Polypharmacology Approaches against the Pseudomonas aeruginosa

Apr 5, 2017 - Polypharmacology Approaches against the Pseudomonas aeruginosa MvfR Regulon and Their Application in Blocking Virulence and Antibiotic T...
0 downloads 14 Views 1MB Size
Subscriber access provided by Karolinska Institutet, University Library

Article

Polypharmacology approaches against the Pseudomonas aeruginosa MvfR regulon and their application in blocking virulence and antibiotic tolerance Damien Maura, Steffen L. Drees, Arunava Bandyopadhaya, Tomoe Kitao, Michele Negri, Melissa Starkey, Biliana Lesic, Sylvain Milot, Eric Deziel, Robert Zahler, Mike Pucci, Antonio Felici, Susanne Fetzner, Francois Lepine, and Laurence G. Rahme ACS Chem. Biol., Just Accepted Manuscript • DOI: 10.1021/acschembio.6b01139 • Publication Date (Web): 05 Apr 2017 Downloaded from http://pubs.acs.org on April 6, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Chemical Biology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

161x229mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2 178x243mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

Figure 3 103x142mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4 98x65mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

Figure 5 130x167mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 27

1

Polypharmacology approaches against the Pseudomonas aeruginosa MvfR

2

regulon and their application in blocking virulence and antibiotic tolerance

3 4

Damien Maura a,b,c, Steffen L. Drees d, Arunava Bandyopadhaya a,b,c, Tomoe Kitao a,b,c, Michele Negri e,

5

Melissa Starkey a,b,c, Biliana Lesic a,b,c,, Sylvain Milot f, Eric Déziel f, Robert Zahler g, Mike Pucci g,

6

Antonio Felici e, Susanne Fetzner d, François Lépine f, and Laurence G. Rahme a,b,c

*

7 8

a

Department of Surgery, b Department of Microbiology and Immunobiology, Harvard Medical School

9

and Massachusetts General Hospital,

Shriners Hospitals for Children Boston, Boston, MA, USA,

d

10

Institute for Molecular Microbiology and Biotechnology, University of Münster, Münster, Germany,

e

11

Aptuit, Verona, Italy,

12

Cambridge, MA, USA.

f

c

INRS Institut Armand Frappier, Laval, QC, Canada,

g

Spero Therapeutics,

13 14

*

Corresponding author: Laurence G. Rahme, Ph.D.

15

E-Mail: [email protected]

16

Tel. +1 (617) 724-5003; Fax +1 (617) 724-8558

17 18

Short title: Dual inhibitors target MvfR and PqsBC

1 ACS Paragon Plus Environment

Page 7 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

19

Abstract

20

Pseudomonas aeruginosa is an important nosocomial pathogen that is frequently recalcitrant to

21

available antibiotics, underlining the urgent need for alternative therapeutic options against this

22

pathogen. Targeting virulence functions is a promising alternative strategy as it is expected to generate

23

less selective resistance to treatment compared to antibiotics. Capitalizing on our non-ligand based

24

benzamide – benzimidazole (BB) core structure compounds reported to efficiently block the activity of

25

the P. aeruginosa multiple virulence factor regulator MvfR, here we report the first class of inhibitors

26

shown to interfere with PqsBC enzyme activity, responsible for the synthesis of the MvfR activating

27

ligands HHQ and PQS, and the first to target simultaneously MvfR and PqsBC activity. The use of these

28

compounds reveals that inhibiting PqsBC is sufficient to block P. aeruginosa’s acute virulence functions,

29

as the synthesis of MvfR ligands is inhibited. Our results show that MvfR remains the best target of this

30

QS pathway, as we show that antagonists of this target block both acute and persistence related

31

functions. The structural properties of the compounds reported in this study provide several insights

32

that are instrumental for the design of improved MvfR regulon inhibitors against both acute and

33

persistent P. aeruginosa infections. Moreover, the data presented offer the possibility of a

34

polypharmacology approach of simultaneous silencing two targets in the same pathway. Such a

35

combined anti-virulence strategy holds promise in increasing therapeutic efficacy, and providing

36

alternatives in the event of a single target’s resistance development.

37 38

Keywords: Anti-virulence, Pseudomonas aeruginosa, Quorum Sensing Inhibitors, MvfR, PqsBC, Dual

39

Inhibitors, Polypharmacology, Antibiotic Tolerance, Acute Infections

2 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 27

40

Pseudomonas aeruginosa is an opportunistic Gram-negative bacterial pathogen responsible for more

41

than 50,000 infections in the US each year, primarily causing nosocomial infections in

42

immunocompromised and cystic fibrosis patients (1). P. aeruginosa rapid expansion of resistance to

43

almost all available antibiotics urges the development of alternative therapeutic strategies (2). One

44

promising strategy, especially in the case of multi-drug resistant or pan antibiotic resistant P. aeruginosa

45

clinical strains, is anti-virulence drugs that target bacterial virulence systems or master virulence

46

regulators (3). One such master virulence regulator is the quorum sensing (QS) cell-to-cell bacterial

47

communication system.

48

P. aeruginosa possesses three major QS systems: LasR (4), RhlR (5, 6) and MvfR (7-10). The MvfR QS

49

system is a promising anti-virulence target due to its critical role in inducing the expression of multiple

50

P. aeruginosa virulence systems that promote both acute and chronic infections (7, 11-14). Moreover, as

51

opposed to LasR no clinical isolates with frequent mutations in MvfR were reported to date (15). The

52

transcriptional regulator MvfR (also known as PqsR) binds to 37 loci and regulates the expression of the

53

associated genes (16) including the pqsABCDE operon, whose encoded proteins catalyze the

54

biosynthesis of MvfR inducers and of ~60 distinct low-molecular-weight compounds (7-9, 17, 18), part of

55

which are the 4-hydroxyl-2-alkyl-quinolines (HAQs)

56

aminoacetophenone (2-AA) (12, 14, 20). This multi-step biosynthetic pathway is summarized in Figure

57

1a. The first step is the conversion of the HAQs precursor anthranilic acid by PqsA and PqsD into 2-

58

aminobenzoylacetyl-CoA (2-ABA-CoA) (21, 22), which then either spontaneously cyclizes into 2,4-

59

dihydroxyquinoline (DHQ) (22-24) or is hydrolyzed by the thioesterase PqsE (or TesB) into 2-

60

aminobenzoylacetate (2-ABA) (23). 2-ABA and octanoyl-CoA are then condensed by the PqsBC enzyme

61

into the MvfR activating ligand 4-hydroxy-2-heptyl-quinoline (HHQ) (24, 25), which is later hydroxylated

62

into the second MvfR activating ligand 3,4-dihydroxy-2-heptyl-quinoline (PQS) by PqsH (26). 2-ABA can

63

also decompose into DHQ (24) or undergo decarboxylation to 2-AA (24) (Fig 1a).

(19)

and the

non-HAQ molecule 2-

3 ACS Paragon Plus Environment

Page 9 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

64

Importantly, HHQ and PQS are critical for MvfR activity in vitro, however, only HHQ appears to be

65

essential for acute infection in vivo, as the absence of PQS assessed by using a pqsH mutant causes WT

66

mortality in mice (9). On the contrary, 2-AA silences the acute infection branch of the MvfR QS system

67

by binding and inhibiting the activity of PqsBC (13, 25) and promotes antibiotic tolerance as well as

68

chronic/persistent infections by interfering with the bacterial translation apparatus and modulating

69

epigenetically the host immune system to promote host tolerance to infection respectively (12-14, 27).

70

The role of the MvfR QS pathway in both acute and chronic infections has motivated several drug

71

discovery studies that generated inhibitors of PqsA (28, 29), PqsD (30, 31) and MvfR (32-35). However,

72

no synthetic PqsBC inhibitor has been identified to date. We previously reported the identification and

73

development of a new family of molecules with a Benzamide-Benzimidazole (BB) core structure as

74

highly cell-permeable inhibitors of the MvfR QS system (32). One of our most potent inhibitors, M64,

75

was shown to target MvfR and inhibit HAQs and 2-AA synthesis with an IC50 in the nanomolar range (32).

76

The present work describes our effort to obtain further mechanistic knowledge on the BB family

77

inhibitors’ capability in inhibiting MvfR circuitry. This work provides novel insights that are critical in the

78

design of improved MvfR regulon inhibitors and reports the identification of the unprecedented first

79

class of dual inhibitors of MvfR and PqsBC activities.

80 81

Results and Discussion

82

HAQs and 2-AA inhibition profiling reveals MvfR regulon inhibitors with dual targets.

83

We previously reported an initial target assessment of few of our most potent BB compounds and

84

demonstrated the physical interaction of the most potent inhibitor, M64, with MvfR (32). However, the

85

profile of a BB compound, M51, was ambiguous which could suggest that this chemical family might

86

target other proteins in the MvfR QS system besides the transcriptional regulator MvfR itself. In order to

4 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 27

87

address this point, we performed a systematic target assessment for representative BB compounds in

88

our collection.

89

First, we assessed the ability of each compound to interfere with the activity of MvfR, PqsA, PqsBC or

90

PqsD by quantifying the production levels of the MvfR-regulated molecules HHQ, PQS, DHQ and 2-AA.

91

We used an mvfR isogenic mutant strain that constitutively expresses the pqsABCDE genes (mvfR-

92

pPqsABCD) (32) and thus has MvfR-independent HHQ, PQS, 2-AA and DHQ production. In this strain,

93

inhibitors targeting PqsA or PqsD enzymes result in decreased production of all four MvfR-regulated

94

molecules, while PqsBC inhibition causes decreased production of HHQ and PQS level and an

95

accumulation of 2-AA or DHQ; in the mvfR- pPqsABCD strain MvfR inhibition has no impact on the

96

production of any of these molecules. Interestingly, our inhibitors’ collection contained compounds

97

exerting at least two types of inhibitory patterns in the mvfR- pPqsABCD strain (Fig. 1b). The first group

98

of molecules (green) exhibits the same inhibition pattern previously described with the MvfR inhibitor

99

M64 (32) that is, no inhibition of HHQ, PQS, DHQ, and 2-AA. These results suggest that inhibitors M50,

100

M62, M34, M61, M53 and M57 do not target the PqsA, PqsBC or PqsD enzymes pointing to MvfR as the

101

potential target. In contrast, however, the second group of compounds (blue), although inhibiting HHQ

102

and PQS, led to a considerable accumulation of 2-AA and DHQ (Fig. 1b), clearly suggesting that they are

103

targeting the PqsBC enzyme.

104 105

The ability of each compound to interfere with the activity of either MvfR or PqsBC was further

106

interrogated in the parental strain PA14 where the expression of the pqs operon is MvfR dependent (Fig.

107

1c). In this strain, inhibition of MvfR results in a reduced production of HHQ, PQS, 2-AA and DHQ

108

whereas PqsBC inhibition only decreases HHQ and PQS production but not 2-AA or DHQ. Figure 1c

109

shows the 2-AA, DHQ, PQS and HHQ levels produced in the WT strain PA14 in presence of each

110

compound. We observed two different inhibition profiles: Compounds of the first profile (red) – M50,

5 ACS Paragon Plus Environment

Page 11 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

111

M62, M34, M61, M53, M57, M59, M58, M51, B1 and M52 – exhibit the inhibition pattern we previously

112

observed with the MvfR inhibitor M64 (32) that is, inhibition of HHQ, PQS, 2-AA and DHQ in this wild

113

type PA14 strain, suggesting they are targeting MvfR. Interestingly, however, five of these compounds –

114

M59, M58, M51, B1 and M52 – were identified as PqsBC inhibitors in Figure 1b, indicating that they may

115

act as dual inhibitory compounds that also target MvfR in addition the PqsBC enzymatic activity.

116

Compounds from the second profile (orange), identified as PqsBC inhibitors in Figure 1b – M27, M26,

117

M23 M4, M8 and M55 – partially inhibit the synthesis of 2-AA and DHQ in addition to fully blocking the

118

synthesis of HHQ and PQS (Fig. 1c). M22 also blocks 2-AA and DHQ production but requires higher

119

concentrations to do so (Fig. S5). These data suggest that compounds M27, M26, M23 M4, M8 and M55

120

and M22 may also interfere with MvfR activity in addition to that of PqsBC. However, their lower

121

efficacy at reducing 2-AA and DHQ production compared to M59, M58, M51, B1 and M52 suggests that,

122

overall, they are less potent MvfR inhibitors (Fig. 1c). While compounds M23, M4 and M8 reduce 2-AA

123

production, they induce the production of DHQ in the wild type PA14 strain (Fig. 1c). Although the

124

reason for this effect is not clear, changes in the bacterial culture pH and inactivation of PqsE have been

125

shown to affect the ratio between 2-AA and DHQ (23, 24).

126 127

Taken together, these data suggest that the BB compounds can be classified in three categories: 1) MvfR

128

inhibitors (M64, M50, M62, M34, M61, M53 and M57); 2) MvfR – PqsBC dual inhibitors with high anti-

129

MvfR and high anti-PqsBC activity (M59, M58, M51, B1 and M52); and 3) MvfR – PqsBC dual inhibitors

130

with low anti-MvfR and high anti-PqsBC activity (M27, M26, M23, M4, M8, M55).

131 132

We report here a combination of two HAQs quantification assays in live cells that provide valuable

133

insights on compound target(s). The mvfR isogenic mutant strain constitutively expressing the pqs

134

operon (mvfR- pPqsABCD) is particularly useful in discriminating between MvfR and PqsA/D inhibitors,

6 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 27

135

and can also help identify PqsBC inhibitors that were overshadowed by the MvfR/PqsA/PqsD inhibition

136

phenotype dominant in the PA14 wild type strain. Accordingly, it is possible that some MvfR, PqsA or

137

PqsD inhibitors reported in previous studies such as (28, 30, 33) might also have other targets within the

138

MvfR QS pathway. We believe that such assays would benefit the community in discriminating the

139

target(s) of MvfR QS system inhibitors and allow a better understanding of the determinants driving the

140

interaction of compounds with various targets in this pathway.

141 142

Target validation

143

To validate the inhibitors’ targets, we selected representative compounds from each category and

144

assessed their ability to bind to MvfR using surface plasmon resonance (SPR). As expected, all tested

145

inhibitors bind to MvfR (Fig. 2a, c and S2). Compounds with a low anti-MvfR activity (M27, M26 and

146

M23) bind 9 to 111 times less efficiently to MvfR than those with high anti-MvfR activity (M64, M50,

147

M62, M59, M51) (Fig. 2a). This lower binding is nonetheless significant since the KD of those inhibitors is

148

in the same order of magnitude as that of HHQ and PQS, two well established MvfR native ligands (10).

149

We then assessed the ability of one compound in each category to interfere directly with PqsBC

150

enzymatic activity by quantifying the in vitro conversion of 2-ABA into HHQ using purified PqsBC protein.

151

As expected, the MvfR – PqsBC dual inhibitors M59 and M27 block the ability of PqsBC to convert 2-ABA

152

into HHQ, with an EC50 of 13.4 µM and 12.5 µM respectively (Fig. 2c, d). Moreover, Figure 2e,f show that

153

M59 binds to PqsBC and displaces 2-AA, a natural PqsBC inhibitor structurally unrelated to BB

154

compounds acting competitively with the physiological substrate 2-ABA, confirming further M59 ability

155

to interfere with PqsBC. These data indicate that the compounds reported here are significantly more

156

potent PqsBC inhibitors than 2-AA whose EC50 for PqsBC inhibition was previously found to be 46 µM

157

(25). Moreover, the anti-PqsBC activity of 2-AA in live cells is weak as its IC50 for HHQ inhibition is around

158

400 µM (13). Interestingly, Figure 2c shows that the MvfR inhibitor M64 also interferes with PqsBC

7 ACS Paragon Plus Environment

Page 13 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

159

activity although the inhibition is weaker (EC50 ~185µM) compared to that of M59 and M27. This

160

suggests that other inhibitors from the first category (M50, M62, M34, M61, M53 and M57) may also be

161

MvfR – PqsBC dual inhibitors with a weak anti-PqsBC activity. To exclude any direct reactions of

162

inhibitors with the PqsBC substrates, we analyzed inhibitor-substrate mixtures by UV spectroscopy and

163

assessed 2-ABA and octanoyl-CoA stability by HPLC. Data verified the absence of reactivity and

164

confirmed the stability of both, 2-ABA and octanoyl-CoA in the presence of any of the inhibitors (Fig. S1

165

and Table S1). Overall, these physical interaction and enzymatic activity data confirm the existence of

166

MvfR – PqsBC dual inhibitors.

167

One puzzling question is how compounds with the same core structure can bind to such different

168

targets, one being a transcriptional regulator and the other a biosynthetic enzyme. It is worth noting

169

that the common point between both proteins is HHQ. Indeed, PqsBC catalyzes HHQ production while

170

MvfR binds to HHQ. Therefore, it is possible that compounds with the BB core structure share some

171

physical properties with HHQ allowing them to bind HHQ related proteins/targets.

172

Polypharmacology, involving a single or multiple drugs acting on the same pathway has been shown to

173

have a significant impact on the treatment efficacy for various diseases (36, 37), including bacterial

174

infections (38, 39). Although there seems to be no synergistic benefit with this series of inhibitory

175

compounds, the possibility of blocking two different targets in the same pathway offers an advantage in

176

the event of a single target’s resistance development (39, 40). For example, if a resistance mutation

177

occurs in PqsBC, a dual inhibitor would retain both anti-virulence and anti-persistence potency via its

178

anti-MvfR activity. In the case of a resistance mutation in MvfR, dual inhibitors would still retain an anti-

179

virulence potency thanks to their anti-PqsBC activity although they might not be able to retain the anti-

180

persistence activity. The current view in the anti-virulence field is that resistance is still expected to

181

occur, but at much lower frequencies than that promoted by traditional antibiotics. By definition

182

virulence contributes to pathogen fitness in vivo implying that an anti-virulence resistant mutant would

8 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 27

183

theoretically outcompete sensitive cells during treatment. However, some important microbial

184

population genetics and ecological concepts such as public/private use of virulence factors, fitness

185

localization and population structure suggest that selective pressure to anti-virulence may only be

186

applied in specific settings, as opposed to that of antibiotics which apply selective pressure in all settings

187

(3, 41, 42). It is worth pointing that thus far no MvfR mutations have been reported in P. aeruginosa

188

clinical isolates sequenced suggesting that a functional MvfR QS system is critical for P. aeruginosa

189

infections. However, one cannot exclude this from occurring once MvfR QS system inhibitors are used in

190

the context of long term treatment regiments (i.e. CF patients). Therefore a polypharmacology anti-

191

virulence approach could present a significant advantage for the treatment of P. aeruginosa infections.

192 193

Inhibition of P. aeruginosa virulence and antibiotic tolerance

194

To assess the potential of our compounds to reduce P. aeruginosa virulence, we evaluated the ability of

195

selected inhibitors from each category to block P. aeruginosa acute virulence against A549 human lung

196

epithelial cells and RAW264.7 macrophages in vitro using cell viability as a readout. Cell survival was

197

quantified 3 hours post-infection with PA14 in the presence or absence of each inhibitor. Infection with

198

P. aeruginosa resulted in 79.2% lung epithelial cell death (Fig. 3a) and 74.4% macrophage cell death (Fig.

199

3b). In contrast, treatment with inhibitors from all three categories increased 3.1 to 3.8 times lung

200

epithelial cell survival and 1.9 to 2.6 times macrophage survival to PA14 infection (Fig. 3a-b). These

201

compounds show no cytotoxic effect (Fig. S4) and importantly reduce P. aeruginosa virulence in an MvfR

202

QS system dependent manner, as they do not significantly change the lung cell survival rate when added

203

to the mvfR mutant cells (Fig. S3), indicating no off-target effects (Fig. S4). Moreover, no significant

204

difference in the survival of lung epithelial cells or murine macrophages was observed between

205

compounds that exhibit low (M27, M26, M23) and high anti-MvfR activity (M64, M50, M59, M58, M51)

206

(Fig. 1c). 9 ACS Paragon Plus Environment

Page 15 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

207

Next, we evaluated the potential of our compounds to inhibit antibiotic tolerance by assessing bacterial

208

survival to the β-lactam antibiotic Meropenem. Data presented in Figure 4 indicate that the dual

209

inhibitors with low anti-MvfR activity (M27, M26, and M23) do not significantly reduce tolerance to

210

Meropenem. However, the dual inhibitors that exhibit high anti-MvfR activity and consequently block

211

2AA production (M64, M50, M59, M58 and M51) reduce antibiotic tolerance by more than 70%

212

compared to vehicle control (Fig. 4).

213

Overall, these data indicate that all three categories of inhibitors have a similar therapeutic potential in

214

the context of acute P. aeruginosa infections likely because they all efficiently block HHQ and PQS

215

production. However, dual inhibitors with a high anti-MvfR activity are more potent at blocking

216

antibiotic tolerance than those with a low anti-MvfR activity since they restrict 2-AA production more

217

efficiently. Therefore, MvfR – PqsBC dual inhibitors with a high anti-MvfR activity have an increased

218

therapeutic potential against pro-acute infection related molecules HHQ and PQS, as well as the pro-

219

persistent and immunomodulatory molecule 2-AA.

220 221

Role of inhibitors structure in target recognition

222

Finally, we assessed whether some compound structural determinants could be associated with

223

selective target recognition. Figure 5 shows the chemical structure for every inhibitor of the three

224

categories. Notably, almost all the compounds with a high anti-MvfR activity (columns 1 and 2) contain a

225

nitro group at the position 5 of the benzimidazole ring. In contrast, this nitro group is lacking from all

226

compounds with a low anti-MvfR activity (column 3). This suggests that the nitro group may play a

227

critical role for the interaction with MvfR. This is most obvious when comparing the binding and

228

inhibitory activity of M26 versus M51, two identical compounds but for the nitro group which is lacking

229

in M26 (Fig. 5). Indeed, the addition of the nitro group on M26 increases 25 times the compound

230

binding to MvfR (Fig. 2a) and 14.8 times the inhibition of 2-AA production in PA14 (Fig. 1c). Moreover,

10 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 27

231

replacing the methyl group at the position 5 of the benzimidazole ring of M27 by a nitro group as in M50

232

(Fig. 5) increases 9.5 times the binding to MvfR (Fig. 2a) and 14.3 times the inhibition of 2-AA production

233

in PA14 (Fig. 1c). Similarly, the addition of the nitro group on M55, as in M50 (Fig. 5), increases 15.4

234

times the inhibition of 2-AA production in PA14 (Fig. 1c). Notably, HHQ ligand based analogues

235

harboring strong electron withdrawing groups such as NO2, CN or CF3 on the benzene moiety of the

236

quinolone structure were also found to be strong MvfR inhibitors (35, 43), supporting our data on the

237

importance of the nitro group for BB compounds to interact with MvfR. Future in depth protein –

238

inhibitor interaction studies will provide more insights on this aspect. Interestingly, the presence of the

239

nitro group also appears to decrease the compounds anti-PqsBC activity. Indeed, adding the nitro group

240

to B1 as in M62, or to M55 as in M50, or replacing the methyl group of M27 by a nitro group as in M50

241

reduce 4.6, 8.9 or 2.3 times respectively HHQ production in the mvfR mutant strain constitutively

242

expressing the pqs operon (Fig. 1b). Overall, these data demonstrate that the nitro group at the position

243

5 of the benzimidazole ring is critical for the compounds to selectively recognize MvfR over PqsBC, and

244

suggest modifications on the benzamide moiety to further modulate target selectivity.

245 246

Conclusions

247

This study provides exciting new insights into the mode of action and therapeutic potential of

248

compounds with a BB core structure in the context of quorum sensing inhibition in P. aeruginosa. We

249

previously reported a series of non-ligand based BB compounds interfering with the P. aeruginosa MvfR

250

QS system (32). In this study, we assessed further the mode of action of this BB compound series in this

251

pathogen. While they all inhibit the MvfR QS system, we discovered that the transcriptional regulator

252

MvfR is not their only target in the MvfR pathway. Our analysis shows that several of our BB compounds

253

also interfere with the PqsBC enzyme, inhibiting its ability to convert 2-ABA into the MvfR activating

254

ligand HHQ. These compounds represent the first class of inhibitors ever reported to interfere with

11 ACS Paragon Plus Environment

Page 17 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

255

PqsBC, and are the first synthetic inhibitors ever shown to interfere with both MvfR and PqsBC. These

256

dual inhibitors harbor an exciting therapeutic potential due to their ability to block both acute and

257

chronic virulence related functions in P. aeruginosa, and to simultaneously inhibit more than one target.

258

As such they offer new avenues to overcome the probability of resistance that might be presented with

259

single target inhibitor in this pathogen.

12 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 27

260

Methods

261

Bacterial strains, plasmids and growing conditions

262

PA14 (UCBPP-PA14) is a P. aeruginosa human clinical isolate (44). The strain mvfR-pPqsABCD which has

263

constitutive and MvfR-independent pqs operon expression was previously described in (32). Unless

264

noted otherwise, all bacterial strains were grown in 5mL LB Lenox medium (Fisher Scientific) at 37˚C

265

under 200rpm orbital shaking using glass tubes (VWR). 75µg/mL Tetracycline was added when growing

266

the mvfR- pPqsABCD strain to maintain the pDN18 plasmid.

267 268

HAQs and 2-AA quantification

269

HAQs and 2-AA levels were quantified in bacterial culture supernatants by LC/MS as described in (9, 45).

270 271

Binding to MvfR via surface plasmon resonance

272

Purification of the MvfR ligand binding domain (MvfRc87) was performed as described in (32). MvfRc87

273

protein (50µg/mL) was diluted in 10mM Sodium Acetate buffer (pH5.5) and immobilized on a CM7

274

Series S Sensor Chip using an Amine Coupling reagent kit (GE Healthcare) at the level of 3,000-5,000

275

Response Units (RU). PBS (pH 7.4) containing 0.05% P20 surfactant was used as the running buffer

276

during protein immobilization.

277

The interactions between test compounds and MvfRc87 ligand binding domain were analyzed by

278

Biacore™ T200 evaluation software 2.0 (GE Healthcare). 10mM HEPES (pH 7.4) containing 150 mM NaCl,

279

3mM EDTA, 0.05% P20 surfactant and 4% DMSO was the running buffer. Injections were performed at a

280

flow rate of 30µL/min with 60s of contact time and 420s (Single-cycle kinetics) or 180s (Multi-cycle

281

kinetics) of dissociation time. Each injection was followed by an extra wash with 50% DMSO. Solvent

282

correction was performed according to sensorgram analysis. The zero-concentration curve was

13 ACS Paragon Plus Environment

Page 19 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

283

subtracted from the other sensorgrams. The affinities of compounds were determined with the “Steady

284

State Affinity” yielding the Binding Affinity Constant (KD) and the maximum binding capacity (Rmax)

285

expressed as Response Units (RU).

286 287

Inhibition of PqsBC enzymatic activity

288

PqsBC was purified and catalytic activity was determined as described previously (25). 2-ABA was

289

synthesized according to (24, 25). Octanoyl-CoA and 2-AA were purchased from Sigma-Aldrich (St. Louis,

290

MO, USA). For evaluating enzyme inhibition, 20-100 nM PqsBC was incubated with various

291

concentrations of the respective inhibitor for 5 minutes before measuring the residual activity. 2-ABA

292

and octanoyl-CoA concentrations were 150-200 µM and 20-40 µM, respectively, and the DMSO

293

concentration was 1 %. All assays were conducted in triplicate.

294 295

Binding to PqsBC via fluorescence spectroscopy

296

Dissociation constants of PqsBC-inhibitor complexes were determined by fluorescence spectroscopy or

297

fluorescence polarization spectrometry, using the fluorescence properties of the respective inhibitor

298

molecule. 2-AA displacement from PqsBC by inhibitors was monitored by the 2-AA fluorescence

299

intensity change resulting from binding of the molecule to PqsBC (25). All assays were conducted in a

300

buffer containing 50 mM HEPES, pH 8.0, 50 mM NaCl and 1 % DMSO, using a Jasco FP-6500 fluorescence

301

spectrometer with polarization accessory.

302 303

Cell viability assays

304

Cells survival to PA14 infection was assessed as previously described in (32). Briefly, bacterial cells were

305

grown until mid-exponential phase (OD600nm = 2) in the presence or absence of each inhibitor at 50 µM,

306

then washed and added to host cells at a MOI of 100. Three hours post-infection, bacterial cells were 14 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 27

307

killed with 500 µg/mL Gentamycin and washed away twice with PBS. Cells were incubated in 100 µg/mL

308

MTT for 16 hours at 37 °C in 5% CO2, then MTT was dissolved in DMSO and OD570nm was measured. All

309

cells were maintained in 5% CO2 at 37 °C. A549 (human lung epithelial cell line, ATCC, USA) and

310

RAW264.7 cells (mouse macrophage cell line, IMGENEX, USA) were maintained in F12K and DMEM

311

medium (Life Technologies, USA) respectively. The media were supplemented with 10% heat-inactivated

312

FBS, penicillin/streptomycin, 2 mM L-glutamine, and 10 mM HEPES (all from Gibco). The cells were

313

seeded in T-75 tissue culture flasks (Falcon, USA) and used between passages 2 and 3.

314 315

Antibiotic tolerance

316

P. aeruginosa cells were grown at 37˚C 200rpm in 10g/L TSB media until mid-exponential phase (OD600nm

317

2) then exposed to 10ug/mL Meropenem (Sandoz, USA) for 24 hour under the same incubating

318

conditions. Before (t=0) and after (t=24h) Meropenem addition, a 200µL sample of each culture was

319

collected, diluted and plated on LB agar plates to quantify the total number of bacteria (t=0) and the

320

surviving bacteria (t=24h). Colony forming units (CFUs) were counted after 24h incubation at 37˚C. The

321

fraction of antibiotic tolerant cells was then calculated as the ratio of the amount of total bacteria (t=0)

322

divided by the amount of surviving bacteria (t=24h). Data are expressed as the percentage of antibiotic

323

tolerant cells relative to the DMSO vehicle control, which represents a survival fraction of 2.3x10-6 cells.

324 325

Statistical analyses

326

Statistical significance was assessed using unpaired One Way ANOVA + Dunnett’s post-test or One Way

327

ANOVA + Tukey post-test as indicated using GraphPad Prism software.

328 329

Author contributions 15 ACS Paragon Plus Environment

Page 21 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

330

DM, SLD, AB, TK, MN, MS, BL, RZ, MP, AF, SF, FL and LGR designed experiments. DM, SLD, AB, TK, MN,

331

SM and ED performed experiments. DM and LGR wrote the manuscript and prepared figures. All authors

332

reviewed the manuscript.

333 334

Acknowledgements

335

This work was supported by Shriners Hospital Postdoctoral Fellowship #84206 to DM and by the

336

research grants, Shriners #8770, Cystic Fibrosis Foundation #11P0, NIAID R33AI105902 to L.G.R and

337

grant FE 383/23-2 from the Deutsche Forschungsgemeinschaft to S.F. Funding sources had no role in

338

study design, data analysis and interpretation or decision to publish.

339 340

Competing financial interest

341

LGR is the scientific founder and scientific advisory board member of Spero Therapeutics LLC. MP is

342

Executive Director, Early Drug Discovery at Spero Therapeutics. RZ is an independent consultant. AF is

343

Director and Head of Microbiology Unit at Aptuit (Verona). MN is Research Scientist, Microbiology

344

Department at Aptuit (Verona). LGR, SF, FL and corresponding lab members received no funding from

345

Spero Therapeutics.

346 347 348

16 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

349

References

350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395

1. 2.

3. 4.

5. 6. 7.

8.

9.

10.

11.

12.

13.

14.

15.

Page 22 of 27

CDC (2013) Antibiotic resistance threats in the United States, 2013, Centres for Disease Control and Prevention, US Department of Health and Human Services. Lister, P. D., Wolter, D. J., and Hanson, N. D. (2009) Antibacterial-resistant Pseudomonas aeruginosa: clinical impact and complex regulation of chromosomally encoded resistance mechanisms, Clin Microbiol Rev 22, 582-610. Maura, D., Ballok, A. E., and Rahme, L. G. (2016) Considerations and caveats in anti-virulence drug development, Curr Opin Microbiol 33, 41-46. Gambello, M. J., and Iglewski, B. H. (1991) Cloning and characterization of the Pseudomonas aeruginosa lasR gene, a transcriptional activator of elastase expression, J Bacteriol 173, 30003009. Schuster, M., and Greenberg, E. P. (2006) A network of networks: quorum-sensing gene regulation in Pseudomonas aeruginosa, Int J Med Microbiol 296, 73-81. Venturi, V. (2006) Regulation of quorum sensing in Pseudomonas, FEMS Microbiol Rev 30, 274291. Cao, H., Krishnan, G., Goumnerov, B., Tsongalis, J., Tompkins, R., and Rahme, L. G. (2001) A quorum sensing-associated virulence gene of Pseudomonas aeruginosa encodes a LysR-like transcription regulator with a unique self-regulatory mechanism, Proceedings of the National Academy of Sciences 98, 14613. Deziel, E., Lepine, F., Milot, S., He, J., Mindrinos, M. N., Tompkins, R. G., and Rahme, L. G. (2004) Analysis of Pseudomonas aeruginosa 4-hydroxy-2-alkylquinolines (HAQs) reveals a role for 4hydroxy-2-heptylquinoline in cell-to-cell communication, Proc Natl Acad Sci U S A 101, 13391344. Xiao, G., Deziel, E., He, J., Lepine, F., Lesic, B., Castonguay, M. H., Milot, S., Tampakaki, A. P., Stachel, S. E., and Rahme, L. G. (2006) MvfR, a key Pseudomonas aeruginosa pathogenicity LTTRclass regulatory protein, has dual ligands, Mol Microbiol 62, 1689-1699. Williams, P., and Camara, M. (2009) Quorum sensing and environmental adaptation in Pseudomonas aeruginosa: a tale of regulatory networks and multifunctional signal molecules, Curr Opin Microbiol 12, 182-191. Deziel, E., Gopalan, S., Tampakaki, A. P., Lepine, F., Padfield, K. E., Saucier, M., Xiao, G., and Rahme, L. G. (2005) The contribution of MvfR to Pseudomonas aeruginosa pathogenesis and quorum sensing circuitry regulation: multiple quorum sensing-regulated genes are modulated without affecting lasRI, rhlRI or the production of N-acyl-L-homoserine lactones, Mol Microbiol 55, 998-1014. Bandyopadhaya, A., Kesarwani, M., Que, Y. A., He, J., Padfield, K., Tompkins, R., and Rahme, L. G. (2012) The quorum sensing volatile molecule 2-amino acetophenon modulates host immune responses in a manner that promotes life with unwanted guests, PLoS Pathog 8, e1003024. Kesarwani, M., Hazan, R., He, J., Que, Y., Apidianakis, Y., Lesic, B., Xiao, G., Dekimpe, V., Milot, S., Deziel, E., Lépine, F., and Rahme, L. G. (2011) A Quorum Sensing Regulated Small Volatile Molecule Reduces Acute Virulence and Promotes Chronic Infection Phenotypes, PLoS Pathogens 7, e1002192. Que, Y., Hazan, R., Ryan, C. M., Milot, S., Lépine, F., Lydon, M., and Rahme, L. G. (2011) Production of Pseudomonas aeruginosa Intercellular Small Signaling Molecules in Human Burn Wounds, Journal of Pathogens 2011, 1-5. D'Argenio, D. A., Wu, M., Hoffman, L. R., Kulasekara, H. D., Deziel, E., Smith, E. E., Nguyen, H., Ernst, R. K., Larson Freeman, T. J., Spencer, D. H., Brittnacher, M., Hayden, H. S., Selgrade, S., Klausen, M., Goodlett, D. R., Burns, J. L., Ramsey, B. W., and Miller, S. I. (2007) Growth 17 ACS Paragon Plus Environment

Page 23 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443

ACS Chemical Biology

16.

17.

18.

19.

20.

21.

22.

23. 24.

25.

26.

27.

28.

29.

30.

phenotypes of Pseudomonas aeruginosa lasR mutants adapted to the airways of cystic fibrosis patients, Mol Microbiol 64, 512-533. Maura, D., Hazan, R., Kitao, T., Ballok, A. E., and Rahme, L. G. (2016) Evidence for Direct Control of Virulence and Defense Gene Circuits by the Pseudomonas aeruginosa Quorum Sensing Regulator, MvfR, Sci Rep 6, 34083. Xiao, G., He, J., and Rahme, L. G. (2006) Mutation analysis of the Pseudomonas aeruginosa mvfR and pqsABCDE gene promoters demonstrates complex quorum-sensing circuitry., Microbiology 152, 1679-1686. Lepine, F., Dekimpe, V., Lesic, B., Milot, S., Lesimple, A., Mamer, O. A., Rahme, L. G., and Deziel, E. (2007) PqsA is required for the biosynthesis of 2,4-dihydroxyquinoline (DHQ), a newly identified metabolite produced by Pseudomonas aeruginosa and Burkholderia thailandensis, Biol Chem 388, 839-845. Lepine, F., Milot, S., Deziel, E., He, J., and Rahme, L. G. (2004) Electrospray/mass spectrometric identification and analysis of 4-hydroxy-2-alkylquinolines (HAQs) produced by Pseudomonas aeruginosa, J Am Soc Mass Spectrom 15, 862-869. Scott-Thomas, A. J., Syhre, M., Pattemore, P. K., Epton, M., Laing, R., Pearson, J., and Chambers, S. T. (2010) 2-Aminoacetophenone as a potential breath biomarker for Pseudomonas aeruginosa in the cystic fibrosis lung, BMC Pulm Med 10, 56. Coleman, J. P., Hudson, L. L., McKnight, S. L., Farrow, J. M., 3rd, Calfee, M. W., Lindsey, C. A., and Pesci, E. C. (2008) Pseudomonas aeruginosa PqsA is an anthranilate-coenzyme A ligase, J Bacteriol 190, 1247-1255. Zhang, Y. M., Frank, M. W., Zhu, K., Mayasundari, A., and Rock, C. O. (2008) PqsD is responsible for the synthesis of 2,4-dihydroxyquinoline, an extracellular metabolite produced by Pseudomonas aeruginosa, J Biol Chem 283, 28788-28794. Drees, S. L., and Fetzner, S. (2015) PqsE of Pseudomonas aeruginosa Acts as Pathway-Specific Thioesterase in the Biosynthesis of Alkylquinolone Signaling Molecules, Chem Biol 22, 611-618. Dulcey, C. E., Dekimpe, V., Fauvelle, D. A., Milot, S., Groleau, M. C., Doucet, N., Rahme, L. G., Lepine, F., and Deziel, E. (2013) The end of an old hypothesis: the pseudomonas signaling molecules 4-hydroxy-2-alkylquinolines derive from fatty acids, not 3-ketofatty acids, Chem Biol 20, 1481-1491. Drees, S. L., Li, C., Prasetya, F., Saleem, M., Dreveny, I., Williams, P., Hennecke, U., Emsley, J., and Fetzner, S. (2016) PqsBC, a Condensing Enzyme in the Biosynthesis of the Pseudomonas aeruginosa Quinolone Signal: CRYSTAL STRUCTURE, INHIBITION, AND REACTION MECHANISM, J Biol Chem 291, 6610-6624. Schertzer, J. W., Brown, S. A., and Whiteley, M. (2010) Oxygen levels rapidly modulate Pseudomonas aeruginosa social behaviours via substrate limitation of PqsH, Mol Microbiol 77, 1527-1538. Bandyopadhaya, A., Tsurumi, A., Maura, D., Jeffrey, K. L., and Rahme, L. G. (2016) A quorumsensing signal promotes host tolerance training through HDAC1-mediated epigenetic reprogramming, Nat Microbiol 1, 16174. Lesic, B., Lepine, F., Deziel, E., Zhang, J., Zhang, Q., Padfield, K., Castonguay, M. H., Milot, S., Stachel, S., Tzika, A. A., Tompkins, R. G., and Rahme, L. G. (2007) Inhibitors of pathogen intercellular signals as selective anti-infective compounds, PLoS Pathog 3, 1229-1239. Calfee, M. W., Coleman, J. P., and Pesci, E. C. (2001) Interference with Pseudomonas quinolone signal synthesis inhibits virulence factor expression by Pseudomonas aeruginosa, Proc Natl Acad Sci U S A 98, 11633-11637. Storz, M. P., Maurer, C. K., Zimmer, C., Wagner, N., Brengel, C., de Jong, J. C., Lucas, S., Musken, M., Haussler, S., Steinbach, A., and Hartmann, R. W. (2012) Validation of PqsD as an anti-biofilm 18 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490

31. 32.

33.

34.

35.

36.

37. 38.

39. 40.

41.

42. 43.

44.

Page 24 of 27

target in Pseudomonas aeruginosa by development of small-molecule inhibitors, J Am Chem Soc 134, 16143-16146. Allegretta, G., Weidel, E., Empting, M., and Hartmann, R. W. (2015) Catechol-based substrates of chalcone synthase as a scaffold for novel inhibitors of PqsD, Eur J Med Chem 90, 351-359. Starkey, M., Lepine, F., Maura, D., Bandyopadhaya, A., Lesic, B., He, J., Kitao, T., Righi, V., Milot, S., Tzika, A., and Rahme, L. (2014) Identification of anti-virulence compounds that disrupt quorum-sensing regulated acute and persistent pathogenicity, PLoS Pathog 10, e1004321. Ilangovan, A., Fletcher, M., Rampioni, G., Pustelny, C., Rumbaugh, K., Heeb, S., Camara, M., Truman, A., Chhabra, S. R., Emsley, J., and Williams, P. (2013) Structural Basis for Native Agonist and Synthetic Inhibitor Recognition by the Pseudomonas aeruginosa Quorum Sensing Regulator PqsR (MvfR) PLoS Pathog 9, : e1003508. doi:1003510.1001371/journal.ppat.1003508. Lu, C., Maurer, C. K., Kirsch, B., Steinbach, A., and Hartmann, R. W. (2014) Overcoming the unexpected functional inversion of a PqsR antagonist in Pseudomonas aeruginosa: an in vivo potent antivirulence agent targeting pqs quorum sensing, Angew Chem Int Ed Engl 53, 11091112. Lu, C., Kirsch, B., Maurer, C. K., de Jong, J. C., Braunshausen, A., Steinbach, A., and Hartmann, R. W. (2014) Optimization of anti-virulence PqsR antagonists regarding aqueous solubility and biological properties resulting in new insights in structure-activity relationships, Eur J Med Chem 79, 173-183. Tolaney, S. M., Barry, W. T., Dang, C. T., Yardley, D. A., Moy, B., Marcom, P. K., Albain, K. S., Rugo, H. S., Ellis, M., Shapira, I., Wolff, A. C., Carey, L. A., Overmoyer, B. A., Partridge, A. H., Guo, H., Hudis, C. A., Krop, I. E., Burstein, H. J., and Winer, E. P. (2015) Adjuvant paclitaxel and trastuzumab for node-negative, HER2-positive breast cancer, N Engl J Med 372, 134-141. Gandhi, M., and Gandhi, R. T. (2014) Single-pill combination regimens for treatment of HIV-1 infection, N Engl J Med 371, 248-259. Diacon, A. H., Pym, A., Grobusch, M., Patientia, R., Rustomjee, R., Page-Shipp, L., Pistorius, C., Krause, R., Bogoshi, M., Churchyard, G., Venter, A., Allen, J., Palomino, J. C., De Marez, T., van Heeswijk, R. P., Lounis, N., Meyvisch, P., Verbeeck, J., Parys, W., de Beule, K., Andries, K., and Mc Neeley, D. F. (2009) The diarylquinoline TMC207 for multidrug-resistant tuberculosis, N Engl J Med 360, 2397-2405. Worthington, R. J., and Melander, C. (2013) Combination approaches to combat multidrugresistant bacteria, Trends Biotechnol 31, 177-184. Thomann, A., de Mello Martins, A. G., Brengel, C., Empting, M., and Hartmann, R. W. (2016) Application of Dual Inhibition Concept within Looped Autoregulatory Systems toward Antivirulence Agents against Pseudomonas aeruginosa Infections, ACS Chem Biol 11, 1279-1286. Gerdt, J. P., and Blackwell, H. E. (2014) Competition studies confirm two major barriers that can preclude the spread of resistance to quorum-sensing inhibitors in bacteria, ACS Chem Biol 9, 2291-2299. Allen, R. C., Popat, R., Diggle, S. P., and Brown, S. P. (2014) Targeting virulence: can we make evolution-proof drugs?, Nat Rev Microbiol 12, 300-308. Lu, C., Kirsch, B., Zimmer, C., de Jong, J. C., Henn, C., Maurer, C. K., Musken, M., Haussler, S., Steinbach, A., and Hartmann, R. W. (2012) Discovery of antagonists of PqsR, a key player in 2alkyl-4-quinolone-dependent quorum sensing in Pseudomonas aeruginosa, Chem Biol 19, 381390. Rahme, L. G., Stevens, E. J., Wolfort, S. F., Shao, J., Tompkins, R. G., and Ausubel, F. M. (1995) Common virulence factors for bacterial pathogenicity in plants and animals, Science 268, 18991902.

19 ACS Paragon Plus Environment

Page 25 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

491 492 493

ACS Chemical Biology

45.

Lepine, F., Deziel, E., Milot, S., and Rahme, L. G. (2003) A stable isotope dilution assay for the quantification of the Pseudomonas quinolone signal in Pseudomonas aeruginosa cultures, Biochim Biophys Acta 1622, 36-41.

494 495

Figure legends

496 497

Figure 1: HAQs and 2-AA biosynthesis pathway and inhibition in live P. aeruginosa cells

498

a) HAQs and 2-AA biosynthesis pathway. PqsABCDE are encoded by the pqs operon. PqsH and TesB

499

encoding genes are located elsewhere in P. aeruginosa chromosome. TesB is another thioesterase also

500

able to convert 2-ABA-CoA into 2-ABA. b, c) HHQ, PQS, 2-AA and DHQ levels measured by LC/MS in an

501

mvfR mutant strain constitutively expressing the pqs operon (b) or in PA14 wild type strain (c) in the

502

presence or absence of various BB inhibitors at 100µM. Levels are normalized to that of the DMSO

503

vehicle control. Results show the average ± SD of at least two independent replicates.

504 505

Figure 2: Target validation

506

Compounds binding intensity to MvfR was measured via SPR (a, b). Binding to MvfR was assessed with a

507

wide range of inhibitors concentrations. Compounds interference with PqsBC was measured via enzyme

508

kinetics by assessing the inhibitory activity on the condensation of 2-ABA and octanoyl-CoA to HHQ by

509

PqsBC (c, d). 2-ABA conversion into HHQ was assessed with a wide range of inhibitors concentrations.

510

Shown is the average ± SEM of three independent replicates. e. Binding of M59 assayed with

511

fluorescence polarization spectrometry. The autofluorescence of M59 was used to probe binding of the

512

molecule to PqsBC (gray fit) and octanoyl-PqsBC (black fit), each of which were titrated stepwise to the

513

inhibitor solution. The calculated dissociation constants were 2.5µM and 2.9µM, respectively, indicating

514

that M59 binds to both forms of the enzyme. f. The displacement of 2-AA, a PqsBC inhibitor competitive

515

with 2-ABA, by M59 was analyzed by measuring the 2-AA fluorescence intensity change in response to 20 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 27

516

PqsBC binding. In an experiment where 2-AA was titrated into a solution containing 1 µM PqsBC, the KD

517

was 6.8µM (upper fit). When 10µM M59 was present in the protein solution, the apparent KD of the

518

PqsBC-2-AA complex increased (lower fit), indicating that M59 interferes with 2-AA binding.

519 520

Figure 3: Anti-virulence efficacy in lung epithelial cells and macrophage infection assays

521

Survival of A549 human lung epithelial cells (a) or RAW264.7 macrophage cells (b) to PA14 infection in

522

presence of 50 µM of dual inhibitors with high anti-MvfR activity and low anti-PqsBC activity (green),

523

dual inhibitors with high anti-MvfR activity and high anti-PqsBC activity (red), dual inhibitors with low

524

anti-MvfR activity and high anti-PqsBC activity (orange) or the DMSO vehicle control (black). Results

525

show the average ± SEM of at least 3 independent replicates. Statistical significance to the DMSO control

526

was assessed using one way ANOVA + Dunnett’s post-test. No statistical difference was observed when

527

comparing the inhibitors with each other (p>0.05, One Way ANOVA + Tukey post-test).

528 529

Figure 4: Inhibition of antibiotic tolerance

530

Tolerance to 10µg/mL of the β-lactam antibiotic Meropenem in presence of 10µM of dual inhibitors

531

with high anti-MvfR activity and low anti-PqsBC activity (green), dual inhibitors with high anti-MvfR

532

activity and high anti-PqsBC activity (red), dual inhibitors with low anti-MvfR activity and high anti-PqsBC

533

activity (orange) or the DMSO vehicle control (black). Results show the average ± SEM of at least 3

534

independent replicates. Statistical significance to the DMSO control was assessed using one way ANOVA

535

+ Dunnett’s post-test.

536 537

Figure 5: Dual inhibitors structures sorted based on their potency on each target

538 21 ACS Paragon Plus Environment

Page 27 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

539

Supporting information

540

Further supporting data on compounds targets, cytotoxicity and off-target assessments.

22 ACS Paragon Plus Environment