Anda di halaman 1dari 18

Article

Heterogeneity of Stop Codon Readthrough in Single


Bacterial Cells and Implications for Population
Fitness
Graphical Abstract Authors
Yongqiang Fan, Christopher R. Evans,
Karl W. Barber, ..., Oleg A. Igoshin,
Jesse Rinehart, Jiqiang Ling

Correspondence
jiqiang.ling@uth.tmc.edu

In Brief
Protein synthesis accuracy is important
for cell physiology and has been mostly
studied at the population level. Fan et al.
develop a dual-reporter system to
quantitate protein synthesis errors in
single bacterial cells and show that
slowing protein synthesis increases some
translational errors, namely readthrough
of stop codons.

Highlights
d Stop codon readthrough is heterogeneous among single cells

d Increased UGA readthrough promotes bacterial growth

d Reducing protein synthesis levels enhances UGA


readthrough

d Fluctuations of translational components lead to UGA


readthrough heterogeneity

Fan et al., 2017, Molecular Cell 67, 826836


September 7, 2017 2017 Elsevier Inc.
http://dx.doi.org/10.1016/j.molcel.2017.07.010
Molecular Cell

Article

Heterogeneity of Stop Codon Readthrough


in Single Bacterial Cells
and Implications for Population Fitness
Yongqiang Fan,1,7 Christopher R. Evans,1,2,7 Karl W. Barber,3,4,7 Kinshuk Banerjee,5 Kalyn J. Weiss,1,2 William Margolin,1,2
Oleg A. Igoshin,5,6 Jesse Rinehart,3,4 and Jiqiang Ling1,2,8,*
1Department of Microbiology and Molecular Genetics, McGovern Medical School, University of Texas Health Science Center, Houston,

TX 77030, USA
2Graduate School of Biomedical Sciences, Houston, TX 77030, USA
3Department of Cellular & Molecular Physiology, Yale University, New Haven, CT 06520, USA
4Systems Biology Institute, Yale University, West Haven, CT 06516, USA
5Center for Theoretical Biological Physics, Rice University, Houston, TX 77005, USA
6Department of Bioengineering, Rice University, Houston, TX 77005, USA
7These authors contributed equally
8Lead Contact

*Correspondence: jiqiang.ling@uth.tmc.edu
http://dx.doi.org/10.1016/j.molcel.2017.07.010

SUMMARY codons by proper aa-tRNAs, and accurate translocation of


mRNA by the ribosome (Fredrick and Noller, 2002; Ling et al.,
Gene expression noise (heterogeneity) leads to 2009; Liu et al., 2015; Mascarenhas et al., 2008; Zaher and
phenotypic diversity among isogenic individual cells. Green, 2009). High translational errors cause growth defects
Our current understanding of gene expression noise and cell death in bacteria (Bacher et al., 2005; Davis, 1987; Kar-
is mostly limited to transcription, as separating trans- khanis et al., 2007), mitochondrial dysfunction in yeast (Reynolds
lational noise from transcriptional noise has been et al., 2010), shortened lifespan in flies (Lu et al., 2014), and neu-
rodegeneration and cardioproteinopathy in mammals (Lee et al.,
challenging. It also remains unclear how translational
2006; Liu et al., 2014). Surprisingly, naturally isolated Escherichia
heterogeneity originates. Using a transcription-
coli strains display a wide range of ribosomal fidelity, suggesting
normalized reporter system, we discovered that that high translational errors may be favored under some natural
stop codon readthrough is heterogeneous among habitats (Mikkola and Kurland, 1992). Recent evidence suggests
single cells, and individual cells with higher UGA that increased translational errors paradoxically provides bene-
readthrough grow faster from stationary phase. Our fits to microorganisms under certain stress conditions (Bezerra
work also revealed that individual cells with lower et al., 2013; Ling et al., 2015; Pan, 2013; Ribas de Pouplana
protein synthesis levels exhibited higher UGA et al., 2014). For example, amino acid misincorporation in the
readthrough, which was confirmed with ribosome- b subunit of RNA polymerase increases resistance of mycobac-
targeting antibiotics (e.g., chloramphenicol). Further teria to rifampicin (Javid et al., 2014; Su et al., 2016), and trans-
experiments and mathematical modeling suggest lational errors improve bacterial tolerance to oxidative stress by
activating the general stress response (Fan et al., 2015; Fredriks-
that varied competition between ternary complexes
son et al., 2007). Interestingly, in such cases only subpopulations
and release factors perturbs the UGA readthrough
of genetically identical cells survive severe stresses (Fan et al.,
level. Our results indicate that fluctuations in the con- 2015; Su et al., 2016), suggesting that stress response activated
centrations of translational components lead to UGA by translational errors may be heterogeneous (noisy) in individual
readthrough heterogeneity among single cells, which cells. However, the noise levels of translational errors have not
enhances phenotypic diversity of the genetically been determined, and how such heterogeneity originates re-
identical population and facilitates its adaptation to mains unknown.
changing environments. Gene expression has been shown to be stochastic and noisy
(Balazsi et al., 2011; Kaern et al., 2005; Levine and Hwa, 2007;
Levine et al., 2013; Newman et al., 2006; Sanchez and Golding,
INTRODUCTION 2013; Taniguchi et al., 2010; Young et al., 2013). The first exper-
imental evidence came from pioneering work by Elowitz et al.
Protein synthesis is a fundamental and essential process in all (2002) showing that transcription is intrinsically noisy. Later
three domains of life. Accurate protein synthesis requires correct studies revealed that transcription is bursty and noncontinuous
matching of amino acids and transfer RNAs (tRNAs) by amino- (Blake et al., 2003; Levine et al., 2013; Sanchez and Golding,
acyl-tRNA synthetases, proofreading of aminoacyl-tRNAs (aa- 2013), and the promoter architecture regulates transcriptional
tRNAs) by trans-editing factors, precise decoding of mRNA noise (Sanchez et al., 2013). Noise in gene expression can be

826 Molecular Cell 67, 826836, September 7, 2017 2017 Elsevier Inc.
Figure 1. Dual-Fluorescent Reporters for
Translational Errors
(A) Dual-fluorescence reporters that measure
translational errors. MCherry (m) and yfp (y) genes
are fused under the control of a constitutive pro-
moter PLtetO-1. At the end of mCherry is a stop
codon, a frameshifting (fs) codon, or no stop
codon. Readthrough of the stop codon or frame-
shifting produces a single mCherry-YFP fusion
protein and yields YFP signal.
(B) The reporters on a low-copy-number plasmid
were expressed for 24 hr in Luria-Bertani broth
(LB) at 37! C in wild-type (MG1655) and ribosomal
error-prone (rpsD*) E. coli strains. Fluorescence
was quantified by spectrometry on a plate reader.
The error rate was calculated as the YFP/mCherry
ratio of the error reporter normalized by the YFP/
mCherry ratio of the m-y control.
(C) Western blotting showing that UGA read-
through of the m-TGA-y reporter. Anti-mCherry
antibody was used to detect both mCherry-YFP
fusion and mCherry. Data are represented as
mean SD.
(D and E) Label-free quantitative mass spectrom-
etry analysis of UGA readthrough in the dual-
fluorescence reporter protein m-TGA-y (D) and
native protein RpsG (E). Extracted ion chromato-
grams (XIC) for the m-TGA-y reporter peptide
WLQTSAGEAAAK (W, or tryptophan, marks the
UGA readthrough site in the peptide) and the
RpsG peptide QPALGYLNWTPK are shown from
two different strains. XIC peak area indicates the
relative abundance of the detected peptide.
Relative peak intensities were fixed at 100% indi-
vidually for comparative purposes.

harmful by disrupting regulatory networks, but increasing evi- motes growth of E. coli cells from stationary phase, indicating
dence shows that such noise can also be beneficial by gener- that UGA readthrough heterogeneity provides an advantage to
ating phenotypic heterogeneity that helps the population to the population during environmental shifts.
quickly adapt to environmental changes through bet-hedging
(Ackermann, 2015; Blake et al., 2006; Sanchez and Golding, RESULTS
2013). Compared with transcriptional noise, translational noise
is extremely poorly understood. Earlier studies used a single Dual-Fluorescence Reporters to Quantitate
fluorescent reporter to determine the noise levels when transla- Translational Errors
tion initiation (Ozbudak et al., 2002) or the codon context (Blake To separate the noise of translational errors from transcriptional
et al., 2003) is varied. It was suggested that in Bacillus subtilis, noise, we have constructed a series of mCherry-YFP (red and
increasing the efficiency of translation initiation enhances trans- yellow fluorescent proteins) fusion reporters that measure the
lational noise (Ozbudak et al., 2002). However, it remains a errors of stop-codon readthrough and frameshifting (Figure 1).
significant challenge to separate translational noise from tran- The mCherry and yfp genes are transcribed as a single mRNA
scriptional noise (Paulsson, 2004). Here, we have developed and translated from the same start codon to yield a single poly-
transcription-normalized dual-fluorescence reporters to quanti- peptide (Figure 1A), therefore minimizing the influence of the
tate the noise levels of stop codon readthrough and investigate noise from transcription and translation initiation in single-cell
the sources of such noise. Our work suggests that reduced analysis. Additionally, both mCherry and YFP are stable in
translation promotes UGA readthrough by altering competition E. coli (Figures S1A and S1B), and degradation of the reporter
between the ternary complex (TC) and release factors and pro- proteins should produce a minimal level of noise. Expression of
vides a model for the heterogeneity of UGA readthrough among the dual-fluorescence reporter on a low-copy-number plasmid
single cells. We also show that increased UGA readthrough pro- does not affect bacterial growth (Figure S1C). With these

Molecular Cell 67, 826836, September 7, 2017 827


interesting to note that for 2D, you see that, even though the units are arbitrary, assuming they are
comparable, there is a lot more M than Y. also note that we could still have R2 value close to one
even in this condition, but a low R value tells us that the read-through levels VARY between individual
cell, which is why the relative level of red and yellow change from cell to cell.

Figure 2. UGA Readthrough Is Heterogeneous among Single Cells


(A and B) YFP and mCherry fluorescence in MG1655 cells carrying either the m-y (A) or m-TGA-y (B) reporter. Cells were grown in LB for 24 hr. The YFP/mCherry
ratio is the relative YFP signal normalized by the mCherry signal. Cells with high and low UGA readthrough are indicated by yellow and red arrows, respectively.
(C and D) In the scatterplots of m-y (C) and m-TGA-y (D), each dot represents a single cell. The fluorescence is background-subtracted and arbitrary units
are shown.
(E and F) Heterogeneity of UGA readthrough among single cells is indicated by CV (E) and noise (F) of the YFP/mCherry ratio. m, the mean of the YFP/mCherry
ratio; s, SD. Data are represented as mean SD. AU, arbitrary units.

reporters, 0.2%3% error rates were detected in wild-type (WT) porter was increased 3-fold in the rpsD* strain compared to the
E. coli MG1655 grown in Luria-Bertani broth (LB) at 37! C using WT (Figure 1D), in line with the fluorescence and western blotting
fluorescence spectrometry (Figures 1B and S1D). Such error results (Figures 1B and 1C). These results demonstrate that our
rates were orders of magnitude higher than DNA mutational dual-fluorescence reporter system is robust for quantitation of
and transcriptional error rates (Gordon et al., 2015; Rosenberg translational errors.
et al., 2012), suggesting that the observed errors directly result
from translation. In support of this notion, a ribosomal ambiguity UGA Readthrough Is Heterogeneous among Single Cells
mutation (rpsD I199N or rpsD*), which affects the ribosomal de- We next applied our dual-fluorescence reporters to determine
coding center to reduce translational fidelity (Bjorkman et al., the heterogeneity of stop codon readthrough among genetically
1999; Fan et al., 2015; Kramer and Farabaugh, 2007), increased identical single cells grown under the same condition. Using fluo-
all four types of errors that we tested (Figure 1B). rescence microscopy, we visualized mCherry and YFP signals in
To validate that the fluorescence of the reporters reflects the single cells (Figures 2 and S1E). Whereas the YFP signal of the
actual protein level, we used western blotting to detect and m-TGA-y reporter was substantially higher than background,
quantitate mCherry and mCherry-YFP fusion protein (Figure 1C). cells without reporters showed no detectable fluorescence
In WT cells expressing the m-TGA-y reporter, the mCherry-YFP signal (Figure S1E), suggesting that the influence of auto fluores-
fusion accounted for 2.5% of total mCherry proteins produced. cence on signal quantification is negligible under our experi-
This was in good agreement with the 2% UGA readthrough level mental setting. Our results showed that the control m-y reporter
determined by fluorescence spectrometry (Figure 1B) and exhibited tight linear correlation between the YFP and mCherry
consistent with previous reports (Devaraj et al., 2009; Eggerts- signals, and the YFP/mCherry ratio was mostly homogeneous
son and Soll, 1988). The fraction of mCherry-YFP in the rpsD* (Figures 2A, 2C, and S3A). In contrast, the YFP/mCherry ratio
strain increased to 8%, which was again consistent with the of the m-TGA-y reporter (indicating UGA readthrough level)
UGA readthrough level determined by fluorescence (Figure 1B). was more dispersed (Figures 2B, 2D, and S3B), suggesting
To identify the amino acid(s) incorporated at UGA sites as well that the concentration of UGA readthrough products was hetero-
as to detect UGA readthrough events in the native proteome, we geneous among single cells.
further applied quantitative mass spectrometry of WT and rpsD* Heterogeneity of gene expression is quantified with coefficient
cells with and without the m-TGA-y reporter (Figures 1D, 1E, and of variation (CV), calculated as the ratio of the SDn (s) over the
S2; Table S1). UGA readthrough in the m-TGA-y reporter and mean (m) or noise (s2/m2) (Blake et al., 2003; Elowitz et al.,
native proteins were indeed detected with liquid chromatog- 2002; Taniguchi et al., 2010). Our results revealed that the CV
raphy-tandem mass spectrometry (LC-MS/MS). Our unbiased of YFP/mCherry ratio in cells with the control m-y reporter was
MS/MS search shows that tryptophan (W) is the predominant "0.1 (Figure 2E). Such a low level of heterogeneity may be
amino acid that suppresses UGA codons in E. coli. The level of caused by noise from fluorophore maturation, partial degrada-
the peptide resulting from UGA readthrough in the m-TGA-y re- tion of mRNA and protein, ribosomal drop-off during translation,

828 Molecular Cell 67, 826836, September 7, 2017


missense errors, and the imaging system. The CV of YFP/ codons are suppressed by Trp-tRNATrp (Figures 1D, 1E, and
mCherry ratio in cells with the m-TGA-y reporter increased to S2; Table S1) and terminated by release factor 2 (RF2) (Engel-
0.2, corresponding to a 4-fold increase in noise (Figures 2E berg-Kulka, 1981; Youngman et al., 2007). It has been reported
and 2F). It has been suggested that when the copy number of that in E. coli K-12 strains (including MG1655), RF2 contains
protein molecules is low (<1,000/cell), partitioning errors and an A246T mutation that decreases termination efficiency
finite-number effect would contribute to the overall protein noise, (Dincbas-Renqvist et al., 2000; Mora et al., 2007). To test
whereby reducing the fluorescence mean would increase the CV whether reduced protein synthesis specifically increases UGA
and noise (Huh and Paulsson, 2011; Kaern et al., 2005). How- readthrough in K-12 strains, we examined E. coli BL21 and
ever, the estimated concentration of mCherry-YFP fusion protein Salmonella enterica Serovar Typhimurium LT2 strains that both
in cells with the m-TGA-y reporter is "6,000 molecules per cell carry wild-type RF2. Like MG1655, both BL21 and S. Typhimu-
(see the STAR Methods). Therefore, partitioning noise and rium strains exhibited strong negative correlation between
finite-number effect is expected to have little contribution to UGA readthrough and mCherry intensity and increased UGA
the observed CV of the YFP/mCherry ratio. In support of this readthrough upon Chl treatment (Figure S5), demonstrating
argument, we found that reducing the mean of mCherry by that reduced protein synthesis enhances UGA readthrough inde-
decreasing the promoter strength of the m-TGA-y reporter did pendent of the A246T mutation.
not further increase the CV (Figure S3C). We thus reason that The next question is how reduced translation may increase
the higher CV and noise of the YFP/mCherry ratio in m-TGA-y UGA readthrough errors, which is surprising given previous ki-
compared to m-y is largely due to the heterogeneity of UGA netic studies showing that there is a trade-off between transla-
readthrough events among individual cells. tional speed and accuracy (Johansson et al., 2008). Using acidic
gel northern blotting, we show that the overall Trp-tRNATrp level
Reduced Protein Synthesis Increases UGA Readthrough increases in the presence of Chl due to an increase in total
To understand how the heterogeneity of UGA readthrough tRNATrp (Figure 4A). Further time course analysis shows that
arises, we analyzed the correlation between UGA readthrough tRNATrp is stable both in the presence and absence of Chl (Fig-
(indicated by the YFP/mCherry ratio) and various cell parameters ure 4B), indicating that the increased level of tRNATrp with Chl is
(Figures 3 and S3). We found that cells with lower mCherry not caused by increased tRNA stability, but rather by enhanced
expression levels exhibited higher UGA readthrough (Figures transcription. We next performed mathematical modeling to un-
3A and 3B). To test whether high UGA readthrough reduces derstand how reduced translation may increase UGA read-
mCherry expression or reduced protein expression increases through (Figures S4B and S4C; Table S2). Our modeling suggests
UGA readthrough, we first used a UGA suppressor tRNASer that increasing TC concentration enhances UGA readthrough
variant with a UCA anticodon (Figures S3G and S3H). The sup- rate. Based on these experimental and modeling results, we
pressor tRNA substantially increased UGA readthrough, but reason that reduced translation increases the effective concen-
did not affect the mCherry protein level. We next tested whether tration of EF-Tu:Trp-tRNATrp:GTP complex to more efficiently
reducing protein expression would enhance UGA readthrough. compete against RF2, thereby promoting UGA readthrough.
A low mCherry level may result from reduced transcription A recent study shows that Chl and Tet decrease the active
or translation. We demonstrated that reducing transcription fraction of ribosomes (Dai et al., 2016). To test the effects of
with rifampicin (Rif), which inhibits RNA polymerase, did not active ribosomes on UGA readthrough, we used mutant strains
significantly affect UGA readthrough (Figure S3I). In contrast, lacking several copies of the rrn operon that encodes ribosomal
treating E. coli with low concentrations of chloramphenicol RNAs (Quan et al., 2015). Deleting six out of seven copies of the
(Chl), which binds to the peptidyl transferase center (PTC) of rrn operon significantly increased UGA readthrough (Figure 3G),
the ribosome and impedes binding of aa-tRNAs and release fac- suggesting that reducing the cellular ribosome concentration
tors (Wilson, 2014), reduced protein synthesis and increased promotes UGA readthrough. We also tested a strain that
UGA readthrough, as determined by fluorescence spectrometry lacks RMF (ribosome modulation factor), which promotes 70S
and western blotting (Figures 3C and 3D). To verify that the ribosomes to form inactive dimers during stationary phase
observed increase in UGA readthrough by Chl was not an artifact and stress conditions (Polikanov et al., 2012; Yoshida and
due to the sequence context of the dual-fluorescence reporter, Wada, 2014). Deleting rmf is expected to increase the concen-
we tested UGA readthrough at a native chromosomal site of trations of active ribosomes. Indeed, deleting rmf mitigates the
the cspC gene (Figure 3E) and confirmed that Chl also signifi- effect of Chl to enhance UGA readthrough (Figure 3H), support-
cantly enhanced readthrough of the native UGA codon (Figures ing that Chl increases UGA readthrough by reducing the fraction
3E and 3F). To determine whether increased UGA readthrough of active ribosomes. It has been suggested that low codon adap-
is a specific effect of Chl, we further tested other ribosomal tation of highly transcribed genes (e.g., our m-TGA-y reporter,
inhibitors tetracycline (Tet, inhibiting aa-tRNA delivery to the expressed at over 300,000 protein molecules per cell) may also
ribosome), spectinomycin (Spc, inhibiting translocation), and sequester ribosomes and reduce the active fraction (Kudla
erythromycin (Ery, blocking the 50S exit tunnel). Tet, Spc, et al., 2009). In contrast, reduced translation initiation of the re-
and Ery also significantly enhanced UGA readthrough as Chl porter is not expected to affect the active ribosome concentra-
(Figure S4A). tion. We therefore reduced expression of the reporter proteins
The efficiency of stop codon readthrough is a direct result of by either introducing non-optimal codons to mCherry or
competition between TC (composed of elongation factor Tu decreasing translation initiation efficiency. Non-optimal codons
[EF-Tu], aa-tRNA, and GTP) and release factors. UGA stop enhanced UGA readthrough as Chl (Figure S4D), but reducing

Molecular Cell 67, 826836, September 7, 2017 829


Just see a) and b). this is an example of finite number effect, as you can see if the protein
concentration is high, there is very low fluctuation in the total abundance but if the protein
concentration is low, then the fluctuations start becoming big. side note - the abundance values
are the same between a and b, because in b) aside from decreasing the transcription rate by 100,
they also decreased the cell volume by 100, to keep the abundance the same, so that we can
compare similar values in a) and b) source - https://www.nature.com/nrg/journal/v6/n6/fig_tab/
nrg1615_F2.html

  



 
 '


  

'
 
&'
  

 
 
 




 




 #'$''!'
they did the E and F to check if the
increased UGA read-through was
because of something specific about the
sequence they were using. So they tried
to check UGA read-through in another
sequence which is always transcribed -
cspC, and still saw UGA read-through

Text

Text

Text

Text

Figure 3. Reducing Protein Synthesis Increases the Level of UGA Readthrough


(A) In single MG1655 cells, higher UGA readthrough correlates (Spearmans rank correlation) with lower mCherry protein level. The mCherry intensity is the
mCherry signal normalized by cell volume with arbitrary units. Relative UGA readthrough is calculated from the YFP/mCherry ratio. Experimental conditions were
the same as in Figure 2.
(B) The y axis (R10%) is the ratio of the average UGA readthrough in the bottom 10% of cells divided by the average UGA readthrough in the top 10% of cells ranked
by mCherry intensity from (A).
(C) MG1655 cells were grown in LB with and without Chl for 24 hr, and UGA readthrough and protein synthesis rates were determined by fluorescence spec-
trometry. Treating MG1655 with low doses of Chl decreases protein synthesis rate and enhances UGA readthrough.
(D) Western blotting confirms that Chl increases UGA readthrough of the m-TGA-y reporter.
(E and F) UGA readthrough of chromosomally encoded CspC increases in the presence of chloramphenicol.
(E) The nucleotide sequence of the 30 end of the chromosomal constitutively expressed cspC gene is shown. Red letters indicate the first and second stop
codons. An in-frame YFP tag is inserted immediately after the first TGA codon of the native cspC gene (CspC-TGA-YFP-1) or before the second stop codon
(CspC-TGA-YFP-2) in MG1655. Western blotting result shows that readthrough of the cspC UGA codon is significantly increased by addition of Chl.
(F) Quantitation of western blotting results in (E).
(G) Reducing ribosome copy number increases UGA readthrough. WT MG1655 and its ribosomal operon deletion mutants (D2, DrrnEG; D4, DrrnGBAD; D6,
DrrnGADBHC) were grown in LB with and without Chl for 24 hr, and UGA readthrough levels were determined with fluorescence spectrometry.
(H) Deleting rmf partially suppresses the effect of Chl to increase UGA readthrough.
(I) Chl treatment decreases the CV of UGA readthrough. Data are represented as mean SD. AU, arbitrary units.

translation initiation efficiency by changing the Shine-Dalgarno Next, we investigated how reduced translation affects the
sequence did not alter the UGA readthrough level (Figure S4E). heterogeneity of UGA readthrough. Both Chl and codon non-
These results are consistent with the notion that reducing the optimization decreases the CV of YFP/mCherry ratio (Figures
concentration of active ribosomes promotes UGA readthrough. 3I and S4F), suggesting that UGA readthrough heterogeneity is

830 Molecular Cell 67, 826836, September 7, 2017


Figure 4. Chloramphenicol Increases Trp-
tRNATrp Level
(A) Chl treatment increases the level of Trp-
tRNATrp shown by acidic northern blotting. Total
RNA was isolated from MG1655 cells grown in LB
with or without 2 mg/mL Chl under acidic condi-
tions andtreated with or without alkaline (OH#)
before acid gel electrophoresis and northern
blotting. Alkaline treatment causes deacylation of
aminoacyl-tRNAs. Almost 100% of tRNATrp was
aminoacylated without alkaline treatment. SsrA
was used as an internal standard to calculate the
relative concentration of Trp-tRNATrp.
(B) Stability of tRNATrp with and without Chl.
MG1655 cells grown in LB in the presence
absence of Chl were treated with a high concen-
tration of rifampicin to stop transcription. RNA
samples were prepared at indicated time points
following addition of Rif and subjected to northern
blotting analysis. Chl increases the overall level of
tRNATrp, but not the stability.
(C) The level of Gly-tRNAGly was determined using
acidic gel northern blotting as in Figure 4.
(D) The protein levels of EF-Tu and FLAG-RF2 with
and without Chl revealed by western blotting.
A FLAG tag is fused to the C terminus of RF2 at the
native chromosomal site. Data are represented as
mean SD.

reduced when translation is attenuated across the population. ing that the RF2 activity was no longer saturating in the absence
Collectively, these results suggest that reduced translation en- of RF3. Consequently, deleting prfC increased both the level and
hances the level of UGA readthrough, and various levels of pro- heterogeneity of UGA readthrough (Figures 5C and 5D). These
tein synthesis among single cells contribute to heterogeneous results suggest that RF2 fluctuations do not significantly
UGA readthrough in a bacterial population. contribute to UGA readthrough heterogeneity in WT E. coli cells,
but defects in translation termination enhance the sensitivity of
Defective Translation Termination Increases UGA UGA readthrough to RF2 fluctuations among single cells,
Readthrough Heterogeneity thereby increasing readthrough heterogeneity.
We next tested how RF2 fluctuations affect UGA readthrough. In
E. coli, the coding region of the prfB gene (encoding RF2) con- UGA Readthrough Promotes Cell Growth from
tains a TGAC frameshifting codon, which autoregulates the pro- Stationary Phase
duction of RF2 (Curran, 1993; Curran and Yarus, 1988). A high Altering translational fidelity results in various phenotypic
level of RF2 decreases the frameshifting efficiency at the changes (Drummond and Wilke, 2009; Pan, 2013). To provide in-
TGAC site by promoting early termination, therefore reducing sights into how UGA readthrough heterogeneity affects bacterial
RF2 translation. To test how RF2 autoregulates UGA read- physiology, we used time-lapse fluorescence microscopy to
through heterogeneity, we used a genome engineering tool monitor division of individual cells with various levels of UGA
(Wang et al., 2009) to change the chromosomal site of TGAC in readthrough. Stationary-phase WT cells carrying the m-TGA-y
prfB to TAGC. RF2 does not terminate translation at UAG co- reporter were tested on agar pad with minimal glucose medium
dons. Therefore, the TAGC mutation in prfB is expected to in the chamber of an automated fluorescence microscope. We
abolish autoregulation of RF2 production and increase fluctua- found that cells with high levels of UGA readthrough required a
tions of RF2 protein levels among single cells. Surprisingly, the shorter time to reach the first division than cells with low UGA
TAGC mutation did not significantly alter the level or CV of readthrough levels (Figure 6A). To test whether increased UGA
UGA readthrough (Figure S6), suggesting that the RF2 activity readthrough is sufficient to cause faster growth from stationary
was possibly close to saturation for translation termination at phase, we expressed the WT and suppressor tRNASer in
UGA codons in the WT cells. In line with this, overexpressing MG1655. The suppressor tRNA indeed improved regrowth of
RF2 in WT cells did not further decrease the level of UGA read- stationary-phase cells in minimal media with various carbon
through (Figures 5A and 5B). However, when prfC (encoding source (Figures 6B, 6C, and S7A). However, growth in rich media
RF3 that facilitates RF2 during termination) was deleted, overex- LB was not affected by the suppressor tRNA (Figure S7A),
pressing RF2 decreased the level of UGA readthrough, suggest- suggesting that the growth advantage provided by UGA

Molecular Cell 67, 826836, September 7, 2017 831


Figure 5. Defective Termination Increases
UGA Readthrough Heterogeneity
(A) ASKA-prfB (Kitagawa et al., 2005) leads to
overproduction of prfB mRNA as determined with
quantitative reverse transcriptase PCR. In the
absence of isopropyl b-D-1-thiogalactopyrano-
side (IPTG), the promoter of the ASKA plasmid is
leaky (Kitagawa et al., 2005).
(B) The m-TGA-y reporter was tested in strains
with and without the RF2 overexpression plasmid.
In MG1655, increasing the RF2 level does not
further decrease UGA readthrough, suggesting
that the RF2 activity is close to saturation. In
contrast, excess levels of RF2 decrease UGA
readthrough in the RF3 (encoded by prfC) deleting
strain, indicating increased sensitivity of UGA
readthrough to RF2 fluctuations when the release
factor activity is compromised. The DprfC result
also suggests the RF2 protein is successfully
overproduced from the plasmid.
(C) Deleting RF3 and introducing the rpsD* muta-
tion both increases UGA readthrough of the
m-TGA-y reporter.
(D) RF3 deletion increases the CV of UGA read-
through. In contrast, the rpsD* mutation decreases
the CV. In addition to recycling release factors,
RF3 also maintains fidelity during translation elongation (Zaher and Green, 2011). The rpsD* result suggests that increased UGA readthrough heterogeneity upon
RF3 deletion is caused by defective termination rather than reduced fidelity during elongation. Data are represented as mean SD.

readthrough is manifested under poor-nutrient conditions. In stop codon by the EF-Tu:Trp-tRNATrp:GTP ternary complex,
addition, regrowth of log-phase cells was not much affected which competes with RF2 that releases the elongating peptide
by the suppressor tRNA even in minimal media (Figure S7B). (Figure 7). Fluctuations of TC and RF2 among single cells would
Collectively, these results suggest that heterogeneous UGA lead to heterogeneity of UGA readthrough and increase its noise.
readthrough allows a subpopulation of cells to grow faster We have shown that reduced translation increases the level of
from the stationary phase, thereby improving the overall fitness UGA readthrough and Trp-tRNATrp (Figures 3, 4, S4, and S5).
of the bacterial population. Because tRNATrp is stable (Figure 4B), the increase in tRNATrp
in the presence of Chl is likely due to enhanced transcription.
DISCUSSION In addition to tRNATrp, the level of tRNAGly is also increased by
Chl (Figure 4C). It is possible that some unknown protein factor
Noise in gene expression has been extensively studied at the is involved in repressing the tRNA promoters, and Chl reduces
transcriptional level, but the levels and sources of noise during synthesis of the repressor to upregulate tRNA transcription.
translation are poorly understood. Here, we have developed a We also show that the protein level of EF-Tu is not significantly
dual-fluorescence reporter system to determine the noise of affected by Chl (Figure 4D). Given that EF-Tu is very abundant
stop codon readthrough with minimal influence from noise pro- and in 6-fold excess relative to the ribosome (Dai et al., 2016),
duced during transcription, translation initiation, and protein an increase in Trp-tRNATrp is expected to drive the formation
degradation. Such reporters will be broadly useful to determine of TC. Another factor that may affect UGA readthrough is the
the levels and heterogeneity of translational errors in single cells RF2 protein level. We show that the FLAG-tagged RF2 level in-
in their native environments, e.g., within biofilms and during host- creases in the presence of Chl (Figure 4D), which is not expected
microbe interactions. We have further provided in-depth ana- to enhance UGA readthrough. Our data thus suggest that
lyses of the regulation of UGA readthrough noise. The sources reduced protein synthesis promotes UGA readthrough by
of gene expression noise include intrinsic sources that result increasing the effective concentration of TC.
from stochastic transcription and translation of individual genes Recently, Dai et al. (2016) have shown that some ribosome-
and extrinsic sources caused by fluctuations of global resources targeting antibiotics, including Chl and Tet, reduce the active
among single cells (Ackermann, 2015; Blake et al., 2003; Elowitz fraction of ribosomes rather than decreasing translation elonga-
et al., 2002). The heterogeneity of UGA readthrough over long tion rates). Our results suggest that reducing the concentration
periods of growth under our experimental conditions is likely of active ribosomes promotes UGA readthrough (Figures 3, 4,
dominated by extrinsic sources of noise, such as fluctuations and S5). Several lines of evidence indicate that protein synthesis
of the concentrations of nutrients and translational components is intrinsically heterogeneous: (1) the mCherry protein level is
among individual cells. For example, amino acid levels may highly heterogeneous among single cells even with the chromo-
reduce or deplete in some cells after 24 hr of growth in LB. somal reporters (Figure S3A), indicating that overall protein pro-
UGA readthrough results from occasional recognition of the duction is heterogeneous; (2) the components of the protein

832 Molecular Cell 67, 826836, September 7, 2017


Figure 6. UGA Readthrough Promotes Cell
Growth from Stationary Phase
(A) Percentage of cells with high and low error
levels that require different time periods to reach
the first cell division. Stationary-phase MG1655
cells were placed on M9 glucose agar pad and
monitored for fluorescence and growth. High and
low error cells are the top and bottom quartiles
of individual cells ranked by the YFP/mCherry
ratio of the m-TGA-y reporter, respectively.
A significantly higher percentage of cells with
high UGA readthrough levels divide within
120 min compared with cells with low UGA
readthrough levels.
(B) Stationary-phase cultures of MG1655 carrying
the WT or UGA suppressor tRNASer were diluted
50-fold, and growth was monitored over time.
(C) UGA suppressor tRNASer improves growth in
minimal media with various carbon sources
(1% each), but not in LB. Data are represented as
mean SD.

synthesis machinery, including ribosomal proteins and elonga- identical single cells due to heterogeneity during expression of
tion factors, vary among single cells at the protein level in ribosomal proteins and rRNAs, ribosome maturation, and inacti-
E. coli (Taniguchi et al., 2010); (3) translation initiation has been vation. This would exemplify fluctuations of TC and contribute to
shown to be noisy in bacteria (Ozbudak et al., 2002); and (4) ribo- the heterogeneity of stop codon readthrough (Figure 7).
some hibernation alters gene expression heterogeneity (Guido Our proteomic analysis identified Trp to be the major amino acid
et al., 2007). Collectively, such evidence suggests that the that suppresses UGA in E. coli, which is in agreement with previous
concentration of active ribosomes may vary among genetically studies (Engelberg-Kulka, 1981). E. coli tRNATrp contains a

Figure 7. Model for UGA Readthrough Het-


erogeneity
(A) The near-cognate Trp-tRNATrp forms a ternary
complex (TC) with EF-Tu and GTP to compete with
RF2 for the UGA stop codon. Recognition of UGA
by RF2 leads to release of the growing peptide
from the ribosome, whereas Trp-tRNATrp sup-
presses UGA by adding Trp to the growing pep-
tide. The UGA readthrough level in a single cell is
determined by the effective concentrations of TC
and RF2.
(B) The concentration of active ribosomes may
vary among single cells due to fluctuations of
global resources and ribosome maturation/inacti-
vation. A low concentration of active ribosomes
decreases global protein synthesis, which in-
creases the effective concentration of TC due to its
reduced usage during translation.
(C) UGA readthrough is sensitive to TC fluctua-
tions. An increase in effective TC leads to
enhanced UGA readthrough.
(D) WT cells with effective translation termination
exhibit nearly saturated RF2 activity, and UGA
readthrough is insensitive to RF2 fluctuations.
Defects in RF2 activity, e.g., due to RF3 deletion,
enhance the sensitivity of UGA readthrough to RF2
fluctuations and thus increase UGA readthrough
heterogeneity. Dashed lines indicate arbitrary
ranges of concentrations. AU, arbitrary units.

Molecular Cell 67, 826836, September 7, 2017 833


modification at A37 that facilitates C-A pairing at the wobble posi- AUTHOR CONTRIBUTIONS
tion (Vacher et al., 1984), which may explain why tRNATrp sup-
Y.F., C.R.E., K.W.B., W.M., O.A.I., J.R., and J.L. designed the experiments.
presses UGA better than other near-cognate tRNAs such as
Y.F., C.R.E., K.W.B., K.J.W., and J.L. performed the experiments. Y.F.,
tRNACys. The suppression efficiency of UAG appears to be lower C.R.E., K.W.B., K.J.W., J.R., and J.L. analyzed the experimental data. K.B.
than UGA, presumably because of the stronger termination and O.A.I. performed mathematical modeling. Y.F., C.R.E., K.W.B., K.B.,
activity of RF1. Indeed, in an RF1 deletion strain, multiple amino W.M., O.A.I., J.R., and J.L. wrote the manuscript.
acids have been detected to suppress UAG (Aerni et al., 2015).
In E. coli, UGA is used by "30% of the protein-coding genes ACKNOWLEDGMENTS
as the stop codon. Readthrough of stop codons may cause mis-
folded protein stress or production of protein isoforms with new We thank Drs. Arvind R. Subramaniam and Philippe Cluzel (Harvard University)
functions (Dunn et al., 2013; True and Lindquist, 2000). For and the National BioResource Project (Japan) for plasmids, Dr. Nicholas De
Lay (The University of Texas Health Science Center at Houston) for strains,
example, sequences following stop codon readthrough may
Dr. Veronica Rowlett (The University of Texas Health Science Center at Hous-
contain novel localization signals (Dunn et al., 2013). In this ton) for technical assistance, and Dr. Marc Lajoie (University of Washington) for
study, we show that increased UGA readthrough in E. coli en- help with bioinformatics. This work was funded by NIGMS (R01GM115431 to
hances growth under poor-nutrient conditions and promotes J.L., R01GM61074 to W.M., and R01GM117230 to J.R.) and NSF (DGE-
growth from stationary phase (Figures 6B, 6C, and S7A). The un- 1122492 to K.W.B. and PHY-1427654 to O.A.I).
derlying mechanism remains to be elucidated. A pathway enrich-
ment analysis suggests that genes with UGA as stop codons are Received: March 20, 2017
Revised: May 22, 2017
most significantly enriched in ABC transporters (Table S3), which
Accepted: July 7, 2017
are involved in nutrient uptake (Hollenstein et al., 2007; Jones Published: August 3, 2017
and George, 2004). It is possible that UGA readthrough may alter
the function of such transporters and provide growth advantage REFERENCES
when nutrients are limited. In line with this notion, UGA suppres-
sion does not enhance growth in rich media (Figure S7A). Such Ackermann, M. (2015). A functional perspective on phenotypic heterogeneity
advantage of high UGA readthrough on cell growth is also in microorganisms. Nat. Rev. Microbiol. 13, 497508.
observed in single cells (Figure 6A). Heterogeneous UGA read- Adamski, F.M., McCaughan, K.K., Jrgensen, F., Kurland, C.G., and Tate,
through among single cells may thus provide benefits to the W.P. (1994). The concentration of polypeptide chain release factors 1 and 2
microbial population by enhancing phenotypic diversity and at different growth rates of Escherichia coli. J. Mol. Biol. 238, 302308.

facilitating adaptation to ever-changing environments. Aerni, H.R., Shifman, M.A., Rogulina, S., ODonoghue, P., and Rinehart, J.
(2015). Revealing the amino acid composition of proteins within an expanded
genetic code. Nucleic Acids Res. 43, e8.
STAR+METHODS Baba, T., Ara, T., Hasegawa, M., Takai, Y., Okumura, Y., Baba, M., Datsenko,
K.A., Tomita, M., Wanner, B.L., and Mori, H. (2006). Construction of
Detailed methods are provided in the online version of this paper Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collec-
and include the following: tion. Mol. Syst. Biol. 2, 2006.0008.
Bacher, J.M., de Crecy-Lagard, V., and Schimmel, P.R. (2005). Inhibited cell
d KEY RESOURCES TABLE growth and protein functional changes from an editing-defective tRNA synthe-
d CONTACT FOR REAGENT AND RESOURCE SHARING tase. Proc. Natl. Acad. Sci. USA 102, 16971701.
d EXPERIMENTAL MODEL AND SUBJECT DETAILS Balazsi, G., van Oudenaarden, A., and Collins, J.J. (2011). Cellular decision
B Bacterial Strains, Plasmids and Growth Conditions making and biological noise: from microbes to mammals. Cell 144, 910925.
d METHOD DETAILS Bezerra, A.R., Simoes, J., Lee, W., Rung, J., Weil, T., Gut, I.G., Gut, M., Bayes,
B Fluorescence Microscopy M., Rizzetto, L., Cavalieri, D., et al. (2013). Reversion of a fungal genetic code
B Time-lapse Microscopy alteration links proteome instability with genomic and phenotypic diversifica-
tion. Proc. Natl. Acad. Sci. USA 110, 1107911084.
B Rate Determination
Bjorkman, J., Samuelsson, P., Andersson, D.I., and Hughes, D. (1999). Novel
B Calculation of CV and Noise
ribosomal mutations affecting translational accuracy, antibiotic resistance and
B Detection of Native UGA Readthrough in cspC
virulence of Salmonella typhimurium. Mol. Microbiol. 31, 5358.
B RNA Isolation
Blake, W.J., KAErn, M., Cantor, C.R., and Collins, J.J. (2003). Noise in eukary-
B qRT-PCR
otic gene expression. Nature 422, 633637.
B Acidic Gel Northern Blotting
Blake, W.J., Balazsi, G., Kohanski, M.A., Isaacs, F.J., Murphy, K.F., Kuang, Y.,
B Protein Digestion and Proteomics
Cantor, C.R., Walt, D.R., and Collins, J.J. (2006). Phenotypic consequences of
B Peptide Database Construction and Searches promoter-mediated transcriptional noise. Mol. Cell 24, 853865.
B Mathematical Modeling Cochella, L., and Green, R. (2005). An active role for tRNA in decoding beyond
d QUANTITATION AND STATISTICAL ANALYSES codon:anticodon pairing. Science 308, 11781180.
Cox, J., and Mann, M. (2008). MaxQuant enables high peptide identification
SUPPLEMENTAL INFORMATION rates, individualized p.p.b.-range mass accuracies and proteome-wide pro-
tein quantification. Nat. Biotechnol. 26, 13671372.
Supplemental Information includes seven figures and four tables and can be Curran, J.F. (1993). Analysis of effects of tRNA:message stability on frameshift
found with this article online at http://dx.doi.org/10.1016/j.molcel.2017. frequency at the Escherichia coli RF2 programmed frameshift site. Nucleic
07.010. Acids Res. 21, 18371843.

834 Molecular Cell 67, 826836, September 7, 2017


Curran, J.F., and Yarus, M. (1988). Use of tRNA suppressors to probe regula- Janssen, B.D., Diner, E.J., and Hayes, C.S. (2012). Analysis of aminoacyl-and
tion of Escherichia coli release factor 2. J. Mol. Biol. 203, 7583. peptidyl-tRNAs by gel electrophoresis. Methods Mol. Biol. 905, 291309.
Dai, X., Zhu, M., Warren, M., Balakrishnan, R., Patsalo, V., Okano, H., Javid, B., Sorrentino, F., Toosky, M., Zheng, W., Pinkham, J.T., Jain, N., Pan,
Williamson, J.R., Fredrick, K., Wang, Y.P., and Hwa, T. (2016). Reduction of M., Deighan, P., and Rubin, E.J. (2014). Mycobacterial mistranslation is neces-
translating ribosomes enables Escherichia coli to maintain elongation rates sary and sufficient for rifampicin phenotypic resistance. Proc. Natl. Acad. Sci.
during slow growth. Nat. Microbiol. 2, 16231. USA 111, 11321137.
Datsenko, K.A., and Wanner, B.L. (2000). One-step inactivation of chromo- Johansson, M., Lovmar, M., and Ehrenberg, M. (2008). Rate and accuracy of
somal genes in Escherichia coli K-12 using PCR products. Proc. Natl. Acad. bacterial protein synthesis revisited. Curr. Opin. Microbiol. 11, 141147.
Sci. USA 97, 66406645.
Jones, P.M., and George, A.M. (2004). The ABC transporter structure and
Datta, S., Costantino, N., and Court, D.L. (2006). A set of recombineering plas- mechanism: perspectives on recent research. Cell. Mol. Life Sci. 61, 682699.
mids for gram-negative bacteria. Gene 379, 109115.
Kaern, M., Elston, T.C., Blake, W.J., and Collins, J.J. (2005). Stochasticity in
Davis, B.D. (1987). Mechanism of bactericidal action of aminoglycosides. gene expression: from theories to phenotypes. Nat. Rev. Genet. 6, 451464.
Microbiol. Rev. 51, 341350.
Kampen, N.G.v. (2007). Stochastic Processes in Physics and Chemistry (North
Devaraj, A., Shoji, S., Holbrook, E.D., and Fredrick, K. (2009). A role for the 30S Holland).
subunit E site in maintenance of the translational reading frame. RNA 15,
Karkhanis, V.A., Mascarenhas, A.P., and Martinis, S.A. (2007). Amino acid tox-
255265.
icities of Escherichia coli that are prevented by leucyl-tRNA synthetase amino
Dincbas-Renqvist, V., Engstrom, A., Mora, L., Heurgue-Hamard, V.,
acid editing. J. Bacteriol. 189, 87658768.
Buckingham, R., and Ehrenberg, M. (2000). A post-translational modification
in the GGQ motif of RF2 from Escherichia coli stimulates termination of trans- Kitagawa, M., Ara, T., Arifuzzaman, M., Ioka-Nakamichi, T., Inamoto, E.,
lation. EMBO J. 19, 69006907. Toyonaga, H., and Mori, H. (2005). Complete set of ORF clones of
Escherichia coli ASKA library (a complete set of E. coli K-12 ORF archive):
Drummond, D.A., and Wilke, C.O. (2009). The evolutionary consequences of
unique resources for biological research. DNA Res. 12, 291299.
erroneous protein synthesis. Nat. Rev. Genet. 10, 715724.
Kramer, E.B., and Farabaugh, P.J. (2007). The frequency of translational
Dunn, J.G., Foo, C.K., Belletier, N.G., Gavis, E.R., and Weissman, J.S. (2013).
misreading errors in E. coli is largely determined by tRNA competition. RNA
Ribosome profiling reveals pervasive and regulated stop codon readthrough in
13, 8796.
Drosophila melanogaster. eLife 2, e01179.
Kudla, G., Murray, A.W., Tollervey, D., and Plotkin, J.B. (2009). Coding-
Eggertsson, G., and Soll, D. (1988). Transfer ribonucleic acid-mediated sup-
sequence determinants of gene expression in Escherichia coli. Science 324,
pression of termination codons in Escherichia coli. Microbiol. Rev. 52,
255258.
354374.
Lajoie, M.J., Rovner, A.J., Goodman, D.B., Aerni, H.R., Haimovich, A.D.,
Elowitz, M.B., Levine, A.J., Siggia, E.D., and Swain, P.S. (2002). Stochastic
Kuznetsov, G., Mercer, J.A., Wang, H.H., Carr, P.A., Mosberg, J.A., et al.
gene expression in a single cell. Science 297, 11831186.
(2013). Genomically recoded organisms expand biological functions. Science
Engelberg-Kulka, H. (1981). UGA suppression by normal tRNA Trp in
342, 357360.
Escherichia coli: codon context effects. Nucleic Acids Res. 9, 983991.
Lee, J.W., Beebe, K., Nangle, L.A., Jang, J., Longo-Guess, C.M., Cook, S.A.,
Fan, Y., Wu, J., Ung, M.H., De Lay, N., Cheng, C., and Ling, J. (2015). Protein
Davisson, M.T., Sundberg, J.P., Schimmel, P., and Ackerman, S.L. (2006).
mistranslation protects bacteria against oxidative stress. Nucleic Acids Res.
Editing-defective tRNA synthetase causes protein misfolding and neurode-
43, 17401748.
generation. Nature 443, 5055.
Ferdaus, M.Z., Barber, K.W., Lopez-Cayuqueo, K.I., Terker, A.S., Argaiz, E.R.,
Levine, E., and Hwa, T. (2007). Stochastic fluctuations in metabolic pathways.
Gassaway, B.M., Chambrey, R., Gamba, G., Rinehart, J., and McCormick, J.A.
Proc. Natl. Acad. Sci. USA 104, 92249229.
(2016). SPAK and OSR1 play essential roles in potassium homeostasis through
actions on the distal convoluted tubule. J. Physiol. 594, 49454966. Levine, J.H., Lin, Y., and Elowitz, M.B. (2013). Functional roles of pulsing in ge-
netic circuits. Science 342, 11931200.
Fredrick, K., and Noller, H.F. (2002). Accurate translocation of mRNA by the
ribosome requires a peptidyl group or its analog on the tRNA moving into Ling, J., Reynolds, N., and Ibba, M. (2009). Aminoacyl-tRNA synthesis and
the 30S P site. Mol. Cell 9, 11251131. translational quality control. Annu. Rev. Microbiol. 63, 6178.
Fredriksson, A., Ballesteros, M., Peterson, C.N., Persson, O., Silhavy, T.J., and Ling, J., ODonoghue, P., and Soll, D. (2015). Genetic code flexibility in micro-
Nystrom, T. (2007). Decline in ribosomal fidelity contributes to the accumula- organisms: novel mechanisms and impact on physiology. Nat. Rev. Microbiol.
tion and stabilization of the master stress response regulator sigmaS upon car- 13, 707721.
bon starvation. Genes Dev. 21, 862874. Liu, Y., Satz, J.S., Vo, M.N., Nangle, L.A., Schimmel, P., and Ackerman, S.L.
Freistroffer, D.V., Kwiatkowski, M., Buckingham, R.H., and Ehrenberg, M. (2014). Deficiencies in tRNA synthetase editing activity cause cardioproteinop-
(2000). The accuracy of codon recognition by polypeptide release factors. athy. Proc. Natl. Acad. Sci. USA 111, 1757017575.
Proc. Natl. Acad. Sci. USA 97, 20462051. Liu, Z., Vargas-Rodriguez, O., Goto, Y., Novoa, E.M., Ribas de Pouplana, L.,
Gordon, A.J., Satory, D., Halliday, J.A., and Herman, C. (2015). Lost in tran- Suga, H., and Musier-Forsyth, K. (2015). Homologous trans-editing factors
scription: transient errors in information transfer. Curr. Opin. Microbiol. with broad tRNA specificity prevent mistranslation caused by serine/threonine
24, 8087. misactivation. Proc. Natl. Acad. Sci. USA 112, 60276032.
Guido, N.J., Lee, P., Wang, X., Elston, T.C., and Collins, J.J. (2007). A pathway Lu, J., Bergert, M., Walther, A., and Suter, B. (2014). Double-sieving-defective
and genetic factors contributing to elevated gene expression noise in station- aminoacyl-tRNA synthetase causes protein mistranslation and affects cellular
ary phase. Biophys. J. 93, L55L57. physiology and development. Nat. Commun. 5, 5650.
Hollenstein, K., Dawson, R.J., and Locher, K.P. (2007). Structure and mecha- Mascarenhas, A.P., An, S., Rosen, A.E., Martinis, S.A., and Musier-Forsyth, K.
nism of ABC transporter proteins. Curr. Opin. Struct. Biol. 17, 412418. (2008). Fidelity mechanisms of the aminoacyl-tRNA synthetases. In Protein
Huh, D., and Paulsson, J. (2011). Non-genetic heterogeneity from stochastic Engineering, U.L. RajBhandary and C. Kohrer, eds. (New York: Springer-
partitioning at cell division. Nat. Genet. 43, 95100. Verlag), pp. 153200.
Indrisiunaite, G., Pavlov, M.Y., Heurgue-Hamard, V., and Ehrenberg, M. Mikkola, R., and Kurland, C.G. (1992). Selection of laboratory wild-type pheno-
(2015). On the pH dependence of class-1 RF-dependent termination of type from natural isolates of Escherichia coli in chemostats. Mol. Biol. Evol. 9,
mRNA translation. J. Mol. Biol. 427, 18481860. 394402.

Molecular Cell 67, 826836, September 7, 2017 835


Mora, L., Heurgue-Hamard, V., de Zamaroczy, M., Kervestin, S., and Soufi, B., Krug, K., Harst, A., and Macek, B. (2015). Characterization of the
Buckingham, R.H. (2007). Methylation of bacterial release factors RF1 and E. coli proteome and its modifications during growth and ethanol stress.
RF2 is required for normal translation termination in vivo. J. Biol. Chem. 282, Front. Microbiol. 6, 103.
3563835645. Su, H.W., Zhu, J.H., Li, H., Cai, R.J., Ealand, C., Wang, X., Chen, Y.X., Kayani,
Newman, J.R., Ghaemmaghami, S., Ihmels, J., Breslow, D.K., Noble, M., M.U., Zhu, T.F., Moradigaravand, D., et al. (2016). The essential mycobacterial
DeRisi, J.L., and Weissman, J.S. (2006). Single-cell proteomic analysis of amidotransferase GatCAB is a modulator of specific translational fidelity. Nat.
S. cerevisiae reveals the architecture of biological noise. Nature 441, 840846. Microbiol. 1, 16147.

Ozbudak, E.M., Thattai, M., Kurtser, I., Grossman, A.D., and van Oudenaarden, Subramaniam, A.R., Pan, T., and Cluzel, P. (2013). Environmental perturba-
A. (2002). Regulation of noise in the expression of a single gene. Nat. Genet. tions lift the degeneracy of the genetic code to regulate protein levels in bac-
31, 6973. teria. Proc. Natl. Acad. Sci. USA 110, 24192424.
Taniguchi, Y., Choi, P.J., Li, G.W., Chen, H., Babu, M., Hearn, J., Emili, A., and
Pan, T. (2013). Adaptive translation as a mechanism of stress response and
Xie, X.S. (2010). Quantifying E. coli proteome and transcriptome with single-
adaptation. Annu. Rev. Genet. 47, 121137.
molecule sensitivity in single cells. Science 329, 533538.
Paulsson, J. (2004). Summing up the noise in gene networks. Nature 427,
True, H.L., and Lindquist, S.L. (2000). A yeast prion provides a mechanism for
415418.
genetic variation and phenotypic diversity. Nature 407, 477483.
Polikanov, Y.S., Blaha, G.M., and Steitz, T.A. (2012). How hibernation factors Vacher, J., Grosjean, H., Houssier, C., and Buckingham, R.H. (1984). The ef-
RMF, HPF, and YfiA turn off protein synthesis. Science 336, 915918. fect of point mutations affecting Escherichia coli tryptophan tRNA on anti-
Quan, S., Skovgaard, O., McLaughlin, R.E., Buurman, E.T., and Squires, C.L. codon-anticodon interactions and on UGA suppression. J. Mol. Biol. 177,
(2015). Markerless Escherichia coli rrn deletion strains for genetic determina- 329342.
tion of ribosomal binding sites. G3 (Bethesda) 5, 25552557. Wang, H.H., Isaacs, F.J., Carr, P.A., Sun, Z.Z., Xu, G., Forest, C.R., and
Reynolds, N.M., Ling, J., Roy, H., Banerjee, R., Repasky, S.E., Hamel, P., and Church, G.M. (2009). Programming cells by multiplex genome engineering
Ibba, M. (2010). Cell-specific differences in the requirements for translation and accelerated evolution. Nature 460, 894898.
quality control. Proc. Natl. Acad. Sci. USA 107, 40634068. Wilson, D.N. (2014). Ribosome-targeting antibiotics and mechanisms of bac-
terial resistance. Nat. Rev. Microbiol. 12, 3548.
Ribas de Pouplana, L., Santos, M.A., Zhu, J.H., Farabaugh, P.J., and Javid, B.
(2014). Protein mistranslation: friend or foe? Trends Biochem. Sci. 39, Xaplanteri, M.A., Andreou, A., Dinos, G.P., and Kalpaxis, D.L. (2003). Effect of
355362. polyamines on the inhibition of peptidyltransferase by antibiotics: revisiting the
mechanism of chloramphenicol action. Nucleic Acids Res. 31, 50745083.
Rosenberg, S.M., Shee, C., Frisch, R.L., and Hastings, P.J. (2012). Stress-
Yoshida, H., and Wada, A. (2014). The 100S ribosome: ribosomal hibernation
induced mutation via DNA breaks in Escherichia coli: a molecular mechanism
induced by stress. Wiley Interdiscip. Rev. RNA 5, 723732.
with implications for evolution and medicine. BioEssays 34, 885892.
Young, J.W., Locke, J.C., and Elowitz, M.B. (2013). Rate of environmental
Rudorf, S., and Lipowsky, R. (2015). Protein synthesis in E. coli: dependence of
change determines stress response specificity. Proc. Natl. Acad. Sci. USA
codon-specific elongation on tRNA concentration and codon usage. PLoS
110, 41404145.
ONE 10, e0134994.
Youngman, E.M., He, S.L., Nikstad, L.J., and Green, R. (2007). Stop codon
Sanchez, A., and Golding, I. (2013). Genetic determinants and cellular con- recognition by release factors induces structural rearrangement of the ribo-
straints in noisy gene expression. Science 342, 11881193. somal decoding center that is productive for peptide release. Mol. Cell 28,
Sanchez, A., Choubey, S., and Kondev, J. (2013). Regulation of noise in gene 533543.
expression. Annu. Rev. Biophys. 42, 469491. Zaher, H.S., and Green, R. (2009). Fidelity at the molecular level: lessons from
Sliusarenko, O., Heinritz, J., Emonet, T., and Jacobs-Wagner, C. (2011). High- protein synthesis. Cell 136, 746762.
throughput, subpixel precision analysis of bacterial morphogenesis and intra- Zaher, H.S., and Green, R. (2011). A primary role for release factor 3 in quality
cellular spatio-temporal dynamics. Mol. Microbiol. 80, 612627. control during translation elongation in Escherichia coli. Cell 147, 396408.

836 Molecular Cell 67, 826836, September 7, 2017


STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE SOURCE IDENTIFIER


Antibodies
Mouse monoclonal anti-mCherry (1C51) NOVUS Biologicals Cat# NBP196752; RRID: AB_11034849
Mouse monoclonal anti-GFP Roche Cat# 11814460001; RRID: AB_390913
Mouse monoclonal anti-RpoB Santa Cruz Biotechnology Cat# SC-101613; RRID: AB_1129055
Mouse monoclonal anti-FLAG Sigma Cat# A8592; RRID: AB_439702
Mouse monoclonal anti-EF-Tu Hycult Biotech Cat# HM6010; RRID: AB_1953571
Bacterial and Virus Strains
rpsD* (Fan et al., 2015) N/A
MG1655 CspC-TGA-YFP-1 This study N/A
MG1655 CspC-TGA-YFP-2 This study N/A
BW25113 Drmf (Baba et al., 2006) N/A
MG1655 DprfC This study N/A
MG1655 attB::Ptet-m-y This study N/A
MG1655 attB::Ptet-m-TGA-y This study N/A
MG1655 D2 (DrrnEG) (Quan et al., 2015) N/A
MG1655 D4 (DrrnGBAD) (Quan et al., 2015) N/A
MG1655 D6 (DrrnGADBHC) (Quan et al., 2015) N/A
MG1655 prfB TAGC This study N/A
MG1655 prfB::3 3 FLAG This study N/A
CR201 N. De Lay N/A
Chemicals, Peptides, and Recombinant Proteins
Chloramphenicol Sigma Cat# C0378
ULTRAhyb Ultrasensitive Hybridization Buffer Invitrogen Cat# AM8670
Rifampicin CHEM-IMPEXINTL INC Cat# 00260
Erythromycin CHEM-IMPEXINTL INC Cat#00126
Tetracycline hydrochloride CHEM-IMPEXINTL INC Cat#00667
Spectinomycin dihydrochloride pentahydrate Sigma Cat# S9007
Critical Commercial Assays
iScript cDNA Synthesis Kit Bio-Rad Cat# 170-8891
SsoAdvanced Universal SYBR Green Supermix Bio-Rad Cat# 172-5271
In-Fusion HD Cloning Plus Takara Bio USA, Inc Cat# 638909
Q5 Hot Start High-Fidelity 2 3 Master Mix NEB Cat# M0494S
TRIzol Reagent Ambion Cat# 15596018
Experimental Models: Organisms/Strains
E. coli: K-12 MG1655 F#; l#; rph-1 E. coli Genetic Stock Center at Yale CGSC# 6300
E. coli: BL21 (DE3) Laboratory strain N/A
S. enterica serovar Typhimurium LT2 ATCC ATCC# 700720
Oligonucleotides
See Table S4 for full list of oligonucleotides used in This study N/A
this study
Recombinant DNA
pZS-Ptet -m-y This study N/A
pZS- Ptet -m-TGA-y This study N/A
pZS- Ptet -m-TAG-y This study N/A
(Continued on next page)

Molecular Cell 67, 826836.e1e5, September 7, 2017 e1


Continued
REAGENT or RESOURCE SOURCE IDENTIFIER
pZS- Ptet -m-y +1 fs This study N/A
pZS- Ptet -m-y #1 fs This study N/A
pZS*11-yfp13 (Subramaniam et al., 2013) N/A
ASKA-prfB (Kitagawa et al., 2005) N/A
pKD46 (Datsenko and Wanner, 2000) N/A
pCP20 (Datsenko and Wanner, 2000) N/A
pSIM6 (Datta et al., 2006) N/A
pKT-tRNASer_GCT (wild-type control) This study N/A
pKT-tRNASer_TCA (UGA suppressor) This study N/A
pZS- Pcp25 -Mut SD-m-y This study N/A
pZS- Pcp25 -Mut SD-m-TGA-y This study N/A
pZS- Pcp25 -Non-opt-m-y This study N/A
pZS- Pcp25 -Non-opt-m-TGA-y This study N/A
Software and Algorithms
MicrobeTracker 0.937 (Sliusarenko et al., 2011) N/A
Slidebook 6.0 3i Intelligent Imaging Innovation N/A
ImageJ 1.47v bundled with 32-bit Java 1.6.0_20 NIH N/A
Image Lab 6.0 for Windows Bio-Rad N/A
MATLAB R2012b 8.0.0.783 The MathWorks N/A

CONTACT FOR REAGENT AND RESOURCE SHARING

Further information and requests for reagents may be directed to, and will be fulfilled by the lead contact Jiqiang Ling (Jiqiang.Ling@
uth.tmc.edu).

EXPERIMENTAL MODEL AND SUBJECT DETAILS

Bacterial Strains, Plasmids and Growth Conditions


All the bacterial strains and main plasmids used in this study are listed in the Key Resources Table. All the oligos for making the mutant
strains and plasmids are listed in Table S4. The mutant strains are derivatives of E. coli K-12 MG1655 (WT). To construct MG1655
CspC-TGA-YFP-1, MG1655 CspC-TGA-YFP-2, MG1655 attB::Ptet-m-y, MG1655 attB::Ptet-m-TGA-y, and MG1655 prfB::3 3
FLAG strains, a cassette (kan-ccdB) containing the toxin encoding gene ccdB under the control of araBAD promoter and a kanamycin
resistance gene was amplified from template genomic DNA of CR201 strain (obtained from N. De Lay, UTHealth), and introduced into
the specific sites on the chromosome of the parental strain harboring plasmid pSIM6 by l red recombinase-mediated gene replace-
ment (Datta et al., 2006). Strains containing pSIM6 were induced for Red expression by growth at 42! C for 15, and electroporated
with PCR fragments containing the kan-ccdB cassette. 1 mL LB was then added and cells were incubated at 32! C for 2 hr. The suc-
cessful transformants were selected by kanamycin resistance. The kan-ccdB cassette was then replaced with respective DNA frag-
ments. The positive clones were selected by growth on 0.5% arabinose LB plate, and verified by colony polymerase chain reaction
(PCR). The MG1655 prfB TAGC strain was generated with a modified multiplex automated genome engineering (MAGE) method
(Wang et al., 2009). Briefly, prfB-MAGE-TAGC oligo was electroporated into the MG1655 strain carrying plasmid pKD46, and cells
were plated onto LB plates after 5 hr incubation at 32! C. Colony PCR was applied with primer pairs prfB-TAGC-MT-F and prfB-
TAGC-R0 to select for the positive clones. All mutant strains were verified by sequencing.
To construct plasmids pZS-Ptet-m-y, pZS-Ptet-m-TGA-y, pZS-Ptet-m-TAG-y, pZS-Ptet-m-y +1 fs, and pZS-Ptet-m-y #1 fs, the pZS-
Ptet-29AA-y plasmid was generated first by ligating a 29 amino acid (LQTSAGEAAAKEAAAKEAAAKEAAAKAAA) linker with plasmid
pZS*11-yfp13 using In-Fusion HD Cloning according to manufacturers protocol. Fragments amplified from optimized mCherry
gBlock by primer pZS-mCherry-29AA-IF paired with pZS-mCherry-29AA-IR, pZS-mCherryTGA-29AA-IR, pZS-mCherryTAA-
29AA-IR, pZS-mCherryTAG-29AA-IR, pZS-mCherryTAAA-29AA-IR, and pZS-mCherryTA-29AA-IR, respectively, were ligated into
the pZS- Ptet-29AA-Y plasmid through In-Fusion HD Cloning.
Unless otherwise noted, E. coli strains were grown in LB at 37! C. For fluorescence microscopy analysis, overnight cultures were
diluted 1: 1,000 and grown for 24 hr. 100 mg/ml Chl was added to immediately stop protein synthesis. After 4 hr incubation to allow full
maturation of mCherry and YFP, samples were subjected to fluorescence microscopy analysis. The minimal medium contains

e2 Molecular Cell 67, 826836.e1e5, September 7, 2017


47.8 mM Na2PO4, 22.0 mM KH2PO4, 8.6 mM NaCl, 18.7 mM NH4Cl, 4 mM MgSO4, 0.2 mM CaCl2, 40 mg/ml each of the 20 amino
acids, and 0.4%1% glucose or indicated sugar.

METHOD DETAILS

Fluorescence Microscopy
Samples were placed on a 2 mL agarose (1.5%) phosphate buffer pad on 15-well Multitest Slides (MP Biomedicals, LLC.). Images
were obtained on an Olympus IX81-ZDC inverted microscope using Slidebook imaging software. Background fluorescence was sub-
tracted from each image using ImageJ Background Subtraction with a 50.0 pixel rolling ball radius. Single-cell fluorescence quan-
titation was completed with MicrobeTracker (Sliusarenko et al., 2011), a MATLAB-based software package. Single-cell fluorescence
data were exported from MATLAB and all further data analysis was completed in Microsoft Excel.

Time-lapse Microscopy
Overnight cultures were diluted 1:1,000 in LB and grown for 24 hr at 37! C. Cultures were placed on a 200 mL 1.5% agarose LB pad.
Fluorescent images were taken at the initial time point for quantitation. Cells were followed for 150 min at room temperature with DIC
images taken at regularly spaced intervals throughout the experiment. Image analysis and editing were performed using ImageJ.

Rate Determination
E. coli cultures were incubated at 37! C in a microplate reader (Synergy HT, BioTek) using 96-well black side plates (Corning). The
signals of mCherry, YFP, and A600 were measured every 20 min with fluorescence spectrometry. To calculate translational error
rates, the YFP/mCherry ratio of m-TGA-y, m-TAG-y, m-y +1 fs, and m-y #1 fs was normalized by the YFP/mCherry ratio of the control
m-y reporter. Protein synthesis rates were calculated as described (Subramaniam et al., 2013) with the following formula:
Protein synthesis rate; S = 1=Absorbance3dfluorescence=dtime.
To calculate translational error rates by western blotting, E. coli strains were grown in LB medium with and without addition of chlor-
amphenicol at 37! C for 24 hr before 10 mL of culture was harvested. The cells were lysed by sonication, and whole protein samples
were analyzed by western blotting with a primary antibody against mCherry. The signals were qualified by Image Lab (Bio-Rad), and
the error rates were calculated as the percentage of mCherry-YFP fusion protein.

Calculation of CV and Noise


CV was calculated as the standard deviation (s) divided by the mean (m) of the YFP/mCherry ratio of each cell in the same microscopic
image frame. Noise was calculated as the ratio of the variance (s2) over the square of the mean (m2).

Detection of Native UGA Readthrough in cspC


The readthrough level of cspC UGA was determined by western blotting. E. coli strains were grown in LB with and without addition of
Chl (2 mg/ml) at 37! C for 24 hr. 10 mL of cell culture were harvested and lysed by sonication, and total proteins were separated by
SDS-PAGE. Antibodies against GFP (Roche) and RpoB (loading control, Santa Cruz) were used in western blotting analysis. The sig-
nals were qualified by Image Lab (Bio-Rad).

RNA Isolation
To prepare total RNA for qRT-PCR, cells in LB medium with or without 10 mM IPTG were grown to mid-log phase (OD600"0.6-0.8).
800 mL of culture was used for total RNA extraction using hot phenol, and residual chromosomal DNA was removed.
To isolate total RNA for acidic gel Northern blotting, overnight cultures were diluted 1: 1,000 in LB and grown for 14 hr at 37! C.
10 mL of cell culture was harvested at 4! C and resuspend with TRIzol reagent immediately. Cells were lysed by the beads beater
with 0.1 mm glass beads (RPI). RNA sample was prepared according to manufacturers protocol, and was stored in 10 mM sodium
acetate buffer (pH 5.0) with 1 mM ethylenediaminetetraacetic acid (EDTA) at #80! C to preserve aminoacyl-tRNAs. Deacylated sam-
ples were obtained by incubating RNA in 200 mM Tris pH 9.0 at 37! C for 30 min.

qRT-PCR
1 mg of total DNA-free RNA was reverse transcribed using iScript cDNA Synthesis Kit (Bio-Rad) according to manufacturers protocol.
cDNA was amplified with the corresponding primers (see Table S4). qPCR was performed using Bio-Rad CFX96 and SsoAdvanced
Universal SYBR Green Supermix (Bio-Rad) according to manufacturers suggestion. mreB transcript level was used for normaliza-
tion. The DDCt method was used to obtain the fold changes of target genes.

Acidic Gel Northern Blotting


Acid urea polyacrylamide gel electrophoresis was performed according to the procedures described in (Janssen et al., 2012). Briefly,
12% Acid urea polyacrylamide gel was prepared freshly before use. 5 mg of RNA sample was loaded onto the gel, and subjected the
electrophoresis with sodium acetate running buffer (100 mM sodium acetate pH5.0, 1 mM EDTA) in cold room for 6 hr. Gel-separated
RNA was transferred to GT membrane (Bio-Rad) by semi-dry electro-blotting at 15 V for 40 min in 0.5 3 TBE. The membrane was

Molecular Cell 67, 826836.e1e5, September 7, 2017 e3


cross-linked by UV and probed with 50 end biotin labeled DNA oligonucleotide probes (see Table S4). Trp-tRNATrp, tRNATrp, Gly-
tRNAGly and tRNAGly were detected by the Northern blotting, respectively, with SsrA as the loading control. Quantitation was per-
formed with Image Lab (Bio-Rad).

Protein Digestion and Proteomics


Protein extraction and digestion from E. coli was performed using cell pellets from 200 mL stationary-phase cultures. Lysis, reduction,
alkylation, trypsin digest, and acid cleavage were performed as in (Lajoie et al., 2013). All resulting peptides were purified and de-
salted using a C18 MacroSpin column (The Nest Group), dried in a rotary vacuum centrifuge, and resuspended in 6 mL 70% formic
acid and 16 mL 0.1% trifluoroacetic acid. Following A280 peptide quantification, samples were diluted to a concentration of 0.5 mg/mL
in the same buffer, and 4 mL of sample (2 mg total) were injected onto an analytical column using ACQUITY UPLC M-Class (Waters) for
mass spectrometry. Column specifications, solvent gradients, and mass spectrometry parameters for the Q Exactive Plus (Thermo)
were performed as described in (Ferdaus et al., 2016).

Peptide Database Construction and Searches


The MG1655 genome and annotated open reading frames (ORFs) were downloaded from https://www.ncbi.nlm.nih.gov/ on
September 16th, 2016. ORFs terminating in TGA were mapped to the reference genome and extended to the next in-frame stop
codon (TAG, TAA or TGA) using an ad hoc Python script. Extended ORFs were translated and separate entries were included in
the search database for every natural amino acid or an in-frame skipping event at the first TGA position. The normal MG1655 pro-
teome and the extended frame protein sequences (sequences beginning at the first tryptic residue N-terminal to the TGA), including
the m-TGA-y reporter sequences, were used concurrently for mass spectrometry data searches using MaxQuant v.1.5.1.2 (Cox and
Mann, 2008). The copy number of mCherry was estimated using Intensity Based Absolute Quantitation as described (Soufi et al.,
2015). The concentration of total mCherry was similar to that of EF-Tu corresponding to "300,000 molecules per cell. The
mCherry-YFP fusion protein was "2% of total mCherry as observed by fluorescence spectrometry and western blotting ("6,000 mol-
ecules per cell).

Mathematical Modeling
Theoretical modeling of the variation of stop codon (UGA) read-through error with antibiotic chloramphenicol (Chl) concentration are
based on the network shown in Figures S4B and S4C. In the scheme, E represents the bare ribosome with UGA mRNA stop codon in
A-site, ER represents the release factor (RF2)-bound state that undergoes a conformational transition to state ER* and PR is the state
generated after RF2-induced pH-dependent hydrolysis of the peptide chain in the ribosome P-site (Indrisiunaite et al., 2015). Here,
RF2 is the right substrate. For the wrong substrate, i.e., the ternary complex (TC) of Trp-tRNATrp, the initial bound state ET undergoes
GTP-hydrolysis to ET*. From state ET*, the system can take two routes: (i) it can accommodate the aminoacyl-tRNA that leads to
peptide chain elongation, i.e., readthrough of the stop codon (state PT) or (ii) it can reset by a proofreading dissociation of the aa-
tRNA from ribosome (Cochella and Green, 2005). No such proofreading appears to be present for the RF2 pathway (Freistroffer
et al., 2000).
A recent study shows that although Chl reduces the overall growth rate of E. coli, the elongation rate actually rises with Chl con-
centration (Dai et al., 2016). This feature is attributed to an increase in RNA/protein ratio that leads to a rise in effective TC concen-
tration. Our experiment also confirms this fact with the Trp-tRNATrp levels going up with [Chl] (see Figure 4). In contrast, the protein
levels go down with [Chl] (Dai et al., 2016). In addition, binding of Chl to ribosome stalls the translation process and this leads to a fall in
the completion probability of translation. We model these three effects of antibiotic addition with the following three schemes.
Model 1
Here, we take the TC concentration to be an increasing function of Chl concentration as
! "
Chl'
TC' = TC'0 1 + f (1)
Chl' + a
Here, [TC]0 denotes TC concentration in absence of Chl. We assume [TC] to be proportional to the Trp-tRNATrp level and fit the
experimentally obtained fold change using Equation (1). This gives f = 20:2; a = 18:6 g/ml. Now, our experimentally measured growth
rate in absence of Chl is "1.9 h#1. With this information and following a recent analysis of TC concentration for various tRNA species
under different growth conditions (Rudorf and Lipowsky, 2015), we set [TC]0 = 3.2 mM. We do not consider here any dependence of
[RF2] on [Chl]. We take [RF2]0 = 18 mM following a report of RF2 level variation at various E. coli growth rates (Adamski et al., 1994)
Also, in this model, we assume the completion probability to be 100%, i.e., all translation initiation events go to completion.
Model 2
Here, [TC] follows Equation (1). The [RF2] dependence on [Chl] is taken as
#1
RF2' = RF2'0 1 + KChl' (2)

e4 Molecular Cell 67, 826836.e1e5, September 7, 2017


The parameter K ("0.4 mM#1) (Dai et al., 2016) is the binding constant of Chl with ribosome. This corresponds to the assumption
that decrease in the fraction of active ribosome leads to proportional decrease in translation rates and correspondingly in the protein
concentrations. Again, we take the completion probability to be 100%.
Model 3
In this case, the [TC] and [RF2] dependences on [Chl] are given by Equation (1) and Equation (2), respectively. However, we now note
that not every read-through even will result in the completed (functional) YFP protein. Therefore, if the translating ribosome binds Chl
before YFP is completed we assume that the translation will be aborted and the read-through wont result in YFP fluorescence signal.
The resulting functional dependence of completion probability on Chl is modeled as
T #1
Pcom = (3)
T #1 + kb Chl'
Here, T is the completion time of YFP protein synthesis, indicator of stop codon readthrough error. It is defined as N=r where
N = 251 is the number of amino acids (aa) in YFP and r is the elongation rate (Dai et al., 2016). We take r = 17 aa/s based on the values
reported in (Dai et al., 2016). The effective binding constant kb "2 X 104 M#1.s#1 is taken from a recent study on Chl inhibition
(Xaplanteri et al., 2003) that indicates that initial binding of Chl is fast followed by a slower isomerization step.
We study the kinetics of the whole network using the standard tools of first-passage technique (Kampen, 2007). The probability to
reach a given end-state (PR or PT) before the other, starting from state-E, is called the splitting probability (9). The UGA read-through
error is defined as the ratio of the splitting probability of reaching the wrong end (PT) to that of reaching the right end (PR). The param-
eter set to generate the nature of error variation as a function of Chl concentration are given in Table S2.

QUANTITATION AND STATISTICAL ANALYSES

Statistical parameters for each experiment are reported in the corresponding figures. All data presented are the mean of at least three
repeats, with error bars showing one standard deviation. The p values were calculated with unpaired t tests.

Molecular Cell 67, 826836.e1e5, September 7, 2017 e5

Anda mungkin juga menyukai