Download Ribosomal Stalk Protein L12: Structure, Function and

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

Signal transduction wikipedia , lookup

Histone acetylation and deacetylation wikipedia , lookup

SR protein wikipedia , lookup

Magnesium transporter wikipedia , lookup

Protein moonlighting wikipedia , lookup

Proteasome wikipedia , lookup

Protein (nutrient) wikipedia , lookup

Protein wikipedia , lookup

Protein phosphorylation wikipedia , lookup

List of types of proteins wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Protein structure prediction wikipedia , lookup

Intrinsically disordered proteins wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

JADE1 wikipedia , lookup

Biosynthesis wikipedia , lookup

Protein–protein interaction wikipedia , lookup

Proteolysis wikipedia , lookup

Epitranscriptome wikipedia , lookup

Transcript
To my Teachers
List of Papers
This thesis is based on the following papers, which are referred to in the text
by their Roman numerals
I
Helgstrand, M., Mandava, C. S., Mulder, F. A., Liljas, A., Sanyal,
S., and Akke, M. (2007) The ribosomal stalk binds to translation
factors IF2, EF-Tu, EF-G and RF3 via a conserved region of the L12
C-terminal domain, J Mol Biol 365, 468-479
II
Huang, C., Mandava, C. S., and Sanyal, S. (2010) The ribosomal
stalk plays a key role in IF2-mediated association of the ribosomal
subunits, J Mol Biol 399, 145-153
III
Mandava, C. S., Ederth, J., Peisker, K., Kumar, R., Szaflarski, W.,
and Sanyal, S. (2011) Bacterial ribosome requires more than one L12
dimer for efficient initiation and elongation of protein synthesis
involving IF2 and EF-G. (Submitted Manuscript)
IV
Ederth, J*., Mandava, C. S*., Dasgupta, S., and Sanyal, S. (2009) A
single-step method for purification of active His-tagged ribosomes
from a genetically engineered Escherichia coli, Nucleic Acids Res 37,
e15. (* equal contribution)
Reprints were made with permission from the respective publishers.
Paper not included in this thesis
I
Tobin, C., Mandava, CS., Ehrenberg, M., Andersson, DI., Sanyal,S.
(2010) Ribosomes lacking protein S20 are defective in mRNA binding and subunit association. Journal of Molecular Biology,
397(3):767–76
Contents
Background ................................................................................................... 11
Ribosome – a macromolecular complex .................................................. 11
Basic components of the ribosome ...................................................... 11
Structural landmarks on the bacterial ribosome .................................. 12
Functional sites on the ribosome ......................................................... 13
Ribosomal stalk ........................................................................................ 15
L7/L12 ................................................................................................. 15
Domains in the L12 dimer ................................................................... 16
The mode of dimerization .................................................................... 16
Structure of the L12 dimers off and on the ribosome .......................... 17
Stoichiometry of the L12 proteins on ribosome .................................. 18
The binding site of the L12 dimers ...................................................... 19
The factor interaction site on L12 ........................................................ 20
Role of L12 in translation .................................................................... 20
Translation................................................................................................ 21
Translation components ....................................................................... 21
The steps in protein synthesis .............................................................. 23
Aim ............................................................................................................... 27
Interaction of L12 CTD with translational G-factors ............................... 27
Role of L12 in translation initiation ......................................................... 29
Why do bacteria need multiple copies of L12 protein on ribosome? ....... 31
Technological application with stalk protein L12 .................................... 34
Current work and future plan ........................................................................ 36
Kinetic studies on translation initiation .................................................... 36
Future plans .............................................................................................. 38
Identification of residues on L12 involved in the interaction of IF2
during subunit association ................................................................... 38
Comparison of E. coli ribosomes with single, double and triple L12
dimers .................................................................................................. 39
Conclusions ................................................................................................... 40
Summary in Swedish .................................................................................... 41
Ribosomala stjälkprotein L12: struktur, funktion och tillämpningar. ...... 41
Bakgrund ............................................................................................. 41
Samverkan mellan L12 och translationella G-faktorer ........................ 41
L12s roll i initieringen av proteinsyntesen .......................................... 42
Varför behöver bakterierna flera kopior av L12 protein på
ribosomen?........................................................................................... 42
Teknisk tillämpning av stjälkproteinet L12 ......................................... 42
Acknowledgement ........................................................................................ 43
References ..................................................................................................... 44
Abbreviations
DNA
RNA
rRNA
mRNA
tRNA
aa-tRNA
r-protein
Mda
kDa
Å
bps
nt
CP
PTC
SD
aSD
TIR
RBS
PIC
IC
A-site
P-site
E-site
IF
EF
RF
RRF
CTD
NTD
GTP
GDP
ATP
FRET
fMet
Met
Leu
Deoxyribonucleic acid
Ribonucleic acid
Ribosomal RNA
Messenger RNA
Transfer RNA
Amino-acyl tRNA
Ribosomal protein
Mega Dalton
Kilo Dalton
Ångström
base pairs
nucleotide
Central Protuberance
Peptidyl-Transferase Centre
Shine-Dalgarno sequence
anti-Shine-Dalgarno sequence
Translation Initiation Region
Ribosome Binding Site
Preinitiation Complex
Initiation Complex
Amino-acyl site
Peptidyl site
Exit site
Initiation Factor
Elongation Factor
Release Factor
Ribosome recycling Factor
C terminal domain
N terminal domain
Guanosine triphosphate
Guanosine diphosphate
Adenosine triphosphate
Fluorescence Resonance Energy Transfer
Formyl Methionine
Methionine
Leucine
cryo-EM
GEF
GAC
E. coli
cryo-Electron Microscopy
Guanine nucleotide Exchange Factor
GTPase associated center
Escherichia coli
Background
Ribosome – a macromolecular complex
The basic mechanism of protein synthesis is similar throughout all
forms of life, where ribosome plays a crucial role together with its accessories. Ribosomes are large ribonucleoprotein complexes with 200- 250 Å in
diameter and with an approximate mass of 2.5 MDa (for review see (Wilson
and Nierhaus 2007; Schmeing and Ramakrishnan 2009)). The macromolecular nature of the ribosome is evident from the diversity of its RNA and protein components as discussed in the section below. The ribosomes are located in the cytoplasm; in eukaryotes they are also found on endoplasmic
reticulum. Ribosomes translate the genetic information encoded in a messenger RNA (mRNA) into the corresponding sequence of amino acids, thereby synthesizing a polypeptide chain or a protein. In Escherichia coli (E.
coli) cells, ribosomes constitute about 50% of the total dry cell mass during
exponential growth, their number varies from 40000 – 70000 per cell.
Basic components of the ribosome
The ribosome is composed of ribosomal proteins (r-proteins) and ribosomal RNAs (rRNAs) arranged in two subunits, the large subunit is about
twice the size of the small subunit. The ribosomes, ribosomal subunits and
rRNAs are called by their sedimentation coefficient (S). The intact ribosome
is called 70S in prokaryotes and 80S in eukaryotes. In prokaryotic and eukaryotic ribosomes the large subunits are called 50S and 60S, and the small
subunits are called 30S and 40S, respectively. These ribosomes, although
very similar in structural organization, differ in their protein and RNA components. In the prokaryotic 70S ribosome, the large subunit (50S) contains
two RNAs, 23S rRNA (~2900 nucleotides) and 5S rRNA (~120 nucleotides)
along with 33 proteins. The small subunit (30S) contains only one RNA
component called the 16S rRNA (~1500 nucleotides) and about 21 proteins.
Eukaryotic ribosomes are larger and much more complex than the prokaryotic ribosomes. The large subunit in eukaryotes is composed of 28S rRNA
(~4700 nucleotides), 5S rRNA (~120 nucleotides) and an additional 5.8S
rRNA (~160 nucleotides) together with 49 r-proteins. In contrast, the small
subunit (40S) is composed of 18S rRNA (~1900 nucleotides) and 33 proteins
(for review see (Wilson and Nierhaus 2003; Marintchev and Wagner 2004)).
11
The ribosomes contain around two- thirds of RNA and one third of protein.
The high RNA content is attributed to the fact that the main functions of the
ribosome i.e. mRNA decoding and peptide bond formation are driven by
rRNA. In each ribosomal subunit rRNA makes the back bone or the core to
which the r-proteins attach. The r-proteins often have distinct functions, yet
in general, they play an important role in maintenance of the structural integrity of the rRNA core.
Ribosomal subunits are recycled and reused through the cycles of
translation. During translation initiation, the subunits associate via interactions along their interface to form the 70S ribosome. However, after completion of synthesis of a peptide the ribosomes split into the individual subunits,
thus completing the cycle. The two subunits perform different roles in protein synthesis. The 30S subunit participates in decoding the information on
mRNA and the 50S subunit catalyzes the peptide bond formation. As mentioned above, both functions are driven by RNA.
Structural landmarks on the bacterial ribosome
Each ribosomal subunit has a unique structure with some ‘landmark’
features on them. The 30S subunit has a relatively simple structure. The upper part of 30S is called the ‘head’, which is connected through a ‘neck’ to
the lower part or ‘body’ carrying a shoulder (Figure 1). The 50S subunit
presents a compact structure consisting of a hemisphere base with three protuberances on it, named L1 stalk, central protuberance (CP) and the L7/L12
stalk (Figure 1). The L1 stalk is composed of L1 protein and its 23S rRNA
binding region. The L7/L12 stalk, also commonly called the ‘ribosomal
stalk’ contains multiple copies of L7/L12 protein (referred hereafter as only
L12) and one L10 protein bound to the rRNA next to the binding site of
another r-protein L11. Protein L11 constitutes the ‘stalk base’, which is a
part of an important region of the ribosome called GTPase associated center
(GAC). The 50S subunit also contains a tunnel through which the nascent
peptide is believed to exit to the cytosol. The tunnel is about 100 Å long and
up to 25 Å wide in diameter, it can accommodate a peptide of approximately
40 amino acids in length (Yonath, Leonard et al. 1987; Nissen, Hansen et al.
2000).
12
Figure 1: Structural landmarks on ribosomal subunits.
Functional sites on the ribosome
tRNA binding sites
Both the ribosomal subunits have three distinct tRNA binding sites
named as A-site, P-site and E-site (Rheinberger, Sternbach et al. 1981). The
A-site accommodates the aminoacyl tRNAs, the P-site holds the peptidyl
tRNAs after translocation and the E-site binds nonspecifically the deacylated
tRNAs (Figure 2).
A-site is the binding site for the incoming aminoacyl-tRNAs (aatRNA). The codon anticodon interactions are monitored in the A-site at the
decoding center (DC) on the 30S subunit, this allows mostly cognate and
some near cognate tRNAs to bind to the mRNA (for review see (Wilson and
Nierhaus 2003)). The A-site is also the binding site for many translation
factors during various steps of translation. The P-site is located next to the
A-site on the ribosome. During initiation of protein synthesis, initiator tRNA
(fMet-tRNAfMet) binds directly to the P-site on the 30S subunit. Other than
that, peptidyl tRNAs occupy the P-site after translocation until the next peptidyl transfer. The deacylated tRNAs when translocated from the P-site move
to the E-site. It has been suggested that binding of aminoacyl tRNA in the
A-site decreases the affinity of deacylated tRNA in the E-site thereby triggering tRNA release from the E-site (Rheinberger and Nierhaus 1986). Thus,
tRNA binding to A-site and E-site are possibly regulated in a coordinated
manner.
13
Figure 2: Thermus thermophilus 70S ribosome with E, P and A site tRNAs, adopted
from (Selmer, Dunham et al. 2006). DC is the decoding center, PTC is the peptidyl
transferase center and GAC is the GTPase associated center.
Decoding center (DC)
The decoding center (Figure 2) is located at the A-site of the 30S subunit which includes the helix 44 and 34 and loop 530 of 16S rRNA
(Schluenzen, Tocilj et al. 2000). Universally conserved bases A1492, A1493
and G530 of 16S rRNA are the important components at the decoding center
(Ogle, Brodersen et al. 2001). The process of decoding is explained in the
later part of the thesis.
Peptidyl transfer center (PTC)
The peptidyl transfer center or PTC is the catalytic heart of the 50S
subunit which is located near the central protuberance and it is almost exclusively composed of RNA (Figure 2). N-terminal tail of L27 is the only protein part that is present close to the PTC. The central loop of domain V of
23S rRNA is involved in the catalytic act of PTC, this loop is also known as
the “PTC ring”. The major catalytic activity of PTC is the peptide bond formation during the elongation of protein synthesis. PTC is also involved in
the release of the polypeptide chain during the termination of protein synthesis.
GTPase associated center (GAC)
The GTPase associated center or GAC (Figure 2) is known to stimulate the GTPase activity of several G-protein factors during translation. GAC
is composed of helices 42-44 located at the domain II of 23S rRNA and the
protein L11 (Li, Sengupta et al. 2006). Protein L10 and L11 although bind to
the same rRNA region but differ in their binding site, L10 interacts with
helices 42 and 43, whereas L11 binds to helices 43 and 44. The GAC together with helix 95 or sarcin resin loop (SRL) of 23S rRNA are responsible for
the binding of all translational GTPases and their GTPase activation.
14
Ribosomal stalk
Ribosomal stalk is a finger like protuberance on the large subunit of
the ribosome (Figure 3) (Strycharz, Nomura et al. 1978; Agrawal, Heagle et
al. 1999) (for review see (Chandra Sanyal and Liljas 2000)), composed
mainly of two to three dimers of L12 protein associated with one L10 protein. In E. coli, two dimers of L12 bind to one L10 protein (Pettersson and
Liljas 1979) and this pentameric complex is known as the L8 complex. L12
is the only multicopy protein on the ribosome (Hardy 1975). The other
unique feature it possesses is that it has no direct contact with rRNA. L12
proteins bind to rRNA via C terminal domain of the L10 protein. During
translation, four translational G-factors (IF2, EF-Tu, EF-G, and RF3) interact
with the ribosomal stalk, which is closely associated with a special region on
the ribosome, known as the GTPase associated center (GAC). This region
also includes the r-proteins L10 and L11 bound on a stretch of rRNA (nt
1028-1124) located at the base of the stalk (Schmidt, Thompson et al. 1981;
Beauclerk, Cundliffe et al. 1984).
(a)
(b)
I
Figure 3: (a) A model of 50S and 30S ribosomal subunit showing relative organization of L12 dimers, L10 and L11 proteins. I, II, and III indicate alternative locations
for the C-terminal domain of L12 on the 50S subunit. The figure is adopted from
(Dey, Bochkariov et al. 1998) (b) High-resolution structure of the ribosome showing
the components of ribosomal stalk and its base. The image was created in PyMol
from the coordinates of T. thermophilus ribosome (PDB id 2WRI and 2WRJ) (Gao,
Selmer et al. 2009).
L7/L12
In E. coli, L12 is composed of 120 amino acids with a molecular mass
of 12 kDa. Protein L12 is also known as L7/L12 as it exists in an N-terminal
acetylated form named L7, where the Serine residue at the N-terminus is
15
modified by acetylation (Terhorst, Moller et al. 1972). The L7:L12 ratio in
the cell changes depending on the growth conditions. L12 acetylation increases during stress such as when cells are grown in minimal media in nutrient depleted condition. It is believed that the acetylation of L12 increases
the stability of the pentameric complex (L8 complex) (Gordiyenko, Deroo et
al. 2008) meaning both dimers of L12 remain stably bound to L10.
Domains in the L12 dimer
L12 is one of the best studied r-protein. In general, L12 proteins form
a strong dimer in solution. The L12 dimer consists of two organized domains
connected by one flexible hinge region. The N-terminal domain spanning
residues 1-37 is responsible for the strong dimer interaction and for binding
to the protein L10 (Bocharov, Gudkov et al. 1998). This domain in the L12
monomer contains two α-helices arranged in a hair pin conformation forming a four helix bundle in the L12 dimer. The C-terminal domain of L12 is
globular, highly conserved and provides sites for interaction with the
GTPase factors (Helgstrand, Mandava et al. 2007). This compact domain
consists of 3 α-helices and 3-β sheets spanning the residues 50-120. The N
and C-terminal domains are connected by the hinge region which includes
the residues from 38 to 49 (Pettersson, Hardy et al. 1976; Gudkov 1997;
Mulder, Bouakaz et al. 2004). The hinge region is highly flexible and can
change conformation from a compact helix to an extended structure. The
flexible nature of the hinge is very important for the function of L12. Due to
the flexible hinge the CTDs of L12 can be seen at different locations on the
ribosome (Figure 3a)(Olson, Sommer et al. 1986; Zecherle, Oleinikov et al.
1992) and also can be extended out of the ribosome for the recruitment of
the translation factors (Liljas and Gudkov 1987; Diaconu, Kothe et al. 2005).
The mode of dimerization
There have been several attempts to describe the dimerization mode of
L12 on and off the ribosome. Following the crystal structure of the incomplete tetramer of L12 from Thermotoga maritima Wahl et al. discussed two
possible dimerization modes for L12 (Wahl, Bourenkov et al. 2000); (i) a
parallel dimer involving N- terminal helices α2 and α3 (hinge in an α-helical
conformation) of two full length molecules arranged side by side (ii) an antiparallel dimer where α1 and α2 at the N-terminus of a full length L12 monomer is linked to the same helices on the N-terminus of the other (incomplete) L12 monomer. In the parallel dimer model, the main interaction between the L12 monomers involved the flexible hinge region (α3), compacted
as helix (Figure 4a). Thus, this model although favoured by the authors was
somewhat unrealistic as the flexible hinge is not always helical in shape.
Instead, Sanyal and Liljas proposed the antiparallel dimer of L12 as the func16
tional dimer, where one of the two hinge regions form a compact α-helix
whereas the other is unstructured and extended (Figure 4b) (Chandra Sanyal
and Liljas 2000). Later, the antiparallel mode of dimerization of L12 was
confirmed by Moens et al. with FRET assays where the distance between the
L12 CTD and NTDs were measured by incorporation of specific fluorescence probes (Moens, Wahl et al. 2005). The four helix bundle N-terminal
dimers, as proposed in the antiparallel dimers, were also seen in the high
resolution crystal structure of T. maritima L10-(L12-NTD)6 complex
(Diaconu, Kothe et al. 2005).
(a)
(b)
Helical
hinge
(c)
Flexible
hinge
α2
α2
α1
α1
α2
α1
α2
α1
Figure 4: (a) Arrangement of two full length L12 molecules in parallel dimerization
mode involving α2 and α3 helices of the monomers. The image was created in PyMol from the coordinates of T. maritima L12 (PDB id 1DD4) (Wahl, Bourenkov et
al. 2000) (b) Model of antiparallel L12 dimer involving α1 and α2 helices where one
hinge is a helix and the other is an extended structure, adopted from (Chandra
Sanyal and Liljas 2000) (c) NMR structure of L12 dimer in solution showing three
distinct regions, adopted from (Bocharov, Sobol et al. 2004).
Structure of the L12 dimers off and on the ribosome
In 2004, three groups have simultaneously reported the structure and
dynamics of L12 dimers in solution based on heteronuclear NMR spectroscopy (Bocharov, Sobol et al. 2004; Christodoulou, Larsson et al. 2004;
Mulder, Bouakaz et al. 2004). In all cases, the CTDs were seen in their usual
globular form matching the crystal structures (Leijonmarck and Liljas 1987;
Wahl, Bourenkov et al. 2000). The NTDs were seen entangled in a four helix
bundle, in good agreement with the earlier NMR based observation
(Bocharov, Gudkov et al. 1998) Most interestingly, both the hinges were
flexible and unstructured, entirely different from the crystal structure of the
incomplete tetramer of L12 (Wahl, Bourenkov et al. 2000).
In contrast to L12 dimers in solution, clear signals only from the Cterminal domain of L12 were seen when NMR was done with intact 70S
ribosome and 50S subunit from E. coli. Somewhat weaker signals from the
hinge residues were also seen while no signal from the N-terminus could be
obtained. Thus it was reconfirmed that L12 CTDs are dynamic even on the
ribosome and the NTDs are tightly bound to the ribosome. Comparing the
17
peak intensities of the residues from the CTD and the hinge region, the number of the extended L12 molecules per ribosome was quantified (Mulder,
Bouakaz et al. 2004). Two out of four L12 monomers were counted as flexible on the ribosome. Based on that, Mulder et al. proposed two models of
L12 dimers on the ribosomes (Figure 5). In one model, both L12 dimers
were arranged as the typical antiparallel dimer (Chandra Sanyal and Liljas
2000) with one L12 molecule extended and the other in a compact shape. In
the other model, one dimer had both the L12 molecules compact as in the
crystal structure (Wahl, Bourenkov et al. 2000) while the second dimer had
both the molecules in extended shape (Mulder, Bouakaz et al. 2004). Further
experiments will be needed to conclude which model is more correct and if
both the models can exist under special conditions.
(a)
(b)
Figure 5: Two models of L12 dimers bound on the ribosome. (a) Both L12 dimers
have one extended and one retracted CTD. (b) One L12 dimer has both CTDs extended, while the other dimer has both CTDs retracted, adopted from (Mulder,
Bouakaz et al. 2004).
Stoichiometry of the L12 proteins on ribosome
The numbers of L12 molecules on the ribosome varies in different
bacteria. In E. coli they are four in number, which exist as two dimers
(Subramanian 1975) forming the pentameric L8 complex together with L10.
In contrast, in thermophiles and archaebacteria, they are six in number forming three dimers on the ribosome (Diaconu, Kothe et al. 2005; Ilag, Videler
et al. 2005; Maki, Hashimoto et al. 2007). The length and sequence of L10
α8 helix determines the number of L12 molecules per ribosome. In recent
studies with mass-spectroscopy, Gordiyenko et al. have observed that the
stoichiometry of the L12 proteins in the stalk complexes (pentametic or heptameric L8 complexes) changes in different growth stages of archaebacteria
(Gordiyenko, Videler et al. 2010). Putting the evidence together one can
conclude that there exists a strong correlation between the number of L12
dimers on the ribosome and the bacterial growth rate.
18
The binding site of the L12 dimers
It has been known for a long time that the L12 dimers are associated
with the ribosome by binding to the C-terminal end of the L10 protein. By
serial deletion from the C-terminal end of L10, Griaznova et al. suggested
that approximately the last 10 amino acids constitute the binding site of one
L12 dimer and the next 10 amino acids, although different in sequence constitute the binding site of the second L12 dimer on the E. coli ribosome
(Griaznova and Traut 2000). It was unclear how two identical L12 dimers
could bind to the unidentical sequences on the two binding sites. Much later,
when high-resolution crystal structure of L10-(L12-NTD)6 complex from T.
maritima was solved the earlier finding about the L12 dimer binding site was
confirmed. It turned out that all L12 dimers bind to the long α8 helix at the
C-terminal end of L10 (Figure 6). These sites although in a different bacteria, spanned approximately 10 amino acids each. It was proposed that hydrophobic interaction plays main role in the association of the L12 dimers on
L10. We have utilized this knowledge about the L12 dimer binding sites in
construction of homogeneous ribosomes containing only one L12 dimer.
Figure 6: Model of L10 and L12 complex, shows the binding of L12 NTDs on α8
helix of L10, and also the binding of two full length L12 molecules with one extended hinge and other with compact (helical) hinge. The image was created in PyMol from the coordinates T.maritima L10 and L12 NTD complex and E. coli L12
dimer (PDB id 1ZAV and 1RQV) (Bocharov, Sobol et al. 2004; Diaconu, Kothe et
al. 2005).
19
The factor interaction site on L12
The sites involved in translation factor interaction are located on the
globular CTD of L12 molecules. Earlier mutational analysis identified some
key residues on L12 CTD (K65, V66, K70, R73 and K84), which affected
the functions of elongation factors EF-Tu and EF-G (Kothe, Wieden et al.
2004; Savelsbergh, Mohr et al. 2005). Many of these sites coincided with
those identified later by means of heteronuclear NMR (Helgstrand, Mandava
et al. 2007). Most interestingly, these residues were concentrated on two
adjoined helices at the L12 CTD, and were common for all four major translation factors IF2, EF-Tu, EF-G and RF3. From this observation, it was suggested that L12 CTDs interact with the translation factors in a general manner.
Other than these direct studies, the L12 CTDs were seen attached to
EF-G and IF2 in the cryo-EM images of the complexes (Agrawal, Penczek et
al. 1998; Allen, Zavialov et al. 2005). The recent high resolution crystal
structure of 70S with EF-G bound on it showed one L12 CTD interacting
directly with EF-G on the ribosome (Gao, Selmer et al. 2009). These observations point towards the functional significance of L12 CTD and translation
factor interaction.
Role of L12 in translation
Protein L12 plays a key role in protein synthesis, L12 protein on ribosome is needed for translation initiation (Kay, Sander et al. 1973; Huang,
Mandava et al. 2010), elongation (Brot, Yamasaki et al. 1972; Agrawal,
Penczek et al. 1998) and termination (Brot, Tate et al. 1974) of protein synthesis. During these steps L12 interacts with different GTPase factors (IF2,
EF-Tu, EF-G and RF3) via its conserved C-terminal domain (Helgstrand,
Mandava et al. 2007). It has been reported that both L12 dimers are required
for the maximal rate of protein synthesis and minimal error frequency of
protein synthesis (Pettersson and Kurland 1980) and in their absence the
binding of elongation factors to the ribosome as well as GTP hydrolysis by
the translation factors are severely impaired (Leijonmarck and Liljas 1987;
Mohr, Wintermeyer et al. 2002). Two mutations in E. coli L12 CTD (G74D
and E82K) led to decreased growth rate and increased error frequency in
vivo (Kirsebom, Amons et al. 1986). Ribosomes from two missense suppressor mutant (mutations in rplLgene) strains (UY134 and UY143) which have
altered L12 protein showed reduced translation efficiency in vitro (Kirsebom
and Isaksson 1985; Bilgin, Kirsebom et al. 1988). Ribosomes with a deletion
in the hinge region of L12 have shown reduced activity in the Poly U translation(Gudkov, Bubunenko et al. 1991), implying that the hinge region in
L12 protein is very important for the mobility of the conserved globular Cterminal domain.
20
Two best studied functions of L12 are GTPase activation and factor
recruitment. It has been shown that these proteins are necessary on the ribosomes for the stimulation of the GTPase factors, particularly for EF-Tu and
EF-G (Hamel, Koka et al. 1972; Savelsbergh, Mohr et al. 2000; Mohr,
Wintermeyer et al. 2002) and such activation is hindered by the mutations on
the factor interaction site of L12 CTD (Kothe, Wieden et al. 2004;
Savelsbergh, Mohr et al. 2005). Also, free L12 protein at high concentrations
could stimulate GTPase activity of EF-G but not of EF-Tu. (Savelsbergh,
Mohr et al. 2000). Recently, using fast kinetics we have shown that unlike
EF-G, the L12 protein cannot stimulate the GTPase activity of IF2, irrespective of whether it is in free state or in a functional ribosomal complex
(Huang, Mandava et al. 2010).
It has also been suggested that highly mobile CTDs of L12 promote
the recruitment of the G-factors, EF-Tu in ternary complex and EF-G, onto
the ribosome (Diaconu, Kothe et al. 2005). Our experiments with IF2 show
that L12 is needed for the recognition of IF2 on 30S preinitiation complex
which plays a key role for rapid subunit association (Huang, Mandava et al.
2010). Thus, although different translation factors interact on a common site
on the L12 CTD, it seems that L12 has specialized functions with different
G-factors. The role of L12 proteins in RF3 function remains to be tested.
Translation
Translation is the process in which ribosomes translate the genetic information from mRNAs into proteins. In bacteria, translation occurs in four
steps, initiation, elongation, termination and recycling. All these steps are
guided by different protein factors in coordination with the ribosome. The
ribosome has been discussed at the beginning of this thesis. Below I describe
other components involved in translation.
Translation components
mRNA (Messenger RNA)
The messenger RNA (mRNA) carries the genetic information from
DNA to the ribosome in three base long genetic codes, also called codons.
In bacteria, mRNAs are polycistronic and contain multiple signal sequences
for translational initiation and termination (Laursen, Sorensen et al. 2005).
During translation, mRNAs interact with both tRNA and 30S ribosomal
subunit.
Bacterial mRNA exists mainly in two forms, leadered mRNA and leaderless mRNA. Canonical mRNAs start with a 5’ untranslated region (5’
UTR), which contains a ribosome binding site called Shine-Dalgarno se21
quence (SD sequence). The SD sequence is normally located six to nine
nucleotides upstream of the translational initiation codon AUG. It interacts
with the anti-SD sequence (aSD) located at the 3’ end of 16S rRNA of the
30S subunit (Shine and Dalgarno 1974). Leaderless mRNAs start either directly from the initiation codon or a few nucleotides upstream from the initiation codon. These mRNAs are recognized by the 30S subunit, initiation
factor IF2 and initiator tRNA (fMet-tRNAfMet) complex. It was shown that
initiation factors IF2 and IF3 regulate the translation efficiency of the leaderless mRNA (Grill, Moll et al. 2001; Moll, Grill et al. 2002). In E. coli, translation starts mainly with AUG initiation codon (90%) although GUG (8%)
and UUG (1%) codons can also act as the initiation codons in some cases
(Schneider, Stormo et al. 1986). The efficiency of translation initiation is
highly influenced by the strength of the SD sequence (determined by the
number of paired bases between SD and aSD sequence) as well as the spacing between the SD sequence and the initiation codon. The mRNA with a
strong SD sequence (7 - 8 base complementarity) initiates translation four
times more efficiently than that with a weak SD sequence (3 – 4 bases).
(Ringquist, Shinedling et al. 1992). The XR7 mRNAs used in our experiments contain a strong SD sequence (7 base complementarity) and six nucleotide long spacing between the SD-sequence and the initiation codon.
tRNA (transfer RNA)
The tRNAs are the adapter molecules that deliver the amino acids to
the ribosome during the elongation of the polypeptide chain. There are 41
different types of tRNAs, which occur in the cell in different frequencies and
carry 20 amino acids in a highly specific manner. The tRNAs carrying the
same amino acid are called tRNA isoacceptors, which are recognized by an
aminoacyl tRNA synthetase (aaRS) that links the corresponding amino acid
to the universally conserved 3’ CCA end in an ATP dependent manner. This
reaction is called aminoacylation or charging of the tRNA (Arnez and Moras
1997; Ibba and Soll 2000). The accuracy of the translation is highly dependent on the correct matching of the amino acid to the cognate tRNA.
The tRNAs carry anticodons on a loop called the anticodon stem loop
(ASL). The anticodons, usually complementary to the codons let the tRNAs
bind specifically to the mRNA by Watson-Crick base pairing. In a cognate
codon – anticodon interaction, the first two positions always complement
each other. Compared to that near-cognate (one of first two bases do not
match) and non-cognate (more than one base mismatch) are much weaker.
The correct codon anticodon interaction is important for adding the correct
amino acid in the chain.
Among all tRNAs, the initiator tRNAs have some special features,
which are important for maintenance of the fidelity in translation initiation.
These tRNAs are charged with methionine in all the organisms, in prokaryotes the methionine is formylated. Initiator tRNA is the only tRNA which
22
can directly bind to the P site on the ribosome, three continuous G-C base
pairs at the bottom of the anticodon stem facilitates this binding (Varshney,
Lee et al. 1993; RajBhandary 1994).
Translation factors
Each step of the translation in bacteria is carried out with the help of
several protein factors. Initiation factors IF1, IF2 and IF3 are involved in the
initiation of protein synthesis to form the 70S initiation complex. Elongation
factors EF-Tu, EF-Ts and EF-G are needed for the elongation of protein
synthesis. Release factors RF1, RF2 and RF3 regulate the steps in termination of the translation. Finally, ribosome release factor (RRF) along with EFG catalyzes the recycling of the ribosomal subunits. The function of the
translation factors has been discussed in detail in a later section. Besides
these commonly known translation factors, there are some other factors too
such as EF-P, which is thought to be important for the formation of the first
peptide bond (Aoki, Xu et al. 2008) and lepA, which is involved in backtranslocation (Qin, Polacek et al. 2006).
IF2, EF-Tu, EF-G and RF-3 are the four GTPase proteins which are
essential for the rapid protein synthesis in bacteria (Zavialov and Ehrenberg
2003). These GTPases belong to a large class of G proteins which switch
between the active GTP and inactive GDP forms and participate in different
cellular functions (Bourne, Sanders et al. 1991). The low intrinsic GTPase
activity of these G proteins gets strongly stimulated upon binding to the
GAC on the ribosome ((Rodnina, Stark et al. 2000), for review see
(Ramakrishnan 2002)).
The steps in protein synthesis
Initiation
After splitting of the post termination complex, IF3 rapidly binds to
the 30S subunit and promotes the dissociation of deacylated tRNAs and
mRNA from it (Karimi, Pavlov et al. 1999; Peske, Rodnina et al. 2005). It
also blocks premature subunit association, this anti association property of
IF3 increases in the presence of IF1 (Grunberg-Manago, Dessen et al. 1975).
The initiation step starts by the binding of a new mRNA to the 30S subunit
(Antoun, Pavlov et al. 2006a), where SD - aSD interaction play a major role
(Shine and Dalgarno 1974; Yusupova, Yusupov et al. 2001). Next, with the
help of the initiation factors, mRNA places the initiation codon AUG into
the P-site (La Teana, Gualerzi et al. 1995). Then the initiator tRNA (fMettRNAfMet) binds to the AUG codon with the help of IF2. Whether or not the
initiator tRNA binds to the ribosome in complex with IF2·GTP has been
debated for a long time (Gualerzi and Pon 1990; Wu and RajBhandary
1997). Recent kinetic studies have resolved the issue confirming that
23
IF2·GTP helps in recruitment of the initiator tRNA, but doesn’t bring it to
the ribosome in a complex (Milon, Carotti et al. 2010). The speed of association of the initiator tRNA to the 30S subunit as well as the accuracy in initiation increases significantly in the presence of all three initiation factors
(Antoun, Pavlov et al. 2006b). The whole complex with mRNA, initiator
tRNA and three initiation factors is called the 30S preinitiation complex.
Association of 50S subunit to 30S preinitiation complex leads to the
formation of 70S initiation complex. IF2 in GTP form is needed for fast association of the subunits (Grigoriadou, Marzi et al. 2007a). We have recently
shown that L12 proteins on the 50S subunit recognize IF2·GTP on the 30S
preinitiation complex, a reaction that leads to rapid subunit association
(Huang, Mandava et al. 2010).
There are two different opinions about the timing and role GTP hydrolysis during translation initiation. According to Antoun et al., IF2 in GTP
form is needed for fast subunit association, and GTP hydrolysis is needed for
the dissociation of IF2·GDP from the 70S initiation complex (Antoun,
Pavlov et al. 2003). In contrary, Grigoriadou et al. claimed that GTP hydrolysis is required for stable subunit association (Grigoriadou, Marzi et al.
2007a). Our recent experiments with two IF2 G-domain mutants (V400G
and H448E), which are impaired in GTP hydrolysis (Luchin, Putzer et al.
1999), did not show any defect in the rate of subunit association when compared with wild type IF2 (Figure 16a). This result is in accordance with Antoun et al. and confirmed that GTP hydrolysis is not needed for subunit association. However, dipeptide formation was really slow, almost insignificant,
when 70S initiation complex was formed with these two mutant IF2s. Thus,
GTP hydrolysis is needed for the release of IF2 from 70S initiation complex.
During the formation of 70S initiation complex all three initiation factors
dissociate in various sequences getting it ready for the elongation step.
Elongation
Elongation of protein synthesis starts with an empty A site and fMettRNAfMet in the P site of the 70S initiation complex. The two crucial steps in
elongation are mRNA decoding and peptide bond formation. Elongation
factors EF-Tu, EF-Ts and EF-G are the three main elongation factors which
facilitate this step.
EF-Tu brings an aminoacyl-tRNA (aa-tRNA) to the ribosomal A site
in a ternary complex with GTP (Stark, Rodnina et al. 1997). An initial selection governed by correct codon anticodon pairing (decoding) stabilizes the
aa-tRNA on the ribosome and triggers GTP hydrolysis by EF-Tu (for review
see (Ramakrishnan 2002)). After EF-Tu·GDP leaves the complex, the aatRNA repositions itself - a process called tRNA accommodation. This step is
followed by rapid peptide bond formation between the amino acids on the
tRNAs at the P site and A site. The peptidyl tranferase reaction is catalyzed
24
mainly by rRNA, the domain V of 23S rRNA being identified as the active
site for this reaction.
After peptidyl transfer, the elongation step continues with tRNA translocation, a step driven by EF-G, when both deacylated and peptidyl tRNAs
move on the mRNA by one codon. In the beginning of translocation the
aminoacyl end of the tRNAs move with 50S subunit, thereby creating E/P
and P/A hybrid states of tRNAs. The ribosome also goes in the ratcheted
stage, with two subunits rotated with respect to each other. In the next step,
EF-G·GTP catalyzes the movement of the ASL end of the tRNAs attached to
the mRNA with respect to 30S (Moazed and Noller 1989). This leads to the
post translocation state of the ribosome with an empty A site and places the
peptidyl and deacylated tRNAs in the classical P/P and E/E state respectively
for the entry of the next aa-tRNA. GTP plays a vital role in EF-G function.
However, whether GTP hydrolysis is essential for tRNA translocation or not
remains an open question (Rodnina, Savelsbergh et al. 1997; Zavialov,
Hauryliuk et al. 2005).
Elongation factor EF-Ts, a guanine nucleotide exchange factor (GEF)
exchanges the GDP on EF-Tu to GTP, which in turn forms a ternary complex with aa-tRNA. The EF-Tu ternary complex reinitiates another round of
elongation. The elongation step repeats until a stop codon on the mRNA
reaches the A site of the ribosome.
Termination
In bacteria, there are three stop codons UAA, UAG and UGA, which
are recognized by two class-I release factors RF1 and RF2 in a semispecific
fashion. Primarily RF1 recognizes UAG and RF2 recognizes UGA codon,
while both factors recognize UAA codon (Kisselev and Buckingham 2000).
These two release factors mimic the aa-tRNA in structure extending from the
DC on the 30S subunit to the PTC on the 50S subunit. RF1 or RF2 binds to
the A site and promotes the hydrolysis and release of nascent peptide from
the peptidyl tRNA in the P site. The GGQ (Gly-Gly-Gln) motif of the class I
release factors interacts with the PTC on the 50S subunit (Zavialov, Mora et
al. 2002). Similarly SPF motifs have been identified on these for recognition
of the stop codons (Ito, Uno et al. 2000). Once the peptide is released, the
class II release factor RF3 prompts the release of RF1 and RF2 from the Asite. RF3 enters the ribosome with GDP and rapidly exchanges it to GTP
(Zavialov, Buckingham et al. 2001). GTP binding on RF3 induces conformational changes on the ribosome which break the interaction of class I release
factors with the ribosome (Gao, Zhou et al. 2007). GTP hydrolysis on RF3
triggers its release from the ribosome (Zavialov, Buckingham et al. 2001).
Recycling
Recycling begins with the post termination complex with a deacylated
tRNA in the P-site. Ribosome recycling factor (RRF) and EF-G modulates
25
the recycling step through many cycles of GTP hydrolysis (Pavlov, Antoun
et al. 2008). The main event in recycling is the splitting of the ribosomal
subunits, which is followed by the binding of the anti-association factor IF3
on the 30S subunit (Karimi, Pavlov et al. 1999; Hirokawa, Nijman et al.
2005).
The main steps of translation are summarized in the following scheme.
Figure 7: Overview of translation cycle, showing initiation, elongation, termination,
and recycling.
26
Aim
The aim of this thesis is to determine the role of L12 in translation, both
from functional and structural angles. In this thesis the following topics are
discussed
1. Ribosomal stalk protein L12 has been implicated as a key player in the
interaction of the translation factors; in Paper-I, with the help of NMR,
we have mapped the interaction sites of IF2, EF-Tu, EF-G and RF-3 on
the CTD of L12.
2. The role of L12 with respect to the function of EF-Tu and EF-G has
been explored by other groups. In Paper-II, with fast kinetic experiments, we have discussed the role of L12 in translation initiation.
3. The numbers of L12 dimers on the ribosome vary in different bacteria
(two to three); in Paper-III we have shown why more than one L12 dimer is required on the bacterial ribosome.
4. Technological application with the stalk protein L12; in Paper-IV we
have described a single step method for purification of active ribosomes
carrying (His)6-tag on the CTD of L12.
Interaction of L12 CTD with translational G-factors
The ribosomal stalk protein L12 is thought to be involved in translation initiation (Kay, Sander et al. 1973), elongation (Brot, Yamasaki et al.
1972) and termination (Brot, Tate et al. 1974) of protein synthesis by interacting with different GTPase factors. Although not fully-evidenced, L12
proteins are thought to play important role in activation of translational
GTPases as well as in their recruitment (Savelsbergh, Mohr et al. 2000;
Mohr, Wintermeyer et al. 2002; Diaconu, Kothe et al. 2005). Using heteronuclear NMR spectroscopy we have shown that L12 directly binds to the
factors IF2, EF-Tu, EF-G and RF-3 and we have mapped the sites on L12
which are involved in these interactions.
Proteins with high purity and activity were used to investigate the
interaction site between L12 and translation factors IF2, EF-Tu, EF-G and
RF3. Each unlabelled translation factor was titrated separately into a solution
of 15N- labeled L12, and the 1H-15N heteronuclear single-quantum coherence
(HSQC) spectrum of the L12 residues were noted at each titration point. The
27
back bone amide group of each residue in 15N- labeled L12 protein corresponds to a single cross-peak in the 1H-15N HSQC spectrum, so residue specific information can be obtained from the chemical shift and line width of
the cross-peaks. By monitoring changes in the chemical shifts and line
widths of the cross peaks in the HSQC spectrum of L12 by addition of the
translation factors, the interaction site of individual translation factor on L12
was identified.
Upon factor binding, NMR signals from the CTD residues of L12
broadened significantly while the signals from the NTD of L12 remained
unchanged confirming the fact that the factors interact with the CTD and not
with the NTD of the L12 proteins. In addition, peak shift was seen in some
specific residues on the CTD, surprisingly common for all four factors. The
following figure (Figure 8) shows the chemical shift of the residues from the
CTD of L12 upon IF2 (K70 and L80) and EF-Tu (V66 and E82) binding.
Right side panel shows cross peaks from two residues (M13 and K108)
which did not change upon factor binding.
(a)
(b)
Figure 8: Close-up views of selected cross-peaks in superimposed 1H-15N HSQC
spectra of the L12 dimer, obtained at different equivalents (molar ratios), x, of added
factor with respect to L12 dimer. Two representative cross-peaks exhibiting significant chemical shift changes are shown for each titration series (left-hand and middle
panels), together with two representative cross-peaks, M17 and K108, that are unaffected by factor addition (right-hand panel). (a) Addition of IF2, showing residues
K70 (left) and L80, (middle); x = 0, 0.4, 0.8, 1.2, 1.6. (b) Addition of EF-Tu, showing residues V66 (left) and E82 (middle); x = 0, 0.1, 0.2, 0.3, 0.4. The cross-peak
contours are colored according to the increasing molar ratio of the added factor in
the order yellow, red, brown, black.
Thus our data indicated that all the four major GTPase factors (IF2,
EF-Tu, EF-G and RF3) interact essentially with the same site on the conserved CTD of the L12 protein. This region includes three strictly conserved
28
residues K70, L80 and E82 and some highly conserved residues, including
V66, A67, V68 and G79 placed on two adjacent α-helices on the L12-CTD
(Figure 9). Based on this result, we have concluded that CTD of L12 directly
interacts with translational G-factors (IF2, EF-Tu, EF-G and RF3) following
a general mechanism.
Figure 9: Binding surfaces of (a) IF2, (b) EF-Tu, (c) EF-G, and (d) RF3 on the CTD
of L12. The residues forming the factor-binding surface are colored in red and labeled with residue number. The figure is adopted from (Helgstrand, Mandava et al.
2007).
Role of L12 in translation initiation
L12 interaction with the elongation factors EF-Tu and EF-G have
been studied in the last decade in some detail (Savelsbergh, Mohr et al.
2000; Mohr, Wintermeyer et al. 2002). However, how the L12 proteins influence the function of the other two translational GTPases IF2 and RF3 is
quite unknown. In Paper-II we have investigated the role of L12 in IF2 mediated initiation of bacterial protein synthesis.
During protein synthesis IF2 in GTP form plays important role in
subunit association (Antoun, Pavlov et al. 2003). To study the role of L12
protein in IF2 mediated initiation steps, we have compared L12 depleted 50S
subunits (Hamel, Koka et al. 1972; Mohr, Wintermeyer et al. 2002) with
native 50S subunits in different steps of initiation using an in vitro translation system with all translation components purified from E. coli. As a control, we have used L12 reconstituted 50S in the similar reactions.
Subunit association reaction, monitored by Rayleigh light scattering in
a stopped flow instrument (Antoun, Pavlov et al. 2004), was performed simultaneously with Pi release or IF2 release, using parallel detectors
equipped with suitable cut-off filters. The depletion of L12 from 50S did not
affect the association of naked subunits, indicating that L12 proteins do not
constitute a structural component of 50S subunit that is needed for its association to 30S subunit. The addition of mRNA, initiator tRNA, and initiation
factors IF1 and IF3 to 30S other than IF2 one at a time or in different combination influenced the rate of subunit association, but did not show any difference between the native and L12 depleted 50S. From these observations
we concluded that L12 proteins do not have any functional or structural interactions with mRNA, initiator tRNA, IF1 and IF3. The presence of IF2-GTP
29
on 30S preinitiation complex resulted in a fast rate of subunit association, ka
= 130±10 µM-1s-1 for native 50S. Upon L12 depletion from the 50S subunit,
this has decreased ~40 -fold giving rise to a rate constant of about 3.5 µM-1s1
, this defect can be restored just with the reconstitution of L12 on to 50S
subunit (Figure 10).
(a)
(b)
Figure 10: Subunit association with different 50S subunits. The time course of association of normal (black trace), L12-depleted (red trace) and L12 reconstituted
(green trace) 50S subunit (0.5 μM), with (a) naked 30S subunit (0.5 μM), (b) 30S
preIC with IF2 but without IF3, was followed by the increase in light scattering at
430 nm. The blue traces in (b) reflect association of the subunits in the absence of
IF2 in otherwise identical conditions.
L12 protein by itself could not activate GTP hydrolysis of IF2 even at
very high concentrations. However, as in the case of EF-Tu, the presence of
L12 on naked 70S was essential for ribosome mediated GTPase activation of
IF2. In this paper we show that L12 depleted 70S is primarily defective in
IF2 binding, which in turn affects the subsequent steps i.e. ribosome stimulated GTP hydrolysis and Pi release even in translation uncoupled reactions.
The association of 30S preinitiation complex with 50S is dependent
on the presence of GTP on IF2. When GTP is removed or replaced with
GDP, the rate of association with native 50S subunits decreases significantly. However, L12 depleted 50S core could not differentiate between GTP or
GDP or no nucleotides present with IF2. Moreover, the slow rate of subunit
association with L12 depleted 50S subunit matched closely with that obtained without IF2. These results demonstrate that rapid subunit association
depends upon the specific interaction between the L12 protein on the 50S
subunit and IF2·GTP on the 30S subunit.
We have also measured the individual rates of GTP hydrolysis, Pi release and IF2 release with native and L12 depleted 50S subunits. In all these
measurements the apparent (or observed) rates for L12 depleted samples
were about 40 times slower than the native ones. However when actual rates
of GTP hydrolysis, Pi release and IF2 release were estimated taking into
account the average time spent in the steps before, there was almost no difference between native and L12 depleted 50S reactions (Figure 11). Thus,
30
we conclude that L12 protein is not a GTPase activating protein (GAP) for
IF2 (Huang, Mandava et al. 2010), but is necessary for the recognition and
interaction of IF2-GTP on 30S, that promotes rapid subunit association. In
this respect L12-IF2 interaction is different from those between L12- EF-G
and EF-Tu as suggested in the earlier reports.
Figure 11: Kinetic scheme of the steps in translation initiation for native and L12depleted 50S subunits. kobs is the observed rate. taverage (average time spent in a
step)=1/kobs (this step)−1/kobs (previous step) and estimated rate kest=1/taverage. Symbols used for L12-depleted 50S subunit are k′obs, k′est and t′average, respectively; and
shown in red in the cartoon.
Why do bacteria need multiple copies of L12 protein on
ribosome?
The numbers of the L12 molecules present on the ribosome are different in different organisms. For example in E. coli they are four in number
and exist as two dimers (Subramanian 1975) but in thermophiles such as
Thermatoga maritima they are six and exist as three dimers on the ribosome
(Diaconu, Kothe et al. 2005). In all cases the L12 dimers bind on a C terminal α8 helix on the protein L10. It has been reported that the binding sites of
different L12 dimers are different in their amino acid sequence, the significance of which is unknown. In this report (Paper III) we describe the construction and characterization of an E. coli strain JE105, which produces
31
ribosomes with single L12 dimer. The strain was created by deletion of 30
nucleotides from the 3’ end of the rplJ gene from the chromosome of the
wild type strain MG1655 and substituted with an ampicillin resistance cassette (Amp) by means of lambda red recombineering (Figure 12) (Yu, Ellis et
al. 2000; Costantino and Court 2003).
Figure 12: The chromosomal replacement of the last 30 nucleotides of rplJ gene
with Amp gene in wild type MG1655 and the resulting JE105 strain (left side), cartoon of L8 complex (right side) from both MG1655 and JE105 strains.
This strain named JE105 holds only one L12 dimer on its ribosomes
as checked by 2D gel and quantitative Western blot. It has been reported that
ribosomes with single L12 dimer are active in poly U dependent poly Phe
synthesis and EF-G dependent GTP hydrolysis (Griaznova and Traut 2000).
In that study a plasmid based construct of L12 and truncated L10 were used,
which were reconstituted on L8 depleted ribosomal cores. The in vivo situation for the single L12 dimer ribosomes was missing in that study, and the
depletion and reconstitution may lead to heterogeneity in the ribosome population. JE105 strain showed significant growth defect in LB media at 37oC.
Further, the single L12 dimer ribosomes were also found defective in synthesis of a full-length protein in a cell-free translation system. These ribosomes were also characterized with fast kinetics assays.
In our previous work we have reported that L12 protein on 50S subunit is not essential for naked subunit association (Huang, Mandava et al.
2010). Similar to that, the single L12 dimer 50S subunits produced the same
rate of association of the naked subunits as the native ones. However, during
proper subunit association in presence of mRNA, fMet-tRNAfMet and initiation factors IF1 and IF2, single L12 dimer 50S subunits showed two to three
fold lower activity than double L12 dimer native 50S (Figure 13a). The difference between the two ribosomes remained unchanged even at high concentrations of 50S subunit. We have also observed an apparent slower rate of
Pi release after subunit association. However, when the actual rate of Pi release was estimated by ‘average time’ analysis (Huang, Mandava et al. 2010)
it became clear that the defect in Pi release was a consequence of the defective subunit association (Table I).
32
Table I: Ratio of various translation steps t (average time spent in a step) between
the native and single L12 dimer ribosomes (last column of the table). kobs is the
observed rate.
In addition we have also checked the dipeptide (ML) and tripeptide
(MLL) formation starting from initiation complex to characterize the single
L12 dimer ribosomes during translation elongation. The rate of dipeptide
formation is similar for both single L12 dimer and normal ribosomes which
is 50 ± 10 s-1. In tripeptide formation starting from initiation complex, where
sequential steps occur (peptide bond formation between the first two amino
acids, translocation of the A site tRNA to P site, and second peptide bond
formation between second and third amino acid), the rate of tripeptide formation is slightly slower for single L12 dimer ribosomes (3.3 s-1) when
compared to normal ribosome’s (5 s-1) (Figure 13 b). As we know that the
single L12 dimer ribosomes are not defective in peptide bond formation, the
defect that we see in tripeptide formation can be from EF-G binding to post
translocation complex or from translocation of the tRNAs.
From the above experiments we can conclude that single L12 dimer
ribosomes are defective in IF2 mediated translation initiation and EF-G mediated elongation.
33
(a)
(b)
Figure 13: (a) Subunit association with different 50S subunits. The time course of
association of normal MRE600 (black trace), single L12 dimer (red trace) and wild
type MG1655 (blue trace) 50S subunit (0.5 μM), with 30S preIC was followed by
the increase in light scattering at 430 nm. Inset shows the same experiment with long
time scale (b) Tripeptide formation experiment starting from initiation complex.
Color codes are same as in the subunit association figure.
Technological application with stalk protein L12
Purification of active ribosomes is the bottle neck step in all in vitro
translation experiments. The conventional method of ribosome purification
is quite expensive in terms of effort, time and cost for chemicals, this method
involves several steps of ultracentrifugation (Antoun, Pavlov et al. 2004), so
a simple and high-throughput method for purification of ribosomes is in
demand.
There have been several attempts to purify the ribosomes with affinity
tags (Gan, Kitakawa et al. 2002; Leonov, Sergiev et al. 2003; Youngman and
Green 2005). RNA fused affinity tag purification was mainly designed for
the purification of E. coli ribosomes (Leonov, Sergiev et al. 2003;
Youngman and Green 2005). These methods involved the plasmid based
over expression of ribosomal components fused with affinity tags and the
success of the method depends upon the over expression and the assembly of
those components.
To circumvent these hurdles we have created a novel E. coli strain
JE28, in which a (His)6-tag has been inserted in frame at the C-terminus of
the r-protein L12 directly on the chromosome using genetic recombineering
method (Ellis, Yu et al. 2001; Costantino and Court 2003) (Figure 14). Since
L12 is present in four copies on the large subunit of E. coli ribosome, the
ribosomes of JE28 are homogeneously tetra-(His)6-tagged. The insertion of
(His)6-tag did not affect the growth of the bacteria.
Further we have developed a simple, fast, single-step affinity purification method to purify the tetra-(His)6-tagged ribosomes from JE28 strain.
34
Affinity purified ribosomes contained only 70S either due to a much shorter
preparation time (1–2 hrs) or due to high Mg2+ concentration during purification. The (His)6-tagged ribosomes were found to be equally active as the
native ones in functional assays such as dipeptide formation and full-length
protein synthesis in a cell free coupled transcription-translation system.
Figure 14: Insertion of linear DNA encoding (His)6, stop codon and Kan gene at the
3’ end of rplL gene in E. coli chromosome by λ Red recombineering.
(a)
(b)
Figure 15: (a) 20-50% sucrose density gradient separation profiles for conventionally purified ribosomes (black line), and affinity column purified ribosomes (blue
line). (b) In vitro dipeptide formation experiment with 70S from MG1655 (black
trace), JE28 conventional purification (red trace) and JE28 column purification
(green trace).
We have also purified 30S and 50S subunits through this method.
These tetra (His)6-tagged ribosomes can be used for the purification of various functional ribosomal complexes for biochemical and structural studies. It
is also useful for the purification of ribosomes carrying mutations on rRNA
or on r-proteins.
35
Current work and future plan
Kinetic studies on translation initiation
We have recently reported the role of ribosomal stalk protein L12 in
IF2 mediated initiation of bacterial translation (Huang, Mandava et al. 2010).
With fast kinetic measurements we have shown that L12 proteins are needed
for the recognition and recruitment of IF2·GTP bound on 30S preinitiation
complex, but not for the activation of its GTPase function. So, unlike in the
cases of EF-Tu and EF-G, L12 is not a GAP (GTPase activating protein) for
IF2. However, this study opened a general question about the sequence of
events in translation initiation. The kinetic steps of translation initiation have
been reported in recent articles (Antoun, Pavlov et al. 2006a; Fabbretti, Pon
et al. 2007; Grigoriadou, Marzi et al. 2007a); there are some contradictory
views for some of the steps in translation initiation. One among them is the
IF2 mediated GTP hydrolysis during the subunit association whether it is
required or not required for the rapid subunit association, if not then what is
the role of GTP hydrolysis at the initiation step?
There are two different opinions about the timing and role of GTP
hydrolysis during the translation initiation. According to Antoun et al., IF2
in GTP form is needed for fast subunit association, and GTP hydrolysis is
needed for the dissociation of IF2·GDP from the 70S initiation complex
(Antoun, Pavlov et al. 2003). In contrary, Grigoriadou et al. from Barry
Cooperman’s group reported that GTP hydrolysis is needed for stable subunit association (Grigoriadou, Marzi et al. 2007a). We thought to check both
models by doing parallel kinetic experiments for different steps in initiation.
To see whether GTP hydrolysis is necessary for subunit association or not,
we have used GDPNP (non hydrolysable analogue of GTP) in subunit association reactions followed by light scattering in stopped flow instrument.
When added in excess, more or less same rate of subunit association was
obtained with GDPNP as with GTP. We have also checked the subunit association by titrating GDP, the rate of subunit association is very low and remained unchanged even with 1mM GDP (Figure 16b). From these results
and the results from Paper-II suggests that GTP hydrolysis is not needed for
subunit association but IF2 in GTP form is needed for rapid subunit association.
To confirm this finding we have used two IF2 G-domain mutants
(V400G & H448E) which are impaired in GTPase activity (Luchin, Putzer et
36
al. 1999) in subunit association and in phosphate release experiments. When
compared with wild type IF2, these two IF2 mutants did not show any defect
in subunit association (Figure 16a), but because of their disability in GTPase
activity there is no Pi released after 70S formation. These mutants again
confirmed that subunit association can take place without any hydrolysis of
GTP.
(a)
(b)
Light Scattering (a.u.)
0.25
0.20
0.15
0.10
WT
V400G
H448E
without IF2
0.05
0.00
0.0
0.2
0.4
0.6
0.8
1.0
Time (sec)
Figure 16: (a). Association of 50S subunits with 30S pre initiation complex containing either WT (black trace) or two GTPase impaired mutant IF2s (V440G and
H448E) (red and blue traces respectively) or without IF2 (green trace) followed by
light scattering in stopped flow instrument. All variants of IF2 produced comparable
rates of subunit association. (b). The rate constants for the subunit association reaction as in (a) with different guanine nucleotides as noted below the bars. The inset
shows the rates against increasing concentration of GTP (black trace), GDP (blue
trace) and GDPNP (green trace).
To investigate the role of GTP hydrolysis we have also checked the
release of IF2 after subunit association in presence of GTP and GDPNP. In
this experiment we have used a fluorescent labeled (rhodamine) IF2 and
followed the decrease of fluorescence after the subunit association when IF2
is released. In the presence of GTP we can see the decrease of fluorescence
indicating IF2 release, whereas with GDPNP we did not see any decrease in
fluorescence. Subunits can associate in presence of GDPNP and GTP, but
the peptide bond formation is very slow with GDPNP when compared to
GTP, indicating that IF2 was able to promote the subunit association in presence of GDPNP but it remained bound to 70S ribosome, thereby blocking
the subsequent binding of EF-Tu ternary complex and peptide bond formation (Antoun, Pavlov et al. 2003). When we have checked the dipeptide formation, both the IF2 mutants showed a slower rate of peptide bond formation when compared to wild type IF2. This indicates the presence of mutant
IF2s on 70S ribosome after subunit association, because of their inactivity in
GTP hydrolysis. Thus the presence of IF2 blocks association of the EF-Tu
ternary complex and peptide bond formation. Based on the above experiments we can conclude that GTP hydrolysis is needed for IF2 release, but
37
not for subunit association. To further confirm this argument, we would like
to check the release of the GTPase impaired IF2 mutants using the rhodamine labeled mutant IF2s. The prediction is that the mutant IF2s will be retained on the ribosome.
We have also checked the maximal rate for each step in translation initiation by titrating up the 50S concentration, further from these rates we
have estimated the time needed for each individual step (Figure 17).
Figure 17: Kinetic scheme of the steps in translation initiation, maximal rate for
individual steps are shown below the arrow of each step and the estimated times are
shown above the arrow.
Future plans
Identification of residues on L12 involved in the interaction of
IF2 during subunit association
We have mapped earlier the IF2 interaction sites on L12 CTD by using heteronuclear NMR spectroscopy (Helgstrand, Mandava et al. 2007).
However, for technical reasons the mapping was done with purified L12
protein and not with L12 bound on the ribosomes. Now that we have found
that L12 plays an important role in IF2 recognition, it is important that we
can trace the identification sites on L12 as well as on IF2 using mutagenesis
approach. Based on the IF2 interaction sites mapped with NMR, we have
now created some mutations on L12 CTD. The next step will be to reconstitute these mutant L12s on ribosomal cores (L12 depleted 50S and 70S) and
the ribosomes with mutant L12 proteins will be tested primarily in subunit
38
association assay and then in GTP hydrolysis and Pi release, covering essentially all the steps in translation initiation.
Comparison of E. coli ribosomes with single, double and triple
L12 dimers
As we mentioned earlier in this report we have created an E. coli
strain (JE105) with single L12 dimer on its ribosomes. This strain has
showed slower growth rate. We have also created another E. coli strain
(CS111) where an additional L12 binding site has been added on the ribosome, which contains three L12 dimer binding sites and therefore expected
to bind three L12 dimers on its ribosomes. With the help of the above two
strains we want to know why the number of L12 proteins per ribosome varies from two to three in different bacterial species and what is the need of
multiple copies of L12 on the ribosome. We have already characterized the
single L12 dimer ribosomes at different steps of translation and found out
that they are defective in translation initiation and elongation when compared to the wild type ribosomes from E. coli strain MG1655. We will purify
the triple L12 dimer ribosomes from CS111 and characterize those for their
content as well as in different steps of translation.
39
Conclusions
From the above discussed works we can conclude that
1. L12 interacts with the translational GTPase factors (IF2, EF-Tu, EF-G
and RF3) involving a common region on its CTD. The residues on L12
CTD involved in the interactions are K70, L80, E82 (strictly conserved),
V66, A67, V68 and G79 (highly conserved). (Paper I).
2. Unlike for EF-G and EF-Tu, L12 is not a GAP for IF2. Instead, during
initiation L12 is needed for the recognition and recruitment of IF2 in
GTP form. (Paper II)
3. Ribosomes with single L12 dimer are defective in IF2 mediated subunit
association and EF-G mediated elongation. Multiple L12 dimers are
needed on the ribosome for efficient translation and fast growth of the
bacteria. (Paper III)
4. His-tagging on multi copy L12 protein provides a good system for single
step purification of active ribosomes. (Paper IV)
40
Summary in Swedish
Ribosomala stjälkprotein L12: struktur, funktion och
tillämpningar.
Bakgrund
Ribosomerna är stora ribonucleoprotein komplex med en approximativ massa på 2,5 MDA och en diameter på 200 - 250 Å. Ribosomerna finns i
cytoplasman, i eukaryoter finns de också på det endoplasmatiska nätverket.
Ribosomer översätter den genetiska information kodad i budbärar-RNA
(mRNA) till motsvarande sekvens av aminosyror och tillverkar därigenom
ett protein.
Ribosomen består av proteiner och ribosomala RNA (rRNA), delat i
två subenheter, som brukar kallas den stora och den lilla subenheten. Prokaryota (70S) och eukaryota (80S) ribosomer, som är mycket lika i organisation och struktur, skiljer sig dock i både protein och RNA-komponenter. Där
finns tre olika tRNA bindningsställen på ribosomen som heter A- (amino
acyl), P- (peptidyl) och E- (exit) stäl.
I bakterier sker proteinsyntesen i fyra steg, initiering, elongering, terminering och återvinning. Flera GTPase protein faktorer (IF2, EF-Tu, EF-G
och RF3) samverkar med den ribosomala stjälken under dessa steg. Stjälken
består av 2-3 dimerer av L12 och ett L10 protein vilket utgör nedre delen av
stjälken. I E. coli finns fyra kopior av L12 vilka existerar som en dimer av
dimerer på ribosomernas 50S-subenhet. I denna avhandling har vi rapporterat följande saker i detalj.
Samverkan mellan L12 och translationella G-faktorer
L12 proteiner är kända för att interagera med flera GTPase faktorer
under translationen. Vi har kartlagt interaktionsplatser för de fyra viktigaste
G-faktorerna (IF2, EF-Tu, EF-G & RF3) på L12 molekylen med heteronukleär NMR-spektroskopi. Våra data visar att alla dessa fyra GTPase faktorer I
huvudsak samverkar med samma plats på den konserverade C-terminala
domänen (CTD) i L12 proteinet. Denna region innehåller tre helt konserverade aminosyrarester K70, L80 och E82 samt några starkt konserverade res-
41
ter på två intilliggande α-helixar. Detta tyder på en generell mekanism för
samspelet mellan L12 och G-faktorer.
L12s roll i initieringen av proteinsyntesen
L12 är känd för att stimulera GTPase aktivering av EF-Tu och EF-G
under elongeringsfasen av proteinsyntesen. I denna rapport har vi klargjort
betydelsen av L12 i IF2 medierad initiering av proteinsyntesen genom att
jämföra 50S utan L12 med vanliga 50S-subenheter i association av subenheterna till 70S ribosomer, GTP hydrolys på IF2, frisläppning av oorganiskt
fosfat efter GTP hydroly samt frisläppning av IF2 från ribosomen. 50S subenheter utan L12 associerar mycket långsammare (~ 40 gånger) med 30S
endast när IF2-GTP finns på 30S innan associeringen. Detta tyder på att en
interaktion mellan L12 protein på 50S och IF2-GTP på 30S behövs för snabb
associering av subenheterna. Avlägsnande av L12 från 50S påverkade inte
de efterföljande stegen. Således har vi kommit fram till att L12 inte är ett
GAP (GTPase aktiverande protein) för IF2.
Varför behöver bakterierna flera kopior av L12 protein på
ribosomen?
Antalet L12 molekyler på ribosomen är olika i olika organismer. L12
binder till ribosomen via C-terminal α8 helix på L10 proteinet. För att undersöka behovet av fler kopior av L12 på ribosomen har vi skapat en genetiskt
modifierad E. coli-stam (JE105), som producerar ribosomer med endast en
L12 dimer. Denna stam uppvisade en betydande tillväxt defekt. In vitro var
dessa ribosomer 2-3 gånger långsammare i initiering, men visade inga fel på
EF-Tu och EF-G medierade steg i peptidelongering. Dessa resultat indikerar
en viktig roll för L12 i initieringen av proteinsyntesen.
Teknisk tillämpning av stjälkproteinet L12
I denna rapport har vi också diskuterat tekniska tillämpningar av flera
kopior av L12 protein. Vi har skapat en genetiskt modifierad E. coli-stam
(JE28) som har en (His)6-tag I slutet av L12 proteinet. Eftersom L12 protein
finns i fyra kopior på ribosomen, innehåller ribosomer som produceras i
denna stam fyra (His)6-tags. Vi har utvecklat en metod för att rena dessa
ribosomer från JE28. Dessa ribosomker var lika aktiva som konventionellt
renade ribosomer från vildtyp E. coli stammar, i in vitro dipeptid bildning
och i cell fri proteinsyntes
42
Acknowledgement
I would like to acknowledge and extend my heart felt gratitude to my supervisor Suparna Sanyal for accepting me as her first PhD student. Suparna
you have been very supportive throughout this period. You are very inspirational and I have always admired your passion for science. Thank you very
much for being there whenever I needed your help in and out of the lab. My
co-supervisor Måns Ehrenberg, thank you Måns for your suggestions and
comments during the group meetings.
My co-authors specially Huang Chenhui, thank you Huang for teaching
me all complicated preps. You’re a nice friend and colleague. Josefine
thanks for generating very stable strains (JE28 and JE105). Chris you are a
good friend and thanks for correcting my thesis and I wish you good luck.
I would like to thank my collaborators Santanu Dasgupta and Mikael
Akke. Santanu thank you for the wonderful discussions during the coffee
time about the Science, Politics, Food… what not
I would like to thank my dearest former and present office members. Suzana you’re a wonderful friend, thanks for the green office room, some of the
plants are still good and Bruce Lee is still hanging on the wall with all the
birthday marks. Venkat garu thanks helping me in the lab during the initial
days. Christiane thanks for your friendship and for the Milk mouse chocolates. Kristin, Mikael and Ravi, thank you guys for the cake and chocolate
supplies. Puli, good luck for your licentiate. Mikael thanks for the Swedish
translations and especially for the Swedish summary in the thesis. Kristin
thanks for the western blots. Chian, thanks for helping me with the red recombineering. Ranjeet, thanks for the comments and corrections for my
thesis and for the spiritual discussions. Gürkan thanks for helping me with
the PyMol. Yanhong, good luck for your licentiate. Liang thanks for the L12
mutations and for the L10 plasmid. Antje good luck. Ray thanks for the
supply of components in the general box and for proof reading my thesis.
Deba da, thanks for the wonderful 'tonfisk’ curry; you’re a good researcher
and I had good time working with you, good luck with the science and life.
I would like to thank present and former members of Måns Ehrenberg
group, especially Lamine for teaching me the ribosome preparation. Misha,
thanks for your suggestions and for the help with stopped flow. Anneli, Doreen, Galyna, Harriet, Ikue, Jingji, Kaweng, and Magnus, I wish you all
good luck for your science. David thanks for showing me the execution of
the curve smoothing program. Thanks to all my friends whom I did not mention here.
Finally to family,, you are the best …!!!
43
References:
Agrawal, R. K., A. B. Heagle, et al. (1999). "EF-G-dependent GTP hydrolysis
induces translocation accompanied by large conformational changes in the 70S
ribosome." Nat Struct Biol 6(7): 643-647.
Agrawal, R. K., P. Penczek, et al. (1998). "Visualization of elongation factor G on
the Escherichia coli 70S ribosome: the mechanism of translocation." Proc Natl
Acad Sci U S A 95(11): 6134-6138.
Allen, G. S., A. Zavialov, et al. (2005). "The cryo-EM structure of a translation
initiation complex from Escherichia coli." Cell 121(5): 703-712.
Antoun, A., M. Y. Pavlov, et al. (2003). "The roles of initiation factor 2 and
guanosine triphosphate in initiation of protein synthesis." EMBO J 22(20):
5593-5601.
Antoun, A., M. Y. Pavlov, et al. (2006a). "How initiation factors tune the rate of
initiation of protein synthesis in bacteria." EMBO J 25(11): 2539-2550.
Antoun, A., M. Y. Pavlov, et al. (2006b). "How initiation factors maximize the
accuracy of tRNA selection in initiation of bacterial protein synthesis." Mol Cell
23(2): 183-193.
Antoun, A., M. Y. Pavlov, et al. (2004). "Ribosome formation from subunits studied
by stopped-flow and Rayleigh light scattering." Biol Proced Online 6: 35-54.
Aoki, H., J. Xu, et al. (2008). "Interactions of elongation factor EF-P with the
Escherichia coli ribosome." The FEBS journal 275(4): 671-681.
Arnez, J. G. and D. Moras (1997). "Structural and functional considerations of the
aminoacylation reaction." Trends in biochemical sciences 22(6): 211-216.
Beauclerk, A. A., E. Cundliffe, et al. (1984). "The binding site for ribosomal protein
complex L8 within 23 s ribosomal RNA of Escherichia coli." The Journal of
biological chemistry 259(10): 6559-6563.
Bilgin, N., L. A. Kirsebom, et al. (1988). "Mutations in ribosomal proteins L7/L12
perturb EF-G and EF-Tu functions." Biochimie 70(5): 611-618.
Bocharov, E. V., A. T. Gudkov, et al. (1998). "Conformational independence of Nand C-domains in ribosomal protein L7/L12 and in the complex with protein
L10." FEBS Lett 423(3): 347-350.
Bocharov, E. V., A. G. Sobol, et al. (2004). "From structure and dynamics of protein
L7/L12 to molecular switching in ribosome." J Biol Chem 279(17): 1769717706.
Bourne, H. R., D. A. Sanders, et al. (1991). "The GTPase superfamily: conserved
structure and molecular mechanism." Nature 349(6305): 117-127.
Brot, N., W. P. Tate, et al. (1974). "The requirement for ribosomal proteins L7 and
L12 in peptide-chain termination." Proc Natl Acad Sci U S A 71(1): 89-92.
Brot, N., E. Yamasaki, et al. (1972). "The properties of an E. coli ribosomal protein
required for the function of factor G." Arch Biochem Biophys 148(1): 148-155.
Chandra Sanyal, S. and A. Liljas (2000). "The end of the beginning: structural
studies of ribosomal proteins." Curr Opin Struct Biol 10(6): 633-636.
44
Christodoulou, J., G. Larsson, et al. (2004). "Heteronuclear NMR investigations of
dynamic regions of intact Escherichia coli ribosomes." Proc Natl Acad Sci U S
A 101(30): 10949-10954.
Costantino, N. and D. L. Court (2003). "Enhanced levels of lambda Red-mediated
recombinants in mismatch repair mutants." Proc Natl Acad Sci U S A 100(26):
15748-15753.
Dey, D., D. E. Bochkariov, et al. (1998). "Cross-linking of selected residues in the
N- and C-terminal domains of Escherichia coli protein L7/L12 to other
ribosomal proteins and the effect of elongation factor Tu." J Biol Chem 273(3):
1670-1676.
Diaconu, M., U. Kothe, et al. (2005). "Structural basis for the function of the
ribosomal L7/12 stalk in factor binding and GTPase activation." Cell 121(7):
991-1004.
Ellis, H. M., D. Yu, et al. (2001). "High efficiency mutagenesis, repair, and
engineering of chromosomal DNA using single-stranded oligonucleotides." Proc
Natl Acad Sci U S A 98(12): 6742-6746.
Fabbretti, A., C. L. Pon, et al. (2007). "The real-time path of translation factor IF3
onto and off the ribosome." Molecular cell 25(2): 285-296.
Gan, X., M. Kitakawa, et al. (2002). "Tag-mediated isolation of yeast mitochondrial
ribosome and mass spectrometric identification of its new components."
European journal of biochemistry / FEBS 269(21): 5203-5214.
Gao, H., Z. Zhou, et al. (2007). "RF3 induces ribosomal conformational changes
responsible for dissociation of class I release factors." Cell 129(5): 929-941.
Gao, Y. G., M. Selmer, et al. (2009). "The structure of the ribosome with elongation
factor G trapped in the posttranslocational state." Science 326(5953): 694-699.
Gordiyenko, Y., S. Deroo, et al. (2008). "Acetylation of L12 increases interactions
in the Escherichia coli ribosomal stalk complex." J Mol Biol 380(2): 404-414.
Gordiyenko, Y., H. Videler, et al. (2010). "Mass spectrometry defines the
stoichiometry of ribosomal stalk complexes across the phylogenetic tree." Mol
Cell Proteomics 9(8): 1774-1783.
Griaznova, O. and R. R. Traut (2000). "Deletion of C-terminal residues of
Escherichia coli ribosomal protein L10 causes the loss of binding of one L7/L12
dimer: ribosomes with one L7/L12 dimer are active." Biochemistry 39(14):
4075-4081.
Grigoriadou, C., S. Marzi, et al. (2007a). "A quantitative kinetic scheme for 70 S
translation initiation complex formation." J Mol Biol 373(3): 562-572.
Grill, S., I. Moll, et al. (2001). "Modulation of ribosomal recruitment to 5'-terminal
start codons by translation initiation factors IF2 and IF3." FEBS letters 495(3):
167-171.
Grunberg-Manago, M., P. Dessen, et al. (1975). "Light-scattering studies showing
the effect of initiation factors on the reversible dissociation of Escherichia coli
ribosomes." J Mol Biol 94(3): 461-478.
Gualerzi, C. O. and C. L. Pon (1990). "Initiation of mRNA translation in
prokaryotes." Biochemistry 29(25): 5881-5889.
Gudkov, A. T. (1997). "The L7/L12 ribosomal domain of the ribosome: structural
and functional studies." FEBS Lett 407(3): 253-256.
Gudkov, A. T., M. G. Bubunenko, et al. (1991). "Overexpression of L7/L12 protein
with mutations in its flexible region." Biochimie 73(11): 1387-1389.
Hamel, E., M. Koka, et al. (1972). "Requirement of an Escherichia coli 50 S
ribosomal protein component for effective interaction of the ribosome with T
and G factors and with guanosine triphosphate." J Biol Chem 247(3): 805-814.
45
Hardy, S. J. (1975). "The stoichiometry of the ribosomal proteins of Escherichia
coli." Mol Gen Genet 140(3): 253-274.
Helgstrand, M., C. S. Mandava, et al. (2007). "The ribosomal stalk binds to
translation factors IF2, EF-Tu, EF-G and RF3 via a conserved region of the L12
C-terminal domain." J Mol Biol 365(2): 468-479.
Hirokawa, G., R. M. Nijman, et al. (2005). "The role of ribosome recycling factor in
dissociation of 70S ribosomes into subunits." RNA 11(8): 1317-1328.
Huang, C., C. S. Mandava, et al. (2010). "The ribosomal stalk plays a key role in
IF2-mediated association of the ribosomal subunits." J Mol Biol 399(1): 145153.
Ibba, M. and D. Soll (2000). "Aminoacyl-tRNA synthesis." Annual review of
biochemistry 69: 617-650.
Ilag, L. L., H. Videler, et al. (2005). "Heptameric (L12)6/L10 rather than canonical
pentameric complexes are found by tandem MS of intact ribosomes from
thermophilic bacteria." Proc Natl Acad Sci U S A 102(23): 8192-8197.
Ito, K., M. Uno, et al. (2000). "A tripeptide 'anticodon' deciphers stop codons in
messenger RNA." Nature 403(6770): 680-684.
Karimi, R., M. Y. Pavlov, et al. (1999). "Novel roles for classical factors at the
interface between translation termination and initiation." Molecular cell 3(5):
601-609.
Kay, A., G. Sander, et al. (1973). "Effect of ribosomal protein L12 upon initiation
factor IF-2 activities." Biochem Biophys Res Commun 51(4): 979-986.
Kirsebom, L. A., R. Amons, et al. (1986). "Primary structures of mutationally
altered ribosomal protein L7/L12 and their effects on cellular growth and
translational accuracy." Eur J Biochem 156(3): 669-675.
Kirsebom, L. A. and L. A. Isaksson (1985). "Involvement of ribosomal protein
L7/L12 in control of translational accuracy." Proc Natl Acad Sci U S A 82(3):
717-721.
Kisselev, L. L. and R. H. Buckingham (2000). "Translational termination comes of
age." Trends in biochemical sciences 25(11): 561-566.
Kothe, U., H. J. Wieden, et al. (2004). "Interaction of helix D of elongation factor
Tu with helices 4 and 5 of protein L7/12 on the ribosome." Journal of molecular
biology 336(5): 1011-1021.
La Teana, A., C. O. Gualerzi, et al. (1995). "From stand-by to decoding site.
Adjustment of the mRNA on the 30S ribosomal subunit under the influence of
the initiation factors." RNA 1(8): 772-782.
Laursen, B. S., H. P. Sorensen, et al. (2005). "Initiation of protein synthesis in
bacteria." Microbiology and molecular biology reviews : MMBR 69(1): 101123.
Leijonmarck, M. and A. Liljas (1987). "Structure of the C-terminal domain of the
ribosomal protein L7/L12 from Escherichia coli at 1.7 A." J Mol Biol 195(3):
555-579.
Leonov, A. A., P. V. Sergiev, et al. (2003). "Affinity purification of ribosomes with
a lethal G2655C mutation in 23 S rRNA that affects the translocation." The
Journal of biological chemistry 278(28): 25664-25670.
Li, W., J. Sengupta, et al. (2006). "Functional conformations of the L11-ribosomal
RNA complex revealed by correlative analysis of cryo-EM and molecular
dynamics simulations." RNA 12(7): 1240-1253.
Liljas, A. and A. T. Gudkov (1987). "The structure and dynamics of ribosomal
protein L12." Biochimie 69(10): 1043-1047.
Luchin, S., H. Putzer, et al. (1999). "In vitro study of two dominant inhibitory
GTPase mutants of Escherichia coli translation initiation factor IF2. Direct
46
evidence that GTP hydrolysis is necessary for factor recycling." J Biol Chem
274(10): 6074-6079.
Maki, Y., T. Hashimoto, et al. (2007). "Three binding sites for stalk protein dimers
are generally present in ribosomes from archaeal organism." The Journal of
biological chemistry 282(45): 32827-32833.
Marintchev, A. and G. Wagner (2004). "Translation initiation: structures,
mechanisms and evolution." Quarterly reviews of biophysics 37(3-4): 197-284.
Milon, P., M. Carotti, et al. (2010). "The ribosome-bound initiation factor 2 recruits
initiator tRNA to the 30S initiation complex." EMBO Rep 11(4): 312-316.
Moazed, D. and H. F. Noller (1989). "Interaction of tRNA with 23S rRNA in the
ribosomal A, P, and E sites." Cell 57(4): 585-597.
Moens, P. D., M. C. Wahl, et al. (2005). "Oligomeric state and mode of selfassociation of Thermotoga maritima ribosomal stalk protein L12 in solution."
Biochemistry 44(9): 3298-3305.
Mohr, D., W. Wintermeyer, et al. (2002). "GTPase activation of elongation factors
Tu and G on the ribosome." Biochemistry 41(41): 12520-12528.
Moll, I., S. Grill, et al. (2002). "Leaderless mRNAs in bacteria: surprises in
ribosomal recruitment and translational control." Molecular microbiology 43(1):
239-246.
Mulder, F. A., L. Bouakaz, et al. (2004). "Conformation and dynamics of ribosomal
stalk protein L12 in solution and on the ribosome." Biochemistry 43(20): 59305936.
Nissen, P., J. Hansen, et al. (2000). "The structural basis of ribosome activity in
peptide bond synthesis." Science 289(5481): 920-930.
Ogle, J. M., D. E. Brodersen, et al. (2001). "Recognition of cognate transfer RNA by
the 30S ribosomal subunit." Science 292(5518): 897-902.
Olson, H. M., A. Sommer, et al. (1986). "Localization of two epitopes of protein
L7/L12 to both the body and stalk of the large ribosomal subunit. Immune
electron microscopy using monoclonal antibodies." J Biol Chem 261(15): 69246932.
Pavlov, M. Y., A. Antoun, et al. (2008). "Complementary roles of initiation factor 1
and ribosome recycling factor in 70S ribosome splitting." EMBO J 27(12):
1706-1717.
Peske, F., M. V. Rodnina, et al. (2005). "Sequence of steps in ribosome recycling as
defined by kinetic analysis." Molecular cell 18(4): 403-412.
Pettersson, I., S. J. Hardy, et al. (1976). "The ribosomal protein L8 is a complex
L7/L12 and L10." FEBS Lett 64(1): 135-138.
Pettersson, I. and C. G. Kurland (1980). "Ribosomal protein L7/L12 is required for
optimal translation." Proc Natl Acad Sci U S A 77(7): 4007-4010.
Pettersson, I. and A. Liljas (1979). "The stoichiometry and reconstitution of a stable
protein complex from Escherichia coli ribosomes." FEBS Lett 98(1): 139-144.
Qin, Y., N. Polacek, et al. (2006). "The highly conserved LepA is a ribosomal
elongation factor that back-translocates the ribosome." Cell 127(4): 721-733.
RajBhandary, U. L. (1994). "Initiator transfer RNAs." Journal of bacteriology
176(3): 547-552.
Ramakrishnan, V. (2002). "Ribosome structure and the mechanism of translation."
Cell 108(4): 557-572.
Rheinberger, H. J. and K. H. Nierhaus (1986). "Allosteric interactions between the
ribosomal transfer RNA-binding sites A and E." The Journal of biological
chemistry 261(20): 9133-9139.
47
Rheinberger, H. J., H. Sternbach, et al. (1981). "Three tRNA binding sites on
Escherichia coli ribosomes." Proceedings of the National Academy of Sciences
of the United States of America 78(9): 5310-5314.
Ringquist, S., S. Shinedling, et al. (1992). "Translation initiation in Escherichia coli:
sequences within the ribosome-binding site." Molecular microbiology 6(9):
1219-1229.
Rodnina, M. V., A. Savelsbergh, et al. (1997). "Hydrolysis of GTP by elongation
factor G drives tRNA movement on the ribosome." Nature 385(6611): 37-41.
Rodnina, M. V., H. Stark, et al. (2000). "GTPases mechanisms and functions of
translation factors on the ribosome." Biological chemistry 381(5-6): 377-387.
Savelsbergh, A., D. Mohr, et al. (2005). "Control of phosphate release from
elongation factor G by ribosomal protein L7/12." EMBO J 24(24): 4316-4323.
Savelsbergh, A., D. Mohr, et al. (2000). "Stimulation of the GTPase activity of
translation elongation factor G by ribosomal protein L7/12." The Journal of
biological chemistry 275(2): 890-894.
Schluenzen, F., A. Tocilj, et al. (2000). "Structure of functionally activated small
ribosomal subunit at 3.3 angstroms resolution." Cell 102(5): 615-623.
Schmeing, T. M. and V. Ramakrishnan (2009). "What recent ribosome structures
have revealed about the mechanism of translation." Nature 461(7268): 12341242.
Schmidt, F. J., J. Thompson, et al. (1981). "The binding site for ribosomal protein
L11 within 23 S ribosomal RNA of Escherichia coli." The Journal of biological
chemistry 256(23): 12301-12305.
Schneider, T. D., G. D. Stormo, et al. (1986). "Information content of binding sites
on nucleotide sequences." Journal of molecular biology 188(3): 415-431.
Selmer, M., C. M. Dunham, et al. (2006). "Structure of the 70S ribosome complexed
with mRNA and tRNA." Science 313(5795): 1935-1942.
Shine, J. and L. Dalgarno (1974). "The 3'-terminal sequence of Escherichia coli 16S
ribosomal RNA: complementarity to nonsense triplets and ribosome binding
sites." Proceedings of the National Academy of Sciences of the United States of
America 71(4): 1342-1346.
Stark, H., M. V. Rodnina, et al. (1997). "Visualization of elongation factor Tu on the
Escherichia coli ribosome." Nature 389(6649): 403-406.
Strycharz, W. A., M. Nomura, et al. (1978). "Ribosomal proteins L7/L12 localized
at a single region of the large subunit by immune electron microscopy." J Mol
Biol 126(2): 123-140.
Subramanian, A. R. (1975). "Copies of proteins L7 and L12 and heterogeneity of the
large subunit of Escherichia coli ribosome." J Mol Biol 95(1): 1-8.
Terhorst, C., W. Moller, et al. (1972). "Amino acid sequence of a 50 S ribosomal
protein involved in both EFG and EFT dependent GTP-hydrolysis." FEBS Lett
28(3): 325-328.
Varshney, U., C. P. Lee, et al. (1993). "From elongator tRNA to initiator tRNA."
Proceedings of the National Academy of Sciences of the United States of
America 90(6): 2305-2309.
Wahl, M. C., G. P. Bourenkov, et al. (2000). "Flexibility, conformational diversity
and two dimerization modes in complexes of ribosomal protein L12." EMBO J
19(2): 174-186.
Wilson, D. N. and K. H. Nierhaus (2003). "The ribosome through the looking glass."
Angewandte Chemie 42(30): 3464-3486.
Wilson, D. N. and K. H. Nierhaus (2007). "The weird and wonderful world of
bacterial ribosome regulation." Critical reviews in biochemistry and molecular
biology 42(3): 187-219.
48
Wu, X. Q. and U. L. RajBhandary (1997). "Effect of the amino acid attached to
Escherichia coli initiator tRNA on its affinity for the initiation factor IF2 and on
the IF2 dependence of its binding to the ribosome." The Journal of biological
chemistry 272(3): 1891-1895.
Yonath, A., K. R. Leonard, et al. (1987). "A tunnel in the large ribosomal subunit
revealed by three-dimensional image reconstruction." Science 236(4803): 813816.
Youngman, E. M. and R. Green (2005). "Affinity purification of in vivo-assembled
ribosomes for in vitro biochemical analysis." Methods 36(3): 305-312.
Yu, D., H. M. Ellis, et al. (2000). "An efficient recombination system for
chromosome engineering in Escherichia coli." Proceedings of the National
Academy of Sciences of the United States of America 97(11): 5978-5983.
Yusupova, G. Z., M. M. Yusupov, et al. (2001). "The path of messenger RNA
through the ribosome." Cell 106(2): 233-241.
Zavialov, A. V., R. H. Buckingham, et al. (2001). "A posttermination ribosomal
complex is the guanine nucleotide exchange factor for peptide release factor
RF3." Cell 107(1): 115-124.
Zavialov, A. V. and M. Ehrenberg (2003). "Peptidyl-tRNA regulates the GTPase
activity of translation factors." Cell 114(1): 113-122.
Zavialov, A. V., V. V. Hauryliuk, et al. (2005). "Guanine-nucleotide exchange on
ribosome-bound elongation factor G initiates the translocation of tRNAs."
Journal of biology 4(2): 9.
Zavialov, A. V., L. Mora, et al. (2002). "Release of peptide promoted by the GGQ
motif of class 1 release factors regulates the GTPase activity of RF3." Molecular
cell 10(4): 789-798.
Zecherle, G. N., A. Oleinikov, et al. (1992). "The proximity of the C-terminal
domain of Escherichia coli ribosomal protein L7/L12 to L10 determined by
cysteine site-directed mutagenesis and protein-protein cross-linking." J Biol
Chem 267(9): 5889-5896.
49