Total synthesis of Xanthoangelol B and its various fragments: towards

3 days ago - As an alternative strategy to fight antibiotic resistance, two-component systems (TCSs) have emerged as novel targets. Among TCSs, master...
0 downloads 0 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Total synthesis of Xanthoangelol B and its various fragments: towards inhibition of virulence factor production of Staphylococcus aureus Pushpak Mizar, Rekha Arya, Truc Kim, Soyoung Cha, Kyoung-Seok Ryu, Won-Sik Yeo, Taeok Bae, Dae Wook Kim, Ki Hun Park, Kyeong Kyu Kim, and Seung Seo Lee J. Med. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jmedchem.8b01012 • Publication Date (Web): 02 Nov 2018 Downloaded from http://pubs.acs.org on November 3, 2018

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 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 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.

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 46 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

Journal of Medicinal Chemistry

Total Synthesis of Xanthoangelol B and Its Various Fragments: Towards Inhibition of Virulence Factor Production of Staphylococcus aureus

Pushpak Mizar,† Rekha Arya,‡ Truc Kim, ‡ Soyoung Cha§, Kyoung-Seok Ryu§, Won-sik Yeo,¶ Taeok Bae, ¶ Dae Wook Kim, ┴ Ki Hun Park, ┴ Kyeong Kyu Kim,*‡ and Seung Seo Lee* †

†Chemistry,

Highfield Campus, University of Southampton, Southampton, SO17 1BJ, UK.

‡Department

of Molecular Cell Biology, Institute for Antimicrobial Resistance and

Therapeutics, Samsung Medical Center, Sungkyunkwan University School of Medicine, Suwon 16419, Republic of Korea. §Protein

structure research group, Korea Basic Science Institute, 162 Yeongudanji-Ro,

Ochang-Eup, Cheongju-Si, Chungcheongbuk-Do 28119, Republic of Korea. ¶Department

of Microbiology and Immunology, Indiana University - School of Medicine-

Northwest, Gary, IN 46408 US ┴Division

of Applied Life Science (BK21 Plus), IALS, Gyeongsang National University, Jinju

52828, Republic of Korea 

These authors contributed equally to this work.

* Corresponding authors: [email protected] & [email protected]

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

Abstract As an alternative strategy to fight antibiotic resistance, two-component systems (TCSs) have emerged as novel targets. Among TCSs, master virulence regulators that control the expression of multiple virulence factors are considered as excellent anti-virulence targets. In Staphylococcus aureus, virulence factor expression is tightly regulated by a few master regulators, including the SaeRS TCS. In this study, we used a SaeRS GFP-reporter system to screen natural compound inhibitors of SaeRS, and identified xanthoangelol B 1, a prenylated chalcone from Angelica keiskei as a hit. We have synthesized 1 and its derivative PM-56, and shown that 1 and PM-56 both had excellent inhibitory potency against the SaeRS TCS, as demonstrated by various in vitro and in vivo experiments. As a mode of action, 1 and PM-56 were shown to bind directly to SaeS and inhibit its histidine kinase activity, which suggests a possibility of a broad spectrum inhibitor of histidine kinases.

ACS Paragon Plus Environment

Page 2 of 46

Page 3 of 46 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

Journal of Medicinal Chemistry

INTRODUCTION Antibiotic resistance has arisen as an extremely serious health problem in all parts of the world. New resistance mechanisms are emerging and spreading globally, threatening our ability to treat common infectious diseases. The list of resistant infections – such as pneumonia, tuberculosis, blood poisoning, gonorrhea, and foodborne diseases – is growing longer and sometimes these are impossible to treat as antibiotics become less effective.1,2 It is estimated that, by 2050, worldwide drug-resistant infections will claim 10 million deaths annually. One particularly problematic pathogen is methicillin-resistant Staphylococcus aureus (MRSA). MRSA is a nosocomial and communal menace that is resistant to many antibacterial drugs and antiseptics.3-5 The U.S. Centers for Disease Control and Prevention recently issued a report to various governments and organizations outlining the concern about MRSA. This report referenced the New Drugs for Bad Bugs (ND4BB) project initiated by the European Union, the United States military, and the World Health Organization (WHO).6,7 The Antimicrobial Resistance Global Report on Surveillance released by the WHO in 2014 outlined a global initiative to tackle the MRSA superbug and rekindle a campaign for new, effective, and safe drugs.8-10 Since they suppress bacterial growth, conventional antibiotics are confined to a limited lifespan. It is urgent to establish a new path forward for medicinal chemists to address successfully this problem. Among many new approaches, the inhibition of virulence factors has shown promising outcomes. This strategy bypasses the route of killing pathogens and thereby has the potential to suppress the development of resistance. Many such concepts have been tested.11-13 Among these, two-component systems (TCSs) that control virulence factors have been of particular interest.14-17 Some TCSs control multiple virulence factors in pathogens and are thus called master virulence regulators.18 These TCSs are attractive targets for novel anti-virulence agents. We identified one such TCS, the SaeRS system of S. aureus, as an excellent anti-virulence

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

target that is less likely to evolve resistance.19 We developed a GFP reporter system that can be applied to a high-throughput assay format. Using this assay, we discovered certain FDAapproved drugs that exert anti-virulence effects on S. aureus.20 In continuing such studies, we identified natural compound inhibitors of the SaeRS TCS by screening plant-derived products21 using the same assay. Of the identified natural products that showed promising inhibitory activity, we chose xanthoangelol B for further study including the synthesis of complete xanthoangelol B, fragments and a derivative, to test for anti-virulence activities. Xanthoangelol B is a prenylated chalcone isolated from Ashitaba (Angelica keiskei Koidzumi, Apiaceae). The Japanese herb Ashitaba is a perennial plant found along the Pacific coast of Japan. Ashitaba is enriched with numerous active compounds such as coumarins, flavanones, and chalcones.22-25 Xanthoangelol B is relatively scarce among these compounds, although it is found in all parts of the plant.26 It shows various therapeutic properties, such as anti-oxidative activity, anti-hypertensive activity, anti-inflammatory activity, and anti-viral activity.27 However, although its related chalcones, xanthoangelol and 4-hydroxyderricin were shown to have an antibacterial activity against gram positive bacteria, 28 its anti-bacterial activity has not been established. Thus, given our intriguing screening results, we set out for an in-depth structure-activity study of xanthoangelol B. We were particularly interested in its effect on the virulence of S. aureus and its feasibility as an anti-virulence agent. Xanthoangelol B has a modular structure composed of chalcone and isoprene moieties. It was deemed important to analyze the biological activity of its fragments to establish any structureactivity relationships. Thus, we synthesized fragments of xanthoangelol B along with the parent compound and tested their activities. During the synthesis, we obtained a derivative of xanthoangelol B (compound PM-56) that had a comparable activity, and tested its activities. Here, we report these processes and describe the anti-virulence activities of the compounds that we synthesized.

ACS Paragon Plus Environment

Page 4 of 46

Page 5 of 46 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

Journal of Medicinal Chemistry

RESULTS AND DISCUSSION Chemistry The synthesis of biologically relevant geranylated flavanones and geranylated chalcones (such as (±)-prostratol F, (±)-6-geranyl-5,7-dihydroxy-3’,4’-dimethoxyflavanone, xanthoangelol, 3geranyl-2,3’,4,4’-tetrahydroxychalcone, and (±)-lespeol) has been reported.29-31 With these precedents, we designed our syntheses using a retrosynthetic analysis, as outlined in Scheme 1. Xanthoangelol B 1 can be disconnected to two main fragments: fragment 3 and 4hydroxybenzaldehyde 2. Fragment 3 can be obtained from 2’,4’-dihydroxyacetophenone 4, which is commercially available. There are two ways to synthesize this fragment 3 (Scheme 1). A synthetic route to 1 was also designed to generate various fragments that could be evaluated for their biological activity. The fragments shown in Scheme 1 can provide insight into the structure-activity relationship of 1 that may help to curtail or enhance its activity. O

OH

HO

1

OH

OH

O

OH

OH

OHC

+

OH

4

O

OH

OH

OH

2

+

3

Br

5 OH O

+

O

OH

7

8

Scheme 1. Retrosynthetic analysis of xanthoangelol B 1

ACS Paragon Plus Environment

O

Journal of Medicinal Chemistry 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

Our synthetic design involved the limited use of protecting groups (Scheme 2). Our initial approach was to introduce 2,6-dimethylhepta-1,6-dien-3-ol at the end of the synthesis to reduce complications arising from the chiral aliphatic hydroxyl group. The 4’-hydroxyl position of 2′,4′-dihydroxyacetophenone was alkylated with allyl bromide that was heated at 220 oC in N, N-diethylaniline for 24 h. This allowed for Claisen rearrangement, giving a 3,3-sigmatropicshifted major product 10 with a 76% yield. The minor product was 5-allyl-2,4-dihydroxy acetophenone. The regioselectivity of Claisen rearrangement giving a 3-allylated major product is well documented.32 10 was then subjected to aldol condensation by refluxing for 18 h in KF-alumina (40%), a solid-support catalyst in toluene. 11 was obtained as the major product, with a yield of approximately 60%. This yield was slightly lower than that of most aldol condensation reactions owing to the two 2’,4’-OH groups. A reaction longer than 18 hours resulted in the formation of a non-polar mixture of spots observed by TLC analysis. The nature of the solvent also drastically influenced the yield of the reaction. 2,6-Dimethylhepta1,6-dien-3-ol was introduced later by initially synthesizing 2,2-dimethyl-3-(3-methylbut-3-en1-yl)oxirane 8 from citral by subjecting it to decarbonylation via heating at 130 oC in the presence of catalytic amounts of palladium acetate (10 mol%) in a sealed tube for 24 h. This produced 2,6-dimethylhepta-1,5-diene 12, with a yield of ca 62%.33 Dimethylhepta-1,5-diene was subjected to selective oxidation using m-chloroperbenzoic acid (mCPBA) to obtain an epoxide 8 with a 45% yield.34 8 was then subjected to cross-olefin metathesis with 11, using a Hoveyda-Grubbs second-generation catalyst to yield 70% of an E-product 13.35 The epoxide ring was subsequently subjected to selective ring opening reactions or isomerization using in situ-generated diethylaluminum 2,2,6,6-tetramethyl-piperidide to yield the desired product xanthoangelol B 1 at a 69% yield.36

ACS Paragon Plus Environment

Page 6 of 46

Page 7 of 46 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

Journal of Medicinal Chemistry

NEt2

O

Br

OH

O

(1.1 equiv.)

K2CO3 (1.5 equiv.) acetone, 14h, reflux

OH

O

OH

76% 9

220 oC, Heat, 24h

O

OH

70% 10 O

+

10

2

OH

OH

KF-Alumina (40%) (3 equiv.) toluene, reflux, 18h

HO

OH

60% 11

Pd(OAc)2 (8 mol%) CHO

11

+

4Ao Cyclohexane 130 oC, sealed tube

8

mCPBA (80% in ether) CHCl3, ehter

62% 12

O

Hoveyda-Grubbs 2nd Gen. (0.05 equiv.) DCM, 40 oC

45% 8

O

OH

HO

OH

O

70% 13

13

DATMP, 0 oC Toluene

1 69%

Scheme 2. Synthesis of xanthoangelol B 1

There were a few interesting observations during the synthesis of 1 (Scheme S1). As an alternative route to xanthoangelol B, the introduction of 2,6-dimethylhepta-1,6-dien-3-ol to the 3’-position of 2,4-dihydroxyacetophenone resulted in decreased reactivity toward condensation reactions. Use of a solid-support reagent (such as KF-alumina) or use of highly basic conditions yielded no product or a complex mixture of products. Heating the reaction in the presence of a high boiling-point solvent (either in highly basic or acidic conditions) yielded no desired product. Finally, the use of L-proline as an organic catalyst37-38 resulted in an aldol

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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 46

addition product 15. However, this product did not react further (to elimination) to yield 1 in any of the acidic or basic conditions attempted. Thus, 15 was subjected to selective epoxide isomerization, and the resulting compound PM-56 was tested for anti-virulence activities as a derivative of 1. HO CHO

O

HO

O

60% KOH/Ethanol +

OH

reflux, 24h

OH OH

OH 80% 17 + mCPBA Br

17 + 18

CHCl3, 0 oC-rt

KOH, reflux, methanol

Br

DATMP, 0 oC

13 34 %

O

18

1

toluene, 2h

80% Ph HN O

nBu

TCCA, CHCl3, -40 oC enationmerically pure 1 34 %

Scheme 3. Alternative route for the synthesis of xanthoangelol B 1. No metal catalysts involved.

Synthesized xanthoangelol B 1 was structurally identical to the natural product, except for the chiral center. Selective epoxidation and subsequent epoxide ring opening resulted in an enantiomeric mixture. We speculated that the number of required steps to synthesize 1 could be reduced by using a longer chain (i.e., geranyl bromide) and carrying out subsequent selective epoxidations, which could be reacted further to yield 1. This would provide us with an

ACS Paragon Plus Environment

Page 9 of 46 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

Journal of Medicinal Chemistry

additional fragment for structure-activity tests and allow us to incorporate the metal-free chemistry. Thus, we devised a shorter synthetic route (without using metal catalysts) by minimizing the use of protecting groups. Specifically, 2,4-dihydroxy acetophenone was subjected to aldol condensation to give an intermediate that was reacted with 1,2-epoxy geranyl bromide. This generated 13, which underwent selective epoxy-isomerization to give 1 as an enantiomeric mixture (Scheme 3). Then, we attempted an optical resolution of synthesized compound 1. We chose a method of enantioselective organocatalyst oxidative kinetic resolution using the 7-Bn-3-n-Bu-4-oxa-5-azahomoadamantane/trichloroisocyanuric acid (TCCA) system to yield a resolved 1.39 The 1H and 13C NMR spectra of the resolved compound 1 were in agreement with spectra previously reported.21 The mass spectrum showed a peak at 409.19 corresponding to [M+H]+, consistent with the calculated mass of the desired product. In a direct comparison using HPLC, the synthetic and resolved 1 was shown to elute at exactly the same retention time as the natural product (Figure S1). The optical rotation for kinetically resolved 1 was D= +13.5 (MeOH, c=0.5). This is consistent with the data reported for naturally-extracted 1,23 which indicates the authenticity of the kinetically resolved synthetic xanthoangelol B 1.

Biological evaluation of xanthoangelol B, its fragments, and its derivative Previously, we constructed a SaeRS GFP reporter strain based on the S. aureus USA300 isolate, and found promising leads among FDA-approved drugs through a high throughput screening.20 The same SaeRS GFP reporter strain was constructed for screening a natural product library in this study. Briefly, as the α-hemolysin promoter (Phla) is a well-characterized target of the
SaeRS TCS, we generated a minimal Phla, termed Phlam, which contains only the SaeR binding sites and the -10/-35 promoter sequences of the hemolysin gene (hla). Phlam was fused to a green fluorescence protein (gfp) gene instead of hla in the single-copy integration plasmid

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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 46

pCL55. The resulting plasmid (pCL-Phlam-gfp) was inserted into S. aureus USA300. Hence, if the expression of GFP in a sample was lower than the control sample, it meant that the SaeRS TCS was inhibited. As the complete inhibition of Phlam requires a complete shutdown of the SaeRS TCS (such as a deletion of saeS), Phlam will detect only potent inhibitors of the SaeRS reporter. Xanthoangelol B 1 was one of hits among the natural products screened with this method. Thus, we assayed 1, its fragments, and a derivative (PM-56). The structures of all the assayed compounds and fragments are presented in Scheme 4 and Scheme S2.

Structure-activity studies of xanthoangelol B and its inhibitory effect on the SaeRS TCS On initial screening, 1 and PM-56 showed an excellent inhibitory activity against SaeRS. These compounds repressed GFP expression while allowing bacterial growth, making them ideal antivirulence candidates. In contrast, all of the fragments (Scheme S2) failed to inhibit the SaeRS TCS, even at 20 µM. This indicates that the whole structure is needed for the inhibitory activity. Furthermore, xanthoangelol and xanthoangelol F (Scheme 4) did not show inhibitory activity at 20 µM. O

HO

OH Xanthoangelol B 1

O

HO

OH

OH O OH

HO

OH PM-56

O

OH

HO

OH Xanthoangelol

OH OH

OH

OCH3 Xanthoangelol F

Scheme 4. Derivatives of xanthoangelol B

These two analogs are also natural prenylated chalcones from Ashitaba, but they lack a hydroxyl group in the terminal isoprene unit. This hydroxyl group appears to be crucial for the

ACS Paragon Plus Environment

Page 11 of 46 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

Journal of Medicinal Chemistry

anti-virulence activity of this class of compounds. Xanthoangelol with an identical structure to that of 1, except for this hydroxyl group, did not have any activity (although methoxylation in the central phenol hydroxyl group of xanthoangelol F may still have an impact). IC50 values against the SaeRS GFP reporter were measured to be 2.1 µM for 1 and 4.3 µM for PM-56. In these experiments (at up to 8 and 16 µM), 1 and PM-56 did not disrupt bacterial growth (Figure 1A and C). To elaborate the inhibitory effect of both compounds on the growth of S. aureus, we measured the growth of S. aureus in the presence of 1 and PM-56 at higher concentrations up to 20 h. From 16 M, 1 started showing growth inhibition, and at 64 M of 1, the growth of S. aureus was severely inhibited (Figure S2). On the other hand, up to 64 M of PM-56, S. aureus still showed comparable growth to the control (Figure S2). The trends shown at 8 h did not change up to 20 hours. This indicates that the rigidity conferred to the chalcone structure by the conjugated double bond is not essential for anti-SaeRS activity, but may lead to higher toxicity (which affects bacterial growth). Together, these data strongly indicate that the isoprene and chalcone moieties are necessary in their entirety for the inhibitory activity.

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

A

B

C

D

Figure 1. Various concentrations of compounds were used to treat the USA300-Phlam GFP strain, and GFP expression was analyzed at 8 h. A, Measurement of bacterial growth in the presence of xanthoangelol B 1; B, measurement of fluorescence in the presence of xanthoangelol B 1; C, measurement of bacterial growth in the presence of PM-56; D, measurement of fluorescence in the presence of PM-56. All the experiments were performed in triplicates and significance was compared with control (no treatment) using one way ANOVA.

Effects of inhibitors on downstream virulence factors Staphylococcal α-hemolysin, which causes lysis of erythrocytes, is the target virulence factor of the SaeRS TCS. Therefore, if the SaeRS TCS is inhibited, the expression of α-hemolysin

ACS Paragon Plus Environment

Page 12 of 46

Page 13 of 46 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

Journal of Medicinal Chemistry

should decrease to a certain degree. When 1 and PM-56 were incubated with erythrocytes in the presence of methicillin-resistant S. aureus USA300, notable protection from hemolysis was observed (Figure 2 and Figure S3). At only 4 µM, both compounds showed visible protection. At 8 µM, a comparable level of protection to that of saeS knockout mutants (SaeS:Tn551) was observed. The hemolysis activity of this saeS mutant was reduced to 82.6% in comparison to the wild type S. aureus. To confirm the relevance of the protection effect observed with 1 and PM-56 to the SaeRS pathway, we also tested 1 and PM-56 against the saeS knockout mutant. This experiment conducted at 8 M of 1 and 15 M of PM-56 showed the reduction in hemolysis by 86.9% and 87.4%, respectively, compared to hemolysis by the wild type (Figure S4). Although apparent further inhibition appeared to exist, the differences were not statistically significant, which thus suggests that the protection against hemolysis by 1 and PM-56 is highly likely attributed to inhibition of the SaeRS pathway. The IC50 values of 1 and PM-56 for the inhibition of erythrocyte hemolysis were then measured to be 4.6 and 4.0 µM, respectively. These values were consistent with the IC50 values of both compounds for the SaeRS TCS GFP reporter. Therefore, these strongly suggests that the protection of erythrocytes from hemolysis in the presence of MRSA can be attributed to inhibition of the SaeRS TCS.

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

A

B

Figure 2. Percentage hemolysis values were determined after treating samples with compounds 1 and PM-56 after an 8 h incubation. A, 1: various concentrations were used to treat human RBCs, and percentage hemolysis values were calculated relative to the control sample; B, PM56: various concentrations were used to treat human RBCs, and percentage hemolysis values were calculated relative to the control sample. All the experiments were performed in triplicates and the data calculated by one way ANOVA. CT: positive control, SaeS:Tn551: saeS knockout.

ACS Paragon Plus Environment

Page 14 of 46

Page 15 of 46 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

Journal of Medicinal Chemistry

A

B

C

D

Figure 3. Effects of 1 and PM-56 on the virulence gene expression were analyzed. A, hla; B, aureolysin (aur); C, ϒ-hemolysin; D, staphylokinase. 16S rRNA was used as a reference control gene. All the experiments were performed in triplicates and the data compared with no treatment control and calculated by one way ANOVA.

We next analyzed the transcription of virulence genes in the presence of 1 and PM-56 (Figure 3). We used concentrations of 8 µM (1) and 10 µM (PM-56), corresponding to an IC90 value for the SaeRS GFP reporter. We observed significant transcriptional suppression of the following genes: α-hemolysin (hla), aureolysin (aur), γ-hemolysin, and staphylokinase. These four genes have a direct SaeR binding site in their promoter regions and are tightly regulated by the SaeRS TCS.40 While it has been reported that hla is regulated by an accessory gene regulator (agr) as well as by sarA, a direct connection between these regulatory systems and

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

the expression of hla remains under debate.41 Their suppression likely reflects the inhibitory effects of both compounds on the SaeRS TCS. Then, we analyzed the transcription of a gene encoding Staphylococcal accessory regulator (SarA). SarA is responsible for the biofilm formation and not regulated by saeS. Thus, it was considered a good negative control, and its transcription analyzed post treatment of xanthoangelol B 1 and PM-56 (Figure S5). As anticipated, it can be seen that the level of sarA transcription was not significantly affected by the two compounds. These results imply that 1 and PM-56 indeed inhibited the SaeRS pathway with specificity, and thus are promising candidates for future development of anti-virulence agents or chemical probes specific for the SaeRS two-component system.

Cytotoxicity In order to explore the potential of xanthoangelol B 1 and PM-56 as initial leads for further translation, we have examined the cytotoxicity of both compounds on Hela cells, quantitatively and qualitatively. Firstly, LDH release assay was performed, in which we have treated Hela cells with 1 and PM-56 at various concentrations (2 M - 128 M) for 72 h and measured the optical density at 450 nm. It was shown that IC50 values for Hela cell survivals were 19.70 M and 28.94 M with 1 and PM-56, respectively (Figure 4 A & B). These values were 10 fold and 7 fold higher than IC50 values for the GFP-reporter inhibition by 1 and PM-56, respectively. Hence, the toxicity/efficacy ratio is reasonable for the initial leads. Next, the live/dead imaging experiment has been performed. The Hela cells were seeded in 96 well plates for 24 h, and after attachment of cells, the culture media (DMEM, 10% FBS, 1% penicillin/streptomycin) were changed, followed by treatment with 1 and PM-56 at various concentrations (from 2 μM to 64 μM) for 72 h. The Hela cells stained with fluorescein diacetate dye (live cell staining, green color) and propidium iodide dye (dead cell staining, red color) were observed using the confocal microscopy (Figure 4 C & D).

ACS Paragon Plus Environment

Page 16 of 46

Page 17 of 46 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

Journal of Medicinal Chemistry

A

B

C

D

Figure 4. Cytotoxicity of xanthoangelol B and PM-56 on Hela cells. A (xanthoangelol B), B (PM-56): LDH release assay. Hela cells were seeded in 5x105 cfu and treated with xanthoangelol B 1 and PM-56 from 2 μM to 128 μM and incubated for 72 h. The LDH release was analyzed at OD450nm. All the experiments were performed in three biological replicates and data are presented as a mean ± SEM. C (xanthoangelol B), D (PM-56), live/dead cell staining. Hela cells were seeded 5x105 cfu in a 96 well plate and incubated 24 h for attachment. The various concentrations of 1 and PM-56 (2 μM to 64 μM) were tested for 72 h. Then, the cells were stained with FDA and PI and imaged under the confocal microscope. Green, live; red,

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

dead, (i) vehicle control; (ii) 2 μM; (iii) 4 μM; (iv) 8 μM; (v) 16 μM; (vi) 32 μM; (vii) 64 μM. Images are based on n=3 independent experiment.

Confocal images were in accordance with the quantitative results as mentioned earlier. Both compounds at their IC90 (approximately 8 M) for the GFP-reporter did not produce substantially visible toxicity to the Hela cells while a reduction in viability was observed only when the cells were treated with concentrations higher than IC50 for cytotoxicity. Cytotoxicity tests demonstrated that 1 and PM-56 indeed had the promising traits as initial leads.

Validation of the anti-virulence activity of compounds 1 and PM56 using a worm infection model With the encouraging in vitro data, we further validated the anti-virulence activity of 1 and PM-56 in a Galleria mellonella infection model. G. mellonella worms possess both a humoral and an innate immune system. The worms provide an attractive, simple, and easy-to-handle infection model with few biosafety or ethical issues. The G. mellonella infection model has previously been used with several human pathogens to test the in vivo efficacy of antibacterial and anti-virulence drugs.42 Thus, G. mellonella larvae was challenged with 5x106 colony forming units (CFUs), a load designed to allow only 20% survival after 72 hours. The larvae was post-treated with various doses of either 1 or PM-56 at 12 h intervals, up to 72 h. Larvae health indices were strictly followed and recorded. A dose-dependent protection from virulence by both compounds (1 and PM-56) was observed in the larvae (Figure 5). Compared to the vehicle control, 1 and PM-56 generated a 3.5-fold higher survival at doses as low as 2.0 and 2.1 mg/kg, respectively. 1 showed 90% survival at a dose of 3.1 mg/kg and complete survival at 4.1 mg/kg. PM-56 showed 90% survival at 3.2 mg/kg and complete survival at 4.3 mg/kg.

ACS Paragon Plus Environment

Page 18 of 46

Page 19 of 46

These survival rates were comparable to or slightly better than survival rates observed in worms infected with saeS knockout strains.

(a) A.

Xanthoangelol B Concentration (mg/kg body weight) *** **

100

Percent survival

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

Journal of Medicinal Chemistry

80

* 60

Vehicle Control SaeS Mutant 1.021 2.042 3.063 4.085

40 20

ns

0 0

12

24

36

48

60

72

Time (h)

B.

Figure 5. The ability of compounds 1 (A) and PM-56 (B) to protect G. mellonella from lethal S. aureus infection. Kaplan-Meier survival curves for G. mellonella challenged with 5x106 CFU S. aureus and treated with compounds 1 and PM-56 at 2 h post-infection. The compounds

ACS Paragon Plus Environment

Journal of Medicinal Chemistry 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

were injected every 12 h for 72 h. The experiments were performed using biological replicate samples (n=10). ***P