You are on page 1of 17

Research Article

ClpXP-mediated Degradation of the TAC


Antitoxin is Neutralized by the SecB-like
Chaperone in Mycobacterium
tuberculosis

Pauline Texier 1, Patricia Bordes 1⇑, Jyotsna Nagpal 2, Ambre Julie Sala 1,
Moise Mansour 1, Anne-Marie Cirinesi 1, Xibing Xu 1, David Andrew Dougan 2⇑ and
Pierre Genevaux 1⇑

1 - Laboratoire de Microbiologie et de Génétique Moléculaires, Centre de Biologie Intégrative (CBI), Université de Toulouse,
CNRS, UPS, Toulouse, France
2 - Department of Biochemistry and Genetics, La Trobe Institute for Molecular Science, La Trobe University, Melbourne,
Victoria 3086, Australia

Correspondence to Patricia Bordes, David Andrew Dougan and Pierre Genevaux: patricia.bordes@univ-tlse3.fr
(P. Bordes), d.dougan@latrobe.edu.au (D.A. Dougan), pierre.genevaux@univ-tlse3.fr (P. Genevaux)
https://doi.org/10.1016/j.jmb.2021.166815
Edited by Urs Jenal

Abstract
Bacterial toxin-antitoxin (TA) systems are composed of a deleterious toxin and its antagonistic antitoxin.
They are widespread in bacterial genomes and mobile genetic elements, and their functions remain lar-
gely unknown. Some TA systems, known as TAC modules, include a cognate SecB-like chaperone that
assists the antitoxin in toxin inhibition. Here, we have investigated the involvement of proteases in the acti-
vation cycle of the TAC system of the human pathogen Mycobacterium tuberculosis. We show that the
deletion of endogenous AAA+ proteases significantly bypasses the need for a dedicated chaperone
and identify the mycobacterial ClpXP1P2 complex as the main protease involved in TAC antitoxin degra-
dation. In addition, we show that the ClpXP1P2 degron is located at the extreme C-terminal end of the
chaperone addiction (ChAD) region of the antitoxin, demonstrating that ChAD functions as a hub for both
chaperone binding and recognition by proteases.
Ó 2021 Elsevier Ltd. All rights reserved.

Introduction treatments and the immune response.9–15 TA sys-


tems are also frequently found on mobile genetic
Toxin-antitoxin (TA) systems are small genetic elements, where they generally contribute to gen-
modules encoding a toxic protein and its ome stability.16 In addition, the highly toxic nature
antagonistic antitoxin, which inhibits activity or of certain toxins has raised the possibility that new
expression of the toxin. They are ubiquitous in antibacterial properties carried by toxins might be
bacterial genomes and have been classified in used to identify new drug targets in pathogens or
different types according to the mode of inhibition directly as antimicrobials, alone or in combination
of the toxin by the antitoxin, which can be a with antibiotherapy.17
protein or an RNA.1–4 The role of chromosomal TA The widespread type II TA systems are
systems remains largely unknown, partly due to composed of a labile antitoxin protein that directly
the lack of phenotype associated with their mutation binds and inhibits its cognate toxin, and also acts
under laboratory conditions.5–8 Yet, in some cases, as a transcriptional repressor of its own promoter.2
it has been found that certain TA systems can help Under normal conditions, the toxin is sequestered
bacteria to survive infection by phages, antibiotic in an inactive state, by the antitoxin and bacterial

0022-2836/Ó 2021 Elsevier Ltd. All rights reserved. Journal of Molecular Biology 433 (2021) 166815
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

growth is not affected. It has been proposed that unknown, its overexpression was shown to be
under certain stress conditions, the less stable anti- highly toxic,39,40 affecting the abundance of a small
toxin is degraded by proteases, leading to either the subset of transcripts mainly related to iron and zinc
direct release of free toxin or an increase in de novo homeostasis.40
toxin synthesis. The resulting active toxin would In contrast with classical two-component TA
then target essential cellular processes, including systems, the TAC toxin-antitoxin pair is tightly
translation, DNA replication or cell wall synthesis, controlled by SecBTA, through a direct interaction
leading to growth inhibition and eventually cell between the chaperone and the antitoxin HigA1.41
death.1,2 Yet, recent work performed in E. coli We previously showed that the TAC antitoxin pos-
showed that antitoxin degradation and transcrip- sesses an unusual C-terminal extension, named
tional activation of TA operons by different stresses ChAD (chaperone-addiction region), which renders
do not lead to any detectable toxin activation,7 indi- the antitoxin aggregation-prone and thus unable to
cating that the specific signals that trigger toxin acti- efficiently inhibit the toxin in absence of chaper-
vation in vivo remain to be identified. one.39,41,42 Indeed, the dedicated SecBTA chaper-
The role of AAA+ (ATPases associated with a one was shown to specifically bind the ChAD
variety of cellular activities) proteases in the region of the antitoxin, with ChAD occupying a dis-
turnover of type II antitoxins has been highlighted tinct zone within the substrate-binding site of the
in several studies. For example, Lon is involved in chaperone,42 to facilitate its folding and subsequent
the turnover of CcdA and PemI,18,19 ClpAP in the toxin inhibition. 39,43
degradation of Kis and MazE, and ClpXP in the Previous data showed that the TAC antitoxin
degradation of PhD and DinJ.20–23 Although certain cannot be detected in whole cell extracts unless
antitoxins possess intrinsically flexible domains or the chaperone is present, strongly suggesting that
carboxy-terminal regions sensitive to proteolysis the aggregation-prone antitoxin might be sensitive
that are protected when bound to the toxin, the to proteolysis in vivo.41 Yet, nothing is known about
degradation signals (degrons) within antitoxins that a possible involvement of proteases in the TAC acti-
determine selective proteolysis are currently vation cycle. In this work, we show that the absence
unknown.24–30 In addition, it remains to be deter- of protease can significantly bypass the need for the
mined whether stress-induced proteolytic adaptors dedicated SecBTA chaperone. In addition, we iden-
could participate in the regulated proteolysis of TA tify the mycobacterial ClpXP protease as the main
systems, as it was observed with the Staphylococ- protease involved in HigA1 degradation and show
cus aureus adaptor protein TrfA, which is required that the ClpXP degron is located at the extreme
for the ClpCP-mediated turnover of the antitoxin C-terminal end of ChAD. Remarkably, this work
MazE.31 shows that the short ChAD region contains all the
The toxin-antitoxin-chaperone (TAC) system of determinants for both chaperone binding and
M. tuberculosis is an atypical TA system recognition by proteases, although different resi-
composed of the three genes Rv1955-Rv1956- dues are involved. The relevance for such intricate
Rv1957 organized into a single operon, which interplay between ClpXP and SecB in the TAC acti-
encodes the toxin-antitoxin pair (Rv1955 and vation cycle is discussed.
Rv1956, respectively) followed by the chaperone
(Rv1957). Given Rv1957 is related to the
canonical SecB chaperone, involved in Sec- Results
dependent export of proteins in E. coli,32 we herein Control of TAC system(s) involves cellular
refer to Rv1957 as SecBTA. Although transcription proteolysis
of the TAC operon is significantly induced in host
phagocytes and by several stress conditions rele- Initially we asked whether altered cellular
vant for M. tuberculosis, including DNA damage, proteolysis could bypass the need for
heat shock, nutrient starvation, hypoxia and drug activation/folding of the TAC antitoxin by the Mtb-
persistence, its role in M. tuberculosis stress- SecBTA chaperone in vivo. To do so, we
adaptive response has yet to be established.33–37 transformed chaperoneless Mtb-HigBA1 in a set
The toxin-antitoxin pair of TAC was originally of single E. coli AAA+ protease mutants, including
named Mtb-HigB1-HigA1 due to its unusual Dlon, DclpP, DhslV (encoding ClpQ) and DftsH,
inverted gene organization, with toxin first and anti- and tested their growth in the in vivo toxin
toxin second, as found in the HigB-HigA TA system inhibition assay. Significantly, in the absence of
of Proteus vulgaris.38 While the HigA1 antitoxin pos- SecBTA only the DclpP strain was capable of
sesses a conserved helix-turn-helix (HTH) DNA restoring partial bacterial growth in the presence
binding domain, which binds to the TAC promoter of Mtb-HigBA1 (Figure 1(A)). These data indicate
region, the HigB1 toxin likely belongs to the Gp49 that ClpP is likely to be the main protease
subclass of the RelE toxin superfamily of involved, as previously observed for several two-
ribosome-dependent endoribonucleases.4 component TA systems.20–23,44,45 Consistently,
Although the influence of HigB1 toxin on M. tubercu- bacterial growth was also partially restored when
losis growth under physiological condition is both clpX and clpA, the ATP-dependent unfoldase
2
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

both subunits contribute to the turnover of the


Mtb-HigA1 antitoxin in E. coli. In addition, we found
that rescue of bacterial growth by DclpA DclpX or by
DclpP mutations was less robust than rescue by the
Mtb-SecBTA chaperone. This finding suggests that,
in the absence of ClpP, other E. coli proteases could
degrade the antitoxin. Therefore, we constructed an
E. coli mutant strain in which the genes encoding
three cytosolic AAA+ proteases (ClpP, Lon and
ClpQ) were deleted (D3 strain) and tested the effect
of Mtb-HigBA1 expression in vivo. In this case, dele-
tion of ftsH was not included with the other three
protease coding genes, since it is known that a sin-
gle ftsH mutation leads to a highly sensitive growth
phenotype and that the mutant readily accumulates
extragenic suppressors needed for survival.46
Importantly, the combined deletion of three cytoso-
lic AAA+ proteases was sufficient to significantly
improve the cells tolerance to expression of the
toxic chaperoneless Mtb-HigBA1 pair in vivo (Fig-
ure 1(A)), without affecting toxin levels (Supplemen-
tary Figure S1). Consistent with our previous
proposal41 these data suggest that Mtb-HigA1 is
highly sensitive to degradation in vivo. Moreover,
in the absence of its cognate chaperone, the lack
of protease likely promotes the accumulation of mis-
folded and aggregated, partially active antitoxin,
which might be sufficient to inhibit the effect of the
toxin. Such a hypothesis is in agreement with the
observation that robust overexpression of the anti-
toxin can also prevent growth without chaper-
one.34,41 Accordingly, we found that in the
Figure 1. The M. tuberculosis TAC system Mtb-HigA1
absence of proteases, the cellular levels of antitoxin
antitoxin is degraded by AAA+ proteases in E. coli (A)
increased (Figure 1(B)), albeit mainly in an insoluble
E. coli AAA + protease deletion allows the chaperone-
form in the absence of chaperone (Supplementary
less Mtb-HigA1 to suppress Mtb-HigB1 toxicity. E. coli
Figure S1). Importantly, reduced proteolysis alone
W3110 wild type (WT), DclpA, DclpX, DclpAX, DclpP,
does not appear to replace the need for a TAC
Dlon, DhslU, DhslV, DftsH and D3 (DclpXP, Dlon, DhslV)
chaperone, at least in the context of E. coli. This is
strains were transformed with pK6-Mtb-HigBA1, and
illustrated by the fact that growth rescue is less effi-
both WT and D3 strains were co-transformed with pK6-
cient in the triple protease mutation than in the pro-
Mtb-HigBA1 and pSE-Mtb-SecBTA. Transformants or
tease containing strain expressing the dedicated
double transformants were grown to mid-log phase,
chaperone (Figure 1(A); compare D3 to and
serially diluted and spotted on LB agar plates containing
WT + SecBTA).
kanamycin (and ampicillin for double transformants) and
We next asked whether the sensitivity of Mtb-
arabinose inducer as indicated. Plates were incubated at
HigA1 to proteolysis in vivo, in the absence of
37 °C overnight, except for the DftsH temperature-
chaperone, is a conserved feature of TAC
sensitive mutant, which was incubated at 30 °C. (B) Mtb-
antitoxins. To address this, four different TAC
HigA1 is stabilized in a strain of E. coli deleted for the
systems from distant bacteria were analyzed: the
main AAA+ proteases. Steady-state level of Mtb-HigA1
previously described (i) Mmet-TAC from
in E. coli W3110 WT and D3 strain co-transformed with
Methylomonas methanica strain MC09 with a TA
pK6-Mtb-HigBA1 and pSE380DNcoI or with pK6-Mtb-
couple belonging to the HicAB family, (ii) Vcho-
HigBA1 and pSE-Mtb-SecBTA. Expression of Mtb-
TAC from Vibrio cholerae strain RC385 with a TA
higBA1 and Mtb-secBTA was induced at mid-log phase
couple belonging to the MqsRA family, (iii) Glov-
with 0.2% Ara and 50 mM IPTG for 1 h. Mtb-HigA1 was
TAC from Geobacter lovleyi strain ATCC BAA-
detected in whole-cell extracts using a rabbit anti-Mtb-
1151 with a TA couple belonging to the HigBA
HigA1 antibody.
family,39,43 and (iv) the newly identified Saci-TAC
from Syntrophus aciditrophicus strain SB with a
subunits that associate with ClpP, are deleted. TA couple belonging to the MqsRA family.32 Strik-
Notably, deletion of clpX or clpA alone was not suf- ingly, in all four cases the absence of cytosolic pro-
ficient to restore toxin inhibition by the chaperone- teases promotes growth of E. coli expressing each
less Mtb-HigBA1 (Figure 1(A)), suggesting that chaperoneless TA pair (Figure 2), as is the case
3
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

for Mtb-TAC (Figure 1(A)). Yet, we noticed that tox- tions together with two different AAA+ regulatory
icity of the Vcho-mqsRA system of V. cholerae was subunits encoded by clpX and clpC1 genes.48–50
somehow less efficiently inhibited by the absence of Note that M. tuberculosis also contains another
E. coli proteases, as judged by the appearance of ClpC protein (ClpC2), which lacks a characteristic
smaller and translucent colonies (Figure 2). AAA+ module (and hence IGL-loop for interaction
Although growth rescue was slightly better at higher with ClpP), and currently there is no evidence that
growth temperatures (Supplementary Figure S1), ClpC2 can cooperate with the ClpP protease, thus
the level of rescue for the Vcho-mqsRA system only bona fide AAA+ partners of ClpP1P2 (i.e. ClpX
was consistently less than that observed in the and ClpC1) were considered here. Initially to deter-
other systems tested, suggesting that the Vcho- mine which M. tuberculosis AAA+ protease
mqsA antitoxin might not be very active in the (ClpXP1P2 or ClpC1P1P2) was responsible for
absence its chaperone when expressed in E. coli Mtb-HigA1 antitoxin degradation, both mycobacte-
and/or that other E. coli proteases might also rial proteases were purified and in vitro degradation
degrade it. Collectively, these data strongly suggest assays of Mtb-HigA1 were performed. Here, the
that the TAC toxin-activation cycle is indeed regu- AAA+ subunits of M. tuberculosis were used
lated by proteolysis. together with the proteolytic complex ClpP1P2 of
the closely related Mycobacterium smegmatis,
M. tuberculosis ClpXP1P2 degrades Mtb-HigA1 which shows 88% identity and can fully cooperate
antitoxin with the AAA+ subunits of M. tuberculosis.51 We
found that the ClpP1P2-dependent turnover of
In light of the above results, we next investigated Mtb-HigA1 was mediated exclusively by Mtb-ClpX,
whether M. tuberculosis ClpP protease was indeed and not by Mtb-ClpC1 (Figure 3(A)). As a control,
involved in the turnover of Mtb-HigA1 antitoxin. In the purified ClpC1P1P2 protease was active
contrast to E. coli, M. tuberculosis lacks both Lon in vitro, as judged by its ability to degrade FITC-
and ClpQ.47 The M. tuberculosis ClpP protease casein (Supplementary Figure S2). Strikingly, addi-
(ClpP1P2) is composed of two different subunits, tion of the TAC chaperone Mtb-SecBTA to the reac-
encoded by the clpP1 and clpP2 genes, and func- tion, completely inhibited antitoxin turnover
(Figure 3(A)). To further confirm the role of Mtb-
ClpXP1P2 in TAC antitoxin degradation, we took
advantage of our E. coli triple protease mutant
and used it as a tool to co-express the Mtb-
ClpXP1P2 protease together with Mtb-HigBA1 to
determine if expression of the protease could
restore growth inhibition by the toxin. Specifically,
Mtb-clpX and clpP1P2 were cloned, into a compat-
ible plasmid, under the control of IPTG-inducible
promoters and expressed in our in vivo growth res-
cue assay. As observed with E. coli proteases, Mtb-
HigBA1 expression only inhibited growth when all
components of Mtb-ClpXP1P2 protease were co-
expressed (Figure 3(B)). In addition, such a Mtb-
ClpXP1P2 dependent growth defect was efficiently
rescued by co-expression of the Mtb-SecBTA chap-
erone (Supplementary Figure S1). Together these
data show that Mtb-HigA1 antitoxin turnover is
specifically mediated by ClpXP1P2 and that this
process is inhibited by the dedicated SecB-like
chaperone.

Residues at the extreme carboxy-terminal end


of ChAD are critical for degradation by Mtb-
Figure 2. Other TAC systems also rely on proteases.
ClpXP1P2
Strains W3110 WT and D3 were transformed with pK6
plasmids harboring the chaperoneless TAC system Next, we sought to identify the degradation motif
HicAB of Methylomonas methanica, MqsRA of Vibrio (degron) recognized by the protease. Given that
cholerae strain RC385, HigBA of Geobacter lovleyi or substrate recognition by many proteases,
MqsRA of Syntrophus aciditrophicus (pK6-Mmet-HicAB, including the Clp-proteases, is generally mediated
pK6-Vcho-MqsRA, pK6-Glov-MqsRA and pK6-Saci- by a degron located at either the N- or C-termini
MqsRA respectively). Transformants were grown to (N-degron and C-degron, respectively)52 we exam-
mid-log phase, serially diluted and spotted on LB agar ined the sequence of HigA for potential degrons.
plates containing kanamycin and arabinose as From this analysis we identified a C-terminal dihy-
indicated. drophobic motif in Mtb-HigA1 (144RQVEVA149),
4
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

A
Mtb-HigA1
Time (min) 0 15 30 60
ClpXP1P2

ClpC1P1P2

ClpXP1P2
+ Mtb-SecBTA

B E. coli Δ3
Mtb-HigBA1 + + + +
ClpP1P2 - - + +
ClpX - + - +

no Ara

Ara 0.2%

-Mtb-HigB1

Figure 3. Mtb-HigA1 is degraded by the ClpXP1P2 protease of M. tuberculosis. (A) Mtb-HigA1 is degraded by M.
tuberculosis ClpXP1P2 in vitro. Degradation of purified Mtb-HigA1 by ClpXP1P2, with (green bars) and without (blue
bars) Mtb-SecBTA, or with ClpC1 of M. tuberculosis (red bars) was assayed in vitro. Time points were taken at 0, 15,
30 and 60 min of incubation at 37 °C, and Mtb-HigA1 was detected by immunoblot using rabbit anti-Mtb-HigA1
antibodies and quantified using GelEval 1.37. (B) Mtb-HigB1 toxicity is restored in E. coli D3 strain when Mtb-HigBA1
is co-expressed with Mtb-ClpXP1P2. E. coli D3 was co-transformed with pK6 plasmids either empty or harbouring
Mtb-HigBA1 (pK6-Mtb-HigBA1) and pSE380DNcoI plasmid harbouring either M. tuberculosis ClpP1P2, ClpX or
ClpXP1P2 (pSE-Mtb-ClpP1P2, pSE-Mtb-ClpX and pSE-Mtb-ClpXP1P2 respectively). Co-transformants were grown
to mid-log phase, serially diluted and spotted on LB agar plates containing kanamycin, ampicillin, arabinose (as
indicated) and IPTG (5 mM). Plates were incubated at 37 °C overnight. Lower panel shows whole-cell extracts of the
same co-transformants obtained after growth and induction at mid-log phase with 0.2% arabinose and 50 mM IPTG for
1 h, Mtb-HigB1 was detected using rabbit anti-Mtb-HigB1 antibody.

which shared similarity with the “classic” bacterial both residues (V148 and A149) and two additional
ClpX C-degron (i.e. the SsrA-tag) (Figure 4(A)). single point mutants of Mtb-HigA1, in which V148
To determine the significance of this motif, we or A149 were replaced with asparate (V148D and
replaced the last two residues of Mtb-HigA1 (resi- A149D, respectively). Consistent with the double
dues 148 and 149) with aspartate (Figure 4(A)) mutant, both the deletion mutant and each single
and tested it in our in vivo assay. Strikingly, the mutant (V148D and A149D) rescued the growth
Mtb-HigA1V148D/A149D mutant rescued growth effi- inhibition phenotype of the toxin in vivo when
ciently in the absence of chaperone (Figure 4(B)), expressed in E. coli (Figure 4(B)). Accordingly, the
suggesting that the last two amino acids of the anti- Mtb-HigA1V148D/A149D mutant was also able to coun-
toxin are indeed key residues for recognition by teract HigB1 toxicity in M. smegmatis in the
Mtb-ClpX. To confirm the importance of this region absence of the chaperone (Figure 4(C)). In this
and the relative importance of each C-terminal resi- case, the levels of the mutant antitoxin were stabi-
due we generated a mutant of Mtb-HigA1 lacking lized significantly in vivo, as judged by the accumu-
5
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

Figure 4. Mutation of the C-degron of Mtb-HigA1 prevent its turnover in vivo. (A) Schematic representation of Mtb-
HigA1 with the middle Helix-Turn-Helix (HTH) DNA binding domain and the C-terminal ChAD region. The last 7
residues (143–149) of Mtb-HigA1 and the underlined V148D/A149D mutations are shown. (B) Mtb-HigA1V148D/A149D
mutants counteract Mtb-HigB1 toxicity in E. coli independent of the presence of Mtb-SecBTA. E. coli W3110 strain was
co-transformed with either pK6-Mtb-HigBA1, pK6-Mtb-HigBA1V148D/A149D, its single mutant derivatives or the degron
deletion mutant pK6-Mtb-HigBA1DV148/A149, together with pSE380DNcoI or pSE-Mtb-SecBTA, grown to mid-log phase,
serially diluted and spotted on LB agar plates containing kanamycin and ampicillin supplemented with IPTG and/or
arabinose as indicated. Plates were incubated at 37 °C overnight. (C) Chaperoneless Mtb-HigA1V148D/A149D
counteracts Mtb-HigB1 toxicity in M. smegmatis. M. smegmatis was electroporated with pLAM-Mtb-HigBA1 and
pLAM-Mtb-HigBA1V148D/A149D. After 4 h of incubation at 37 °C, 1/100 of the transformations were directly plated on LB
agar supplemented with kanamycin (5 mg/ml) and 0.2% acetamide. Plates were incubated at 37 °C for 48 h. (D) Mtb-
HigBA1V148D/A149D is stabilized in vivo in M. smegmatis. Cultures of M. smegmatis transformed with pLAM-Mtb-
HigBA1 or pLAM-Mtb-HigBA1V148D/A149D were induced 6 h in the presence (+) or absence (-) of acetamide inducer
(0.2%) at 37 °C. Steady state level of Mtb-HigA1 was monitored by western blot using rabbit anti-HigA1 antibody.

lation of Mtb-HigA1V148D/A149D in M. smegmatis To further investigate the role of the last two
cells lacking the chaperone (Figure 4(D)). We next residues of Mtb-HigA1 in the native host of Mtb-
co-expressed either Mtb-HigBA1 or Mtb- TAC (M. tuberculosis), we generated two
HigBA1V148D/A149D together with ClpXP1P2 in the chimeras of Photinus pyralis luciferase fused to
D3 protease mutant strain (Figure 5(A)). Consistent the C-terminus of HigA as a reporter system.
with an important role for antitoxin turnover by Specifically, the last 15 residues of Mtb-HigA1 or
ClpXP1P2 in the restoration of growth inhibition by Mtb-HigA1V148D/A149D, grafted to the C-terminal
the toxin (Figure 3(B)), expression of ClpXP1P2 in end of luciferase, were cloned into an integrative
the presence of Mtb-HigBA1V148D/A149D was unable vector (pGMC) under the control of a tetracycline
to decrease the steady state levels of antitoxin (Fig- inducible promoter, and expressed in the
ure 5(B)) or restore growth inhibition by the toxin pathogenic M. tuberculosis H37Rv strain.
(Figure 5(A)). Finally, the Mtb-HigA1V148D/A149D Luciferase activity was monitored directly following
mutant was purified and tested in our in vitro degra- cell lysis of strains grown to mid-log phase and
dation assay, in the presence of Mtb-ClpXP1P2. induced with Atc for one week. In agreement with
Consistent with our in vivo findings (above), and in the above work performed in E. coli and M.
contrast to wild type antitoxin (Figure 3(A)), the smegmatis, the chimera containing the grafted C-
Mtb-HigA1V148D/A148D mutant was resistant to Mtb- terminal residues of Mtb-HigA1V148D/A149D showed
ClpXP1P2 mediated degradation (Figure 5(C)). a strong luminescent signal, comparable to that of
6
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

E. coli
Mtb
Mtb

Mtb
Mtb
Mtb
Mtb

E. coli

Mtb

Figure 5. Mtb-HigA1V148D/A149D is not degraded by ClpXP1P2. (A) Mtb-HigB1 toxicity is prevented in E. coli D3
strain when Mtb-HigBA1V148D/A149D is co-expressed with Mtb-ClpXP1P2. E. coli D3 strain was co-transformed with
pK6-Mtb-HigBA1 or pK6-Mtb-HigBA1V148D/A149D together with pSE380DNcoI or pSE-Mtb-ClpXP1P2. Double
transformants were grown to mid-log phase, serially diluted and spotted on LB agar plates containing kanamycin,
ampicillin, arabinose and IPTG as indicated. Plates were incubated at 37 °C overnight. (B) Mtb-HigBA1, Mtb-
HigBA1V148D/A149D and Mtb-HigB1 steady state level in the presence of Mtb-ClpXP1P2. Whole-cell extracts of E. coli
D3 strain double transformants were obtained after growth and induction at mid-log phase with 0.2% arabinose and
50 mM IPTG for 1 h. Mtb-HigB1, Mtb-HigA1 and Mtb-HigA1V148D/A149D were detected using rabbit anti-Mtb-HigB1 and
anti-Mtb-HigA1 antibodies respectively. (C) Mtb-HigA1V148D/A149D is not degraded by ClpXP1P2 in vitro. The in vitro
turnover of purified Mtb-HigA1 (blue bars) or Mtb-HigA1V148D/A149D (red bars) was monitored in the presence of
M. smegmatis ClpP1P2 and M. tuberculosis ClpX, as performed in Figure 3(A).

7
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

luciferase alone, while the chimera with the grafted affecting the destabilizing effect of ChAD on the
C-terminal residues of wild type Mtb-HigA1 antitoxin, both in vivo and in vitro.39 We took advan-
displayed very weak luminescence tage of this mutation and showed in vivo in M. smeg-
(Supplementary Figure S3). Collectively these matis within the context of the tripartite TAC operon,
data strongly suggest that, as observed both that the single Y114A mutant is indeed unable to
in vitro and in vivo with E. coli and M. smegmatis, rescue bacterial growth unless it is combined with
the last 15 residues of Mtb-HigA1 are critical for the V148D/A149D mutation (Supplementary
recognition by M. tuberculosis proteases in vivo. Figure S4), further suggesting that chaperone-
binding and recognition by proteases do not occur
within the same stretch of residues in ChAD.
Protease resistant mutations in Mtb-
HigA1 do not affect chaperone binding
Discussion
As stated above, the TAC chaperone specifically
binds to the ChAD region of the TAC antitoxin and in This work shows that proteolysis plays a key role
its absence, the ChAD region induces aggregation in the control of the TAC system of M. tuberculosis
of the antitoxin both in vivo and in vitro.39 Therefore, and identifies the ClpXP1P2 protease as a main
we first asked whether mutations that affect ClpX player in this process. We demonstrated that the
recognition also affect chaperone (SecBTA) binding. degron (for recognition by ClpX) is located at the
To do so, we monitored complex formation of Mtb- extreme C-terminal end of the chaperone-binding
SecBTA chaperone with either Mtb-HigA1 or Mtb- region of HigA and that both chaperone binding
HigA1V148D/A149D by native gel separation in vitro, and protease recognition involve unique residues
as described.39 Under these conditions, we found of ChAD. Notably, we found that in contrast to
that chaperone binding to the protease resistant chaperone binding, the last two hydrophobic
mutant was unaffected (Figure 6(A)). Importantly, residues of the antitoxin are crucial for recognition
and consistent with published findings,39 the nega- by ClpX and that substitution of these residues
tive control mutant (Mtb-HigA1W108A/W137A) failed with aspartate inhibited antitoxin turnover. In
to interact with SecBTA (Figure 6(A)). We next used support of this, the extreme C-terminal residues of
an in vitro coupled transcription/translation assay to several ClpXP substrate proteins have been
determine if the protease resistant mutations in shown to play a crucial role in ClpX
Mtb-HigA1 altered the destabilizing effect of ChAD recognition,53,54 as is the case with the well-
on the antitoxin. Our data from Figure 6(B) clearly characterized ClpX-degron, the SsrA-tag, where
show that newly synthesized Mtb-HigA1V148D/A149D substitution of the last two alanine residues with
is largely insoluble in vitro, albeit slightly less than aspartate inhibits recognition by ClpX.54 Notably,
wild type Mtb-HigA1. This indicates that replace- both the position and composition of the Mtb-
ment of the dihydrophobic motif (Val-Ala) with the HigA1 degron resembles that of several other M.
dihydrophilic sequence (Asp-Asp) only weakly tuberculosis ClpP1P2 substrates identified so far,
impairs the overall destabilizing effect of Mtb- namely RsdA, SsrA, ClgR, WhiB1, CarD, VapB20,
ChAD in vitro (Figure 6(B)). Notably, as observed RelB1 and Hsp20.49,55–58 The fact that ClpXP1P2,
for wild type Mtb-HigA1, the presence of chaperone but not ClpC1P1P2, degrades Mtb-HigA1 in vitro
significantly enhances Mtb-HigA1V148D/A149D solu- suggests that these substrates are likely to be rec-
bility (Figure 6(B)), further supporting the view that ognized by the ClpX regulatory subunit. However,
these mutations do not affect chaperone binding. we cannot exclude the possibility that ClpC1 may
This is in sharp contrast with the Mtb-HigA1DC42 still play a part in Mtb-HigA1 recognition in vivo in
mutant, which lacks the entire ChAD region, and M. tuberculosis. Analysis of the last 15 amino acid
remains soluble in the absence of chaperone.39 In residues of all the known and putative M. tuberculo-
agreement with the weak but significant increased sis antitoxins show high sequence variability, even
solubility of Mtb-HigA1V148D/A149D compared to wild for antitoxins belonging to the same TA family
type in vitro, we found that the steady state levels of (Table S1). Only VapB19, VapB41, VapB44,
the soluble Mtb-HigA1V148D/A149D mutant accumu- VapB45, Rv2017 and HigA2 possess at least two
lates when overexpressed in wild type E. coli lack- hydrophobic residues at their extreme C-terminus.
ing the Mtb-SecBTA chaperone (Figure 6(C)). This These hydrophobic residues are mainly non-polar
suggests that the fraction of soluble HigA1V148D/ aliphatic ones (A, V, L, I) as is usually observed in
A149D
might be sufficient to rescue toxicity in vivo. C-terminal degrons. In addition, VapB41, VapB44,
The fact that the protease resistant mutant fully and VapB45 also possess acidic residues before
binds to the chaperone is supported by previous the hydrophobic end, as found in Mtb-HigA1. This
findings that show chaperone binding mainly suggests that these antitoxins might also be recog-
occurs within the region encompassing residues nized by M. tuberculosis ClpX. Remarkably, deple-
105 to 116 of Mtb-ChAD, with residues W108, tion of endogenous ClpP1P2 in M. tuberculosis was
R110 and Y114 largely involved in this shown to stabilize several antitoxins that harbor dif-
interaction.42 Notably, the Y114A mutation in ChAD ferent C-terminal ends, including MazE10, VapB22,
was shown to inhibit chaperone binding without VapB9, VapB41 and HigA149 (Table S1). Whether
8
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

Mtb
Mtb

Mtb

Mtb

Mtb
Mtb

Mtb Mtb

Figure 6. Degron mutation does not affect Mtb-HigA1 interaction with Mtb-SecBTA . (A) Mtb-HigA1 V148D/A149D
interacts with Mtb-SecB in vitro. Native PAGE separation of Mtb-SecB (16 lM) incubated with either Mtb-HigA1 or
TA TA

Mtb-HigA1 V148D/A149D (2, 4 or 8 lM). As a negative control, the non-interacting mutant Mtb-HigA1 W108A/W137A (8 lM) is
also shown. (B) Solubilization of Mtb-HigA1 V148D/A149D requires Mtb-SecBTA. Mtb-HigA1 and Mtb-HigA1 V148D/A149D
were expressed in a cell-free expression system with and without Mtb-SecBTA (8 lM). Translation products were labelled
with [35S]-methionine. After 1 h of incubation at 37 °C, the total (t) and soluble (s) fractions were separated on SDS/
PAGE and quantified by phosphorimager. The mean solubility values (%) and the standard deviation are indicated. (C)
Mtb-HigA1V148D/A149D is partially soluble in vivo. E. coli wild type strain was co-transformed with pSE380DNcoI or pSE-
Mtb-SecBTA, and Mtb-HigBA1 or pK6-Mtb-HigBA1V148D/A149D. Double transformants were grown to midlog phase, and
the expression of Mtb-HigBA1 or Mtb-HigBA1 V148D/A149D and Mtb-SecBTA was induced for 1 h using 0.1% arabinose
and 50 mM IPTG. Soluble and insoluble fractions were separated by SDS/PAGE and analyzed by Western Blot using
anti-Mtb-HigA1 antibody. (D) Schematic representation of the ChAD region (ChAD) of the Mtb-HigA1 TAC antitoxin (A)
as a hot spot for interaction with the SecB-like chaperone (C; folding pathway) and the ClpXP1P2 protease (X/P1/P2;
degradation pathway), leading to toxin (T) inhibition or activation, respectively.

these antitoxins are ClpX- or ClpC1-dependent search for Mtb-ClpC1 interactors in vivo using a
remains to be determined. Interestingly, a recent bacterial two hybrid system has identified TA sys-
9
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

tems a major substrate class of this protease, region required for interaction with the chaperone39
including 8 antitoxins and 16 toxins.56 These data and an extreme C-terminal region important for
indicate that indeed, activation of certain TA in M. AAA+ proteases recognition (Figure 6(D)). The
tuberculosis could rely on ClpC1. Yet, the fact that proximity of these motifs suggests that chaperone
toxins, as well as antitoxins, were identified as inter- binding could inhibit ClpXP1P2 binding, somehow
actors suggests that protease-mediated toxin masking the C-terminal degron from recognition
degradation could contribute to the inhibition of by ClpX. In this case, exposure of the chaperone-
toxin activity. Likewise, it is currently unknown if bound ChAD C-terminal end to stress could tran-
adaptor proteins such as ClpS play a role in the siently expose its degron to ClpXP, leading to anti-
recognition of these toxins and/or antitoxins.56,59,60 toxin degradation and subsequent growth
To date, AAA+ proteases are the only proteases inhibition. This proposal is reminiscent of the recog-
that have been involved in antitoxin degradation.61 nition of the C-terminal degron of DinJ by Lon, which
Yet, M. tuberculosis also encodes a eukaryotic- was shown to trigger dissociation of the DinJ/YafQ
like proteasome that could also participate to such complex in vitro.71 Alternatively, in the absence of
degradation.62,63 This proteasome was originally chaperone, destabilization of Mtb-HigA1 by the
described to drive the ATP-dependent degradation unprotected ChAD region39 could also trigger
of pupylated substrates. To date, no antitoxin has recognition by ClpXP1P2. In addition, we cannot
been reported to be pupylated.64 However, the M. exclude that ClpXP1P2 might also need to bind
tuberculosis pupylome was analyzed under stan- other unfolded parts of the Mtb-HigA1 antitoxin out-
dard laboratory growth conditions and since it has side the ChAD region in order to efficiently degrade
been shown that changes in growth conditions mod- it. Such regions would thus be folded and protected
ify the abundance of pupylated proteins we cannot in the presence of the chaperone, thus indirectly
exclude that antitoxins could be pupylated under interfering with ClpXP recognition.
certain stresses.65 Interestingly, the 20S core pro- The mechanism proposed in this work would
teasome can also be activated by PafE (also called represent a sensitive switch for the post-
Bpa) a dodecameric activator.66,67 Since PafE is an translational activation of TAC toxins. The fact that
ATP-independent activator of proteasomal degra- all the SecBTA chaperones tested so far can
dation, it is largely believed to mediate the turnover replace the E. coli export chaperone SecB in vivo
of partially or completely unfolded proteins and as suggests that the accumulation of aggregated pre-
such it has been proposed that it may replace the proteins (or the synthesis of specific exported
Lon protease in this bacterium.67 Interestingly, the proteins) under certain stress conditions could
fact that many antitoxins have been reported to be hijack the chaperone and subsequently induce
unstructured suggest that they could be putative antitoxin degradation by ClpXP1P2, which in turn,
targets of this degradation pathway. could lead to a transient activation of the Mtb-
In contrast with the antitoxins, four toxins, namely HigB1 toxin until normal growth conditions resume.
PemK, PhoH2, VapC17 and VapC31, were shown
to be pupylated.64,68 The relevance of such toxin Materials and methods
degradation is currently unknown, but it suggests
that some toxins could be differently regulated by Bacterial strains and culture conditions
proteasomal degradation and antitoxin inhibition
depending on stress conditions.69 Interestingly, E. coli strains W3110, BL21(kDE3) (Novagen),
the toxins, MazF3 and MazF7 were recently found BL21 AI(kDE3) (Novagen), DhslV DclpPX-lon::
to be Mtb-ClpC1-interaction partners56 although CmR,39 Mycobacterium smegmatis MC2155 (strain
the significance of this observation on MazEF regu- ATCC 700084) and Mycobacterium tuberculosis
lation is currently unknown. H37Rv (strain ATCC 25618) were previously
Our data show that the Mtb-SecBTA chaperone described. E. coli DclpA, DclpX, Dlon, DhslV and
efficiently protects the Mtb-HigA1 antitoxin from DhslU strains were constructed as follows. The cor-
degradation by ClpXP1P2 protease. Yet, it responding DclpA::KanR, DclpX::KanR, Dlon::
remains to be determined whether ClpXP1P2 KanR, DhslV::KanR or DhslU::KanR allele from
directly contribute to the Mtb-HigB1 toxin strain JWK3903 (Keio collection) was first moved
activation in vivo in the context of the native TAC in W3110 using P1-mediated transduction and the
operon. Interestingly, the E. coli SecB chaperone kanamycin resistance cassette was then removed
was previously shown to protect its pre-secretory using plasmid pCP20, as described.72 The E. coli
protein clients from degradation by the AAA+ DclpAX strain was obtained by moving the DclpX::
protease Lon, both in vivo and in vitro.70 Since SecB KanR allele from strain JWK3903 (Keio collection)
binding motifs and Lon recognition signals share in W3110 DclpA strain and by removing the kana-
significant similarities (enriched in hydrophobic mycin resistance cassette as described above. All
and aromatic residues), SecB binding to pre- DNA cloning experiments were carried out in
secretory substrates was proposed to mask Lon DH5a (NEB). E. coli strains were routinely grown
recognition signals.70 Here, we show that the C- in LB medium supplemented with kanamycin
terminal ChAD region combines both a middle (50 lg/ml), tetracycline (15 lg/ml) or ampicillin
(50 lg/ml) as required. Mycobacterium smegmatis
10
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

strains were grown in 7H9 medium supplemented fragment was ligated into p29SEN (TetR) digested
with Tween 80 (0.05% (v/v)) and kanamycin (5 lg/ with the same enzymes.
ml) and acetamide (0.2% (w/v)) as needed. To construct pLAM-Mtb-HigBA1 and pLAM-Mtb-
Mycobacterium tuberculosis H37Rv strain was HigA1 plasmids, higBA1 and higA1 were PCR
grown in 7H9 medium supplemented with 10% amplified using oligonucleotide primers HigB1_
OADC (oleic albumin dextrose catalase) supple- For_Fusion: 50 -aagggagtccacatatgccgccccctgatc
ment and Tween 80 (0.05% (v/v)), and zeocin cagccgccatg-30 and HigA1_Rev_Fusion: 50 -ataagc
(50 mg/ml) or anhydrotetracycline (Atc; 200 ng/ml) ttcgaattctcatgccacctcaacctgccgaac-30 or HigA1_
if necessary. For_Fusion: 50 -aagggagtccacatatgagcattgacttccctt
tgggtgac-30 and HigA1_Rev_Fusion: 50 -ataagcttc
gaattctcatgccacctcaacctgccgaac-30 , respectively,
Plasmids construct and pK6-Mtb-HigBA1 or pK6-Mtb-HigA141 as tem-
plates. The PCR fragments were then cloned by
Plasmids pMPMK6,73 pSE-Mtb-SecBTA, homologous recombination into the pLAM12 plas-
pET15b-His6HigA1, pET15b-His6Rv1957, pK6-Mtb- mid linearized by PCR with pLAM_For_Fusion: 50 -
HigBA1, pK6-Mtb-HigB1 and pK6-Mtb-HigA1,41 gaattcgaagcttatcgatg-30 and pLAM_Rev_Fusion:
pSE380DNcoI,74 pK6-Mmet-HicAB, pK6-Vcho- 50 -catatgtggactccctttc-30 primers using the In-
MqsRA, pK6-Glov-MqsRA, pET15b-His6 Fusion PCR cloning system (Clontech). The
HigA1W108A/W137A and pET-15b- His6HigA1Y114A,39 pLAM-Mtb-HigBA1 and pLAM-Mtb-HigA1 were
pET15b (Novagen), pET10C,75 pHUE76 and then used as templates to construct pLAM-Mtb-
pLAM1277 were previously described. The HigBA1V148D/A149D and pLAM-HigA1V148D/A149D
pET30a-Mtb-6HisClpC1 was kindly provided by Dr. respectively using Quickchange mutagenesis with
Dileep Vasudevan (Institute of Life Sciences, Bhu- higA1 codon change gtg to gac for V148D and
baneswar 751023, India) and was constructed as gca to gac for A149D.
described.78 The pSE-Mtb-ClpXP1P2 plasmid, encoding M.
Plasmids pK6-Mtb-HigA1 and pET15b-His6HigA1 tuberculosis ClpX (Rv2457c) and ClpP1P2 operon
were used as templates to construct pK6-Mtb- (Rv2461c-Rv2460c) under the control of two
HigBA1V148D/A149D, pK6-Mtb-HigBA1D(V148/A149), different Ptrc, was constructed as follows.
V148D
pK6-Mtb-HigBA1 , pK6-Mtb-HigBA1A149D and Rv2457c and the Rv2461c-Rv2460c operon were
V148D/A149D
pET15b-His6HigA1 respectively, using amplified by PCR using ClpX_For_Fusion: 50 -tctaga
Quickchange mutagenesis (Stratagene) with ctcctcacaattgatggcgcgcataggagacggtggtga-30 ’ and
higA1 codon change gtg to gac for V148D, gtg of ClpX_Rev_Fusion: 50 -caaaacagccaagcttctacgcgct
V148 to taa stop for D(V148/A149), and gca to cttgtcgcggcgctcc-30 or ClpP1P2_For_Fusion: 50 -a
gac for A149D. Plasmid pK6-Mtb-HigB1K95A and caggagaattccagatgagccaagtgactgacatgcgttcga-30
pET-20b-Mtb-HigB1optK95A were obtained by and ClpP1P2_Rev_Fusion: 50 -caattgtgaggagtcta
Quickchange mutagenesis of pK6-Mtb-HigB1 and gatcaggcggtttgcgcggagagcttccg-30 , respectively,
pET-20b-Mtb-HigB1opt respectively, with higB1 and M. tuberculosis genomic DNA as a template.
codon change aag to gcg for K95A. Plasmid pET- These two PCR fragments were then cloned
20b-Mtb-HigB1opt encoding higB1 sequence using the In-Fusion PCR cloning system into the
optimized for E. coli codon usage between NdeI pSE380DNcoI linearized by PCR using pSE_
and XhoI restriction site was synthesized by For_Fusion: 50 -aagcttggctgttttggcgg-30 and pSE_
Genscript. Rev_Fusion: 50 -ctggaattctcctgtgtgaa-30 primers.
Plasmid pK6-Saci-MqsRA was obtained by PCR To add the second Ptrc upstream of Rv2457c, the
amplifying the genomic region comprising resulting plasmid was transformed into an E. coli
the mqsR toxin SYN_00016 and the mqsA antitoxin dam-/dcm- strain, purified, digested with XbaI-
SYN_00015 using primers T_Saci_FOR_E: 50 -ttgaa MfeI and ligated to the DNA region of the
ttcatgatacaaaaaatcagtgaaaaacgc-30 and A_Saci_ pSE380DNcoI plasmid encoding the Ptrc that was
REV_H/Bam: 50 -ttaagcttggatccctatgatccatatgtcgcc amplified by PCR using the oligonuceleotide
tctttgg-30 , and S. aciditrophicus SB genomic DNA primers Ptrc_For: 50 -gactctagacgcaacgcaattaatgtg
(gift from Michael J. McInerney, University of a-30 and Ptrc_Rev: 50 -gtccaattgttctcctgtgtgaaattgt-
Oklahoma, USA) as a template. The PCR product 30 and digested with the same restriction
was digested with EcoRI–HindIII and was ligated enzymes. To construct the pSE-Mtb-ClpP1P2, the
into pK6 digested with the same enzymes. To pSE-Mtb-ClpXP1P2 was digested with XbaI-
construct pK6-Mtb-SecBTA, Mtb-SecBTA was HindIII in order to remove the Ptrc-ClpP1P2
amplified from the pSE-Mtb-SecBTA using 1957_F_ encoding region. Extremities were blunted using
(ERI/NdeI):50 -gagaattcatatgactgaccgaaccgacgccgac DNA polymerase I, Large (Klenow) fragment
-30 and 1957_R_(HIII/BHI): 50 -gaaagcttggatcct (NEB) and ligated. To construct the pSE-Mtb-
cagggcgttcctctcgttgccggc-30 primers. The PCR ClpX, Rv2457c was amplified by PCR using the
product was then digested by EcoRI and HindIII oligonucleotide primers ClpX_For_NcoI: 50 -gacc
and ligated into the pK6 vector digesting with the catggcgcgcataggagacgg-30 and ClpX_Rev_
same enzymes. To generate plasmid p29T-Mtb- HindIII: 50 -gtcaagcttctacgcgctcttgtcgcggc-30 and
SecBTA, the EcoRI/HindIII-digested Rv1957
11
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

M. tuberculosis genomic DNA as a template. The pH 7.5, 150 mM NaCl, 1 mM DTT, 20% (v/v) glyc-
PCR product was then digested with NcoI-HindIII erol, 8 M urea) were obtained following over-
and ligated into the pSE380 plasmid digested with expression from pET15b-His6HigA1, pET15b-His6
the same enzymes. HigA1V148D/A149D and pET15b-His6
To construct the pGMC-Luc plasmid, the luciferase HigA1W108A/W137A in E. coli BL21(kDE3) from the
gene from P. pyralis was PCR amplified using aggregated fractions.41 Complexes were formed
Luc_For_Fusion: 50 -gaagacaggctgcccatggaagacgc by mixing 2, 4 or 8 mM of unfolded Mtb-HigA1,
caaaaacataaag-30 and Luc_Rev_Fusion: 50 -tgtataa Mtb-HigA1V148D/A149D or Mtb-HigA1W108A/W137A
taaagttgtcatctagacacggcgatctt-30 primers, and pK6- and 16 mM of Mtb-SecBTA at 37 °C for 15 min in
Luc39 as a template. The PCR product was then 10 ml reaction buffer (30 mM Hepes pH 7.5,
cloned into the pGMC plasmid linearized using 40 mM KCl, 7 mM MgAc and 50 mM NaCl). Reac-
pGMC_For_Fusion: 50 -caactttattatacatagttgataattc- tion mixtures were separated using native PAGE
30 and pGMC_Rev_Fusion: 50 -gggcagcctgtcttcctc-30 on 4–20% Mini-PROTEAN TGX precast gels (Bio-
primers with the In-Fusion PCR cloning system. Rad) using Tris/Glycine buffer, and run at 125 V
The pGMC-Luc-CtHigA1 plasmid, encoding the luci- for 2 h at room temperature. Gels were stained with
ferase in fusion with the last 15 residues of Mtb- InstantBlue (Expedeon).
HigA1, was constructed as follows. The pK6-Luc-
Mtb-ChAD39 was first amplified by PCR using
oligonucleotide primers Luc_CtHigA1_For: 50 -acgaa Cell-free protein synthesis and centrifugation-
ctgggcccgccacatttc-30 and Luc_CtHigA1_Rev: 50 -tct based aggregation assay
agacacggcgatctttccg-30 in order to remove the coding Cell-free transcription/translation-coupled in vitro
region of the first 27 ChAD residues, extremities were assay using the PURE system (NEB) was carried
phosphorylated with T4 polynucleotide kinase (NEB) out as described.39 Briefly, DNA of Mtb-HigA1 and
and ligated. The resulting pK6-Luc-CtHigA1 was Mtb-HigA1V148D/A149D were amplified by PCR using
then used as a template to subclone Luc-Mtb- primers containing T7 promoter and terminator and
CtHigA1 into the pGMC vector with the In-Fusion added at a final concentration of 7.2 ng/ll to the
PCR cloning system using the Luc_For_Fusion and PURE system. When indicated, purified Mtb-
CtHigA1_Rev_Fusion: 50 -tgtataataaagttgtcatgccacct SecBTA was added to a final concentration of
caacctgccg-30 . Plasmid pGMC-Luc-CtHigA1V148D/ 8 mM. Protein synthesis was performed at 37 °C
A149D
was constructed using the same protocol but for 60 min in the presence of 0.4 lCi/ll of 35S-
using the pK6-Luc-Mtb-ChADV148D/A149D as a tem- methionine. After removing an aliquot (total frac-
plate. Note that the pK6-Luc-Mtb-ChADV148D/A149D
tion), the rest of the reaction was centrifuged at
was obtained by Quickchange mutagenesis with
21,600g for 30 min and the supernatant fraction
higA1 codon change gtg to gac for V148D and gca
was collected. Both the total and supernatant frac-
to gac for A149D using pK6-Luc-Mtb-ChAD as a
tions were separated by SDS/PAGE on 4–20%
template. To construct the pGMC-TAC plasmid, the
TAC operon from M. tuberculosis was amplified by Mini-Protean TGX gels (Bio-Rad). Proteins were
PCR using HigB1_For_Fusion 50 -gaagacaggc visualized using a Typhoon phosphorimager (GE
tgcccatgccgccccctgatccagccgccatg-30 and Rv1957_ Healthcare) and analyzed with Multigauge software
Rev_Fusion: 50 - tgtataataaagttgtcagggcgttcctctcgttg (Fuji). Results are presented as the mean and stan-
ccgg-30 primers, and genomic DNA as template. dard deviation (s.d.) of triplicate experiments.
The PCR product was then cloned into the pGMC
plasmid linearized using pGMC_For_Fusion and Protein purification
pGMC_Rev_Fusion primers with the In-Fusion
PCR cloning system. To construct the pET-15b- Mtb-ClpX and Mtb-ClpC1 were overexpressed
Mtb-His6ClpX plasmid, Rv2457c gene was amplified from a BL21 (kDE3) strain containing plasmid
from M. tuberculosis genomic DNA using pET-15b-Mtb-His6ClpX and pET-30a-Mtb-His6ClpC1
pET_ClpX_For: 50 -gaccaattgcatatgatggcgcgcatagga respectively. Cells were grown to an OD600 of 0.8
gac-30 and pET_ClpX_Rev: 50 - gacaagcttag in LB medium containing ampicillin at 37 °C, and
atctctacgcgctcttgtcgcg-30 . The resulting PCR product Mtb-ClpX and Mtb-ClpC1 expression was induced
was then digested with NdeI and BglII restriction with 0.5 mM IPTG overnight at 16 °C for Mtb-ClpX
enzymes and subsequently ligated into the pET- and 4 h at 37 °C for Mtb-ClpC1. Cells were
15b vector digested with NdeI and BamHI restriction harvested by centrifugation, resuspended in lysis
enzymes. The clpP1 and clpP2 genes were cloned buffer [50 mM Tris-HCl pH 8.0, 200 mM KCl, 10%
into pET10C and pHUE, respectively as described.51 (v/v) glycerol, 1 mM EDTA, 0.05% (v/v) Triton
X-100, 1 mM DTT, protease inhibitors (Roche)]
and lysed by successive steps, snap frozen in
Native gel separation of purified protein
liquid nitrogen and sonicated after addition of
complexes
1 mg/ml lysozyme and 1 ml/ml benzonase. The
Mtb-SecBTA was purified following lysate was clarified by centrifugation at 30 000g
overexpression from pET15b in E. coli BL21 for 40 min at 4 °C. A Ni-NTA agarose column
(kDE3) as described.41 Purified Mtb-HigA1 WT, (Qiagen) pre-equilibrated with buffer A [20 mM
V148D/A149D and W108A/W137A (in 20 mM Tris, Tris-HCl pH 8.0, 300 mM NaCl, 10 mM imidazole,
12
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

1 mM DTT] was loaded with the supernatant and cells and the supernatant was kept as the whole-
washed with 5 bead volumes (BV) of buffer A and cell extract fraction. Proteins were separated on
5 BV of buffer B [20 mM Tris-HCl pH 8.0, 300 mM Mini-Protean TGX gels (Bio-Rad) by SDS/PAGE
NaCl, 20 mM imidazole, 1 mM DTT]. Elution and transferred to polyvinylidene difluoride
fractions were collected with buffer C [20 mM Tris- membranes (Bio-Rad) using the Trans-BlotÒ
HCl pH 8.0, 300 mM NaCl, 250 mM imidazole, TurboTM transfer system (Bio-Rad). Membranes
1 mM DTT, 20% (v/v) glycerol]. Fractions were blocked 1 h at room temperature or
containing the Mtb-ClpX and Mtb-ClpC1 proteins overnight at 4 °C in 5% (w/v) nonfat dry milk in
were pooled and snap-frozen in liquid nitrogen. PBS containing 0.05% (v/v) Tween 20. Primary
Msm-ClpP1 was overexpressed from a BL21 antibodies used in this study were anti-Mtb-HigA1
(kDE3) strain containing plasmid pET10C-Msm- (1:1000), anti-Mtb-HigB1 (1:1000) and anti-Mtb-
ClpP1 and purified as described above for Mtb- SecBTA (1:1000).39,41 Horseradish peroxidase-
ClpX except that an additional washing step was conjugated rabbit IgG was used as a secondary
performed with 5 BV of buffer D [20 mM Tris-HCl antibody. Blots were developed by chemilumines-
pH 8.0, 300 mM NaCl, 65 mM imidazole, 1 mM cence using Clarity western ECL substrate (Bio-
DTT] before the bound proteins were eluted with Rad) with the ChemidocTM Touch imaging system
buffer E [20 mM Tris-HCl pH 8.0, 300 mM NaCl, (Bio-Rad) and analysed with Image Lab software
500 mM imidazole, 1 mM DTT, 20% (v/v) (Bio-Rad).
glycerol]. Msm-ClpP2 was purified in its processed
form (lacking the first 17 residues) and was
overexpressed from a BL21 CodonPlus (kDE3) In vitro degradation assay
strain containing plasmid pHUE-Msm-pClpP2
allowing expression of a His6-Ub-fusion protein. Protein degradation assays using Mtb-HigA1 or
Msm-ClpP2 was first purified using Ni-NTA Mtb-HigA1V148D/A149D as a substrate were
agarose (Qiagen) as described above for Mtb- performed as follows. 0.1–0.5 mM of substrate
ClpX. Then, the His6-Ub was cleaved using His6- (estimated by dilution of the denaturated stock)
Usp2cc as described76 and the untagged protein was mixed in buffer A (50 mM K3PO4, 100 mM
was isolated by reverse affinity chromatography.51 KCl, 5% (v/v) glycerol, 50 mM MgCl2, 2 mM ATP,
0.5 mM Z-Leu-Leu) with 2.4 mM Mtb-ClpX or Mtb-
Western blot analysis ClpC1, and 2.8 mM of Msm-ClpP1 and Msm-
ClpP2, and incubated at 37 °C. 5 mM of Mtb-
For E. coli fractionation, overnight cultures of SecBTA was added when indicated. The reaction
fresh transformants were diluted and grown to was stopped, at the indicated time points, by the
mid-log phase at 37 °C. After induction, OD600 addition of Laemmli buffer and the samples were
was measured and 1 ml aliquot of the culture was heated at 95 °C for 5 min. Proteins were
harvested at 5 000 g. Whole-cell extract was separated by SDS/PAGE and analyzed by
prepared by pellet resuspension in ¼ volume of western-blot using anti-Mtb-HigA1 antibodies.
the initial OD600 of 1  SDS loading buffer. To Proteins were quantified using FrogDance
separate the insoluble and soluble fractions, 10 ml software GelEval 1.37 or GelAnalzyer software
of cells were collected by centrifugation at 5000g 2010a and the results are presented as the mean
for 10 min at 4 °C and resuspended in 300 ll of and SEM of triplicate experiments.
fresh lysis buffer [10 mM Tris (pH 8), 100 mM KCl,
300 mM NaCl, 5 mM b-mercaptoethanol, 0.5% (v/
v) Triton X-100, and 0.6 mg/ml lysozyme].
Luminescence assay
Samples were incubated on ice for 20 min and the
cells were lysed by three freeze–thaw cycles Luminescence assay were carried out on freshly
using liquid nitrogen followed by sonication. The prepared whole-cell extracts of M. tuberculosis
insoluble and soluble fractions were then H37Rv strain containing the integrative plasmid
separated by centrifugation for 30 min at 4 °C (16 pGMC-Luc, pGMC-LucCtHigA1 or pGMC-
000g). The pellet was resuspended in 150 ll of LucCtHigA1 V148D/A149D. To this purpose, mid-log
1  SDS loading buffer, and 5  SDS loading phase cultures were induced with 200 ng/ml
buffer was added to the supernatant. anhydrotetracycline for 1 week and 5 ml aliquots
M. smegmatis whole-cell extracts were prepared were harvested at 4000 g. Cells were
as follows. 48 h cultures of fresh transformants resuspended in 400 ml of PBS, mixed with glass
were diluted and grown to mid-log phase. After beads and mechanically lysed using a bead
6 h of induction with 0.2% acetamide, OD600 beater (Bertin technologies precellys 24). After
were measured and aliquots of the cultures were double filtration (0.22 mm), 20 ml of the resulting
harvested at 4 000 g for 10 min. Pellets were whole-cell extracts were mixed with 16.6 ml of
resuspended in 1x PBS to an OD600 per ml of 5, Luciferase Assay Substrate (Promega) diluted in
and cells mechanically lysed using a bead beater 33.3 ml of 1x PBS, and luminescence was directly
(Bertin technologies precellys 24). Lysates were monitored at an exposure time of 100 milliseconds
centrifuged at 1000g for 3 min to pellet unbroken using the VarioskanFlash.
13
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

CRediT authorship contribution References


statement
1. Goeders, N., Van Melderen, L., (2014). Toxin-antitoxin
Pauline Texier: Conceptualization, Investigation, systems as multilevel interaction systems. Toxins, 6, 304–
Funding acquisition. Patricia Bordes: 324. https://doi.org/10.3390/toxins6010304.
Conceptualization, Investigation, Supervision. 2. Harms, A., Brodersen, D.E., Mitarai, N., Gerdes, K., (2018).
Jyotsna Nagpal: Conceptualization, Investigation. Toxins, targets, and triggers: an overview of toxin-antitoxin
biology. Mol. Cell., 70, 768–784. https://doi.org/10.1016/
Ambre Julie Sala: Conceptualization,
j.molcel.2018.01.003.
Investigation. Moise Mansour: Conceptualization,
3. Page, R., Peti, W., (2016). Toxin-antitoxin systems in
Investigation. Anne-Marie Cirinesi:
bacterial growth arrest and persistence. Nat. Chem. Biol.,
Conceptualization, Investigation. Xibing Xu: 12, 208–214. https://doi.org/10.1038/nchembio.2044.
Conceptualization, Investigation. David Andrew 4. Sala, A., Bordes, P., Genevaux, P., (2014). Multiple toxin-
Dougan: Conceptualization, Supervision, Funding antitoxin systems in Mycobacterium tuberculosis. Toxins,
acquisition. Pierre Genevaux: Conceptualization, 6, 1002–1020. https://doi.org/10.3390/toxins6031002.
Supervision, Funding acquisition. 5. Goormaghtigh, F., Fraikin, N., Putrinš, M., Hallaert, T.,
Hauryliuk, V., Garcia-Pino, A., Sjödin, A., Kasvandik, S.,
et al., (2018). Reassessing the role of Type II toxin-antitoxin
systems in formation of Escherichia coli Type II persister
Acknowledgements cells. MBio., 9 https://doi.org/10.1128/mBio.00640-18.
6. Pontes, M.H., Groisman, E.A., (2019). Slow growth
We thank Marie-Pierre Castanié-Cornet for determines nonheritable antibiotic resistance in
plasmid gift. This work was supported by the Salmonella enterica. Sci. Signal., 12 https://doi.org/
Agence Nationale de la Recherche (ANR-13- 10.1126/scisignal.aax3938.
BSV8-0010-01 and ANR-19-CE12-0026) to P.G., 7. LeRoux, M., Culviner, P.H., Liu, Y.J., Littlehale, M.L., Laub,
the Fondation pour la Recherche Médicale to P.T. M.T., (2020). Stress can induce transcription of toxin-
antitoxin systems without activating toxin. Mol. Cell., 79,
(FRM-FDT20160435132) and to M.M. (FRM
280–292.e8. https://doi.org/10.1016/j.molcel.2020.05.028.
FDT201805005796) and an EMBO short term
8. Harms, A., Fino, C., Sørensen, M.A., Semsey, S., Gerdes,
fellowship to P.T.
K., (2017). Prophages and growth dynamics confound
experimental results with antibiotic-tolerant persister cells.
Author contributions MBio, 8 https://doi.org/10.1128/mBio.01964-17.
9. Agarwal, S., Sharma, A., Bouzeyen, R., Deep, A., Sharma,
Conceptualization, all authors. Investigation, P. H., Mangalaparthi, K.K., Datta, K.K., Kidwai, S., et al.,
T., P.B., J.N., A.J.S., M.M., A-M.C., X.X. Writing, (2020). VapBC22 toxin-antitoxin system from
P.T, P.B, D.A.D, P.G. Funding acquisition, P.T., Mycobacterium tuberculosis is required for pathogenesis
D.A.D, P.G Supervision, P.B., D.A.D., P.G. and modulation of host immune response. Sci. Adv., 6,
eaba6944. https://doi.org/10.1126/sciadv.aba6944.
Declaration of Competing Interest 10. Correia, F.F., D’Onofrio, A., Rejtar, T., Li, L., Karger, B.L.,
Makarova, K., Koonin, E.V., Lewis, K., (2006). Kinase
The authors declare that they have no known activity of overexpressed HipA is required for growth arrest
competing financial interests or personal and multidrug tolerance in Escherichia coli. J. Bacteriol.,
relationships that could have appeared to 188, 8360–8367. https://doi.org/10.1128/JB.01237-06.
11. Helaine, S., Cheverton, A.M., Watson, K.G., Faure, L.M.,
influence the work reported in this paper.
Matthews, S.A., Holden, D.W., (2014). Internalization of
Salmonella by macrophages induces formation of
Appendix A. Supplementary material nonreplicating persisters. Science, 343, 204–208. https://
doi.org/10.1126/science.1244705.
Supplementary data to this article can be found 12. Kwan, B.W., Lord, D.M., Peti, W., Page, R., Benedik, M.J.,
online at https://doi.org/10.1016/j.jmb.2021. Wood, T.K., (2015). The MqsR/MqsA toxin/antitoxin
166815. system protects Escherichia coli during bile acid stress.
Environ. Microbiol., 17, 3168–3181. https://doi.org/
Received 5 November 2020; 10.1111/1462-2920.12749.
Accepted 5 January 2021; 13. Pecota, D.C., Wood, T.K., (1996). Exclusion of T4 phage
Available online 13 January 2021 by the hok/sok killer locus from plasmid R1. J. Bacteriol.,
178, 2044–2050. https://doi.org/10.1128/jb.178.7.2044-
Keywords: 2050.1996.
Toxin-antitoxin; 14. Ronneau, S., Helaine, S., (2019). Clarifying the link
HigB1-HigA1; between toxin-antitoxin modules and bacterial
SecB; persistence. J. Mol. Biol., 431, 3462–3471. https://doi.org/
ClpX; 10.1016/j.jmb.2019.03.019.
ClpC1 15. Talwar, S., Pandey, M., Sharma, C., Kutum, R., Lum, J.,
Carbajo, D., Goel, R., Poidinger, M., Dash, D., Singhal, A.,
Pandey, A.K., (2020). Role of VapBC12 toxin-antitoxin
14
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

locus in cholesterol-induced mycobacterial persistence. bacterial antitoxin CcdA. J. Mol. Biol., 364, 170–185.
MSystems, 5 https://doi.org/10.1128/mSystems.00855-20. https://doi.org/10.1016/j.jmb.2006.08.082.
16. Fraikin, N., Goormaghtigh, F., Van Melderen, L., (2020). 30. Overgaard, M., Borch, J., Gerdes, K., (2009). RelB and
Type II toxin-antitoxin systems: evolution and revolutions. RelE of Escherichia coli form a tight complex that
J. Bacteriol.,. https://doi.org/10.1128/JB.00763-19. represses transcription via the Ribbon–Helix–Helix Motif
17. Chan, W.T., Balsa, D., Espinosa, M., (2015). One cannot in RelB. J. Mol. Biol., 394, 183. https://doi.org/10.1016/j.
rule them all: Are bacterial toxins-antitoxins druggable?. jmb.2009.09.006.
FEMS Microbiol. Rev., 39, 522–540. https://doi.org/ 31. Donegan, N.P., Marvin, J.S., Cheung, A.L., (2014). Role of
10.1093/femsre/fuv002. adaptor TrfA and ClpPC in controlling levels of SsrA-tagged
18. Tsuchimoto, S., Nishimura, Y., Ohtsubo, E., (1992). The proteins and antitoxins in Staphylococcus aureus. J.
stable maintenance system pem of plasmid R100: Bacteriol., 196, 4140–4151. https://doi.org/10.1128/
degradation of PemI protein may allow PemK protein to JB.02222-14.
inhibit cell growth. J. Bacteriol., 174, 4205–4211. 32. Sala, A., Calderon, V., Bordes, P., Genevaux, P., (2013).
19. Van Melderen, L., Bernard, P., Couturier, M., (1994). Lon- TAC from Mycobacterium tuberculosis: a paradigm for
dependent proteolysis of CcdA is the key control for stress-responsive toxin-antitoxin systems controlled by
activation of CcdB in plasmid-free segregant bacteria. SecB-like chaperones. Cell Stress Chaperones, 18, 129–
Mol. Microbiol., 11, 1151–1157. 135. https://doi.org/10.1007/s12192-012-0396-5.
20. Aizenman, E., Engelberg-Kulka, H., Glaser, G., (1996). An 33. Stewart, G.R., Wermisch, L., Stabler, R., Mangan, J.A.,
Escherichia coli chromosomal “addiction module” regulated Hinds, J., Laing, K.G., Butcher, P.D., Young, D.B., (2002).
by guanosine corrected 3’,5’-bispyrophosphate: a model The heat shock response of Mycobacterium tuberculosis:
for programmed bacterial cell death. Proc. Natl. Acad. Sci. linking gene expression, immunology and pathogenesis.
U.S.A, 93, 6059–6063. Comp. Funct. Genom., 3, 348–351.
21. Diago-Navarro, E., Hernández-Arriaga, A.M., Kubik, S., [34]. Ramage, H.R., Connolly, L.E., Cox, J.S., (2009).
Konieczny, I., Dı́az-Orejas, R., (2013). Cleavage of the Comprehensive functional analysis of mycobacterium
antitoxin of the parD toxin-antitoxin system is determined tuberculosis toxin-antitoxin systems: implications for
by the ClpAP protease and is modulated by the relative pathogenesis, stress responses, and evolution. PLoS
ratio of the toxin and the antitoxin. Plasmid, 70, 78–85. Genet., 5, e1000767. https://doi.org/10.1371/journal.
https://doi.org/10.1016/j.plasmid.2013.01.010. pgen.1000767.
22. Lehnherr, H., Yarmolinsky, M.B., (1995). Addiction protein 35. Betts, J.C., Lukey, P.T., Robb, L.C., Mcadam, R.A.,
Phd of plasmid prophage P1 is a substrate of the ClpXP Duncan, K., (2002). Evaluation of a nutrient starvation
serine protease of Escherichia coli. Proc. Natl. Acad. Sci. model of Mycobacterium tuberculosis persistence by gene
U.S.A, 92, 3274–3277. and protein expression profiling. Mol. Microbiol., 43, 717–
23. Prysak, M.H., Mozdzierz, C.J., Cook, A.M., Zhu, L., Zhang, 731.
Y., Inouye, M., Woychik, N.A., (2009). Bacterial toxin YafQ 36. Keren, I., Shah, D., Spoering, A., Kaldalu, N., Lewis, K.,
is an endoribonuclease that associates with the ribosome (2004). Specialized persister cells and the mechanism of
and blocks translation elongation through sequence- multidrug tolerance in Escherichia coli. J. Bacteriol., 186,
specific and frame-dependent mRNA cleavage. Mol. 8172–8180. https://doi.org/10.1128/JB.186.24.8172-
Microbiol., 71, 1071–1087. https://doi.org/10.1111/j.1365- 8180.2004.
2958.2008.06572.x. 37. Fivian-Hughes, A.S., Davis, E.O., (2010). Analyzing the
24. Garcia-Pino, A., Sterckx, Y., Vandenbussche, G., Loris, R., regulatory role of the HigA antitoxin within Mycobacterium
(2010). Purification and crystallization of Phd, the antitoxin tuberculosis. J. Bacteriol., 192, 4348–4356.
of the phd/doc operon. Acta Crystallograph Sect. F Struct. 38. Schureck, M.A., Dunkle, J.A., Maehigashi, T., Miles, S.J.,
Biol. Cryst. Commun., 66, 167–171. https://doi.org/ Dunham, C.M., (2015). Defining the mRNA recognition
10.1107/S1744309109051550. signature of a bacterial toxin protein. Proc. Natl. Acad. Sci.
25. Hansen, S., Vulić, M., Min, J., Yen, T.-J., Schumacher, M. U.S.A, 112, 13862–13867. https://doi.org/10.1073/
A., Brennan, R.G., Lewis, K., (2012). Regulation of the pnas.1512959112.
Escherichia coli HipBA toxin-antitoxin system by 39. Bordes, P., Sala, A.J., Ayala, S., Texier, P., Slama, N.,
proteolysis. PLoS ONE, 7 https://doi.org/10.1371/journal. Cirinesi, A.-M., Guillet, V., Mourey, L., et al., (2016).
pone.0039185. Chaperone addiction of toxin-antitoxin systems. Nat.
26. Kamada, K., Hanaoka, F., (2005). Conformational change Commun., 7, 13339. https://doi.org/10.1038/
in the catalytic site of the ribonuclease YoeB toxin by YefM ncomms13339.
antitoxin. Mol. Cell., 19, 497–509. https://doi.org/10.1016/ 40. Schuessler, D.L., Cortes, T., Fivian-Hughes, A.S.,
j.molcel.2005.07.004. Lougheed, K.E.A., Harvey, E., Buxton, R.S., Davis, E.O.,
27. Kamada, K., Hanaoka, F., Burley, S.K., (2003). Crystal Young, D.B., (2013). Induced ectopic expression of HigB
structure of the MazE/MazF complex: molecular bases of toxin in Mycobacterium tuberculosis results in growth
antidote-toxin recognition. Mol. Cell., 11, 875–884. inhibition, reduced abundance of a subset of mRNAs and
28. Loris, R., Marianovsky, I., Lah, J., Laeremans, T., cleavage of tmRNA. Mol. Microbiol., 90, 195–207. https://
Engelberg-Kulka, H., Glaser, G., Muyldermans, S., Wyns, doi.org/10.1111/mmi.12358.
L., (2003). Crystal structure of the intrinsically flexible 41. Bordes, P., Cirinesi, A.M., Ummels, R., Sala, A., Sakr, S.,
addiction antidote MazE. J. Biol. Chem., 278, 28252– Bitter, W., Genevaux, P., (2011). SecB-like chaperone
28257. https://doi.org/10.1074/jbc.M302336200. controls a toxin-antitoxin stress-responsive system in
29. Madl, T., Van Melderen, L., Mine, N., Respondek, M., Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. U.S.A,
Oberer, M., Keller, W., Khatai, L., Zangger, K., (2006). 108, 8438–8443. https://doi.org/10.1073/
Structural basis for nucleic acid and toxin recognition of the pnas.1101189108.

15
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

42. Guillet, V., Bordes, P., Bon, C., Marcoux, J., Gervais, V., carboxy-terminal peptide tails added by the SsrA-tagging
Sala, A.J., Dos Reis, S., Slama, N., et al., (2019). Structural system. Genes Dev., 12, 1338–1347.
insights into chaperone addiction of toxin-antitoxin 55. Yamada, Y., Dick, T., (2017). Mycobacterial caseinolytic
systems. Nat. Commun., 10, 782. https://doi.org/10.1038/ protease gene regulator ClgR Is a substrate of caseinolytic
s41467-019-08747-4. protease. MSphere, 2 https://doi.org/10.1128/
43. Sala, A.J., Bordes, P., Ayala, S., Slama, N., Tranier, S., mSphere.00338-16.
Coddeville, M., Cirinesi, A.-M., Castanié-Cornet, M.-P., 56. Ziemski, M., Leodolter, J., Taylor, G., Kerschenmeyer, A.,
et al., (2017). Directed evolution of SecB chaperones Weber-Ban, E., (2020). Genome-wide interaction screen
toward toxin-antitoxin systems. Proc. Natl. Acad. Sci. U.S. for Mycobacterium tuberculosis ClpCP protease reveals
A, 114, 12584–12589. https://doi.org/10.1073/ toxin-antitoxin systems as a major substrate class. FEBS
pnas.1710456114. J.,. https://doi.org/10.1111/febs.15335.
44. Koga, M., Otsuka, Y., Lemire, S., Yonesaki, T., (2011). 57. Barik, S., Sureka, K., Mukherjee, P., Basu, J., Kundu, M.,
Escherichia coli rnlA and rnlB compose a novel toxin- (2010). RseA, the SigE specific anti-sigma factor of
antitoxin system. Genetics, 187, 123–130. https://doi.org/ Mycobacterium tuberculosis, is inactivated by
10.1534/genetics.110.121798. phosphorylation-dependent ClpC1P2 proteolysis. Mol.
45. Wang, X., Kim, Y., Hong, S.H., Ma, Q., Brown, B.L., Pu, M., Microbiol., 75, 592–606. https://doi.org/10.1111/j.1365-
Tarone, A.M., Benedik, M.J., et al., (2011). Antitoxin MqsA 2958.2009.07008.x.
helps mediate the bacterial general stress response. Nat. 58. Lunge, A., Gupta, R., Choudhary, E., Agarwal, N., (2020).
Chem. Biol., 7, 359–366. https://doi.org/10.1038/ The unfoldase ClpC1 of Mycobacterium tuberculosis
nchembio.560. regulates the expression of a distinct subset of proteins
46. Ogura, T., Inoue, K., Tatsuta, T., Suzaki, T., Karata, K., having intrinsically disordered termini. J. Biol. Chem., 295,
Young, K., Su, L.H., Fierke, C.A., et al., (1999). Balanced 9455–9473. https://doi.org/10.1074/jbc.RA120.013456.
biosynthesis of major membrane components through 59. Dougan, D.A., Alver, R., Turgay, K., (2020). Exploring a
regulated degradation of the committed enzyme of lipid A potential Achilles heel of Mycobacterium tuberculosis:
biosynthesis by the AAA protease FtsH (HflB) in defining the ClpC1 interactome. FEBS J.,. https://doi.org/
Escherichia coli. Mol. Microbiol., 31, 833–844. 10.1111/febs.15430.
47. Knipfer, N., Seth, A., Roudiak, S.G., Shrader, T.E., (1999). 60. Schuenemann, V.J., Kralik, S.M., Albrecht, R., Spall, S.K.,
Species variation in ATP-dependent protein degradation: Truscott, K.N., Dougan, D.A., Zeth, K., (2009). Structural
protease profiles differ between mycobacteria and basis of N-end rule substrate recognition in Escherichia coli
protease functions differ between Mycobacterium by the ClpAP adaptor protein ClpS. EMBO Rep., 10, 508–
smegmatis and Escherichia coli. Gene, 231, 95–104. 514. https://doi.org/10.1038/embor.2009.62.
48. Leodolter, J., Warweg, J., Weber-Ban, E., (2015). The 61. Muthuramalingam, M., White, J.C., Bourne, C.R., (2016).
Mycobacterium tuberculosis ClpP1P2 protease interacts Toxin-antitoxin modules are pliable switches activated by
asymmetrically with its ATPase partners ClpX and ClpC1. multiple protease pathways. Toxins, 8 https://doi.org/
PLoS ONE, 10 https://doi.org/10.1371/journal. 10.3390/toxins8070214.
pone.0125345. 62. Becker, S.H., Darwin, K.H., (2017). Bacterial proteasomes:
49. Raju, R.M., Unnikrishnan, M., Rubin, D.H.F., mechanistic and functional insights. Microbiol. Mol. Biol.
Krishnamoorthy, V., Kandror, O., Akopian, T.N., Rev. MMBR, 81 https://doi.org/10.1128/MMBR.00036-16.
Goldberg, A.L., Rubin, E.J., (2012). Mycobacterium 63. Striebel, F., Imkamp, F., Özcelik, D., Weber-Ban, E.,
tuberculosis ClpP1 and ClpP2 function together in protein (1843). Pupylation as a signal for proteasomal
degradation and are required for viability in vitro and during degradation in bacteria. Biochim. Biophys. Acta, 2014,
infection. PLoS Pathog., 8, https://doi.org/10.1371/journal. 103–113. https://doi.org/10.1016/j.bbamcr.2013.03.022.
ppat.1002511 e1002511. 64. Festa, R.A., McAllister, F., Pearce, M.J., Mintseris, J.,
50. Schmitz, K.R., Carney, D.W., Sello, J.K., Sauer, R.T., Burns, K.E., Gygi, S.P., Darwin, K.H., (2010). Prokaryotic
(2014). Crystal structure of Mycobacterium tuberculosis ubiquitin-like protein (Pup) proteome of Mycobacterium
ClpP1P2 suggests a model for peptidase activation by AAA tuberculosis corrected. PloS One, 5, https://doi.org/
+ partner binding and substrate delivery. Proc. Natl. Acad. 10.1371/journal.pone.0008589 e8589.
Sci. U.S.A, 111, E4587–E4595. https://doi.org/10.1073/ 65. Becker, S.H., Li, H., Heran Darwin, K., (2019). Biology and
pnas.1417120111. biochemistry of bacterial proteasomes. Subcell. Biochem.,
51. Nagpal, J., Paxman, J.J., Zammit, J.E., Alhuwaider, A., 93, 339–358. https://doi.org/10.1007/978-3-030-28151-
Truscott, K.N., Heras, B., Dougan, D.A., (2019). Molecular 9_11.
and structural insights into an asymmetric proteolytic complex 66. Delley, C.L., Laederach, J., Ziemski, M., Bolten, M.,
(ClpP1P2) from Mycobacterium smegmatis. Sci. Rep., 9, Boehringer, D., Weber-Ban, E., (2014). Bacterial
18019. https://doi.org/10.1038/s41598-019-53736-8. proteasome activator bpa (rv3780) is a novel ring-shaped
52. Varshavsky, A., (2019). N-degron and C-degron pathways interactor of the mycobacterial proteasome. PloS One, 9,
of protein degradation. Proc. Natl. Acad. Sci. U.S.A, 116, https://doi.org/10.1371/journal.pone.0114348 e114348.
358–366. https://doi.org/10.1073/pnas.1816596116. 67. Jastrab, J.B., Wang, T., Murphy, J.P., Bai, L., Hu, K.,
53. Flynn, J.M., Neher, S.B., Kim, Y.-I., Sauer, R.T., Baker, T. Merkx, R., Huang, J., Chatterjee, C., et al., (2015). An
A., (2003). Proteomic discovery of cellular substrates of the adenosine triphosphate-independent proteasome activator
ClpXP protease reveals five classes of ClpX-recognition contributes to the virulence of Mycobacterium tuberculosis.
signals. Mol. Cell., 11, 671–683. https://doi.org/10.1016/ Proc. Natl. Acad. Sci. U.S.A, 112, E1763–E1772. https://
S1097-2765(03)00060-1. doi.org/10.1073/pnas.1423319112.
54. Gottesman, S., Roche, E., Zhou, Y., Sauer, R.T., (1998). 68. Chi, X., Chang, Y., Li, M., Lin, J., Liu, Y., Li, C., Tang, S.,
The ClpXP and ClpAP proteases degrade proteins with Zhang, J., (2018). Biochemical characterization of mt-

16
P. Texier, P. Bordes, J. Nagpal, et al. Journal of Molecular Biology 433 (2021) 166815

PemIK, a novel toxin-antitoxin system in Mycobacterium 74. Genevaux, P., Keppel, F., Schwager, F., Langendijk-
tuberculosis. FEBS Lett., 592, 4039–4050. https://doi.org/ Genevaux, P.S., Hartl, F.U., Georgopoulos, C., (2004). In
10.1002/1873-3468.13280. vivo analysis of the overlapping functions of DnaK and
69. Burns, K.E., Cerda-Maira, F.A., Wang, T., Li, H., Bishai, W. trigger factor. EMBO Rep., 5, 195–200. https://doi.org/
R., Darwin, K.H., (2010). “Depupylation” of prokaryotic 10.1038/sj.embor.7400067.
ubiquitin-like protein from mycobacterial proteasome 75. Truscott, K.N., Kovermann, P., Geissler, A., Merlin, A.,
substrates. Mol. Cell., 39, 821–827. https://doi.org/ Meijer, M., Driessen, A.J.M., Rassow, J., Pfanner, N.,
10.1016/j.molcel.2010.07.019. Wagner, R., (2001). A presequence- and voltage-sensitive
70. Sakr, S., Cirinesi, A.-M., Ullers, R.S., Schwager, F., channel of the mitochondrial preprotein translocase formed
Georgopoulos, C., Genevaux, P., (2010). Lon protease by Tim23. Nat. Struct. Mol. Biol., 8, 1074–1082. https://doi.
quality control of presecretory proteins in escherichia coli org/10.1038/nsb726.
and its dependence on the SecB and DnaJ (Hsp40) 76. Catanzariti, A.-M., Soboleva, T.A., Jans, D.A., Board, P.G.,
chaperones. J. Biol. Chem., 285, 23506. https://doi.org/ Baker, R.T., (2004). An efficient system for high-level
10.1074/jbc.M110.133058. expression and easy purification of authentic recombinant
71. Ruangprasert, A., Maehigashi, T., Miles, S.J., Dunham, C. proteins. Protein Sci. Publ. Protein Soc., 13, 1331–1339.
M., (2017). Importance of the E. coli DinJ antitoxin carboxy https://doi.org/10.1110/ps.04618904.
terminus for toxin suppression and regulated proteolysis. 77. van Kessel, J.C., Hatfull, G.F., (2007). Recombineering in
Mol. Microbiol.,. https://doi.org/10.1111/mmi.13641. Mycobacterium tuberculosis. Nat. Methods, 4, 147–152.
72. Datsenko, K.A., Wanner, B.L., (2000). One-step https://doi.org/10.1038/nmeth996.
inactivation of chromosomal genes in Escherichia coli K- 78. Vasudevan, D., Rao, S.P.S., Noble, C.G., (2013).
12 using PCR products. Proc. Natl. Acad. Sci. U.S.A, 97, Structural basis of mycobacterial inhibition by cyclomarin
6640–6645. A. J. Biol. Chem., 288, 30883–30891. https://doi.org/
73. Mayer, M.P., (1995). A new set of useful cloning and 10.1074/jbc.M113.493767.
expression vectors derived from pBlueScript. Gene, 163,
41–46.

17

You might also like