1%»! .5”. h Wa_nj§m.m_$1‘ L. . : .9... .5 area. .. _ . , M144} MES... . . ‘ .. V. y $135 at 439 .11. 6",. _ . .. V. . 1. Jam? .m 7 . 5.3. than . 1nd $51.53. “N a. b.3521. ii. .v. ban.» 4... tan. “VI. in .40... .V .3. i. .I.) 1.13? s kniamwlu . 4.. , . “hp!!! ; l? ‘ . . 9.: V v...‘ F . . . vnurfaanAh-uluhvi I! a!!! evarhtkL-Rnll .L . V .1? 1421”“me . )7 . .. ‘ Esp-3.! i. r. . . J1}: .. . . , , V. . V . . a C! nvAr‘toinmehvvau .r . . . . . ”(an V3913. ..f 1 451...... .5... . x , V . . n. *5... 3 . v. . , . . A 1 1 V ‘9 umrr, . 0'. 1:: I . aIJ Lu . . (.7 .uu 5.? ‘n . 1“, 9:3.me Haw... Bun?“ . - . . :5... ..uAu.....a$. ‘ L i 1w??? . V . .. . .. : i . .. :5 :2 33nd,} N... . . . . A i .13. V IV . .. W:J$..U§,§.w .35.. V ‘ V , .. . V . , . .. ..... 2.2.5,: 3.... .. , .. . .. . V .... .V V . , V. V, . x s .m'fixfllbpny .: , V . , _ V , _ V . 5.. Iié...,vu. - 5.5 . V . .. V 416'... V t V . .l...r__uys£x.ny.. Titan“... . . V .er 4n! . .2 ‘ . .. n . . . w i . Ida)“, .! nail». g1 .. u 7rbi9i «.6421; V 2‘ l , BJIL. . .ifl a. 1,.le ind. v 3 ft. is , 1.51m... ‘15:)..5'3 V0.7 i... i: .21.! 3;». I; f: ., i... v. . . U10 . . . [1.1 . 5.11 I A 157. .1! r... 51)}; . V . .o .. .91.)... 1 u»!§..AsVlALH .{sl ‘ .u )4! 3 «3; #9112 1139. ..r. I ‘ ’tos-vufi’l ‘ 151‘ tl"~rl(,‘ltn.‘n 22-1.11Iu “1.1 My LL. .L,v .1 , K: t'i' Ark: titll! l4). ... an»... n. O . “gggmmgfigflfifiéfii THESIS :30 i. V' r C? k; ||l||||l|llll||llllllllllllllllIlllllllllllllllllllllllll 1 3 1293 02050 5“ fr“ T32 “rev ..... ~ 1 a .La' :- I h m .3, a 2M? aw it“ a- .fi: * I», n, ' IEUH: U55 .7 \.-t‘" Unswersatv [ This is to certify that the dissertation entitled RNA Recombination between Viral Transgenes and Infecting Viruses presented by Min Deng has been accepted towards fulfillment of the requirements for Ph. D. , Plant Pathology degree in W¥m Major professor Date 9/21/99 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE moo www.mu RNA RECOMBINATION BETWEEN VIRAL TRANSGENES AND INFECTING VIRUSES By Min Deng A DISSERTATION Submitted to Michigan State University in partial fiilfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Botany and Plant Pathology 1999 ABSTRACT RNA RECOMBINATION BETWEEN VIRAL TRANSGENES AND INFECTING VIRUSES Min Deng RNA recombination is a natural process that occurs during viral RNA replication, and it contributes to RNA virus evolution. In transgenic plants expressing viral genes, RNA recombination between viral transgenes and infecting viruses could generate chimeric viruses with distinct properties. Using the bromovirus model system, Greene and Allison (1994) demonstrated that a bromovirus capsid transgene with a complete 3’ untranslated region (UTR) was involved in RNA recombination with an infecting virus and the recombinant recovery rate was 3%. A decrease in recombinant recovery rate was found when the 3’ UTR was deleted (Greene and Allison, 1996). Experiments carried out in this study were built upon these two studies and attempted to expand our understanding of RNA recombination in transgenic plants expressing viral genes. These experiments have demonstrated that 1) a complete 3’ UTR in the viral transgene allowed the synthesis of minus strand copies of the viral transgene, and thus permitted RNA recombination to happen more frequently, presumably during both plus and minus strand synthesis; 2) a polymerase transgene could also be involved in RNA recombination with an infecting virus, and the recombinant recovery rate (4%) of the polymerase transgene is comparable with that of the capsid transgene; 3) a translatable transgene is less likely to be involved in RNA recombination, especially if the transgene provides resistance to the infecting virus; 4) truncated bromovirus capsid and polymerase transgenes provided virus resistance to transgenic plants. Such smaller viral transgenes may be more desirable than complete genes as they should less likely contribute to viral evolution through RNA recombination. Copyright by MIN DENG 1 999 ACKNOWLEDGMENTS I would like to thank my major professor Dr. Richard Allison for his support and guidance during the last four years. I would like to acknowledge my guidance committee members Dr. Gustaaf de Zoeten, Dr. Rebecca Grumet, and Dr. Suzanne Thiem for challenge me to become a better scientist. I would like to say thanks to Dr. Barbara Sears for helping me to change my career to molecular sciences. I would like to thank Dr. Jihad Skaf, Dr. Heng Zhong, and Dr. Sergei Mekhedov for valuable discussions and help on lab techniques. I would like to say thanks to Jeanne, Al and Bill for their help. Thanks to current lab members Adam, David and Cheryl for their suggestions and corrections on my thesis. Thanks to Jan Szyren for her help on clonally propagating transgenic plants. I would like to thank Shirley Owens and Nancy Veenstra, who always care about me. I would like to thank MSU writing center, especially Cathy Lydia Fleck, for helping me with my thesis writing. Thank you, Cathy, for your encouragement and help. I would like to thank Dr. Aureal Cross and Dr. Ralph Taggart for their kindness to me. I would like to thank Haibo for his support and encouragement. Thanks to Joshua, who understand that Mommy is busy and has to work at such an early age. I am so proud of him. I would like to thank my parents, who always wanted me to reach high and do the best I could. TABLE OF CONTENTS List of Tables ........................................................................................ vii List of Figures ..................................................................................... viii Introduction ........................................................................................... 1 Chapter 1. Synthesis of minus strand copies of the viral transgene during viral infections of transgenic plants Introduction ........................................................................ 16 Materials and Methods ............................................................ 18 Results .............................................................................. 23 Discussion .......................................................................... 3 1 Chapter 2. RNA recombination between a brome mosaic virus deletion mutant and a polymerase transgene Introduction ..... ' ................................................................... 36 Materials and Methods ............................................................ 38 Results .............................................................................. 46 Discussion .......................................................................... 57 Chapter 3. RNA recombination in transgenic plants expressing translatable or untranslatable viral transgenes Introduction ........................................................................ 66 Materials and Methods ............................................................ 67 Results .............................................................................. 74 Discussion .......................................................................... 80 Chapter 4. Test of virus resistance in transgenic plants expressing bromovirus transgenes Introduction ........................................................................ 85 Materials and Methods ............................................................ 86 Results .............................................................................. 92 Discussion .......................................................................... 98 Summary and Conclusions ....................................................................... 102 Appendix ............................................................................................ 107 List of References ................................................................................. 1 18 vi LIST OF TABLES Chapter 3 Table 1. Test of virus resistance in transgenic plants .......................................... 78 Chapter 4 Table]. Test of virus resistance ................................................................... 97 vii LIST OF FIGURES Chapter 1 Figure 1. Plasmid construction .................................................................... 22 Figure 2. Dot blot assay quantifying the sensitivity of radiolabeled RA518(-) probe. . . ...25 Figure 3. Northern blot analysis of dsRNA ...................................................... 27 Figure 4. Agarose gel electrophoresis showing RT-PCR amplified negative-sense CCMV RN A3 .......................................................... 29 Figure 5. Agarose gel showing Noll treated RT-PCR products ............................... 30 Chapter 2 Figure 1. Plasmid construction .................................................................... 39 Figure 2. Detection of viral transgenic RNA transcripts in transgenic plants ............... 48 Figure 3. Test of BMV resistance in transgenic plants ......................................... 50 Figure 4. Northern blots showing the detection of recombinants ............................. 53 Figure 5. RT-PCR results of recombinants established in nontransgenic N. benthamiana ........................................................ 55 Figure 6. Sequence analysis of the recombinants ............................................... 58 Figure 7. Representative mutations recovered in the polymerase gene of recombinant viruses ................................................................ 59 Figure 8. Polymerase amino acid sequence of recovered recombinants ..................... 60 Chapter 3 Figure 1. Plasmid construction .................................................................... 69 Figure 2. Northern blot analysis ................................................................... 75 Figure 3. Schematic diagram of CCMV RNA3 and the position of the two primers ...... 79 Figure 4. Diagram showing mutations recovered in the RN A3 of recombinants ........... 82 Figure 5. Capsid protein amino acid sequence of recovered recombinants .................. 83 Chapter 4 Figure 1. Plasmid construction for plant transformation ....................................... 88 Figure 2. Northern blot analysis reflecting viral transgene transcription in transgenic plants ......................................................................... 93 Figure 3. Dot blot analysis of virus resistance in transgenic plants ........................... 96 Appendix Figure 1. Infection of cowpea with virus A69-1 ............................................... 1 11 Figure 2. Model for RNA recombination to generate recombinant A69-1 ................. 112 Figure 3. Nucleotide and amino acid changes in virus A69-1 ............................... 114 Figure 4. Plasmid construction for making infectious transcripts .......................... l 15 viii INTRODUCTION Since the advent of Agrobacterium-mediated transformation techniques to genetically modify tobacco plants (Matzke and Chilton, 1981), a variety of genetic modification technologies have been developed. This has enabled plant breeders to develop new lines of crops at a faster pace than was possible with traditional methods. A large range of plants, including cereals, vegetables, trees, and ornamental crops, was genetically modified for the purpose of introducing novel and economically important traits. For example, the Flavr SavrTM tomato contains an antisense copy of the polygalacturonase gene, which delays the ripening of the tomatoes (Kramer and Redenbaugh, 1994), and ZW-20 squash contains the coat protein genes from watermelon mosaic virus 2 (WMV2) and zucchini yellow mosaic virus (ZYMV), which provide resistance to these two viruses (Fuchs and Gonsalves, 1995; Tricoli et al., 1995). These two varieties, plus dozens of other genetically modified (GM) major crop species, have reached or are in the process of reaching the market. The appearance of GM food in the market gives rise to concerns and controversy. One major concern in the plant virology community is the risk associated with the viral transgenes, which includes transcapsidation and RNA recombination (de Zoeten, 1991; 1995). The following study assesses the possibility that all or part of a viral transgene may be usefully incorporated into a challenging virus through RNA recombination. Additionally, this study begins to elucidate the mechanism by which the transgenes become involved in viral RNA recombination events. Dynamic features of RNA virus population contributes to RNA virus evolution The genetic information of most plant viruses is stored in the form of single- stranded positive sense RNA. These plant RNA viruses encode their own replicative proteins which, through association with cellular proteins and membranes, form replication complexes which carry out the replication of the viral RNA genome (Kao et al., 1992; Restrepo-Hartwig and Ahlquist, 1996). The central part of this replication complex is the viral encoded RNA-dependent RNA polymerase (RdRp). RdRp has two unique characteristics which differentiate it from its counterpart in DNA-based organisms, the DNA-dependent DNA polymerase (Dde); these characteristics also contribute to the dynamic features of RNA virus population and the rapid evolutionary drifi of RNA viruses. One unique characteristic of RdRp is its high error rate. Error rates of RdRps have been estimated to be in the range of 10“ to 10 ‘5, whereas those of Ddes are between 10'8 to 10'9 (Drake, 1993). This is due to the lack of 3’—->5’ exonucleolytic proofreading activity in RdRp. Theoretically, for an RNA virus with a genome of 10 Kb, one error could be introduced during each replication cycle. Unlike the mutations in DNA-based organisms which may be repaired, these RNA mutations are incorporated into the genome of the progeny viruses during RNA replication. Thus, the population of an RNA virus is not genetically homogeneous, but represents a dynamic distribution which has been described as mutant swarms (Domingo et al., 1995). Another characteristic of RdRp is its ability to switch templates during the process of RNA replication, which is referred to as RNA recombination (for review, see Lai, 1992; 1995). RNA recombination has been observed frequently in animal RNA viruses, such as poliovirus and coronaviruses (Lai, 1992). The recombination fi'equency was 2% for a particular segment of the poliovirus genome (Cooper, 1968). For plant RNA viruses, RNA recombination was observed in members of several different virus groups, including bromoviruses (Bujarski and Kaesberg, 1986; Bujarski and Dzianott, 1991; Allison et al., 1990), turnip crinkle virus (Cascone et al., 1990), and alfalfa mosaic virus (Huisman et al., 1989; Van der Kuyl et al., 1991). RNA recombination is thought to be a three-step process, during which the replication complex first pauses at a strong secondary structure in the template RNA. Secondly, the nascent RNA-replication complex dissociates from the original template, and reassociates with a different RNA template. Following the polymerization, a chimeric progeny RNA, which results from copying two distinct templates, is created. This chimeric RNA is referred to as a recombinant (Lai, 1992; 1995). The generally favored model for RNA recombination is the copy-choice model, which proposes that the viral RNA polymerase switches templates during negative strand synthesis (Kirkegaard and Baltimore, 1986). RNA recombination is believed to play a major role in RNA virus evolution. Sequence similarities among RNA viruses and arrangements of viral genetic segments indicate that RNA viruses evolved through the exchange of functional modules between related and possibly unrelated viral RNAs. This evolutionary process, which is referred to as modular evolution of viruses (Koonin and Dolja, 1993; de Zoeten, 1995), is obviously dependent on RNA recombination. In addition to high polymerase error rates and RNA recombination, other mechanisms also reshape the RNA virus population. Gene segment reassortment and transcapsidation perturb the populations of multipartite RNA viruses (Domingo and Holland, 1994; de Zoeten, 1991). During host-to-host transmission, bottleneck events happen frequently in nature due to random sampling of viral genomes (Domingo et al., 1996). For a bottlenecked RNA virus p0pulation which has high mutation rate, deleterious mutations will accumulate in an irreversible manner. Repeated bottlenecked transmission of RNA viruses will fix sub0ptimal classes of genomes and decrease the viral fitness. Only compensatory mechanisms such as RNA recombination can restore the initial, mutation-free class of a genome (Domingo et al., 1996). Natural selection plays a major role in reshaping the perturbed RNA virus populations, since most of the mutations generated during the process of RNA replication and recombination result in such dramatic fitness loss that the variants fail to compete well with their wild type progenitors, and eventually disappear (Domingo et al., 1978; Duarte et al., 1994). This explains why, despite high mutation rates, RNA viruses maintain major consensus sequences that typify each virus (Domingo et al., 1985). Current studies on viral population genetics show that an RNA virus population is not a genetically defined population, but rather a dynamic genetic organization in which individual genomes differ in one or more positions from the consensus or average sequence of the population (Domingo and Holland, 1994). Thus, an RNA virus species is considered a quasispecies, rather than a true species consisting of a more homogeneous population (Duarte et a1, 1994). A striking feature of the RNA virus quasispecies is its ability to not only create variations, but also to tolerate them. The genetic flexibility of quasispecies provides a selective advantage, and allows RNA viruses to penetrate various types of host tissue under normal conditions (Morimoto et al., 1998), and upon fast environmental changes, to adapt to new ecological niches (Domingo and Holland, 1994). Thus, RNA viruses, including bromoviruses, are different from all other organisms. Bromoviruses Bromoviruses are among the most extensively studied RNA viruses, and they have been used as a model system to study viral RNA replication, encapsidation, movement, and RNA recombination. Bromoviruses include brome mosaic virus (BMV) and cowpea chlorotic mottle virus (CCMV), which are closely related and share extensive sequence similarities (Allison et al., 1989). The natural hosts of BMV are monocots such as barley (Hordeum vulgare L.), whereas those of CCMV are dicots, especially legumes. A common host for both viruses is Nicotiana benthamiana Domin (Lane, 1981). Both BMV and CCMV are single-stranded positive sense RNA viruses, which belong to the plant RNA virus family Bromoviridae. Bromoviruses have three genomic RNA3: RNA 1 (3.2 Kb), RNA 2 (2.9 Kb) and RNA 3 (2.1 Kb), and a subgenomic RNA 4 (0.8 Kb). Genomic RNA] and 2 are monocistronic and in BMV they encode for the 1a (109 KDa) and 2a (94 KDa) proteins, respectively, which together are sufficient to direct viral RNA replication in barley protoplasts (French et al., 1986; Kiberstis et al., 1981; Allison et al., 1988). The 1a protein contains two conserved domains within the N- terrninal and C-terminal regions. The N-terminal domain has sequence homology to the Sindbis virus nsPl methyltransferase protein, which is believed to be involved in capping the viral RNAs (Ahlquist et al., 1985; Mi and Stollar, 1991). The C-terminal portion of the la protein has homology to viral and cellular helicases, which typically unwind double-stranded nucleic acid molecules (Evans et al., 1985). The 2a protein, a putative RNA-dependent RNA polymerase (RdRp), is characterized by the GDD (Gly-Asp-Asp) motif which is present in all known RdRps. The RdRp of BMV has a central polymerase core region flanked by 5’ and 3’ regions. The central polymerase core region is highly conserved and is indispensable for RNA replication, whereas the 5’ and 3’ flanking regions are not required for replication (Traynor et al., 1991).; The la and 2a proteins, through association with cellular proteins and membranes, form a replication complex which is responsible for viral RNA replication (Kao et al., 1992; Restrepo-Hartwig and Ahlquist, 1996). .Genomic RNA 3 of bromoviruses is dicistronic and encodes for a putative movement protein 3a (32 KDa) and the coat protein (CP) (20 KDa). The 3a protein is translated directly from genomic RNA 3, whereas the coat protein is translated from the subgenomic RNA 4, which is derived from the minus-strand RNA 3 during replication (Allison et al., 1988, Ahlquist, 1994). Molecular and genetic studies revealed that, for both BMV and CCMV, the movement protein is critical for cell-to-cell movement (Mise et al., 1993; Misc and Ahlquist, 1995). However, despite the sequence similarities between these two viruses, BMV and CCMV appear to utilize different mechanisms to move from cell to cell. BMV requires a fimctional coat protein for cell to cell movement in addition to the movement protein (Schmitz and Rao, 1996), whereas CCMV moves between epidermal cells without a coat protein (Rao, 1997). The replication of bromoviruses, like most other positive sense RNA viruses, can be divided into three distinct steps: minus-sense RNA synthesis, positive-sense genomic RNA synthesis, and subgenomic RNA synthesis (Dinant et al., 1993). Each step is differentially regulated at various stages of viral infection. Evidence suggests that the 3’ untranslated region (UTR), which is conserved for all three genomic BMV RNAs, contains the replication initiation site for minus-sense RNA synthesis (Ahlquist et al., 1981a; Dreher and Hall, 1988). The viral replication complex first utilizes the positive- sense viral RNA as a template to make a small amount of minus-sense RNA, which serves as a template for synthesis of large quantities of positive-sense genomic RNA (Marsh et al., 1991b). The minus-sense copy of the genomic RNA 3 has an intercistronic region, which bears the promoter for synthesis of subgenomic RNA 4 (French and Ahlquist, 1987; Pogue et al., 1992). Minus-sense RNA is associated with viral positive-sense RNA in the cytoplasm. Upon extraction, they form a double-stranded viral RNA called replicative form, which is resistant to mild RNase digestion at high salt concentrations, but susceptible at low salt concentrations (Hardy et al., 1979). In BMV infected barley protoplasts, negative-strand RNA accumulation plateaus at 8 hours post inoculation (hpi), while positive-strand genomic RNA and subgenomic mRN A continue to accumulate until 20 hpi (Kroner et al., 1989). In the presence of all three genomic BMV RNAs, positive and negative strand RNA synthesis is asymmetric. The negative-sense RNA accumulates 100-fold lower than the positive sense RNA in BMV infected barley protoplasts at 24 hpi (Rao et al., 1990b; Marsh et al., 1991a). The quantity of negative-strand RNA in wt CCMV infections is even lower (Pacha and Ahlquist, 1991). The 3’ terminal sequences of BMV and CCMV are distinct from each other and are highly conserved within each virus. However, the predicted secondary structures of the 3’ terminal 200 bases of BMV and CCMV are strikingly similar (Ahlquist et al., 1981a). A genomic RNA reassortment experiment in N. benthamiana protoplasts indicated that the replication complex of BMV could replicate CCMV RNAs, and vice versa (Allison et al., 1988). The genomic RNA reassortment experiment also indicated that, using either BMV or CCMV replicative proteins, BMV genomic RNAs are preferentially replicated and the accumulation level of BMV RN As was higher than that of CCMV RNAs (Allison et al., 1988). Using an in vitro replication system of BMV, Miller et a1. (1986) found that the 3’ terminal 134-nt fragment, which folds into a tRNA-like structure, contains all of the signals required for initiation of minus-strand replication. Minus-strand initiation occurs opposite the penultimate 3'-nucleotide (the 3’ C of CCAOH), and the terminal adenosine is added to the plus-strand RNA afier replication through interaction of viral RNA and the host tRNA nucleotidyl transferase (Miller et al., 1986). Virus resistant transgenic plants Following the theoretic prediction of pathogen-derived resistance in 1985 (Sanford and Johnston), virus resistant transgenic plants were first generated in Roger Beachy’s laboratory in 1986 (Powell-Abel et al., 1986). In this experiment, tobacco plants were transformed with a cDNA copy of the coat protein (CP) gene of tobacco mosaic virus (TMV). These transgenic plants had an increased level of resistance to TMV virion inoculation. This phenomenon was named coat protein-mediated resistance (CPMR). CPMR has since been demonstrated for many other plant RNA viruses, including alfalfa mosaic virus (Loesch-Fries et al., 1987), cucumoviruses (Cuozzo et al., 1988), potyviruses (Lawson et al., 1990; Stark and Beachy, 1989), potexviruses (Hemenway et al., 1988), tobravirus (van Dun and B01, 1988), tospoviruses(Gie1en et al., 1991), and luteoviruses (Kawchuk et al., 1990; for review on CPMR, see Grumet, 1995). In a majority of the cases, virus accumulation is reduced or absent, in some cases however, the level of protection is related to the level of CP expression. The transgenic plants were protected against only the virus from which the CP gene was derived, with some exceptions where the transgenic plants were also protected against closely related viruses (Nejidat and Beachy, 1990). Currently, several coat protein-mediated virus resistant crops are on the market, and numerous others are being field tested prior to commercial release (Fuchs et al., 1997; Yeh et al., 1998). The ZW-20 squash mentioned above was the first commercialized CPMR crop. The theory of pathogen-derived resistance was also applied to several other non- structural viral protein genes, including the replicase gene and the movement protein gene. Replicase-mediated resistance was first demonstrated with TMV (Golemboski et al., 1990) and later in several other virus groups, including potexviruses (Braun and Hemenway, 1992; Longstaff et al., 1993), tobraviruses (MacFarlane and Davies, 1992), cucumoviruses (Carr et al., 1994; Zaitlin et a1, 1994), tombusviruses (Rubino and Russo, 1995), and potyviruses (Audy et al., 1994; Guo and Garcia, 1997). In many cases, the level of resistance provided by the replicase transgene is greater than that provided by the coat protein gene (Braun and Hemenway, 1992). Compared to CPMR the spectrum of replicase-mediated resistance was sometimes broader (Donson et al., 1993), with some exceptions where it was restricted (Rubino and Russo, 1995). In some cases, a translatable replicase transgene is required before virus resistance is observed (MacFarlane and Davies, 1992), while in other cases, the resistance is believed to be mediated by the transgenic mRN A only (Guo and Garcia, 1997). An interesting phenomenon associated with replicase-mediated resistance is that a higher level and/or a broader spectrum of resistance was observed when the replicase gene used for generating transgenic plants was modified or truncated. For example, a 10 truncated cucumber mosaic virus (CMV) replicase gene provided absolute resistance to CMV virion and RNA inoculum at concentrations up to 500 ug/ml or 50 ug/ml, respectively (Anderson et a1, 1992). In another study, tobacco plants transformed with the replicase gene of potato virus X (PVX) containing a modified GDD motif were extremely resistant to PVX in that no virus systemic movement was observed at inoculum concentrations as high as 250 ug/ml PVX RNA (Longstaff et al., 1993). Although the mechanism of pathogen-derived resistance remains unknown, many studies indicate several distinct mechanisms. Based on the observation that transgenic tobacco plants expressing TMV CP were more resistant to the TMV virion than to the TMV RNA inoculum, the inhibition of virion disassembly in the initially infected cells was thought to be responsible for CPMR (Register and Beachy, 1988). Specific interactions between the transgenic CP and the CP of the challenging virus as well as an interference with virus movement could also be necessary components of CPMR against TMV (Bendahmane et al., 1997) Replicase mediated resistance is likely RNA-based (Mueller et al., 1995), and post-transcriptional gene silencing may provide the mechanism (Baulcombe, 1996). Plants naturally use homologous or complementary nucleic acid sequences to control excess production of host mRNA or the replication of RNA pathogens (Matzke and Matzke, 1995). Overproduction of a transgene can trigger post-transcriptional gene silencing of both the transgene as well as the homologous host gene (Wassenegger and Pelissier, 1998). If the transgene is derived from part of the genome of an RNA virus, 11 plants show post-transcriptional silencing of the viral transgene as well as the RNA of infecting viruses (Lindbo et al., 1993). The consequence is that the accumulation of the virus in the cytoplasm is strongly reduced and plants display resistance against the corresponding virus. Risk assessment and the possibility of RNA recombination between viral transgene and infecting viruses Although pathogen-derived resistance appears to provide a promising way for combating virus infection in transgenic crops, the dynamic features of RNA viruses suggest that potential risks may be associated with the release of virus resistant transgenic plants in the environment. In 1991, de Zoeten proposed two possible risks of the release of transgenic plants expressing the viral CP gene, namely transcapsidation and RNA recombination. Transcapsidation occurs in mixed infections when the genomic RNA of one virus is encapsidated into the capsid of another virus, or a phenotypically mixed capsid with protein subunits from both viruses (de Zoeten, 1991). For the viruses that require aphids for their transmission, transcapsidation may lead to an altered virus-vector specificity because the viral RNA of an aphid-nontransmissible virus could be encapsidated into the capsid of an aphid-transmissible virus, and subsequently, be transmitted to a new host (Creamer and Falk, 1990). RNA recombination, as discussed previously, occurs when the viral replication complex changes templates during the natural process of RNA replication (Kirkegaard and Baltimore, 1986). If CP-transgenic plants become infected with viruses related to the 12 one which provides the CP transgene, or unrelated viruses, template switching might occur between the transgenic mRNA and the replicating viral RNA (de Zoeten, 1991). The progeny virus would have segments of genes from both the viral transgene and the infecting virus and thus have properties different fiom the parental virus. Template switching and transcapsidation, coupled with the other factors which perturb the RNA virus population, could generate “new” viruses with altered vectors, host ranges, and new combinations of genes (de Zoeten, 1991). RNA recombination in transgenic plants expressing viral genes was first discovered on red clover necrotic mosaic dianthovirus (RCNMV) and was reported during a conference in 1991 (Lommel and Xiong, 1991). In this study, the RCNMV movement protein transgene lacking the 5’ untranslated region (UTR) acquired a fimctional 5’ UTR through RNA recombination with the inoculated RNA 1, which has a complete 5’ UTR. The second demonstration of transgene recombination was in a plant DNA pararetrovirus cauliflower mosaic virus (CaMV), where template switching occurred between the gene VI transgene and a movement defective CaMV during reverse transcription of the CaMV RNA to DNA (Gal et al., 1992; Schoelz and Wintermantel, 1993). In 1994, Greene and Allison conducted a sensitive bioassay experiment using the bromovirus model system and further demonstrated that RNA recombination involving viral transgenes is a frequent phenomenon. In their experiment, three single nucleotide substitutions, which also generate a unique N011 restriction site, were introduced near the 3’ end of the CCMV coat protein ORF. The 3’ 2/3 of the CCMV CP ORF with the NotI l3 restriction site, along with the complete 3’ UTR was introduced into the N. benthamiana plants using Agrobacterium-mediated transformation. Transgenic plants with a high accumulation of transgene transcripts were selected for the recombination experiments and were clonally propagated to ensure genetic homogeneity of the plant materials. These transgenic plants were inoculated with a movement defective mutant, AG3, which lacked the 3’ 1/3 of the CCMV CP ORF. The design of this experiment allowed Greene and Allison to screen for recombinants by looking for the systemic movement of the virus in transgenic plant cuttings, since the systemic movement of the virus indicated the recovery of the deletion in the CP ORF through RNA recombination between the CP transgene and the deletion mutant AG3. Out of the 235 transgenic plant cuttings inoculated, seven recombinants were recovered, which gave a 3% rate of recombinant recovery. Sequence analysis indicated that all recombinants had restored the deletion at the 3’ end of the CP ORF, and the molecular marker N011 site was also present. These recombinants also bore nucleotide and amino acid substitutions, deletions, and insertions, and several recombinants produced atypical symptoms in the cowpea host. In a similar study, Greene and Allison (1996) inoculated this same movement defective mutant AG3 on a set of transgenic plants transformed with the same 3’ 2/3 of the CP ORF but an incomplete 3’ UTR. No recombinant was recovered from any of the 479 transgenic plant cuttings. The lack of recombinant recovery in this experiment leads to the conclusion that the transgene 3’ UTR played a significant role in RNA recombination in transgenic plants (Greene and Allison, 1996). 14 My research is built upon the results of Greene and Allison (1994; 1996) and attempts to answer four general questions: 1) Why is there a decrease in the RNA recombination rate when the 3’ UTR is incomplete? What is the role of 3’ UTR in RNA recombination? 2) Can RNA recombination occur between a polymerase transgene and an infecting virus? Is the RNA recombination rate in the polymerase transgene comparable to that in the CP transgene? 3) Is a translatable transgenic mRNA more or less likely to be involved in RNA recombination? 4) Does a full-length CCMV CP transgene confer virus resistance to transgenic plants? We hope that answers to these questions will lead to a better understanding of RNA recombination in transgenic plants expressing viral genes. The ultimate goal of this research is to provide information to genetically modified (GM) food regulators and developers for the purpose of developing safer GM products, especially virus resistant transgenic crops in which the transgenes are less likely to be involved in RNA recombination events. 15 CHAPTER 1 Synthesis of Minus Strand Copies of the Viral Transgene during Viral Infections of Transgenic Plants INTRODUCTION Coat protein and replicase mediated resistance are effective in limiting viral infections in a variety of plant species (F itchen and Beachy, 1993; Grumet, 1995). While the mechanisms of pathogen-derived resistance remain unsolved, many studies indicate that transcription of the viral transgene and in some cases expression of the viral protein is required for effective resistance (Fitchen and Beachy, 1993). While several studies have shown that a full-length coat protein gene is sufficient to provide virus resistance (Cuozzo et al., 1988; Smith et al., 1995; Tricoli et al., 1995; Hefi‘eron et al., 1997), numerous constructs have included the 3’ UTR along with the full-length coat protein gene (Powell-Abel et al., 1986; Stark and Beachy, 1989; Zaccomer et al., 1993). Presumably inclusion of the 3’ UTR was intended to stabilize the transgene transcript in the cytoplasm. The 3’ UTR of most RNA viruses bears the replicase recognition site and also serves as an initiation site for negative-sense RNA synthesis. This has been demonstrated in bromoviruses (Ahlquist et al., 1981a; Bujarski et al., 1986), potexviruses (Sriskanda et al., 1996; White et al., 1992), and ilarviruses (Bachman et al., 1994). Our previous data indicated that exclusion of the 3’ UTR in the viral transgene discouraged RNA 16 recombination between viral transgenic transcripts and infecting viruses (Greene and Allison, 1996). When transgenic N. benthamiana plants transcribing the 3’ two-thirds of the capsid gene and the complete 3’ UTR of CCMV were inoculated with a movement defective CP deletion mutant, recombinant virus was recovered from 3% of the inoculated plants (Greene and Allison, 1994). However, when transgenic N. benthamiana plants, transcribing the same viral gene fiagment but lacking all or part of the 3’ UTR, were inoculated with the same CP deletion mutant, no recombinant was recovered from 479 transgenic plants (Greene and Allison, 1996). One hypothesis to explain these results is that inclusion of replication recognition sequences in the transgene leads to partial amplification of the viral transgene by the viral replicase. This amplification produces a minus-strand copy of the transgene which is also available for recombination within the transgenic plants. In this chapter, we report the use of Northern blot analysis and RT-PCR techniques to detect the complementary copy of a viral transgene in transgenic plants infected with either BMV or CCMV. Two transgenic plant lines, one with an intact 3’ UTR, and one with a deleted 3’ UTR, were inoculated with either wild type BMV or CCMV. In the PCR experiment, we took advantage of the uniqueness of the 3’ UTRs of BMV and CCMV genomic RNA and also the ability of BMV to recognize and replicate CCMV RNAs (Allison et al., 1988). Our data indicated that a fiall-length complementary cOpy of the viral transgene was synthesized only when the 3’ UTR of the viral transgene was intact. 17 MATERIALS AND METHODS Plasmid construction Full-length cDNA clones of wild type CCMV RNAI, RNA2 and RNA3 are maintained in plasmid pCClTPl, pCC2TP2, and pCC3TP4, respectively (Allison et al., 1988). Infectious transcripts, namely C1, C2, and C3, were synthesized from the Xbal digested DNA templates using T7 RNA polymerase (Epicenter). Plasmid pCC3AG1 was generated fi'om plasmid pCC3TP4 by introducing three single nucleotide substitutions including a Natl restriction site near the 3’ end of the CP ORF (Greene and Allison, 1994; Fig. 1). Full-length cDNA clones of wild type BMV RNAl, RNA2 and RNA3 were maintained in plasmids pBlTP3, pB2TP5, and pB3TP8, respectively (Janda et al., 1987). Infectious transcripts, namely B1, B2, and B3, were synthesized from EcoRI digested DNA templates using T7 RNA polymerase (Epicenter). Plant materials The transgenes in transgenic plants 3-57 and A69 were derived from plasmid pCC3AG1 which bears three marker mutations and the unique Natl restriction site near the 3’ end of the CCMV CP gene (Fig. 1). Transgenic N. benthamiana plant 3-57 was transformed with the 3’ 2/3 of the modified CCMV CP gene and the complete 3’ UTR. Transgenic N. benthamiana plant A69 was transformed with the same 3’ 2/3 of the CCMV CP gene, but the 3’ terminal 69-nucleotide of the 3’ UTR were deleted (Greene and Allison, 1996; Fig. 1). Transgenic 3-57 and A69 plants used in this study were either clonally propagated F0 plants, or their F1 seedlings. To ensure the transgenic status of the 18 seedlings, they were maintained in MS media containing 100 mg/L kanamycin for one month before being transferred to soil. Nontransgenic N. benthamiana plants were grown in soil. Plant inoculation Wild type BMV or CCMV infection was established using the full-length infectious transcripts. Infected N. benthamiana leaf tissue was ground in a 10 mM sodium phosphate buffer (pH 7.0) with a mortar and pestle. This viral inoculum was applied to two fully expandedleaves dusted with carborundum. All of the plants were inoculated at the 5-leaf stage. Newly inoculated plants were left in the dark for one day before transferring to the grth chamber maintained at 26°C under florescent light with a 16-hour day period. Detection of viral infection To determine the most appropriate time to sample viral RNA, dot blot analysis was used to determine the progression of the infection and the viral RNA concentrations in BMV and CCMV inoculated N. benthamiana. Leaves above the inoculated leaves were numbered and two leaf discs 8 mm in diameter were collected from each leaf at 10, 12, 14, and 16 days post inoculation (dpi). The leaf samples were ground in 40 [.11 of deionized distilled water in a microcentrifiige tube. After centrifugation at 13,000 rpm for two minutes, the supernatant was collected and treated in 1% SDS at 65°C for five minutes. The tubes were again centrifuged at 13,000 rpm for two minutes to remove all the plant debris. Five microliters of the supernatant was dotted directly onto Hybond N 19 nylon membrane (Amersham). . Afier baking the membrane at 80°C for two hours under vacuum, the blot was hybridized with dig-labeled RA518(+) (Allison et al., 1990) or HEl RNA probe (Ahlquist et al., 1981). These probes are specific for the 3’ UTR of CCMV or BMV RN As, respectively. The hybridization, washing and immunodetection were as recommended by the manufacturer (Boehn'nger-Mannheim). Total RNA extraction and treatment Total RNA was extracted from BMV or CCMV inoculated leaf tissue at 14 days post inoculation (dpi) (Puissant and Houdebine, 1990). To recover the double-stranded RNA (dsRNA), total RNA fiom one gram of plant tissue was treated in a 100 ul solution containing 10 mM Tris-HCl (pH 7.2), 10 mM MgCl;, 1 mM EDTA, 300 mM NaCl, 10ug/ml RNase A, 30 units/ml RNase T1 and 500 units/ml RNase fiee DNase I at 30°C for 2 hours (Hardy et al., 1979; Sethna et al., 1989). The added nucleases were destroyed by adding 2.5 pl 20% SDS and 2.5 pl 20 mg/ml proteinase K to each tube and incubating them at 37°C for one hour. After seven extractions with phenol-chloroform- isoamylalcohol, and twice with chloroform-isoamylalcohol, the double-stranded RNA was precipitated with 2.5 volumes of ethanol. Gel electrophoresis and Northern blot analysis The double-stranded RNA pellets were resuspended in 11 ul of RNase-free water and denatured at 55°C for 15 minutes in a 50 ul solution containing 1X MOPs buffer (pH 7.0), 2.2M formaldehyde, 50% forrnamide. Gel preparation, electrophoresis, and Northern transfer were as described by Brown and Mackey (1997). Digoxigenin- or 32P- 20 labeled RA518(-) probes were synthesized using Pvul linearized pCC3RA518 (Allison et al., 1990) plasmid DNA in the presence of T7 RNA polymerase (Epicenter). RA518(-) anneals to a 179-nucleotide region in the 3’ UTR of all four negative-sense CCMV RNAs (Allison et al., 1990). The procedure of hybridization and washing was as described by Allison et a1. (1990). Plasmid DNA pCC3TP4 (Allison et al., 1989) was denatured by the same method as double-stranded RNA and dotted directly on nylon membrane. The position of negative sense RNA 4 and the viral transgene were calculated according the migration pattern of negative sense RN As 1, 2, and 3 (Maniatis et al., 1982). RT-PCR analysis and restriction digestion The primer used for first strand cDNA synthesis was a negative sense CCMV RNA3 specific primer RA83 (5’ AAGTGGATCCCCTCTTGTGCGGCTGC 3’), which anneals at nucleotides 1519-1544. An additional primer used for PCR analysis was RA84 (5’ ACTCCAAAGAGTTCTTCCG 3’), which anneals at nucleotide 2072-2090 of positive sense CCMV RN A3. Total RNA extracted from the plant tissue was first treated with DNase I in a buffer containing 50 mM Tris-Cl, pH 7.5, 10 mM MgC12, 50 1.1ng BSA, and 0.05 U/ul RNase fiee DNase I (Boehringer—Mannheim) at 37°C for an hour prior to first strand cDNA synthesis. Reverse transcription was carried out as described by the manufacturer (Boehringer-Mannheim). Complementary DNA synthesized from lug of total RNA was used for each PCR reaction as described by Greene and Allison (1994). For the control reaction, 0.1 ug pCC3AG1 (Fig. 1) plasmid DNA was used in each reaction. RT- PCR products were purified by extraction twice in phenol-chloroform- isoamylalcohol followed by ethanol precipitation. NotI restriction enzyme digestion was 21 pCC3TP4 1476 StyI , l I 3a | L | CF I SM :33; , I I | 3a | I | or | Natl I Transgenic plant 3-57 [ I + ’4- RA83 RA84 Not] I Transgenic plant A69 -> <- RA83 RA84 Figure 1. Plasmid construction. Wild type CCMV RNA3 has a 3a gene ORF and a coat protein (CP) ORF, and it is maintained in plasmid pCC3TP4. A Not I restriction site was introduced near the 3’ end of the CP gene ORF in plasmid pCC3AG1 as well as the viral transgenes in transgenic plants 3-57 and A69 (Greene and Allison, 1994;1996). Transgenic plant 3-57 was transformed with the 3’ 2/3 of the CP ORF and the full-length 3’ UTR. Transgenic plant A69 was transformed with the same viral gene, but the 3’ UTR bears a 69-nucleotide deletion at the 3’ end. The negative sense RNA specific primer RA83 (5’ AAGTGGATCCCCTCTTGTGCGGCTGC 3’) anneals at nucleotides 1519-1544, and was used for first strand cDNA synthesis. An additional primer RA84 (5’ ACTCCAAAGAGTTCTTCCG 3’) anneals at nucleotides 2072-2090, and was used for PCR. 22 as described by the manufacturer (Boehringer-Mannheim). RESULTS Three sets of plant materials, namely nontransgenic N. benthamiana and transgenic 3-57 and A69 N. benthamiana, were used in this study. As indicated in Fig. 1, a Natl restriction site was engineered into mutant AG] and was therefore included in the viral transgenes of transgenic plants 3-57 and A69 (Greene and Allison, 1994; 1996). Both 3-57 and A69 transgenic plants cany the 3’ two thirds of the CP ORF with the molecular marker Natl site. The only difference between these plants is that 3-57 plants have an intact 3’ UTR, whereas A69 plants have a 69 nucleotide deletion at the 3’ end of the 3’ UTR (Greene and Allison, 1994; 1996). Detection of minus-sense viral RNA in planta As single-stranded positive sense RNA viruses, bromoviruses replicate in the cytoplasm, and a small amount of minus-sense RNA is synthesized prior to the synthesis of positive-sense RNA (Marsh et al., 1991a). The minus-sense RNA is constantly associated with a plus-sense RNA, and upon extraction, becomes a double stranded form, which is resistant to mild ribonuclease treatment under high salt conditions (Hardy et al., 1979). The minus-sense RNA accumulates to a level 100-fold lower than the positive- sense RNA in the total RNA extracted from protoplasts at 24 hour post inoculation (Rao et al., 1990a; Marsh et al., 1991b). 23 For in planta detection of minus strand genomic RNAs, it was important to determine the time when most plant cells were at an early stage of infection. To track the movement of BMV and CCMV in N. benthamiana, leaf tissue was collected from each leaf above the inoculated leaves at 10, 12, and 14 days post inoculation (dpi) and viral RNA accumulation was analyzed in dot blots. Our data indicated that at 10 dpi, viral RNA was only present in the youngest visible leaves, but absent in the other fully expanded leaves. Viruses moved to the first fully expanded leaf at 12 dpi. By 14 dpi, viral RNA was present in all leaves tested (data not shown). Thus, in this study, total RNA was extracted at 14 dpi. At the initial stage of this study, attempts were made to detect minus-sense viral RNA directly from total RNA extracted from leaf tissue. However, none of them was successfiJl. This is probably because minus-sense RNA represents such a small fraction of the total RNA that the majority single-stranded RNA prevented it from hybridization with the RNA probe. RNase treatment was used later to remove the single-stranded cellular and viral RNAs, which make up the majority of the total RNA. The double- stranded RNA which was resistant to RNase-treatment was denatured and analyzed in the Northern blots using 32P-labeled RNA probe RA518(-), which recognizes the 3’ UTR of CCMV RN As (Allison et al., 1990). The sensitivity of this probe was optimized to detect 10 pg of denatured plasmid DNA pCC3TP4 (Fig. 2). Using this RNA-RNA hybridization system, we were able to routinely detect the negative-sense genomic RNAs of CCMV from 0.5 to 1.0 gram of the infected nontransgenic N. benthamiana plant tissue at 14 dpi. 24 100 ng 10 ng 1 ng 100 pg 10 pg 1 pg Figure 2. Dot blot assay quantifying the sensitivity of radiolabeled RA518(-) probe. Serial dilutions of denatured plasmid pCC3TP4 were spotted mechanically onto Hybond N nylon membrane. Each dot received 100 ng, 10 ng, 1 ng, 100 pg, 10 pg or 1 pg denatured plasmid DNA, respectively. 25 The complementary copy of the viral transgene accumulated at a much lower level than that of the genomic RNAs A total of four experiments were carried out to detect the complementary copy of the viral transgene by Northern blot analysis. Fig. 3 shows the representative results. Double-stranded RNA fi'om two grams of plant tissue was loaded in each lane; lanes 1-6 and lanes 7-10 contain double-stranded RNA extracted from CCMV and BMV infected 3-57 plant tissue, respectively. Negative-sense RNAs l, 2, and 3 of CCMV were detected by the negative-sense CCMV RNA specific probe in CCMV infected plants, but not in BMV infected plants. The amount of negative-strand RNA migrating at the position of subgenomic RNA4 is very small compared with the negative-strand genomic RNA3 of CCMV. This data agrees with the negative-sense RNA accumulation patterns in CCMV infected barley protoplasts (Pacha and Ahlquist, 1991). The expected site of the complementary strand of the viral transgene was calculated according to the migration patterns of the negative sense RN As 1, 2, and 3 of CCMV. This fi'agment was not observed in either BMV or CCMV infected plants (Fig. 3). Although F1 seedlings of 3-57 plants were used in this experiment, similar results were obtained when F0 plant samples were used in repeated experiments. Despite the fact that in this Northern blot, the probe was optimized to detect 10 pg of denatured plasmid DNA, and dsRNA from two grams of the plant tissue was loaded in each lane, the complementary copy of the viral transgene was not observed. This indicates that if the complementary copies of the viral transgene were present in the double-stranded form, 26 RNA3 (-) RNA4 (-) Figure 3. Northern blot analysis of dsRNA. Double-stranded RNA extracted from either CCMV or BMV infected transgenic plant 3-57 was probed with the negative-sense CCMV RNA specific probe RA518(-). Lanes 1-6 contain samples from CCMV infected plant tissue, lanes 7-10 contain samples from BMV infected plant tissue. Positions of negative-sense CCMV RNAs l, 2, 3, and 4 were labeled on the left of the picture. The arrow, —>, indicates the calculated position of negative-sense viral transgene which was not detected. 27 they were beyond the detection limit of our conventional RNA-RNA hybridization technique and below the concentration of ten picograms per two grams of leaf tissue. Complementary copies of the viral transgene were detected in the RT-PCR analysis To better search for the complementary copies of the viral transgene, a more sensitive technique, RT-PCR, was used. In this procedure, total RNA from the above mentioned plant tissue was first treated with RNase free DNase I to remove the plant genomic DNA including the chromosomal copy of the viral gene. A minus-strand CCMV RNA3 specific primer RA83 (Fig. l) was used for first-strand cDNA synthesis. PCR was completed using an additional primer RA84, which anneals near the 5’ terminus of the 69-nucleotide deletion (Fig. 1). The size of the expected RT-PCR fi'agment was 572 bp. As shown in Fig. 4, lanes 7, 10 and 13, a minus-strand RNA was amplified in all of the CCMV infected 3-57, A69, and nontransgenic plants. This minus-sense RNA was not observed in any mock inoculated transgenic plants or in BMV infected A69 or non- transgenic plants (Fig. 4, lanes 5, 6, 8, 9 and 11). A band of the correct size was also present in the BMV infected 3-57 transgenic plants (Fig. 4, lane 12). A Natl restriction digestion was used to determine the origin of the amplified minus-strand CCMV RNAs. As mentioned before, the Natl site was present only in the viral transgene and not the wild-type CCMV inoculum (derived from pCC3TP4). As shown in Fig. 5, lane 3, RT-PCR product of AGI mutant containing the Natl site was completely cleaved by the Natl restriction enzyme, but that of wild type CCMV was not cleaved (Fig. 5, lane 5). The RT-PCR product of CCMV infected A69 plants was not 28 Nontransgenic A69 3—57 AGl neg E le M B C l 2 3 4 5 6 7 “E cu: m 3 w n Figure 4. Agarose gel electrophoresis showing RT-PCR amplified negative-sense CCMV RNA3. Total RNA from virus infected or mock-inoculated transgenic and nontransgenic N. benthamiana plant tissue was treated with RNase free DNase I to remove the genomic DNA. A negative-sense CCMV RNA3 specific primer RA83 (5’ AAGTGATCCCCTCTTGTGCGGCTGC 3’), which anneals at nucleotide 1519-1544, was used for first-strand cDNA synthesis. An additional primer M84 (5’ ACTCCAAAGAGTTCTTCCG 3’), which anneals at nucleotide 2072-2090, was used in PCR. Lanes 5-7, 8-10, and 11-13 were from nontransgenic, transgenic A69, and 3-57 plants, respectively. Samples in lanes 5, 8, and 11 were mock-inoculated (M); lanes 6, 9, and 12 were inoculated with BMV (B); lanes 7, 10, and 13 were inoculated with CCMV (C). Lane 1 contains PCR product using 0.1 pg pCC3AG1 plasmid DNA. Lane 2 is the negative control of PCR where water was added to the PCR mixture. Lane 3 is empty (E). Lane 4 contains 1 Kb size marker (GibcoBRL). 29 nontransgenic A69 3-57 3-57 M pCC3AG1 CCMV CCMV BMV CCMV 1 2 3 4 5 6 7 8 9 10 ll Figure 5. Agarose gel showing Natl treated RT-PCR products amplified from total RNA extracted from virus infected plant tissue. Lane 1 contains the 1 Kb size marker (GibcoBRL). Lanes 2-3 contain PCR products amplified from pCC3AG1. RT-PCR amplified products were from CCMV infected nontransgenic N. benthamiana plants (lanes 4-5); CCMV infected A69 plants (lanes 6-7); BMV infected 3-57 plants (lanes 8-9); CCMV infected 3-57 plants (lanes 10-11). RT- PCR products in lanes 2, 4, 6, 8, and 10 were not treated with Natl restriction enzyme, whereas those in lanes 3, 5, 7, 9, and 11 were treated with Natl. Arrow “(—” indicates undigested fragments. Arrow “4 ” indicates digested fragments. Note the small digested bands in lane 11. 3O cleaved by Natl (Fig. 5, lane 7), which indicates that the negative strand CCMV RNA amplified in CCMV infected A69 transgenic plants was from the wild type CCMV. The RT-PCR product of CCMV infected 3-57 plants was partially cleaved by Natl (Fig. 5, lane 11), indicating that for CCMV infected 3-57 transgenic plants, the negative CCMV RNA3 which was PCR amplified was of two origins: most of the RNA was from wild type CCMV inoculum, while some was from the transgene mRNA. For BMV infected 3- 57 transgenic plants, the RT-PCR product was completely cleaved by Natl, which indicates that this RT-PCR amplified fragment was derived from the viral transgene, the only CCMV RNA fi'agment with the marker mutation Natl site. Note that no PCR fragment is present in the BMV inoculated A69 plants (Fig. 4). The only explanation for this data is that the replication complex of BMV recognized the CCMV viral transgene and synthesized a complementary copy. DISCUSSION Previous studies revealed that negative-sense RNA of BMV and CCMV accumulates to levels 100-fold' lower than positive-sense RNA in protoplasts at 24 hour post inoculation (Marsh et al., 1991a; Pacha and Ahlquist, 1991). Considering that synchronized infection of viruses is impossible in a plant, and the negative-strand synthesis happens at the very initial stage of viral infection in a single cell, the negative- sense RNA could accumulate at a much lower level in planta than in protoplasts. Thus in planta detection of negative-sense viral RNA of bromoviruses usually requires total RNA extracted from several grams of the plant tissue rather than a small amount of protoplast, and a series of treatments to remove the single-stranded cellular and viral RNA from the 31 double-stranded replicative forms (Bastin and Kaesberg, 1976). More recent studies on RNA replication of bromoviruses were exclusively conducted in protoplasts rather than in planta (Loesch-Fries and Hall, 1980; Marsh et al., 1991b). However, since our previous RNA recombination experiments implied that initiation of negative RNA synthesis on the viral transgene in planta could increase the incidence of RNA recombination in transgenic plants (Greene and Allison, 1994; 1996), we chose to demonstrate in planta synthesis of the complementary copies of the viral transgene during viral infections of transgenic plants. Viral transgene transcripts were recognized by viral replication complex, and a complementary copy of the viral transgene was synthesized The RT-PCR experiment provided convincing data that the viral transgene can be recognized by the viral replication complex when the viral transgene contains an intact 3’ UTR. The minus-strand promoter on the CCMV viral transgene is not only recognized by the viral replication complex of wild type CCMV, but also by that of another closely related virus, BMV. Thus in both cases, a full-length complementary copy of the viral transgene was obtained. Effects of deletion on the recognition of viral transgene Minus-strand promoter was mapped to within the 134 3’ terminal bases of genomic BMV RNAs (Miller et al., 1986), and the minus-strand promoter of CCMV RNA is expected to reside in a similar location (Ahlquist et al., 1981a; Bujarski et al., 1986). Deletions within the 3’ terminal sixty bases can cause dramatic loss of the minus- 32 strand promoter activity in BMV (Marsh et al., 1991b). In this study, the complementary copy of the viral transgene with 69 nucleotides deleted 3’ UTR was not detected in either BMV or CCMV infected plants. This indicated that the minus-strand promoter at the 3’ UTR of the viral transgene is essential for viral transgenic mRN A recognition and synthesis of minus-sense RNA by the viral replication complex. The viral transgene as well as the subgenomic RNA4 were not the preferred templates for negative-sense RNA synthesis in planta Despite the high accumulation level of the viral transgene in 3—57 transgenic plants (Greene and Allison, 1994; 1996), the nearly 700 bp viral transgene is not proportionately represented in the pool of negative-sense RNA. Even though the detection limit of the RNA-RNA hybridization system was optimized to detect 10 pg of denatured DNA and the negative-sense copies of the genomic RNAs 1, 2, and 3 showed strong signals, the complementary copy of the viral transgene was still not observed in the Northern blot (Fig. 3). The negative-sense copy of the subgenomic RNA 4 provided a weak band (Fig. 3). This agrees with the observation of Pacha and Ahlquist (1991), that the amount of negative-sense subgenomic RNA 4 is small for both BMV and CCMV. The negative-sense RNA 4 of CCMV was barely visible in the Northern blot when either BMV or CCMV replicative proteins in combination with CCMV RNA3 were used to infect barley protoplasts (Pacha and Ahlquist, 1991). An explanation for these observations is that, negative RNA synthesis in planta is a regulated process. Although bearing the same minus strand promoter, subgenomic RNA 4, as well as the CCMV transgenes, were not recognized as readily as the genomic RNA 3. 33 Miller et al. (1986) reported that polyadenylation of either the 3’—terminal 161- nucleotide fragment of BMV RNA or the full-length RNA4 had little effect on minus strand promoter activity in vitra when the poly(A) tail was between 15-30 nucleotides long. Since the size of the replication product was the same as that obtained from unmodified templates, Miller et a1. (1986) concluded that initiation occurred at the same site in both cases. The poly(A) tail added to the viral transgenic mRNA in the 3-57 plants may be up to several hundred nucleotides long. Whether the length of the poly(A) tail had any effect on the viral RNA recognition and initiation during minus strand synthesis remains unknown. This study provides an explanation on the difference in recovery of recombinant virus in transgenic plants with a complete or incomplete 3’ UTR. Several studies on different viruses indicated that RNA recombination can happen during either plus (Carpenter et al., 1995) or minus (Kirkegaard and Baltimore, 1986; Mueller and Wimmer, 1998) strand synthesis. Detection of the complementary copy of the viral transgene in transgenic plants with full-length 3’ UTR offers an explanation for the reason why the recombination rate is higher when the 3’ UTR of the viral transgene is intact. The presence of both plus and minus strand copies of the viral transgene in a single infected cell increases the incidence of the interaction between the viral transgene and the infecting viruses, and also allows the recombination event to happen during both plus and minus strand RNA synthesis. This study also offers fiirther support to the conclusion of Greene and Allison (1996), that excluding the viral replication initiation 34 site from the viral transgene is an effective way of reducing, although not abolishing, RNA recombination in transgenic plants. 35 CHAPTER 2 RNA Recombination betWeen a Brome Mosaic Virus Deletion Mutant and a Polymerase Transgene INTRODUCTION Transgenic plants that express a viral capsid or a polymerase gene are, in many cases, resistant to infection by the virus from which the genes were derived, and closely related viruses (Grumet, 1995). Following the discovery of coat protein-mediated resistance (CPMR) (Powell-Abel et al., 1986), replicase-mediated resistance was demonstrated in tobacco plants transformed with the TMV replicase gene (Golemboski et al., 1990). In many cases, it appears that replicase mediated resistance is stronger than CP mediated resistance (Braun and Hemenway, 1992; Donson et al., 1993). While virus resistant transgenic plants (VRTPs) appear to be an effective means of controlling viral diseases, several studies have indicated that the viral transgene transcripts provide a unique opportunity for RNA recombination with challenging viruses (reviewed in Allison et al., 1997). The case study by Greene and Allison (1994) indicated that a capsid transgene of CCMV was involved in RNA recombination with a movement defective CCMV CP deletion mutant used to infect transgenic plants, and 3% of the plants yielded recombinants in which the deletion was repaired through recombination with the transgene. Examination of the recovered recombinants revealed numerous 36 nucleotide and amino acid substitutions, deletions, and additions. Several recombinant viruses caused symptoms that were distinctly different from wild type CCMV. In this study, we sought to determine if RNA recombination could also involve polymerase transgenes, and if so, to compare the recovery rate of polymerase recombinants with that of the capsid recombinants obtained by Greene and Allison (1994). The design of this experiment was similar to that of Greene and Allison (1994). Nicatiana benthamiana Domin were transformed with a segment of the BMV polymerase gene which bore a silent marker mutation which provided an Eca47III restriction site. The challenging BMV included a 307-nt deletion near the 3’ end of the polymerase gene open reading fi'ame (ORF) in BMV RNA 2. This deletion mutant, PTSO, remained capable of replication in both barley (Hordeum vulgare L.) and N. benthamiana (Traynor et al., 1991; Smirnyagina et al., 1996), but systemic movement in N. benthamiana was inhibited. Transgenic lines contained the segment of the BMV polymerase gene that was absent in the BMV deletion mutant. Thus, systemic movement of the virus in PTSO infected transgenic plants signaled the restoration of the polymerase gene through RNA recombination. Our data indicates that the viral polymerase transgene is actively involved in RNA recombination events and the recovery of polymerase recombinants is comparable with that of capsid recombinants. 37 MATERIALS AND METHODS Plasmids for the synthesis of infectious transcripts Full-length cDNA clones of wild type BMV RNAl, RNA2 and RNA3 were maintained in plasmids pBlTP3, pB2TP5, and pB3TP8, respectively (Janda et al., 1987). Transcripts from these clones are referred to as B], B2, and B3 hereafier. The BMV RNA2 deletion mutant clone pBZPTSO was a generous gift from Dr. Patricia Traynor (Virginia PolyTech, Blacksburg, Va.). This clone lacks 307 nucleotides at the 3’ end of the 2a gene ORF which results in a 125 amino acid deletion at the C-terminal of the 2a protein (Traynor et al., 1991; Figure 1). Transcripts from pB2PT50 are referred to as PT50 hereafter. Plasmid pB2MD4 is a full-length BMV RNA2 clone containing an A—9G silent mutation at position 2458. This mutation was generated through site-directed mutagenesis of pB2TP5 using primers RA76 and B2. Mutagenesis primer RA76 (5’ ATCGTCGACCTTCTTTTTCWGTGTGGTC 3’) anneals at nucleotides 2449- 2482 and it introduces the A——>G silent mutation (bold and italicized) which also creates a unique Eco47HI restriction site (double underlined). The existing restriction site Sail (single underlined, Fig. 1) is also included in RA76 for the convenience of cloning. Primer B2 (5’ GTAAACCACGGAACG 3’) anneals at nucleotides 1-15. Plasmid pB2MD4 was constructed by replacing the 800 bp MIul-SalI fragment of pBZTPS with the same restriction fragment containing the PCR generated silent mutation (Fig. 1). Transcripts from pB2MD4 are referred to as MD4 hereafter. The mutation in plasmid pB2MD4 and the 307-nucleotide deletion in plasmid pB2PT50 were confirmed by dideoxy-sequencing (Amersham). 38 1252 1681 104 Kpnl MM 2195 San 2502 2569 2865 PBZTP5 I 1 1 3 ' / I fir-if l / I / / I I pB2PT50 1 1 ’ UAA pB2MD4 Transgene Figure 1. Plasmid construction. Plasmid pB2TP5 is the full-length wild type BMV RNA2 clone. Plasmid pB2PT50 is a deletion mutant of BMV RNA2, which lacks 307- nt at the 3’ end of the polymerase gene (deleted nucleotides 2195-2501) (Traynor et al., 1991). Plasmid pB2MD4 is the full-length BMV RNA2 clone with an A-—>G marker mutation at position 2458. The primer used for site-directed mutagenesis was RA76 (5’ ATCGTCGACCTTCTTTTTCWGTGTGGTC 3’), which adds a unique Eca47III restriction site (double underlined) through this single A->G substitution at position 2458 (bold and italicized). The 3’ 1.6 Kb KpnI-BamH I fragment was excised from pB2MD4 and inserted into binary vector pTETl downstream of the CaMV 358 promoter to make pTETMD3. This 1.6 Kb transgenic mRNA was detected in several transgenic plants selected for the recombination experiments (Fig. 2). The boxes indicate the polymerase open reading frame (ORF). The shaded area indicates the central polymerase core region. 39 Infectious transcripts, namely B1, B2, B3, PT50, and MD4, were synthesized using T7 RNA polymerase (Epicenter) and EcoRI digested DNA plasmid templates (Janda et al., 1987). Plasmid for the Agrabacterium—mediated transformation To construct an Agrabacterium-mediated transformation vector, the 3’ 1.6 Kb KpnI-BamHI fragment of pB2MD4 was excised and inserted into Kpnl-BamHI digested binary vector pTETl, which is a derivative of pBin19 (Frisch et al., 1995), at a position downstream of the CaMV 358 promoter to make pTETMD3 (Fig. 1). The NPT II gene is also located within the left (LB) and right (RB) border sequences of the binary vector and it provides kanamycin resistance to regenerated shoots. The correct orientation and joining of the KpnI-BamI-ll fragment in pTETMD3 was confirmed by dideoxy- sequencing (Amersham). Bacterial strains Escherichia cali strain JM101 cells were grown at 37°C for 16 hours prior to plasmid DNA extraction. LB medium containing 100 mg/L kanamycin was used to select JMlOl cells containing pTETl or pTETMD3. For all other plasmids in this study, 50 mg/L ampicillin was included in the media. Agrabacterium tumefaciens strain LBA4404, which contains the helper Ti plasmid pRAL4404 (Hoekema et al., 1983), was maintained on solid YEP medium (10 mg/ml peptone, 10 mg/ml yeast extract, 5 mg/ml NaCl, 1.5% Bacto-Agar) containing 50 4O myL rifampicin (Tavazza et al., 1988). The transformation vector pTETMD3 was introduced into LBA4404 cells using the freeze-thaw method (Holsters et al., 1978). Transformed cells were selected and maintained on solid YEP medium containing 50 mg/L rifampicin and 100 mg/L kanamycin. LBA4404 cell lines were grown at 28-30°C. Agrabacten‘um—mediated transformation The Agrabacterium-mediated transformation was adopted from Benvenuto et a1. (1991), Tavazza et a1. (1988) and Horsch et al. (1985), with minor modifications. MS basal salt mix (GibcoBRL, Cat. No. 11117-066) was used to make both MS liquid and solid media, which were adjusted to pH 5.0 and pH 5.7, respectively, using KOH or HCl. Three types of MS solid media were used in this study, and they all contained Gamborg’s B-S vitamins (100 mg/L myo-Inositol, 1 mg/L nicotinic acid, 1 mg/L pyridoxine—HCI, and 10 mg/L thiamine-HCI), 3% sucrose, and 0.8% agar. The cocultivation medium contained 1 mg/L IBA (3-indolebutyric acid) and 1 mg/L BAP (6-benzylaminopurine), but lacked the antibiotics. The shoot induction medium contained the same phytohormones as the cocultivation medium, plus 100 mg/L kanamycin for the selection of transgenic shoots and 200 mg/L cefotaxime for the elimination of bacterial contamination. The root induction medium lacked the phytohormones, but contained both kanamycin and cefotaxime at the same concentration as shoot-induction medium. For plant transformation, leaves from four-week old N. benthamiana were first sterilized in 10% Chlorox containing 0.1% Tween 20 for ten minutes. The leaves were rinsed in sterilized water three times, and blotted dry between two sheets of sterilized 41 Whatman 3MM paper. Steriliz‘ed leaf tissue was cut into one centimeter leaf discs and inoculated by floating them for 5 min in diluted LBA4404 cells containing pTETMD3 plasmids (one milliliter overnight cell culture diluted in 20 ml MS liquid medium, pH 5.0). Inoculated leaf discs were blotted dry on sterilized Whatman 3MM paper and incubated upside down (abaxial surface up) on cocultivation medium in the dark for two days. After incubation, the discs were inverted and transferred to shoot-induction medium. Afler one to two months, the regenerated shoots were excised and transferred to root induction medium. Transformed shoots would usually regenerate roots within twenty days and could be transferred to soil. These plant cell cultures were maintained in petri dishes or magenta boxes and were housed in a grth chamber which was maintained at 26°C with a 16-hour day period. NPT II ELISA testing for transformants Transformants were initially identified by ELISA screening for the NPT II marker gene which provides kanamycin resistance. A leaf disc 8 mm in diameter was collected fi'om each plant and ground in a microcentrifuge tube in 100111 of 0.25 M Tris-Cl, pH 7.8 containing 1.0 mM phenylmethylsulfonylfluoride (PMSF). After centrifirgation at 7500 g for 30 minutes, eight microliters of supernatant were used for the ELISA test (NPT ll ELISA Kit, 5 prime —> 3 prime, Inc). ELISA readings were collected using Minireader II from Dynatech. 42 Plant maintenance and inoculation Transgenic plants (F0) were maintained in soil and clonally propagated by removing auxiliary shoots, dipping them into root induction mixture (200 mg IBA in 100 m1 Talc), and rooting them in medium-sized vermiculite. Transgenic plant cuttings were incubated at 26°C under strong fluorescent light 24 hours each day for two weeks or until the roots grew two centimeters or longer. Rooted cuttings were transferred to soil and maintained in a 26°C growth room with a 16-hour day period. For the recombination experiment, these clonally propagated F0 transgenic plants and non-transgenic N. benthamiana seedlings (used as controls) were inoculated at the five-leaf stage with 0.5 ug each of B1 and B3 transcripts plus 1.0 11g of either B2 or PT50 transcripts. These transcripts were first mixed in a buffer which contained 5 mg/L bentonite, 0.25 M NaCl, 25 mM Tris-Cl, 25 mM NaHZPO4 and SmM EDTA (pH 8.0) and rub-inoculated on two leaves dusted with carborundum (Traynor et al., 1991). Inoculated plants were lefi in the dark for one day before being transferred to the growth room. To test the symptom development of the MD4 mutant, barley (Hordeum vulgare L.) plants were inoculated with 0.2 11g of each B1 and B3 transcripts plus 0.5 ug of either B2 or MD4 transcripts. The procedure was similar as for inoculating N. benthamiana plants. 43 To transfer the potential recombinant viruses to nontransgenic N. benthamiana and barley plants, 0.2 g of leaf tissue was ground in 1 ml of 10 mM sodium phosphate buffer (pH 7 .0) with a mortar and pestle. This inoculum was rub-inoculated to two fiJlly expanded leaves of nontransgenic N. benthamiana at the S-leaf stage, or the first leaf of 7-day-old barley plants. RNA extraction and Northern blot analysis To analyze for the transgene transcripts in regenerated shoots, total RNA was extracted from plant leaf tissue using the technique of Puissant and Houdebine (1990). Total RNA from approximately 0.2 g of leaf tissue was separated in a 1.0% agrose gel in TBE buffer and transferred to a Hybond N Nylon membrane (Amersham). A digoxigenin-labeled HEI probe (Ahlquist et al., 1981b) which anneals to the 3’ 200 nucleotides of BMV positive strand RN As was used for hybridization as described by the manufacturer (Boehringer-Mannheim). For dot blot analysis, one square inch of leaf tissue was ground in a microcentrifiJge tube in 40111 of distilled water. After centrifugation at 13,000 rpm for two minutes, the supernatant was transferred to a fresh microcentrifuge tube and treated in 1.0% SDS at 65°C for five minutes. Following centrifugation at 13,000 rpm for two minutes, 5111 of the supernatant was dotted directly onto Hybond N nylon membrane. The probe and the procedure of hybridization were the same as for the Northern blot analysis. 44 Screening and analysis of recombinant viruses To screen for all potential recombinant viruses, viral RNA was extracted from PT50 inoculated transgenic N. benthamiana showing a positive signal in the dot blot analysis using the same procedure as Traynor et al. (1991). Plants with even the faintest signals were included in the analysis. Viral RNA fi'om 0.2 g of plant tissue was loaded in 1.0% TBE gel and the procedure of RNA-RNA hybridization was similar as for analyzing the total RNA. RT-PCR was also used to screen for potential recombinant viruses. Viral RNA from 0.2 g of leaf tissue was used in the first strand cDNA synthesis using primer QE (5’ CGGAATTCTGGTCTCTTTTAAGAGAT 3’) as described by the manufacturer (Boehringer-Mannheim). Primer OE anneals at nucleotides 2849-2865, and it introduces an EcoRI restriction site (underlined). Primer RA96 (5’ AGTTTCTAGACTCGTCGAAACTGAAATGGG 3’) was added for PCR amplification as described by Greene and Allison (1994). Primer RA96 anneals at nucleotides 1895- 1914, and it adds an XbaI site (underlined). RT-PCR products were purified twice through phenol-chloroform-isoamylalcohol extraction followed by ethanol precipitation, and the products resulting from recombination with the viral transgene were identified by Eca47III digestion (MBI Fermentas). ' These potential recombinant viruses were transferred to nontransgenic N. benthamiana and viral RNA was extracted at 30 dpi. RT- PCR amplified cDNA of viral RNA of potential recombinants was used for cloning. Purified RT—PCR products were first digested with EcaRl and Xbal, and cloned into pBS/KS+ using T4 DNA ligase (GibcoBRL). Clones were sequenced manually by 45 dideoxy sequencing (Amersham). Student’s t-test was used for comparison of the recombinant recovery rate in the study with that in Greene and Allison (1994). RESULTS Several preliminary experiments were required to ensure the logistic design of the recombination experiment. First, we needed to know the movement capability of the PT50 mutant in N. benthamiana plants. Secondly, we wanted to ensure that the single nucleotide mutation at position 2458, which generates the unique Eca47IH restriction site, did not affect the BMV infectivity or symptom development. Thirdly, we wanted to confirm the transgenic nature of the plant materials by detecting the polymerase transgene transcripts within the total RNA. And lastly, we wanted to ensure that the transgenic plants chosen for RNA recombination experiment were not resistant to BMV RNA transcript inoculation. The movement of PT50 mutant is inhibited in N. benthamiana plants Traynor et al. (1991) demonstrated that, in spite of the 307 nucleotides deletion, PT50 directed viral RNA replication in barley protoplasts at nearly wild type level. However, the deletion affected the movement capability of PTSO and caused 20 to 30- fold decrease in viral RNA accumulation in barley plants. In this experiment, we assessed the systemic movement capability of PT50 mutant in non-transgenic N. benthamiana plants. A total of ten non-transgenic N. benthamiana plants and ten barley plants were inoculated with Bl, BB, and PT50 transcripts and the viral RNA was extracted from the non-inoculated leaves at 30 dpi. Despite two tries, viral RNA accumulation was not 46 detected in any N. benthamiana plants, while it was detected in barley using RT-PCR (data not shown). Mutation introduced in the RNA2 gene did not affect the infectivity and symptom development of BMV Plasmid pBZTPS was successfully modified to contain an A—->G silent mutation at position 2458, thus creating a unique Eca47III restriction site in the new plasmid pB2MD4. To ensure that the introduced mutation in the ORF did not affect the infectivity and symptom development of BMV, infectious transcript MD4, together with B1 and B3, were rub-inoculated onto both N. benthamiana and barley. Control plants were simultaneously inoculated with wild type transcripts B1, B2, and B3. At 14 dpi, MD4 inoculated barley plants showed the same symptoms as B2 inoculated barley plants. Electrophoresis and Northern blot analysis of viral RNA indicated that mutant MD4 accumulated at the same level as wt BMV in both N. benthamiana and barley (data not shown). The Eca47lII restriction site in the mutant RNA was confirmed by restriction digestion of RT-PCR amplified products from MD4 inoculated plants only (data not shown). Transgene transcripts were detected from transgenic N. benthamiana plants. One hundred shoots were regenerated from the tissue culture of Agrabacterium inoculated leaf discs. ELISA tests using anti-NPT 11 antibody indicated that half of the shoots were transformed. To compare these transgenic plants on the basis of viral 47 Transgenic lines 3-5 4-1 6-9 7-1 7-2 7-3 9-1 Lane 1 .2“ .. _ 3 4 5 6 7 Figure 2. Detection of viral transgenic RNA transcripts in transgenic N. benthamiana plants. N. benthamiana plants were transformed with the 3’ 1.6 Kb fragment of BMV RNA2 using Agrabacterium-mediated transformation techniques. This 1.6 Kb fragment was first modified using site-directed mutagenesis technique to introduce a unique molecular marker Eca47IH restriction site, and then cloned into a binary vector to make pTETMD3. Total RNA was extracted from transformants 3-5, 4-1, 6-9, 7-1, 7-2, 7-3, and 9-1. Dig-labeled riboprobe HEl, which is specific for a 200-nt fragment at the 3’ end of all four BMV RNAs, was used in RNA-RNA hybridization. The arrow “ 5 ” indicates the site of the transgenic mRNA of 1.6 Kb in size. The “*9, asterisk marks the plant that was resistant to BMV. 4s transgene transcript accumulation, total RNA was extracted from ten transgenic plants with the highest ELISA readings, and analyzed by hybridization in a Northern blot. The probe HEl (Ahlquist et al., 1981b), which is specific for the 3’ terminal 200 nucleotides of BMV RNAs, detected a single RNA band of predicted size (~1.6 Kb). Seven transgenic plant lines, 3-5, 4-1, 6-9, 7-1, 7-2, 7-3, and 9-1, had viral transgene accumulation (Fig. 2) and were chosen for the resistance test described below. Transgenic plants susceptible to BMV infections were selected for the recombination experiments To select transgenic plants susceptible to BMV viral RNA inoculation, clonally propagated cuttings of these seven transgenic plants were challenged with wt BMV RNA transcripts (B1, B2, and B3). Gel electrophoresis and Northern blot analysis of the viral RNA extracted at 14 dpi showed that six of the seven transgenic plants accumulated BMV RNAs in non-inoculated leaves similar to non-transgenic N. benthamiana plants (Fig. 3, lanes 1-2, 4-8). Short exposure of the Northern blot indicated no sign of resistance (data not shown). One transgenic plant, namely 4-1, did not show any viral RNA accumulation in the Northern blot (Figure 3, lane 3). This experiment was repeated twice using either viral RNA transcripts or virion as inoculum, and dot blot analysis was used to detect the systemic movement of the virus at 30 and 60 dpi. Viral RNA accumulation was not detected in any of the 4-1 transgenic plants at 60 dpi. Based on these above observations, transgenic plant 4-1, which was highly resistant to BMV virion and viral RNA inoculation, was excluded from the recombination experiments. Six transgenic plants, namely 3-5, 6-9, 7-1, 7-2, 7-3, and 9-1, were chosen for the 49 .— Control m—rasuNM—r Plants «lessees. 2345678 Figure 3. Test of BMV resistance in transgenic plants. Transgenic plants with the highest accumulation level of transgenic mRNA were inoculated with infectious RNA transcripts of BMV. Viral RNA was extracted at 14 dpi and was examined in Northern blot analysis using dig-labeled riboprobe HEl (Ahlquist et al., 1981b). The gel was overloaded with viral RNA in order to detect virus accumulation in the resistant plant 4-1, and each lane contained viral RNA from 0.2 gram of the plant tissue. Plant samples used were: nontransgenic control (lane 1), transgenic 3-5 (lane 2), 4-1 (lane 3), 6-9 (lane 4), 7-1 (lane 5), 7-2 (lane 6), 7-3 (lane 7), and 9-1 (lane 8). 50 recombination experiments. This resistance data is also pertinent to Chapter 4 of this thesis. Recombinants detected in PT50 inoculated transgenic plants The above mentioned transgenic plants were clonally propagated through cuttings. At the 5-leaf stage, established cuttings were inoculated with wild type B1, B3, and mutant PTSO transcripts. Dot blot analysis was used for the initial screening of recombinants at 30, 45 and 60 dpi. Of the 202 transgenic plants inoculated, 75 provided dot blot hybridization signals. Viral RNA from plants providing even the faintest hybridization signals was further examined by Northern blot analysis, RT-PCR, or both. Three potential recombinants, namely 711 (Fig. 4A, lane 5), 911 (Fig. 4A, lane 11), and 7239 (Fig. 4B, lane 1), were detected in the Northern blot analysis (Figure 4). Another five potential recombinants, namely 7210, 7246, 733, 919, and 9110, were detected using RT-PCR techniques. All eight recombinants were detected by 30 dpi. Viral RNA extracted at a later time did not give rise to additional recombinants. For the convenience of nomenclature, these potential recombinants were named after the transgenic plant lines in which they developed, followed by the cutting number. For example, recombinant 911 was detected from the transgenic plant 9-1 cutting number one. These eight potential recombinants were transferred to non-transgenic N. benthamiana plants, and the viral RNA was extracted at 30 dpi. The BMV 3’ terminal primer QE was used for first-strand cDNA synthesis. Primer RA96 was added for PCR amplification (see Fig. SA). All eight potential recombinants provided 989 bp RT-PCR 51 fiagments (Fig. 5B, lanes 1-8), which are equal in size to the wild type BMV generated fragment (Fig. 5B, lane 9). Sequence analysis of the cDNA clones indicated that the 307-nucleotide deletion at the 3’ end of the polymerase gene was restored in all eight recombinants, and seven of the recombinants contained the single marker mutation at position 2458 (Fig. 6). In addition to the marker mutation, several nucleotide substitutions were observed in the 2a gene ORF (Fig. 7), which resulted in amino acid changes in all but one recombinant virus (recombinant 9110) (Fig. 8). No deletion or insertion was present in the sequenced area of any recombinant. For virus 711, the marker mutation was not present, but one nucleotide substitution was found in the sequenced area of 2a gene ORF. This virus also accumulated at a much lower level than wild type BMV (Fig. 4A, lane 5). Thus, virus 711 could either be a recombinant with the marker mutation reverted to wild type, or a variant of wt BMV. Counting that eight recombinant viruses were generated from a total of 202 challenged plants, 4% of the transgenic plants yielded recombinant virus. RT-PCR from recombinants 711, 7246, and 911 resulted in an additional product of around 700 bp (Fig. 5B, lanes 1, 4, and 6), the same size as the RT-PCR fragment of the inoculated PT50 mutant (Fig. 5B, lane 10). This indicates that PT50, although incapable of systemic movement by itself in N. benthamiana, moved systemically in N. benthamiana in the presence of recombinant viruses, and was transmitted along with the recombinant viruses to new hosts. 52 E: a man a“ :b 1 2345678910111213 2 m a 4—3 4A a % eVNb oan 2 w e (-——3 (—4 4B 53 Figure 4. (A) & (B) Northern blots showing the detection of recombinant viruses. Transgenic plants were inoculated with B1, PTSO, and B3 transcripts. At 30 dpi, viral RNA was extracted from each plant and analyzed in Northern blot using dig-labeled riboprobe HE] (Ahlquist et al., 1981b). Three recombinants were detected in the Northern blot, and were designated recombinant 711 (Fig. 4A, lane 5), 911 (Fig. 4A, lane 11), 7239 (Fig. 4B, lane 1). Recombinants 733 (4A, lane 10) and 7246 (4B, lane 2) did not show up in the Northern blot, but were amplified from systemically infected leaves using RT-PCR techniques (see Fig. 5). Positive control lanes, Fig. 4A, lane 13, and Fig. 4B, lanes 4 and 5, contained virion RNA from wt BMV inoculated nontransgenic N. benthamiana plants. Remaining lanes repreSent attempts to recover recombinant viral RNA from PT50 inoculated transgenic plants. Arrows “9” mark the position of the corresponding genomic and subgenomic BMV RNAs. Asterisks “*” mark the recombinant viruses that carried PT50 mutant for systemic movement. The sign “#” marks the recombinant viruses that did not show up in the Northern blot. 54 (.- 2865 (2E _) 1895 IUA96 5A Lotta— cmhh arezm_rs c——a a—a fi—a nab eVNb «nah c—Nb -b 6 7 8 9 10 11 5 4 5B 55 Figure 5. RT-PCR results of recombinants established in nontransgenic N. benthamiana plants. (A) Schematic drawing of the BMV RN A2 and the site of two primers used for RT-PCR. Primer QE (5’ CGGAATTCTGGTCTCT’ITTAAGAT 3’) was used for first-strand cDNA synthesis. For the PCR reaction, an additional primer RA96 (5’ AGTTTCTAGACTCGTCGAAACTGAAATGGG 3’) was used. (B) Agarose gel electrophoresis of RT-PCR products amplified from viral RNA extracted from recombinants. Order of loading: recombinant 711 (lane 1), 7210 (lane 2), 7239 (lane 3), 7246 (lane 4), 733 (lane 5), 911 (lane 6), 919 (lane 7), 9110 (lane 8), wt BMV (lane 9), and PTSO mutant (lane 10). Arrow f indicates RT-PCR product from full-length BMV RNA2. Arrow >> indicates RT-PCR product from BMV RNA2 deletion mutant PT50. Lane 11 contains 1 Kb size marker (GibcoBRL). 56 Symptom development and virus accumulation of the recombinants Systemic infection of all eight recombinants in barley was confirmed using RT- PCR techniques (data not shown). All eight recombinants accumulated. at very low levels in both barley and N. benthamiana as compared to wt BMV infections. Mild chlorotic local lesions were present on the inoculated barley leaves, and sporadic chlorotic areas were found in uninoculated leaves. No stunting of barley was observed in any of the recombinant viruses. Thus, these recombinant viruses cause symptoms which are distinct from the PT50 deletion mutant, the MD4 mutant carrying the Eca47III mutation, and the wild type BMV in barley. DISCUSSION Involvement of polymerase transgene in RNA recombination Previous studies indicated that viral coat protein transgene could be involved in RNA recombination with an infecting virus (Greene and Allison, 1994). In this study, our data indicated that the polymerase transgene could also be involved in RNA recombination events. Since the PTSO mutant is defective in systemic movement in N. benthamiana, systemic movement of the virus in transgenic plants indicates the recovery of the 307-nucleotide deletion at the 3’ end of the polymerase gene through RNA recombination. Sequence analysis of cDNA clones of the recombinant viruses clearly indicated that the 307-nucleotide deletion was recovered through RNA recombination between the polymerase transgene of BMV RNA2 and PT50 mutant, because the marker mutation Eca47III site, which is only present in the transgene, was present in seven of the eight recombinant viruses. 57 Recombinants 91 l 733 7239 GAUC GAUC GAUC 2458 + Figure 6. Sequence analysis of the recombinants showing that the marker mutation Eco47III is present. RT-PCR fragments amplified from viral RNA extracted fi'om recombinant virus infected plant tissue were cloned into pBS/KS+, and sequenced using the dideoxy sequencing technique. Sequence analysis indicated that seven out of the eight recombinants have the marker mutation at position 2458. Recombinant virus 711 is the only one without this silent mutation. Arrows indicate the single A to G marker mutation found in the recombinant viruses. 58 BMV RNA2 1900 2865 — —---J- - - - 711 l 2438 our: . 7210 l 2255 INK; - 7239 : 2318 WC U/A . 7246 I n I I 1950 1958 NC AC/CA A/G U/C . 733 l l » l l 1991 2002 2104 2215 10m; . (3AA 911 g : 2331 2695 IVCI . 919 l 2082 lUflS . 9110 . 2215 Figure 7. Representative mutations recovered in the polymerase gene of recombinant viruses. The open box represents the polymerase gene open reading frame. The gray shaded box represents the polymerase core region. The nature and extent of each nucleotide substitution is denoted above the horizontal line, and the position is marked below the horizontal line. The asterisk, * , indicates the Eca47III marker mutation present in the recombinant viruses. 59 62.: 00:80 :05 83:3 805.8000: 3:00:05 8?. 0:00.“ 33 0320 500 0580 0: 80:3 :5w0: 0:: 8:00:05 ..\ \.. :0083: 505 80:00 8:58: 500 0580 05 0>000 800802 0500 0580 00w:0:0:= 80850: 30D 8:505:33 80:00 0: 08: 0:0 8000 500 0580 8:2 205m .2 3 Ba .25 .5 as .38 .28 ._ E .3 ._ a 3585888 5 80520 028262. 65 9 05288 0005000 0:03 80:85:38 500 058< 35508000: 00.0500: :0 00:03.8 200 0580 00.80839: .m 8&3 e o e e o o see oo o o o o oo o e o o o o e o a o e o o o e o e o e o e o o o o o o o as o o o o o e o e o e e o e o e o e 00 e e OHHQ e o o e o o o e e e o o e e e e e o o o o e o e e e e o e o o o e a o o o e e e o o e o o e o o o e o o e o o o e e o e o o o e o e o e mHm o o o o o o 0000. o e e o e o o oo o o e e e 000 o o e o o 0 000 00 o o e o o o o o o o a e e o e o o o o o o o e e e .0 e e HHm e o e o o o e o o e o e o o o o o e o o o e o o o e o e e e o e o o o e o o e o e e e o e o o o o o o o o o e e o o o o o e o o e o e o mmb o o e o o o oo 00 e o e o o e o 0 cos a e o o 000 e e e o e e o o a o e>e o o o o 000 e e e e o e o e e o o o o o o o o o e 0 0mm“- 0 e o o o o 000 o o e o o o o o o oo o o e o e o o e o e o o o o o e e o o o o o o e 0 see 0 o o o e o o o e o o o o e e 0 see 0 mQNh o o o o o o oo o oo o o o o e a”. o e o o e e o e o o o e o o o e e o o o o o e o o o 00 o o o o o o o o o o e. o o o e o oo o o OHNN- e a e e a 000000 e e e o e e 0000 o e e a 0000 o o e e e o co so 0. o e o 000 o o ee 0 o o e on 0000 o o e 0000 HAHN- Héfiflfl>¥¥¥4¢098dfi02¢fl¥flmu“m4(mm\\HHMM>¥QA>UQN¥QZMBZ¢®BDUQBMMWHBBMMMKU BB mmh . . bmb we: . . . HHb .................................................................... oHHm ...........................e¢....................................... mam .................................................................... Ham ............................................z..n>................... mmn .................................................................... mmmh ...........................................................H..a..... mum: .................................................................... camp .................................................................... Han ¢4m>mmhHBAMMMUMMAthmUABBHZXNQAOZHhKZMDUhthfidMAZOMQMAHMMMdAMOHmmnma 83 wow . . . . . . OHw 6O In this study, eight recombinants were recovered from 202 transgenic plants inoculated with the PTSO mutant. Sequence analysis of the 989 bp cDNA clones indicated that the marker mutation Eca47III was present in seven of the recombinants. Although virus 711 does not have this single nucleotide mutation, it accumulates at a very low level compared to the wild type, and contains a point mutation typical of other recombinant molecules. Thus, we are considering 711 among the recombinants and assume that the marker mutation reverted to wild type. If 711 were a point mutant that had generated in an accidental wild type inoculation, we would have expected the wild type infection to have out competed 711 which accumulates at a lower titer than wild type. Nucleotide substitutions were present in the sequenced area of all of the eight recombinant viruses. Although the single U to C change did not result in any amino acid change in virus 9110, the low accumulation level and the mild symptom of the virus in barley suggested that other mutations could exist in the 5’ region of the RNA 2 ORF, or other BMV RN As. Our study gave a 4% recombinant recovery rate in the polymerase transgene. Based on student’s t-test at the 95% confidence level, this is not significantly different from the 3% recombinant recovery rate in the coat protein transgene (Greene and Allison, 1994). Previous studies indicated that the concentration of viral RNA and the size of homologous fi'agment positively affect the rate of RNA recombination (Allison et al., 1997; Lai, 1992). Since in a bromovirus infected plant, RNA3 accumulates several folds higher than RNA2, it appears that RNA3 should be more frequently involved in RNA recombination events. However, the polymerase transgene used in this study was 1.6 Kb, 61 900 bases longer than the coat protein transgene used in Greene and Allison’s study (1994) Many studies have indicated that a full-length polymerase gene could provide high resistance to virus infection (Audy et al., 1994; Braun and Hemenway, 1992). In this study, transgenic plants expressing a truncated BMV polymerase gene displayed two distinct phenotypes. Most of the transgenic plants were completely susceptible to wt BMV transcript inoculation, and one plant (4-1) was highly resistant to either BMV virion or RNA inoculation (Fig. 3). This is similar to the result of Tenllado et al. (1996), where a truncated 54 kDa gene of pepper mild mottle tobamovirus (PMMoV) provided high resistance to PMMoV. A recent study demonstrated RNA recombination between the replicase genes of two related tripartite RNA viruses, cucumber mosaic virus (CMV) and tomato aspermy virus (TAV) upon mixed infection (Masuta et al., 1998). The involvement of the viral polymerase transgene in RNA recombination with infecting viruses indicates that further cautions should be taken when designing virus resistant transgenic plants expressing polymerase transgenes. Since truncated polymerase transgenes were found to provide high resistance to viral infection (Tenllado et al., 1996 and this study), and the length of the viral transgene could affect the involvement of the transgene in RNA recombination (Lai, 1992), a truncated polymerase gene is a better choice for generating VRTPs. Thus the results of this study firrther support the suggestion of Allison et al. (1997), which 62 recommends that the transgene should be the smallest viral fragment that provides useful resistance. For the purpose of comparing RNA recombination rates in coat protein and polymerase genes, the complete 3’ UTR was included in the transgene in this study. The 3’ UTR of many plant RNA viruses, including bromoviruses, bears the replication initiation site (Alquist et al., 1981a; Sriskanda et al., 1996; Bachman et al., 1994). Greene and Allison (1996) demonstrated that a truncated 3’ UTR discouraged the involvement of transgene in RNA recombination. The recovery of recombinant viruses in this study further implied the importance of the 3’ UTR in RNA recombination in transgenic plants. Exclusion of the regions for replication initiation is a must for generating VRTPs. RNA 2 plays a role in the systemic spread of the virus PTSO mutant was found to direct RNA replication at nearly wild type levels in barley protoplasts (Traynor et al., 1991). Regarding the replication capability of PTSO mutant in N. benthamiana protoplasts, Smirnyagina et al. (1996) showed that BMV 1a-2a fusion protein, whose 2a protein had the same 125 a C-terminal truncation as PTSO mutant, could support negative- and positive-strand RNA synthesis to approximately 30% of wt level in the presence of the la protein. This implied that, in the presence of wild type 1a protein, PTSO mutant could support viral RNA synthesis in N. benthamiana. The detection of several recombinant viruses in this study fithher demonstrates that PT50 could indeed replicate in N. benthamiana, because RNA recombination requires the involvement of viral RNA polymerase (Kirkegaard and Baltimore, 1986). 63 The 307-nt deletion affected the movement capability of PT50 and caused a 20 to 30-fold decrease in viral RNA accumulation in barley (Traynor et al., 1991). In this study, PTSO mutant was found to be defective in movement in N. benthamiana plants. However, with the recovery of the 3’ deletion, recombinant viruses regained the capability to move in N. benthamiana. This agrees with the results of previous studies, which indicate that replicative proteins play roles in systemic spread and/or host specificity of bromoviruses (Allison et al., 1988; Traynor et al., 1991). Thus, BMV RNA2 is multifirnctional. The 3’ terminal 307-nucleotide RNA encodes firnctions supporting virus movement. A previous study indicated that, in most of the recombinant viruses generated with the coat protein transgene, nucleotide deletions and insertions were fiequently observed in the recovered region (Greene and Allison, 1994). However, in this study, only nucleotide substitutions were present in the cDNA clones of the recombinants. This agrees with the observation that BMV polymerase gene is very conservative. Deletions in the central polymerase core region abolish RNA replication in the barley protoplasts, and deletions in the 5’ and 3’ regions cause decrease in RNA replication, virus accumulation, and symptom development in barley plants (Traynor et al., 1991). Favorable interaction between PT50 mutant and recombinant viruses A discovery of this study is the recovery of the deletion mutant PTSO along with the recombinant viruses in the upper uninoculated leaves. PTSO was shown in this study 64 to be defective in systemic movement in N. benthamiana. However, in the presence of the recombinant viruses, PT50 moved systemically in N. benthamiana, and it was also transmitted along with the recombinant viruses to new hosts (both barley and N. benthamiana). This indicates a favorable interaction between these two viruses. Whether the PTSO mutant played a role in the replication of these recombinant viruses remains unknown. Thus, from a population genetics point of view, RNA recombination between the polymerase transgene and PT50 mutant is a favored event because the generation of viable recombinants which can support the movement of mutant PTSO could lead to a dramatic fitness gain for the mutant virus PTSO. 65 CHAPTER 3 RNA Recombination in Transgenic Plants Expressing Translatable or Untranslatable Viral Transgenes INTRODUCTION Transgenic expression of viral genes was proven to be an effective way of combating viral infections (Grumet, 1995). However, the presence of viral transgenes provides a unique opportunity for RNA recombination with infecting viruses (de Zoeten, 1991). Following the demonstration of RNA recombination between a CCMV coat protein transgene and an infecting virus (Greene and Allison, 1994), studies were carried out to look for ways of generating virus resistant transgenic crops (VRTPs) whose viral transgene would be less likely to be involved in RNA recombination. The study of Greene and Allison (1996) demonstrated that excluding the 3’ UTR in the viral transgene is one effective way of discouraging the involvement of viral transgene in RNA recombination events. Experiments carried out in this study were to determine if a translatable transgene was more or less likely to be involved in RNA recombination than a non-translatable transgene. In the study of Greene and Allison (1994), the transgene transcripts were derived from the 3’ 2/3 of the CCMV CP ORF with a complete 3’ UTR. This fiagment of viral gene was directly excised from the cDNA clone of CCMV RNA3, and no AUG start codon was added at the 5’ end of the viral gene. Thus, translation of the viral gene was 66 unlikely. In this study, N. benthamiana plants were transformed with two modified constructs of the CCMV CP gene used in the RNA recombination experiments of Greene and Allison (1994). One construct had an inframe start codon added at the 5’ end of the truncated CP gene. Theoretically, this viral transgene was translatable. The other construct had an out-of-frame start codon followed by a stop codon, and this viral transgene was untranslatable. For both constructs, three single nucleotide substitutions near the 3’ end of the CCMV CP gene characterized the transgene, and the 144- nucleotide translational enhancer TEV leader sequence (Gallic et al., 1995) was added at the 5’ end of the transgene. Our data indicated that transgenes with an initiation codon were less likely to be involved in recombination events than the transgene containing no initiation codon. Additionally, the translatable transgene was less likely to be involved in a recombination event than a nontranslatable transgene. MATERIALS AND METHODS Plasmid construction Full-length cDNA clones of wild type CCMV RNA], RNA2 and RNA3 are maintained in plasmid pCClTPl, pCC2TP2, and pCC3TP4, respectively (Allison et al., 1988). Infectious transcripts, namely C1, C2, and C3, were synthesized using T7 RNA polymerase (Epicenter) fi'om the Xbal digested DNA plasmids. Plasmid pCC3AG1 was generated through site-directed mutagenesis which introduced three single nucleotide substitutions including a Natl restriction site near the 3’ end of the CP open reading frame (ORF) (Greene and Allison, 1994). Plasmid pCC3 AG3 was generated by removing the SacII-Natl restriction fragment from pCC3AG1 and ligating the Klenow filled ends 67 (Greene and Allison, 1994). The transcripts synthesized from Xbal linearized pCC3 AG3, AG3, were used to inoculate the transgenic plants together with transcripts C1 and C2. Plasmid construction for plant transformation Plasmid pGAMD201(+) is the transformation vector containing the untranslatable viral gene which bears a stop codon seven codons down stream of the start codon (Fig. 1). Primer RA38 (5’ CCATCGCTCCATGGCCAATGGGCAA 3’) was designed to introduce a start codon (bold) and a NcaI restriction site (underlined) at position 1473 of CCMV RNA3. Addition of two nucleotides (bold and italicized) provided a frameshift and introduced a UAA stop codon seven codons down stream of the start codon. An additional primer CCMV 3’ (5’ CAGTCTAGATGGTCTCCTTAGAGAT 3’), which introduces an Xbal site at the 3’ end (underlined), was used in conjunction with RA38 to amplify a 700-bp fragment from pCC3AG1. After Neal-Xbal treatment, the PCR fragment was inserted into a modified plasmid pTL37 (Carrington and Dougherty, 1987), which has an added Ncal site downstream of the tobacco etch virus (TEV) 5’ leader sequence. The resulting plasmid has a 144-bp TEV leader sequence upstream of the 700- nucleotide CCMV CP fragment (from nucleotide 1473-2173). Primer RG26 (5’ AAGTCTAGAAAATAACAAATCTCAACA 3’) anneals to the 5’ end of the TEV leader sequence and it also adds an Xbal site (underlined). Using an additional primer CCMV 3’, the 850-nucleotide PCR fi'agment was amplified and cloned into Xbal digested binary vector pGA643 (An et al., 1988). 68 pCC3TP4 StyI SacH Xbal pCC3AG1 I 3a I pCC3AG3 l 3a I SaeII vNatI pGACCMV I r Transgenic line :.-;.I.€ as: 3-57 pGAMD201(+) ’ * SacII vNatI Transgenic A and B lines, untranslatable GATAGCC ATG GCC AAT GGG CAA GGC TAT IAA NcaI STOP ’ SacH vNatI pGAMD202(+) Transgenic G and H lines, traiTW/ZYK ACCCATCGCTCC ATG GGC CAA GGC AAG GCT NcaI 69 Figure 1. Plasmid construction. Plasmid pCC3TP4 is the full-length clone of CCMV RNA3 (Allison et al., 1989). Plasmid pCC3AG1 was generated through site-directed mutagenesis which introduced three single nucleotide substitutions introducing a Natl restriction site near the 3’ end of the CP ORF (Greene and Allison, 1994). Plasmid pCC3AG3 was generated by removing the SacH-Natl restriction fragment from pCC3AG1 and ligating the Klenow filled ends (Greene and Allison, 1994). The transcripts synthesized from Xbal linearized pCC3AG3, AG3, were used to inoculate the transgenic plants together with wild type transcripts C1 and C2. Plasmid pGACCMV was generated by inserting the StyI-Xbal fiagment of pCC3AG1 into pGA643 and it was used generate transgenic line 3-57 (Greene and Allison, 1994). Plasmid pGAMD201(+) is the transformation vector containing the untranslatable viral gene which bears a stop codon seven codons down stream of the start codon. Primer RA38 (5’ CCATCGCTCCATfiCCAATGGGCAA 3’) was designed to introduce the start codon (bolded) and a Ncal restriction site (underlined) at position 1473. Addition of two nucleotides (bolded and italicized) provided a frameshift and introduced a UAA stop codon. Plasmid pGAMD202(+) is the transformation vector containing the translatable viral gene which bears an inframe start codon. It was constructed using a similar strategy, and differed from pGAMD201(+) in that the introduced start codon was inframe, and no stop codon was introduced. The primer used for mutagenesis was RA45 (5’ ACCCATCGCTCCATGGGCCAAGGCAA 3’) instead of RA38. Transgenic A and B lines were generated using pGAMD201(+), and G and H lines were generated using pGAMD202(+). Open boxes indicate CCMV 3a or CP ORF. Shaded area indicates the central part of the CP ORF. Black boxes indicate the 5’ TEV leader sequence. “ I ” indicates the introduced start codon. “v” indicates the introduced marker mutation in the CCMV CP ORF. An asterisk, *, indicates the introduced stop codon. The nucleotide sequences of the 5’ end of the tmncated viral gene are below the diagram of pGAMDZO 1 (+) and pGAMD202(+). 7O Plasmid pGAMD202(+) is the transformation vector containing the translatable viral gene which bears an infi'ame start codon (Fig. 1). It was constructed using a similar strategy, and is different from pGAMD201(+) in that the introduced start codon was inframe with the remainder of the coat protein gene, and no stop codon was introduced. The primer used for this mutagenesis was RA45 (5’ ACCCATCGCTCCATGGGCCAAGGCAA 3’) instead of RA38. Primer RA45 introduces a Ncal site (underlined) and a start codon (bold). The final plasmids pGAMD201(+) and pGAMD202(+) were sequenced to confirm the correct orientation and joining of the fragments and the correct start and stop codons. The NPT II gene which is located within the left (LB) and right (RB) border sequences of both binary vectors, provides kanamycin resistance to transformed shoots. Bacterial strains The bacterial cell lines used in this experiment were the same as those used in Chapter 2. Since vector pGA643 was used in this study, tetracycline was added to LB and YEP medium to select JM101 and LBA4404 cells containing pGA643 and its derivatives. The tetracycline concentrations used were 10 mg/L for selecting JM101 cells, and 3 mg/L for selecting LBA4404 cells. Agrabacterium-mediated transformation and ELISA tests of regenerated plants The procedures for Agrabacterium-mediated transformation and anti-NPT II ELISA tests of the regenerated plants were the same as in Chapter 2. Four-week-old 71 Nicatiana benthamiana Domin was used for leaf disc transformation. For each transformation construct, two LBA4404 colonies were used for plant transformation. Colonies were designated A and B (untranslatable construct) or G and H (translatable construct). Independent transformants fi'om each colony were numbered, eg. A7 (Fig. 2). Transgenic A and B lines were generated from pGAMD201(+) and were referred to as non-translatable lines. Transgenic G and H lines were generated using pGAMD202(+) and were referred to as translatable lines. Transgenic plant maintenance and inoculation Transgenic plants were maintained and clonally propagated using the same techniques as described in Chapter 2. To test the virus resistance of each transgenic line, 0.5ug (20ug/ml) or 311g (120ug/ml) of each C1, C2 and C3 transcripts was applied to each plant in an inoculation mix of 25111 (see Chapter 2 for procedure). For the recombination experiments, 311g of each C1, C2 and AG3 transcripts was applied to each plant in a mix of 25111 (120ug/ml). The procedure of sap inoculation was the same as for Chapter 1. Viral RNA extraction, Northern and Western blot analysis Procedures for total RNA extraction, viral RNA extraction, Northern blot and dot blot analyses were as described in Chapter 2. To test the CCMV resistance in transgenic plants, dot blot analysis was performed at 14 and 30 dpi. For screening for recombinants, dot blot analysis was performed at 30 and 45 dpi. Dig-labeled probe RA518(+), which is 72 specific for the 3’ UTR of CCMV RNAs, was used to detect viral transgenes in the total RNA, and the CCMV viral RNA accumulation in the dot blot analysis. Procedures for total protein extraction, quantification, and Western blot analysis were as described (Ausubel et al., 1987). Approximately 100 11g of total protein was loaded in each lane. Crude anti-CCMV capsid protein polyclonal antibody was diluted 1:10,000 in the blotting reagent. RT-PCR, cloning and sequence analysis Potential recombinants detected by dot blot analysis were transferred to both nontransgenic N. benthamiana and cowpea (Vigna sinensis (Tamer) Savi), and the viral RNA was extracted at 14 dpi. Viral RNA from 0.2 g of plant tissue was used in the first strand cDNA synthesis using primer CCMV 3’ and the procedure was as described by the manufacture (Boehringer-Mannheim). Primer CCMV 3’ also adds a 3’ terminal Xbal site. An additional primer CCMV 3L (5’ TAAATCGCCGTAACCGC 3’) which anneals at nucleotides 1078-1095 was used for PCR amplification and the procedure was as described (Greene and Allison, 1994). The RT-PCR fragment was first digested with BglII and Xbal, and cloned into the BamI-H and Xbal digested pBS/KS+ using T4 DNA ligase (GibcoBRL). Clones were sequenced manually by dideoxy-sequencing (Amersham). 73 RESULTS Viral transgenic mRNA detection For each plasmid construct, fifty transgenic plants were regenerated from callus and transferred to soil for further examination. Anti-NPT II ELISA tests were used to screen for the expression of the NPT II marker gene which was located adjacent to the viral gene in the transformation vector. Twenty nine transgenic plants with the highest ELISA readings were selected to examine the viral transgene transcription in the total RNA using dig-labeled RA518(+) probe, which is specific for the 3’ UTR of the CCMV RNAs. As shown in Fig. 2, most of the plants gave a single band the size of the transgene (~850 bp). For the untranslatable construct, plant lines A7, B2, B6, B8, and B9 were selected for the recombination experiments. For the translatable construct, plant lines G1, G18, H1, H3 and H4 were selected (Fig. 2). Viral protein detection Total protein was extracted from the ten transgenic plants mentioned above and analyzed by Western blot analysis to evaluate the translation of the viral protein using CCMV polyclonal antibody. The detection limit of the Western blot was quantified as 1 ng of CCMV virion (data not shown). However, when 100 11g of total protein was loaded in each lane, no viral protein was detected from either translatable or untranslatable lines (data not shown). 74 Figure 2. Northern blot analysis. Total RNA was extracted from N. benthamiana transformed with the 3’ 700-nucleotide fragment of CCMV RNA3. Transgenic A and B lines were generated using binary vector pGAMD201(+), whose 700-nucleotide viral gene was modified to have a stop codon seven amino acids down stream of the start codon. Thus transgenic A and B lines are referred to as non-translatable lines. Transgenic G and H lines were generated using binary vector pGAMD202(+), whose 700-nucleotide viral gene was modified to have an inframe start codon. Thus, transgenic G and H lines are referred to as translatable lines. An equal amount of total RNA was loaded in each lane. Dig-labeled RA518(+) probe which is specific for the 3’ UTR of CCMV RNAs was used in the RNA-RNA hybridization. An asterisk, *, indicates the transgenic plants selected for the RNA recombination experiments. 75 Test of virus resistance in transgenic plants Selected transgenic lines with high levels of viral transgene expression were tested for virus resistance prior to the RNA recombination experiments. Clonally propagated cuttings of translatable lines A7, 82, B6, B8, and B9, and untranslatable lines G1, G18, H1, H3, and H4 (indicated with “*” in Fig. 2), were inoculated with CCMV RNA transcripts synthesized in separate reactions. Dot blot analysis was used to detect viral RNA accumulation in the inoculated plants at 14 and 30 dpi. When 0.5ug of each viral RNA was used to inoculate each plant and the concentration of the inoculum was 20 ug/ml of each viral RNA, 100% infection rate was achieved for the nontransgenic control plants. However, for both translatable and untranslatable lines, the infection rate was 20% (Table 1). When 3ug of each CCMV viral RNA was used to inoculate each plant and the concentration of the inoculum was 120 11ng of each viral RNA, the infection rate was 50%, and in one of the three experiments, it reached 80% (Table 1). The CCMV viral RNA in both translatable and untranslatable lines accumulated at a lower level than in the nontransgenic control plants, with a 50% reduction in viral RNA accumulation in untranslatable lines, and an 80-90% reduction in the translatable lines (Table 1). The reduced level of virus accumulation was not observed in the transgenic plants generated by Greene and Allison (1994; 1996), which was transcribing the same 3’ 2/3 of the CCMV CP gene but without the addition of the AUG start codon (observed in Chapter 1 and data not shown). 76 Table 1: Test of virus resistance in transgenic plants. N. benthamiana plants were transformed with the two modified CCMV viral gene which contained the 3’ 2/3 of the CCMV CP gene with a complete 3’ UTR. The translatable construct has an inframe start codon added at the 5’ end of the truncated CP gene. The untranslatable construct has a stop codon following the added start codon. Five transgenic lines were chosen from each construct and tested for CCMV viral RNA resistance. Both dot blot and Northern blot analysis was used to assess the viral RNA accumulation at 14 and 30 dpi. Inoculum CCMV vRNA CCMV vRNA 0.5 ug/plant at 3 ug/plant at concentration concentration 20 ug/ml 120 ug/ml Resistance tests Infection rate Infection rate Viral RNA accumulation level Nontransgenic 100% 100% 100% control Translatable lines G1, 20% 50% 10-20% G18, H1, H3, and H4 Untranslatable lines 20% 50% 50% A7, B2, B6, B8, and B9 77 RNA recombination in transgenic plants Based on the data obtained from the resistant test, 311g of each viral RNA transcripts C1, C2, and AG3 was used at the concentration of 120 ug/ml to inoculate each transgenic plant at the 5-leaf stage. A total of 232 cuttings from the untranslatable lines and 221 cuttings from the translatable lines were inoculated with the AG3 mutant together with wild type C1 and C2. Dot blot analysis at 30 and 45 dpi detected the systemic movement of virus, which indicated potential recombinant viruses. Viral RNA was extracted from plants showing hybridization signals in the dot blot analysis and was examined using RT-PCR techniques. Out of the 453 transgenic plant cuttings inoculated with AG3, around 150 were examined using RT-PCR techniques. Systemic movement of the virus was detected in onlytwo cuttings from the untranslatable lines. No virus was recovered for the translatable lines. Both viruses were detected at 30 dpi. After passing and establishing the potential recombinant virus infections in the nontransgenic N. benthamiana plants, viral RNA was extracted, and cDNA was synthesized, amplified, and cloned. Sequence identification of the transgene marker mutation in the resulting clones indicated that both viruses had recovered the deleted region of the AG3 mutant fiom the transgene. Counting the number of plant cuttings inoculated, the rate of recovery of recombinant virus was 0.85% for the untranslatable lines, and 0% for the translatable lines. These two recombinant viruses were named after the transgenic lines, followed by the number of the cutting fiom which they were derived, B2-13 and B9-5. 78 BgIII Xbal , l l 3a 1 CCMV 3L CCMV 3’ Figure 3. Schematic diagram of CCMV RN A3 and the positions of the two primers used for RT-PCR analysis. Boxes represent wild type CCMV movement protein gene (3a) and capsid gene (CP). Single linesrepresent non-coding sequences. Oligonucleotide CCMV 3’ (5’ CAGTCTAGATGGTCTCCTTAGAGAT 3’) anneals at nucleotides 2158-2173 and was used for first strand cDNA synthesis, and it also introduces an Xbal site (underlined). An additional primer CCMV 3L (5’ TAAAATCGCCGTAACCGC 3’) which anneals at nucleotides 1078-1095, was used for the PCR reaction. Restriction site BglII is at the 3’ end of the 3a gene ORF. RT-PCR fragment was first digested with BglII and Xbal before being cloned into BamHI and Xbal digested pBS/KS+. 79 The complete CP genes and the intercistronic regions of the two recombinants were sequenced. Four nucleotide substitutions and three nucleotide deletions were found in recombinant B9-5, and two nucleotide substitutions were found in recombinant B2-13 (Fig. 4). The nucleotide changes also resulted in several deduced amino acid substitutions (Fig. 5). Dot blot analysis indicated that both recombinant viruses accumulated to levels several fold lower as compared to wild type CCMV in cowpea and N. benthamiana plants (data not shown). In contrast to wt CCMV which provides symptoms in cowpea at 10 dpi, recombinant B2-13 caused mild chlorotic symptoms at around 20 dpi, and recombinant B9-5 remained symptomless till senecence. DISCUSSION The 3’ 2/3 of the CCMV CP gene used by Greene and Allison (1994; 1996) did not provide any virus resistance to transgenic plants. However, the reduced accumulation of viral RNA in the translatable and nontranslatable lines of this study indicated that the addition of an AUG start codon and the TEV translational enhancer sequence at the 5’ of the truncated CP gene conferred moderate virus resistance to the transgenic plants. The translatable lines appeared to be more resistant because the viral RNA accumulated at a much lower level than in the untranslatable lines. This implied that translation of the coat protein might play some role in providing virus resistance to the transgenic plants. However, the truncated viral coat protein was not detected in any transgenic plant in the Western blot when the detection limit was 1 ng CCMV virion. This is similar to the observations of Huntley and Hall (1996), where no BMV coat protein was detected in transgenic rice protoplasts expressing the full-length BMV capsid 80 gene. Possible explanations could be that the coat proteins were made at such low levels that they were below the limit of detection, or the coat protein was very labile and rapidly turned over in the cell, or the antigenic epitopes of the virions were not present on the C- terminal portion of the protein. . Despite considerable efforts to detect recombinant virus, only two recombinants were recovered from the 232 cuttings from the untranslatable lines, and none was recovered from the 221 cuttings from the translatable lines. This gives a 0.85% recombinant recovery rate for the untranslatable lines, and 0% for the translatable lines. Thus, there is a decrease in the recombinant recovery rate for both constructs compared to the 3% recombinant recovery rate of Greene and Allison (1994), with a lower rate for translatable lines. One reason for the overall decrease in the recombinant recovery rate could be the lower infection rate, which was around 50% for both constructs. Another reason could be the strong interference of the viral transgenes on viral RNA accumulation. Huntley and Hall (1996) observed an 80-90% reduction in viral RNA replication in transgenic rice protoplasts expressing the full-length BMV CP gene. This is similar to our results, where viral RNA accumulation was reduced to 50% in the untranslatable lines, and 80-90% in the translatable lines. The purpose of this study was to see if the association of the transgenic RNA with ribosomes during translation would reduce the involvement of the viral transgene in RNA 81 1150 1360 1473 1812 1932 Xbal \ U/C A/G - l l B2 13 I l 1649 1862 A3 A/U A/C U/A C/G - I I ll l B95 I I II l 1290 1315 1552 1587 1766 Figure 4. Diagram showing mutations recovered in the RNA3 of recombinant viruses B2-13 and B9-5. The open boxes represent the 3a gene and capsid gene open reading frames. Shaded box represent the homologous regions between the transgene and inoculum AG3. The nature and extent of each nucleotide mutation is denoted above the horizontal line. “A” indicates deleted nucleotides. Substitution mutations are separated by a slanted line with the wild type to the left and substitution to the right 82 WT 89-5 82-13 WT 89-5 B2-13 60 ' 100 AAAEAKVTSAITISLPNELSSERNKQLKVGRVLLWLGLLPS 133 . 173 VVAAMYPEAFKGITLEQLTADLTIYLYSSAALTEGDVIVHL Figure 5. Capsid protein amino acid sequence of recovered CCMV recombinants B2-13 and B9-5. The amino acid sequences of wild type CCMV (WT) are shown at the top. For each recombinant, single letter amino acid codes are used to denote substitutions, and dots represent unchanged amino acids. 83 recombination. Ribosomes are generally observed to protect mRNA from degradation, and binding of ribosomes to mRNA could exclude the entry of endonucleases (Petersen, 1993). The association of the viral transgene with ribosomes may exclude the binding of the replication complexes. Subsequently, the translatable viral transgene could be less likely to be involved in RNA recombination. Data provided in this experiment did show a decrease in the recombinant recovery rate in the translatable lines comparing to the untranslatable lines. However, since the translatable lines are more resistant than untranslatable lines to CCMV inoculation, it is unclear whether the lower recombinant recovery rate in translatable lines was due to the association of ribosomes during the process of translation, or due to. reduced viral RNA replication. A second hypothesis is that RNA recombination is less likely when the transgene provides resistance to the infecting virus, especially if this viral transgene interferes with the replication of that virus. Transgenes that show no interference with virus replication would be more likely to be involved in RNA recombination, such as the 3% recovery rate of recombinants by Greene and Allison (1994). Transgenes that interfere with the virus infection would be less likely to be involved in RNA recombination, which is shown in the translatable lines in this study. 84 Chapter 4 Test of Virus Resistance in Transgenic Plants Expressing Bromovirus Transgenes INTRODUCTION Coat protein mediated resistance (CPMR) was first demonstrated in tobacco plants transformed with the coat protein gene of TMV (Powell-Abel et al., 1986). Since then, CPMR has been applied to a long list of viruses in a variety of difi‘erent plant species. In the majority of cases, translation of the CP gene leads to high virus resistance, with some exceptions where the virus resistance was mediated by the viral transgene alone. Generally, transcription of the viral transgene is driven by the constitutive 35S promoter of cauliflower mosaic virus (CaMV), which leads to high accumulation of the viral transgene transcripts. In 1991, de Zoeten suggested that, RNA recombination between the constitutively transcribed viral transgene and an infecting virus could create viruses with unique properties. Using bromoviruses as a model system, Greene and Allison (1994) demonstrated that viral transgenes could be involved in RNA recombination with challenging viruses. The transgenic plants used in the Greene and Allison (1994) study were transformed with the 3’ 2/3 of the CCMV CP open reading frame (ORF), and they were not resistant to CCMV infection. However, the addition of an AUG start codon at the 5’ 85 end of this same viral gene provided some CCMV resistance to the transgenic plants (Chapter 3). In Chapter 2, a transgenic line (4-1) which was transformed with a truncated BMV replicase gene was highly resistant to BMV viral RNA and virion inoculation. A remaining question is, can the lull-length CP gene of CCMV provide virus resistance to transgenic plants? This experiment was designed to answer this question. N. benthamiana were transformed with either a translatable full-length CCMV CP gene, an antisense full- length CCMV CP gene, or an antisense truncated CCMV CP gene. Twenty-three transgenic lines were tested for CCMV resistance and the data indicated that CCMV CP gene provided CCMV resistance to transgenic plants at various levels and in different constructs. The results of this resistance test are compared with the results of previous studies (Chapters 2 and 3; Greene and Allison, 1994). MATERIALS AND METHODS Plasmids for the synthesis of infectious transcripts Full-length cDNA clones of wild type CCMV RNA], RNA2 and RNA3 are maintained in plasmid pCClTPl, pCC2TP2, and pCC3TP4, respectively (Allison et al., 1988). Infectious transcripts, namely C1, C2, and C3, were synthesized using T7 RNA polymerase (Epicenter) and the Xbal digested plasmid DNA (Allison et al., 1988). Plasmid pCC3 AGl was generated through site-directed mutagenesis of pCC3TP4 by introducing three nucleotide substitutions which also create a unique Noll site near the 3’ end of the CCMV CP gene (Greene and Allison, 1994). 86 Plasmid construction for plant transformation Plasmid pGAWS4 is the transformation vector containing the firll-length CCMV CP gene. To generate plasmid pGAWS4 (Fig. 1), pCC3TP10 (Pacha and Ahlquist, 1992), a full-length CCMV RNA3 cDNA clone which contains SaII site at the 3’ end of the CP gene, was digested with SalI and the sticky ends were filled using Klenow DNA polymerase. This plasmid was treated with Xbal to excise an 800-nucleotide fragment containing the full-length CCMV CP and the complete 3’ UTR. This 800-nucleotide fragment was cloned into pTL37 (Carrington and Dougherty, 1988), which was first digested with NcoI, blunt-ended with Klenow, and further digested with Xbal (underlined). Plasmid pTL37 contains the TEV leader sequence for translational enhancement upstream of the polylinker. Molecular markers of pCC3AGl were introduced into the resulting plasmid by replacing the SacH-Xbal fragment with that of pCC3AG1. The resulting plasmid had the l44-nucleotide TEV leader sequence joined with the full-length CCMV CP gene followed by the fiJll-length 3’ UTR. Primer RG26 (5’ AAGTCTAGAAAATAACAAATCTCAACA 3’) anneals to the 5’ end of the TEV leader sequence and it also adds an Xbal site (underlined). Using an additional primer CCMV 3’ (5’ CAGTCTAGATGGTCTCCTTAGAGAT 3’) which anneals to the 3’ end of CCMV RNA3 and also adds an Xbal site, the 950-nucleotide PCR fi'agment was amplified and cloned into Xbal digested binary vector pGA643 (An et al., 1988). 87 pCC3TP4 3a I pCC3AGl Styl SacII Natl r 3a 1 pGACCMV Plant 3-57 pGAWS4 Transgenic M and N lines, messenger-sense full-length CP gene pGAMD101(-) . I Transgenic S lines, antisense full-length CP gene pGAMD202(—) Transgenic X and Y lines, antisense truncated CP gene l [3:233:12 NotI Sad] 88 Figure l. Plasmids construction for plant transformation. Three nucleotide substitutions, which also create a unique N011 site, were engineered into plasmid pCC3TP4 to create pCC3AGl (Greene and Allison, 1994). Transformation vector pGACCMV was used in Greene and Allison (1994) to generate transgenic plant 3-57. Plasmid pGAWS4, the transformation vector containing the full-length CCMV CP gene, contains the 144- nucleotide TEV leader sequence (black box) placed 5’ of the full-length CCMV CP gene. The 3’ UTR of CCMV RNA3 was retained in pGAWS4, as were the marker mutations which generate the unique NotI site. Plasmid pGAMDlOl(-) is an antisense construct of pGAWS4 without the TEV leader sequence. Plasmid pGAMD202(-) is the transformation vector containing the antisense 3’ 2/3 of the CCMV CP gene. 89 Plasmid pGAMD101(-) is an antisense construct of pGAWS4 without the translational enhancing TEV leader sequence (Fig. 1). This plasmid was constructed by inserting the Sail-Xbal fragment of pCC3TPlO (Pacha and Ahlquist, 1992) into HpaI and Xbal treated pGA643, with the sticky SalI end blunted with Klenow. Plasmid pGAMDlOl(-) contains no distinguishing marker mutations within the transgene. Plasmid pGAMD202(-), the transformation vector containing the antisense truncated CCMV CP gene, was generated by digesting pCC3AG1 with SIyI at nucleotide 1477 and filling the sticky ends using Klenow DNA polymerase (GibcoBRL). Following phenol-chloroform extraction and ethanol precipitation, the DNA was treated with Xbal to excise the 700-nucleotide fragment containing the 3’ two thirds of the CP gene plus the 3’ UTR. This 700-nucleotide fragment was purified fi'om an agarose gel using GeneClean Kit (BiolOl), and ligated into HpaI-Xbal digested pGA643 (An et al., 1988). These three transformation vectors have the NPT II marker gene adjacent to the viral gene and within the lefi (LB) and right (RB) Agrobacterium border sequences. This NPT II gene provides kanamycin resistance to the transformants. The correct nucleotide sequence and orientation of the inserts of all three binary vector plasmids was confirmed by dideoxy-sequencing (Amersham). Bacterial strains The bacterial cell lines used in this experiment were the same as those used. in Chapter 2. Since vector pGA643 was used in this study, tetracycline was used to make 90 LB and YEP medium to select JM101 and LBA4404 cells containing pGA643 or its derivatives. The concentrations used were 10 mg/L for selecting JM101 cells, and 3 mg/L for selecting LBA4404 cells. Agrobacten'um-mediated transformation and ELISA tests of regenerated plants The Agrobacterium-mediated transformation procedure and the anti-NPT II ELISA tests of regenerated plants were as described in Chapter 2. F our-week-old Nicotiana benthamiana Domin plants were used for leaf disc transformation. For each transformation construct, several LBA4404 colonies were selected for plant transformation. Each colony was given a letter name, such as M. Independent transformants generated using this colony were designated M followed by a transformant number, such as transgenic line M2 (Fig. 2A and 3A). Transgenic M and N lines were transformed with the full-length translatable CP gene using the transformation vector pGAWS4. Transgenic S line was transformed with the antisense lull-length CCMV CP gene using transformation vector pGAMDlOl(-). Transgenic X and Y lines were transformed with the antisense 3’ 2/3 of the CCMV CP gene using transformation vector pGAMD202(-). Plant maintenance and inoculation for CCMV resistance test Transgenic plants were maintained and clonally propagated as in Chapter 2. To inoculate transgenic plants with infectious CCMV RNA transcripts, approximately three micrograms of each viral RNA was applied to each plant at the concentration of 91 lZOug/ml (see Chapter 2 for procedure). The procedure of sap inoculation was the same as for Chapter 1. Total RNA extraction, Northern and Western blot analysis Procedures for total RNA extraction, Northern and dot blot analysis were as described in Chapter 2. Total RNA was extracted from transgenic plants and analyzed for viral transgene transcripts by Northern blots. To test the CCMV resistance in transgenic plants, dot blot analysis was performed at 14 and 30 dpi. Dig-labeled probe RA518(+), which is specific for the 3’ UTR of CCMV RN As, was used in the Northern blot analysis to detect the viral transgene in M and N lines, and in the dot blot analysis to detect the viral RNA accumulation. Dig-labeled probe RA518(-), which is specific for the 3’ UTR of negative sense CCMV RNAs, was used to detect the antisense viral transgene in X, Y, and S lines. Total protein was extracted from the transgenic X and Y lines expressing the full-length CCMV CP gene and analyzed by Western blot. The procedures of protein extraction and Western blot analysis were as described in Chapter 3. RESULTS Viral transgenic mRNA detection Trangenic M and N lines were transformed with the full-length translatable CP gene using the transformation vector pGAWS4. Transgenic S line was transformed with the antisense full-length CCMV CP gene using transformation vector pGAMDlOl(-). Transgenic X and Y lines were transformed with the antisense 3’ 2/3 of the CCMV CP gene using transformation vector pGAMD202(-). For each plasmid construct, fifty transgenic plants were regenerated from the callus tissue and transferred to soil for 92 XI X1 X3 X4 X5 X6 # X7 X9 # Xll # Y1 # Y2 Y3 # Y4 Y5 Y6 Y7 Y9 Y12 Sl # $2 # $6 # S9 # 810 # $12 # 516 # Sl7 # Figure 2. Northern blot analysis reflecting viral transgene transcription in transgenic plants. (A) Transgenic M and N lines were transformed with a full-length translatable CP gene using transformation vector pGAWS4. (B) Transgenic X and Y lines were transformed with an antisense constnrct of the 3’ 2/3 of the CCMV CP gene using transformation vector pGAMD202(-); transgenic S lines were transformed with an antisense construct of the lull-length CCMV CP gene using transformation vector pGAMD101(-).Total RNA from an equal amount of transgenic plant leaf tissue with high anti-NPT H ELISA readings was loaded in each lane. Dig-labeled probe RA518(+), which is specific for the 3’ UTR of the positive sense CCMV RNAs, was used to detect the viral transgene in M and N lines. Dig-labeled probe RA518(-), which was specific for the 3’ UTR of the negative sense CCMV RNAs, was used to detect the antisense viral transgene in X, Y, and S lines. Transgenic plant lines marked with a “#” were used to test virus resistance. 93 further examination. Anti-NPT H ELISA tests were used to screen for expression of the NPT II marker gene which was located adjacent to the viral gene in the transformation vector. A total of thirty-six transgenic lines were examined for viral transgene accumulation in the total RNA. As shown in Fig. 2, a band at the expected transgene size was detected in most of the plants. For transgenic lines M and N, 29% of the transgenic plants regenerated from tissue culture were sterile. Although all of these plants had flowers, no seeds were produced. Viral protein detection For transgenic lines expressing the full-length CCMV CP, Western blot analysis was used to analyze the translation of viral protein. As discussed in Chapter 3, the detection limit of the Western blot was approximately 1 ng wild type CCMV virion. A small amount of viral protein (~20 KDa) was detected in transgenic lines M7 and N1 when lOOug of the total protein was loaded in each lane. These two bands are less strong than the control lane with 1 ng CCMV virion, thus the concentration of the viral protein in lines M7 and N] was less than 1 ng/lOOug total protein. No band was observed in the Western blot analysis of the other transgenic lines (data not shown). Test of virus resistance in transgenic plants Transgenic plants with a range of viral transgene accumulation (indicated with “#” in Fig. 2) were selected for the resistance test. These included: five lines transformed with antisense 3’ 2/3 of the CCMV CP, eight lines transformed with antisense full-length CCMV CP, and five lines transformed with messenger sense full-length CCMV CP. 94 Three micrograms of each wild type CCMV RNA transcripts (C1, C2, and C3) were applied to each transgenic plant at the concentration of 120ug/ml. Systemic CCMV infections were analyzed by dot blot at 14 and 30 dpi. The dot blots shown in Fig. 3 were carried out at 14 dpi, and similar viral RNA accumulation patterns were observed at 30 dpi. The results of this resistance test were compared with the results of previous studies (Chapters 2 and 3; Greene and Allison, 1994; 1996) and are described in Table 1. Translatable CP transgene and virus resistance Transgenic plants expressing full-length translatable CP gene displayed two types of virus resistance, and this resistance was not correlated with the expression of the coat protein. Viral coat protein expression was observed only in lines M7 and N1. Both showed reduced level of viral RNA accumulation. However, for several other transgenic lines such as M2 and N2, no viral protein expression was detected, but both plant lines showed a reasonable level of resistance (Fig. 3A). For most of the M and N lines of transgenic plants, levels of viral transgenic mRNA correlated with levels of virus resistance (see Fig. 2 and 3, transgenic lines M2, M3, M7, M9, N1, N3, N4, and N5). One exception is transgenic line N2, which displayed a high level of resistance despite its low accumulation of transgenic transcripts (Fig. 2A and 3A). In the experiments testing virus resistance in transgenic plants expressing a truncated CCMV CP gene (Chapter 3, G and H lines), a level of resistance similar to that in transgenic plant M2, M7, and N2, was observed (data not shown). Thus, a truncated 95 M2 M3 M7 M9 N2 N3 N4 N5 S1 82 S6 S9 S 10 S 12 S 16 S 17 X6 X9 x11 Y1 Y3 (+) (-) (-) M1 M1 N1 (+) A B Figure 3. Dot blot analysis of virus resistance in transgenic plants expressing viral genes. (A) & (B) Clonally propagated cuttings of transgenic plants were inoculated at the 5-leaf stage with 3 ug each of CCMV RNA transcripts at the concentration of 120 ug/ml. At 14 dpi, viral RNA accumulation was detected by dot blot analysis with the dig-labeled RA518(+) probe. In Fig. 3A, the order of the samples in the dot blot is shown in the grid above the blot, and in Fig. 38, each sample number is above the blot. Transgenic M and N lines were transformed with a full-length translatable CP gene. Transgenic S line was transformed with an antisense construct of the full-length CCMV CP gene. Transgenic X and Y lines were transformed with an antisense construct of the 3’ 2/3 of the CCMV CP gene. The same patterns of viral RNA accumulation were observed at 30 dpi. (+) designates RNA fi'om CCMV infected nontransgenic plants. (-) designates RNA from non-inoculated N. benthamiana. 96 Table 1: Test of Virus Resistance. Transgenic plants were challenged with either BMV C“) or CCMV and analyzed by dot blots. Five “+” indicate full resistance, one “+” indicate low level of resistance. Minus “-“ indicates fully susceptible plants. Transgene Transgenic Lines Transgene Resistance Transcription 3’ 1.6 Kb BMV polymerase 3-5“ H .- gene. . 4-1“ H- +++-H- Potential translation could start 6-9“ ++ - at the beginning of the 7-1 * +++ - transgene where an inframe 7.2!!R ++ - start codon is present 7.3:: +1- _ (nucleotide 1259). 94: +++ - 3’ 2/3 CCMV CP gene with no 3-57 ++ - start codon and TEV leader A69 ++ - sequence added (G&A, 94;96) A83 ++ _ 3’ 2/3 CCMV CP gene with a A7 H- + stop codon following the 82 +++ ++ added start codon. TEV leader B6 +++ ++- sequence was added at the 5’ 133 +++ .H. of the transgene.Untranslatable 139 +++ ++ 3’ 2/3 CCMV CP gene with G1 +++ +++ inframe start codon. TEV G18 +++ +++ leader sequence was added at H] +++ +++ the 5’ end of the transgene. H3 +++ +++ Translatable. H4 +++ +++ Full-length translatable CCMV Ml - + CP gene, with TEV leader M2 ++ ++ sequence added at the 5’ end M3 - +- of the transgene. Messenger M7# H H sense. M9 ++ - N l # +++—l- ++ # - coat protein detected in N2 4., +++ Western blots N3 +_ _ N4 +- +- N5 -H- + F ull-length antisense CCMV Sl - + CP gene. 82 + +++- S6 +—+ +++ No TEV leader sequence 39 +++ +++ added. S 10 +++ .. S 12 ++ - S 16 ++ - S 17 + ++ 3’ 2/3 CCMV CP gene. X6 +++- +- Antisense. X9 +++ +- X11 +++ - No TEV leader sequence Y1 +++ +- added. Y3 +++ +- 97 CCMV CP gene with added AUG start codon consistently provided CCMV resistance while the selected full-length CCMV CP lines provided various levels of resistance. Antisense transgene and virus resistance CCMV resistance in transgenic plants expressing full-length antisense CP gene provided variable results. For transgenic lines S2, S6, S9, and 817, high viral transgene accumulation seemed to correlate with high level of virus resistance because the accumulation level of viral RNA was highly reduced (Fig. ZB, Fig 3A). However, this correlation was not observed in transgenic lines 810, $12 and 816 where a highlevel of viral transgene accumulation resulted in low virus resistance (Fig. 3A). For transgenic line 81, low viral transgene accumulation resulted in an intermediate level of resistance. Transgenic lines transformed with tmncated antisense CP gene, lines X and Y, had slightly reduced viral RNA accumulation and displayed low virus resistance (Fig. 3A). Compared to the antisense lull-length CP gene (S lines), antisense truncated CP gene (X and Y lines) provided lower resistance. DISCUSSION Data provided in this experiment indicate that the CP gene of CCMV provides virus resistance to transgenic plants. Several transgenic plants expressing the translatable full-length CP also had a reduced level of viral RNA accumulation (Fig. 3A). This resistance was not correlated with the translation of the CP protein. Although the 3’ 2/3 of the CCMV CP gene did not provide any resistance to transgenic plants (Greene and Allison, 1994; 1996), data in Chapter 3 indicated that the addition of an AUG start codon 98 resulted in a lower infection rate (50% success) and lower viral RNA accumulation levels. This resistance was observed in both translatable (G and H lines) and untranslatable transgenic lines (A and B lines), although viruses accumulated even lower in translatable lines (BO-90% reduction in translatable lines, 50% reduction in untranslatable lines). Several transgenic plants expressing the full-length CP gene (M and N lines) exhibited a similar level of resistance as the plants expressing the truncated translatable CP gene (G and H lines). However, when antisense CP gene was used, full- length CP gene provided better resistance than truncated CP gene. The data provided in this study and previous studies (Greene and Allison, 1994; Chapter 3) suggested that the mechanisms of CPMR in CCMV are complex. Although the addition of a start codon at the 5’ end of the truncated CCMV CP gene provided virus resistance, truncated CP transgene expression did not provide a detectable amount of protein. Therefore, the presence of the viral protein may not be required for virus resistance. This is especially true in transgenic plants expressing the full-length CP gene, where several transgenic lines lacking detectable coat protein expression had high virus resistance (Fig. 3A). Using transgenic rice expressing the fiill-length BMV CP gene, Huntley and Hall (1996) found that a viral protein product was not likely to be the bioactive agent conferring resistance to BMV because transgenic lines with 80 to 90% reduction in progeny virus accumulation did not have a detectable amount of coat protein. Huntley and Hall (1996) concluded that virus resistance in transgenic plants expressing BMV CP gene was more likely mediated through viral RNA rather than through the coat protein. However, results in this study indicated that, for the two constructs with a 99 translatable 3’ 2/3 CP transgene, translatable lines showed a higher level of resistance than untranslatable lines because the viral RNA accumulated at a lower level in translatable lines (Chapter 3; Table 1, this chapter). Thus, we are unable to rule out the possibility that trace amounts of coat protein could also play some role in virus resistance in CCMV. Substantial evidence indicates that segments of the BMV genome regulate plus- strand (Pogue et al., 1990), minus-strand (Dreher and Hall, 1988; Dreher et al., 1989), and subgenomic RNA synthesis (Marsh et al., 1988). A strong interference of transgenic BMV mRNA on BMV viral RNA replication was observed in transgenic rice protoplasts expressing the BMV polymerase gene and the antisense BMV RNA3 (Huntley and Hall, 1996). In each case, viral RNA accumulation was decreased to 5 to 30% of that in wild type (Huntley and Hall, 1996). In another study, sense and antisense transcripts of the (-) strand promoter of BMV interfered with progeny RNA accumulation in a concentration- dependent manner (Huntley and Hall, l993). These observations implied that the interference of transgenic mRNA on viral RNA replication could play a major role in the resistance observed in transgenic plant 4-1 (Chapter 2) and transgenic S lines (this chapter). Transgenic plant 4-1 which was transformed with a truncated BMV 2a gene was completely resistant to BMV in that no viral RNA accumulation was observed in repeated experiments (Chapter 2). For transgenic S lines which was transformed with an antisense CCMV CP gene, high levels of viral transgene accumulation were associated with high levels of virus resistance. 100 In plants expressing firll-length or truncated CP genes, the viral RNA accumulation was generally reduced, but not completely eliminated. This was different from the resistance provided by the truncated BMV polymerase gene in transgenic plant 4-1, where the viral RNA accumulation was completely eliminated (Chapter 2). This data agrees with the general observations that a polymerase transgene can provide stronger resistance than a CP transgene. It is also another example that a truncated polymerase transgene is sufficient to provide strong virus resistance to transgenic plants. These above observations did not give direct evidence that gene silencing was involved in and was responsible for the resistance provided by the viral transgene, although we could not rule out the possibility that gene silencing played some role in virus resistance observed in this study. In conclusion, our data indicates that CCMV CP gene can provide virus resistance to transgenic plants. Although high levels of viral transgene accumulation is generally correlated with high levels of virus resistance, one exception was found in line N2, where a low level of transgene expression resulted in a high level of resistance. While full- length polymerase and CP genes were used in many other studies to generate VRTPs, truncated polymerase and CP transgenes were shown in this study to provide strong resistance to transgenic plants. Thus, our data indicates that a smaller viral transgene fragment and a lower level of viral transgene expression may provide useful viral resistance. Viral transgenes with both properties would be less likely involved in RNA recombination with infecting viruses. lOl SUMMARY AND CONCLUSIONS The development of Agrobacterium-mediated transformation techniques has enabled plant breeders to generate virus resistance transgenic crops (VRTPs) ' by introducing viral genes into plants. These viral genes are usually cDNA fragments of plant RNA viruses, and transcription of these viral transgenes driven by constitutive promoter leads to accumulation of the viral transgene transcripts in the cytoplasm. RNA recombination is a natural process that occurs during viral RNA replication and has the capacity to unite unrelated RNA fi'agments. Upon viral infections of transgenic plants, RNA recombination between the viral transgene and an infecting virus could generate chimeric viruses with distinct properties. Using a bromovirus model system, Greene and Allison (1994) demonstrated RNA recombination between a bromovirus capsid transgene with a complete 3’ untranslated region (UTR) and an infecting virus. Three percent of the plants produced recombinant virus, but a decrease in recombinant recovery rate was found when the 3’ UTR was deleted (Greene and Allison, 1996). Experiments in this study were based upon these two previous studies and attempted to expand our knowledge of RNA recombination in plants with viral transgenes. An additional goal was to understand and predict the risks of releasing virus resistant transgenic plants to the field. In Chapter 1, a full-length minus strand copy of the viral transgene was detected in virus infected transgenic plants transcribing the CCMV CP gene with a complete 3’ UTR. This data provides an explanation for the difference in recovery of recombinant 102 virus in transgenic plants with a complete or deleted 3’ UTRs (Greene and Allison, 1994; 1996). The presence of both plus and minus strand copies of the viral transgene in a single infected cell would increase the incidence of interactions between the viral transgene and an infecting virus. Additionally, the presence of both the transgene and its complement would enable recombination to occur during either plus or minus strand RNA synthesis. In Chapter 2, a bioassay experiment similar to Greene and Allison (1994) was designed to assess the involvement of a viral polymerase transgene in RNA recombination. Nicotiana benthamiana plants were transformed with a segment of the BMV replicase gene which bore a silent marker mutation that provided an Eco47III restriction site. The challenging virus, PT50, included a 307-nucleotide deletion near the 3’ end of the polymerase gene’open reading frame (ORF) in BMV RNA2. While PT50 remained capable of replication in N. benthamiana, it was unable to move systemically. Transgenic plant lines contained the segment of the BMV polymerase gene that was absent in PT50 mutant. Thus, systemic movement of the virus in PT50 infected transgenic plants signaled the restoration of the polymerase gene through RNA recombination. Eight of 202 challenged transgenic plants yielded systemic infection. Sequence analysis of cDNA clones of the RNA2 components from these infections revealed molecular marker Eco47III site which was originally present only in the viral transgene. Thus, these viruses were generated through recombination between the viral transgene and the infected PT50. The recovery rate of recombinant was 4%, which is comparable to the 3% recovery rate observed with the capsid transgene. 103 Experiments in Chapter 3 were designed to determine if a translatable transgenic mRN A is more or less likely to be involved in RNA recombination than a nontranslatable version of the same transgene. The viral gene fragment used for transformation by Greene and Allison (1994) was excised directly from the CCMV RNA3 cDNA clone and no AUG start codon was added to encourage translation of the transcript. Thus, translation of the viral gene was unlikely. In this study, N. benthamiana plants were transformed with two modified constructs of the CCMV CP gene used by Greene and Allison (1994). One construct had an inframe start codon added at the 5’ end of the truncated CP gene and theoretically, this viral transgene was translatable. The other construct had an out-of-frame start codon followed by a stop codon, and this viral transgene was untranslatable. For both constructs, three single nucleotide substitutions near the 3’ end of the CCMV CP gene characterized the transgene, and the translational enhancer TEV leader sequence was added at the 5’ end of both transgenes. The infecting virus was the AG3 mutant used by Greene and Allison (1994), which lacks the 3’ 1/3 of the CCMV CP ORF. Only two recombinant viruses were recovered from 232 clonally propagated plants from untranslatable lines, and none was recovered from 221 translatable plant lines. This gave a 0.85% recombinant recovery rate for untranslatable lines, and 0% for translatable lines. Thus, there was an overall decrease in the recombinant recovery rate compared to the 3% rate of Greene and Allison (1994), with a lower recovery in translatable lines. 104 It is complicated to explain the data because both translatable and untranslatable lines were resistant to CCMV viral RNA inoculation and a 50% infection rate was observed for both lines. This could be one reason of the overall decrease in recombinant recovery. The translatable lines showed higher resistance because the viral RNA accumulation was reduced to 10-20% of that in nontransgenic N. benthamiana plants. For untranslatable lines, a 50% reduction in viral RNA accumulation was observed. Thus, it is unclear whether the lower recombinant recovery in translatable lines was due to the association the transgenes with ribosomes during translation, or due to reduced viral RNA replication. Additionally, this study indicated that RNA recombination is less likely when the transgene provides resistance to the infecting virus, especially if this viral transgene could interfere with the replication of that virus. Chapter 4 experiments established that CCMV CP gene could provide virus resistance to transgenic plants. N. benthamiana plants were transformed with either a full- length CP gene, an antisense lull-length CP gene, or an antisense truncated CP gene. Twenty three transgenic lines were tested for CCMV resistance. Resistance test results in Chapters 2 and 3 were also analyzed and compared with the new data described in Chapter 4. While full-length messenger sense and antisense CP genes provided virus resistance, truncated capsid and polymerase genes also provided good resistance. Variations in resistance were observed in different transgenic lines. Based on this thesis research data and previous results obtained in our lab (Greene and Allison, 1994; 1996), several conclusions are drawn below: 105 1) 2) 3) 4) A complete 3’ UTR which bears the replication initiation site allows the synthesis of minus strand copies of the viral transgene, and thus permit RNA recombination to occur at an easily detectable frequency, presumably during both plus and minus strand synthesis. Thus, exclusion of the regions required for replication initiation is a must for generating VRTPs with a low risk for generating recombinant viruses. Both capsid and polymerase viral transgenes can be involved in RNA recombination with infecting viruses and their recombinant recovery rates are comparable. Thus, cautions should be taken in generating VRTPs using either viral gene. A translatable transgene is less likely to be involved in RNA recombination, especially if the transgene provides resistance to the infecting virus. Truncated bromovirus capsid and polymerase transgenes provided virus resistance to transgenic plants. Such smaller viral transgenes may be more desirable than complete genes as they should less likely contribute to viral evolution through RNA recombination. 106 APPENDIX 107 Appendix A Recombinant Virus Isolated from Transgenic Plants Expressing CCMV Coat Protein Gene with a Deleted 3’ UTR The 3’ UTR of bromoviruses bear the replication recognition and initiation site. We have demonstrated that when the 3’ UTR is included in the viral transgene, it may be recognized by the replication complex of infecting viruses, and thus a negative-sense copy of the transgene may be synthesized (Chapter 1). When the 3’ UTR of the viral transgene was deleted, a decrease in the rate of RNA recombination was observed (Greene and Allison, 1994). However, the exclusion of the 3’ UTR from the viral transgene does not guarantee that the transgenic transcripts will not be involved in RNA recombination. Sequence analysis of several recombinant viruses generated in our laboratory indicated that cellular mRNA, which does not have any sequence homology with the 3’ UTR of bromoviruses, could be incorporated into the genome of recombinant viruses through RNA recombination (Greene, 1995). Transgenic plant A69 transcribes the 3’ 2/3 of the CCMV CP gene. Characteristics of the A69 transgene are the deletion of the 3’ 69 nucleotides of the 3’ UTR. While I was working on the project for detecting minus-sense copies of the viral transgene (Chapter 1), a transgenic A69 plant infected with CCMV looked sick (probably because of senesce). Assumed that a recombinant virus which made the plant sick might have been generated through RNA recombination between the viral transgene and 108 CCMV, I transferred the virus to both nontransgenic N. benthamiana and cowpea. While no symptom was observed in N. benthamiana, the inoculated cowpea plants showed severe chlorotic symptoms which were different from wild type CCMV at 10 days post inoculation (dpi). This virus was named A69-1. At 14 dpi, viral RNA was extracted from virus A69-1 and wt CCMV infected cowpea. Viral RNA from 0.1g of the plant tissue was loaded in each lane. Northern blot analysis using dig-labeled probe RA518(+) indicated that virus A69-l had all four RNAs of CCMV and thus was a CCMV variant. Virus A69- 1 also accumulated around three to five- fold higher than wt CCMV in cowpea (Fig. 2). RT-PCR was used to amplify the RNA3 of virus A69-1. Primer CCMV 3’ (5’ CAGTCTAGATGGTCTCCTTAGAGAT 3’), which anneals at the 3’ end of CCMV RN A3, was used for first strand cDNA synthesis (Boehringer-Mannheim). Primer 5’ CC- D/3 (5’ CGCTGCAGTAATACGACTCACTATAGTAATCTTTACCTAAC 3’) which anneals to the 5’ end of CCMV RNA3, was added for PCR amplification. Primers CCMV 3’ and 5’ CC-D/3 add Xbal and P311 sites at the ends of the cDNA, respectively. A 2.1 Kb cDNA was amplified and cloned into PstI-Xbal digested pBS/KS+. The complete coat protein gene was sequenced and only one single nucleotide substitution (G—9C at 1926) was found in the CP ORF. This mutation was one of the three mutations introduced into the transgene of A69 plants (Greene and Allison, 1996). The other two mutations in the transgene of A69 plants were absent in virus A69-l. The presence of the single nucleotide substitution indicated that virus A69-1 could be a recombinant virus which was generated through RNA recombination between the viral transgene of A69 plants and the wild type CCMV. To acquire a marker mutation 109 which is only present near the 3’ end of the CP transgene, two crossovers have to occur between the transgene and CCMV RNA3, and this can only happen during the minus- strand synthesis (see Fig. 3 for the model). However, we could not rule out the possibility that this single G—+C substitution was introduced through polymerase errors during the process of RNA replication. This single G—>C substitution at 1926 is a silent mutation which does not cause any amino acid change in the CCMV coat protein. This mutation is unlikely the genetic determinant of the severe chlorotic symptom, because AGl mutant which bears this mutation causes symptoms similar to wild type CCMV in cowpea (Greene and Allison, 1994). Previous studies indicate that the movement protein of BMV influences virus-host interactions that are crucial to systemic infection. A single Ser 180—)Arg mutation in the 3a cell-to-cell movement protein of a CCMV hybrid conferred compatibility for systemic infection of a new host, and also affected the symptom development (Fujita et al., 1996). In another study, a single amino acid change (Val 266—)Ile) in the 3a movement protein of BMV altered symptom expression in N. benthamiana (Rao and Grantham, 1995). To search for the genetic determinant of the severe chlorotic phenotype of this CCMV variant A69-l, the movement protein gene of the cDNA clone was sequenced and the data is presented in Fig. 3. Two G to A substitutions were found in the 3a gene ORF. Both nucleotide substitutions resulted in amino acid changes, one is Val 17——)Ile, and the other one is Glu 41—)Lys. 110 Figure 1. Infection of cowpea with virus A69-1. Trifoliate leaves fi'om cowpea inoculated with wild type CCMV show typical chlorotic mottling symptom (right). Leaves from cowpea inoculated with virus A69-l produce severe chlorotic symptoms (left). 111 WT CCMV RNA3 I 3a I I A69 Transgene Figure 2. Model for RNA recombination to generate recombinant A69-1. The transgenic mRNA in A69 plants has a 69-nucleotide deletion at the 3’ terminus of the UTR. Thus, initiation of negative-strand synthesis is unlikely. To generate a recombinant virus with a marker mutation which originally exists only in the transgenic mRNA of A69 plants, RNA recombination has to happen during the negative-strand RNA synthesis through two template switches. Filled triangles, A and v, indicate the G—)C marker mutation within the CP transgene (nucleotide 1926). The other two introduced mutations in the transgene of A69 plants create a NotI restriction site near the 3’ end of the CCMV CP gene in the 3’ UTR. These two mutations were not present in the RNA3 of virus A69-1. 112 To determine the amino acid(s) responsible for the severe chlorotic phenotype, experiments were designed to introduce the mutations into wild type CCMV RNA3 clone pCC3TP4 by replacing the restriction fragment of wild type RNA3 with the same fragment bearing the mutations. Plasmid pCC3MD6901 contains both 3a gene mutations and was made by replacing the 1054-nucleotide EcoRV-Sunl fragment of pCC3TP4 with the same fragment of the A69-1 clone. Plasmid pCC3MD6902 has the G to A mutation at position 287 and was made by replacing the EcoRV-Narl fragment of pCC3TP4 with that of the A69-1 clone. Plasmid pCC3MD6903 has the G to A mutation at position 359 and was made by replacing the 950-nucleotide NarI-Sunl fragment of pCC3TP4 with that of the A69-l clone. Following dideoxy sequence analysis to confirm the correct orientation and joining of the fragments, infectious RNA transcripts will be made from the Xbal linearized clones. The mutant RNA3 transcripts (MD6901, MD6902 or MD6903) will be applied to cowpea together with wild type CCMV RNA] and RN A2 transcripts (C1 and C2). Comparing the development of the symptoms of the three mutants with that of virus A69-l, we will establish which amino acid is responsible for the symptom change of CCMV. If the above mentioned experiments fail to determine the genetic determinant of the symptom alteration of A69-1, alternative experiments are: 113 WT CCMV RNA3 F San I A69 transgenic plant A69-l l 3a I l l 1 CR1 G/A G/A 27 287 359 A 17 41 1 . l . . . 50 WT MSNTTFRPFTGSSRTVVEGEQAGAQDDMSLLQSLFSDKSREEFAKECKLG A69—1 ................ I ....................... K ......... B Figure 3. Nucleotide and amino acid changes in virus A69-1 compared with wild type CCMV. (A) Nucleotide changes in RNA3 of virus A69-l compared with wild type CCMV. Only one single nucleotide mutation (G—)C at 1926, indicated by triangle v) was found in the CP ORF of A69-1. This was one of the three marker mutation present originally in the transgene. Two single nucleotide substitutions were present in the 3a gene ORF, and one single nucleotide substitution was found in the 5’ UTR. (B) Deduced amino acid changes in the 3a movement protein of A69-1. Two amino acid changes were deduced according to the nucleotide changes in the RNA3 of A69-1, Val 17—>Ile and Glu 41—+Lys. 114 pCC3TP4 202 306 1255 EcoRV NarI SunI II 3a I A69“ 202 306. 1255 EcoRV NarI SunI v i 3. l I 1 cm G/A G/A 27 287 359 pCC3MD6901 3a I I I G/A G/A 287 359 pCC3MD6902 l 3a l 1 CIA 237 pCC3MD6903 G/A 359 115 Figure 4. Plasmid construction for making infectious transcripts bearing the mutations of virus A69-1. Plasmid pCC3TP4 is the lull-length clone of CCMV RNA3. Plasmid A69-l is the full-length clone the virus A69-1 RNA3. Two G to A substitutions were present in the 3a gene of A69-l virus, and these two nucleotide substitutions result in two amino acid substitutions. To determine the mutations responsible for the altered symptoms in cowpea, three clones were created by replacing the restriction fragment of pCC3TP4 with the same fragment of A69-1. Plasmid pCC3MD690l bears both mutations, and plasmids pCC3MD6902 and pCC3MD6903 bear one of the two mutations. 116 1) Transfer virus A69-1 to local lesion host Chenopodium quinoa and transfer viruses from several local lesions to individual cowpea plants. This allows the isolation of a single virus which contributes to the symptom alteration in cowpea. 2) Clone the three genomic RNAs of the virus (A69-l). By exchanging the RNAs of virus A69-1 with wild type CCMV, we would identify the RNA responsible for the altered symptoms. 3) Sequence analysis would firrther locate the mutation(s) contributing to the virus-host interaction. 4) Inoculate cowpea plants with infectious transcripts containing the mutations and identify the mutation(s) responsible for the symptom development. 5) Host range tests of virus A69-l Previous studies indicate that bromovirus RNA3 is involved in virus-host interactions and contributes to symptom development (Fujita et al., 1996; Rao and Grantham, 1995). BMV RNA2 was found to play a role in host specificity and systemic infection (Traynor et al., 1991). Thus, experiments should be concentrated on the RNA3. 117 LIST OF REFERENCES 118 LIST OF REFERENCES Ahlquist, P., R. Dasgupta, and P. Kaesberg. 1981a. Near identity of 3’ RNA secondary structure in bromoviruses and cucumber mosaic virus. Cell 23: 183-189. Ahlquist, P., Luckow, V., and P. Kaesberg. 1981b. Complete nucleotide sequence of brome mosaic virus RNA 3. J. Mol. Biol.. 153223-38. Ahlquist, P., E.G. Strauss, C.V. Rice, J.H. Strauss, .J. Haseloff, and D. Zimmern. 1985. Sindbis virus proteins nsPl and nsP2 contain homology to nonstructural proteins from several RNA plant viruses. J. Virol. 53:536-542. Ahlquist, P. 1994. Bromoviruses. In: Encyclopedia of Virology. Webster, RG. and A. Granoff, (eds), Vol. 1, Academic Press. Allison, R. F., M. Janda, and P. Ahlquist. 1988. Infectious in vitro transcripts from cowpea chlorotic mottle virus cDNA clones and exchange of individual RNA components with brome mosaic virus. J. Virol. 62:3581-3588. Allison, R. F., M. Janda, and P. Ahlquist. 1989. Sequence of cowpea chlorotic mottle virus RNAs 2 and 3 and evidence of a recombination event during bromovirus evolution. Virology 172:321-330. Allison, R., C. Thompson, P. Ahlquist. 1990. Regeneration of a functional RNA virus genome by recombination between deletion mutants and requirement for cowpea chlorotic mottle virus 3a and coat genes for systemic infection. Proc. Natl. Acad. Sci. USA. 87: 1820-1824. Allison, R. F., W. L. Schneider, and A. E. Greene. 1997. Recombination in plants expressing viral transgenes. Seminars in Virology 7:417-422. An, G., P.R. Ebert, A. Mitra, and SB. Ba. 1988. Binary vectors. Plant Molecular Biology A3: 1-19. Anderson, J.M., P. Palukaitis, and M. Zaitlin. 1992. A defective replicase gene induces resistance to cucumber mosaic virus in transgenic tobacco plants. Proc. Natl. Acad. Sci. USA. 89:8759-8763. Audy, P., P. Palukaitis, S.A. Slack, and M. Zaitlin, 1994. Replicase-mediated resistance to potato virus Y in transgenic tobacco plants. Molecular Plant-Microbe Interactions 7: 15-22. Ausubel, F.M., R. Brent, R.E. Kingston, D.D. Moore, J.G. Seidman, J.A. Smith, and K. Struhl. 1987. Current protocols in molecular biology. Vol. 2. 119 Bachman, E.J., S.W. Scott, G. Xin, and VB. Vance. 1994. The complete nucleotide sequence of prune dwarf ilarvirus RNA 3: implication for coat protein activation of genome replication in ilarviruses. Virology 201: 127-131. Bastin, M., and P. Kaesberg. 1976. A possible replicative form of brome mosaic virus RNA 4. Vriology 72:536-539. Baulcombe, D.C.. 1996. MeChanisms of pathogen-derived resistance to viruses in transgenic plants. Plant Cell 8: 1833-1844. Bendahmane, M., J.H. Fitchen, G. Zhang, R.N. Beachy. 1997. Studies of coat protein- mediated resistance to tobacco mosaic tobamovirus: correlation between assembly of mutant coat proteins and resistance. J. Virol. 71:7942-7950. Benvenuto, E., RJ. Ordas, R. Tavazza, G. Ancora, S. Biocca, A. Cattaneo, and P. Galeffi. 1991. ‘Phytoantibodies': a general vector for the expression of immunoglobulin domains in transgenic plants. Plant Mol. Biol. 17:865-874. Braun, C.J., and CL. Hemenway. 1992. Expression of amino-terminal portions or full- length viral replicase genes in transgenic plants confers resistance to potato virus X infection. Plant Cell 42735-744. Brown, T., and K. Mackey. 1997. In: Current Protocols in Molecular Biology, Unit 4.9. Ausubel, et al (Ed). John Wiley & Sons, Inc. Bujarski, J. J., and A. M. Dzianott. 1991. Generation and analysis of nonhomologous RNA-RNA recombinants in brome mosaic virus: sequence complementarities at crossover sites. J. Virol. 65:4153-4159. Bujarski, .J. J., and P. Kaesberg. 1986. Genetic recombination between RNA components of a multipartite plant virus. Nature (London) 321:528-531. Bujarski, J. J., T. W. Dreher, and T.C. Hall. 1985. Deletions in the 3'-terminal tRNA- like structure of brome mosaic virus RNA differentially affect aminoacylation and replication in vitro. Proceedings National Academic of Science, USA 82:5636-5640. Carpenter, C.D., J.W. Oh, C. Zhang, and A.E. Simon. 1995. Involvement of a stem- loop structure in the location of junction sites in viral RNA recombination. J. Mol. Biol. 245:608-22. Carr, J.P., L.E. Marsh, G.P. Lomonossoff, M.E. Sekiya, and M. Zaitlin. 1992. Resistance to tobacco mosaic virus induced by the 54 kDa gene sequence requires expression of te 54 kDa protein. Mol. Plant-Microbe Interact. 52397-404. Carr, J.P., A. Gal-on, P. Palukaitis, and M. Zaitlin, 1994. Replicase-mediated resistance to cucumber mosaic virus in transgenic plants involves suppression of both 120 virus replication in the inoculated leave and long-distance movement. Virology 199:434- 447. Cascone, P. J., C. D. Carpenter, X. H. Li, and A. E. Simon. 1990. Recombination between satellite RNAs of turnip crinkle virus. EMBO J. 9: 1709-1715. Cooper, RD. 1968. A genetic map of poliovirus temperature-sensitive mutants.Virology. 35:584-596. Creamer, R., and B.W. Falk. 1990. Direct detection of transcapsidated barley yellow dwarf luteoviruses in doubly infected plants. J. Gen. Virol. 23:41-49. Cuozzo, M., K.M. O’Connell, W. Kaniewski, R.X. Fang, N.H. Chua and N.E. Turner. 1988. Viral protection in transgenic tobacco plants expressing the cucumber mosaic virus coat protein or its antisense RNA. Bio/Technology 6:549—557. De Zoeten, G.A.. 1991. Risk assessment: Do we let history repeat itself? Phytopathology 81:585-586. De Zoeten, GA. 1995. Virus resistant plants: heterologous encapsidation and recombination. Virus resistance: biosafety research needs. In: Key biosafety aspects of genetically modified organisms. Landsmann, J., and R. Casper (eds). 11-15. Blackwell Wissenschafis-Verlag. Dinant, S., M. Janda, P.A. Kroner, and P. Ahlquist. 1993. Bromovirus RNA replication and transcription require compatibility between the polymerase- and helicase- like viral RNA synthesis proteins. J. Virol. 67:7181-7189. Domingo, E., D. Sabo, T. Taniguchi, and C. Weissmann, 1978. Nucleotide sequence heterogeneity of an RNA phage population. Cell 13:73 5-744. Domingo, E., E. Martinez-Salsa, F. Sobrino, J.C. de la Torre, A. Portela, J. Ortin, C. Lopez-Galindez, P. Perez-Brena, N. Villanueva, R. Najera, S. VandePol, D. Steinhauer, N. DePolo and J.J. Holland, 1985. The quasispecies (extremely heterogeneous) nature of viral RNA genome populations: biological relevance: a review. Gene 40: 1-8. Domingo, E., and J. Holland. 1994. Mutation rates and rapid evolution of RNA viruses. In: Morse, S. (ed) The Evolutionary Biology of Viruses, Raven Press, New York, pp 161- 183. Domingo, E., J. Holland, C. Biebricher, and M. Eigen, 1995. Quasi-species: the concept and the word. In: Molecular Basis of Virus Evolution. Gibbs, A., C.H. Calisher and F. Garcia-Arenal (Ed) 181-191. Cambridge. 12] Domingo, E., C. Escarmis, N. Sevilla, A. Moya, S.F. Elena, J. Quer, I.S. Novella, and J.J. Holland. 1996. Basic concepts in RNA virus evolution. The FASEB J. 10:859-864. Donson, J., C.M. Kearney, T.H. Turpen, I.A. Khan, G. Kurath, A.M. Turpen, G.E. Jones, W.O. Dawson, and DJ. Lewandowski. 1993. Broad resistance to tobamoviruses is mediated by a modified tobacco mosaic virus replicase transgene. Mol. Plant-Microbe Interactions 6:63 5-642. Drake, J.W. 1993. Rates of spontaneous mutations among RNA viruses. Proc. Natl. Acad. Sci. USA. 90:4171-4175. Dreher, T. W., and T. C. Hall. 1988. Mutational analysis of the sequence and structural requirements in brome mosaic virus RNA for minus strand promoter activity. Journal of Molecular Biology 201:31-40. Dreher, T.W., Rao, A.L.N., and T.C. Hall. 1989. Replication in vivo of mutant brome mosaic virus RN As defective in aminoacylation. J. Mol. Biol. 206:425-43 8. Duarte, E.A., I.S. Novella, S. Ledesma, D.K. Clarke, A. Moya, S.F. Elena, E. Domingo, and J.J. Holland, 1994. The subclonal components of an RNA virus fitness. J. Virol. 67:4295-4301. Evans, R.K., B.E. Haley, and D.A. Roth. 1985. Photoaffinity labeling of a viral induced protein from tobacco mosaic virus. Characterization of nucleotide binding properties. J. Biol. Chem. 260:7800-7804. Fitchen, J.H. and RN. Beachy. 1993. Genetically engineered protection against viruses in transgenic plants. Annu. Rev. Microbiol. 47:739-763. French, R., M. Janda, and P. Ahlquist. 1986. Bacterial gene inserted in an engineered RNA virus: efficient expression in monocotyledonous plant cells. Science 231:1294- 1297. French, R. and P. Ahlquist. 1987. Intercistronic as well as terminal sequences are required for efficient amplification of brome mosaic virus RNA3. J. Virol. 61:1457-1465. Frisch, D.A., L.W. Harris-Haller, N.T. Yokubaitis, T.L.Thomas, S.H. Hardin, and T.C. Hall. Complete sequence of the binary vector Bin 19. Plant Mol. Biol. 27: 405-409. Fuchs, M., M. Fuchs, J.R. McFerson, D.M. Tricoli, J.R. McMaster, R.Z. Deng, M.L. Boeshore, J.F. Reynolds, P.F.'Russe|1, H.D. Quemada, and D. Gonsalves, 1997. Cantaloupe line CZW-30 containing coat protein genes of cucumber mosaic virus, zucchini yellow mosaic virus, and watermelon mosaic virus-2 is resistant to these three viruses in the field. Mol. Breed. 3:279-290. 122 Fuchs, M., D. Gonsalves. 1995. Resistance of transgenic hybrid squash ZW-20 expressing the coat protein genes of zucchini yellow mosaic virus and watermelon mosaic virus 2 to mixed infections by both potyviruses. Bio/technology 13: 1466-1473. Fujita, Y., K. Mise, T. Okuno, P. Ahlquist, and I. Furusawa. 1996. A single codon change in a conserved motif of a bromovirus movement protein gene confers compatibility with a new host. Virology 223:283-291. Gal, 8., B. Pisan, T. Hohn, N. Grimsley, and B. Hohn. 1992. Agroinfection of transgenic plants leads to viable cauliflower mosaic virus by intermolecular recombination. Virology 187:525-533. Gallic, D.R., R.L. Tanguar, and V. Leathers. 1995. The tobacco etch viral 5’ leader and poly(A) tail are functionally synergistic regulators of translation. Gene 165:233-23 8. Gielen, J.J.L., P. de Haan, A.J. Kool, D. Peters, M.Q.J.M. van Grinsven, and RW. Goldbach. 1991. Engineered resistance to tomato spotted wilt virus, a negative-strand RNA virus. Bio/Technology 9: 1363-1367. Golemboski, D.B., G.P. Lomonossoff, and M. Zaitlin. 1990. Plants transformed with a tobacco mosaic virus nonstructural gene sequence are resistant to the virus. Proc. Natl. Acad. Sci. USA 87:6311-6315. Greene, A.E. 1995. Risk assessment of transgenic plants: RNA recombination in transgenic plants expressing viral sequences. Dissertation for the degree of Ph.D. Michigan State University. Greene, A. E. and R. F. Allison. 1994. Recombination between viral RNA and transgenic plant transcripts. Science 263: 1423-5. Greene A.E. and RF. Allison. 1996. Deletions in the 3’ untranslated region of Cowpea Chlorotic Mottle Virus transgene reduce recovery of recombinant viruses in transgenic plants. Virology 225:231-234. Grumet, R.. 1995. Genetic engineering for crop virus resistance. HortScience 30:449- 456. Guo, H.S., and J.A. Garcia, 1997. Delayed resistance to plum pox potyvirus mediated by a mutated RNA replicase gene: involvement of a gene-silencing mechanism. Molecular Plant-Microbe Interactions 10:160-170. Hardy, S.F., T.L. German, L. Sue Loesch-Fries, and T.C. Hall. 1979. Highly active template-specific RNA-dependent RNA polymerase from barley leaves infected with brome mosaic virus. Proc. Natl. Acad. Sci. USA. 76:4956-4960. 123 Hefferon, K.L., H. Khalilian, and M.G. AbouHaidar. 1997. Expression of the PVYO coat protein (CP) under the control of the PVX CP gene leader sequence: protection under greenhouse and field conditions against PVY0 and PVYN infection in the three potato cultivars. Theor. Appl. Genet. 94:287-292. Hemenway, C., R.X. Fang, W.K. Kaniewski, N.H. Chua, and N.E. Tumer. 1988. Analysis of the mechanism of protection in transgenic plants expressing the potato virus X coat protein or its antisense RNA. EMBO J. 7: 1273-1280. Hoekema, A., P.R. Hirsch, P.J.J. Hooykaas, and R. A. Schlperoort. 1983. A binary plant vector strategy based on separation of vir- and T-region of the Agrobacterium rumefaciens Ti-plasmid. Nature 303: 179-180. Holsters, M., D. de Waele, A. Depicker, E. Messens, M.Van Montagu, J. Schell. 1978. Transfection and transformation of A. tumefaciens. Mol. Gen. Genet. 163:181-187. Horsch, R.B., R.T. Fraley, S.G. Rogers, P.R. Sanders, A. Lloyd, and N. Hoffmann. 1984. Inheritance of functional foreign genes in plants (genetic engineering with T-DNA from Agrobacterium tumefaciens, Nicotiana plumbaginifolia). Science 223: 496-498. Huisman, M. J., B. J. C. Cornelissen, C. F. M. Groenendijk, J. F. B01, and L. van Vloten-Doting. 1989. Alfalfa mosaic virus temperature sensitive mutants. V. The nucleotide sequence of TBTS 7 RNA 3 shows limited nucleotide changes and evidence for heterologous recombination. Virology 171:409-416. Huntley, CC. and T.C. Hall. 1993. Interference with brome mosaic virus replication by targeting the minus strand promoter. Journal of General Virology 74:2445-2452. Huntley, CC. and T.C. Hall. 1996. Interference with brome mosaic virus replication in transgenic rice. Mol. Plant-Microbe Interactions 9: 164-170. Janda, M., R. French, and P. Ahlquist. 1987. High efficiency T7 polymerase synthesis of infectious RNA from cloned brome mosaic virus cDNA and effects of 5’ extensions on transcript infectivity. Virology 158:259-262. Kao, C. C., R. Quadt, R. P. Hershberger, and P. Ahlquist. 1992. Brome mosaic virus RNA replication proteins 1a and 2a form a complex in vitro. J. Virol. 66:6322-6329. Kawchuk, L.M., RR. Martin, and J. Mcpherson. 1990. Resistance in transgenic potato expressing the potato leafroll virus coat protein gene. Mol. Plant-Microbe Interact. 3:301-307. Kiberstis, P., L. S. Loesch-Fries, and T. C. Hall. 1981. Viral protein synthesis in barley protoplasts inoculated with native and fractionated brome mosaic virus RNA. Virology 112:804-808. 124 King, A.M.Q., McCahon, D., Slade, W.R. and Newman, J.W.I. 1982. Recombination in RNA. Cell 29:921-928. Kirkegaard, K, and D. Baltimore, 1986. The mechanism of RNA recombination in poliovirus. Cell 47:433-443. Koonin, E.V. and V.V. Dolja, 1993. Evolution and taxonomy of positive-strand RNA viruses: implications of comparative analysis of amino acid sequences. Critical reviews in biochemistry and molecular biology, 28:375-430. Kramer, M.G., K. Redenbaugh. 1994. Commercialization of a tomato with an antisense polygalacturonase gene: the FLAVR SAVR tomato story. Euphytica. 79:293-297. Kroner, P., D. Richards, P. Traynor, and P. Ahlquist. 1989. Defined mutations in a small region of the brome mosaic virus 2a gene cause diverse temperature-sensitive RNA replication phenotypes. J. Virol. 63:5302-5309. Kroner, P.A., B. Young, and P. Ahlquist. 1990. Analysis of the role of brome mosaic virus 1a protein domains in RNA replication, using linker insertion mutagenesis. J. Virol. 64:6110-6120. Lai, M. M.. 1992. RNA recombination in animal and plant viruses. Microbiological reviews 56:61-79. Lai, M.M. 1995. Recombination and its evolutionary effect on viruses with RNA genomes. InMolecular Basis of Virus Evolution. Gibbs et al. (ed.). Cambridge. 119-132. Lane, L. 1981. Bromoviruses, in Handbook of Plant Virus Infections and Comparative Diagnosis, Kurstak, E., (ed). Elsevier Biomedical Press, Amsterdam. P333. Lawson, C., W. Kaniewski, L. Haley, R. Rozman, C. Newell, P. Sanders, and N.E. Turner. 1990. Engineering resistance to mixed virus infection in a commercial potato cultivar: Resistance to potato virus X and potato virus Y in transgenic Russet Burbank. Bio/Technology 8: 127-134. Lindbo, J.A., L. Silvarosales, W.M. Proebsting, and W.G. Dougherty. 1993. Induction of a highly specific antiviral state in transgenic plants - implications for regulation of gene expression and virus resistance. Plant Cell, 5:1749-1759. Loesch-Fries, LS. and T.C. Hall. 1980. Synthesis, accumulation and encapsidation of individual brome mosaic virus RNA components in barley protoplasts. J. Gen. Virol. 47 2323-332. Loesch-Fries, L.S., D. Merlo, T. Zinnen, L. Burhop, K. Hill, K. Krahn, N. Jarvis, S. Nelson, and E. Halk. 1987. Expression of alfalfa mosaic vims RNA 4 in transgenic plants confers virus resistance. EMBO J. 6: 1845-1851. 125 Lommel, S.A., and Z. Xiong. 1991. Reconstitution of a functional red clover necrotic mosaic virus by recombinational rescue of the cell-to-cell movement gene expressed in a transgenic plant. Journal of Cellular Biochemistry Abstracts, 20th Annual Meetings 15A:151. Longstaff, M., G. Brigneti, F. Boccard, S. Chapman, and D. Baulcombe. 1993. Extreme resistance to potato virus X infection in plants expressing a modified component of the putative viral replicase. EMBO J. 12:379-3 86. MacFarlane, S.A. and J.W. Davies. 1992. Plants transformed with a region of the 201- kilodalton replicase gene from pea early browning virus RNA 1 are resistant to virus infection. Proc. Natl. Acad. Sci. USA 89:5829-5833. Maniatis, T., E.F. Fritsch, and J. Sambrook. 1982. Molecular cloning. A laboratory manual. Agarose gel electrophoresis. Page 150-152. Cold Spring Harbor, New York. Marsh, L.E., T.W. Dreher, and T.C. Hall. 1988. Mutational analysis of the core and modulator sequences of the BMV RNA3 subgenomic promoter. Nucleic Acids Research 16:981-995. Marsh, L.E., C.C. Huntley, G.P. Pogue, J.P. Connell, T.C. Hall. 1991a. Regulation of (+):(-)-strand asymmetry in replication of brome mosaic virus RNA. Virology 182:76-83. Marsh, L.E., G.P. Pogue, C.C. Huntley, and T.C. Hall. 1991b. Insight to replication strategies and evolution of (+) strand RNA viruses provided by brome mosaic virus. Oxford Surveys of Plant Molecular & Cell Biology 7 :297-334. Masuta, C., S. Ueda, M. Suzuki, and I. Uyeda. 1998. Evolution of a quadripartite hybrid virus by interspecific exchange and recombination between replicase components of two tripartite RNA viruses. Proc. Natl. Acad. Sci. USA. 95: 10487-10492. Matzke, A.J.M. and M.D. Chilton. 1981. Site-specific insertion of genes into T-DNA of the Agrobacterium tumor-inducing plasmid: an approach to genetic engineering of higher plant cells. J. Mol. Appl. Genet. 1:39-44. Matzke, M.A., and A.J.M. Matzke. 1995. How and why do plants inactivate homologous (trans)gene? Plant Physiol. 107:679-685. Mi, S., and V. Stollar. 1991. Expression of Sindbis virus nsPl and methyltransferase activity in Escherichia coli. Virology 184:423-427. Miller, W.A., J.J. Bujarski, T.W. Dreher, and T.C. Hall. 1986. Minus-strand initiation by brome mosaic virus replicase within the 3’ tRNA-like structure of native and modified RNA templates. J. Mol. Biol. 187:537-546. 126 Mise, K., and Ahlquist, P. 1995. Host specificity restriction by bromovirus cell-to-cell movement protein occurs after initial cell-to-cell spread of infection in nonhost plants. Virology 206:276-286. Misc, K., R.F. Allison, M. Janda, and P. Ahlquist. 1993. Bromovirus movement protein genes play a crucial role in host specificity. J. Virol. 67 :2815-2823. Morimoto, K., D.C. Hooper, H. Carbaugh, Z.F. Fu, H. Koprowski, and B. Dietzschold. 1998. Rabies virus quasispecies: implications for pathogenesis. Proc. Natl. Acad. Sci. USA. 95:3152-3156. Mueller, E., J., Gilbert, G. Davenport, G., Brigneti, and D. Baulcombe. 1995. Plant J. 721001-1013. Mueller, S., and E. Wimmer. 1998. Expression of foreign proteins by poliovirus polyprotein fiJsion: analysis of genetic stability reveals rapid deletions and formation of cardioviruslike open reading frames. J. Virol. 72:20-31. Nejidat, A. and RN. Beachy. 1990. Transgenic tobacco plants expressing a coat protein gene of tobacco mosaic virus are resistant to some other tobamoviruses. Mol. Plant- Microbe Interact. 3:247-251. Pacha, RF. and P. Ahlquist. 1991. Use of bromovirus RNA3 hybrid to study template specificity in viral RNA amplification. J. Virol. 65:3693-3703. Petersen, C. 1993. Translation and mRNA stability in bacteria: a complex relationship. In Control of messenger RNA stability. Belasco, J. and G. Brawennan (eds). Academic Press, Inc. Pogue, G.P., L.E. Marsh, J.F. Connell, and T.C. Hall. 1992. Requirement for ICR-like sequences in the replication of brome mosaic virus genomic RNA. Virology 1882742- 753. Powell-Abel, P.A., RS. Nelson, B. De, N. Hoffmann, S.G. Rogers, R.T. Fraley and RN. Beachy. 1986. Delay of disease development in transgenic plants that express the tobacco mosaic virus coat protein gene. Science. 232: 738-43 Puissant, C. and L. M. Houdebine. 1990. An improvement of the single-step method of RNA isolation by acid guanidinium thiocyanate—phenol-chloroform extraction. Biotechniques 8: 148-149. Rao, A. L. N., B. P. Sullivan and T. C. Hall. 1990a. Use of Chenopodium hybridum facilitates isolation of brome mosaic virus RNA recombinants. J. Gen. Virol. 71:1403- 1407. 127 Rao, A.L.N., C.C. Huntley, L.E. Marsh, and T.C. Hall. 1990b. Analysis of RNA stability and (-) strand content in viral infections using biotinylated probes. J. Virol. Methods 30:23 9-250. Rao, A.L.N., and G.L. Grantham. 1995. A spontaneous mutation in the movement protein gene of brome mosaic virus modulates symptom phenotype in Nicotiana benthamiana. J. Virol. 69:2689-2691. Rao, A.L.N. 1997. Molecular studies on bromovirus capsid protein. 111 Analysis of cell- to-cell movement competence of coat protein defective variants of cowpea chlorotic mottle virus. Virology 232:385-395. Register, J.C. IH, and RN. Beachy. 1988. Resistance to TMV in transgenic plants results from interference with an early event in infection. Virology 166:524-532. Restrepo-Hartwig, M. A. and P. Ahlquist. 1996. Brome mosaic virus helicase- and polymerase-like proteins colocalize on the endoplasmic reticulum at sites of viral RNA synthesis. J. Virol. 70:8908-8916. Rubino, L., and M. Russo, 1995. Characterization of resistance to cymbidium ringspot virus in transgenic plants expressing a full-length viral replicase gene. Virology 212:240- 243. Sanford, J.C., and S.A. Johnston, 1985. The concept of parasite-derived resistance - deriving resistance genes from the parasite’s own genome. J. Theor. Biol. 113:395-405. Schmitz, 1., and A.L.N. Rao. 1996. Molecular studies on bromovirus capsid protein. 1. Characterization of cell-to-cell movement-defective RNA3 vriants of brome mosaic virus. Virology 226:281-293. Schoelz, J.E. and W.M. Wintermantel. 1993. Expansion of viral host range through complementation and recombination in transgenic plants. Plant Cell 5: 1669-1679. Sethna, P.B., S.L. Hung and D.A. Brian. 1989. Coronavirus subgenomic minus-strand RNAs and the potential for mRN A replicons. Proc. Natl. Acad. Sci. USA 86:5626-5630. Smirnyagina, E., N. Lin, and P. Ahlquist. 1996. The polymerase-like core of brome mosaic virus 2a protein, lacking a region interacting with viral 1a protein in vitro, maintains activity and la selectivity in RNA replication. J. Virol. 70:4729-4736. Smith, H.A., H. Powers, S. Swaney, C. Brown, and W.G. Dougherty. 1995. Transgenic potato virus Y resistance in potato: evidence for an RNA-mediated cellular response. Phytopathology 85:864-870. 128 Sriskanda,V.S., G. Pruss, X.Ge, V.B. Vance. 1996. An eight-nucleotide sequence in the potato virus X 3' untranslated region is required for both host protein binding and viral multiplication. J. Virol. 70:5266-5277. Stark, D.M. and RN. Beachy. 1989. Protection against potyvirus infection in transgenic plants: Evidence for broad spectrum resistance. Bio/1‘ echnology 7: 1257-1262. Tavazza, R. R.J. Ordas, M. Tavazza, G. Ancora, and E. Benvenuto. Genetic transformation of Nicotiana clevelandii using a Ti plasmid derived vector. J. Plant Physiol. 133: 640-644. Tenllado, F., I. Garcia-Luque, M.T. Serra, J.R. Diaz-Ruiz. 1995. Nicotiana benthamiana plants transformed with the 54-KDa region of the pepper mild mottle tobamovirus replicase gene exhibit two types of resistance response against viral infection. Virology 21 1:170-183. Tenllado,F., L. Garcia-Luque, M.T. Serra, J.R. Diaz-Ruiz. Resistance to pepper mild mottle tobamovirus conferred by the 54-kDa gene sequence in transgenic plants dose not require expression of the wild-type 54-kDa protein. Virology 219:330-335. Traynor, P., B.M. Young, and P. Ahlquist. 1991. Deletion analysis of brome mosaic virus 2a protein: effects on RNA replication and systemic spread. J. Virol. 65: 2807- 2815. Tricoli, D.M., K.J. Carney, P.F. Russell McMaster, D.W. Groffl, K.C. Hadden, P.T. Himmel, J.P. Hubbard, M.L. Boeshore, and H.D. Quemada. 1995. Field Evaluation of Transgenic Squash containing single or multiple virus coat protein gene constructs for resistance to cucumber mosaic virus, watermelon mosaic virus 2, and zucchini yellow mosaic virus. Bio/Technology 13:1458-1465. Van der Kuyl, A.C., L. Neeleman and J.F. B01. 1991. Complementation and recombination between alfalfa mosaic virus RNA 3 mutants in tobacco plants. Virology 183:731-738. Van Dun, C.M.P. and J.F. B01. 1988. Transgenic tobacco plant accumulating tobacco rattle virus coat protein resist infection with tobacco rattle virus and pea early browning virus. Virology 167:649-652. Wassenegger, M., and T. Pelissier. 1998. A model for RNA-mediated gene silencing in higher plants. Plant Mol. Biol. 37:349-362. White, K.A., J.B. Bancroft, and G.A. Mackie. 1992. Mutagenesis of a hexanucleotide sequence conserved in potexvirus RNAs. Virology 1892817-820. 129 Yeh, S.D., H.J. Bau, Y.H. Cheng, T.A. Yu, J.S. Yang, 1998. Greenhouse and field evaluations of coat-protein transgenic papaya resistant to papaya ringspot virus. Acta Hortic. 4612321-328. Zaccomer, B., F. Cellier, J-C Boyer, A-L. Haenni, and M. Tepfer. 1993. Transgenic plants that express genes including the 3’ untranslated region of the turnip yellow mosaic virus (TYMV) genome are partially protected against TYMV infection. Gene 136:87-94. Zaitlin, M., J.M. Anderson, K.L. Perry, L. Zhang, and P. Palukaitis, 1994. Specificity of replicase-mediated resistance to cucumber mosaic virus. Virology 201:200- 205. 130 llllllllllllllll