. ...
... The slippage can determine (1) the production of an extended, functional polypeptide from an alternative reading frame with efficiencies varying from very low to as high as 80%, and (2) the production of a non-functional polypeptide as the ribosome encounters a stop codon in the new reading frame (Figure 4) [94,103,104]. In archaea, only one case of −1PRF has been reported to date [98,104]. In the thermoacidophilic archaeon S. solfataricus strain P2, the fucA1 gene was found to be organized in two open reading frames (ORFs) SSO11867 and SSO3060 of 81 and 426 amino acids, respectively, which are separated by a −1 frameshifting in a 40-base overlap. ...
... The framefucA mutant produced a fully functional α-L-fucosidase, named Ssα-fuc, which was thermophilic, thermostable, and had an unusual nonameric structure [105][106][107]. The interrupted gene fucA1 is translated by −1PRF in both E. coli and S. solfataricus, producing a full-length protein showing for the first time that this kind of recoding is present in archaea [98]. Moreover, only the wild-type slippery sequence in S. solfataricus resulted in being functional, as shown by the in vitro translation of fucA1 and the mutant gene in the slippery sequence [98]. ...
Article
Full-text available
Archaea represents the third domain of life, displaying a closer relationship with eukaryotes than bacteria. These microorganisms are valuable model systems for molecular biology and biotechnology. In fact, nowadays, methanogens, halophiles, thermophilic euryarchaeota, and crenarchaeota are the four groups of archaea for which genetic systems have been well established, making them suitable as model systems and allowing for the increasing study of archaeal genes’ functions. Furthermore, thermophiles are used to explore several aspects of archaeal biology, such as stress responses, DNA replication and repair, transcription, translation and its regulation mechanisms, CRISPR systems, and carbon and energy metabolism. Extremophilic archaea also represent a valuable source of new biomolecules for biological and biotechnological applications, and there is growing interest in the development of engineered strains. In this review, we report on some of the most important aspects of the use of archaea as a model system for genetic evolution, the development of genetic tools, and their application for the elucidation of the basal molecular mechanisms in this domain of life. Furthermore, an overview on the discovery of new enzymes of biotechnological interest from archaea thriving in extreme environments is reported.
... For a comprehensive review on the genes expressed by PRF in Bacteria, Eukarya and viruses see Atkins et al. (2016); Rodnina et al. (2020). Among PRF, -1 frameshifting is more widespread with examples in all three domains of life (Luthi et al., 1990;Tsuchihashi and Kornberg, 1990;Cobucci-Ponzano et al., 2006;Wills et al., 2006;Belew et al., 2014), many of which are phylogenetically conserved. ...
... In Archaea only one case of -1 PRF has been reported (Cobucci-Ponzano et al., 2006). In the thermoacidophilic archaeon Saccharolobus solfataricus (formerly Sulfolobus solfataricus) (Sakai and Kurosawa, 2018) strain P2 the fucA1 gene was found to be organized in two open reading frames (ORFs) SSO11867 and SSO3060 of 81 and 426 amino acids, respectively, which are separated by a -1 frameshifting in a 40 bases overlap. ...
... The framefucA mutant encoded for a polypeptide of 495 amino acids, that, remarkably, in recombinant form produced a fully functional α-L-fucosidase, named Ssα-fuc, which was thermophilic, thermostable and had an unusual non-americ structure (Cobucci-Ponzano et al., 2003b, 2005aRosano et al., 2004). The full-length protein FucA was expressed by -1 PRF in both E. coli and S. solfataricus showing for the first time that this kind of recoding is present in Archaea (Cobucci-Ponzano et al., 2006). The observation that the fucA1 interrupted gene directed the expression of low α-L-fucosidase activity in E. coli led to the isolation and characterization of the polypeptides expressed in the recombinant form demonstrating that the fucA1 gene produced in E. coli a mixture of two full-length polypeptides, both functional, with a total efficiency of about 5% (Xu et al., 2004;Cobucci-Ponzano et al., 2006). ...
Article
Full-text available
Genetic code decoding, initially considered to be universal and immutable, is now known to be flexible. In fact, in specific genes, ribosomes deviate from the standard translational rules in a programmed way, a phenomenon globally termed recoding. Translational recoding, which has been found in all domains of life, includes a group of events occurring during gene translation, namely stop codon readthrough, programmed ± 1 frameshifting, and ribosome bypassing. These events regulate protein expression at translational level and their mechanisms are well known and characterized in viruses, bacteria and eukaryotes. In this review we summarize the current state-of-the-art of recoding in the third domain of life. In Archaea, it was demonstrated and extensively studied that translational recoding regulates the decoding of the 21st and the 22nd amino acids selenocysteine and pyrrolysine, respectively, and only one case of programmed –1 frameshifting has been reported so far in Saccharolobus solfataricus P2. However, further putative events of translational recoding have been hypothesized in other archaeal species, but not extensively studied and confirmed yet. Although this phenomenon could have some implication for the physiology and adaptation of life in extreme environments, this field is still underexplored and genes whose expression could be regulated by recoding are still poorly characterized. The study of these recoding episodes in Archaea is urgently needed.
... In Archaea, recoding, which was deeply studied only recently, was unequivocally demonstrated only for termination codon readthrough events that regulate the incorporation of the unusual amino acids selenocysteine and pyrrolysine [8,15], and −1 PRF that allow the expression of a fully functional α-L-fucosidase in the crenarchaeon Saccharolobus solfataricus [16][17][18][19][20][21]. This gene, named fucA, is organized in two open reading frames (ORFs) SSO11867 and SSO3060 of 81 and 426 amino acids, respectively, which are separated by a −1 frameshifting in a 40 bases overlap ( Figure 1A). ...
... Remarkably, we demonstrated that a full-length mutant of gene, named framefucA, obtained by inserting specific site-directed mutations in the fucA gene in the positions that were predicted to generate by −1 PRF a complete polypeptide ( Figure 1B) led to a functional enzyme α-L-fucosidase, named Ssα-fuc, of 495 amino acids, which resulted in it being thermophilic, thermostable, and having an unusual nonameric structure [16][17][18][19]22]. In addition, we showed that fucA is expressed by −1 PRF in both E. coli and S. solfataricus demonstrating, for the first time, that this kind of recoding is present in Archaea [20]. To date, only 8 archaeal α-L-fucosidases are reported and that from S. solfataricus is the only one characterized. ...
... By comparing the growth curves, we observed that all strains were viable, but the mutants had a slightly longer latency phase than WT (Figure 2A). Western Blot analysis ( Figure 2B) performed on the cellular extracts of both wild type and the two mutants, using antibodies against α-L-fucosidase, confirmed that the higher molecular band revealed in WT cellular extracts corresponded to the oligomeric form of the α-L-fucosidase as previously reported [20,22]. As expected, a more intense signal was observed in the full-length mutant FFuc strain. ...
Article
Full-text available
Genetic decoding is flexible, due to programmed deviation of the ribosomes from standard translational rules, globally termed “recoding”. In Archaea, recoding has been unequivocally determined only for termination codon readthrough events that regulate the incorporation of the unusual amino acids selenocysteine and pyrrolysine, and for −1 programmed frameshifting that allow the expression of a fully functional α-l-fucosidase in the crenarchaeon Saccharolobus solfataricus, in which several functional interrupted genes have been identified. Increasing evidence suggests that the flexibility of the genetic code decoding could provide an evolutionary advantage in extreme conditions, therefore, the identification and study of interrupted genes in extremophilic Archaea could be important from an astrobiological point of view, providing new information on the origin and evolution of the genetic code and on the limits of life on Earth. In order to shed some light on the mechanism of programmed −1 frameshifting in Archaea, here we report, for the first time, on the analysis of the transcription of this recoded archaeal α-l-fucosidase and of its full-length mutant in different growth conditions in vivo. We found that only the wild type mRNA significantly increased in S. solfataricus after cold shock and in cells grown in minimal medium containing hydrolyzed xyloglucan as carbon source. Our results indicated that the increased level of fucA mRNA cannot be explained by transcript up-regulation alone. A different mechanism related to translation efficiency is discussed.
... Cases of PRF have been reported in many viruses and domains of life such as on the bacterial Escherichia coli dnaX gene (Tsuchihashi and Kornberg, 1990), in archaea like in Sulfolobus solfataricus on the α-l-fucosidase fucA1 mRNA (Cobucci-Ponzano et al., 2006), as well as in eukaryotes on the human embryonic Paternally Expressed Gene 10 (PEG-10) (Manktelow et al., 2005;Clark et al., 2007). Movement of ribosomes during PRF can occur in both the + or − direction relative to the 5′ end of the mRNA by one to even six nucleotides (Weiss et al., 1987;Lainé et al., 2008;Fang et al., 2012;Yan et al., 2015). ...
Article
Full-text available
Translation facilitates the transfer of the genetic information stored in the genome via messenger RNAs to a functional protein and is therefore one of the most fundamental cellular processes. Programmed ribosomal frameshifting is a ubiquitous alternative translation event that is extensively used by viruses to regulate gene expression from overlapping open reading frames in a controlled manner. Recent technical advances in the translation field enabled the identification of precise mechanisms as to how and when ribosomes change the reading frame on mRNAs containing cis -acting signals. Several studies began also to illustrate that trans -acting RNA modulators can adjust the timing and efficiency of frameshifting illuminating that frameshifting can be a dynamically regulated process in cells. Here, we intend to summarize these new findings and emphasize how it fits in our current understanding of PRF mechanisms as previously described.
... Research on extremophiles and their enzymes (extremozymes) has not only reshaped our understanding of the origin and evolution of life [5] and the potential for life on other planetary bodies [6], but also it has simultaneously led to numerous advances in molecular biology, medicine, and biotechnology [7][8][9][10]. In fact, extremozymes represent interesting cases of protein adaptation under conditions where conventional enzymes quickly denature [11][12][13]. Thus, extremozymes are ideal tools for industrial applications where harsh chemical and physical conditions are encountered. ...
Article
Full-text available
Terrestrial hot springs are of great interest to the general public and to scientists alike due to their unique and extreme conditions. These have been sought out by geochemists, astrobiologists, and microbiologists around the globe who are interested in their chemical properties, which provide a strong selective pressure on local microorganisms. Drivers of microbial community composition in these springs include temperature, pH, in-situ chemistry, and biogeography. Microbes in these communities have evolved strategies to thrive in these conditions by converting hot spring chemicals and organic matter into cellular energy. Following our previous metagenomic analysis of Pisciarelli hot springs (Naples, Italy), we report here the comparative metagenomic study of three novel sites, formed in Pisciarelli as result of recent geothermal activity. This study adds comprehensive information about phylogenetic diversity within Pisciarelli hot springs by peeking into possible mechanisms of adaptation to biogeochemical cycles, and high applicative potential of the entire set of genes involved in the carbohydrate metabolism in this environment (CAZome). This site is an excellent model for the study of biodiversity on Earth and biosignature identification, and for the study of the origin and limits of life.
... 52) (GH116). These activities are involved in the hydrolysis and removal of sugar appendages of (hemi)cellulose polysaccharides, in starch/glycogen mobilization, and in the turnover of the oligosaccharides of Nglycosylated proteins[49][50][51][52][53][54][55][56][57][58][59][60][61][62][63][64][65]. On the other hand, several GH families in Pool1 and 2 (GH2, GH65, GH78, ...
Article
Full-text available
The enzymes from hyperthermophilic microorganisms populating volcanic sites represent interesting cases of protein adaptation and biotransformations under conditions where conventional enzymes quickly denature. The difficulties in cultivating extremophiles severely limit access to this class of biocatalysts. To circumvent this problem, we embarked on the exploration of the biodiversity of the solfatara Pisciarelli, Agnano (Naples, Italy), to discover hyperthermophilic carbohydrate‐active enzymes (CAZymes) and to characterize the entire set of such enzymes in this environment (CAZome). Here, we report the results of the metagenomic analysis of two mud/water pools that greatly differ in both temperature and pH (T = 85 °C and pH 5.5; T = 92 °C and pH 1.5, for Pool1 and Pool2, respectively). DNA deep sequencing and following in silico analysis led to 14 934 and 17 652 complete ORFs in Pool1 and Pool2, respectively. They exclusively belonged to archaeal cells and viruses with great genera variance within the phylum Crenarchaeota, which reflected the difference in temperature and pH of the two Pools. Surprisingly, 30% and 62% of all of the reads obtained from Pool1 and 2, respectively, had no match in nucleotide databanks. Genes associated with carbohydrate metabolism were 15% and 16% of the total in the two Pools, with 278 and 308 putative CAZymes in Pool1 and 2, corresponding to ~ 2.0% of all ORFs. Biochemical characterization of two CAZymes of a previously unknown archaeon revealed a novel subfamily GH5_19 β‐mannanase/β‐1,3‐glucanase whose hemicellulose specificity correlates with the vegetation surrounding the sampling site, and a novel NAD⁺‐dependent GH109 with a previously unreported β‐N‐acetylglucosaminide/β‐glucoside specificity. Databases The sequencing reads are available in the NCBI Sequence Read Archive (SRA) database under the accession numbers SRR7545549 (Pool1) and SRR7545550 (Pool2). The sequences of GH5_Pool2 and GH109_Pool2 are available in GenBank database under the accession numbers MK869723 and MK86972, respectively. The environmental data relative to Pool1 and Pool2 (NCBI BioProject PRJNA481947) are available in the Biosamples database under the accession numbers SAMN09692669 (Pool1) and SAMN09692670 (Pool2).
... PRF was initially identified in viral genomes, where it plays an important role in viral propagation by modulating synthesis of viral proteins in specific stoichiometric ratios (81,82). Examples of -1PRF were found in all three domains of life (83)(84)(85)(86)(87)(88). In eukaryotes, frameshifting can regulate the stability of an mRNA. ...
Article
Full-text available
During canonical translation, the ribosome moves along an mRNA from the start to the stop codon in exact steps of one codon at a time. The collinearity of the mRNA and the protein sequence is essential for the quality of the cellular proteome. Spontaneous errors in decoding or translocation are rare and result in a deficient protein. However, dedicated recoding signals in the mRNA can reprogram the ribosome to read the message in alternative ways. This review summarizes the recent advances in understanding the mechanisms of three types of recoding events: stop-codon readthrough, -1 ribosome frameshifting and translational bypassing. Recoding events provide insights into alternative modes of ribosome dynamics that are potentially applicable to other non-canonical modes of prokaryotic and eukaryotic translation.
... This is particularly true for the conversion of plant lignocellulosic material including cellulose and hemicelluloses (xylans, xyloglucans, pectins, etc.) that are the two most abundant polymers on Earth (global cellulose production estimates range between 9 9 10 12 and 1.5 9 10 12 tons/year Pinkert et al. 2009) and are remarkably stable to spontaneous hydrolysis (half-life of the glycosidic bond is 4.7 9 10 6 years, Wolfenden et al. 1998). Thus, carbohydrate active enzymes (cazymes) from (hyper)thermophiles have interesting biotechnological potential (Cobucci-Ponzano et al. 2006, 2010a. In fact, their impressive stability (Ausili et al. 2004) at the conditions at which plant lignocellulose is pretreated in second generation biorefineries (steamexplosion at extremes of temperatures and pHs), make them the ideal catalysts for the hydrolysis of (hemi)cellulose into fermentable sugars for the production of bioethanol and plastic precursors (Castiglia et al. 2016;Cobucci-Ponzano et al. 2013Iacono et al. 2016;Aulitto et al. 2017). ...
Article
Full-text available
Geothermally heated regions of Earth, such as terrestrial volcanic areas (fumaroles, hot springs, and geysers) and deep-sea hydrothermal vents, represent a variety of different environments populated by extremophilic archaeal and bacterial microorganisms. Since most of these microbes thriving in such harsh biotopes, they are often recalcitrant to cultivation; therefore, ecological, physiological and phylogenetic studies of these microbial populations have been hampered for a long time. More recently, culture-independent methodologies coupled with the fast development of next generation sequencing technologies as well as with the continuous advances in computational biology, have allowed the production of large amounts of metagenomic data. Specifically, these approaches have assessed the phylogenetic composition and functional potential of microbial consortia thriving within these habitats, shedding light on how extreme physico-chemical conditions and biological interactions have shaped such microbial communities. Metagenomics allowed to better understand that the exposure to an extreme range of selective pressures in such severe environments, accounts for genomic flexibility and metabolic versatility of microbial and viral communities, and makes extreme- and hyper-thermophiles suitable for bioprospecting purposes, representing an interesting source for novel thermostable proteins that can be potentially used in several industrial processes.
... In the genome of S. solfataricus, only an α-L-fucosidase presumably regulated by translational frameshifting was identified so far (Cobucci-Ponzano et al., 2003;Cobucci-Ponzano et al., 2006). In addition no homologs to any of the known enzymes of the E. coli L-fucose degradation pathway were identified. ...
Article
Archaea are characterised by a complex metabolism with many unique enzymes that differ from their bacterial and eukaryotic counterparts. The thermoacidophilic archaeon Sulfolobus solfataricus is known for its metabolic versatility and is able to utilize a great variety of different carbon sources. However, the underlying degradation pathways and their regulation are often unknown. In this work, we analyse growth on different carbon sources using an integrated systems biology approach. The comparison of growth on L-fucose and D-glucose allows first insights into the genome-wide changes in response to the two carbon sources and revealed a new pathway for L-fucose degradation in S. solfataricus. During growth on L-fucose we observed major changes in the central carbon metabolic network, as well as an increased activity of the glyoxylate bypass and the 3-hydroxypropionate/4-hydroxybutyrate cycle. Within the newly discovered pathway for L-fucose degradation the following key reactions were identified: (i) L-fucose oxidation to L-fuconate via a dehydrogenase, (ii) dehydration to 2-keto-3-deoxy-L-fuconate via dehydratase, (iii) 2-keto-3-deoxy-L-fuconate cleavage to pyruvate and L-lactaldehyde via aldolase and (iv) L-lactaldehyde conversion to L-lactate via aldehyde dehydrogenase. This pathway as well as L-fucose transport shows interesting overlaps to the D-arabinose pathway, representing another example for pathway promiscuity in Sulfolobus species. This article is protected by copyright. All rights reserved.
Thesis
Full-text available
Programmed ribosomal frameshifting (PRF) represents an important mechanism for translational genetic recoding, especially in viruses. The components of a PRF stimulator have been well characterized, though accounting for the variation in the frameshift stimulating efficiency has thus far been elusive. Frameshift efficiencies at known PRF sites vary from a few percent to 70-80%, and several studies have been undertaken to determine what distinguishes a high efficiency PRF site from a low efficiency PRF site via structural characterization of the stimulatory structure. Observations suggest that conformational plasticity, the ability of a certain sequence to adopt multiple conformations, is correlated with frameshift efficiency. We examine a very high efficiency (70%) PRF stimulatory structure responsible for the NS1′ frameshift in West Nile virus (WNV) to determine its characteristics. We find a high degree of structural plasticity and heterogeneity; the PRF signal exhibits multiple different starting states and unfolds via two main pathways. Furthermore, we characterize the structures involved in these pathways, and find that they correspond to predicted structures using bioinformatic predictions and SHAPE analysis. Moreover, we suggest a new operational metric of conformational plasticity, one that obviates two existing problems with the previous method for defining conformational plasticity, namely the requirement to specify a native state, and the insensitivity to multiple conformations. Additionally, we extend this definition to be force dependent, and find that the value of this conformational plasticity metric in the force range of ribosomal stalling correlates highly with frameshifting efficiency. These results may elucidate the process of frameshifting by illustrating the relationship between conformational plasticity within a specific force range and frameshift efficiency. In addition, the characterization of a high efficiency frameshift signal allows for a better understanding of the structural dynamics underlying frameshifting.
Article
Full-text available
The expression of some genes requires a high proportion of ribosomes to shift at a specific site into one of the two alternative frames. This utilized frameshifting provides a unique tool for studying reading frame control. Peptidyl-tRNA slippage has been invoked to explain many cases of programmed frameshifting. The present work extends this to other cases. When the A-site is unoccupied, the P-site tRNA can be repositioned forward with respect to mRNA (although repositioning in the minus direction is also possible). A kinetic model is presented for the influence of both, the cognate tRNAs competing for overlapping codons in A-site, and the stabilities of P-site tRNA:mRNA complexes in the initial and new frames. When the A-site is occupied, the P-site tRNA can be repositioned backward. Whether frameshifting will happen depends on the ability of the A-site tRNA to subsequently be repositioned to maintain physical proximity of the tRNAs. This model offers an alternative explanation to previously published mechanisms of programmed frameshifting, such as out-of-frame tRNA binding, and a different perspective on simultaneous tandem tRNA slippage. Keywords • translation • recoding • kinetic model • frameshifting • ribosome
Article
Full-text available
Pyrrolysine is the 22nd amino acid. An unresolved question has been how this atypical genetically encoded residue is inserted into proteins, because all previously described naturally occurring aminoacyl-tRNA synthetases are specific for one of the 20 universally distributed amino acids. Here we establish that synthetic L-pyrrolysine is attached as a free molecule to tRNACUA by PylS, an archaeal class II aminoacyl-tRNA synthetase. PylS activates pyrrolysine with ATP and ligates pyrrolysine to tRNACUA in vitro in reactions specific for pyrrolysine. The addition of pyrrolysine to Escherichia coli cells expressing fylT (encoding tRNACUA) and pylS results in the translation of UAG in vivo as a sense codon. This is the first example from nature of direct aminoacylation of a tRNA with a non-canonical amino acid and shows that the genetic code of E. coli can be expanded to include UAG-directed pyrrolysine incorporation into proteins.
Article
Full-text available
Synthesis of the gamma-subunit of DNA polymerase III holoenzyme depends on precise and efficient translational frameshifting to the -1 frame at a specific site in the dnaX gene of Escherichia coli. In vitro mutagenesis of this frameshift site demonstrated the importance of an A AAA AAG heptanucleotide sequence, which allows two adjacent tRNAs to retain a stable interaction with mRNA after they slip to the -1 position. The AAG lysine codon present in the 3' half of this heptanucleotide was a key element for highly efficient frameshifting. A tRNA(Lys) with a CUU anticodon, which has a strong affinity for AAG lysine codons, is present in eukaryotic cells but absent in E. coli. Expression in E. coli of a mutant tRNA(Lys) with a CUU anticodon specifically inhibited the frameshifting at the AAG codon, suggesting that the absence of this tRNA in E. coli contributes to the efficiency of the dnaX frameshift.
Article
Full-text available
The lacS gene from the extremely thermoacidophilic archaebacterium Sulfolobus solfataricus encodes an enzyme with beta-galactosidase activity that, like other enzymes from this organism, is exceptionally thermophilic (optimal activity above 90 degrees C), thermostable, and resistant to common protein denaturants and proteases. Expression of the gene in mesophilic hosts is needed to uncover the molecular nature of these features. We have obtained expression of beta-galactosidase in Saccharomyces cerevisiae under the control of the galactose-inducible upstream activating sequence of the yeast genes GAL1 and GAL10. The expressed enzyme is identical in molecular mass, thermostability, and thermophilicity to the native enzyme, showing that these features are intrinsic to the primary structure of the enzyme. We also present a new promoter for the expression of thermostable proteins in S. cerevisiae. This promoter contains a sequence isolated from the nematode Caenorhabditis elegans that works as a strong, heat-inducible upstream activating sequence in S. cerevisiae. Transcription of the lacS gene under the control of this sequence is rapidly and efficiently induced by heat shock. The availability of a plate assay for monitoring beta-galactosidase activity in S. cerevisiae may allow screening for mutants affecting the efficiency and activity of the enzyme.
Article
Full-text available
The dnaX gene (previously called dnaZX) of Escherichia coli has only one open reading frame for a 71-kDa polypeptide from which two distinct DNA polymerase III holoenzyme subunits, tau (71 kDa) and gamma (47 kDa), are produced. To determine how the gamma subunit is generated, we examined the influence of mutations in the dnaX gene on the pattern of tau and gamma production in overproducing cells. Important structural elements in dnaX mRNA include a stretch of six adenines (nucleotides 1425-1430), a stable hairpin structure (nucleotides 1437-1466), and a UGA stop codon in a -1 frame (nucleotides 1434-1436) between the stretch of adenines and the hairpin structure. Disruption of this stop codon generates a slightly larger gamma subunit, indicative of the use of a -1 stop codon farther downstream (nucleotides 1470-1472). These results suggest that a -1 frameshift during translation allows the use of this UGA codon to terminate translation of the gamma polypeptide. The amino acid composition, sequence, and mass spectra of a C-terminal peptide from mild digestion of the purified gamma protein with endoproteinase Lys-C confirms that this frameshift occurs at either of the two lysine codons in the region of the adenine stretch. Remarkable features of this frameshifting are its high frequency (i.e., about 80% in an overproducing cell) and the striking structural similarity to the frameshifting signal responsible for expression of the pol and pro genes in many retroviruses.
Article
A protein determination method which involves the binding of Coomassie Brilliant Blue G-250 to protein is described. The binding of the dye to protein causes a shift in the absorption maximum of the dye from 465 to 595 nm, and it is the increase in absorption at 595 nm which is monitored. This assay is very reproducible and rapid with the dye binding process virtually complete in approximately 2 min with good color stability for 1 hr. There is little or no interference from cations such as sodium or potassium nor from carbohydrates such as sucrose. A small amount of color is developed in the presence of strongly alkaline buffering agents, but the assay may be run accurately by the use of proper buffer controls. The only components found to give excessive interfering color in the assay are relatively large amounts of detergents such as sodium dodecyl sulfate, Triton X-100, and commercial glassware detergents. Interference by small amounts of detergent may be eliminated by the use of proper controls.
Article
Genes encoding methanogenic methylamine methyltransferases all contain an in-frame amber (UAG) codon that is read through during translation. We have identified the UAG-encoded residue in a 1.55 angstrom resolution structure of the Methanosarcina barkeri monomethylamine methyltransferase (MtmB). This structure reveals a homohexamer comprised of individual subunits with a TIM barrel fold. The electron density for the UAG-encoded residue is distinct from any of the 21 natural amino acids. Instead it appears consistent with a lysine in amide-linkage to (4R,5R)-4-substituted-pyrroline-5-carboxylate. We suggest that this amino acid be named l-pyrrolysine.
Article
A protein determination method which involves the binding of Coomassie Brilliant Blue G-250 to protein is described. The binding of the dye to protein causes a shift in the absorption maximum of the dye from 465 to 595 nm, and it is the increase in absorption at 595 nm which is monitored. This assay is very reproducible and rapid with the dye binding process virtually complete in approximately 2 min with good color stability for 1 hr. There is little or no interference from cations such as sodium or potassium nor from carbohydrates such as sucrose. A small amount of color is developed in the presence of strongly alkaline buffering agents, but the assay may be run accurately by the use of proper buffer controls. The only components found to give excessive interfering color in the assay are relatively large amounts of detergents such as sodium dodecyl sulfate, Triton X-100, and commercial glassware detergents. Interference by small amounts of detergent may be eliminated by the use of proper controls.