identification of a novel rna splicing pattern as a basis of restricted cell tropism of erythrovirus...

8
Identification of a Novel RNA Splicing Pattern as a Basis of Restricted Cell Tropism of Erythrovirus B19 John Brunstein, 1 Maria So ¨derlund-Venermo, and Klaus Hedman Department of Virology, Haartman Institute PL 21, University of Helsinki, Helsinki FIN-00014, Finland Received March 4, 2000; returned to author for revision May 15, 2000; accepted June 8, 2000 Prior studies on the transcription of erythrovirus B19 have identified a short leader sequence associated with all spliced viral transcripts. While some variability has been observed in the acceptor for this first intron, studies to date in both permissive and nonpermissive cell types have reported a unique splice donor site. In the semipermissive MB-02 cell line, we have found that splicing of this first intron proceeds almost exclusively via a cryptic CT donor downstream of the previously reported GT donor at nucleotide 406. The resulting messages for the viral structural proteins and 11-kDa protein are thereby made bicistronic, with the first expressible polypeptide being a 34 amino acid fusion of the NS-1 and 7.5-kDa proteins. The presence of an upstream open-reading frame on these messages is likely to block effective translation of the downstream structural protein products. We propose this as a significant mechanism in determining B19’s tropism on the basis of host cell splicing machinery, and present evidence in support of this model. Additionally, this is the first report of usage of a noncanonical splice donor in B19, and to our knowledge the first report of a CT-AG splice in any system. © 2000 Academic Press INTRODUCTION Erythrovirus (human parvovirus) B19 is the causative agent of the common childhood illness erythema infec- tiosum (Anderson et al., 1983), as well as a number of other less common conditions including fetal hydrops (Brown et al., 1984) and miscarriage (Knott et al., 1984), transient aplastic crises (Saarinen et al., 1986), and tran- sient postinfection arthropathy (White et al., 1985). In common with other parvoviruses, the virus has a small, unenveloped, icosahedrally symmetric capsid and a sin- gle-stranded DNA genome of approximately 5.6 kb (As- tell et al., 1997). Two major open-reading frames are encoded, with that on the left of negative-strand ge- nomes coding for the viral nonstructural protein NS-1 and that on the right coding for two structural proteins, VP-1 and VP-2. The two structural proteins are encoded by the same reading frame such that VP-2 is completely contained within VP-1, and differential expression of the two depends on alternative splicing to select the initia- tion codon (Ozawa et al., 1987). In addition, B19 has coding potential for two small polypeptides, one of 7.5 kDa around the middle of the genome and one of 11 kDa near the extreme right side (St. Amand et al., 1991). Spliced transcripts carrying these small open-reading frames have been found to be abundant in infected cells and expression of both proteins has been detected by Western blot, immunoprecipitation, and immunofluores- cence studies of both infected human leukemic and transfected COS-7 cell systems; however, the biological significance of these polypeptides remains unclear (As- tell et al., 1997; St. Amand et al., 1991; Luo et al., 1993). Two previous studies (Ozawa et al., 1987; St. Amand et al., 1991) have examined the transcriptional map of B19, which is unique among parvoviruses in using a single promoter to drive differential expression of nonstructural and structural genes via differential splicing (as summa- rized briefly in Fig. 5A). While these two previous studies, incorporating both permissive and nonpermissive cell types, demonstrated some minor differences in splice site usage between the cell types none of the differences were of significance with regard to coding potential of the structural gene or 11-kDa mRNAs. Both studies ob- served that all spliced messages (that is, all encoding anything other than the NS-1 protein) start with a 56-bp leader sequence corresponding to viral nt 350 through 406 (nucleotide numbering throughout this report follows the scheme of Shade et al. (1986). B19 demonstrates a markedly restricted host range, with fully permissive infections only occurring in late erythroid cell precursors and burst-forming erythroid pro- genitors (BFU-E) (Mortimer et al., 1983). Culture of the virus is generally limited to growth-factor stimulated bone marrow aspirates (Srivastava et al., 1988; Ozawa et al., 1986) or to fetal cord blood (Sosa et al., 1992; Srivas- tava et al., 1992) or to similarly treated “semipermissive” cell lines such as MB-02 which allow for very limited viral production (Munshi et al., 1993). While some other cell lines of erythroid lineage have been shown to allow for 1 To whom reprint requests should be addressed at current address: Department of Biochemistry and Molecular Biology, 2146 Health Sci- ences Mall, University of British Columbia, Vancouver, BC V6T 1Z3, Canada. Fax: (530) 687-9346. E-mail: [email protected]. Virology 274, 284–291 (2000) doi:10.1006/viro.2000.0460, available online at http://www.idealibrary.com on 0042-6822/00 $35.00 Copyright © 2000 by Academic Press All rights of reproduction in any form reserved. 284

Upload: helsinki

Post on 22-Apr-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

tscugtenaVbcttcknSfa

DeC

Virology 274, 284–291 (2000)doi:10.1006/viro.2000.0460, available online at http://www.idealibrary.com on

Identification of a Novel RNA Splicing Pattern as a Basisof Restricted Cell Tropism of Erythrovirus B19

John Brunstein,1 Maria Soderlund-Venermo, and Klaus Hedman

Department of Virology, Haartman Institute PL 21, University of Helsinki, Helsinki FIN-00014, Finland

Received March 4, 2000; returned to author for revision May 15, 2000; accepted June 8, 2000

Prior studies on the transcription of erythrovirus B19 have identified a short leader sequence associated with all splicedviral transcripts. While some variability has been observed in the acceptor for this first intron, studies to date in bothpermissive and nonpermissive cell types have reported a unique splice donor site. In the semipermissive MB-02 cell line, wehave found that splicing of this first intron proceeds almost exclusively via a cryptic CT donor downstream of the previouslyreported GT donor at nucleotide 406. The resulting messages for the viral structural proteins and 11-kDa protein are therebymade bicistronic, with the first expressible polypeptide being a 34 amino acid fusion of the NS-1 and 7.5-kDa proteins. Thepresence of an upstream open-reading frame on these messages is likely to block effective translation of the downstreamstructural protein products. We propose this as a significant mechanism in determining B19’s tropism on the basis of host

cell splicing machinery, and present evidence in support of this model. Additionally, this is the first report of usage of a

first re

WctstTawparitswtsal4t

weg

atc

noncanonical splice donor in B19, and to our knowledge the

INTRODUCTION

Erythrovirus (human parvovirus) B19 is the causativeagent of the common childhood illness erythema infec-tiosum (Anderson et al., 1983), as well as a number ofother less common conditions including fetal hydrops(Brown et al., 1984) and miscarriage (Knott et al., 1984),ransient aplastic crises (Saarinen et al., 1986), and tran-ient postinfection arthropathy (White et al., 1985). Inommon with other parvoviruses, the virus has a small,nenveloped, icosahedrally symmetric capsid and a sin-le-stranded DNA genome of approximately 5.6 kb (As-

ell et al., 1997). Two major open-reading frames arencoded, with that on the left of negative-strand ge-omes coding for the viral nonstructural protein NS-1nd that on the right coding for two structural proteins,P-1 and VP-2. The two structural proteins are encodedy the same reading frame such that VP-2 is completelyontained within VP-1, and differential expression of the

wo depends on alternative splicing to select the initia-ion codon (Ozawa et al., 1987). In addition, B19 hasoding potential for two small polypeptides, one of 7.5Da around the middle of the genome and one of 11 kDaear the extreme right side (St. Amand et al., 1991).pliced transcripts carrying these small open-reading

rames have been found to be abundant in infected cellsnd expression of both proteins has been detected by

1 To whom reprint requests should be addressed at current address:epartment of Biochemistry and Molecular Biology, 2146 Health Sci-

pl

nces Mall, University of British Columbia, Vancouver, BC V6T 1Z3,anada. Fax: (530) 687-9346. E-mail: [email protected].

0042-6822/00 $35.00Copyright © 2000 by Academic PressAll rights of reproduction in any form reserved.

284

port of a CT-AG splice in any system. © 2000 Academic Press

estern blot, immunoprecipitation, and immunofluores-ence studies of both infected human leukemic and

ransfected COS-7 cell systems; however, the biologicalignificance of these polypeptides remains unclear (As-

ell et al., 1997; St. Amand et al., 1991; Luo et al., 1993).wo previous studies (Ozawa et al., 1987; St. Amand etl., 1991) have examined the transcriptional map of B19,hich is unique among parvoviruses in using a singleromoter to drive differential expression of nonstructuralnd structural genes via differential splicing (as summa-

ized briefly in Fig. 5A). While these two previous studies,ncorporating both permissive and nonpermissive cellypes, demonstrated some minor differences in spliceite usage between the cell types none of the differencesere of significance with regard to coding potential of

he structural gene or 11-kDa mRNAs. Both studies ob-erved that all spliced messages (that is, all encodingnything other than the NS-1 protein) start with a 56-bp

eader sequence corresponding to viral nt 350 through06 (nucleotide numbering throughout this report follows

he scheme of Shade et al. (1986).B19 demonstrates a markedly restricted host range,

ith fully permissive infections only occurring in laterythroid cell precursors and burst-forming erythroid pro-enitors (BFU-E) (Mortimer et al., 1983). Culture of the

virus is generally limited to growth-factor stimulatedbone marrow aspirates (Srivastava et al., 1988; Ozawa et

l., 1986) or to fetal cord blood (Sosa et al., 1992; Srivas-ava et al., 1992) or to similarly treated “semipermissive”ell lines such as MB-02 which allow for very limited viral

roduction (Munshi et al., 1993). While some other cell

ines of erythroid lineage have been shown to allow for

(s7a

U1211

dt

Us

US B1

viral attachment, entry, and apparently normal viralmRNA expression as judged by Northern blot analysis(UT7 cells (Leruez et al., 1994)), or at least to allowfunction of the viral promoter (K562 cells (Gareus et al.,1998)) these cell lines do not appear to support produc-tive replication (in the case of UT7 cells, some authorshave reported very low levels of virus production (Shi-momura et al., 1992)). The mechanism(s) behind thisblock in replication have not been elucidated, but onesuggestion has been a block in transcript maturation,specifically in active 39 processing and recognition ofpolyadenylation signals (Liu et al., 1992).

In this study we report our observation that the splic-ing of the first viral intron is altered in MB-02 cells suchthat the leader sequence associated with all splicedtranscripts does not generally end at nt 406 as previouslyobserved. The vast majority of spliced transcripts in thissystem instead have a leader extending to viral nt 441,and utilize a CT splice-donor nucleotide pair in conjunc-tion with the previously reported nt 2030 AG acceptor forthe first intron. These messages thus contain a shortopen-reading frame (ORF) upstream of their proper mes-sage and, in the lack of a functional internal ribosomalentry site (IRES), result in an effective block in the trans-lation of viral structural proteins required for productiveinfection.

RESULTS

While setting up a system for the detection and anal-ysis of B19 transcription in patient samples, we chose touse infection of the semipermissive MB-02 cell line as acontrol source of viral transcripts. In order to minimizethe risk of false positive RT-PCR products arising frompotential contaminating viral DNA in RNA isolates, prim-ers B19L2 and B19S2O were designed to span the viralfirst intron as based on published B19 transcriptionmaps. A third primer B19S2I was designed to anneal just

TABLE 1

Predicted B19 Amplicon Sizes

Template

First-round(RT-PCR)product

expected, bp

Second-round(seminested PCR)

productexpected, bp

nspliced pre-mRNA or DNA 1702 1686910 acceptor 199 183030 acceptor 79 63925 acceptor 184 168952 acceptor 157 141

Note. Predicted sizes of amplicons derived from B19 templatesescribed in the literature. All sizes are calculated based on usage of

he nt 406 splice donor for the viral first intron, and acceptors as noted.

SPLICING PATTERN IN PARVOVIR

upstream of B19S20 and be used with B19L2 for semin-ested reamplification of weak first-round products. t

Based on previously published results, a single splicedonor at nt 406 is used for this first viral intron, with somevariability observed in choice of splice acceptor. In per-missive cells, splice acceptors at nt 2030 and nt 1910 arereported, with that at nt 1910 being more commonly usedOzawa et al., 1987). In the nonpermissive COS-7 system,plice acceptors in the small viral mRNAs coding for the.5- and 11-kDa proteins have been reported at nt 1925nd nt 1952 in addition to nt 1910 and nt 2030 (St. Amand

et al., 1991). Our primer set was designed to amplify all ofthese alternative products. The expected first (RT-PCR)and second (seminested PCR) amplicon sizes for thesedifferent possible templates are shown in Table 1.

As a control, we also designed a set of primers toamplify the cellular retinoblastoma mRNA with primersspanning intron 7 of the gene. These primers were de-signed to have similar annealing temperature and prod-uct amplicon sizes as the B19 primers used in the study.As for the B19 system, a total of three primers (Rb1, Rb20,and Rb2I) was designed to be used in RT-PCR andsubsequent seminested PCR amplification of templates.The expected product sizes from mature spliced mes-sage or from DNA or unspliced pre-mRNA templates aregiven in Table 2.

Following infection of differentiated MB-02 cells withB19 and RNA isolation we performed RT-PCR and theseminested PCR as described on the RNA extracts. Asshown in Fig. 1, the RT-PCR for the cellular retinoblas-toma message proceeded cleanly and yielded a singleproduct at the expected size (lane 2). Results of a semi-nested second round reaction were also as predicted(data not shown; as the products of the RT-PCR wereclearly visible this was generally not conducted). RT-PCRfor B19 message (lane 4) yielded a single strong unpre-dicted product of approximately 115 bp, as well as a verymuch weaker mixture of poorly resolving bands in the150–200 bp size range corresponding to a mixture ofsome of the predicted amplicon products. Second-roundseminested PCR amplification of these RT-PCR productsretained all of these bands with the expected decreasein length of each product (lane 5). Successful second-round amplification indicated that all of these products,

TABLE 2

Predicted Retinoblastoma Amplicon Sizes

Template

First-round(RT-PCR)product

expected, bp

Second-round(seminested PCR)

productexpected, bp

nspliced pre-mRNA or DNA ;1800 ;1800pliced mRNA 175 121

2859 RESTRICTS ITS HOST RANGE

Note. Predicted sizes of amplicons derived from retinoblastomaemplates, based on data of McGee et al. (1989).

(sfw

¨ D-VEN

including the major unexpected product, were derivedfrom B19 sequence.

In order to determine if the unexpected product couldpotentially be amplified from some viral DNA speciessuch as defective interfering genomes, we attempted anidentical set of amplifications for B19 products usingviremic patients directly as the template. As shown inFig. 2, the RT-PCR (lane 2) only yielded one distinctproduct, which is the predicted 1.7 kb size for amplifica-tion with this primer set from viral DNA. No products inthe 100–200 bp size range are observed. Upon reampli-fication with the seminested PCR (lane 3), the full-lengthgenomic-derived product was not reamplified due to avery short extension time in the thermocycling (optimizedto favor short products). Only faint diffuse products werevisible in the second-round reaction with most beingapproximately 500 bp in length; these are likely to be a

FIG. 1. RT-PCR and seminested PCR on RNA extracts of B19-infectedMB-02 cells. Products of RT-PCR and seminested second-round PCRon RNA extracts of B19-infected differentiated MB-02 cells as analyzedby 2% agarose gel electrophoresis and ethidium bromide staining of10-ml reaction samples. Left panel: RT-PCR for retinoblastoma mRNALane 2). Right panel: RT-PCR for B19 mRNA (Lane 4) and subsequenteminested PCR (Lane 5); Lane 6 is product of the second-round PCR

or B19 without template added. Lanes 1 and 3 are 100-bp DNA ladderith sizes as marked.

FIG. 2. RT-PCR and PCR results on viremic serum-derived nucleicacids. RT-PCR amplification (Lane 2) and seminested second-roundPCR (Lane 3) on serum control, analyzed by 2% agarose gel electro-phoresis and ethidium bromide staining of 10-ml reaction samples.

286 BRUNSTEIN, SODERLUN

Lane 4 is a PCR on same template using p8f/p5r primer set; Lane 5 isa p8f/p5r no template (2) control. Lanes 1 is a 100-bp ladder.

range of single-stranded products arising from nonexpo-nential extension of both first- and second-round primersets off B19 DNA present in the sample. In contrast,amplification of an identical template sample with primerset p8f/p5r under conditions very similar to those used inthe second-round reactions yielded a single sharp prod-uct band of expected size (393 bp, lane 4) and demon-strated the presence of amplifiable B19 DNA in the sam-ple.

To further examine whether the unexpected majorproduct was amplified from an RNA template, we pre-treated samples of RNA from infected MB-02 cell culturewith RNase A prior to RT-PCR amplification. As shown inFig. 3, the RNase predigestion completely abolished am-plification of both the retinoblastoma control target andall B19 targets. Based on this and the previous experi-ment, we concluded that the template for the unexpectedproduct was in fact an RNA species apparently of viralorigin.

Prior studies which generated the available viral tran-scription maps had demonstrated some variability in theacceptor site for the splicing of the viral first intron whichour amplicon should span. A total of four acceptors hasbeen observed in all cell types, at nts 1910, 1925, 1952,and 2030, and in all cases the acceptor is a canonical AGsplice acceptor dinucleotide. As this variability suggestssome lack of precision in recognition of this splice ac-ceptor by cellular machinery, we examined the viral se-quence between nt 1910 and nt 2030 for other potentialAG splice acceptor sites whose usage might result in amessage serving as template for our observed majorproduct. A total of 13 AG dinucleotides was found in thisregion in the B19-Au reference sequence (GenBank Ac-cession Number M13178), including the four already

FIG. 3. Effect of RNase A treatment of templates prior to RT-PCR. RNAextracts from infected MB-02 cells were either pretreated with RNaseA (Lane 3) or left untreated (Lane 2) prior to RT-PCR amplification forB19 (upper panel) or retinoblastoma (lower panel) mRNAs. Productamplification in both cases is only observed in the absence of RNase

ERMO, AND HEDMAN

digestion. Results shown are of agarose gel electrophoresis andethidium bromide staining of 10-ml reaction samples.

icaf

fn

ading

known to act as splice acceptors. In order to examinewhether our product arose from a message using one ofthese sites, we gel-isolated our observed major RT-PCRproduct and performed DNA sequencing of it. As shownin Fig. 4A, results of this demonstrate that the product infact had arisen from an RNA species in which the firstintron is indeed spliced in a novel manner; however,contrary to our preliminary expectations, it is the splicedonor site (at nt 441) and not the splice acceptor (at nt2030) which differs from published transcription maps.Based on these observed splice boundaries, the ex-pected size of our first-round (RT-PCR) amplicon is 114bp and that of the second round (seminested PCR) is 98bp; these values agree well with their observed sizes inagarose gel electrophoresis (see Fig. 1, lanes 4 and 5).

Based on the Au reference sequence for B19 (Shade etal., 1986), the splice donor in this case is an unusual CTbase pair. In order to check whether this held true for thevirus present in our infectious serum, the majority of theviral genome was cloned from the same serum sampleused for infecting cells in previous experiments (seeMaterials and Methods). We sequenced the genomicregion in question and found that our viral genomeagreed with the Au reference for these nucleotides; thesplice donor for our observed major intron is thus a “CT”dinucleotide. Over the total region we sequenced (.600

FIG. 4. Partial sequence of the major amplified product and its translarom B19 RT-PCR as described. Nucleotide numbers above the sequucleotide numbering of Shade et al. (1986). Inferred translational produ

sequence. (B) Sequence of the genomic splice donor and splice accejunction is indicated by “/”. (C) Full translational potential of the open-reof the splice junctions observed in this study.

SPLICING PATTERN IN PARVOVIR

nt; submitted to the GenBank nucleotide sequence da-tabase and assigned Accession Number AF272985) we

nm

observed better than 98% identity with other B19 se-quences in GenBank, suggesting that our infectious se-rum does not represent an unusual strain.

DISCUSSION

Human parvovirus B19 has a remarkably narrow hostrange limited to a subset of erythroid lineage cells in-cluding burst-forming and colony-forming erythroid pro-genitors from bone marrow, fetal liver, chronic myeloge-nous leukemia, erythroleukemia, and peripheral blood(for review, see Astell et al., 1997). While all of these canbe used as primary cultures for virus propagation in thelaboratory, they require addition of expensive growthfactors and are inconvenient to handle. Efforts to identifya continuous cell line culture system have met with littlesuccess (Brown et al., 1994); the megakaryoblastoidlines MB-02 and UT-7 have both been shown to havevery limited capacity to support viral replication in thepresence of growth factors. To date these are the mostconvenient system for viral culture; however, a very re-cent report (Miyagawa et al., 1999) suggests that a newlysolated cell line, KU812Ep6, may afford an improvedulture system. A major cause of restricted viral tropismppears to be an inability of the viral p6 promoter to

unction in nonerythroid cells (Kurpad et al., 1999; Pon-

otential. (A) Sequence data obtained from the major amplified productindicate the position of specific nucleotides relative to the genomicthis part of the sequence are given by one-letter code beneath the DNAes. Twenty nucleotides to either side of the junction are given; spliceframe created by fusion of the NS-1 ORF to the 7.5-kDa ORF by usage

2879 RESTRICTS ITS HOST RANGE

tional pence

cts forptor sit

US B1

azhagen et al., 1995). Other similar cell lines whichight be expected to support viral growth, such as K-562

onor ar served

2 ¨

cells with or without treatment to induce erythroid differ-entiation, do not allow for detectable productive viralreplication (M. Soderlund-Venermo and J. Brunstein, un-published observations), despite apparently normal ca-pacity to support expression of the viral p6 promoter(Gareus et al., 1998). On the basis of observed differ-ences in accumulation of structural versus nonstructuraltranscripts in permissive and nonpermissive cell types,Liu et al. have suggested that differences in active 39processing of viral pre-mRNAs may account in part forviral tropism, with nonpermissive cell types overproduc-ing message for the nonstructural gene relative to struc-tural gene mRNAs (Liu et al., 1992). As there is goodevidence to suggest the nonstructural gene productNS-1 is cytotoxic (Moffat et al., 1998), this is proposed tocause cell death before productive replication has oc-curred. While such a model has merits, it is somewhatunsatisfying as it predicts that at least low levels of viralstructural protein should be expressed and suggeststhat correspondingly low levels of infectious progenyvirions should be liberated. More recently, it has beenshown that the 39 untranslated region of spliced viralmessages is capable of inhibiting its own translation insome nonpermissive cell types (HeLa, COS-7) by block-ing ribosome loading (Pallier et al., 1997), although thebasis for this has not been elucidated. Thus, to date noconvincing mechanism has been put forward to accountfor B19’s limited tropism and its lack of replication in cellswhich support viral entry and expression.

We report here on studies conducted in the MB-02 cell

FIG. 5. Previous and proposed variant transcriptional maps. Multiple lintron sequences, and solid thick lines indicate open-reading frames ofamilies after Ozawa et al. (1987) and St. Amand et al. (1991), with depresentation of the same transcripts as resulting from the splicing ob

88 BRUNSTEIN, SODERLUN

line. Prior studies have demonstrated that following dif-ferentiation in growth-factor-supplemented media, this

cell line produces viral mRNAs (as shown by Northernblot) and low levels of infectious progeny virions asdemonstrated by their ability to be passaged and infectnormal bone marrow cells (Munshi et al., 1993). Whileattempting to use this culture system as a source ofstandard B19 mRNAs, we have observed that in viralmessages which incorporate a spliced first intron (thatis, all messages other than that coding for the viralnonstructural protein) the vast majority of such splicesutilize a novel splice donor site at nt 441 in conjunctionwith the previously known major acceptor at nt 2030, andthereby alter the viral transcriptional map. Examination ofthe coding potential of these altered messages showsthat this splicing pattern fuses the initiating and secondcodon of the nonstructural protein gene in frame with the32 C-terminal residues of the poorly characterized 7.5-kDa protein ORF (see Fig. 4B). In doing so it makes allspliced viral messages bicistronic, with this short NS-1/7.5-kDa fusion ORF being the 59 member of the pair(Fig. 5).

Rather than producing a single genomic mRNA spe-cies and translating all protein products from that viausage of an IRES, or a polyprotein expression scheme asused by some other viruses, evidence to date suggeststhat B19 uses differential splicing to generate monocis-tronic mRNAs for each of its expressed proteins as itsonly protein expression tactic. The levels of each specificmRNA, as determined by differential splicing, in turncontrol the levels of the different protein products. Someevidence in the literature exists that in the case of the

icates noncoding regions of mature mRNA, thin dashed lines indicatese mRNA; “mu” is map units. (A) Simplified schematic of B19 transcriptnd acceptor sites for the first viral intron as marked. (B) Equivalent

in this report. The 39 terminii of these messages are inferred from A.

ERMO, AND HEDMAN

ines indf matur

D-VEN

viral 11-kDa message, three very closely spaced AUGcodons in the same frame at the start of the message

1bbVcdf

d(hawttsugrr

MH

US B1

may each act as an initiation codon (St. Amand et al.,991); however, aside from that example there appears toe no evidence for initiation of translation from anythingut the first in-frame AUG of viral mRNAs. In particular,P-1 and VP-2 are coded for in the same frame with VP-1ontaining all of VP-2 plus additional N-terminal resi-ues; yet translation of VP-2 does not appear to occur

rom VP-1 mRNA (Kajigaya et al., 1991), and insteadrequires messages in which the VP-1 initiation codonhas been lost by a splicing event. This observation wouldstrongly suggest that B19 messages, like the overwhelm-ing majority of cellular messages, only express their firstfunctional open-reading frame. If this holds true, thenusage of the splice site as reported here might be ex-pected to effectively block translation of the viral capsidproteins as well as the 11-kDa protein. Such a resultwould come about if the small first fusion ORF wereefficiently recognized by ribosomal machinery. Given thatthis ORF utilizes the ribosome binding site and initiationcodon of the well expressed viral NS-1 protein, it wouldbe surprising if these same initiation signals were notrecognized in their new fusion context.

Based on these observations, we propose a modelwhereby in some nonpermissive cell types, host RNAsplicing machinery does not recognize the previouslyreported canonical GT splice donor at nt 406 but insteadchooses an unusual CT donor at nt 441. Usage of this inconjunction with the nt 2030 acceptor as previously iden-tified in the majority of spliced viral transcripts function-ally inactivates the viral structural genes and effectivelyblocks production of progeny virions. As examined byNorthern blot analysis, all viral transcript families wouldbe indistinguishable from those of a permissive cell line,as the total difference in length between these aberrantlyspliced messages and the productive ones is 35 nucle-otides and would not be differentiated from transcriptsize variation due to differences in polyadenylation. In-triguingly, there exists in the literature very strong evi-dence in support of this hypothesis. In particular, Leruezet al. showed that in the UT7/EPO megakaryoblastic cellline (similar to the MB-02 cell line examined here) B19can enter cells and that as detected by hybridization, allviral message families are produced as a consequence;however, while NS-1 protein is expressed, viral capsidgene products were not detectable in this study (Leruezet al., 1994).

In our study, we do observe a very small number ofviral RT-PCR products of the sizes expected from theprevious transcription maps (Fig. 1, lanes 4 and 5). Wesuspect that this accounts for the “semipermissivity” ofthe MB-02 cell line, as a small percentage of primarytranscripts are processed as expected and can give riseto functional capsid messages. We do not believe thedifference in band intensities are due to preferential

SPLICING PATTERN IN PARVOVIR

amplification of our product due to its length, as the sizedifferences are very small and one of the predicted

1c

products of previously reported splice junctions isshorter than our observed major product (see Table 1,2030 acceptor).

Our observation of a CT splice donor is uncommon,being to our knowledge only the second report of such adinucleotide functioning in this manner (French et al.,1999) and the first report of a CT-AG acceptor/donor pair.Comparison of the flanking sequences of our observedsplice donor (AG/CTATTT) with that of the common spli-ceosomal U2 target of (AG/GTRAGT), the less commonU12-type introns (/ATATCTT), and that of the only previ-ously published report of a CT donor (AC/CTGGGA)[where R is any purine, /indicates the splice site, andinvariant residues are indicated in boldface; DNA se-quence corresponding to RNA of message given](French et al., 1999; Stephens et al., 1992) shows a poormatch to any of them and suggests that a novel splicingfactor or pathway may be involved. Interestingly, thesequence flanking the nt 406 splice donor previouslyreported for this intron (AG/GTATTT) varies from our ob-served donor site only at the 11 position relative to thesplice site. This conservation suggests that in bothcases B19 may be utilizing an uncharacterized splicingmotif; whether this requires specific host factors and/orcis-acting elements remote from the splice junction issomething which further studies will be required to re-solve. An intriguing possibility of using B19 with its smallgenomic size and genetic simplicity to study this unusualsplicing suggests itself.

If our proposed mechanism is indeed a determiningfactor for viral tropism, it clearly cannot explain B19’sinability to replicate in the nonpermissive cells systemswhere the mRNA splice junctions were previouslymapped; in these cases the block to viral replicationmust arise via a mechanism such as the interference ofcapsid protein message’s 39 UTR with translation as

emonstrated by Pallier et al. (1997). As these cell typesCOS-7, HeLa) are however more distant from the nativeost cell of the virus than are such erythroid lineage cellss MB-02, K562, and UT7, we believe that the mechanisme propose here is the crucial one in restricting viral

ropism among bone-marrow-derived cells. Experimentso further validate our findings by examination for expres-ion of the putative 34-residue NS-1/7.5-kDa fusion prod-ct at the expense of capsid proteins are currently on-oing and should help to resolve outstanding questions

egarding the influence of different mechanisms for host-ange restriction in B19.

MATERIALS AND METHODS

Cell culture. The human megakaryoblastoid cell lineB-02 (Munshi et al., 1993) (a gift of Doris Morgan,ahnemann University, U.S.A.) was cultured in RPMI

2899 RESTRICTS ITS HOST RANGE

640 supplemented with glutamine, penicillin, streptomy-in, gentamicin (Sigma-Aldrich Co., St. Louis, MO; 20

mbhMa

i

¨ D-VEN

mg/ml), GM-CSF (Boehringer Mannheim Biochemica,Mannheim Germany; 200 U/ml), and 10% fetal calf serum(GIBCO BRL, Gaithersburg, MD) at 37°C and 5% CO2.

Viral infection and RNA isolation. Infectious virus wasin the form of serum from a B19-infected patient duringthe viremic phase (a gift of Bernard Cohen, PHLS, Colin-dale, UK) and containing approximately 1013 genomes/

l. MB-02 cells were induced to erythroid differentiationy addition of Erythropoietin (Epo, Boehringer Mann-eim, 4 U/ml) and stem cell factor (SCF, Boehringerannheim, 25 ng/ml) to the culture media 6 days prior to

ddition of 0.02–0.1 ml infectious virus (approximatem.o.i. 0.1–0.5). No significant differences were observedbetween results obtained from infections done with dif-ferent viral doses. Infections were allowed to proceed for4 days prior to isolation of total culture RNA by Trizolreagent (GIBCO BRL) in accordance with the supplier’sinstructions. RNA isolates were resuspended in a finalvolume of 30 ml per each 1.5 ml original cell culture.

RT-PCR and PCR. RT-PCR was performed with theTitan one-tube system (Boehringer Mannheim) as persupplier’s instructions, using 5 ml of RNA extract astemplate per reaction. For B19 transcripts, the primer pairB19L2 (59-GCTGTTTTTTGTGAGCTAACTAAC-39) andB19S2O (59-ACTGGTCCCGGGGATG-39) were used andcycled as follows: 42°C for 30 min, 94°C for 2 min, 10cycles of (94°C for 30 s, 46°C for 30 s, and 68°C for 45 s),25 cycles of (94°C for 30 s, 46°C for 30 s, and 68°C for50 s), 68°C for 3 min, and then held at 15°C prior tosubsequent processing. For retinoblastoma transcripts,the primer pair Rb1 (59-CTCACCTCCCATGTTGCTCA-39)and Rb2O (59-CTCATCTATATTACATTCATGTTCTTTAC-39)were used and cycled as for the B19 messages exceptthat all annealing steps (46°C, above) were conducted at51°C.

Seminested second-round PCR amplifications of theproducts of the RT-PCR were performed with AmplitaqGold polymerase (Perkin-Elmer, Branchburg, NJ) as perthe supplier’s general instructions. Briefly, for B19 prod-ucts, 2 ml of B19 RT-PCR product was used as templaten a 50-ml reaction containing 1X supplier’s buffer, 200mM dNTPs, 0.5 mM each of primers B19L2 and B19S2I(59-GGCGTACTAGAGCGCGG-39), and 1.25 U of polymer-ase. Reactions were thermocycled as follows: 95°C for10 min, 40 cycles of (95°C for 30 s, 49°C for 30 s, 72°Cfor 20 s), cooled to 15°C for 1 min, and then allowed torest at ambient temperature prior to analysis. For retino-blastoma products, the same protocol was followed withprimers Rb1 and Rb2I (59-ATTTTCTAGTTGTTTTGC-TATCCG-39) being used in place of the B19-specific prim-ers.

B19 DNA in viremic serum extracts was detected byPCR using Amplitaq Gold polymerase and primers p8f(59-TGTGCTTACCTGTCTGGATTG-39) and p5r (59-AGGCT-

290 BRUNSTEIN, SODERLUN

TGTGTAAGTCTTCAC-39) (Soderlund et al., 1997). Vol-umes and reagent concentrations were as per the semi-

nested PCR above, and thermocycling was performed as95°C for 10 min, 40 cycles of (94°C for 15 s, 55°C for 30 s,72°C for 40 s), 72°C for 5 min, and then held at 4°Cbefore further processing.

Serum to be used as template directly for PCR andRT-PCR was prepared by heating 5-ml samples of serumdiluted 1/100 in TE, pH 8, to 94°C for 5 min prior toaddition to PCR or RT-PCR.

DNA sequencing. DNA sequencing was conducted byan in-house sequencing facility (Haartman Institute, Uni-versity of Helsinki), utilizing ABI instruments and dye-terminator chemistry.

RNase A digestions. Five-microliter aliquots of totalcellular RNA extract were mixed with four ml of TE, pH8.0, buffer and 1 ml (2.8 Kuntz units) of DNAse-free RNaseA (Sigma-Aldrich) and allowed to digest for 30 min atroom temperature. Control RNA samples for these ex-periments were mixed with 5 ml of TE, pH 8.0, andallowed to sit at room temperature for 30 min, without theaddition of RNase A.

Direct cloning of B19 partial genome. The majority ofthe B19 genome present in our infectious serum wasdirectly cloned briefly as follows. One hundred microli-ters of the infectious serum was mixed with proteinase Kbuffer and subjected to overnight proteinase K digestion.Liberated sample DNA was purified with the Prep-A-Gene kit (BioRad Laboratories, Hercules, CA), and posi-tive and negative polarity viral genomes were allowed toanneal at room temperature. This DNA isolate was di-gested with BssHII (New England Biolabs, Beverly, MA)and ligated to BssHII-digested, dephosphorylated Litmus29 vector (New England Biolabs). Ligation products weretransformed into competent Escherichia coli (strainDH5a) and transformants selected on media containingampicillin. Resulting clones containing the B19 genomeminus the extreme hairpin termini were selected basedon restriction digest and partial sequence analysis.

ACKNOWLEDGMENTS

This study was supported by grants from the Center for InternationalMobility, the Finnish Science Academy, the Finnish Technology Ad-vancement Fund, and the Helsinki University Central Hospital Re-search and Education Fund. The authors express their very sinceregratitude to Drs. Doris Morgan and Bernard Cohen for their generousgifts of the MB-02 cell line and infectious patient serum, respectively,without which this study could not have been conducted.

REFERENCES

Anderson, M. J., Jones, S. E., Fisher-Hoch, S. P., Lewis, E., Hall, S. M.,Bartlett, C. L. R., Cohen, B. J., Mortimer, P. P., and Pereira, M. S. (1983).Human parvovirus, the cause of erythema infectiosum (fifth dis-ease)? Lancet i, 1378.

Astell, C. R., Luo, W., Brunstein, J., and St. Amand, J. (1997). B19parvovirus: Biochemical and molecular features. In “Monographs inVirology” (L. J. Anderson and N. Young, Eds.), Vol. 20, pp. 16–41.

ERMO, AND HEDMAN

Karger, Switzerland.Brown, K., Young, N., and Liu, J. (1994). Molecular, cellular, and clinical

F

K

K

K

L

L

L

M

M

M

M

M

O

O

P

P

S

S

S

S

S

S

S

S

S

US B1

aspects of parvovirus B19 infection. Crit. Rev. Oncol. Hematol. 16,1–31.

Brown, T. A., Anand, A., Ritchie, L. D., Clewley, J. P., and Reid, T. M. S.(1984). Intrauterine parvovirus infection associated with hydrops fe-talis. Lancet ii, 1033–1034.

rench, C., Menegazzi, P., Nicholson, L., Macaulay, H., DiLuca, D., andGompels, U. (1999). Novel, nonconsensus cellular splicing regulatesexpression of a gene encoding a chemokine-like protein that showshigh variation as is specific for human herpesvirus 6. Virology 262,139–151.

Gareus, R., Gigler, A., Hemauer, A., Leruez-Ville, M., Morinet, F., Wolf, H.,and Modrow, S. (1998). Characterization of cis-acting and NS1 pro-tein-responsive elements in the p6 promoter of parvovirus B19.J. Virol. 72, 609–616.

ajigaya, S., Fujii, H., Field, A., Anderson, S., Rosenfeld, S., Anderson, L.,Shimada, T., and Young, N. (1991). Self-assembled B19 parvoviruscapsids, produced in a baculovirus system, are antigenically andimmunogenically similar to native virions. Proc. Natl. Acad. Sci. USA88, 4646–4650.

nott, P. D., Welply, G. A. C., and Anderson, M. J. (1984). Serologicallyproven intrauterine infection with parvovirus. Br. Med. J. 289, 1660.

urpad, C., Mukherjee, P., Wang, X. S., Ponnazhagen, S., Li, L., Yoder,M. C., and Srivastava, A. (1999). Adeno-associated virus 2-mediatedtransduction and erythroid lineage-restricted expression from parvo-virus B19p6 promoter in primary human hematopoietic progenitorcells. J. Hematother. Stem Cell Res. 8, 585–592.

eruez, M., Pallier, C., Vassias, I., Elouet, J., Romeo, P., and Morinet, F.(1994). Differential transcription, without replication, of nonstructuraland structural genes of human parvovirus B19 in the UT7/EPO cellline as demonstrated by in situ hybridization. J. Gen. Virol. 75, 1475–1478.

iu, J., Green, S., Shimada, T., and Young, N. (1992). A block in full-lengthtranscript maturation in cells nonpermissive for B19 parvovirus. J. Vi-rol. 66, 4686–4692.

uo, W., and Astell, C. R. (1993). A novel protein encoded by small RNAsof parvovirus B19. Virology 195, 448–455.cGee, T., Yandell, D., and Dryja, T. (1989). Structure and partialgenomic sequence of the human retinoblastoma susceptibility gene.Gene 80, 119–128.iyagawa, E., Yoshida, T., Takahashi, H., Yamaguchi, K., Nagano, T.,Kiriyama, Y., Okochi, K., and Sato, H. (1999). Infection of the erythroidcell line, KU812Ep6 with human parvovirus B19 and its application totitration of B19 infectivity. J. Virol. Methods 83, 45–54.offat, S., Nobuo, Y., Kohtaro, T., Noboyuki, T., and Sugamura, K. (1998).Human parvovirus B19 nonstructural (NS1) protein induces apopto-sis in erythroid lineage cells. J. Virol. 72, 3018–3028.ortimer, P. P., Humphries, R. K., Moore, J. G., Purcell, R. H., and Young,N. S. (1983). A human parvovirus-like virus inhibits haematopoietio

SPLICING PATTERN IN PARVOVIR

colony formation in vitro. Nature 302, 426–429.unshi, N., Zhou, S., Woody, M., Morgan, D., and Srivastava, A. (1993).

W

Successful replication of parvovirus B19 in the human megakaryo-cytic cell line MB-02. J. Virol. 67, 562–566.

zawa, K., Ayub, J., Yu-Shu, H., Kurtzman, G., Shimada, T., and Young, N.(1987). Novel transcription map for the B19 (human) pathogenicparvovirus. J. Virol. 61, 2395–2406.

zawa, K., Kurtzman, G., and Young, N. (1986). Replication of the B19parvovirus in human bone marrow cultures. Science 233, 883–886.

allier, C., Greco, A., Le Junter, J., Saib, A., Vassias, I., and Morinet, F.(1997). The 39 untranslated region of the B19 parvovirus capsidprotein mRNAs inhibits its own mRNA translation in nonpermissivecells. J. Virol. 71, 9482–9489.

onnazhagen, S., Woody, M., Wang, X-S., Zhou, S., and Srivastava, A.(1995). Transcriptional transactivation of parvovirus B19 promoters innonpermissive human cells by adenovirus type-2. J. Virol. 69, 8096–8101.

aarinen, U. M., Chorba, T. L., Tattersall, P., Young, N. S., Anderson, L. J.,Palmer, E., and Coccia, P. F. (1986). Human parvovirus B19-inducedepidemic acute red cell aplasia in patients with hereditary hemolyticanemia. Blood 67, 1411–1417.

hade, R. O., Blundell, M. C., Cotmore, S. F., Tattersall, P., and Astell,C. R. (1986). Nucleotide sequence and genome organization of hu-man parvovirus B19 isolated from the serum of a child during aplasticcrisis. J. Virol. 58, 921–936.

himomura, S., Komatsu, N., Frickhofen, N., Anderson, S., Kajigaya, S.,and Young, N. (1992). First continuous propagation of B19 parvovirusin a cell line. Blood 79, 18–24.

osa, C., Mahony, J., Luinstra, K., Sternbach, M., and Chernesky, M.(1992). Replication and cytopathology of human parvovirus B19 inhuman umbilical cord blood erythroid progenitor cells. J. Med. Virol.36, 125–130.

rivastava, A., and Lu, L. (1988). Replication of B19 parvovirus in highlyenriched hematopoietic progenitor cells from normal human bonemarrow. J. Virol. 62, 3059–3063.

rivastava, C., Zhou, S., Munshi, N., and Srivastava, A. (1992). Parvovi-rus B19 replication in human umbilical cord blood cells. Virology 189,456–461.

t. Amand, J., Beard, C., Humphries, K., and Astell, C. R. (1991). Analysisof splice junctions and in vitro and in vivo translational potential ofthe small, abundant B19 parvovirus RNAs. Virology 183, 133–142.

tephens, M., and Schneider, T. (1992). Features of spliceosome evo-lution and function inferred from an analysis of the information athuman splice sites. J. Mol. Biol. 228, 1124–1136.

oderlund, M., Ruutu, P., Ruutu, T., Asikainen, K., Franssila, R., andHedman, K. (1997). Primary and secondary infections by humanparvovirus B19 following bone marrow transplantation: Characteriza-tion by PCR and B-cell molecular immunology. Scand. J. Infect. Dis.29, 129–135.

2919 RESTRICTS ITS HOST RANGE

hite, D. G., Woolf, A. D., Mortimer, P. P., Cohen, B. J., Blake, R., andBacon, P. A. (1985). Human parvovirus arthropathy. Lancet i, 419–421.