Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

13
JOURNAL OF VIROLOGY, Nov. 2010, p. 11542–11554 Vol. 84, No. 21 0022-538X/10/$12.00 doi:10.1128/JVI.00595-10 Copyright © 2010, American Society for Microbiology. All Rights Reserved. Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex Molecules Esther Schnettler,# Hans Hemmes,§# Rik Huismann, Rob Goldbach,† Marcel Prins,¶ and Richard Kormelink* Laboratory of Virology, Wageningen University, Droevendaalsesteeg 1, 6708 PB Wageningen, Netherlands Received 18 March 2010/Accepted 17 August 2010 The tospovirus NSs protein was previously shown to suppress the antiviral RNA silencing mechanism in plants. Here the biochemical analysis of NSs proteins from different tospoviruses, using purified NSs or NSs containing cell extracts, is described. The results showed that all tospoviral NSs proteins analyzed exhibited affinity to small double-stranded RNA molecules, i.e., small interfering RNAs (siRNAs) and micro-RNA (miRNA)/miRNA* duplexes. Interestingly, the NSs proteins from tomato spotted wilt virus (TSWV), impatiens necrotic spot virus (INSV), and groundnut ringspot virus (GRSV) also showed affinity to long double-stranded RNA (dsRNA), whereas tomato yellow ring virus (TYRV) NSs did not. The TSWV NSs protein was shown to be capable of inhibiting Dicer-mediated cleavage of long dsRNA in vitro. In addition, it suppressed the accumulation of green fluorescent protein (GFP)-specific siRNAs during coinfiltration with an inverted-repeat-GFP RNA construct in Nicotiana benthamiana. In vivo interference of TSWV NSs in the miRNA pathway was shown by suppression of an enhanced GFP (eGFP) miRNA sensor construct. The ability to stabilize miRNA/miRNA* by different tospovirus NSs proteins in vivo was demonstrated by increased accumulation and detection of both miRNA171c and miRNA171c* in tospo- virus-infected N. benthamiana. All together, these data suggest that tospoviruses interfere in the RNA silencing pathway by sequestering siRNA and miRNA/miRNA* molecules before they are uploaded into their respective RNA-induced silencing complexes. The observed affinity to long dsRNA for only a subset of the tospoviruses studied is discussed in light of evolutional divergence and their ancestral relation to the animal-infecting members of the Bunyaviridae. In recent years, RNA silencing has become known as one of the major defense mechanisms acting against viruses in plants and insects (15, 20, 21, 54). During a virus infection, double- stranded RNA (dsRNA) molecules arise as replicative inter- mediates or due to the formation of secondary RNA struc- tures. These are recognized by specific Dicer-like proteins (Dicer in insects and predominantly DCL-4 in plants) and processed into 21-nucleotide (nt) small interfering RNA (siRNA) molecules (6, 14, 27). One strand of this molecule, the guide strand, is incorporated in the RNA-induced silencing complex (RISC) and enables RISC to recognize and degrade complementary (viral) target RNA molecules through the ac- tion of the core Argonaute (AGO) protein (28, 44). In plants, these 21-nt primary siRNAs have been shown to serve as prim- ers for the host-encoded RNA-dependent RNA polymerase (RDR) to convert RNA target sequences into new long dsRNAs. These in turn are processed into secondary siRNAs. In this way, silencing not only is being amplified but also is spread along the entire RNA target sequence (transitive si- lencing) (53). Two other plant DCL proteins (DCL-2 and DCL-3) also play a role in processing of dsRNA, but these generally lead to 22- to 26-nt siRNAs which are suggested to function in a range of other purposes unrelated to viral defense (1). DCL-1 pro- duces micro-RNAs (miRNAs) that are structurally similar to siRNA molecules but originate after processing of long host-carried RNA transcripts called primary miRNA (pri- miRNA) via precursor miRNA (pre-miRNA) into miRNA/ miRNA* duplexes. After unwinding of miRNA/miRNA* du- plexes, one strand is incorporated into RISC, whereas the other strand, miRNA*, is rapidly degraded after cleavage by an AGO protein (56). MicroRNAs are regulatory factors that are loaded into RISC to silence host-carried genes by either RNA degrada- tion or translational inhibition of their target (55). In insects, Dicer-2 is required for the siRNA and Dicer-1 for the miRNA processing steps that are divided over multiple DCLs in plants (1). As a response to antiviral RNA silencing, many plant and insect viruses have been shown to express RNA silencing sup- pressor (RSS) proteins (2, 40). Most of these RSS proteins specifically bind either siRNAs or long dsRNA (14, 37, 42), whereas some RSS proteins, like the poleroviral p0 (8, 47, 59), interact with key proteins of the RNA silencing pathway, such as DCL or AGO. A few RSS proteins are able to interact at multiple points in the RNA silencing pathway. One such example is that of cucumber mosaic virus 2b, which not only binds RNA molecules but also interacts with AGO (24, 59). Besides their interference in the antiviral siRNA pathway, viral RSS proteins also effect the miRNA pathway, likely by sequestering miRNA/ * Corresponding author. Mailing address: Wageningen University, Laboratory of Virology, Droevendaalsesteeg 1, 6708 PB Wageningen, Netherlands. Phone: 31 317 484498. Fax: 31 317 484820. E-mail: [email protected]. § Present address: Rockefeller University, Laboratory of Plant Mo- lecular Biology, 1230 York Avenue, New York, NY 10065. ¶ Present address: Keygene N.V., Agro Business Park 90, 6708 PW Wageningen, Netherlands. # Both authors contributed equally to the manuscript. † R.G. died in an accident on 7 April 2009. Published ahead of print on 25 August 2010. 11542 Downloaded from https://journals.asm.org/journal/jvi on 02 December 2021 by 177.44.91.107.

Transcript of Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

Page 1: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

JOURNAL OF VIROLOGY, Nov. 2010, p. 11542–11554 Vol. 84, No. 210022-538X/10/$12.00 doi:10.1128/JVI.00595-10Copyright © 2010, American Society for Microbiology. All Rights Reserved.

Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins,NSs, for Various RNA Duplex Molecules�

Esther Schnettler,# Hans Hemmes,§# Rik Huismann, Rob Goldbach,†Marcel Prins,¶ and Richard Kormelink*

Laboratory of Virology, Wageningen University, Droevendaalsesteeg 1, 6708 PB Wageningen, Netherlands

Received 18 March 2010/Accepted 17 August 2010

The tospovirus NSs protein was previously shown to suppress the antiviral RNA silencing mechanismin plants. Here the biochemical analysis of NSs proteins from different tospoviruses, using purified NSsor NSs containing cell extracts, is described. The results showed that all tospoviral NSs proteins analyzedexhibited affinity to small double-stranded RNA molecules, i.e., small interfering RNAs (siRNAs) andmicro-RNA (miRNA)/miRNA* duplexes. Interestingly, the NSs proteins from tomato spotted wilt virus(TSWV), impatiens necrotic spot virus (INSV), and groundnut ringspot virus (GRSV) also showed affinityto long double-stranded RNA (dsRNA), whereas tomato yellow ring virus (TYRV) NSs did not. The TSWVNSs protein was shown to be capable of inhibiting Dicer-mediated cleavage of long dsRNA in vitro. Inaddition, it suppressed the accumulation of green fluorescent protein (GFP)-specific siRNAs duringcoinfiltration with an inverted-repeat-GFP RNA construct in Nicotiana benthamiana. In vivo interferenceof TSWV NSs in the miRNA pathway was shown by suppression of an enhanced GFP (eGFP) miRNAsensor construct. The ability to stabilize miRNA/miRNA* by different tospovirus NSs proteins in vivo wasdemonstrated by increased accumulation and detection of both miRNA171c and miRNA171c* in tospo-virus-infected N. benthamiana. All together, these data suggest that tospoviruses interfere in the RNAsilencing pathway by sequestering siRNA and miRNA/miRNA* molecules before they are uploaded intotheir respective RNA-induced silencing complexes. The observed affinity to long dsRNA for only a subsetof the tospoviruses studied is discussed in light of evolutional divergence and their ancestral relation tothe animal-infecting members of the Bunyaviridae.

In recent years, RNA silencing has become known as one ofthe major defense mechanisms acting against viruses in plantsand insects (15, 20, 21, 54). During a virus infection, double-stranded RNA (dsRNA) molecules arise as replicative inter-mediates or due to the formation of secondary RNA struc-tures. These are recognized by specific Dicer-like proteins(Dicer in insects and predominantly DCL-4 in plants) andprocessed into 21-nucleotide (nt) small interfering RNA(siRNA) molecules (6, 14, 27). One strand of this molecule, theguide strand, is incorporated in the RNA-induced silencingcomplex (RISC) and enables RISC to recognize and degradecomplementary (viral) target RNA molecules through the ac-tion of the core Argonaute (AGO) protein (28, 44). In plants,these 21-nt primary siRNAs have been shown to serve as prim-ers for the host-encoded RNA-dependent RNA polymerase(RDR) to convert RNA target sequences into new longdsRNAs. These in turn are processed into secondary siRNAs.In this way, silencing not only is being amplified but also is

spread along the entire RNA target sequence (transitive si-lencing) (53).

Two other plant DCL proteins (DCL-2 and DCL-3) alsoplay a role in processing of dsRNA, but these generally lead to22- to 26-nt siRNAs which are suggested to function in a rangeof other purposes unrelated to viral defense (1). DCL-1 pro-duces micro-RNAs (miRNAs) that are structurally similar tosiRNA molecules but originate after processing of longhost-carried RNA transcripts called primary miRNA (pri-miRNA) via precursor miRNA (pre-miRNA) into miRNA/miRNA* duplexes. After unwinding of miRNA/miRNA* du-plexes, one strand is incorporated into RISC, whereas the otherstrand, miRNA*, is rapidly degraded after cleavage by an AGOprotein (56). MicroRNAs are regulatory factors that are loadedinto RISC to silence host-carried genes by either RNA degrada-tion or translational inhibition of their target (55). In insects,Dicer-2 is required for the siRNA and Dicer-1 for the miRNAprocessing steps that are divided over multiple DCLs in plants (1).

As a response to antiviral RNA silencing, many plant andinsect viruses have been shown to express RNA silencing sup-pressor (RSS) proteins (2, 40). Most of these RSS proteinsspecifically bind either siRNAs or long dsRNA (14, 37, 42),whereas some RSS proteins, like the poleroviral p0 (8, 47, 59),interact with key proteins of the RNA silencing pathway, suchas DCL or AGO. A few RSS proteins are able to interact atmultiple points in the RNA silencing pathway. One such exampleis that of cucumber mosaic virus 2b, which not only binds RNAmolecules but also interacts with AGO (24, 59). Besides theirinterference in the antiviral siRNA pathway, viral RSS proteinsalso effect the miRNA pathway, likely by sequestering miRNA/

* Corresponding author. Mailing address: Wageningen University,Laboratory of Virology, Droevendaalsesteeg 1, 6708 PB Wageningen,Netherlands. Phone: 31 317 484498. Fax: 31 317 484820. E-mail:[email protected].

§ Present address: Rockefeller University, Laboratory of Plant Mo-lecular Biology, 1230 York Avenue, New York, NY 10065.

¶ Present address: Keygene N.V., Agro Business Park 90, 6708 PWWageningen, Netherlands.

# Both authors contributed equally to the manuscript.† R.G. died in an accident on 7 April 2009.� Published ahead of print on 25 August 2010.

11542

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 2: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

miRNA* duplexes, and thereby cause developmental defects inArabidopsis after transgenic expression (11, 16).

Tomato spotted wilt virus (TSWV) is the type species of theTospovirus genus within the family of arthropod-borne Bunya-viridae. In contrast to all other mammal-infecting members ofthe Bunyaviridae, tospoviruses specifically infect plants and aretransmitted in a propagative manner by thrips (19, 57). Tospo-viruses are therefore a likely target of antiviral RNA silencingin both plant and insect hosts. Previously, the TSWV NSsprotein has been shown to suppress RNA silencing in plantsand insects (10, 23, 48, 51). Accumulation of this protein, atleast in plants, coincides with increased virulence of the virus(36). It is hypothesized that the RSS mode of action of tospo-virus NSs proteins is accomplished by interacting with a com-ponent of the antiviral RNA silencing pathway that is sharedbetween plants and insects, i.e., dsRNA. To test this hypothe-sis, the affinity of TSWV NSs to a range of dsRNA moleculesof the si- and miRNA pathway was analyzed in vitro and ver-ified by reporter-based assays in vivo. The analysis was broad-ened to other tospoviral NSs proteins in order to determine ifthe observed mode of action is a general feature for tospoviralNSs proteins.

MATERIALS AND METHODS

Plasmid constructs. Agrobacterium expression plasmids for maltose bindingprotein (MBP), TSWV NSs, inverted repeat of GFP (IR-GFP), and GFP weredescribed previously (10, 30, 42).

The coding sequence for tomato yellow ring virus tomato strain (TYRV-t) NSsand groundnut ringspot virus (GRSV) NSs was PCR amplified and cloned intothe binary pK2GW7 (32) vector using Gateway technology. Baculovirus express-ing TSWV NSs or GFP has been described previously (31, 36). Recombinantbaculoviruses expressing GRSV NSs and TYRV NSs were constructed usingGateway technology and the Bac-to-Bac system (Invitrogen), following the man-ufacturers’ protocol.

The insect expression vectors of MBP and TSWV NSs were constructed usingGateway technology in pIB-GW (Invitrogen).

The 3� untranslated region (3�UTR) of the par6 gene was PCR amplified froman existing insect expression plasmid (18) to provide the restriction sites SstII andXbaI for convenient cloning and 3� fusion to the enhanced GFP (eGFP) codingsequence. This reporter construct, denoted eGFP-3�UTR, was subsequentlycloned into the binary expression vector pK2GW7 (32), using Gateway technol-ogy. In a similar fashion, the par6 gene 3�UTR was cloned behind the fireflyluciferase coding sequence (Fluc-miRNA1) in pMT-Fluc (54).

The binary vector expressing primary miRNA1 was made by PCR amplifica-tion of pri-miRNA1 from a previously described insect expression vector (18)followed by Gateway technology recombination into the binary pK2GW7 vector(32).

Inducible expression vectors carrying pri-miRNA1 or pri-miR12 were con-structed by excision of the respective fragments by either NotI-XbaI or KpnI-XbaI from the vectors tub-miRNA1 and pAc-miR12 (18) and subsequent cloninginto pMT-B (Invitrogen).

Cell culture and transfection. Schneider 2 (S2) cells were grown in Schneidermedium (Invitrogen) supplemented with 10% heat-inactivated fetal calf serum(FCS) (Gibco) at 28°C. To reach a confluence of 60 to 70% at the time oftransfection, cells were seeded 24 h prior to transfection in a 96-well plate at aconcentration of 5 � 104 cells per well. Transfections were performed usingCellfectine II (Invitrogen) reagent according to the manufacturer’s instructions.Cells were cotransfected with 100 ng RSS expression plasmid (MBP or TSWVNSs), 12.5 ng pMT-Fluc-miRNA1, 3 ng pMT-Rluc, and 2.5 ng pMT-miRNA,either pri-miRNA1 or pri-miRNA12. Expression of the inducible constructs wasinduced 48 h posttransfection by 5 �M CuSO4 and assayed 24 h postinduction.Luciferase expression was determined using self-made buffers for the dual-luciferase reporter assay (17).

Bacterial expression and purification of thioredoxin-TSWV NSs and -MBP.The coding sequence of TSWV NSs or MBP was PCR amplified to introduceGateway specific recombination sites and cloned into pDONR207 (Invitrogen).For expression and subsequent purification, the NSs or MBP coding sequence

was cloned in frame with His-patched thioredoxin (HP-thioredoxin) in pDest49-BAD (Invitrogen) by Gateway reaction. Proteins were expressed in DH10betacells (Qiagen) according to the manufacturer’s recommendations. After induc-tion for 6 h at 37°C with 0.2% (wt/vol) L-arabinose, cells were harvested bycentrifugation for 15 min at 4,000 rpm (Sorvall GSA rotor) at 4°C. Cells werelysed by sonication on ice three times with 30-s intervals in lysis buffer (50 mMK2PO4, 400 mM NaCl, 100 mM KCl, 10% [vol/vol] glycerol, 1% [vol/vol] TritonX-100 and EDTA-free protease inhibitor cocktail [Roche]). The soluble fractionwas recovered by centrifugation at 4,000 rpm (Sorvall GSA rotor) for 30 min at4°C. Recombinant protein was purified using a Talon metal affinity resin column(Clontech) and eluted with 2.5 packed bed volumes (PBV) elution buffer (50 mMNaH2PO4, 300 mM NaCl, 200 mM imidazole, 10% [vol/vol] glycerol) afterwashing with 15-PBV lysis buffer. Protein fractions were instantly frozen inaliquots in liquid nitrogen and stored at �80°C until use. Protein concentrationsof eluted fractions were determined using the standard procedure of the Bio-Radprotein assay according to the manufacturer’s recommendations, and the puri-fication process was analyzed by SDS-PAGE and subsequent staining with Coo-massie brilliant blue.

Preparation of virus-infected plant extracts and baculovirus-infected insectcell extracts. Groundnut ringspot virus (GRSV), impatiens necrotic spot virus(INSV), tomato yellow ring virus tomato strain (TYRV-t), tomato spotted wiltvirus (TSWV), and cymbidium ringspot virus (CymRSV) were mechanicallyinoculated on Nicotiana benthamiana, and extracts were prepared from sys-temically infected leaves essentially as described previously (42) with minormodifications. Virus accumulation was verified either by enzyme-linked im-munosorbent assay (ELISA) or NSs-specific Western blot analysis prior to prep-aration of virus-infected plant extracts. The total protein concentration wasdetermined using the standard procedure of the Bio-Rad protein assay accordingto the manufacturer’s recommendations. Crude extracts were centrifuged twiceat 14,000 rpm for 15 min at 4°C. Extracts were immediately frozen in liquidnitrogen and stored at �80°C until use.

TSWV NSs, GRSV NSs, TYRV NSs, and GFP were expressed in Trichoplusiani Hi5 cells, using baculovirus expression vectors expressing the genes under thecontrol of the polyhedrin promoter (29).

Hi5 cells were infected with baculoviruses at a multiplicity of infection (MOI)of 10 and incubated for 48 h at 28°C. Cells were detached, harvested by low-speed centrifugation (1,500 rpm), and washed with phosphate-buffered saline(PBS) prior to lysis by sonication during three intervals of 30 s in lysis buffer (100mM NaCl, 20 mM Tris 7.4, 2 mM MgCl2, 1 mM dithiothreitol [DTT], 10%[vol/vol] glycerol). The infection was monitored either by GFP fluorescence orSDS-PAGE and Western immunoblot analysis for TSWV NSs, GRSV NSs, andTYRV NSs. The total protein concentration was determined by the Bio-Radprotein assay according to the manufacturer’s protocol.

Expression analysis. Expression of different NSs proteins was monitored byWestern immunoblot analysis. Samples of the extracts were mixed with 2� SDSloading buffer, incubated for 5 min at 95°C, and centrifuged for 3 min at 14,000rpm. Proteins were separated by SDS-PAGE and transferred to an Immobilion-P(Millipore) membrane by semidry blotting. TSWV and GRSV NSs proteins weredetected using an NSs-specific polyclonal antibody. The TYRV NSs protein wasdetected using a monoclonal antibody (kindly provided by S.-D. Yeh). Pro-tein-antibody complexes were detected by an alkaline phosphatase-conju-gated secondary antibody and visualized with nitroblue tetrazolium–5-bromo-4-chloro-3-indolyl- phosphate (NBT-BCIP) as a substrate (Roche) accordingto the manufacturer’s recommendations.

dsRNA preparation. A 114-nt dsRNA molecule was generated by T7 RNApolymerase (Promega) transcription on a gel-purified (High Pure PCR purifica-tion kit; Roche) eGFP template in the presence of [�-32P]CTP (Perkin Elmer).The template was provided with T7 RNA polymerase promoter sequences atboth ends by PCR amplification using the DNA oligonucleotides T7_dsRNA114F (5� GTA ATA CGA CTC ACT ATA GGG GGC GTG CAG TGC TTC AGCCGC 3�) and T7_ds114 R (5� GTA ATA CGA CTC ACT ATA GGG GCC GTCGTC CTT GAA GAA GAT GG 3�). Precursor miRNA 2b was generated by T7RNA polymerase transcription in the presence of [�-32P]CTP (Perkin Elmer) ona template obtained after annealing of two long primers: 5� GTA ATA CGACTC ACT ATA GGC GTT GCG AGG AGT TTC GAC CGA CAC TAT ACTTAT AAC AAC TGT TGT ACA GTG ACG GTG AAA CTT CTG TCA ACTTC 3� and 5� GAA GTT GAC AGA AGT TTC ACC GTC ACT GTA CAACAG TTG TTA TAA GTA TAG TGT CGG TCG AAA CTC CTC GCA ACGCCT ATA GTG AGT CGT ATT AC 3�. Following T7 transcription, reactionmixtures were incubated at 70°C for 10 min and cooled down to room temper-ature. Template DNA was removed by treatment with DNase I, and dsRNA wasgel purified from an 8% PAGE, 0.5� Tris-borate-EDTA (TBE) native gel.Labeling of custom-made RNA oligonucleotides targeting the GFP sequence or

VOL. 84, 2010 TOSPOVIRUS NSs EXHIBITS dsRNA AFFINITY 11543

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 3: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

corresponding to the Arabidopsis thaliana micro-RNA 171a sequence was per-formed by end labeling of the GFP siRNA guide strand or miRNA171a strandusing [�-32P]ATP (Perkin Elmer) and T4 polynucleotide kinase. These radiola-beled strands were annealed to the RNA oligonucleotide corresponding to therespective GFP siRNA passenger or miRNA171* a strand and PAGE purifiedessentially as described previously (25).

Electrophoretic mobility shift assay and Western blot analysis. In a bindingreaction, radiolabeled RNA (0.5 nM) was incubated with �2 �g total proteinfrom virus-infected leaf or cell extracts per 10 �l reaction mixture and incubatedfor 20 min at room temperature as previously described (30, 42). As controls,RNA was loaded without plant extracts, with healthy plant extracts, or withGFP-expressing baculovirus-infected insect cell extracts. The same reaction wascarried out with serial dilutions of the bacterium-expressed HP-thioredoxin-NSsor -MBP proteins. The complexes were separated on a 0.5� TBE native PAGEgel. For 114-nt dsRNA and pre-miRNA 2b, a 5% gel was used, and an 8% gelwas used for siRNA and miRNA/miRNA* molecules. Following electrophoresis,gels were dried, exposed overnight to a phosphor screen, and scanned (Molec-ular Dynamics Typhoon PhosphorImager; Amersham Biosciences). A represen-tative picture of at least three independent experiments was shown.

To determine the presence of NSs in the RNA-protein complex, the excisedgel slices were ground in 2� SDS loading buffer and PBS. After denaturation,the solution was loaded on an SDS-PAGE gel and blotted, and a Western blotanalysis was performed using either polyclonal TSWV NSs or monoclonal anti-body detecting Asian tospoviral NSs (supplied by S. D. Yeh).

Dicer cleavage reactions. Drosophila melanogaster embryo extract preparationwas described previously (25). In Dicer-mediated cleavage reactions, embryoextracts were incubated for 3 h at 25°C in reaction mixtures as described previ-ously (25), where KCl was omitted from the reaction mixture. In a typical 10-�lreaction mixture, 2 �l Drosophila embryo extract, 0.5 ng dsRNA, and 2 �gvirus-infected extract were mixed. Samples were deproteinized, and RNA wasanalyzed on a 12% denaturing gel. Electrophoresis gels were dried, exposed to aphosphor screen, and scanned (Molecular Dynamics Typhoon PhosphorImager;Amersham Biosciences).

Agrobacterium tumefaciens transient transformation assay (ATTA). Agrobacte-rium infiltration was performed as previously described (10). N. benthamianaleaves were coinfiltrated with Agrobacterium (at an optical density at 600 nm(OD600) of 0.25) harboring binary vectors encoding IR-GFP, GFP, and differentconstructs coding for MBP, CymRSV P19, turnip crinkle virus (TCV) coatprotein (CP), TYRV NSs, or TSWV NSs. Expression of GFP in the leaves wasmonitored 3 days postinfiltration (dpi) with a hand-handled UV lamp, andphotos were taken with a Canon Power Shot A710IS digital camera, using thehigh fluorescence setting. For the miRNA-based sensor constructs, Agrobacte-rium (at an OD600 of 0.1) harboring eGFP-3�UTR, miRNA1, and either MBP orTSWV NSs was coinfiltrated in N. benthamiana leaves. Green fluorescence wasvisualized 5 dpi with a Leica binocular microscope (type S) and the GFP Plusfluorescence module 10446143.

Northern blot analysis. RNA extraction was performed as described previ-ously (9), and 7 �g of total RNA was mixed with formaldehyde loading buffer,heated for 5 min at 70°C, and separated on an 1% agarose gel. The RNA wastransferred onto a Hybond-N membrane (Pharmacia-Biotech), followed by UV-cross-linking.

For miRNA1 and siRNA detection, 5 �g RNA enriched for small RNAs (27)was separated on a 20% 0.5� TBE denaturing acrylamide gel. Following sepa-ration, the RNA was electroblotted onto Hybond-N� (Pharmacia-Biotech) andcross-linked by UV light. Hybridization was performed overnight at 48°C inmodified church buffer (0.36 M Na2HPO4, 0.14 M NaH2PO4, 7% [wt/vol] SDS,1 mM EDTA) with either an eGFP or miRNA1-specific digoxigenin (DIG)-labeled DNA probe. The blots were washed briefly three times with 2� SSC (1�SSC is 0.15 M NaCl plus 0.015 M sodium citrate) and three times for 15 min with2� SSC supplemented with 0.2% (wt/vol) SDS at 48°C. The labeled probe wasdetected by Western blot analysis using a DIG-specific antibody conjugated toalkaline phosphatase in blocking buffer (maleic acid buffer plus 1% blockingreagent) and CSPD as a substrate (Roche) according to the manufacturer’srecommendations.

For the detection of miRNA171 in tospovirus-infected extracts, 5 to 15 �gsmall RNA was loaded onto a 12%, 1� TBE denaturing gel, electroblotted ontoa Hybond-N� membrane (Pharmacia-Biotech), and hybridized overnight at 50°Cin hybridization buffer (1 mM EDTA, 0.36 M Na2HPO4, 0.14 M NaH2PO4, 7%[wt/vol] SDS) using locked nucleic acid probes (2 �g). Probes specific formiRNA171c or miRNA171c* were labeled using polynucleotide kinase and(�-32P)ATP. Following hybridization, blots were washed briefly with 2� SSC–0.2% (wt/vol) SDS, twice for 20 min with 2� SSC–0.2% (wt/vol) SDS, and oncefor 20 min with 1� SSC–0.1% (wt/vol) SDS at 50°C. Blots were exposed to a

phosphor screen and scanned (Molecular Dynamics Typhoon PhosphorImager;Amersham Biosciences). Stripping of blots was performed at 85°C using 200 mlbuffer containing 1 mM EDTA and 0.1% (wt/vol) SDS for 15 min and used forsubsequent hybridization experiments.

RESULTS

TSWV NSs binds long and short dsRNA in vitro. Previouslyit was reported that of the TSWV proteins tested, only the NSsprotein was able to suppress RNA silencing in Agrobacteriuminfiltration assays (10, 51). Furthermore, silencing suppressionby NSs revealed significantly low levels of target siRNA incomparison to suppression by tombusvirus P19 and rice hojablanca virus (RHBV) NS3, both known to specifically bind onlysmall double-stranded RNAs (i.e., siRNAs) (10, 30, 37, 51).These results implied that TSWV NSs interferes in the RNAsilencing pathway upstream of siRNA synthesis, e.g., by bind-ing to long dsRNA and thereby preventing these from becom-ing processed into siRNAs by Dicer (like) proteins.

To analyze whether TSWV NSs exerts its suppressor func-tion by sequestering dsRNA, the affinity to various dsRNAmolecules was analyzed. In vivo, the NSs protein tends to formlarge insoluble aggregates (36), and earlier attempts to expressand purify NSs from Escherichia coli failed due to solubilityproblems. Thioredoxin has been reported to increase transla-tion efficiency and solubility of eukaryotic proteins expressed inE. coli (39). For these reasons and because N-terminal fusionsto NSs were shown not to hamper RSS activity (data notshown), NSs was expressed in E. coli, fused at its N terminus toHis-patched thioredoxin (HP-thioredoxin). After purification,the HP-thioredoxin-NSs fusion protein was incubated withradiolabeled 114-nt dsRNA or 21-nt siRNA molecules andsubsequently analyzed by electrophoretic mobility shift as-says (EMSA) on native acrylamide gels (38). These analysesrevealed that upon increasing NSs concentrations, both siRNA(Fig. 1A) and 114-nt dsRNA (Fig. 1C) showed a retardation inelectrophoretic mobility, together indicating that TSWV NSswas able to bind both siRNA and long dsRNA. As negativecontrols, dsRNAs or siRNAs were incubated with purifiedHP-thioredoxin N-terminally fused to the inert maltose bind-ing protein (MBP), and in both cases, even at the highestconcentrations tested, no complex formation was observed(Fig. 1B and D). While the mobility of siRNAs was still re-tarded when they were incubated in the presence of a ratherlow concentration (14.8 nM) of NSs, only binding to 114-ntdsRNA was observed at significantly higher concentrations ofNSs (237.5 nM). These results indicated a higher affinity ofTSWV NSs for siRNA molecules than for long dsRNA.

To verify these results, experiments were repeated using thenative NSs protein produced from the eukaryotic baculovirus-insect cell expression system. Due to the lack of a tag forconvenient purification purposes, entire insect cell extractscontaining the expressed NSs were used. To determine thebinding affinity of NSs, extracts were incubated with radiola-beled 21-nt siRNAs (Fig. 1E) or long (114-nt) dsRNA mole-cules (Fig. 1F). The expression and presence of NSs in thesoluble fraction of insect cell extracts used were verified byWestern blot analysis (Fig. 1G). The EMSA analyses (Fig. 1Eand F) showed that the NSs-containing insect cell extractsexhibited dsRNA affinity profiles similar to the results obtained

11544 SCHNETTLER ET AL. J. VIROL.

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 4: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

FIG. 1. Affinity of TSWV NSs for 21-nt siRNAs and 114-nt dsRNA. Different concentrations of either bacterially purified HP-thioredoxin-NSsprotein or insect cell extract infected with a recombinant baculovirus expressing NSs or GFP were incubated for 20 min at room temperature with100 pM of 32P-labeled siRNA (A and E) or 114-nt dsRNA (panels C and F). RNA-protein complexes were separated on a native polyacrylamidegel, and a representative picture is shown from at least two independent experiments. As negative controls, RNA was incubated in the presenceof elution buffer (A and C) (first lane), GFP-containing insect cell extracts (E and F) (first lane), or HP-thioredoxin-MBP (B and D). In the caseof HP-thioredoxin-MBP (B and D), HP-thioredoxin-NSs was used as a positive control. Expression of NSs in infected insect cell extracts wasverified by Western blot analysis using an NSs-specific polyclonal antibody (G).

VOL. 84, 2010 TOSPOVIRUS NSs EXHIBITS dsRNA AFFINITY 11545

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 5: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

with purified HP-thioredoxin NSs. Similar binding affinitieswere also observed with longer, 400-nt dsRNA molecules (datanot shown). Although complex formation with 114-nt dsRNA(and 400-nt dsRNA) was also observed in the negative-control,insect cell extracts infected with a recombinant baculovirusexpressing GFP (Fig. 1F, asterisk), the complex clearly showeda different mobility from the NSs-RNA complex. Due to theabsence of this complex when using siRNA in EMSA (Fig. 1E),the origin of this complex was not investigated further.

To confirm the higher affinity of NSs for siRNAs versus longdsRNA, an in vitro affinity competition was performed. To thisend, a fixed amount of baculovirus-infected cell extract wasmixed with radiolabeled siRNAs in the presence of an increas-ing concentration of nonlabeled long dsRNA competitor (Fig.2A) and vice versa (Fig. 2B). For easier comparison, the per-centage of bound versus unbound labeled RNA molecules wasquantified and is presented in a graph (Fig. 2C). Results showthat whereas the amount of siRNAs present in an NSs complex(approximately 91%) hardly changed upon addition of longdsRNA (Fig. 2A), in the reciprocal situation, the amount ofNSs-long dsRNA complexes was readily reduced (from anapproximately 100% bound status in the absence of siRNAcompetitor to an approximately 58% bound status at a 100�molar excess of siRNA competitor) by the addition of siRNAcompetitors (Fig. 2B and C). The latter was confirmed by anincreased signal of unbound dsRNA (Fig. 2) (from 0% to90%). These results further supported a higher affinity of

TSWV NSs for siRNA molecules over long dsRNA. As de-scribed above (Fig. 1F, asterisk), again a distinct mobility shiftwas observed when, as a negative control, the long dsRNAmolecules were incubated with insect cell extracts infected witha recombinant baculovirus expressing GFP (Fig. 2B).

TSWV NSs inhibits Dicer-mediated dsRNA processing. SinceTSWV NSs exhibited affinity to long dsRNA (Fig. 1C and F)and lower levels of target siRNAs from silenced genes wereobserved in previous experiments (10), it was tempting to as-sume that TSWV NSs prevents long dsRNA from becomingcleaved by DCL proteins. To test whether TSWV NSs indeedwas able to interfere with Dicer-mediated dsRNA processing,Dicer cleavage assays using Drosophila embryo extracts wereperformed. As a substrate, 114-nt dsRNA was used, and theproduction of 21-nt siRNAs was monitored in the presence ofincreasing amounts of insect cell extracts containing recombi-nant baculovirus-expressed TSWV NSs. As a negative control,extract of insect cells infected with a recombinant baculovirus-GFP was included. Only when extracts harboring TSWV NSswere added, the formation of siRNAs was significantly reduced(Fig. 3A), indicating that TSWV NSs interfered with Dicercleavage of dsRNA in vitro. For comparative purposes, therelative percentage of processed dsRNA was quantified forthree independent Dicer cleavage reactions (Fig. 3B).

To verify whether the in vitro-observed inhibition of Dicercleavage by TSWV NSs also occurred in vivo, plants werecoinfiltrated with Agrobacterium harboring a plasmid encoding

FIG. 2. Competition experiments with recombinant baculovirus-NSs-infected cell extracts for siRNAs and 114-nt dsRNA. Fixed concentrationsof insect cell extracts infected with a baculovirus expressing TSWV NSs were incubated with 32P-labeled siRNA and increasing amounts (0, 100�,200�, 250�, and 300� molar excess) of unlabeled 114-nt dsRNA competitor molecules (A) or 32P-labeled 114-nt dsRNA and increasing amounts(0, 100�, 200�, 250�, and 300� molar excess) of unlabeled siRNA competitor molecules (B). Samples were loaded and resolved on a nativeacrylamide gel. As a negative control, extracts from cells infected with a GFP-expressing baculovirus were used. In the case of 114-nt dsRNA, alower retardation complex is formed for the negative control (*). The percentage of bound RNA was quantified by GeneTools (SynGene) and isrepresented versus the concentration of competitor (114-nt dsRNA or 21-nt siRNA) (C).

11546 SCHNETTLER ET AL. J. VIROL.

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 6: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

an inverted repeat of GFP (IR-GFP) and/or GFP and eitherTSWV NSs, tomato yellow ring virus (TYRV) NSs, or cym-bidium ringspot virus (CymRSV) P19, the latter as a negativecontrol. Turnip crinkle virus (TCV) coat protein (CP), previ-ously shown to bind long dsRNA and to inhibit Dicer cleavage(42), was used as a positive control. An Agrobacterium strainexpressing the maltose binding protein (MBP) was included asa negative control. The presence of green fluorescence (Fig.3D) during all coinfiltrations, except for the negative controlMBP, indicated that all RSS proteins analyzed (TSWV NSs,TYRV NSs, CymRSV P19, and TCV CP) were able to inhibitRNA silencing induced by IR-GFP. To analyze whether thiswas due to inhibition of Dicer cleavage of the IR-GFP RNA,leaves were harvested 3 days postinfiltration, and RNA wasisolated, enriched for small RNA, and investigated by North-ern blot analysis for the presence of GFP siRNA (10). Relativeto the loading controls, GFP-specific siRNAs were observed insignificantly smaller amounts in samples from leaves coinfil-trated with TSWV NSs or TCV CP (Fig. 3 C) than in thosefrom leaves coinfiltrated with TYRV NSs, CymRSV P19, orMBP. These findings supported the idea that TSWV NSs,like TCV CP, was able to associate with long dsRNA andthereby inhibited Dicer cleavage of these long dsRNAs. SinceCymRSV P19 is known to exclusively bind siRNAs and sup-presses RNA silencing downstream of Dicer, the similar levelsof GFP siRNAs observed for TYRV NSs and CymRSV P19

suggested that TYRV NSs was not able to inhibit Dicer cleav-age of long dsRNA.

Binding of dsRNA is a common feature of tospoviral NSsproteins. The observation that TYRV NSs was not able toinhibit Dicer cleavage suggested a very low affinity of TYRVNSs for long dsRNA, and it is therefore distinct from TSWVNSs. To test whether the affinity for differently sized dsRNAmolecules, as was demonstrated for TSWV NSs (Fig. 1), isshared among other tospoviruses, a comparative EMSA anal-ysis was performed with several tospoviruses. To this end,infected plant extracts from a range of different tospovirusspecies (TSWV), groundnut ringspot virus (GRSV) and impa-tiens necrotic spot virus (INSV) from the American clade andTYRV from the Eurasian clade (29, 45), were incubated withradiolabeled molecules (siRNA or 114-nt dsRNA) and ana-lyzed by EMSA (42). The results showed that NSs-containingprotein extracts of all tospoviruses were able to shift and thusbind siRNA molecules (Fig. 4A). Surprisingly, the extractsfrom INSV and GRSV showed an additional strong affinity forlonger dsRNA molecules, while the extract from TYRV NSsdid not (Fig. 4C). The latter was in agreement with the earlier-observed lack of indirect inhibition on Dicer cleavage byTYRV NSs (Fig. 3C). Surprisingly, nearly no affinity to 114-ntdsRNA could be observed for TSWV NSs in this assay (Fig.4C), in contrast to previous results with E. coli- and baculovi-rus-expressed TSWV NSs (Fig. 1). As already mentioned for

FIG. 3. Analysis of Dicer-mediated dsRNA cleavage in the presence of TSWV-NSs. Radioactively labeled dsRNA was cleaved into siRNAusing Drosophila embryo extract in lysis buffer but was inhibited in the presence of TSWV NSs-expressing baculovirus-infected insect cell extracts.Controls used were GFP-expressing baculovirus-infected insect cell extracts as a negative control and undiluted and 0.5�-diluted (0.5� GFP and0.5� TSWV NSs) extracts. A representative picture of at least two independent experiments is presented (A). The fraction of cleaved dsRNA wasquantified by GeneTools (SynGene) and is represented as a relative percentage by setting the amount of cleaved dsRNA in GFP extract as 100%for undiluted (gray bars) or 0.5�-diluted (black bars) extracts (B). Agrobacterium strains harboring vectors encoding IR-GFP were coinfiltratedin N. benthamiana leaves with MBP (negative control), TCV CP (positive control), CymRSV P19, TSWV NSs, or TYRV NSs constructs. Thecorresponding GFP siRNAs levels were detected by Northern blot analysis (C). Ethidium bromide-stained RNA was used as a loading control. Toensure silencing suppressor activity of the tested RSS proteins, their gene constructs were coinfiltrated with a GFP construct in N. benthamianaleaves and monitored for GFP expression 5 days postinfiltration (D).

VOL. 84, 2010 TOSPOVIRUS NSs EXHIBITS dsRNA AFFINITY 11547

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 7: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

the baculovirus-infected extract (Fig. 1F, asterisk), a lowercomplex was observed in the case of TSWV and TYRV with114-nt dsRNA. This retardation was not due to NSs, since asimilar retardation complex was observed with uninfectedplant extracts as a negative control (Fig. 4C, asterisk), andtherefore, it was not further investigated.

To quantify the affinity to longer and shorter dsRNA, aserial dilution series of INSV- and GRSV-infected plant ma-terial was made and used in dsRNA binding assays. Resultsfrom this analysis showed that binding to long dsRNA waslost at dilutions of INSV- and GRSV-infected plant extractsthat contained �0.25 �g total protein per 10 �l whereasbinding to 21-nt siRNA molecules was not observed below�0.06 �g total protein per 10 �l (compare Fig. 4E to F).This indicated a slightly higher affinity to siRNA moleculesof GRSV NSs and INSV NSs (data not shown), as similarlyobserved for baculovirus- and E. coli-expressed TSWV NSs(Fig. 1A and C).

It was assumed that the tospoviral NSs proteins provided thedsRNA binding activity in the infected plant extract. Due tothe lack of any NSs-specific antibody (INSV) or the incapacityof the present antibodies to detect native NSs proteins effi-ciently (GRSV, TSWV, and TYRV), supershift and immuno-precipitation experiments (data not shown) showed negativeresults.

Therefore, the presence of NSs in the retarded dsRNAcomplexes formed in tospoviral plant-infected extract wasdemonstrated. TSWV-, GRSV-, TYRV-, and CymRSV-(negative control) infected plant extracts and Agrobacterium-infiltrated leave extract expressing TSWV NSs (positive con-trol) were incubated with radiolabeled siRNA molecules andanalyzed by EMSA, and shifted siRNA complexes were ex-cised from the gel. These were subsequently resolved in SDS-PAGE, using excised unbound siRNA molecules as a negativecontrol, and after Western blotting screened for the presenceof NSs. The mobility-shifted siRNA complexes from TSWV-,

FIG. 4. Affinity analysis of NSs from tospovirus-infected plant extracts for dsRNA. Electrophoretic mobility shift assays were performed usingsystemically infected N. benthamiana leaf extracts containing CymRSV, GRSV, INSV, TSWV, or TYRV, incubated with radioactively labeled21-nt siRNA, and subsequently resolved on a 8% native gel (A). Except for CymRSV, the experiment was repeated with radioactively labeled114-nt dsRNA and resolved on a 5% native gel (C). Uninfected plant extract (uninfected) and RNA only (�) were included as negative controls(A and C, first two lanes). The experiments shown in panels A and C were repeated using recombinant baculovirus-infected extract containingGRSV NSs, TYRV NSs, TSWV NSs, or GFP (siRNAs [B] or 114-nt dsRNA [D]). Complexes formed in the negative controls (uninfected or GFPexpressing) are indicated (*) (C, lane 2, and D, lane 1). A serial dilution of GRSV-infected N. benthamiana extracts (2 �g to 15.6 ng proteincontent) was tested for affinity to 21-nt siRNA (E) or 114-nt dsRNA (F). As a negative control, the RNA duplex in extraction buffer was included(E and F, first lane). Western blot analysis was performed on excised and denatured siRNA-protein complexes as observed in panel A using eitheran antibody against TSWV NSs (G) or a monoclonal NSs antibody detecting Asian tospoviral NSs (H). The siRNA-protein complex of TSWV NSsexpressed in Agrobacterium-infiltrated leaf extract (TSWV NSs ATTA) was used as a positive control. As a negative control, gel slices of unboundsiRNAs were used (unbound RNA complex).

11548 SCHNETTLER ET AL. J. VIROL.

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 8: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

GRSV-, and TYRV-infected or TSWV-NSs agroinfiltrated ex-tracts clearly showed the presence of NSs (Fig. 4G and H),which was not detected in the unbound siRNA- or CymRSV-infected samples. This supports the hypothesis that the re-tarded complexes consist of the tospoviral NSs proteins thatprovide the dsRNA binding activity in the infected plant ex-tracts.

To further substantiate the observed differences in bindingaffinities of tospovirus NSs proteins to long dsRNA and tostrengthen the idea that the NSs proteins were binding thedsRNA molecules, another expression system was used. EMSAanalyses were performed using extracts from recombinant bacu-lovirus-infected insect cells expressing NSs from TSWV, GRSV(American clade), and TYRV (Eurasian clade). As expected, allNSs extracts showed retardation of siRNA molecules (Fig. 4B)but not of single-stranded RNA molecules (data not shown). Aretardation of long, either 114-nt or 400-nt dsRNA was againobserved in the presence of NSs from TSWV and GRSV butnot with TYRV NSs containing extracts (Fig. 4D and data notshown). This is in agreement at least with the results obtainedwith infected plant extract containing GRSV and TYRV NSs.In dilution series and competition EMSA analysis, baculo-virus-expressed GRSV NSs again revealed higher affinitiesfor siRNA than for long dsRNA molecules (data not shown),as already observed in infected plant extract (Fig. 4E and F)and for baculovirus-expressed TSWV NSs (Fig. 1 and 2).

Tospoviral NSs proteins interfere with the miRNA pathway.The reported loss of leaf polarity and interference with theformation of plant reproduction organs after constitutive ex-pression of plant viral RSS in Arabidopsis have been explainedas a result of suppression of miRNA-mediated gene regulation(11, 16). Since siRNAs and miRNA/miRNA* duplexes sharestructural similarities, it was not surprising that the RHBV NS3and tombusvirus P19 RSS proteins, both able to bind siRNAs,exhibited similar affinities to miRNA/miRNA* duplexes (16,30). To investigate whether the TSWV NSs protein was alsoable to interfere with miRNA-mediated gene regulation, theaffinity to either miRNA/miRNA* duplexes or the longer pre-miRNA was analyzed using infected insect cell extracts contain-ing baculovirus-expressed TSWV NSs. EMSAs with increasingamounts of NSs-containing cell extracts showed retardation of themiRNA/miRNA* duplex slightly lower than that observed forsiRNAs (Fig. 5A and 1A). For the pre-miRNA molecules, noshift was observed even at the highest concentration of cell ex-tracts used (Fig. 5B), indicating that in this assay, TSWV NSsshows affinity only for miRNA/miRNA* duplexes.

The affinity of TSWV NSs for miRNA/miRNA* duplexes invitro implied that NSs could potentially interfere with themiRNA pathway in plants by sequestering miRNA/miRNA*molecules. To test this hypothesis, a miRNA-based sensor con-struct encoding eGFP harboring a 3�UTR with target sitesfor miRNA1 (eGFP-3�UTR) was agroinfiltrated in RDR6-silenced N. benthamiana together with either TSWV NSs orthe negative control MBP. Use of the miRNA1-dependentsensor construct instead of previously described ones (46) en-sured that the observed miRNA/miRNA* duplex binding byNSs is not sequence specific for miRNA171.

Normally, during Agrobacterium infiltration of this eGFP-miRNA1 sensor construct, the host-encoded RDR6 convertsfunctional RNA transcripts into dsRNA, resulting in silencing

of eGFP and production of eGFP-specific siRNAs (Fig. 5D).In RDR6-knockdown plants, no silencing of this constructoccurs unless it is coinfiltrated with pri-miRNA1 (12, 46). Asexpected, in the absence of pri-miRNA1, RDR6 knockdownplants showed similar eGFP fluorescence in the presence orabsence of the RSS protein (Fig. 5C, upper panels), while adrastic decrease in the eGFP fluorescence level was observedwhen these plants were coinfiltrated with eGFP-3�UTR andpri-miRNA1 (Fig. 5C, lower left panel). Enhanced GFP fluo-rescence was restored by the addition of TSWV NSs (Fig. 5C,lower right panel) and not when using MBP as a negativecontrol (Fig. 5C, lower left panel), demonstrating that TSWVNSs was able to suppress miRNA-induced silencing. Whereasno eGFP-specific siRNAs were observed in the absence ofmiRNA1 (Fig. 5D, left panel, lanes 2 and 3), only a slightamount of siRNAs was observed in the presence of miRNA1(Fig. 5D, left panel, lanes 4 and 5). In contrast, elevated levelsof eGFP-specific siRNAs were produced when the sensor con-struct was infiltrated into wild-type N. benthamiana plants (Fig.5D, lane 6, left panel). This strongly indicates that eGFP si-lencing was most likely the result of translational repression.

Similar results were obtained in insect cells expressing TSWVNSs and a Firefly luciferase-based miRNA1 sensor construct(Fig. 5E). This organism-independent suppressor effect sup-ported the idea that the observed interference with themiRNA pathway is possibly due to miRNA/miRNA* se-questering and not to a protein-specific interaction.

To determine if the capacity to suppress the miRNA pathwaywas shared among tospoviral NSs proteins, the EMSAs us-ing miRNA/miRNA* duplex or pre-miRNA were repeated forthe other tospoviruses. Next to the positive-control CymRSV,miRNA/miRNA* retardation was observed only when in-fected plant extracts containing TYRV were used and not withGRSV, INSV, and TSWV (Fig. 6A). In contrast, retardation ofpre-miRNA complexes was observed only with infected leafextracts containing GRSV and INSV and not with TSWV andTYRV (Fig. 6B). Strikingly, no affinity to miRNA/miRNA*duplexes was observed for TSWV NSs when crude extracts ofvirus-infected plants were used (Fig. 6A), whereas binding wasobserved when using recombinant baculovirus-TSWV NSs-in-fected cell extracts (Fig. 6C) or plant extracts agroinfiltratedwith TSWV NSs (data not shown) (Fig. 5A and 6C). Similarresults were obtained for Agrobacterium-infiltrated leaf ex-tracts and baculovirus-infected cell extract of GRSV NSs andTYRV NSs (data not shown) (Fig. 6C).

These results indicated that the NSs proteins of the tospo-viruses analyzed interfered with the miRNA pathway. How-ever, their mode of action seemed to differ depending on theexpression system used. To substantiate these findings withevidence from the natural situation, RNA was isolated fromtospovirus-infected plant material and assayed for the presenceof miRNA171c and miRNA171c* molecules. If the tospovirusNSs protein binds miRNA/miRNA* duplexes, it would preventRISC loading of the miRNA guide strand, the subsequenttarget cleavage in plants, and degradation of the miRNA*strand (11, 16). Detection of miRNA/miRNA* duplexes andspecifically of the miRNA* strand in plants containing NSsthus would be indicative for direct association of the RSSwith miRNA/miRNA* duplexes. Indeed, both miR171c andmiRNA171c* were readily detected in RNA samples from

VOL. 84, 2010 TOSPOVIRUS NSs EXHIBITS dsRNA AFFINITY 11549

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 9: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

TSWV- and TYRV-infected plant material (Fig. 6D), whereasin uninfected plants, only the miR171c strand could be de-tected under the same conditions. The miRNA* strand wasalso readily detected in RNA samples of GRSV- and INSV-

infected plant material (Fig. 6D). This indicates that their NSsproteins also interfere with the RISC loading step, most likelyby sequestering the miRNA/miRNA* duplexes.

This would assume that GRSV and INSV NSs exhibit affinity

FIG. 5. Analysis of TSWV NSs interference with the miRNA pathway. EMSA analysis of radioactively labeled miRNA171/miRNA171* duplex(A) or pre-miRNA2b (B) in the presence of increasing amounts of recombinant baculovirus-TSWV NSs-infected insect cell extracts. A repre-sentative picture of at least three independent repetitions is shown. TSWV NSs interference on the miRNA pathway in RDR6 knockdown N.benthamiana plants as visualized by eGFP fluorescence from an eGFP-miRNA sensor construct (eGFP-3�UTR) 5 days postcoinfiltration (C) isshown. As controls, leaves were infiltrated with Agrobacterium harboring vectors for eGFP-3�UTR and MBP (C, upper left) or eGFP-3�UTR andTSWV NSs (C, upper right). Silencing was induced by coinfiltration of pri-miRNA1 (C, lower left and right) and suppressed in the presence ofTSWV NSs (C, lower right). Levels of eGFP siRNA (D, left panel) and processed miRNA1 (D, right panel) were detected by Northern blothybridization. RNA of a wild-type N. benthamiana plant infiltrated with Agrobacterium harboring the eGFP-3�UTR sensor was used as a positivecontrol (D, last lane, top panel). Noninfiltrated N. benthamiana was used as a negative control (D, first lane, top panel) (wt). As a loading control,RNA was stained by ethidium bromide. Suppression of miRNA1-induced silencing of a firefly luciferase-miRNA1 sensor construct was investigatedin insect cells (E). Drosophila S2 cells were cotransfected with a pMT-Renilla luciferase (Rluc), pMT-firefly luciferase (Fluc)-miRNA1 sensorconstruct, either specific (miRNA-1) or unspecific (miRNA-12) primary miRNA in concert with either pIB-MBP, -TSWV NSs, or -carnationItalian ringspot virus (CIRV) P19. After induction at 48 h posttransfection (hpt), relative luciferase expression (firefly/Renilla) was determined 72hpt and the mean of at least two independent experiments is shown with standard error.

11550 SCHNETTLER ET AL. J. VIROL.

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 10: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

for miRNA/miRNA* duplexes in vivo, in contrast to the in vitrodata with infected plant extract (Fig. 6A).

DISCUSSION

A common strategy employed by many plant viruses to coun-teract antiviral RNA silencing is to sequester and inactivate(antiviral) siRNAs through their viral RSS protein. Size-selective binding of siRNAs has been described for severalplant viral suppressors, such as RHBV NS3, tobacco etchvirus (TEV) HC-Pro, beet yellows virus (BYV) P21, P19 ofseveral tombusviruses, peanut clump virus P15, and barleystripe mosaic virus �B (30, 37, 42). TSWV NSs has been re-ported previously to act as an RSS in plants and insect cells (10,23, 48); however, its mode of action remained unclear. Resultsshown in this work demonstrate that E. coli- and recombinantbaculovirus-expressed TSWV NSs can bind both siRNA andlong dsRNA molecules. This enables TSWV to block antiviralRNA silencing at two stages, i.e., before and after Dicer-me-diated dsRNA cleavage. Similar results were obtained for NSsfrom GRSV and INSV, two other tospovirus species that to-gether with TSWV belong to the American clade of tospovi-ruses (45) and share 49.7 to 82.2% amino acid sequence sim-ilarity. In contrast, the NSs protein of the more distantlyrelated Eurasian clade tospovirus TYRV (15 to 22% proteinsimilarity) (29) revealed affinity only to small dsRNA mole-cules in all used extracts. The observed lack of a TYRV NSs to

bind long dsRNA is unclear. The same is true for the lowaffinity to long dsRNA binding in TSWV-infected plant ex-tract, which is in contrast to the results with baculovirus- or E.coli-expressed TSWV NSs. Considering the large quantities ofdsRNA at least in infected plants, a significant part of NSscould be preloaded with dsRNA, and a difference in the re-maining soluble and free NSs protein levels within the infect-ed-plant extract cannot be excluded. It is not known if all usedtospoviruses produce similar amounts of viral siRNAs duringinfections, possibly resulting in differences in NSs preloading.On the other hand, the fact that no Dicer inhibition could beobserved for TYRV NSs, in contrast to the case with TSWVNSs in an Agrobacterium infiltration assay, strengthens the ob-servation that TYRV NSs does not bind long dsRNA mole-cules, in contrast to TSWV NSs.

Whereas all tospovirus NSs proteins analyzed did exhibit aclear affinity for siRNAs, the EMSAs showed a clear differencein siRNA-NSs complex mobility between those from GRSV,INSV, and TSWV versus TYRV. However, this was only incases where plant extracts were used and was not found withbaculovirus-infected cell extract (Fig. 4A). In plant extracts, alltested tospovirus NSs proteins (from GRSV, TSWV, andTYRV) showed, in addition to the monomeric form, higher-molecular-weight bands corresponding in size to dimers, trim-ers, and multimers (data not shown). This was irrespective of(non)denaturing conditions and independent of the presenceof RNA. Similar multimers were observed under semidenatur-

FIG. 6. Affinity of tospovirus NSs for duplex RNA molecules from the miRNA pathway. Electrophoretic mobility assay analysis using leafextracts from N. benthamiana systemically infected with TSWV, TYRV, CymRSV, GRSV, or INSV and incubated with radioactively labeledmiRNA171/miRNA171* duplexes (A) or pre-miRNA2b (B). Baculovirus-infected cell extracts expressing GFP, TSWV NSs, TYRV NSs, or GRSVNSs were incubated for 20 min with radioactively labeled miRNA171/miRNA171* and loaded on an 8% native gel (C). As negative controls, RNAwas incubated with extracts from uninfected plants (A) (first lane) or GFP-expressing baculovirus-infected extract (C) (second lane). Northern blotdetection of miRNA171c and miRNA171c* (after stripping) in RNA samples from N. benthamiana leaves systemically infected with GRSV,TSWV, INSV, and TYRV (D) is shown. As a negative control, RNA from uninfected leaves was included (D) (first lane).

VOL. 84, 2010 TOSPOVIRUS NSs EXHIBITS dsRNA AFFINITY 11551

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 11: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

ing conditions in recombinant baculovirus NSs-infected cellextracts in the presence or absence of RNA (data not shown).Therefore, the discrepancy in stoichiometry of siRNA-RSScomplexes between the two groups of tospoviruses observedonly with NSs from plant extracts was likely not to be attrib-uted to differences between NSs oligomerization in plant ver-sus insect cells. Instead, differences in posttranslational modi-fication (e.g., phosphorylation) and/or interaction with hostproteins or viral proteins may account for the observed differ-ences in siRNA-NSs complex mobility. Predictions for post-translational modifications revealed several potential phosphor-ylation sites. To our knowledge, no viral or host-encodedinteraction partner has yet been identified for tospovirus NSs.

The reason for the additional affinity to long dsRNA tocounterdefend against RNA silencing for TSWV, GRSV, andINSV remains intriguing. The question remains if this affinityto long dsRNA molecules represents an ancestral activity lostby TYRV NSs over time or newly gained by the Americantospovirus clade. Recently, long dsRNA has been described toinduce an antiviral response in Drosophila different from RNAsilencing (33). Therefore, it is tempting to speculate that to-spoviruses from the American clade benefit from their longdsRNA affinity in order to counteract two different antiviralpathways in insects. Whether a similar antiviral response ispresent in the thrips insect vector is still unknown. Hitherto, asimilar divergence in the affinity of RSS for longer and shorterdsRNAs has been observed in only one other well-studiedbut completely unrelated family of viruses, the Tombusviri-dae (42, 43).

Although binding of TSWV NSs to longer dsRNA moleculeshas not been observed, or only to a small degree, in all used cellextracts, since it is nearly lacking in infected plant extracts, thisbinding property is supported by its in vitro inhibitory effect onDicer-mediated processing of dsRNA molecules into siRNAs.A similar inhibitory effect on Dicer cleavage of an invertedrepeat of GFP has also been observed in plants that transientlyexpressed TSWV NSs. Previous research reported that TSWVNSs is not able to suppress IR-induced RNA silencing (51),suggesting a mode of action upstream of DCL in the RNAsilencing pathway. However, in our hands, analyses haveconsistently shown that TSWV NSs was perfectly capableof suppressing IR-induced silencing in Agrobacterium-infil-trated plants.

Among plant viruses, binding of viral RSS proteins to longerdsRNA and subsequent inhibition of Dicer-mediated dsRNAcleavage so far has been observed for only two members ofthe positive-stranded Tombusviridae, i.e., Turnip crinkle virus(TCV) CP and aureusvirus p14 (42, 43). This property, though,is more common to RSS proteins of insect- and mammal-infecting viruses, like Flock house virus B2, Drosophila C virus1A, and Ebola virus VP35, which all have been shown toexhibit a high affinity to long dsRNA (34, 41–43, 54).

Tospoviruses are the only plant-infecting members of theBunyaviridae family and are transmitted by thrips, in whichthey also replicate (57). As a result, tospoviruses are targetedby antiviral RNA silencing in plants as well as insects. Byinteracting with both long and short dsRNA molecules, tospo-viruses are able to interfere with multiple steps in the antiviralRNA silencing machinery, notably Dicer-mediated processes,assembly of active RISC complexes, and possibly the amplifi-

cation of the silencing signal in plants. The NSs protein of theanimal-infecting bunyavirus La Crosse virus (LACV) has beenshown to be an active suppressor of RNA silencing in humancells by interfering with siRNA-mediated RNA silencing (49).Recently, however, contradictory results have been published(7). Since vertebrates possess an effective antiviral defensesystem based on interferon induction that also involves dsRNAspecies, the antiviral activity of the ubiquitous RNA silencingmachinery in vertebrate systems is still being debated. For thesame reason, the biological relevance of RSS activity of someproteins from animal-infecting viruses is being disputed, sincemany of these proteins also have shown interferon antagonisticproperties (4, 26). Besides being an inducer of RNA silencingand a substrate for Dicer, long dsRNA molecules of cellular orviral origin also activate the dsRNA-dependent protein kinase(PKR) and thereby trigger the interferon-induced antiviral de-fense mechanism in mammals (22). Sequestering long dsRNAmolecules would therefore be an effective way for vertebrateviruses to simultaneously suppress antiviral RNA silencing andinterferon induction. The capacity of NSs proteins of sometospoviruses to bind long dsRNA could thus reflect a propertyinherited from a common ancestor shared between the plant-and mammal-infecting bunyaviruses (13, 35). Whether thislong dsRNA binding is redundant or still has biological rele-vance to tospovirus infections in plants or insects remains to beinvestigated. Comparison of mutated NSs proteins lacking longdsRNA binding ability with wild-type NSs proteins could shedlight on the question of whether siRNA binding is sufficient forthe RSS activity of tospoviral NSs proteins. For this, moreinformation regarding the RNA binding domain(s) in the to-spoviral NSs proteins would be required.

The property of tospovirus NSs proteins of binding miRNA/miRNA* duplexes and thereby interfering in the regulation ofhost gene expression has previously been shown for a few otherplant viral RSS proteins, such as potyviral HC-Pro and tom-busviral P19 (11, 16). The affinity of TSWV, like that of TYRV,for miRNA/miRNA* but not for pre-miRNA molecules, asobserved for GRSV and INSV, is somewhat intriguing in lightof earlier-mentioned similarities and differences between theNSs proteins studied and needs further analysis. The observedlack of affinity of TSWV-, GRSV-, and INSV-infected plantextracts for miRNA/miRNA* duplexes is likely due to reasonsdiscussed for the observed low binding affinity of infectedTSWV plant extract to long dsRNA, i.e., a significant part ofNSs is already loaded with (viral) dsRNA molecules that ariseduring a viral infection. This is supported by the ability of alltested NSs proteins (from TSWV, GRSV, and TYRV) to bindmiRNA/miRNA* duplexes in Agrobacterium-infiltrated plantextracts (data not shown) and baculovirus-infected cell extracts(Fig. 6C). The small amount of NSs-siRNA complexes ob-served for TSWV-, INSV-, and GRSV-infected plant extractscompared to baculovirus-NSs-infected cell extracts (Fig. 4Aand B) and Agrobacterium-infiltrated plant extract (data notshown) strengthens this explanation. Furthermore, in light ofthe slightly lower affinity to miRNA/miRNA* duplexes (ap-proximately 5% bound at the lowest used concentration) thanto siRNAs (approximately 90% bound at lowest used concen-tration) in the case of baculovirus-infected extract expressingTSWV NSs, maybe it is not that surprising that no miRNA/

11552 SCHNETTLER ET AL. J. VIROL.

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 12: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

miRNA* retardation was observed for TSWV-, GRSV-, andINSV-infected plant extract.

Despite these observations, the accumulation of miRNA171c/miRNA171c* duplexes in leaf material infected with any to-spovirus strengthens the idea that all tospoviral NSs proteinsstabilize the miRNA/miRNA* duplexes and prevent their up-loading into RISC. This possibly occurs by sequestering andthereby interfering with miRNA-mediated gene regulation inplanta. Since miRNA171c is predicted to target transcripts forthe SCARECROW-like transcription factor (50, 58) and abeneficial effect of this gene on virus replication is not evident,it is tempting to speculate that NSs interference with themiRNA pathway is most likely due to the high structural sim-ilarity of miRNA/miRNA* molecules to antiviral siRNA mol-ecules. Further support for this idea comes from the observedsilencing suppression effect of TSWV NSs on a “randomly”selected miRNA1 sensor construct during Agrobacterium infil-trations on N. benthamiana leaves. Whereas the interference ofplant viral RSS in the miRNA pathway might reflect an aber-rancy due to structural similarities between siRNA andmiRNA/miRNA*s, the interference by RSS proteins of hu-man-infecting viruses (e.g., HIV-1) with miRNA silencing hasbeen proposed to down- or upregulate genes involved in anti-viral (defense) responses (5, 52). This is elegantly exemplifiedby the higher expression level of the miRNA-regulated histoneacetylase p300/CBP-associated factor (PCAF) during HIV in-fections, a host factor that has been shown to be required as acofactor for the HIV transactivator of transcription (Tat) pro-tein. Until recently, no such case has been reported for virusinfections in plants and insects. In plants, interference of TCVCP with AGO1 has been reported, resulting in changes inmiRNA levels that in turn create a virus-favorable environ-ment in the plant (3). Whether this interaction is TCV specificor a global plant viral characteristic is not yet known. Thehere-reported interaction of tospoviral NSs proteins with themiRNA pathway leaves the possibility that interference withthe plant miRNA pathway could be a characteristic shared byall plant viruses. Whether this interaction with the NSs proteinand the miRNA pathway occurs during tospovirus infectionsand results in an environment beneficial to viruses remains tobe investigated.

ACKNOWLEDGMENTS

We are grateful to Meltem Isik for constructing the binary vectorscarrying eGFP-3�UTR and pri-miRNA1 and to Sjoerd van Deventerfor the initial help with EMSA analysis using virus-infected mate-rial. Furthermore, we thank Elisa Izaurralde for providing the in-sect miRNA-expressing constructs, Lorant Lakatos for the kind gift ofthe miRNA171 and miRNA171* LNA probes, Ko Verhoeven of theDutch plant protection service (LNV-PD) for providing CymRSV in-fectious material, Afshin Hassani-Mehraban and Marcio Hedil for thebinary vector construct encoding TYRV NSs or GRSV NSs, DanielSilhavy for the binary vectors IR-GFP, GFP, and TCV-CP, MartinMarek for constructing the GFP-encoding baculovirus, Dick Lohuisfor RNA isolation from TSWV-infected plant material and help withNorthern blot analysis, S. D. Yeh for the monoclonal antibody againstAsian tospovirus NSs, and David Baulcombe for providing RDR6knockdown N. benthamiana.

This work was financially supported by the Netherlands Organiza-tion for Scientific Research, section Earth and life Sciences (NWO/ALW).

REFERENCES

1. Aliyari, R., and S. W. Ding. 2009. RNA-based viral immunity initiated by theDicer family of host immune receptors. Immunol. Rev. 227:176–188.

2. Alvarado, V., and H. B. Scholthof. 2009. Plant responses against invasivenucleic acids: RNA silencing and its suppression by plant viral pathogens.Semin. Cell Dev. Biol. 20:1032–1040.

3. Azevedo, J., D. Garcia, D. Pontier, S. Ohnesorge, A. Yu, S. Garcia, L. Braun,M. Bergdoll, M. A. Hakimi, T. Lagrange, and O. Voinnet. 2010. Argonautequenching and global changes in Dicer homeostasis caused by a pathogen-encoded GW repeat protein. Genes Dev. 24:904–915.

4. Basler, C. F., and A. Garcia-Sastre. 2002. Viruses and the type I interferonantiviral system: induction and evasion. Int. Rev. Immunol. 21:305–337.

5. Berkhout, B., and K. T. Jeang. 2007. RISCy business: MicroRNAs, patho-genesis, and viruses. J. Biol. Chem. 282:26641–26645.

6. Bernstein, E., A. A. Caudy, S. M. Hammond, and G. J. Hannon. 2001. Rolefor a bidentate ribonuclease in the initiation step of RNA interference.Nature 409:363–366.

7. Blakqori, G., S. Delhaye, M. Habjan, C. D. Blair, I. Sanchez-Vargas, K. E.Olson, G. Attarzadeh-Yazdi, R. Fragkoudis, A. Kohl, U. Kalinke, S. Weiss,T. Michiels, P. Staeheli, and F. Weber. 2007. La Crosse bunyavirus nonstruc-tural protein NSs serves to suppress the type I interferon system of mam-malian hosts. J. Virol. 81:4991–4999.

8. Bortolamiol, D., M. Pazhouhandeh, and V. Ziegler-Graff. 2008. Viral sup-pression of RNA silencing by destabilisation of ARGONAUTE 1. PlantSignal Behav. 3:657–659.

9. Bucher, E., H. Hemmes, P. de Haan, R. Goldbach, and M. Prins. 2004. Theinfluenza A virus NS1 protein binds small interfering RNAs and suppressesRNA silencing in plants. J. Gen. Virol. 85:983–991.

10. Bucher, E., T. Sijen, P. De Haan, R. Goldbach, and M. Prins. 2003. Negative-strand tospoviruses and tenuiviruses carry a gene for a suppressor of genesilencing at analogous genomic positions. J. Virol. 77:1329–1336.

11. Chapman, E. J., A. I. Prokhnevsky, K. Gopinath, V. V. Dolja, and J. C.Carrington. 2004. Viral RNA silencing suppressors inhibit the microRNApathway at an intermediate step. Genes Dev. 18:1179–1186.

12. Dalmay, T., A. Hamilton, S. Rudd, S. Angell, and D. C. Baulcombe. 2000. AnRNA-dependent RNA polymerase gene in Arabidopsis is required for post-transcriptional gene silencing mediated by a transgene but not by a virus.Cell 101:543–553.

13. de Haan, P., R. Kormelink, R. de Oliveira Resende, F. van Poelwijk, D.Peters, and R. Goldbach. 1991. Tomato spotted wilt virus L RNA encodes aputative RNA polymerase. J. Gen. Virol. 72:2207–2216.

14. Deleris, A., J. Gallego-Bartolome, J. Bao, K. D. Kasschau, J. C. Carrington,and O. Voinnet. 2006. Hierarchical action and inhibition of plant Dicer-likeproteins in antiviral defense. Science 313:68–71.

15. Ding, S. W., and O. Voinnet. 2007. Antiviral immunity directed by smallRNAs. Cell 130:413–426.

16. Dunoyer, P., C. H. Lecellier, E. A. Parizotto, C. Himber, and O. Voinnet.2004. Probing the microRNA and small interfering RNA pathways withvirus-encoded suppressors of RNA silencing. Plant Cell 16:1235–1250.

17. Dyer, B. W., F. A. Ferrer, D. K. Klinedinst, and R. Rodriguez. 2000. Anoncommercial dual luciferase enzyme assay system for reporter gene anal-ysis. Anal. Biochem. 282:158–161.

18. Eulalio, A., J. Rehwinkel, M. Stricker, E. Huntzinger, S. F. Yang, T. Doerks,S. Dorner, P. Bork, M. Boutros, and E. Izaurralde. 2007. Target-specificrequirements for enhancers of decapping in miRNA-mediated gene silenc-ing. Genes Dev. 21:2558–2570.

19. Falk, B. W., and J. H. Tsai. 1998. Biology and molecular biology of virusesin the genus Tenuivirus. Annu. Rev. Phytopathol. 36:139–163.

20. Galiana-Arnoux, D., C. Dostert, A. Schneemann, J. A. Hoffmann, and J. L.Imler. 2006. Essential function in vivo for Dicer-2 in host defense againstRNA viruses in Drosophila. Nat. Immunol. 7:590–597.

21. Galiana-Arnoux, D., and J. L. Imler. 2006. Toll-like receptors and innateantiviral immunity. Tissue Antigens 67:267–276.

22. Gantier, M. P., and B. R. Williams. 2007. The response of mammalian cellsto double-stranded RNA. Cytokine Growth Factor Rev. 18:363–371.

23. Garcia, S., A. Billecocq, J. M. Crance, M. Prins, D. Garin, and M. Bouloy.2006. Viral suppressors of RNA interference impair RNA silencing inducedby a Semliki Forest virus replicon in tick cells. J. Gen. Virol. 87:1985–1989.

24. Goto, K., T. Kobori, Y. Kosaka, T. Natsuaki, and C. Masuta. 2007. Charac-terization of silencing suppressor 2b of cucumber mosaic virus based onexamination of its small RNA-binding abilities. Plant Cell Physiol. 48:1050–1060.

25. Haley, B., G. Tang, and P. D. Zamore. 2003. In vitro analysis of RNAinterference in Drosophila melanogaster. Methods 30:330–336.

26. Haller, O., and F. Weber. 2009. The interferon response circuit in antiviralhost defense. Verh. K. Acad. Geneeskd Belg. 71:73–86.

27. Hamilton, A. J., and D. C. Baulcombe. 1999. A species of small antisenseRNA in posttranscriptional gene silencing in plants. Science 286:950–952.

28. Hammond, S. M., E. Bernstein, D. Beach, and G. J. Hannon. 2000. AnRNA-directed nuclease mediates post-transcriptional gene silencing in Dro-sophila cells. Nature 404:293–296.

VOL. 84, 2010 TOSPOVIRUS NSs EXHIBITS dsRNA AFFINITY 11553

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.

Page 13: Diverging Affinity of Tospovirus RNA Silencing Suppressor Proteins, NSs, for Various RNA Duplex

29. Hassani-Mehraban, A., J. Saaijer, D. Peters, R. Goldbach, and R.Kormelink. 2005. A new tomato-infecting tospovirus from Iran. Phytopa-thology 95:852–858.

30. Hemmes, H., L. Lakatos, R. Goldbach, J. Burgyan, and M. Prins. 2007. TheNS3 protein of Rice hoja blanca tenuivirus suppresses RNA silencing inplant and insect hosts by efficiently binding both siRNAs and miRNAs. RNA13:1079–1089.

31. Kaba, S. A., V. Nene, A. J. Musoke, J. M. Vlak, and M. M. van Oers. 2002.Fusion to green fluorescent protein improves expression levels of Theileriaparva sporozoite surface antigen p67 in insect cells. Parasitology 125:497–505.

32. Karimi, M., D. Inze, and A. Depicker. 2002. GATEWAY vectors forAgrobacterium-mediated plant transformation. Trends Plant Sci. 7:193–195.

33. Kemp, C., and J. L. Imler. 2009. Antiviral immunity in Drosophila. Curr.Opin. Immunol. 21:3–9.

34. Kimberlin, C. R., Z. A. Bornholdt, S. Li, V. L. Woods, Jr., I. J. Macrae, andE. O. Saphire. 2010. Ebolavirus VP35 uses a bimodal strategy to bind dsRNAfor innate immune suppression. Proc. Natl. Acad. Sci. U. S. A. 107:314–319.

35. Kormelink, R., P. de Haan, C. Meurs, D. Peters, and R. Goldbach. 1992. Thenucleotide sequence of the M RNA segment of tomato spotted wilt virus, abunyavirus with two ambisense RNA segments. J. Gen. Virol. 73:2795–2804.

36. Kormelink, R., E. W. Kitajima, P. De Haan, D. Zuidema, D. Peters, and R.Goldbach. 1991. The nonstructural protein (NSs) encoded by the ambisenseS RNA segment of tomato spotted wilt virus is associated with fibrousstructures in infected plant cells. Virology 181:459–468.

37. Lakatos, L., T. Csorba, V. Pantaleo, E. J. Chapman, J. C. Carrington, Y. P.Liu, V. V. Dolja, L. F. Calvino, J. J. Lopez-Moya, and J. Burgyan. 2006. SmallRNA binding is a common strategy to suppress RNA silencing by severalviral suppressors. EMBO J. 25:2768–2780.

38. Lakatos, L., G. Szittya, D. Silhavy, and J. Burgyan. 2004. Molecular mech-anism of RNA silencing suppression mediated by p19 protein of tombusvi-ruses. EMBO J. 23:876–884.

39. LaVallie, E. R., E. A. DiBlasio, S. Kovacic, K. L. Grant, P. F. Schendel, andJ. M. McCoy. 1993. A thioredoxin gene fusion expression system that cir-cumvents inclusion body formation in the E. coli cytoplasm. Biotechnology(NY) 11:187–193.

40. Li, F., and S. W. Ding. 2006. Virus counterdefense: diverse strategies forevading the RNA-silencing immunity. Annu. Rev. Microbiol. 60:503–531.

41. Lingel, A., B. Simon, E. Izaurralde, and M. Sattler. 2005. The structure ofthe flock house virus B2 protein, a viral suppressor of RNA interference,shows a novel mode of double-stranded RNA recognition. EMBO Rep.6:1149–1155.

42. Merai, Z., Z. Kerenyi, S. Kertesz, M. Magna, L. Lakatos, and D. Silhavy.2006. Double-stranded RNA binding may be a general plant RNA viralstrategy to suppress RNA silencing. J. Virol. 80:5747–5756.

43. Merai, Z., Z. Kerenyi, A. Molnar, E. Barta, A. Valoczi, G. Bisztray, Z.Havelda, J. Burgyan, and D. Silhavy. 2005. Aureusvirus P14 is an efficientRNA silencing suppressor that binds double-stranded RNAs without sizespecificity. J. Virol. 79:7217–7226.

44. Nykanen, A., B. Haley, and P. D. Zamore. 2001. ATP requirements and small

interfering RNA structure in the RNA interference pathway. Cell 107:309–321.

45. Pappu, H. R., R. A. Jones, and R. K. Jain. 2009. Global status of tospovirusepidemics in diverse cropping systems: successes achieved and challengesahead. Virus Res. 141:219–236.

46. Parizotto, E. A., P. Dunoyer, N. Rahm, C. Himber, and O. Voinnet. 2004. Invivo investigation of the transcription, processing, endonucleolytic activity,and functional relevance of the spatial distribution of a plant miRNA. GenesDev. 18:2237–2242.

47. Pfeffer, S., P. Dunoyer, F. Heim, K. E. Richards, G. Jonard, and V. Ziegler-Graff. 2002. P0 of beet Western yellows virus is a suppressor of posttran-scriptional gene silencing. J. Virol. 76:6815–6824.

48. Reavy, B., S. Dawson, T. Canto, and S. A. MacFarlane. 2004. Heterologousexpression of plant virus genes that suppress post-transcriptional gene si-lencing results in suppression of RNA interference in Drosophila cells. BMCBiotechnol. 4:18.

49. Soldan, S. S., M. L. Plassmeyer, M. K. Matukonis, and F. Gonzalez-Scarano.2005. La Crosse virus nonstructural protein NSs counteracts the effects ofshort interfering RNA. J. Virol. 79:234–244.

50. Sunkar, R., and J. K. Zhu. 2004. Novel and stress-regulated microRNAs andother small RNAs from Arabidopsis. Plant Cell 16:2001–2019.

51. Takeda, A., K. Sugiyama, H. Nagano, M. Mori, M. Kaido, K. Mise, S. Tsuda,and T. Okuno. 2002. Identification of a novel RNA silencing suppressor, NSsprotein of Tomato spotted wilt virus. FEBS Lett. 532:75–79.

52. Triboulet, R., B. Mari, Y. L. Lin, C. Chable-Bessia, Y. Bennasser, K. Leb-rigand, B. Cardinaud, T. Maurin, P. Barbry, V. Baillat, J. Reynes, P. Cor-beau, K. T. Jeang, and M. Benkirane. 2007. Suppression of microRNA-silencing pathway by HIV-1 during virus replication. Science 315:1579–1582.

53. Vaistij, F. E., L. Jones, and D. C. Baulcombe. 2002. Spreading of RNAtargeting and DNA methylation in RNA silencing requires transcription ofthe target gene and a putative RNA-dependent RNA polymerase. Plant Cell14:857–867.

54. van Rij, R. P., M. C. Saleh, B. Berry, C. Foo, A. Houk, C. Antoniewski, andR. Andino. 2006. The RNA silencing endonuclease Argonaute 2 mediatesspecific antiviral immunity in Drosophila melanogaster. Genes Dev. 20:2985–2995.

55. Voinnet, O. 2009. Origin, biogenesis, and activity of plant microRNAs. Cell136:669–687.

56. Wang, Y., G. Sheng, S. Juranek, T. Tuschl, and D. J. Patel. 2008. Structureof the guide-strand-containing argonaute silencing complex. Nature 456:209–213.

57. Wijkamp, I., J. van Lent, R. Kormelink, R. Goldbach, and D. Peters. 1993.Multiplication of tomato spotted wilt virus in its insect vector, Frankliniellaoccidentalis. J. Gen. Virol. 74:341–349.

58. Xie, Z., E. Allen, N. Fahlgren, A. Calamar, S. A. Givan, and J. C. Carrington.2005. Expression of Arabidopsis MIRNA genes. Plant Physiol. 138:2145–2154.

59. Zhang, X., Y. R. Yuan, Y. Pei, S. S. Lin, T. Tuschl, D. J. Patel, and N. H.Chua. 2006. Cucumber mosaic virus-encoded 2b suppressor inhibits Arabi-dopsis Argonaute1 cleavage activity to counter plant defense. Genes Dev.20:3255–3268.

11554 SCHNETTLER ET AL. J. VIROL.

Dow

nloa

ded

from

http

s://j

ourn

als.

asm

.org

/jour

nal/j

vi o

n 02

Dec

embe

r 20

21 b

y 17

7.44

.91.

107.