fulltext

Digital Comprehensive Summaries of Uppsala Dissertations
from the Faculty of Science and Technology 1306
Accuracy of mRNA Translation in
Bacterial Protein Synthesis
JINGJI ZHANG
ACTA
UNIVERSITATIS
UPSALIENSIS
UPPSALA
2015
ISSN 1651-6214
ISBN 978-91-554-9383-7
urn:nbn:se:uu:diva-262901
Dissertation presented at Uppsala University to be publicly examined in B10:2, BMC,
Husargatan 3, Uppsala, Friday, 4 December 2015 at 13:00 for the degree of Doctor of
Philosophy. The examination will be conducted in English. Faculty examiner: Professor Olke
C. Uhlenbeck (Northwestern University).
Abstract
Zhang, J. 2015. Accuracy of mRNA Translation in Bacterial Protein Synthesis. Digital
Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and
Technology 1306. 49 pp. Uppsala: Acta Universitatis Upsaliensis. ISBN 978-91-554-9383-7.
Reading of messenger RNA (mRNA) by aminoacyl-tRNAs (aa-tRNAs) on the ribosomes in the
bacterial cell occurs with high accuracy. It follows from the physical chemistry of enzymatic
reactions that there must be a trade-off between rate and accuracy of initial tRNA selection
in protein synthesis: when the current accuracy, the A-value, approaches its maximal possible
value, the d-value, the kinetic efficiency of the reaction approaches zero. We have used an in
vitro system for mRNA translation with purified E. coli components to estimate the d- and
A-values by which aa-tRNAs discriminate between their cognate and near cognate codons
displayed in the ribosomal A site. In the case of tRNALys, we verified the prediction of a
linear trade-off between kinetic efficiency of cognate codon reading and the accuracy of codon
selection. These experiments have been extended to a larger set of tRNAs, including tRNAPhe,
tRNAGlu, tRNAHis, tRNACys, tRNAAsp and tRNATyr, and linear efficiency-accuracy trade-off was
observed in all cases. Similar to tRNALys, tRNAPhe discriminated with higher accuracy against
a particular mismatch in the second than in the first codon position. Remarkably high d-values
were observed for tRNAGlu discrimination against a C-C mismatch in the first codon position (70
000) and for tRNAPhe discrimination against an A-G mismatch in the second codon position (79
000). At the same time, we have found a remarkably small d-value (200) for tRNAGlu misreading
G in the middle position of the codon (U-G mismatch).
Aminoglycoside antibiotics induce large codon reading errors by tRNAs. We have studied the
mechanism of aminoglycoside action and found that the drug stabilized aminoacyl-tRNA in a
codon selective in relation to a codon non-selective state. This greatly enhanced the probability
of near cognate aminoacyl-tRNAs to successfully transcend the initial selection step of the
translating ribosome. We showed that Mg2+ ions, in contrast, favour codon non-selective states
and thus induce errors in a principally different way than aminoglycosides.
We also designed experiments to estimate the overall accuracy of peptide bond formation
with, including initial selection accuracy and proofreading of tRNAs after GTP hydrolysis on
EF-Tu. Our experiments have now made it possible to calibrate the accuracy of tRNA selection
in the test tube to that in the living cells. We will now also be able to investigate the degree to
which the accuracy of tRNA selection has been optimized for maximal fitness.
Keywords: protein synthesis, genetic code, misreading, error hot spots, kinetics,
aminoglycoside
Jingji Zhang, Department of Cell and Molecular Biology, Structure and Molecular Biology,
596, Uppsala University, SE-751 24 Uppsala, Sweden.
© Jingji Zhang 2015
ISSN 1651-6214
ISBN 978-91-554-9383-7
urn:nbn:se:uu:diva-262901 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-262901)
To my princess
List of Papers
This thesis is based on the following papers, which are referred to in the text
by their Roman numerals.
I
II
III
IV
Johansson, M., Zhang, J., Ehrenberg, M. (2012) Genetic code
translation displays a linear trade-off between efficiency and
accuracy of tRNA selection. Proc Natl Acad Sci U S A,
109(1):131–136
Zhang, J., Ieong, K.W., Johansson, M., Ehrenberg, M. (2015)
Accuracy of initial codon selection by aminoacyl-tRNAs on the
mRNA-programmed bacterial ribosome. Proc Natl Acad Sci U
S A, 112(31):9602–9607
Zhang, J., Pavlov, M., Ehrenberg, M. (2015) On the mechanism of translation error induction by aminoglycoside antibiotics. (Manuscript)
Zhang, J.*, Ieong, K.W.*, Ehrenberg, M. (2015) Enhanced
proofreading neutralizes potential error hot spots in genetic
code translation by transfer RNAs. (Manuscript) *Co-first author
Reprints were made with permission from the respective publishers.
Contents
Introduction ................................................................................................... 11 Early kinetic models of translation accuracy............................................ 12 Linus Pauling’s accuracy model .......................................................... 12 Bacterial protein synthesis ................................................................... 14 High and tunable intracellular accuracy of mRNA translation ............ 15 Kinetics and single step substrate selection ......................................... 16 Proofreading selection ......................................................................... 18 The present work........................................................................................... 22 Recent two step kinetic modeling of translation accuracy ....................... 22 In vivo accuracy studies............................................................................ 23 Mg2+ dependent efficiency and accuracy trade-off (Paper I and II) ......... 24 Aminoglycoside dependent efficiency-accuracy trade-off (Paper III) ..... 31 Accuracy enhancement by proofreading (Paper I and IV) ....................... 37 Conclusions ................................................................................................... 41 Summary in Swedish .................................................................................... 43 Acknowledgements ....................................................................................... 45 References ..................................................................................................... 47 Abbreviations
30S
50S
70S
A
A
aa
aa-tRNA
A site
ASL
ATP
C
E. coli
EF-G
EF-Tu
fMet
G
GTP
IC
IF
IleRS
Mg
mRNA
PTC
P site
PEP
rRNA
R group
RNA
T3
tRNA
U
ValRS
The small subunit of a bacterial ribosome
The large subunit of a bacterial ribosome
The complete bacterial ribosome
Adenosine
Accuracy
Amino acid
Aminoacyl-tRNA
Aminoacyl-tRNA site
Anticodon stem loop
Adenosine 5’-triphosphate
Cytidine
Escherichia coli
Elongation factor G
Elongation factor Tu
Formylmethionine
Guanosine
Guanosine 5’-triphospate
Initiation complex
Initiation factor
Isoleucyl-tRNA synthetase
Magnesium
Messenger RNA
Peptidyl transferase center
Peptidyl-tRNA site
Phosphoenolpyruvate
Ribosomal RNA
Amino acid side chain
Ribonucleic acid
Ternary complex (aa-tRNA·EF-Tu·GTP)
Transfer RNA
Uridine
Valyl-tRNA synthetase
Introduction
In the living cell, genetic information stored in molecules of DNA is transcribed into messenger RNA (mRNA) molecules and then translated into
protein by the ribosome. The ribosome reads the genetic code, composed of
base triplets (codons) in the mRNA sequence. Each codon is decoded by a
transfer RNA (tRNA) carrying the corresponding amino acid. There are 61
codons to code for the 20 natural amino acids and three stop codons, which
tell the ribosome to end translation of the message. It is vital for life that this
process of translating the genetic code in mRNA into proteins is both fast
and highly accurate.
This thesis concerns the accuracy of bacterial protein synthesis and is split
into two sections. First there is a brief history of accuracy research including
Linus Pauling’s first accuracy model from 1957, a short overview of bacterial protein synthesis, Robert Loftfield’s in vivo accuracy measurements in
1972, Luigi Gorini’s ribosomal accuracy mutants in 1971, Jaques Ninio’s
kinetic interpretation of Gorini’s data and John Hopfield’s proofreading
mechanism in 1974, as well as experimental evidence for proofreading from
Anne Norris Baldwin and Paul Berg (1966), John Hopfield (1976), Robert
Thomson and Pamela Stone (1977) and Alan Fersht’s double sieve mechanism (1979).
The second section concerns the present work and includes Marina Rodnina’s kinetic model for tRNA selection by the ribosome in 2004 and Philip
Farabaugh’s in vivo measurements of accuracy in 2014. It focuses on our
measurements of the Mg2+ concentration (Paper I and II) and aminoglycoside (Paper III) dependent speed and accuracy trade-off in initial codon selection as well as our measurements of the total accuracy (Paper IV).
11
Early kinetic models of translation accuracy
Linus Pauling’s accuracy model
Already in the 1950s Linus Pauling formulated essential concepts for the
later understanding of the accuracy of protein synthesis. He used his experimental estimates of the precision by which antibodies can discriminate between a cognate antigen and its near cognate competitors along with physical
chemical modeling of ligand binding to proteins to discuss the frequency of
errors in protein synthesis. He postulated the existence of template proteins
designed to recognize their cognate amino acids (aac) and reject near cognate
amino acids (aanc). Discrimination was modeled as an equilibrium between a
template protein, E, and its amino acid ligands (Pauling, 1957):
Kc
E  aa c nc
nc
 c nc
C

Cc is the cognate and Cnc the near cognate complex, Kc/nc the dissociation
equilibrium constant for cognate and near cognate complex formation, i.e.
the ratio between the dissociation rate constant and the association rate constant.
Pauling defined the affinity difference between cognate and near cognate
substrates as the discrimination ratio Knc/Kc, the d-value. He focused on how
well a template protein with a binding pocket for a methyl group could possibly discriminate against a hydrogen atom and, vice versa, how well a template protein cognate to a hydrogen atom (group) could discriminate against
a methyl group (Table 1).
12
Table 1. Comparison of binding constant for antibody to cognate and
near cognate antigens (d-value).
Experimental data
London theory of
electronic dispersion forces
Forces between antibody and
hapten
Template
Van der
Waals repulsion
Antibody
Cognate
substrate
Near cognate
dsubstrate
value1
Benzoic acid
with methyl
group
Antigen with
Antigen with
Antibody
hydrogen
methyl group
group
Antigen with
Van der
Acetylcholine
Antigen with
hydrogen
Waals attrac- esterase inhibitor methyl group
group
tion
Protein
Benzoic acid
with hydrogen group
Alanine
100
Glycine
3-6
6-7
4.3
1
The ratio of near cognate and cognate dissociation constants (d-value) are from
(Pauling, 1957).
Linus Pauling identified van der Waals attraction, van der Waals repulsion, hydrogen-bond formation along with attraction between positively and
negatively charged groups as important factors for amino acid recognition by
template proteins (Pauling, 1957). He expected discrimination against near
cognate amino acids with smaller R group than the cognate amino acid to be
more difficult than when the near cognate amino acid had a larger R group.
This hypothesis was based on the idea that steric exclusion by van der Waals
repulsion would be a better discriminator than the differential binding energy
provided by van der Waals attraction to the larger side group. He found that
an antibody, evolved for cognate binding to benzoic acid with a hydrogen
group, had 100 times larger dissociation constant for benzoic acid with the
hydrogen replaced by a methyl group (Table 1). From this he postulated that
the machinery for protein synthesis in the living cell would not be able to
discriminate against near cognate amino acids with a methyl group replacing
the hydrogen of the cognate amino acid better than by a factor of 100, which
would lead to amino acid substitution errors in the 1% range in such cases.
He exemplified with the amino acids alanine (methyl group) and glycine
(hydrogen group). However, in the case where a near cognate amino acid has
smaller R group than the cognate one, then even larger error frequencies
could be expected. In this case it is the van der Waals attraction energy that
is important. Here, he calculated the error to be around 20% for incorpora13
tion of near cognate glycine instead of alanine, meaning that the accuracy of
this mis-incorporation would be about 4 (Table 1). A similar example is
incorporation of valine as the near cognate amino acid with a hydrogen
group instead of isoleucine with a cognate methyl group.
Bacterial protein synthesis
Synthesis of peptide chains on the mRNA programmed ribosome includes
four stages: initiation of protein synthesis, peptide chain elongation, termination of protein synthesis and recycling of ribosomes. In brief, these stages
are as follows.
Initiation
To initiate protein synthesis, initiation factor 3 (IF3) binds to the small (30S)
subunit, preventing it from premature docking with the large (50S) ribosomal subunit (Antoun et al., 2006b). An mRNA binds to the 30S subunit and
is positioned with its initiation codon (AUG) in the ribosomal P site by base
pairing between its Shine-Dalgarno sequence (SD-sequence) and the anti-SD
sequence of the 16S ribosomal RNA (rRNA). With the help of initiator factors 1 (IF1) and 2 (IF2) initiator tRNA (fMet-tRNAfMet) binds to the initiation codon (Antoun et al., 2006a). Then the 50S subunit docks with the 30S
subunit, the initiation factors dissociate from the ribosome after GTP hydrolysis on IF2 and the 70S initiation complex with initiator tRNA in P site is
ready for entry of the first elongator tRNA into the A site.
Peptide Elongation and Translocation
Aminoacyl-tRNA in ternary complex with EF-Tu and GTP binds to the A
site of the 70S post-translocation complex with peptidyl-tRNA in the P site
(Schmeing et al., 2009). Cognate interaction between the anticodon of the
incoming aminoacyl-tRNA and the mRNA codon in the A site rapidly leads
to GTP hydrolysis on EF-Tu and release of EF-Tu in the GDP form, to be
recycled back to EF-Tu·GDP by elongation factor Ts (EF-Ts). Then the nascent peptide chain is transferred from the P site tRNA by peptide bond formation with the amino acid of the A-site accommodated aminoacyl-tRNA.
Then EF-G·GTP binds to the pre-translocation ribosome with peptidyltRNA in hybrid A/P state and deacylated tRNA in hybrid P/E state
(Ermolenko et al., 2007; Spiegel et al., 2007), hydrolyses GTP and translocates the mRNA by one codon and the tRNAs to P/P and E/E states, respectively (Rodnina et al., 1997). After release of EF-G·GDP the now empty A
site is ready for yet another round of peptide elongation.
14
Termination
When the A site displays one of the three stop codons (UAA, UAG or
UGA), a class one release factor (RF1 or RF2) binds to the A site (Scolnick
et al., 1968) and promotes ester bond hydrolysis and release of the completed peptide chain from the P site bound peptidyl-tRNA. The GTP form of the
class two release factor (RF3) induces the ribosome conformational change
that releases the class one release factor from the post-termination ribosome
(Gao et al., 2007). After GTP hydrolysis on RF3 it leaves the posttermination ribosome (Zavialov et al., 2001; Zavialov et al., 2002).
Recycling
Ribosome recycling factor (RRF) and EF-G·GTP split the post-termination
complex into ribosomal subunits in a GTP hydrolysis dependent manner
(Karimi et al., 1999; Zavialov et al., 2005). Then IF3 binding to the 30S
subunit removes its remaining deacylated tRNA, making the ribosomal subunits ready for yet another round of initiation (Karimi et al., 1999; Peske et
al., 2005).
High and tunable intracellular accuracy of mRNA translation
Robert Loftfield developed assays to measure the accuracy of protein synthesis in the living cell, which made it possible for him to seriously test Pauling’s prediction of sloppy intracellular messenger RNA translation
(Loftfield and Vanderjagt, 1972). He studied isoleucine-containing peptide
fragments from rabbit haemoglobin and estimated the frequency of valine
substitution for isoleucine as between two and six per 10 000. This frequency is about a thousand times smaller than Pauling’s predictions and Loftfield
speculated that cells may have a “scavenging mechanism for eliminating
errors, but more probably Nature has found means for accentuating the small
differences among amino acids and the similarly small differences among
trinucleotide code-words”. These words were remarkably prophetic in that
we now know that single step discrimination can be much better than the
poor d-values estimated by Pauling and that there indeed exist if not scavenging so at least editing mechanisms that allow for repeated selection of the
same “small difference” between cognate and near cognate substrates.
A plethora of antibiotic drugs bind to the bacterial ribosome (Yonath,
2005) and some of these induce high amino acid substitution errors in the
living cell as well as in the test tube (Ogle and Ramakrishnan, 2005). One of
these is streptomycin (Sm), a bacteriocidal drug which has generated a large
number of Sm-resistant, Sm-dependent, Sm-pseudo dependent and Sm-hyper
sensitive mutants in clever selection schemes pioneered by Luigi Gorini. He
found that some Sm-resistant bacteria (strA) appeared to have more accurate
mRNA translation than wild type strains due to amino acid substitutions in
15
ribosomal protein S12 of the small ribosomal subunit. He also found that
other ribosomal mutants displayed not only increased sensitivity to Sm but
also reduced accuracy of mRNA translation. These low accuracy mutations
were localized to alterations in proteins S4 and S5, of the small ribosomal
subunit (Ozaki et al., 1969). Gorini was fascinated by his finding that a single ribosome alteration in a ribosomal protein could leave cognate codon
reading virtually unchanged but greatly affect the propensity of near cognate
reading up or down.
Kinetics and single step substrate selection
Jacques Ninio not only introduced chemical kinetics and used this tool to
tentatively interpret Gorini’s complex data sets (Ninio, 1974) but also forged
generally applicable concepts to explain selective enzymatic pathways in
general. Here, we describe a modified version of Ninio’s kinetics albeit
keeping in line with his general ideas. The mRNA programmed ribosome
selects aminoacyl-tRNAs in ternary complex with elongation factor Tu (EFTu) and guanosine-5’-triphosphate (GTP) for binding to the ribosomal A site
and subsequent hydrolysis of GTP by EF-Tu.
A simple aa-tRNA selection scheme that summarizes the essence of Ninio’s contribution can be written:
k1
T3  R
c nc


q1c
c nc
C1
kc
nc
2


nc
Here the same type of ternary complex (T3) binds to a ribosome programmed with its cognate, R c , or a near cognate, R nc , codon with the same
association rate constant (k1) to form a cognate, C1c, or a near cognate ribosome complex, C1nc. Products are formed with rate constants k2c/nc. We define a discard parameter a (q1c /k2c) which determines the efficiency of cognate codon reading. The discrimination value (d) is (q1nc /q1c)· (k2c /k2nc). The
steady state flows to product formation are given by:
 
k
j  T3   R   cat
 Km
c
c
 
c

k
k1
  T3  R c 
 T3  R c  1
c
1 a
q

1  1c
k2
k 
j nc  T3   R nc   cat 
 Km 
16
nc
 T3  R nc 
k1
k1
 T3  R nc 
1 d  a
q1nc
1  nc
k2
The ratio between cognate and near cognate flow is given by:
c
 
 
T3  R c
jc

j nc T3   R nc
 k cat 


K m 
T3  R c  1  d  a



nc
T3  R nc 1  a
 k cat 


 Km 
 
 
If in both cognate and near cognate cases the T3 and R concentrations are
equal, then
c
jc
j nc
 k cat 


K m 
1 d  a



A
nc
1 a
 k cat 


 Km 
The accuracy (A) is defined as the ratio between cognate, (kcat/Km)c, and
near cognate efficiency, (kcat/Km)nc, of product formation.
The d-value is normally much larger than one. If it is smaller than one,
the definition of cognate and near cognate would change. It is interpreted as
the maximal possible accuracy of cognate in relation to near cognate product
formation (Ehrenberg and Blomberg, 1980; Ehrenberg and Kurland, 1984)
and can be written as
d  eG
c ,nc
/ RT
Here, ΔΔGc,nc is the standard free energy difference between cognate and
near cognate ribosome complex in the transition state of the GTP hydrolysis
reaction. R is the gas constant and T the absolute temperature. The a-value
can be understood as an accuracy tuning parameter that determines how
much of the d-value that the enzyme uses in its current A-value. Now we
come back to Ninio’s kinetic explanation of the phenotypes of Gorini’s ribosome mutants (Ninio, 1974). The Ram mutation decreases the q1c/nc parameter by the same factor for cognate and near cognate substrate, meaning that
the phenotype of Ram is reduced a-parameter. In the extreme case, when a is
close to zero, the A goes to its minimum value of 1. So the Ram mutation is
an error prone mutant as Gorini’s data showed. The StrA mutation decreases
the k2c/nc parameter which increases the a parameter. In the extreme case,
when a is very large, then A approaches its maximum value d. So the phenotype of the StrA mutation is hyper accurate code translation. This means, in
other words, that Ninio suggested that these mutations affected the a-
17
parameter up or down but left the d-parameter unaltered. In this case there is
a linear relation between (kcat/Km)c and A:
 kcat

K
 m
c

dA
  k1

d 1

Plotting (kcat/Km)c versus A leads to a straight line:
By changing the a-value to change (kcat/Km)c and A, the d-value can be obtained from the intercept of the line with the x-axis. The accuracy A is essentially the inverse of the error frequency E:
E
j nc

j nc  j c
1
1
c
j
j nc

1
1
~
1 A A
Proofreading selection
Linus Pauling’s predictions of the accuracy of protein synthesis in living
cells had turned out to be severe underestimates. The deviation between prediction and reality could mean that he underestimated the d-values that can
be achieved by the aminoacyl-tRNA synthetases that couple the amino acids
to their cognate tRNAs (Loftfield et al., 1973). Another option would be that
the living cell could, somehow, use the same (small) intrinsic discrimination
capacity, the d-value, several times and in this way supersede the physical
chemical limitation of selective enzymes.
18
One intriguing experiment suggested that the latter option could be true. It
was found by Anne Norris Baldwin and Paul Berg in 1966 (Baldwin and
Berg, 1966) that, after activation of Ile to Ile-AMP by isoleucyl-tRNA synthetase (IleRS), addition of tRNAIle led to rapid formation Ile-tRNAIle:
IleRS  Ile  AMP  tRNA Ile  Ile  tRNA Ile  IleRS  AMP
However, after activation of the near cognate amino acid Val to Val-AMP
by IleRS, addition of tRNAIle just leads to hydrolysis of Val-AMP:
IleRS  Val  AMP  tRNA Ile  Val  tRNA Ile  IleRS  AMP
In the mid-seventies, John Hopfield (Hopfield, 1974) and Jacques Ninio
(Ninio, 1975) proposed a kinetic proofreading step after initial selection, that
would allow the same difference in standard free energy between near cognate or cognate substrate to be used several times. A generic scheme for
proofreading applied to mRNA translation may be written as:
k1
T3  R c nc
 c nc k 2c nc
C1 

q1c
c nc
C2
qc
nc
R c nc
kc
nc
3


nc
 2
 aa  tRNA  EF  Tu  GDP
In this scheme, initial selection of ternary complex ends through GTP hydrolysis by EF-Tu in complex C1c/nc leading to complex C2c/nc and release of
inorganic phosphate (Pi). From complex C2c/nc the reaction can either lead
back to ribosomal complex Rc/nc through release of EF-Tu·GDP and aatRNA with rate constant q2c/nc or, alternatively to A-site accommodation and
peptide bond formation with rate constant k3c/nc. We note that in this scheme
the influx from ribosome complex Rc/nc over the vertical proofreading path
has been neglected. The reason why this can be done and, in fact, why proofreading works at all is that the ternary complex concentration is very far
above equilibrium with EF-Tu·GDP and aa-tRNA. In the living cell such out
of equilibrium shifts have been estimated as 108 (Hopfield, 1974), which
provide yet another upper limit to how much proofreading can enhance the
accuracy of enzymatic selection (Ehrenberg and Blomberg, 1980). At equal
concentrations of cognate and near cognate ribosomes the accuracy, A, is
given by:
19
c
 kcat 


c
Km 
1  d1  a1 1  d 2  a2
j

A  nc 


nc
1  a1
1  a2
j
 kcat 


 Km 
Here a1=q1c /k2c, a2=q2c /k3c, d1=(q1nc /q1c)· (k2c /k2nc) and d2=(q2nc
/q2c)· (k3c /k3nc). The maximum accuracy is d1d2. This scheme shows in
concrete terms that the total accuracy of an enzyme system can be much
higher than the d-value of initial selection, here equal to d1. Referring to the
above question; the A could be larger than d1. It also follows from this type
of scheme that the proofreading contribution to the accuracy is equal to the
ratio between the numbers of GTPs (or ATPs in other cases) hydrolyzed per
near cognate (fnc) and cognate (fc) product formation (Hopfield, 1974). In
1976, Hopfield and Yamane proved experimentally that on average 1.5
ATPs are hydrolyzed per Ile-tRNAIle formed by IleRS in the ATP driven
aminoacylation reaction. In contrast, 270 ATPs were consumed per formed
Val-tRNAIle by IleRS. This demonstrated that proofreading exists in this
aminoacylation reaction and suggested the proofreading contribution to the
overall accuracy of 180 (270/1.5) (Hopfield and Yamane, 1976). We note
that the probabilities that an aminoacyl-tRNA “survives” the proofreading
step in aminoacylation is 1/fnc=P2nc in the near cognate and 1/fc = P2c in the
cognate case, so the F=P2c/P2nc =fnc/fc. Now, the accuracy A is the initial
selection, I , multiplied with the proofreading selection F: A=I·F.
Thompson and Stone (1977) were the first to demonstrate proofreading of
aa-tRNAs in bacterial protein synthesis (Thompson and Stone, 1977). Later,
Ruusala et al. (1982) demonstrated proofreading of aa-tRNAs in a more
complete in vitro system for bacterial protein also containing elongation
factor G (EF-G) (Ruusala et al., 1982).
Now, we have seen that proofreading of aa-tRNAs enhances the accuracy
of protein synthesis on the ribosome and that proofreading of amino acids by
aminoacyl-tRNA synthetases enhance the accuracy of aminoacylation. What
about the d-value estimates provided by Pauling? Are they accurate? Alan
Fersht measured kcat/Km for activation of Val (cognate) and Ile (near cognate)
by valyl-tRNA synthetase (ValRS) (Fersht and Dingwall, 1979), and showed
that the cognate initial selection reaction was 60 000 times more efficient
than the near cognate one, i.e. that I=60 000 (Fersht and Dingwall, 1979).
This means a d-value at least 600 times larger than Linus Pauling’s estimate
of d=100 for a template protein we can now identify as ValRS (Pauling,
1957). Fersht also measured the kcat and Km for activation of Val and Ile by
isoleucyl-tRNA synthetase (IleRS) (Fersht, 1977). From the kcat/Km values
for both cognate Ile and near cognate Val, the initial selection is about 200
20
(Fersht, 1977; Fersht, 1979), i.e. about 50 times larger than Linus Pauling’s
prediction, which was 4.3 (Pauling, 1957). These comparisons show that
Pauling’s d-value estimates for common amino acids were far below the
mark. This means, in conclusion, that the discrepancy between Pauling’s
prediction that protein synthesis in living cells is permeated with amino acid
substitution errors (Pauling, 1957) and Loftfield’s experimental observations
that such errors are extremely rare stems from his underestimated d-values
and from mother Nature’s invention of proofreading. One reason why Pauling’s d-values came out so low, could be that they were based on equilibrium constants, while Fersht’s estimates were based on the efficiency of catalytic reactions. In the latter case it is likely that substrates are positioned in
precise positions to minimize the entropy of activation of reactions.
Fersht found that Ile activation by IleRS to Ile-AMP complex is a fast, initial step and Ile transfer to tRNAIle is a slow, rate limiting step (Fersht and
Kaethner, 1976). He measured the efficiency of ValRS activation of Ile
(large side group), Val (cognate) and Thr (isosteric side group) (Fersht and
Dingwall, 1979). He found that the aa-activation site can discriminate
against amino acids with side chains larger than that of Val, but not against
those with smaller side chains. Discrimination against substrates larger than
the cognate constitutes the first sieve in the double sieve mechanism proposed by Fersht to explain amino acid selection by aminoacyl-tRNA synthetases (Fersht and Dingwall, 1979). The editing site, in contrast, could discriminate against amino acids with side groups smaller than that of the cognate substrate. This constitutes the second sieve of the double sieve mechanism suggested by Fersht.
21
The present work
Recent two step kinetic modeling of translation
accuracy
A kinetic model for initial codon selection was proposed by Marina Rodnina
and collaborators (Gromadski and Rodnina, 2004):
k1
T3  R
c nc


q1
k2
C1c nc


k c nc
3
C2c nc 

q 2c nc
In the above scheme, R is the ribosome, T3 is the ternary complex, c
stands for cognate, nc for near cognate, C1 is the initial binding complex.
The first binding step is not codon-anticodon selective, and thus with the
same dissociation constant, Kd (q1/k1), for cognate and near cognate ternary
complex (Rodnina et al., 1994). When the Mg2+ concentration was varied
from 5 mM to 10 mM, Kd decreased ~ 100 fold (Rodnina et al., 1996). At
Mg2+ concentrations less than 3 mM, Kd is above 30 µM, leading to rapid
dissociation from C1 (rate constant q1) after binding to (rate constant k1) this
state. Under these low Mg2+ conditions, there is negligible inhibition of protein synthesis by near cognate ternary complex (Johansson et al., 2008).
The forward rate constant, k2, leading to codon recognition in complex C2
is the same for cognate and near cognate reactions. The C2 complex is much
more stable for cognate than for near cognate ternary complex: the ratio q2nc
/q2c ratio is about 1000 (Gromadski and Rodnina, 2004).
GTPase activation and hydrolysis are irreversible steps as indicated in the
above scheme. The rate of conformation change of EF-Tu after GTP hydrolysis has been analyzed by fluorescence from mant-GTP replacing GTP in the
ternary complex. It was found that k3c is larger than 260 s-1 and that k3nc is
0.4 s-1 (Gromadski and Rodnina, 2004). The much smaller value of k3nc than
of k3c has been used to characterize GTPase activation as an induced fit
mechanism (Zaher and Green, 2009).
To measure the proofreading contribution to accuracy, purified ternary
complex was mixed with ribosomes programmed with cognate or near cognate codon for the aminoacyl-tRNA and the relative yield (1/fc or 1/fnc) of
22
peptide bond formation was estimated in each case (Gromadski and Rodnina,
2004).
Rodnina and co-workers estimated K2c/nc-values for the equilibrium between complexes C1 and C2, containing Phe-tRNAPhe interacting with either
one cognate or six near cognate codons in the ribosomal A site. They found
similar K2nc values in all near cognate cases and suggested uniform accuracy
in ternary complex recognition (Gromadski et al., 2006).
Using pre-steady kinetics, Hani S. Zaher and Rachel Green investigated
the above mentioned classical ribosome mutants StrA and Ram found by
Gorini (Zaher and Green, 2009). The Ram mutation mainly increased the
error of initial ternary complex selection, whereas the StrA mutation mainly
decreased the error in proofreading selection of aminoacyl-tRNA.
In vivo accuracy studies
The accuracy of the steps in the central dogma displays great variation. In
DNA replication, transcription and translation accuracy levels of 108-1010
(Kunkel and Bebenek, 2000), 104 (Rosenberger and Foskett, 1981) and 103104 (Bouadloun et al., 1983; Edelmann and Gallant, 1977; Kramer and
Farabaugh, 2007; Laughrea et al., 1987), respectively, have been estimated.
Codon reading accuracy measured in vivo using enzyme (firefly luciferase or
beta-galactosidase) mutants that need missense errors for residual activity
have an error background level between 10-6 and 10-5 (Manickam et al.,
2014). With these in vivo assays Farabaugh and co-workers have estimated
different types of missense errors when tRNAGlu, tRNATyr or tRNAAsp misread their near cognate codons (Kramer and Farabaugh, 2007; Manickam et
al., 2014). Concerning the first codon position, the U36-U1 mismatch is
error prone (tRNALys misreading stop codon UAA). U-C, U-G, A-A, A-C,
A-G, C-C, C-U and C-A mismatches do not lead to significant errors. The
most error frequent codon reading concerns the second and third codon positions. In the second codon position, U35-G2 (tRNATyr misreading codons
UGC, UGU; tRNALys misreading codons AGA, AGG; tRNAAsp misreading
codons GGU, GGC; tRNAGlu misreading codons GGA, GGG) is a mismatch
with high error propensity. In the third codon position, mnm5s2U34-U3/C3
(tRNALys misreading codons AAU, AAC; tRNAGlu misreading codon GAU,
GAC) is more error prone than Q34-A3/G3 (tRNAAsp misreading codon
GAA, GAG; tRNATyr misreading codon UAA, UAG). In contrary, our biochemical system with E. coli components of high purity measures error levels down to 10-8 (Paper IV).
23
Mg2+ dependent efficiency and accuracy trade-off
(Paper I and II)
The following generic model for initial selection of codons by tRNAs may
serve as the starting point for a discussion about the accuracy of bacterial
protein synthesis:
Fig 1. Kinetic scheme of initial codon selection on the mRNA programmed
ribosome.
The cognate (c) and near cognate (nc) kcat/Km-values (efficiencies) for
GTP hydrolysis are given by:
 kcat / Km 
c , nc
 ka
kcc ,nc
ka

c , nc
c , nc
c , nc
kc  k d
1  kd / kcc , nc
The (normalized) accuracy of initial selection is defined as the ratio between cognate and near cognate kcat/Km-values so that:
A   kcat / Km  /  kcat / Km  
c
nc
1  kdnc / kcnc 1  da

1  kdc / kcc
1 a
It follows that the discrimination parameter, d, and discard parameter, a,
are given by:
d
kcc kdnc
 eG / RT ; a  kdc / kcc
kcnc kdc
In an experiment where a is varying, e.g. by changing the free Mg2+ concentration, at constant ka and d-values, there is a linear trade-off between
cognate efficiency and accuracy:
 kcat / K m 
24
c
 ka
dA
d 1
We measured the cognate and near cognate efficiency by performing GTP
hydrolysis experiments (Fig 2)
Fig 2. Measurements of kcat/Km-parameters for GTP hydrolysis during cognate
or near cognate codon reading. Time evolution of the level of [3H]GDP in response to [3H]GTP·EF-Tu·Cys-tRNACys binding to 70S ribosomes programmed
with cognate codon or near cognate codons. For the cognate reaction in short timeframe (top left panel), data was fitted to a single exponential function. Each near
cognate experiment was performed in parallel with a cognate experiment (in black).
The very same ternary complex mixture was here used for both cognate and near
cognate reactions, and both curves were jointly fitted with sharing of parameters to
increase precision of the measurement. In all experiments ribosomes were in excess
over ternary complexes, and kcat/Km-values were calculated from the apparent GTP
hydrolysis rate constant divided by the active ribosome concentration (here, 0.7 µM
and 1.8 µM ribosomes were used for cognate and near cognate reactions, respectively). The decrease in [3H]GDP level in the long time-frame is due to spontaneous
dissociation of [3H]GDP from EF-Tu followed by its rapid regeneration to [3H]GTP
by pyruvate kinase. All experiments were performed in polymix buffer with the
addition of 2 mM extra Mg(OAc)2.
25
From such experiments the initial selection accuracy can be calculated
from ratios between cognate and near cognate kcat/Km-values. With this experimental tool, the GTP hydrolysis efficiency for cognate and near cognate
codon reading by tRNAGlu, tRNAPhe, tRNAHis, tRNACys, tRNATyr, tRNAAsp
and tRNALys was studied. Increasing kcat/Km-values for cognate and near
cognate codon reading at different Mg2+ concentrations are displayed in Figs
3A and 3B, respectively. Selected linear trade-off plots are shown in Fig. 3C
and all the seven association rate constants, ka, for cognate codon reading are
shown in Fig. 3D.
26
Fig 3. The rate-accuracy trade-off. (A) Efficiency of cognate GTP hydrolysis,
(kcat/Km)c, for different tRNAs (see panel B for symbol legend) reading their fully
matched codons at varying Mg2+ concentration. (B) Efficiency of near cognate GTP
hydrolysis, (kcat/Km)nc, for different tRNAs reading single mismatch codons at varying Mg2+ concentration. (C) Efficiency of cognate GTP hydrolysis, (kcat/Km)c, versus
the accuracy (calculated as the ratio (kcat/Km)c/(kcat/Km)nc) for different tRNAs reading single mismatch codons as indicated in panel B. In each tRNA misreading case,
cognate and near cognate (kcat/Km) values were measured at different Mg2+ concentrations as shown in panels A and B. The x-axis intercept gives the maximal accuracy, d, for each misreading case, and the y-axis intercept gives the rate constant for
association of each cognate tRNA to the ribosome. (D) The rate constant, ka, for
association of different aa-tRNAs in ternary complexes to ribosomes, estimated from
the linear dependence between cognate GTP hydrolysis efficiency and accuracy.
Data in panels A, B and C represent weighted averages from at least two experiments ± propagated standard deviation. Error bars in panel D represent the standard
deviation estimates from the parameter fitting procedure, where experimental errors,
such as in panel A and B, are used as weights.
27
As seen, all cognate kcat/Km values saturate at high Mg2+ concentration but
not the near cognate ones, as explained in paper III. The cognate saturation
values, ka , are in the range from 75 µM-1s-1 to 200 µM-1s-1 among the seven
studied tRNAs. In all cases, we observe linear trade off plots with y-axis
intercepts estimating ka-values and x-axis intercepts estimating d-values:
28
Fig 4. Maximal accuracy variation dependent on codon position, mismatch
identity, and aa-tRNA. Maximal accuracy values, d, for single mismatch readings
by different tRNAs, summarized with respect to mismatch codon position (columns)
and mismatch identities (rows for the anticodon bases and colours for the codon
bases as indicated in the figure).
29
The d-values varied from 200 to 84,000 with different types of mismatch,
tRNA and codon position. In general, with the same type of tRNA and mismatch, the second codon position had highest and the third codon position
the lowest d-value. Pyrimidine-purine mismatches show the lowest d-values
in general. There are also error hot spots: tRNAGlu misreading GGA (Gly)
with d-value 200, GAU (Asp) with d-value 240, GAC (Asp) with d-value
650; tRNAHis misreading CGC (Arg) with d-value 250, CAA (Gln) with dvalue 200.
30
Aminoglycoside dependent efficiency-accuracy tradeoff (Paper III)
Taking advantage of quench-flow techniques and data analysis, previously
applied to initial codon selection (Paper II), we estimated the efficiency
(kcat/Km) by which Phe-tRNAPhe in ternary complex with EF-Tu and GTP
selects a cognate (UUC) and a near cognate codon (CUC) at different Mg2+
concentrations in the presence and absence of paromomycin (Figure 5A).
Fig 5. Mg2+ concentration dependence of the efficiency of ribosome induced
GTP hydrolysis in initial selection of ternary complex. Efficiency of GTP hydrolysis for tRNAPhe-containing ternary complex reading cognate (UUC) codon
(kcat/Km)c in absence ( ) or presence ( ) of paromomycin or reading near cognate
(CUC) codon in absence ( ) or presence ( ) of paromomycin at varying Mg2+ concentration.
The cognate efficiency was identical in the absence and presence of
paromomycin and varied about fourfold from its lowest to its plateau value
as the free Mg2+ concentration increased from 1 to 22 mM. The near cognate
efficiency, in contrast, increased by four orders of magnitude in the absence
and by two orders of magnitude in the presence of paromomycin. All efficiencies reached the same plateau value in the limit of high Mg2+ concentration. The observations that paromomycin did not affect the variation of cognate kcat/Km and that all efficiency curves reached the same efficiency plateau
can be accounted for by a simple three step model for initial selection:
31
Fig 6. Schematic and efficiency of initial codon selection on the mRNA programmed ribosome. (A) Kinetic scheme of initial codon selection on the mRNA
programmed ribosome. (B) Simplified kinetic scheme of panel corresponding to the
cartoon in (A). (C) Efficiency of cognate and near cognate initial codon selection.
R2 is a ternary complex-containing ribosome complex without codonanticodon contact, as suggested (Rodnina et al., 1994), while R3 is a ternary
complex-containing ribosome complex with codon-anticodon contact in
which the monitoring bases A1493, A1492 and G530 are activated and in
contact with the minor groove of the codon-anticodon helix (Carter et al.,
2000; Lynch et al., 2003). Hence, the only codon specific parameters are
a2=q2c/k3c and a2d2=q2nc/k3nc so that
c
 k cat

 Km

k1
 
1

a
1 1  a 2 

 k cat

 Km



nc

k1
1  a1 1  a2 d 2 
In all cases, parameter a1, unaffected by amino glycoside binding decreased with increasing Mg2+ concentration from about three towards zero.
Parameter a2, unaffected by Mg2+ concentration, decreased by a large factor
upon paromomycin addition from a value much smaller than one to an even
smaller value. Parameter a2d2, unaffected by Mg2+ concentration, was greatly
reduced by aminoglycoside binding from a very large to a smaller but still
large number. Since in the cognate case a2 was much smaller than one and
32
became even smaller upon drug addition, the cognate curves in Fig. 5 were
insensitive to the presence or absence of the drug. Since, in contrast, a2d2
was much larger than one, its large decrease with paromomycin was clearly
seen.
Cognate and near cognate kcat/Km-values were used to construct efficiency-accuracy trade-off lines for ternary complex-containing Phe-tRNAPhe
misreading CUC (Fig. 7A), UCC (Fig. 7B) and UUA (Fig. 7C) in absence of
aminoglycosides or presence of paromomycin, gentamicin or neomycin. For
each near cognate codon, different aminoglycosides resulted in different
slopes of the straight lines, whereas the intercepts with the y-axis, ka, representing the cognate rate constant for association of the specific aa-tRNAcontaining ternary complex to the ribosome, was the same as without durgs.
All d-values obtained in this work are summarized in Table 2 along with
previously obtained tRNAPhe data (Zhang et al., 2015).
33
Fig 7. Efficiency-accuracy trade-off and rate constants for cognate ternary
complex dissociation from the ribosome. Efficiency of cognate GTP hydrolysis,
(kcat/Km)c, versus the accuracy (calculated as the ratio (kcat/Km)c/(kcat/Km)nc) for
tRNAPhe reading CUC (A), UCC (B) and UUA (C) codons, in each case with no
drug, paromomycin, gentamicin or neomycin. In each tRNA misreading case, cognate and near cognate (kcat/Km) values were measured at different Mg2+ concentrations. The x-intercept gives the effective discrimination parameter, de, for each misreading case, and the y-intercept gives the rate constant for association of each cognate tRNA to the ribosome. (D) Time evolution of dipeptide bond formation in experiment, where A-site bound ternary complex EF-Tu(H84A)·GTP·Phe-tRNAPhe
with GTPase deficient EF-Tu mutant His84A is chased by ternary complex with
wild type EF-Tu.
34
Table 2. Rate constant q2 of cognate reaction compared with d-values for
different near cognate codons with and without antibiotics
No
Paromomycin
Gentamicin
Neomycin
drug
chase of ternary
1.04
0.08
0.016
0.008
complex, q2c (s-1)
Fold change by
1
14
70
170
drug
1760 ±
d-vaue tRNAPhe
75 ± 10
17 ± 4
3.9 ± 0.6
on CUC
180
Fold change by
1
23
104
451
drug
5900 ±
d-vaue tRNAPhe
420 ± 76
103 ± 20
30 ± 5
on UCC
970
Fold change by
1
14
57
197
drug
1200 ±
d-vaue tRNAPhe
111 ± 22
14,6 ± 0,8
9 ± 1.7
on UUA
190
Fold change by
1
11
82
133
drug
Form the Mg2+ dependent steps each a1-value at each Mg2+ concentration
could be estimated. The effective d-value (a2d2) decreased 23-, 14-, and 11fold upon paromomycin addition in cases of tRNAPhe reading CUC (Fig.
7A), UCC (Fig. 7B) or UUA (Fig. 7C) codon (Table 1). If we take the average of these three numbers, the mean d-value decrease by paromomycin was
16 fold (Table 3). Upon gentamicin addition the corresponding d-values
decreased by factors of 104, 57 and 82 with an average of 80. Upon neomycin addition these d-value reductions were 451, 197 and 133 with an average
of 260 (Table 3). The smallest effective d-value in Table 1 is 4, obtained
from tRNAPhe reading CUC codon in the presence of neomycin. This result
shows that any residual codon selectivity of the parameter ratio q1/k2 in Fig.
6 must be smaller than 4, in line with the hypothesis that complex R2 lacks
codon-anticodon contact.
From the above section, we know that aminoglycosides profoundly affect
the effective d-value a2d2. To find out if aminoglycosides affect a2, d2 or
both parameters we performed experiments in which cognate ternary complex with a GTPase deficient EF-Tu mutant was chased from the ribosomal
A site with a native ternary complex (complex III, Fig. 6A).
35
For this, we used the GTPase deficient EF-Tu mutant (H84A) (Daviter et
al., 2003) to form the mutant ternary complex (EF-Tu(H84A)·GTP·PhetRNAPhe) for pre-incubation with ribosomes with UUC programmed A sites
UUC for formation of complex III (Fig. 6A). Addition of the native ternary
complex leads to peptide bond formation at a rate determined by the compounded rate constant, qdiss, for dissociation of the mutated ternary complex
(Fig. 7D):
1
1
 c 1  a1 
c
qdiss q2
Since we have estimated a1 as 1.0 at 2.3 mM free Mg2+ concentration we
could estimate q2c as 1 s-1 in the absence of aminoglycoside (Table 2).
Using the same method, estimated q2c values in the presence of paromomycin, gentamicin or neomycin as 0.08, 0.016 and 0.008 s-1, respectively
(Table 2).
Table 3. Average d-value change compared with ternary complex dissociation rate change
-1
<d>-value change
q2c (s ) change
No drug
1
1
Paromomycin
16
14
Gentamicin
80
70
Neomycin
260
170
Table 3 shows that the fold change in effective d-value in response to addition of each one of the drugs closely corresponds to the fold change in the
cognate back reaction parameter q2 (Table 2). This result strongly suggests
that the effective d-value change in response to aminoglycoside addition is
caused by uniform reduction of the back rate constant from complex C2 (recent two step kinetic modeling translation accuracy section) for cognate and
near cognate ternary complex in an aminoglycoside specific manner. This
would mean that the reduction in effective d-value observed here does not
reduce the intrinsic accuracy of the reaction (d2) but only how much of it is
expressed in ternary complex selection for GTP hydrolysis (a2). This conclusion is in sharp contrast to previous suggestions that aminoglycosides greatly
enhance the rate constant k3 for near cognate but not cognate ternary complex, which would imply a great reduction in the intrinsic discrimination
parameter d2 (Pape et al., 2000).
36
Accuracy enhancement by proofreading (Paper I and
IV)
We also designed experiments to estimate the overall accuracy of peptide
bond formation, which includes the accuracy enhancement provided by
proofreading of tRNA following initial selection of ternary complex. We
used the following generic model for overall accuracy selection of codons by
tRNAs:
Fig 8. Kinetic scheme of peptide bond formation on the mRNA programmed
ribosome. aa-tRNAs can be rejected during initial selection or at the proofreading
step. The two selection steps are separated by hydrolysis of EF-Tu bound GTP.
In the experiments we used Lys-tRNALys, Glu-tRNAGlu and Phe-tRNAPhe
with either fully matched cognate codons or single mismatched near cognate
codons:
Fig 9. The overall accuracy of tRNA selection was measured for Lys-tRNALys,
Glu-tRNAGlu, and Phe-tRNAPhe reading all possible single-mismatch codons.
Compared to their fully matched codons AAA (tRNALys), GAA (tRNAGlu) and
UUC (tRNAPhe), mismatch codon positions are underlined.
We estimated the kcat/Km values for dipeptide formation when the same
ternary complex reacted with cognate and near cognate programmed ribosomes at varying Mg2+ concentration.
37
Figure 10. Measurements of cognate and near cognate kcat/Km-values for dipeptide formation. (A) Time evolution of f[3H]Met-Glu formation. Ternary complexes
EF-Tu·GTP·Glu-tRNAGlu were reacted with 70S initial complexes programmed
with f[3H]Met-tRNAfMet in the P site and a cognate codon GAA (black) or near cognate codon GGA (red) in the A site. Reactions were performed at increasing complex concentration as indicated in the figure. Ternary complexes were in excess over
ribosomes so that the rate of dipeptide formation kdip was limited by ternary complex
concentration. (B) Concentration dependence of the rate of dipeptide formation kdip
estimated from (A). Insert: near cognate reaction. Experiments were performed in
Polymix buffer with 2.3 mM free Mg2+.
From such experiments the overall accuracy can be calculated from ratios
between cognate and near cognate kcat/Km-values. With this experimental
tool, the dipeptide bond formation efficiency for cognate and near cognate
codon reading by tRNALys, tRNAGlu and tRNAPhe was studied. Increasing
kcat/Km-values for cognate and near cognate codon reading at different Mg2+
concentrations cause the overall accuracy decrease as shown (Fig. 11):
38
Figure 11. The rate-accuracy trade-off in overall selection. Rate-accuracy tradeoff plots in log-log scale for overall accuracy for tRNAGlu selection of GAA in relation to the GGA codon.
The trade-off line for overall accuracy of codon reading now makes it
possible to calibrate accuracy data obtained in vitro with the translation accuracy in vivo from the previous section (in vivo accuracy studies).
39
Figure 12. Comparison of in vivo and in vitro misreading error frequency. In
vivo data (black stars) from the Farabaugh lab (Manickam et al., 2014) are based on
induction of bioluminescence by mistranslation by tRNAGlu ternary complex from E.
coli strains with β-galactosidase mutants. In vitro measurements (red squares) were
performed at 2.3 mM free Mg2+ and calibrated to the in vivo condition according to
the abundance of the 2 competing tRNA species in vivo (Dong et al., 1996) (see
Materials and Methods) and assuming different ternary complexes as well as release
factors have similar efficiencies for binding to ribosomes in the living cell. Mismatch codon positions are underlined.
The best correspondence between our in vitro accuracy measurements and
those in vivo is at 2.3 mM free Mg2+ concentration in our in vitro system.
Here our measurements and those in vivo agree well at the error levels from
10-5 and higher. It is seen that when our error estimates go down below 10-6,
the in vivo estimates does not respond, a result very likely due to an error
background in this range. In the case of tRNAGlu reading G in second and U
or C in third position (Zhang et al., 2015), the overall accuracy is in a range
below 1/3 000. So the proofreading remedies those initial selection errors to
a tolerable level.
40
Conclusions
Like the children of moles, which always inherit the skills to dig holes, the
children of humans inherit the merits and demerits of their parents. Like
parent, like offspring. An implication of these proverbs of Chinese and English origin is that the accuracy of genetic code translation is an essential feature of life. The accuracy of messenger RNA translation is high in the living
cell and depends on two consecutive steps: initial reading of the genetic code
by transfer RNA followed by transfer RNA guided proofreading of the initial
code reading. This thesis is about speed and accuracy of genetic code translation.
In brief, we have used biochemical experiments to demonstrate a linear
trade-off between the speed and accuracy of messenger RNA translation on
the bacterial ribosome (Papers I and II). We have clarified the mechanism by
which antibiotic drugs of the aminoglycoside type decrease the accuracy of
messenger RNA translation (Paper III). We have shown that the accuracy
amplification by proofreading is essential to remedy error hot spots in initial
readings of messenger RNA (Paper IV).
Over the past six decades scientists have been brooding about how gene
expression to protein by messenger RNA transcription and translation can
have both high speed and high accuracy. Speculative explanations have been
given, but hard experimental evidence on how the speed-accuracy trade-off
is resolved in living cells has been hard to come by.
Here, quench-flow and other techniques have been used to monitor the
speed of correct messenger RNA reading performed at different Mg2+ ion
concentrations to tune the level of misreading (Paper I). Plots of correct initial reading speed versus accuracy of initial genetic code reading, defined as
the inverse of the misreading frequency, were used to estimate the maximal
reading accuracy at zero reading speed and the maximal reading speed in the
low accuracy limit.
We studied the speed-accuracy trade-off for initial messenger RNA reading by seven transfer RNAs. These experiments involved fourteen correct
genetic code letter readings and fifty six of the most common misreadings by
these transfer RNAs, and covered about 15% of all correct and incorrect
messenger RNA readings. There was a 400-fold variation in the accuracy of
initial code reading, in which we identified distinct error hot spots. We suggested that such error hot spots forced Nature to evolve a proofreading step
for genetic code translation (Paper II).
41
We subjected the effects of aminoglycoside antibiotics on initial code
reading to an in depth biochemical study (Paper III). We identified two fundamental reading steps. First, transfer RNA binds to the ribosome in complex with its protein factor in a state that is blind for the genetic message.
Then, transfer RNA moves to a reading activated state. We found that Mg2+
ions shift the equilibrium from free transfer RNA equally to the blind and
reading activated states. Aminoglycosides, in contrast, only shift the equilibrium from the reading blind to the reading activated state. These findings
have vast implications for the mechanism of initial code reading as well as
for the mechanism by which aminoglycosides corrupt the accuracy of genetic code translation to protein (Paper III).
We calibrated the total genetic code reading accuracy as obtained in our
biochemical assay to that in the living bacterial cell. It follows from our data
that the error frequency of genetic code reading varies by five orders of
magnitude within a subset of three transfer RNA types. As predicted (Paper
II), the potentially deleterious error hot spots in initial codon reading are
neutralized by enhanced proofreading (Paper IV).
42
Summary in Swedish
Precis som råttors avkomma som alltid gräver hål, så även med människor,
söner ärver sina fäders egenskaper, både bra och dåliga. Sådan far sådan son.
Både det kinesiska och det svenska ordspråket antyder att noggrannheten i
överföringen av informationen i den genetiska koden är viktig i naturen. I
den här avhandlingen undersöks noggrannheten i mRNA translationen på
djupet, både initialselektionen och korrekturläsningssteget.
Direkta mätningar av den operativa diskrimineringskonstanten i initialselektionen hur hastigheten i kodontranslationen minskar när noggrannheten
går mot sitt maximala värde (Artikel I och II). Aminoglykosider kan minska
den operationella, men inte den verkliga, diskrimineringskonstanten (Artikel
III). Förstärkningen av noggrannheten genom korrekturläsningssteget är
speciellt viktig för att hantera de ”hot spots” där noggrannheten i initialselektionen är särskilt låg (Paper IV).
Under de senaste 60 åren har forskare försökt förstå hur levande celler
kan uppnå både hög noggrannhet och hög hastighet när de uttrycker sina
gener. En modell för både hastigheten och noggrannheten i bakteriers proteinsyntes publicerades i tidskriften Current opinion in microbiology år 2008
men experimentellt stöd för denna hypotes saknas fortfarande.
Med hjälp av både quench-flow och manuella tekniker upptäckte vi att
genom att variera koncentrationen av Mg2+ joner så kunde vi demonstrera
det linjära hastighet-noggrannhet sambandet experimentellt (Paper I). Från
de här linjära sambanden kan vi få fram både de operationella diskrimineringskonstanterna, noggrannheten när hastigheten med kognat substrat är
noll, och associationshastighetskonstanten för kognat substrat när noggrannheten är 1. Dessa samband bidrar även med en visualisering av hur hastigheten minskar när noggrannheten ökar.
Genom mätningar av hastigheten och noggrannheten hos 7 kognata, 7
wobble, och 56 när-kognata kodon-antikodon par, motsvarande ungefär 15
% av den genetiska koden, upptäckte vi en stor variation (~400 ggr) i noggrannheten för initialselektionen av aminoacyl-tRNA av den bakteriella ribosomen (Artikel II). Vi identifierade även flera fel ”hot spots” där noggrannheten är särskilt låg. Vi föreslår att korrekturläsningssteget har utvecklats för att korrigera för dessa ”hot spots”.
Vi har undersökt denna operationella diskriminering mellan rätt och fel
tRNA i initialselektionen med hjälp av aminoglykosid antibiotika för att få
vidare förståelse för hur den fungerar i levande celler (Artikel III). Det visar
43
sig att det finns två trippelkomplexbundna ribosomala stadier. Först ett stadie
som inte är selektivt med avseende på kodon-antikodon interaktionen. Sen
ett stadie som är selektivt med avseende på kodon-antikodon interaktionen.
Mg2+ joner skiftar jämvikten mellan fritt trippelkomplex och det första
bundna stadiet mot det bundna stadiet men lämnar det andra stadiet ostört.
Aminoglykosidantibiotika skiftar jämvikten mellan de två bundna stadier
mot det selektiva stadiet men lämnar jämvikten mellan fritt och bundet
trippelkomplex ostört.
Genom att jämföra våra mätningar in vitro med in vivo data fann vi att
ungefär 50% av både det operationella diskrimineringsvärdet och den kognata hastigheten utnyttjas av levande celler (Artikel IV). Korrekturläsningen
är viktig för att kompensera för den oanvända delen av den operationella
diskrimineringskonstanten för att minimera skada på det bakteriella proteomet från felläsning av den genetiska koden under proteinsyntesen, speciellt
vid fel ”hot spots” i initialselektionen.
44
Acknowledgements
During the last six years of my PhD studies, many people were there for me
when I studied here, when I worked out scientific problems, when I needed
friends and support, and when I laughed and grieved. They brought something
wonderful to my life and without them my life in Sweden would be significantly different. I would like to take this opportunity to thank all of them.
First and foremost I want to express my very sincere and special appreciation to my main supervisor, Måns Ehrenberg. His professional knowledge,
excelsior scientific attitude, inclusive mind and high prestige have created
such an inspiring academic environment. Not only did it give me space to
learn and grow, but provided me with the chance to communicate with some
of the world’s top scientists. All of these make him to be one of the best
supervisors I have ever had. I thank him for giving me the opportunity of
being his student, for trusting me gradually and for his support at any moment. Without his guidance and constant feedback this Ph.D. would not have
been achievable.
I’d also like to give a heartfelt, special thanks to my co-supervisor, Suparna
Sanyal, who offered help of all sorts, useful tips with experimental design
and interesting talks. Great thanks to my co-supervisor, Michael Y. Pavlov,
for equation setup and insightful comments on drafts. I would like to thank
Professor Anthony C. Forster too, for the useful tips on presentation.
My thanks also go out to all the former and present members in the Molecular Biology program. I am especially grateful to Magnus Johansson
who shared his experimental knowledge when I was a newcomer to our lab,
and gave me lots of useful comments on manuscripts. He witnessed my
growth in scientific research. I am also very grateful to Ka-Weng Ieong for
collaboration. Anneli Borg, thank you for chipper laughter. Thanks to Harriet Mellenius for dissertation party information. Gabriele Indrisiunaite,
we had many scientific discussions. Onur Ercan’s jokes lightened my daily
work. Ajith Harish shared his career experience. I can’t forget Raymond
Fowler who was behind me every time when I need components and chemical orders. Chandra Sekhar Mandava, a good-tempered person. Ravi
Kiran Koripella’s introduction of Indian language helped me to learn about
another ancient country. Xueliang Ge, the nicest and most patient roommate
45
I have ever had. Ram Gopal Nitharwal interpreted some knowledge of
tRNA synthetase for me. Yanhong Pang, her joy, happiness and laugher
kept all throughout our chats. A lot of thanks to Mikael Holm, I got many
benefits after our every discussion about projects, and he gave me a big hand
in the end of writing my thesis. Petar Kovachev, gave me his jokes. Gürkan Korkmaz, gave me a chance to experience X-ray in akademiska sjukhuset together. I couldn’t forget Matteo Libero Baroni’s very special good
food too. And I also remembered it was Kristin Peisker told me how to prepare ATP solution, and I had a very interesting talk with Sunanda Chatterjee. Ranjeet Kumar always welcomed me with nice juice when I went to
his apartment. There were many scientific questions in Neelanjan Vishnu’s
brain which led me to wonder his cranial capacity. Marek Kwiatkowski’s
chemistry knowledge was helpful to understand my work. Josefine Liljeruhm shared her synthetic biology book. Sofia Pytharopoulou, we shared
the same bench. I knew more about American from Tyson Shepherd’s talking. Jinfan Wang, our planting and competing on HPLC made me cherish
my working time more than before. Thank all of you for coloring my working days. I grew more and learned more every day with all of you working in
the same lab. I also give my thanks to the reliable Chinese friends who
made my life in Sweden much more convenient and entertaining.
Leena Tirkkonen, my Mom in Sweden. Thanks for giving me a room to
live when I first arrived in Sweden, for driving me out to feel the country
scenes of Sweden, and for everything we experienced together.
I would also like to say a heartfelt thank you to my parents, for helping in
whatever way they could during this challenging period. My hard-working
parents have sacrificed their lives for myself and provided unconditional
love and care. I love them so much, and I would not have made it this far
without them. A very special thank you to Mr. Xuejun Zhang and Mrs.
Xiaojuan Gao for allowing me to take care of their little princess, for always being so supportive of our love. I am thankful for it and appreciate it.
There are so many very special people who have touched my life in past six
years, but there is one that stands out – my lovely princess, Bo. She saw
something in me that I didn’t know was there; she believed in me until I
learned to believe in myself; she brought me to sense wonderful parts of the
world and explored something which would happen in the future; she accompanied me in laughter and facing my fears; she gave me a colorful life
and she brightened my days and soul. There were many possibilities for us to
miss each other, while I’m lucky that I could find her crown and she would
like to share her life with me. Thank you, my princess, for being in my life.
And finally to myself, thankful for me, never give up, ever.
46
References
Antoun, A., Pavlov, M.Y., Lovmar, M., and Ehrenberg, M. (2006a). How initiation
factors maximize the accuracy of tRNA selection in initiation of bacterial
protein synthesis. Mol Cell 23, 183-193.
Antoun, A., Pavlov, M.Y., Lovmar, M., and Ehrenberg, M. (2006b). How initiation
factors tune the rate of initiation of protein synthesis in bacteria. Embo J 25,
2539-2550.
Baldwin, A.N., and Berg, P. (1966). Transfer Ribonucleic Acid-Induced Hydrolysis
of Valyladenylate Bound to Isoleucyl Ribonucleic Acid Synthetase. J Biol
Chem 241, 839-&.
Bouadloun, F., Donner, D., and Kurland, C.G. (1983). Codon-specific missense
errors in vivo. Embo J 2, 1351-1356.
Carter, A.P., Clemons, W.M., Brodersen, D.E., Morgan-Warren, R.J., Wimberly,
B.T., and Ramakrishnan, V. (2000). Functional insights from the structure of the
30S ribosomal subunit and its interactions with antibiotics. Nature 407, 340-348.
Daviter, T., Wieden, H.J., and Rodnina, M.V. (2003). Essential role of histidine 84
in elongation factor Tu for the chemical step of GTP hydrolysis on the
ribosome. J Mol Biol 332, 689-699.
Dong, H.J., Nilsson, L., and Kurland, C.G. (1996). Co-variation of tRNA abundance
and codon usage in Escherichia coli at different growth rates. J Mol Biol 260,
649-663.
Edelmann, P., and Gallant, J. (1977). Mistranslation in E. coli. Cell 10, 131-137.
Ehrenberg, M., and Blomberg, C. (1980). Thermodynamic Constraints on Kinetic
Proofreading in Biosynthetic Pathways. Biophys J 31, 333-358.
Ehrenberg, M., and Kurland, C.G. (1984). Costs of Accuracy Determined by a
Maximal Growth-Rate Constraint. Quarterly reviews of biophysics 17, 45-82.
Ermolenko, D.N., Majumdar, Z.K., Hickerson, R.P., Spiegel, P.C., Clegg, R.M., and
Noller, H.F. (2007). Observation of intersubunit movement of the ribosome in
solution using FRET. J Mol Biol 370, 530-540.
Fersht, A.R. (1977). Editing Mechanisms in Protein-Synthesis - Rejection of Valine
by Isoleucyl-Transfer-Rna Synthetase. Biochemistry-Us 16, 1025-1030.
Fersht, A.R. (1979). Editing mechanisms in the aminoacylation of tRNA. Cold
Spring Harbor Monograph Archive 9, 247-254.
Fersht, A.R., and Dingwall, C. (1979). Evidence for the Double-Sieve Editing
Mechanism in Protein-Synthesis - Steric Exclusion of Isoleucine by ValylTransfer Rna-Synthetases. Biochemistry-Us 18, 2627-2631.
Fersht, A.R., and Kaethner, M.M. (1976). Mechanism of Aminoacylation of
Transfer-Rna - Proof of Aminoacyl Adenylate Pathway for Isoleucyl-TransferRna and Tyrosyl-T-Rna Synthetases from Escherichia-Coli-K12. BiochemistryUs 15, 818-823.
47
Gao, H.X., Zhou, Z.H., Rawat, U., Huang, C., Bouakaz, L., Wang, C.H., Cheng,
Z.H., Liu, Y.Y., Zavialov, A., Gursky, R., et al. (2007). RF3 induces ribosomal
conformational changes responsible for dissociation of class I release factors.
Cell 129, 929-941.
Gromadski, K.B., Daviter, T., and Rodnina, M.V. (2006). A uniform response to
mismatches in codon-anticodon complexes ensures ribosomal fidelity. Mol Cell
21, 369-377.
Gromadski, K.B., and Rodnina, M.V. (2004). Kinetic determinants of high-fidelity
tRNA discrimination on the ribosome. Mol Cell 13, 191-200.
Hopfield, J.J. (1974). Kinetic proofreading: a new mechanism for reducing errors in
biosynthetic processes requiring high specificity. Proceedings of the National
Academy of Sciences of the United States of America 71, 4135-4139.
Hopfield, J.J., and Yamane, T. (1976). Experimental-Evidence for Kinetic
Proofreading in Aminoacylation of Transfer-Rna-Ile
Biophys J 16, A16-A16.
Johansson, M., Bouakaz, E., Lovmar, M., and Ehrenberg, M. (2008). The kinetics of
ribosomal peptidyl transfer revisited. Mol Cell 30, 589-598.
Karimi, R., Pavlov, M.Y., Buckingham, R.H., and Ehrenberg, M. (1999). Novel
roles for classical factors at the interface between translation termination and
initiation. Mol Cell 3, 601-609.
Kramer, E.B., and Farabaugh, P.J. (2007). The frequency of translational misreading
errors in E. coli is largely determined by tRNA competition. Rna 13, 87-96.
Kunkel, T.A., and Bebenek, K. (2000). DNA replication fidelity. Annu Rev
Biochem 69, 497-529.
Laughrea, M., Latulippe, J., Filion, A.M., and Boulet, L. (1987). Mistranslation in
12 Escherichia-Coli Ribosomal-Proteins - Cysteine Misincorporation at Neutral
Amino-Acid-Residues Other Than Tryptophan. Eur J Biochem 169, 59-64.
Loftfield, R.B., Pastuszy.A, and Lovgren, T. (1973). Mechanism of Aminoacylation
of Transfer-Rna. Fed Proc 32, 460-&.
Loftfield, R.B., and Vanderjagt, D. (1972). The frequency of errors in protein
biosynthesis. The Biochemical journal 128, 1353-1356.
Lynch, S.R., Gonzalez, R.L., and Puglisi, J.D. (2003). Comparison of x-ray crystal
structure of the 30S subunit-antibiotic complex with NMR structure of decoding
site oligonucleotide-paromomycin complex. Structure 11, 43-53.
Manickam, N., Nag, N., Abbasi, A., Patel, K., and Farabaugh, P.J. (2014). Studies of
translational misreading in vivo show that the ribosome very efficiently
discriminates against most potential errors. Rna 20, 9-15.
Ninio, J. (1974). Semi-Quantitative Treatment of Missense and Nonsense
Suppression in Stra and Ram Ribosomal Mutants of Escherichia-Coli Evaluation of Some Molecular Parameters of Translation Invivo. J Mol Biol 84,
297-313.
Ninio, J. (1975). Kinetic Amplification of Enzyme Discrimination. Biochimie 57,
587-595.
Ogle, J.M., and Ramakrishnan, V. (2005). Structural insights into translational
fidelity. Annu Rev Biochem 74, 129-177.
Ozaki, M., Mizushim.S, and Nomura, M. (1969). Identification and Functional
Characterization of Protein Controlled by Streptomycin-Resistant Locus in E
Coli. Nature 222, 333-&.
Pape, T., Wintermeyer, W., and Rodnina, M.V. (2000). Conformational switch in
the decoding region of 16S rRNA during aminoacyl-tRNA selection on the
ribosome. Nat Struct Biol 7, 104-107.
48
Pauling, L. (1957). The Probability of Errors in the Process of Synthesis of Protein
Molecules (Birkhäuse, Basel), pp597-602.
Peske, F., Rodnina, M.V., and Wintermeyer, W. (2005). Sequence of steps in
ribosome recycling as defined by kinetic analysis. Mol Cell 18, 403-412.
Rodnina, M.V., Fricke, R., and Wintermeyer, W. (1994). Transient Conformational
States of Aminoacyl-Transfer-Rna during Ribosome Binding Catalyzed by
Elongation-Factor Tu. Biochemistry-Us 33, 12267-12275.
Rodnina, M.V., Pape, T., Fricke, R., Kuhn, L., and Wintermeyer, W. (1996). Initial
binding of the elongation factor Tu•GTP•aminoacyl-tRNA complex preceding
codon recognition on the ribosome. J Biol Chem 271, 646-652.
Rodnina, M.V., Savelsbergh, A., Katunin, V.I., and Wintermeyer, W. (1997).
Hydrolysis of GTP by elongation factor G drives tRNA movement on the
ribosome. Nature 385, 37-41.
Rosenberger, R.F., and Foskett, G. (1981). An estimate of the frequency of in vivo
transcriptional errors at a nonsense codon in Escherichia coli. Mol Gen Genet
183, 561-563.
Ruusala, T., Ehrenberg, M., and Kurland, C.G. (1982). Is There Proofreading during
Polypeptide-Synthesis. Embo J 1, 741-745.
Schmeing, T.M., Voorhees, R.M., Kelley, A.C., Gao, Y.G., Murphy, F.V., Weir,
J.R., and Ramakrishnan, V. (2009). The Crystal Structure of the Ribosome
Bound to EF-Tu and Aminoacyl-tRNA. Science 326, 688-694.
Scolnick, E., Tompkins, R., Caskey, T., and Nirenber.M (1968). Release Factors
Differing in Specificity for Terminator Codons. P Natl Acad Sci USA 61, 768&.
Spiegel, P.C., Ermolenko, D.N., and Noller, H.F. (2007). Elongation factor G
stabilizes the hybrid-state conformation of the 70S ribosome. Rna 13, 14731482.
Thompson, R.C., and Stone, P.J. (1977). Proofreading of Codon-Anticodon
Interaction on Ribosomes. P Natl Acad Sci USA 74, 198-202.
Yonath, A. (2005). Antibiotics targeting ribosomes: Resistance, selectivity,
synergism, and cellular regulation. Annu Rev Biochem 74, 649-679.
Zaher, H.S., and Green, R. (2009). Fidelity at the molecular level: lessons from
protein synthesis. Cell 136, 746-762.
Zavialov, A.V., Buckingham, R.H., and Ehrenberg, M. (2001). A posttermination
ribosomal complex is the guanine nucleotide exchange factor for peptide release
factor RF3. Cell 107, 115-124.
Zavialov, A.V., Hauryliuk, V.V., and Ehrenberg, M. (2005). Splitting of the
posttermination ribosome into subunits by the concerted action of RRF and EFG. Mol Cell 18, 675-686.
Zavialov, A.V., Mora, L., Buckingham, R.H., and Ehrenberg, M. (2002). Release of
peptide promoted by the GGQ motif of class 1 release factors regulates the
GTPase activity of RF3. Mol Cell 10, 789-798.
Zhang, J., Ieong, K.W., Johansson, M., and Ehrenberg, M. (2015). Accuracy of
initial codon selection by aminoacyl-tRNAs on the mRNA-programmed
bacterial ribosome. Proceedings of the National Academy of Sciences of the
United States of America 112, 9602-9607.
49
Acta Universitatis Upsaliensis
Digital Comprehensive Summaries of Uppsala Dissertations
from the Faculty of Science and Technology 1306
Editor: The Dean of the Faculty of Science and Technology
A doctoral dissertation from the Faculty of Science and
Technology, Uppsala University, is usually a summary of a
number of papers. A few copies of the complete dissertation
are kept at major Swedish research libraries, while the
summary alone is distributed internationally through
the series Digital Comprehensive Summaries of Uppsala
Dissertations from the Faculty of Science and Technology.
(Prior to January, 2005, the series was published under the
title “Comprehensive Summaries of Uppsala Dissertations
from the Faculty of Science and Technology”.)
Distribution: publications.uu.se
urn:nbn:se:uu:diva-262901
ACTA
UNIVERSITATIS
UPSALIENSIS
UPPSALA
2015