• 검색 결과가 없습니다.

Systematic editing of synthetic RIG-I ligands to produce effective antiviral and anti-tumor RNA immunotherapies

N/A
N/A
Protected

Academic year: 2021

Share "Systematic editing of synthetic RIG-I ligands to produce effective antiviral and anti-tumor RNA immunotherapies"

Copied!
13
0
0

로드 중.... (전체 텍스트 보기)

전체 글

(1)

Systematic editing of synthetic RIG-I ligands to

produce effective antiviral and anti-tumor RNA

immunotherapies

Janghyun Lee

1,

, Eun-Byeol Park

2,

, Jiyoun Min

2,

, Si-Eun Sung

1,

, Yejin Jang

3

, Jin

Soo Shin

3

, Dongmin Chun

2

, Ki-Hun Kim

1

, Jihyun Hwang

1

, Mi-Kyung Lee

1

, Yun Young Go

3

,

Dohyeong Kwon

2

, Meehyein Kim

3,*

, Suk-Jo Kang

2,*

and Byong-Seok Choi

1,*

1Department of Chemistry, Korea Advanced Institute of Science and Technology (KAIST), 291 Daehak-ro,

Yuseong-gu, Daejeon 34141, South Korea,2Department of Biological Sciences, Korea Advanced Institute of Science and Technology (KAIST), 291 Daehak-ro, Yuseong-gu, Daejeon 34141, South Korea and3Center for Virus Research

and Testing, Korea Research Institute of Chemical Technology, 141 Gajeong-ro, Yuseong-gu, Daejeon 34114, South Korea

Received August 11, 2017; Revised December 22, 2017; Editorial Decision January 15, 2018; Accepted January 17, 2018

ABSTRACT

Retinoic acid-inducible gene I (RIG-I) recognizes double-stranded viral RNAs (dsRNAs) containing two or three 5 phosphates. A few reports of 5 -PPP-independent RIG-I agonists have emerged, but little is known about the molecular principles un-derlying their recognition. We recently found that the bent duplex RNA from the influenza A panhan-dle promoter activates RIG-I even in the absence of a 5-triphosphate moiety. Here, we report that non-canonical synthetic RNA oligonucleotides contain-ing G-U wobble base pairs that form a bent helix can exert RIG-I-mediated antiviral and anti-tumor ef-fects in a sequence- and site-dependent manner. We present synthetic RNAs that have been systemati-cally modified to enhance their efficacy and we out-line the basic principles for engineering RIG-I ago-nists applicable to immunotherapy.

INTRODUCTION

RIG-I is a cytosolic ATP-dependent RNA sensor in the in-nate immune system (1–3). When RNA viruses enter the cytosol, RIG-I binds viral RNAs and triggers a signal via mitochondrial antiviral-signaling protein (MAVS) that ac-tivates interferon regulatory factor 3 (IRF3) and NF-␬B.

This induces the type I interferons and suppresses viral propagation (4,5). Aside from its role in host defense against viruses, RIG-I activation leads to immunogenic cell death in carcinomas through the TNF-related apoptosis-inducing ligand (TRAIL) and Noxa (6,7). Immunogenic cell death is a type of programmed cell death that enhances tumor-targeting immunity by stimulating cytokine production and by releasing a number of intracellular molecules (e.g. cal-reticulin, heat shock proteins and HMGB1) that activate antigen-presenting cells and cytotoxic lymphocytes (8,9). Thus, RIG-I is a promising target not only for antiviral drug development but also for cancer immunotherapy (6,8) and as a vaccine adjuvant (10).

The specific molecular structures in viral RNAs that trig-ger RIG-I activation reside at their 5 ends. RIG-I recog-nizes dsRNAs containing uncapped 5-PPP (11) or 5-PP (12) which are found in RNA virus genomes and therefore produced during viral replication. Recent publications re-ported that RIG-I can also be activated by dsRNA (cap-0) containing m7G-capped-5-PPP without 2-OH methyla-tion (13,14). Conversely, RIG-I is not activated by the more abundant endogenous RNAs whose 5 ends have m7G-capped-triphosphates with methylation (cap-1 or cap-2) or monophosphates (5-P). Thus, the structure of 5-PPP acts as a molecular switch in the activation of RIG-I-mediated signaling. RIG-I also recognizes polyinosinic:polycytidylic acid (poly (I:C)) (15), poly (A:U) (16), and poly U tracts

*To whom correspondence should be addressed. Tel: +82 42 350 2828; Fax: +82 42 350 5828; Email: byongseok.choi@kaist.ac.kr

Correspondence may also be addressed to Suk-Jo Kang. Tel: +82 42 350 2611; Fax: +82 42 350 2610; Email: suk-jo.kang@kaist.ac.kr Correspondence may also be addressed to Meehyein Kim. Tel: +82 42 860 7540; Fax: +82 42 861 4246; Email: mkim@krict.re.kr

These authors contributed equally to this work as first authors.

Present addresses:

Mi-Kyung Lee, Disease Target Structure Research Center, Korea Research Institute of Bioscience and Biotechnology, 125 Gwahak-ro, Yuseong-gu, Daejeon 34141, South Korea.

Jiyoun Min, Drug Discovery Center, Life Sciences R&D, E8 Block LG Science Park, 30 Magokjungang 10-ro, Gangseo-gu, Seoul, 07796, South Korea.

C

The Author(s) 2018. Published by Oxford University Press on behalf of Nucleic Acids Research.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original work is properly cited. For commercial re-use, please contact

(2)

(17), but the molecular mechanism by which this activation occurs is unknown.

We recently reported that the bent structure in the con-served promoter region of the influenza A virus (IAV pan-handle promoter) activates RIG-I-mediated immune re-sponses without requiring 5-triphosphates (18). The RNA helix in this structure contains two G-U wobble base pairs that sandwich an A-U base pair. This configuration bends at an angle of 46◦, causing the G–U pairs to pivot around the A–U pair (19). Each G–U base pair causes the RNA he-lix to twist clockwise by∼23◦, forming an interhelical bend (19) (Supplementary Figure S1). Here, we show that this in-terhelical bend formed by the two G–U wobble base pairs is a novel structurally-defined molecular pattern recognized by RIG-I.

MATERIALS AND METHODS RNA oligonucleotides

All 5-OH RNA oligonucleotides without chemical modi-fications were purchased from Integrated DNA Technolo-gies or Bioneer. CBS-7, CBS-13-BPS, CBS-21, CBS-22, CBS-23 and CBS-31 were purchased from Dharmacon or Bioneer. 5-Fluorescein isothiocyanate (FITC)-labeled CBS RNAs were purchased from Bioneer. All RNA oligonu-cleotides were purified by HPLC. RNA oligonuoligonu-cleotides were dissolved in water and subsequently dialyzed against a storage buffer (10 mM potassium phosphate, pH 6.6, 100 mM sodium chloride, 0.1 mM EDTA). Prior to experi-ments, all RNA oligonucleotides except CBS-13-BPS were re-annealed by heating at 95◦C for 5 min followed by cool-ing on ice for at least 30 min. CBS-13-BPS was re-annealed by heating at 95◦C for 5 min followed by slow cooling to room temperature for an hour. The annealed RNA oligonu-cleotides were then dialyzed against the desired buffer for each experiment.

5-PPP RNA (PPP-CBS-4 and CBS-10) was prepared by in vitro transcription using double-stranded DNA encod-ing the RNA sequence of CBS-4 or CBS-10 and containencod-ing the T7 promoter at its 5-end (Integrated DNA Technolo-gies). T7 RNA polymerase (prepared in house) was used to transcribe the DNA sequence in the presence of nucleotide triphosphates. The RNA was then purified by 20% (w/v) polyacrylamide gel electrophoresis (PAGE) in 7 M urea and 1× Tris–borate–EDTA (TBE). The RNA was electroeluted in 0.5× TBE buffer, precipitated in ethanol, dissolved in wa-ter, and desalted using dialysis in water. The desalted RNA was then exchanged into storage buffer.

Protein expression and purification

The gene encoding human RIG-I C-terminal domain (hRIG-I CTD; 803–925) was amplified by polymerase chain reaction (PCR) and cloned into the 2M-T vector (Ad-dgene). The expression plasmid was transformed into the Escherichia coli BL21 (DE3) strain. The protein was over-expressed in lysogeny broth (LB) medium with 0.4 mM isopropyl ␤-D-1-thiogalactopyranoside (IPTG) until the OD600 reached∼0.6. After incubation at 17◦C for an

ad-ditional 18 h, the cells were harvested and lysed in Ni-NTA binding buffer (20 mM Tris–HCl buffer, pH 7.9,

500 mM sodium chloride, 5 mM imidazole, 5 mM 2-mercaptoethanol) by sonication. The resulting cellular de-bris was removed by centrifugation and the supernatant was collected for affinity chromatography purification us-ing Ni-NTA agarose (Qiagen). The protein was then eluted with elution buffer (20 mM Tris–HCl, pH 7.9, 500 mM sodium chloride, 250 mM imidazole). The maltose bind-ing protein (MBP) tag was removed by TEV protease di-gestion and dialyzed against buffer (20 mM Tris–HCl, 100 mM sodium chloride, 5 mM 2-mercaptoethanol, pH∼8.0) overnight at 4◦C. The protein was further purified using Ni-NTA agarose (Qiagen) and ion exchange chromatography using an SP-HP column (GE Healthcare). The eluted pro-tein was then dialyzed against the final buffer (10 mM potas-sium phosphate buffer, 50 mM sodium chloride, 5 mM 2-mercaptoethanol, pH 6.87)

The human RIG-I gene encoding the helicase and its C-terminal domains (amino acids 200–925 residues) was amplified by PCR using complementary DNA (pEF-BOS-RIG-I, Addgene). The amplified product was cloned into the pET-His6-GST-TEV-LIC vector (Addgene). The pro-tein was then expressed in the BL21(DE3) strain. The bac-terial cells were grown in LB medium with 100␮g/ml of ampicillin. Gene expression was induced by the addition of 1 mM IPTG until the OD600reached 0.4∼0.6. After

incu-bation at 16◦C for an additional 20 h, the cells were har-vested by centrifugation at 5000 rpm for 15 min, resus-pended in 20 mM Tris–HCl, 500 mM sodium chloride, 10 mM imidazole at pH 7.9, and lysed by sonication in the presence of phenylmethanesulfonyl fluoride (PMSF). The resulting cellular debris was removed by centrifugation at 15 000 rpm for 1 h. The proteins were then purified from the supernatant by His-Binding Ni-NTA affinity chromatog-raphy (EMD Biosciences). The N-terminal glutathione S-transferase (GST) tag was removed with TEV protease at 4◦C. The digested protein was dialyzed against a buffer con-taining 25 mM HEPES, 100 mM sodium chloride, 5 mM 2-mercaptoethanol, pH 7.4 and purified further in a Q-Sepharose column (GE Healthcare) with a linear gradient from 0 to 1 M NaCl in the same HEPES buffer. The protein-containing fractions were concentrated using an Amicon Ultra-15 (Millipore, MWCO 30,000) and purified with a Superdex 75 (GE Healthcare) gel-filtration column equili-brated with a buffer containing 20 mM Tris–HCl, 100 mM sodium chloride and 1 mM dithiothreitol at pH∼7.5. Cells and viruses

Cell lines of human lung adenocarcinoma (A549), Madin-Darby canine kidney (MDCK) and Calu-3 were purchased from the American Type Culture Collection (ATCC). They were maintained in RPMI 1640 medium (Hyclone), min-imum essential medium (MEM; Hyclone), and Eagle’s MEM (EMEM; ATCC) respectively, each supplemented with 10% Fetal Bovine Serum (FBS) (Gibco).

The influenza A virus, A/Puerto Rico/8/34 (PR8; H1N1), was obtained from ATCC. The mouse-adapted PR8 (maPR8) was kindly provided by Prof. H.-J. Kim (Chung-Ang University, Seoul, Korea). Recombinant PR8-EGFP and PR8(H275Y)-PR8-EGFP viruses were constructed as previously reported (20–22). The viruses were amplified

(3)

in 10-day-old embryonated chicken eggs for 3 days. Allan-toic fluid was harvested and infectious virus titers were de-termined by plaque assays in MDCK cells (20). The stocks were stored at –70◦C before use.

Human embryonic kidney cell lines (HEK293 and HEK293T) were purchased from ATCC. They were main-tained in DMEM medium (Welgene) supplemented with 10% FBS (ThermoFisher Scientific).

The murine panc02 cell line was kindly provided by Prof. Christiane Bruns at University Hospital Magdeburg, Germany. The panc02 cells were maintained in DMEM (Welgene) with NEAA (non-essential amino acid solution, Sigma-Aldrich), vitamin (ThermoFisher Scientific) and 10% FBS (ThermoFisher Scientific).

The human liver cancer cell line, SNU886 (Korean Cell Line Bank), was maintained in RPMI medium (Welgene) supplemented with 10% FBS. 293-hTLR3 cells (Invivo-Gen), which are HEK293 cells stably transfected with the human TLR3 gene, were maintained in DMEM supple-mented with 10% FBS. Huh 7 and Huh 7.5 cells (kindly pro-vided by Prof. Eui-Cheol Shin, KAIST) were maintained in DMEM supplemented with 10% FBS.

Mice

Female C57BL/6J mice (6–7 weeks of ages) were purchased from DBL (Seoul, Korea) and used for tumor experiments. All procedures were performed with the approval of the In-stitutional Animal Care Committee of KAIST.

Female BALB/c mice (Orientbio) were maintained until 7 weeks of age before being immunized with the inactivated influenza A virus vaccine. All animal experiments with in-fectious influenza virus described here were approved by the Institutional Animal Care and Use Committee (IACUC) at KRICT.

NMR spectroscopy

CBS-4 was exchanged into NMR buffer (10 mM potassium phosphate, pH∼6.6, 50 mM NaCl, 1 mM dithiothreitol, 10% v/v D2O) by dialysis. CBS-7 was exchanged into the

same NMR buffer, except with 1 mM DTT replacing the 0.1 mM EDTA. 1D imino proton spectra of CBS-4 and CBS-4-Min were measured with jump-return pulses for wa-ter suppression at 15◦C. All experiments were performed in 700 MHz Avance Bruker or Agilent spectrometers equipped with 5 mm triple-resonance cryogenic probes. All NMR spectra were processed using NMRPipe (23) and SPARKY 3 (24).

Isothermal titration calorimetry

Isothermal titration calorimetry was performed at 25◦C us-ing a Microcal VP-ITC instrument (Malvern Instruments Ltd.). The hRIG-I CTD and the RNAs (CBS-4 and CBS-7) were dialyzed against ITC buffer (10 mM potassium phos-phate, 50 mM sodium chloride, 5 mM 2-mercaptoethanol, pH∼6.87). The cell chamber contained 5–10 ␮M of RNA sample, and the titration syringe was loaded with 50–100 ␮M of hRIG-I CTD. The concentration of the hRIG-I CTD in the syringe was∼10 times higher than the concen-tration of the RNA sample in the chamber (e.g. 10␮M of

RNA and 100␮M of hRIG-I CTD). hRIG-I CTD in the syringe was injected at 30 successive points into the cham-ber containing the RNA sample. A blank experiment for titrating the hRIG-I CTD into a chamber without an RNA sample was also conducted for each RNA sample. These blank data were subtracted from the actual titration data to eliminate miscellaneous heat contributions including the heat of dilution. The resulting titration data were then pro-cessed using the Microcal software (Malvern Instruments Ltd.) that was provided with the VP-ITC instrument. The reported thermodynamic parameters including the dissoci-ation constant and the stoichiometry were obtained by fit-ting the data to a one-site binding model.

Electrophoretic mobility shift assay

The human RIG-I CTD without its CARD domains (hRIG-ICARD) was incubated with 0.1 mM of ATP and 20␮M of CBS-4, CBS-7, PPP-CBS-4 in a buffer (20 mM Tris–HCl, 100 mM sodium chloride, 1 mM dithiothreitol, pH∼7.5) at increasing molar ratios of hRIG-I CARD to RNA (i.e. 0:1, 0.5:1, 1:1, 2:1, 3:1). The reaction mixture was incubated on ice for 30 min. The 2× loading buffer (40 mM Tris–HCl, 200 mM sodium chloride, 2 mM dithiothreitol, 80% glycerol by volume, pH∼7.5) was added to the reaction mixture prior to loading on a non-denaturing gel (9% poly-acrylamide, 1× TBE). The reaction mixture was resolved by running the electrophoresis in the cold room (4◦C). The gel was then fixed in 1 M acetic acid for 30 min and stained with methylene blue for visualization.

RNA interference and quantitative real-time PCR

HEK293T cells were seeded at 3 × 106 cells per 100

mm dish in DMEM (Welgene) supplemented with 10% FBS. The cells were incubated at 37◦C and 5% CO2

for 24 h. After 1 day, the cells were transfected with 200 pmol hRIG-I siRNA (Bioneer) or control siRNA (Bioneer) using 20␮l of lipofectamine LTX (ThermoFisher Scientific) and 8 ␮l of plus-reagent (ThermoFisher Sci-entific) in OPTI-MEM (Thermo Fisher SciSci-entific). The transfection efficiencies of CBS-1, CBS-4 and CBS-7 were measured by using fluorescently labeled RNA con-structs and flow cytometry (Supplementary Figure S2). We found that their transfection efficiencies were compa-rable to each other. The hRIG-I siRNA sequences are 5-ACGGAUUAGCGACAAAUUUAA dTdT-3 (sense) and 5-UUAAAUUUGUCGCUAAUCCGU dTdT-3 (an-tisense). After 4 h of siRNA treatment, the OPTI-MEM was replaced with 7 ml of DMEM with 10% FBS and incubated at 37◦C and 5% CO2 for an additional 20 h.

The siRNA-treated cells were detached with 0.25% trypsin– EDTA (ThermoFisher Scientific) and then split into six-well plates, with each six-well containing 1.5× 105 cells. The

cells were then incubated at 37◦C and 5% CO2for 42 h prior

to experiments.

After 42 h of incubation, the siRNA-treated cells were transfected with 1 ␮M of each CBS RNA in 4 ␮l of lipofectamine LTX and 1.6 ␮l of plus-reagent in OPTI-MEM. After 4 h of transfection, the cells were incubated in DMEM with 10% FBS for another 2 h. The cells were

(4)

harvested using the TRI-reagent (ThermoFisher Scientific). Their RNA was extracted using chloroform and precip-itated with isopropanol. The RNA pellets were washed with 75% ethanol and dissolved in DEPC-treated wa-ter. The dissolved RNA was then treated with DNase (Promega) to remove any residual DNA. This RNA was then subjected to reverse transcription using Superior-Script III reverse transcriptase (Enzynomics). Quantita-tive real-time PCR analysis (Rotorgene-Q, Qiagen) was performed using EvaGreen Dye (Biotium), h-taq DNA polymerase (Solgent), and the following primers: hu-man interferon-stimulated gene 56 (ISG56) forward, 5 -GCCTCCTTGGGTTCGTCTACAA-3, ISG56 reverse, 5 -GAGGCCATGTGGGCCATGAG-3, human glyceralde-hyde 3-phosphate dehydrogenase (GAPDH) forward, 5 -GCATTGCCCTCAACGACCAC-3, GAPDH reverse, 5 -GAGGCCATGTGGGCCATGAG-3 RIG-I forward, 5 -GCATTGCCCTCAACGACCAC-3, RIG-I reverse, 5 -GAGGCCATGTGGGCCATGAG-3. ISG56 and RIG-I expression data were normalized to GAPDH.

Huh 7 or Huh 7.5 cells were seeded in six-well plates at 3 × 105 cells per well in DMEM (Welgene) supplemented

with 10% FBS. The cells were incubated at 37◦C and 5% CO2overnight prior to transfection of CBS RNA using the

same procedure as described above. Western blot analysis

A549 cells were seeded at 1 × 106 cells per well in

six-well plates. The next day, synthetic CBS RNAs (100 nM) or polyinosine–polycytidylic acid [poly (I:C)] (0.1 ␮g/ml) were transfected into cells using the Lipofectamine 2000 reagent (Invitrogen) according to the manufacturer’s in-structions. After 24 h, the cell culture supernatants were re-moved and the cell monolayers were mock-infected or in-fected with PR8 at a multiplicity of infection (MOI) of 0.1 for 1 h at 37◦C. After the unabsorbed virus was removed, the cells were further maintained in culture medium containing 0.1␮g/ml TPCK-trypsin (Sigma-Aldrich). On day 1 or 2 post-infection (p.i.), the cells were lysed with the M-PER reagent (Thermo Scientific). Then, 30 ␮g of total protein was loaded onto a 10% SDS-PAGE gel for electrophore-sis and transferred to a PVDF membrane (Immobilon-P membrane; Millipore). Viral NP protein was detected with an anti-NP antibody (11675-MM03; Sino Biological) and HRP-conjugated goat anti-mouse IgG (Sigma-Aldrich). Vi-ral NS1 protein was detected with an anti-NS1 antibody (sc-17596; Santa Cruz Biotechnology) and HRP-conjugated donkey anti-goat IgG. Beta-actin was detected as a loading control with an anti-␤-actin antibody (Sigma-Aldrich) and an HRP-conjugated goat anti-mouse antibody. The protein bands were visualized with SuperSignal West Pico Chemi-luminescent Substrate (Pierce) or Western ECL Femto Kit (LPS Solution) using an image analyzer (LAS-4000; Fuji-film).

For immunoblot of ISG56, HEK 293T cells were seeded at 1× 106cells per well in 6-well plates. The next day,

syn-thetic CBS RNAs (1␮M) or polyinosine-polycytidylic acid [poly (I:C)] (2 ␮g/ml) were transfected into cells using 4 ␮l of Lipofectamine LTX and 1.6 ␮l of Plus reagent (In-vitrogen). After 6 hours, the cells were harvested by using

trypsin–EDTA (Gibco) and lysed with 1× SDS-reducing sample buffer. Cell lysates were analyzed by 10% SDS-PAGE followed by immunoblotting with antibodies against ISG56 (Cell Signaling) and tubulin (Sigma).

Live-cell imaging of influenza A virus infected cells

Calu-3 cells were seeded onto six-well plates (6× 105cells

per well). After two days, they were transfected with syn-thetic RNAs using the Lipofectamine 2000 reagent. The next day, the cells were washed with PBS and infected with PR8-EGFP or PR8(H275Y)-EGFP at an MOI of 0.05 for 1 h at 37◦C. The media was exchanged with fresh EMEM with or without 20␮M oseltamivir carboxylate (OSV-C; US Biological). Green fluorescent images were captured every 2 h for 3 days using an IncuCyte FLR (Essen BioScience). Fluorescent objects per well were counted using the built-in software.

Vaccination and virus challenge

The influenza virus vaccine was prepared by ultracentrifu-gation of PR8 virus inactivated with 0.02% formalin (25). Seven-week-old female mice (five per group) were immu-nized intranasally with 0.5␮g vaccine alone or in combi-nation with 5␮g RNA adjuvant, CBS-13-PBS complexed in vivo with jetPEI (Polyplus-transfection) or with poly (I:C). Three weeks after their vaccination, the mice were in-tranasally challenged with 100-fold excess of 50% minimal lethal dose (MLD50) of maPR8. Then, they were observed

for changes in body weight and mortality for an additional 11 days.

Measurement of IgG level in mouse serum by ELISA Two weeks after immunization, sera were collected from mice and treated with receptor-destroying enzyme (RDE). The levels of influenza A virus (IAV)-specific IgG were determined by ELISA. 96-well ELISA plates (MaxiSorp; Nunc) were coated with the vaccine (0.5 ␮g per well). They were incubated with serum samples 10-fold diluted in PBS containing 5% horse serum and 0.05% Tween 20. IgG specifically bound to IAV was detected using HRP-conjugated goat mouse IgG Fab fragment anti-body (Southern Biotech). After addition of 2,2’-azinobis (3-ethylbenzothiazoline-6-sulfonic acid)-diammonium salt (ABTS) and ABTS peroxidase stop solution, optical den-sity was measured at 405 nm using an ELISA microplate reader (SpectraMax M3; Molecular Devices).

In vitro apoptotic activity of RIG-I agonists

0.5× 105 panc02, SNU886, MEF or HEK293 cells were

plated per well on 24 well-plates. The cells were then trans-fected with 1␮g/ml of the appropriate RNA molecule or poly (I:C) (InvivoGen) using lipofectamine LTX and incu-bated for 24 h at 37◦C and 5% CO2. Note that 1␮g/ml

of CBS-1, CBS-7, CBS-13-BPS and CBS-31 corresponds to ∼156, ∼149, ∼35 and ∼34 nM, respectively. The cells were then harvested and washed with 1× binding buffer (10 mM HEPES, 140 mM NaCl, 2.5 mM CaCl2). The cells

(5)

were stained with FITC-labeled annexin-V (Biolegend) and 7-aminoactinomycin D (7-AAD; BD Bioscience) in 100␮l of 1× binding buffer for 15 minutes at room temperature. An additional 400␮l of 1× binding buffer was added, and the cells were analyzed on an LSRFortessa flow cytometer (BD Bioscience) with the FlowJo software (Tree Star, Inc.). Prophylactic anti-tumor effect of apoptotic panc02 cells Before treatment with RNA oligonucleotides, 2 × 106

murine panc02 cells were plated on 100 mm plates. The cells were transfected with 1␮g/ml CBS-13-BPS as described in the in vitro apoptotic activity assay section. After 24 h of incubation with the RNAs, the apoptotic––floating––cells were harvested and quantified by trypan blue staining. C57BL/6J mice, anesthetized by isoflurane inhalation, were injected subcutaneously on the right flank with 5 × 105

apoptotic cells induced by CBS-13-BPS. Seven days later, the mice were injected subcutaneously on the left flank with 5× 105viable panc02 cells. The volume of the resulting

tu-mors was monitored beginning on day 5 after tumor cell in-jection. Tumor volume was calculated by V= (L × W2)/2,

where V is tumor volume, L is length in millimeters and W is width in millimeters.

In vivo anti-tumor experiment by intratumoral injection of

RNA

C57BL/6J mice were subcutaneously injected with 2.5 × 105viable panc02 cells on the left flank. Then, 7 days after

cancer cell transplantation, 25␮g of CBS-7 were admixed with in vivo jetPEI®(Polyplus) and 5% glucose, incubated

for 15 min at room temperature, and injected directly into the tumor mass. RNA was injected twice a week for 2 weeks for a total of 4 injections. Tumor growth was monitored be-ginning on day 6 after tumor injection by measuring tumor volume. Tumor volume was calculated using the same equa-tion as above: [V= (L × W2)/2].

RNase I reaction

CBS-4 and CBS-7 (1.2 ␮g each) were incubated with 20 units of RNase I (ThermoFisher Scientific) in a buffer (20 mM Tris–HCl, 100 mM sodium chloride, 1 mM dithiothre-itol, pH 7.5) at 37◦C for 1, 6, 24 or 48 h. After the incu-bation, loading dye (0.01% w/v xylene cyanol, 0.01% w/v bromophenol blue, 7 M urea, 0.5× TBE) was added to the reaction mixture. Then, the reaction mixture was heated at ∼95◦C for∼5 min and stored at room temperature. All

re-action mixtures stopped at different time points were heated again prior to loading on a denaturing gel (20% polyacry-lamide, 7 M urea, 1× TBE). CBS-4 and CBS-7 untreated with RNase I were loaded on the gel in the same manner. After fixation in 1 M acetic acid, the RNA bands were vi-sualized by staining with methylene blue.

Dual luciferase reporter gene assay

293-hTLR3 cells (InvivoGen) were seeded in 12-well plates (1 × 105 cells/well). After 24 hours, the cells were

trans-fected with 1 ␮g of pGL3-IFN-␤-FLuc, that was con-structed in-house on the basis of the previous report (26),

or pNF-␬B-Luc (Stratagene) together with 10 ng of phRL-CMV (Promega) using Lipofectamine 2000 (ThermoFisher Scientific). After washing with PBS, the cells were incu-bated with 1 ml fresh DMEM/10% FBS containing CBS-13-BPS or poly (I:C) in the absence of any transfection reagents for 18 hours. The activities of firefly and Renilla luciferases were then measured using the dual luciferase as-say kit (Promega).

RESULTS

RIG-I-dependent immunogenicity of the Bent RNA duplex To determine the minimal motif in bent RNAs necessary for RIG-I activation, we synthesized CBS-4, a short hairpin RNA of 8 base pairs containing two G-U wobble base pairs flanking an A-U base pair (Figure1A). We generated CBS-4 by chemical synthesis to ensure that it lacked 5phosphates. We also incorporated a UUCG loop (27) to help the RNA form a hairpin structure. By NMR spectroscopy, we ob-served two characteristic nuclear Overhauser effect (NOE) cross peaks corresponding to the two G–U base pairs and assigned imino protons appropriately (Supplementary Fig-ure S3). Thus, we were able to confirm that CBS-4 folds into the secondary structure depicted in Figure1A.

We then examined the immunogenicity of CBS-4 by mon-itoring its ability to induce interferon-stimulated gene 56 (ISG56), an IRF-3 target gene that is activated by RIG-I. Indeed, CBS-4 induces ISG56 expression in a RIG-I-dependent manner (Figure 1B) to levels similar to those induced by a 5-PPP dsRNA (CBS-10, Figure1A). As ex-pected, CBS-1, a 5-OH hairpin duplex RNA formed en-tirely from canonical Watson-Crick base pairs (Figure1A), fails to induce ISG56 expression (Figure1B). CBS-1 was therefore used as a negative control throughout this study. We also confirmed that the induction of ISG56 by CBS-4 cannot be attributed to the UUCG loop, since a duplex RNA without the loop (CBS-4-Min) induces ISG56 equally well (Figure1A and B). We confirmed that CBS-4-Min ex-isted in the double-stranded form, and two single-strands did not confer the immunogenicity (Supplementary Figure S4).

We next asked which structural features of CBS-4 allow it to act as a RIG-I agonist despite lacking a 5-PPP moi-ety. To determine whether both G-U base pairs are neces-sary for RIG-I activation, we designed two mutants––CBS-4-U3C and CBS-4-U16C––in which each of the G-U wob-ble base pairs is replaced with a G–C Watson–Crick base pair (Figure1A). Neither CBS-4-U3C nor CBS-4-U16C in-duce ISG56 expression (Figure1C), indicating that a single G-U base pair is insufficient for RIG-I activation. This is consistent with our previous finding that mutant influenza panhandle promoters containing only one G–U base pair do not form the bent helix (28). To determine whether the location of the G-U pairs affects CBS-4’s ability to acti-vate RIG-I, we moved its G-U base pairs further from the open end of the duplex, generating CBS-4-Mid (Figure1A). Again, CBS-4-Mid treatment does not induce expression of ISG56 (Figure1C). Similarly, replacement of its G–U wob-ble base pairs with U–U or A–G wobwob-ble base pairs also pre-vents CBS-4 from inducing ISG56 expression (CBS-50 and 51; Supplementary Figure S5). Thus, the immunogenicity

(6)

Figure 1. Bent RNAs induce RIG-I-dependent ISG56 expression. (A) Secondary structures of the RNA constructs used in the experiments. The dashed

red boxes denote wobble base pairs. (B) RIG-I-dependent induction of ISG56 in HEK293T cells by CBS-4 or CBS-4-Min, as monitored by RT-PCR. CBS-1 and CBS-10 were used as negative and positive controls, respectively. Cells were transfected with 1␮M of the indicated RNA constructs 6 hours prior to measurement. Medium refers to cells transfected with lipofectamine alone. Expression levels of ISG56 and RIG-I were normalized to GAPDH. (C) The induction of ISG56 by 1␮M CBS-4-U3C, CBS-4-U16C, or CBS-4-Mid in HEK293T cells was measured as described in Figure1B. (B, C) The boxed graphs show the RIG-I siRNA knockdown efficiency. The black and red asterisks indicate significant differences in comparisons to medium and control (Ctrl) siRNA, respectively. Data are representative of two independent experiments (N= 3, mean ± SD). *P < 0.05; **P < 0.01.

of CBS-4 depends on the presence of two G–U base pairs located close to the terminus of the RNA duplex.

We next asked whether other types of interhelical bends lacking 5-PPP moieties also activate RIG-I. We synthesized two bent RNA constructs each containing a helix-junction-helix (HJH) motif. One contains a Kt-7 K-turn motif (29) with a GAA bulge (Kinked RNA, Figure2A); the helices of the K-turn motif bend at an angle of∼60◦around the bulge (Supplementary Figure S1) (29). The other HJH RNA is derived from the transactivation response (TAR) element of human immunodeficiency virus type-1 (HIV-1) (30) and contains a UCU bulge (TAR-Bulge; Figure 2A); the he-lices of HIV-1 TAR bend at an angle of∼47◦ (Supplemen-tary Figure S1) (31,32). Remarkably, both Kinked RNA and TAR-Bulge are capable of inducing RIG-I-dependent ISG56 expression (Figure2B). Reducing the angle of TAR’s interhelical bend by deleting one or two of its bulge nu-cleotides (TAR-2 and TAR-1, Supplementary Figure S6A) (33) blocks TAR’s induction of ISG56 (Supplementary Fig-ure S6B). This suggests that interhelical bends are impor-tant structural cues in the RIG-I activation.

Backbone modification of the bent RNA duplex to Gain re-sistance to degradation by RNase

We next explored the potential of synthetic bent RNAs as immunotherapeutic agents. To increase the stability of these RNAs inside the body, we replaced the phosphate back-bone of CBS-4 with phosphorothioate linkages to create the nuclease-resistant hairpin CBS-7 (Figure3A and Sup-plementary Figure S7). We confirmed by NMR that the secondary structure of CBS-7 is similar to that of CBS-4 (Supplementary Figure S3). Previous structural studies sug-gest that the binding of a given dsRNA to RIG-I is in-fluenced by electrostatic interactions between RIG-I and the dsRNA phosphate backbone (34–39). In an isothermal titration calorimetry (ITC) analysis, we found that the bind-ing affinity of CBS-7 to the C-terminal domain of human RIG-I (hRIG-I CTD, Supplementary Figure S8A and S8B) is 31-fold lower than that of CBS-4. This may be due to the reduced polarity of phosphorothioate linkages compared to phosphate linkages. Still, we confirmed using an elec-trophoretic mobility shift assay (EMSA) that CBS-7 does form a complex with a human RIG-I mutant lacking two CARD domains (hRIG-ICARD) (Supplementary Fig-ure S8A and S8C). Furthermore, CBS-7 and CBS-4 induce similar levels of ISG56 expression (Figure3B). CBS-1-PS,

(7)

Figure 2. RIG-I-dependent immunogenicity of the bulged RNA. (A) Secondary structures of the bulged RNA (green; Kinked RNA and TAR-Bulge). The

hairpin RNA without a bulge (CBS-3) used as a control in the experiments. (B) The induction of ISG56 expression by 1␮M Kinked RNA, TAR-Bulge, CBS-3 or CBS-10 was measured as described in Figure1B. The boxed graphs show the RIG-I siRNA knockdown efficiency. The black and red asterisks indicate significant differences in comparisons to medium and control (Ctrl) siRNA, respectively. Data are representative of two independent experiments (N= 3, mean ± SD). *P < 0.05; **P < 0.01.

the phosphorothioate-linked version of the CBS-1 control, does not induce ISG56 expression (Figure 3B), excluding the possibility that phosphorothioate backbones in general activate RIG-I. These results suggest the bend structure of our synthetic RNAs permits significant backbone alter-ations without substantially compromising their ability to activate RIG-I.

Elongation of the bent RNA duplex to enhance the immuno-genicity

Poly (A–U) stretches with 5-PPP moieties have been shown to enhance the immunogenicity of RIG-I (16). Given our finding that the UUCG loop is dispensable (CBS-4-Min, Figure 1A), we synthesized CBS-13-BPS, a modified ver-sion of CBS-7 with 35 Watson–Crick base pairs containing 26 consecutive A–U base pairs and a UAA loop in place of the original UUCG loop (Figure 3A). Remarkably, CBS-13-BPS induces 7-fold more ISG56 expression than CBS-7 (Figure3C) and the immunogenicity was RIG-I-dependent (Figure 3D). CBS-13-BPS also increased the expression level of ISG56 protein (Supplementary Figure S9). This ac-tivity of CBS-13-BPS, however, is independent of toll-like-receptor 3 (TLR3; Supplementary Figure S10), whose acti-vation requires at least 40 bp of poly (A:U) dsRNA (40,41). Antiviral activity of the chemically modified bent RNA duplex We next examined the antiviral activity of 7 and CBS-13-BPS. We treated a human lung epithelial cancer cell line, A549, with CBS-7 or CBS-13-BPS for 24 hours before in-fecting them with an H1N1 strain, PR8, of the influenza A virus for an additional 24 hours. We then monitored vi-ral replication by immunoblotting for the vivi-ral nucleopro-tein (NP) and non-structural pronucleopro-tein 1 (NS1) (Figure4A). Compared to pre-treatment with CBS-1 and to a lesser ex-tent poly (I:C), pre-treatment with CBS-7 produces a

con-sistent dose-dependent reduction in NP and NS1 (Figure

4A). CBS-13-BPS, in contrast, exhibits a very potent an-tiviral effect––more so than CBS-7––completely suppress-ing viral replication at low nanomolar concentrations (Fig-ure4A).

We then used a PR8 virus expressing enhanced green fluo-rescent protein (EGFP) [EGFP-PR8 (WT)] to compare the effectiveness of CBS-13-BPS to that of the anti-influenza drug oseltamivir carboxylate in altering the time-course of viral replication. Throughout the course of infection, we found CBS-13-BPS suppresses viral replication at a level comparable to oseltamivir, but at a 200-fold lower molar concentration. Moreover, we also found CBS-13-BPS is ac-tive against an oseltamivir-resistant PR8 variant [EGFP-PR8(H275Y)] (Figure 4B), emphasizing the potential of CBS-13-BPS as an antiviral agent. The controls, CBS-1 and CBS-1-PS, which do not induce ISG56 expression, are inca-pable of suppressing viral replication.

CBS-13-BPS’s strong antiviral effect on the influenza virus suggested that it may also be useful as a vaccine ad-juvant. We therefore mixed an influenza vaccine contain-ing inactivated PR8 with CBS-13-BPS complexed with the transfection agent jetPEI®or poly (I:C). We then

adminis-tered these intranasally to mice (Figure4C). Three weeks later, we challenged the vaccinated mice with a mouse-adapted PR8 virus at its lethal dose and monitored their weight and survival daily for 11 days. The mice given the CBS-13-BPS-admixed vaccine (Vac.CBS13BPS) com-pletely survived this lethal influenza infection, while mice treated with vaccine alone or with the poly (I:C)-admixed vaccine (Vac.Poly(I:C)) suffered 60% or 20% death, respec-tively (Figure4C). The influenza A virus (IAV)-specific IgG levels were higher in mice treated with Vac.CBS13BPS than in mice treated with the vaccine alone (Figure4D). Thus, CBS-13-BPS shows significant potential for use as an adju-vant in vaccine development.

(8)

Figure 3. Backbone modification and elongation of the bent RNA do not deteriorate the immunogenicity. (A) Secondary structures of the bent RNA

used in this experiment. The red dashed boxes denote G-U wobble motifs. Orange nucleotides denote the nucleotides with phosphorothioate backbone. The boxed A–U base pairs in CBS-13-BPS represent 26 consecutive base-pairs. (B) RIG-I-dependent induction of ISG56 by 1␮M CBS-1, CBS-1-PS, CBS-4 or CBS-7 was monitored by RT-PCR as described in Figure1B. The boxed graph shows the RIG-I siRNA knockdown efficiency. (C) The induction of ISG56 by 1␮M CBS-1, CBS-4, CBS-7, or CBS-13-BPS in HEK293T cells was measured as described in Figure1B. (B, C) Data are representative of two independent experiments. The error bars represent the standard deviations of triplicate experiments. The black and red asterisks indicate significant differences in comparisons to medium only and control (Ctrl) siRNA, respectively. *P< 0.05. (D) The induction of ISG56 by 1 ␮M CBS-1, CBS-13-BPS, or low molecular weight poly (I:C) in Huh 7 and Huh 7.5 cells was measured under the same conditions as described in Figure1B. Data are representative of two independent experiments. The error bars represent the standard deviations of triplicate experiments. **P< 0.01.

Anti-tumor activity of the bent RNA duplex

Previous studies have suggested that RIG-I agonists can induce apoptosis of cancer cells (6–8). We therefore asked whether our synthetic RNAs also induce cancer cell death. Twenty-four hours after transfecting the mouse pancreatic cancer cell line panc02 and human liver cancer cell line SNU886 with CBS-1, CBS-7, CBS-13-BPS or poly (I:C), we stained the transfected cells with fluorescently-labeled annexin-V and 7-aminoactinomycin D (7-AAD). Apop-totic cells display phosphatidylserine on the outer leaflet of their plasma membranes. This phosphatidylserine can be detected by annexin-V. Necrotic cells, in contrast, incorpo-rate the membrane impermeable dye 7-AAD. When we de-tected apoptotic cells (annexin-V+7-AAD) by flow

cytom-etry, we found that although CBS-7 did induce apoptosis

in panc02 and SNU886 cells, it did so less than poly (I:C) (Figure5A and B). CBS-13-BPS, in contrast, induced apop-tosis 4-fold more than CBS-7 and 2-fold more than poly (I:C) in panc02 cells, although it induced less apoptosis than poly (I:C) in SNU886 cells. This suggests that the induc-tion of apoptosis by CBS RNAs may vary depending on the cell types. Importantly, neither CBS-7 nor CBS-13-BPS induced significant apoptosis in the non-cancerous human embryonic kidney 293 (HEK293) (Figure5C) and mouse embryonic fibroblast (MEF) cell lines (Supplementary Fig-ure S11), suggesting enhanced death of tumor cells triggered by CBS-RNAs.

The tumor cell death induced by RIG-I agonists was re-cently shown to enhance anti-tumor immunity in vivo (8). To determine whether our synthetic RNAs can also boost

(9)

anti-Figure 4. Antiviral activities of CBS-7 and CBS-13-BPS in vitro and in vivo. (A) Antiviral activities of CBS-7 and CBS-13-BPS monitored by western

blot. Mock refers to cells transfected with lipofectamine alone. CBS-1 (100 nM), CBS-1-PS (100 nM), poly (I:C) (0.1␮g/ml), and oseltamivir carboxylate (OSV-C; 10␮M) were used as controls. (B) Antiviral activity of CBS-13-BPS against oseltamivir-resistant influenza virus. A549 cells were treated with CBS-1, CBS-13-BPS or OSV-C, and infected with EGFP-expressing wild-type (WT) influenza virus [EGFP-PR8 (WT)] or its oseltamivir-resistant variant [EGFP-PR8 (H275Y)] 12 h later. Infected cells were counted at 2-h intervals for 36 h using a live cell imaging system. Mean fluorescence values were normalized to the maximum fluorescence value 24 h post-infection. Images at 24 h post-infection were visualized by fluorescence microscopy. (C) In vivo activity of CBS-13-BPS as an adjuvant for the influenza vaccine. 0.5␮g of inactivated PR8 (A/H1N1; Vaccine) was administered alone, with 5 ␮g CBS-13-BPS (Vac.CBS13CBS-13-BPS), or with 5␮g poly (I:C) (Vac.Poly(I:C)). Mice were infected with mouse-adapted PR8 (A/H1N1) influenza virus 3 weeks later. The survival rate and weight changes were monitored post-infection for 11 days. Mock refers to mice without any treatment. The mock trace (black) is overlaid with the Vac.CBS13BPS trace (purple). The change of weight: N= 5 (except for Mock, N = 3), mean ± SD. Statistical significance with respect to virus,

*P< 0.05, **P < 0.01, ***P < 0.001. (D) IAV-specific antibody production enhanced by CBS-13-BPS. The IAV-specific serum IgG levels were determined

by ELISA. Mice were immunized with the inactivated influenza A virus vaccine alone or together with CBS-13-BPS (5␮g per head) or with poly (I:C) (5 ␮g per head). The asterisks represent statistical significance compared to the vaccine only group. N = 5, mean ± SD. *P < 0.05, ***P < 0.001.

tumor immunity, we transfected panc02 cells with CBS-13-BPS in vitro and then injected the resulting apoptotic cells subcutaneously into the right flank of mice (Figure

5D). Seven days later, we injected viable panc02 cells sub-cutaneously into the left flank of mice and then monitored tumor growth for 17 days. We observed a clear suppres-sion of tumor growth in mice vaccinated with

CBS-13-BPS-induced apoptotic cells, but not in un-vaccinated mice (Fig-ure5D). We also found that two weeks of CBS-7 injections into panc02-derived mouse tumors induces substantial tu-mor regression with one mouse cleared of tutu-mors (Figure

5E). Together, these results indicate that our RNA-based RIG-I agonists have tremendous potential as cancer im-munotherapeutics.

(10)

Figure 5. Anti-tumor activity of CBS-7 and CBS-13-BPS in vitro and in vivo. (A) Induction of apoptosis in the pancreatic cancer cell line panc02 by CBS

RNAs. Panc02 cells were transfected with 1␮g/ml of CBS-1, CBS-7, CBS-13-BPS, and poly (I:C), admixed with lipofectamine. Twenty-four hours later, apoptotic cells were detected by flow cytometry after staining the cells with annexin-V and 7-AAD. Annexin-V+/7-AADcells were considered apoptotic

and their percentages are shown as a bar graph. Data are representative of two independent experiments (N= 3, mean ± SD). *P < 0.05, ***P < 0.001. (B) Induction of apoptosis in the human liver cancer cell line SNU886 by CBS RNAs. SNU886 cells were treated and analyzed as described in Figure

5A. Data are representative of two independent experiments (N= 3, mean ± SD). ***P < 0.001. (C) No significant induction of apoptosis in the normal human cell line HEK293. HEK293 was treated and analyzed as described in Figure5A. Data are representative of two independent experiments (N= 3, mean± SD). (D) Prophylactic anti-tumor effect of apoptotic panc02 cells induced by CBS-13-BPS. 5 × 105apoptotic panc02 cells, induced in vitro by

CBS-13-BPS, were injected subcutaneously to the right flank of mice. A week later, 5× 105viable panc02 cells were seeded onto the contralateral side.

Injected tumor volumes were then measured for 17 days. Data are representative of three independent experiments (N= 5, mean ± SEM). *P < 0.05. (E) Anti-tumor effect of CBS-7. Five days after tumor seeding, 25␮g of CBS-7 was injected intratumorally twice a week (marked with an asterisk). Changes in tumor volume were then monitored for 16 days. jetPEI (N= 5) and CBS-7 (N = 4). Mean ± SEM. *P < 0.05.

DISCUSSION

Here, we present our results in systematically editing a group of RIG-I agonists to alter their ligand recognition, stability, and potency. We found the incorporation of a terminal interhelical bend facilitates recognition by RIG-I even in the absence of a 5-PPP moiety. The RNA back-bone can be replaced with phosphorothioate linkages to enhance cytosolic delivery and stability (42,43). The hair-pin structure of these agonists facilitates duplex formation

and helps maintain their stability. We also found additional poly (A:U) base pairs increase the potency with which these RNAs act as RIG-I agonists. We were able to demonstrate the immunogenicity of our synthetic RNA oligonucleotides against viruses and tumor cells both in vitro and in vivo, ver-ifying the rationality of our design process.

The fact that the 5-end of our RNA agonists can be mod-ified gives them a distinct advantage over classical RIG-l agonists, which must harbor 5-PPP or 5-PP. To further confirm this, we coupled biotin to the 5-end of CBS-4

(11)

Figure 6. Chemical modifications of CBS RNAs at their 5-end do not affect their activity. (A) Secondary structures of 21, 22, 23 and CBS-31. The dashed red boxes denote the bent duplex motif. Orange nucleotides denote phosphorothioate modifications. The boxed A–U base pairs in CBS-31 represent 26 consecutive base pairs. Biotin and cholesterol are abbreviated as bi and chol, respectively. Chemical structures of 5-biotin, 3-cholesterol, 5-cholesterol are shown. (B) The induction of ISG56 by 1␮M CBS-1, CBS-1-PS, CBS-4, CBS-21 or CBS-22 in HEK293T cells was measured as described in Figure1B. Mean± SD. *P < 0.05, **P < 0.01. (C) The induction of ISG56 by 1 ␮M CBS-1, CBS-4, or CBS-23 in HEK293T cells was measured as described in Figure1B. Mean± SD. *P < 0.05. (D) Induction of apoptosis in the human liver cancer cell SNU886 by treatment with 1 ␮g/ml of CBS-1, CBS-7, CBS-13-BPS or CBS-31 analyzed as described in Figure5A. Data are representative of two independent experiments (N= 3, mean ± SD). ***P

< 0.001.

and CBS-13-BPS to synthesize CBS-21 and CBS-31, re-spectively (Figure 6A). We also coupled cholesterol to the 5-end of CBS-7 to synthesize CBS-23 (Figure6A). Then, we measured their immunogenicity. We found CBS-21 and CBS-23 induce ISG56 expression at levels comparable to CBS-4 (Figure6B and C, respectively) and CBS-31 induces apoptosis in the hepatic cancer cell line (SNU886) at lev-els comparable to CBS-13-BPS (Figure6D). This confirms the flexibility of our RIG-I agonists because the chemical modification of their 5-ends does not seem to hinder their activity. In contrast to the situation with their 5-ends and consistent with a previous report that showed 5-PPP

dsR-NAs containing 3-overhangs induce reduced expression of interferon-␤ (44), we found 3 conjugation of cholesterol (CBS-22) prevents the induction of ISG56 expression (Fig-ure6B). From these observations, we infer that the 5-end of our RNA agonists tolerates chemical conjugations that may improve their pharmacokinetic properties.

The quintessential feature of 5-PPP-independent RIG-I agonists is a non-canonical RNA structure that forms an interhelical bend near the terminal end of the RNA du-plex. Here, we provide evidence that dsRNAs with a termi-nal interhelical bend are recognized by RIG-I in eukaryotic cells. Most cytosolic RNAs––including ribosomal RNAs

(12)

(rRNA), transfer RNAs (tRNA) and precursor microR-NAs (pre-miRNA)––do not conform to the structural qual-ifications we have discovered for 5-PPP-independent RIG-I agonists. For example, pre-miRNAs and tRNAs have two- and four-nucleotide 3-overhangs, respectively. These should block RIG-I activation even in the presence of a terminal interhelical bend. The 3D structures of rRNAs are finely-tuned to interact with ribosomal proteins for the hierarchical assembly of the ribosome (45), which makes rRNAs unlikely targets for recognition by RIG-I. Our find-ings thus highlight the unique ability of RIG-I to discern the structural features of pathogen-derived RNAs from the variety of endogenous host RNA structures.

SUPPLEMENTARY DATA

Supplementary Dataare available at NAR Online.

ACKNOWLEDGEMENTS

We would like to thank the Korea Basic Science Institute for access to its NMR facility. We would like to thank Prof. Christiane Bruns at University Hospital Magdeburg in Germany for generously providing the panc02 cell line. We would like to thank Prof. Eui-Cheol Shin at KAIST for kindly providing the Huh 7 and Huh 7.5 cell lines.

FUNDING

National Research Foundation of Korea [2015R1A2A2 A01005688, 2015R1A2A1A15055329 to B.-S.C., 2017R1 A2B4007817 to S.-J.K.]; Korea Drug Development Fund funded by MSIT, MOTIE, and MOHW [KDDF-201706-06 to B.-S. C.]; KRICT [KK1703-E00 to M.K.]. Funding for open access charge: Korea Drug Development Fund [KDDF-201706-06 to B.-S. C.].

Conflict of interest statement. J.L., E.-B.P., J.M., S.-E.S., M.-K.L., S.-J.K. and B.-S.C. have filed patent applications re-lated to the findings presented in this paper. The other au-thors declare no competing financial interests.

REFERENCES

1. Rawling,D.C., Fitzgerald,M.E. and Pyle,A.M. (2015) Establishing the role of ATP for the function of the RIG-I innate immune sensor.

eLife, 4, e09391.

2. Civril,F., Bennett,M., Moldt,M., Deimling,T., Witte,G., Schiesser,S., Carell,T. and Hopfner,K.-P. (2011) The RIG-I ATPase domain structure reveals insights into ATP-dependent antiviral signalling.

EMBO Rep., 12, 1127–1134.

3. L¨assig,C., Matheisl,S., Sparrer,K.M.J., de Oliveira Mann,C.C., Moldt,M., Patel,J.R., Goldeck,M., Hartmann,G., Garc´ıa-Sastre,A., Hornung,V. et al. (2015) ATP hydrolysis by the viral RNA sensor RIG-I prevents unintentional recognition of self-RNA. eLife, 4, e10859.

4. Yoneyama,M., Kikuchi,M., Natsukawa,T., Shinobu,N., Imaizumi,T., Miyagishi,M., Taira,K., Akira,S. and Fujita,T. (2004) The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nat. Immunol., 5, 730–737. 5. Xu,L.-G., Wang,Y.-Y., Han,K.-J., Li,L.-Y., Zhai,Z. and Shu,H.-B.

(2005) VISA is an adapter protein required for virus-triggered IFN-␤ signaling. Mol. Cell, 19, 727–740.

6. Besch,R., Poeck,H., Hohenauer,T., Senft,D., H¨acker,G., Berking,C., Hornung,V., Endres,S., Ruzicka,T. et al. (2009) Proapoptotic signaling induced by RIG-I and MDA-5 results in type I

interferon–independent apoptosis in human melanoma cells. J. Clin.

Invest., 119, 2399–2411.

7. Matsushima-Miyagi,T., Hatano,K., Nomura,M., Li-Wen,L., Nishikawa,T., Saga,K., Shimbo,T. and Kaneda,Y. (2012) TRAIL and Noxa are selectively upregulated in prostate cancer cells downstream of the RIG-I/MAVS signaling pathway by nonreplicating Sendai virus particles. Clin. Cancer Res., 18, 6271.

8. Duewell,P., Steger,A., Lohr,H., Bourhis,H., Hoelz,H.,

Kirchleitner,S.V., Stieg,M.R., Grassmann,S., Kobold,S., Siveke,J.T.

et al. (2014) RIG-I-like helicases induce immunogenic cell death of

pancreatic cancer cells and sensitize tumors toward killing by CD8+ T cells. Cell Death Differ., 21, 1825–1837.

9. K ¨ubler,K., tho Pesch,C., Gehrke,N., Riemann,S., Daßler,J., Coch,C., Landsberg,J., Wimmenauer,V., P ¨olcher,M., Rudlowski,C. et al. (2011) Immunogenic cell death of human ovarian cancer cells induced by cytosolic poly(I:C) leads to myeloid cell maturation and activates NK cells. Eur. J. Immunol., 41, 3028–3039.

10. Beljanski,V., Chiang,C., Kirchenbaum,G.A., Olagnier,D.,

Bloom,C.E., Wong,T., Haddad,E.K., Trautmann,L., Ross,T.M. and Hiscott,J. (2015) Enhanced influenza virus-like particle vaccination with a structurally optimized RIG-I agonist as adjuvant. J. Virol., 89, 10612–10624.

11. Hornung,V., Ellegast,J., Kim,S., Brz ´ozka,K., Jung,A., Kato,H., Poeck,H., Akira,S., Conzelmann,K.-K., Schlee,M. et al. (2006) 5-Triphosphate RNA Is the ligand for RIG-I. Science, 314, 994–997. 12. Goubau,D., Schlee,M., Deddouche,S., Pruijssers,A.J., Zillinger,T.,

Goldeck,M., Schuberth,C., Van der Veen,A.G., Fujimura,T., Rehwinkel,J. et al. (2014) Antiviral immunity via RIG-I-mediated recognition of RNA bearing 5-diphosphates. Nature, 514, 372–375. 13. Devarkar,S.C., Wang,C., Miller,M.T., Ramanathan,A., Jiang,F.,

Khan,A.G., Patel,S.S. and Marcotrigiano,J. (2016) Structural basis for m7G recognition and 2-O-methyl discrimination in capped RNAs by the innate immune receptor RIG-I. Proc. Natl. Acad. Sci.

U.S.A., 113, 596–601.

14. Schuberth-Wagner,C., Ludwig,J., Bruder,Ann K., Herzner,A.-M., Zillinger,T., Goldeck,M., Schmidt,T., Schmid-Burgk,Jonathan L., Kerber,R., Wolter,S. et al. (2015) A conserved histidine in the RNA sensor RIG-I controls immune tolerance to N1-2O-methylated self RNA. Immunity, 43, 41–51.

15. Kato,H., Takeuchi,O., Mikamo-Satoh,E., Hirai,R., Kawai,T., Matsushita,K., Hiiragi,A., Dermody,T.S., Fujita,T. and Akira,S. (2008) Length-dependent recognition of double-stranded ribonucleic acids by retinoic acid–inducible gene-I and melanoma

differentiation–associated gene 5. J. Exp. Med., 205, 1601–1610. 16. Chiang,C., Beljanski,V., Yin,K., Olagnier,D., Ben Yebdri,F., Steel,C.,

Goulet,M.-L., DeFilippis,V.R., Streblow,D.N., Haddad,E.K. et al. (2015) Sequence-specific modifications enhance the broad spectrum antiviral response activated by RIG-I agonists. J. Virol., 89, 8011–8025.

17. Schnell,G., Loo,Y.-M., Marcotrigiano,J. and Gale,M. Jr (2012) Uridine composition of the poly-U/UC tract of HCV RNA defines non-self recognition by RIG-I. PLoS Pathog., 8, e1002839. 18. Lee,M.-K., Kim,H.-E., Park,E.-B., Lee,J., Kim,K.-H., Lim,K.,

Yum,S., Lee,Y.-H., Kang,S.-J., Lee,J.-H. et al. (2016) Structural features of influenza A virus panhandle RNA enabling the activation of RIG-I independently of 5-triphosphate. Nucleic Acids Res., 44, 8407–8416.

19. Bae,S.-H., Cheong,H.-K., Lee,J.-H., Cheong,C., Kainosho,M. and Choi,B.-S. (2001) Structural features of an influenza virus promoter and their implications for viral RNA synthesis. Proc. Natl. Acad. Sci.

U.S.A., 98, 10602–10607.

20. Kim,M., Kim,S.-Y., Lee,H.W., Shin,J.S., Kim,P., Jung,Y.-S., Jeong,H.-S., Hyun,J.-K. and Lee,C.-K. (2013) Inhibition of influenza virus internalization by (−)-epigallocatechin-3-gallate. Antiviral Res.,

100, 460–472.

21. Manicassamy,B., Manicassamy,S., Belicha-Villanueva,A.,

Pisanelli,G., Pulendran,B. and Garcia-Sastre,A. (2010) Analysis of in vivo dynamics of influenza virus infection in mice using a GFP reporter virus. Proc. Natl. Acad. Sci. U.S.A., 107, 11531–11536. 22. Tran,T.T., Kim,M., Jang,Y., Lee,H.W., Nguyen,H.T., Nguyen,T.N.,

Park,H.W., Le Dang,Q. and Kim,J.C. (2017) Characterization and mechanisms of anti-influenza virus metabolites isolated from the Vietnamese medicinal plant Polygonum chinense. BMC Complement

(13)

23. Delaglio,F., Grzesiek,S., Vuister,G.W., Zhu,G., Pfeifer,J. and Bax,A. (1995) NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J. Biomol. NMR, 6, 277–293.

24. Goddard,T.D. and Kneller,D.G. (2008) SPARKY 3. University of California, San Francisco, https://www.cgl.ucsf.edu/home/sparky/. 25. Hwang,S.D., Shin,J.S., Ku,K.B., Kim,H.S., Cho,S.W. and Seo,S.H.

(2010) Protection of pregnant mice, fetuses and neonates from lethality of H5N1 influenza viruses by maternal vaccination. Vaccine,

28, 2957–2964.

26. Poole,E., He,B., Lamb,R.A., Randall,R.E. and Goodbourn,S. (2002) The V Proteins of Simian Virus 5 and Other Paramyxoviruses Inhibit Induction of Interferon-␤. Virology, 303, 33–46.

27. Varani,G. (1995) Exceptionally Stable Nucleic Acid Hairpins. Annu.

Rev. Biophys. Biomol. Struct., 24, 379–404.

28. Cheong,H.K., Cheong,C., Lee,Y.S., Seong,B.L. and Choi,B.S. (1999) Structure of influenza virus panhandle RNA studied by NMR spectroscopy and molecular modeling. Nucleic Acids Res., 27, 1392–1397.

29. McPhee,S.A., Huang,L. and Lilley,D.M.J. (2014) A critical base pair in k-turns that confers folding characteristics and correlates with biological function. Nat. Commun., 5, 5127.

30. Rosen,C.A., Sodroski,J.G. and Haseltine,W.A. (1985) The location of cis-acting regulatory sequences in the human T cell lymphotropic virus type III (HTLV-III/LAV) long terminal repeat. Cell, 41, 813–823.

31. Aboul-ela,F., Karn,J. and Varani,G. (1996) Structure of HIV-1 TAR RNA in the absence of ligands reveals a novel conformation of the trinucleotide bulge. Nucleic Acids Res., 24, 3974–3981.

32. Casiano-Negroni,A., Sun,X. and Al-Hashimi,H.M. (2007) Probing Na+-induced changes in the HIV-1 TAR conformational dynamics using NMR residual dipolar couplings: new insights into the role of counterions and electrostatic interactions in adaptive recognition.

Biochemistry, 46, 6525–6535.

33. Bailor,M.H., Sun,X. and Al-Hashimi,H.M. (2010) Topology links RNA secondary structure with global conformation, dynamics, and adaptation. Science, 327, 202–206.

34. Wang,Y., Ludwig,J., Schuberth,C., Goldeck,M., Schlee,M., Li,H., Juranek,S., Sheng,G., Micura,R., Tuschl,T. et al. (2010) Structural

and functional insights into 5-ppp RNA pattern recognition by the innate immune receptor RIG-I. Nat. Struct. Mol. Biol., 17, 781–787. 35. Luo,D., Ding,S.C., Vela,A., Kohlway,A., Lindenbach,B.D. and

Pyle,A.M. (2011) Structural Insights into RNA Recognition by RIG-I. Cell, 147, 409–422.

36. Jiang,F., Ramanathan,A., Miller,M.T., Tang,G.-Q., Gale,M., Patel,S.S. and Marcotrigiano,J. (2011) Structural basis of RNA recognition and activation by innate immune receptor RIG-I. Nature,

479, 423.

37. Kowalinski,E., Lunardi,T., McCarthy,A.A., Louber,J., Brunel,J., Grigorov,B., Gerlier,D. and Cusack,S. (2011) Structural basis for the activation of innate immune pattern-recognition receptor RIG-I by viral RNA. Cell, 147, 423–435.

38. Cui,S., Eisen¨acher,K., Kirchhofer,A., Brz ´ozka,K., Lammens,A., Lammens,K., Fujita,T., Conzelmann,K.-K., Krug,A. and

Hopfner,K.-P. (2008) The C-terminal regulatory domain is the RNA 5-triphosphate sensor of RIG-I. Mol. Cell, 29, 169–179.

39. Lu,C., Xu,H., Ranjith-Kumar,C.T., Brooks,M.T., Hou,T.Y., Hu,F., Herr,A.B., Strong,R.K., Kao,C.C. and Li,P. (2010) The structural basis of 5triphosphate double-stranded RNA recognition by RIG-I C-terminal domain. Structure, 18, 1032–1043.

40. Liu,L., Botos,I., Wang,Y., Leonard,J.N., Shiloach,J., Segal,D.M. and Davies,D.R. (2008) Structural basis of Toll-like receptor 3 signaling with double-stranded RNA. Science, 320, 379.

41. Alexopoulou,L., Holt,A.C., Medzhitov,R. and Flavell,R.A. (2001) Recognition of double-stranded RNA and activation of NF-[kappa]B by Toll-like receptor 3. Nature, 413, 732–738.

42. Eckstein,F. (2014) Phosphorothioates, essential components of therapeutic oligonucleotides. Nucleic Acid Ther., 24, 374–387. 43. Dowdy,S.F. (2017) Overcoming cellular barriers for RNA

therapeutics. Nat Biotech., 35, 222–229.

44. Schlee,M., Roth,A., Hornung,V., Hagmann,C.A., Wimmenauer,V., Barchet,W., Coch,C., Janke,M., Mihailovic,A., Wardle,G. et al. (2009) Recognition of 5triphosphate by RIG-I helicase requires short blunt double-stranded RNA as contained in panhandle of negative-strand virus. Immunity, 31, 25–34.

45. Fromont-Racine,M., Senger,B., Saveanu,C. and Fasiolo,F. (2003) Ribosome assembly in eukaryotes. Gene, 313, 17–42.

수치

Figure 1. Bent RNAs induce RIG-I-dependent ISG56 expression. (A) Secondary structures of the RNA constructs used in the experiments
Figure 2. RIG-I-dependent immunogenicity of the bulged RNA. (A) Secondary structures of the bulged RNA (green; Kinked RNA and TAR-Bulge)
Figure 3. Backbone modification and elongation of the bent RNA do not deteriorate the immunogenicity
Figure 4. Antiviral activities of CBS-7 and CBS-13-BPS in vitro and in vivo. (A) Antiviral activities of CBS-7 and CBS-13-BPS monitored by western
+3

참조

관련 문서

In the present study we investigated the proliferation effect of human oral cancer KB cell treated with pulsatilla koreana extract.. We analyzed the effects of this

The aim of this study was to examine the effects and mechanisms of the new metformin derivative HL156A in human multidrug-resistant cancer cell lines (FADU/PTX,

The PIG3 was interacted with the CTSB in human prostate cancer cell line PC3, which was verified by co-immunoprecipitation analysis using specific

Inhibitory effects of classified methanol extracts of Gastrodia elata Blume on the COX-2 and iNOS expression of human colorectal cancer cell lines...

In this study, the expression profiles of miRNAs were compared and analyzed for establishment of miRNAs related cancer cell growth inhibition in normal human oral

1) Effects of methanol extracts of Capsella bursa-pastoris on cyclooxygenase-2 (COX-2) and Inducible Nitric oxide synthase (iNOS) expression in human prostate cancer cell

Inhibitory effects of classified methanol extracts of Smilax china L on the COX-2 and iNOS expression of human colorectal cancer cell lines... Inhibitory

Results: The human breast cancer cell subline MCF-7/MX5 cells selected in the presence of 5 µg/ml mitoxantrone (MX) were more resistant to MX (15.7... Western blot and