key: cord-0944556-eo3bei9x authors: Fondong, Vincent N. title: Geminivirus protein structure and function date: 2013-04-25 journal: Molecular Plant Pathology DOI: 10.1111/mpp.12032 sha: 80dbdfbf10174a13cded449973fab8e6c12281c2 doc_id: 944556 cord_uid: eo3bei9x Geminiviruses are a family of plant viruses that cause economically important plant diseases worldwide. These viruses have circular single‐stranded DNA genomes and four to eight genes that are expressed from both strands of the double‐stranded DNA replicative intermediate. The transcription of these genes occurs under the control of two bidirectional promoters and one monodirectional promoter. The viral proteins function to facilitate virus replication, virus movement, the assembly of virus‐specific nucleoprotein particles, vector transmission and to counteract plant host defence responses. Recent research findings have provided new insights into the structure and function of these proteins and have identified numerous host interacting partners. Most of the viral proteins have been shown to be multifunctional, participating in multiple events during the infection cycle and have, indeed, evolved coordinated interactions with host proteins to ensure a successful infection. Here, an up‐to‐date review of viral protein structure and function is presented, and some areas requiring further research are identified. Geminiviruses (family Geminiviridae) are among the most devastating plant pathogens worldwide and, together with potyviruses (family Potyviridae), constitute the two largest and most important plant virus families (Gibbs and Ohshima, 2010; Scholthof et al., 2011) . By causing heavy losses on food and cash crops, such as cassava, tomatoes, grain legumes, vegetables, maize and cotton, geminiviruses represent a new threat to global food security and sustainability. During the last two decades, epidemics of re-emerging and newly emerging geminiviruses have caused huge crop losses and threatened crop production, particularly in the tropics and subtropics. Because of their agricultural importance, geminiviruses have been characterized extensively at the molecular level. Geminivirus genomes consist of one (monopartite) or two (bipartite) circular single-stranded DNA (ssDNA) molecules, which are packaged in icosahedral twinned particles (Böttcher et al., 2004; Krupovic et al., 2009; Zhang W. et al., 2001) . These viruses are divided into four genera-Begomovirus, Curtovirus, Topocuvirus and Mastrevirus-based on genome organization, nucleotide sequence similarities and biological properties (Brown et al., 2012) . Begomoviruses (type species Bean golden mosaic virus, BGMV) have either monopartite or bipartite genomes and are transmitted by whiteflies (Bemisia tabaci) (Brown, 1994) . The two genomic components of bipartite begomoviruses are designated DNA-A and DNA-B (Fig. 1a) . DNA-A has six open reading frames (ORFs), two in the virion sense (AV1 and AV2) and four in the complementary sense (AC1, AC2 and AC3 and AC4). DNA-B has two ORFs, the virion-sense BV1 and complementary-sense BC1. The AV2 ORF is found in Old World bipartite begomoviruses, but not in New World viruses. Except for an approximately 200nucleotide segment of the 5′ intergenic region (IR), designated the common region (CR), the two genomic components of bipartite begomoviruses are different in sequence. The CR contains an origin of replication organized modularly, including a stem-loop structure containing the invariant nonanucleotide TAATATTAC sequence, which is required for the cleavage and joining of the viral DNA during replication (Laufs et al., 1995) . The ORFs of monopartite begomoviruses include the virion-sense V1 and V2 genes and the complementary-sense C1, C2, C3 and C4 genes (Fig. 1b) . Mastreviruses (type species Maize streak virus, MSV) are transmitted by leafhoppers (order Hemiptera, family Cicadellidae) and have a single genome component. Except for a recent identification of mastreviruses in sweet potato in Peru (Kreuze et al., 2009) and in dragonflies (suborder Anisoptera) in Puerto Rico (Rosario et al., 2013) , these viruses have been found only in the Old World, where they infect a range of monocotyledonous plants (reviewed in Brown et al., 2012) and, from recent reports, dicotyledonous plants (Hadfield et al., 2012; Liu et al., 1999; Nahid et al., 2008) . Mastreviruses have four ORFs, V1 and V2 on the virion-sense strand, and C1 and C2 on the complementary-sense strand, separated by two IRs on opposite sides of the genome, designated as large (LIR) and small (SIR) (Kammann et al., 1991) (Fig. 1c) . The SIR contains termination and polyadenylation sequences for both the complementary-sense and virion-sense transcripts, and is probably the origin of replication of the complementary-sense strand (Hayes et al., 1988) . For mastreviruses, but not for viruses of other genera, an approximately 80-nucleotide primer is present in the virion and specifically anneals to the SIR (Kammann et al., 1991; Shen and Hohn, 1991) . Curtoviruses (type species Beet curly top virus, BCTV) infect dicotyledonous hosts, are transmitted by the beet leafhopper (Circulifer tenellus) and have monopartite genomes (Fig. 1d) . The genome has seven ORFs and a single IR (Hormuzdi and Bisaro, 1995; . The three virion-sense ORFs are V1, V2 and V3. The four complementary-sense ORFs are C1, C2, C3 and C4. Like mastreviruses, curtovirus genomes contain a polyadenylation signal of approximately 20 nucleotides between the converging V1 and C3 ORFs. The genus Topocuvirus, which so far has only one member, Tomato pseudo-curly top virus (TPCTV), has a monopartite genome and is transmitted to dicotyledonous hosts by treehoppers (Micrutalis malleifera) (Briddon et al., 1996) . Topocuviruses have six ORFs (Fig. 1e) : the virion-sense strand contains V1 and V2 and the complementary-sense strand contains C1, C2, C3 and C4, which are homologues of AC1, AC2, AC3 and AC4, respectively. Previous comprehensive reviews (Gutierrez et al., 2004; Jeske, 2009; Stanley, 2004) have focused on the biological and molecular properties of geminiviruses. In this article, a review of current information on the molecular biology of geminiviruses and their satellites is provided, with special emphasis on viral protein structure, function and interactions with host proteins. Important areas needing additional research are also identified. The replication-associated protein (Rep) encoded by the AC1 ORF (also called AL1) in bipartite geminiviruses and by C1 (also called L1) in monopartite geminiviruses (except mastreviruses) is conserved in sequence, position and function (Hanley-Bowdoin et al., 2004) and is expressed under the control of a bidirectional core promoter in the IR (Hanley-Bowdoin et al., 1999) . In the mastreviruses, the full-length Rep is translated from a spliced transcript of the C1 and C2 ORFs. The multifunctional nature of begomovirus Reps and the functional domains have been characterized (Heyraud-Nitschke et al., 1995; Orozco et al., 1997) (Fig. 2) . Rep is essential for rolling circle replication (RCR) and is involved in the modulation of gene expression (Elmer et al., 1988; Etessami et al., 1991; Haley et al., 1992; Hong and Stanley, 1995; Saunders et al., 1991) , analogous to some animal and bacterial ssDNA viruses (Stenger et al., 1991) and plasmids (Oshima et al., 2001) , suggesting a strong evolutionary link between these proteins (Ilyina and Koonin, 1992) . The first step in geminivirus replication is the synthesis of the complementary strand from the genomic ssDNA using a stillunknown mechanism that is thought to be catalysed entirely by host factors. Rep, an early gene, confers virus-specific recognition of its cognate origin of replication (Fontes et al., 1994) and initiates DNA replication (Laufs et al., 1995; Lazarowitz et al., 1992; Orozco et al., 1997) . The recognition mechanism appears to differ between mastreviruses and begomoviruses, based on the architecture of the origin of replication (Gutierrez et al., 2004) . For mastreviruses, the origin consists of a large cis-acting region containing several Rep binding sites (Castellano et al., 1999; Sanz-Burgos and Gutiérrez, 1998) , whereas the begomovirus origin contains a single Rep binding site (Fontes et al., 1992 (Fontes et al., , 1994 Lazarowitz et al., 1992; Orozco and Hanley-Bowdoin, 1998) . In either case, however, the Rep oligomerizes into a complex as illustrated for MSV (Horvath et al., 1998) , Wheat dwarf virus (WDV) (Castellano et al., 1999; Sanz-Burgos and Gutiérrez, 1998) and Tomato golden mosaic virus (TGMV) (Orozco and Hanley-Bowdoin, 1998) from mutational analyses of the oligomerization domain. In planta Rep oligomerization has been elucidated recently using Abutilon mosaic virus (AbMV) Rep bimolecular fluorescence complementation (BiFC) assay in Nicotiana benthamiana plants, where Rep oligomers accumulated in the nucleoplasm and replicated AbMV DNA-B (Krenz et al., 2011) . To initiate the RCR, the begomovirus Rep binds to the Rep complex binding site, which contains a directly repeated sequence between the TATA box and the transcription start site (Fontes et al., 1992) . For the mastreviruses, the complex binds presumably to three sites located in the proximity of the two divergent TATA boxes upstream and downstream of the DNA replication initiation site and at the base of the conserved stem-loop (Castellano et al., 1999; Laufs et al., 1995) . On binding, a process which involves motif I (Fig. 2a) , the Rep complex cleaves the phosphodiester bond between the last T and A in the invariant loop sequence in a biochemical process mediated by motif II, which is a metal binding site that may be involved in protein conformation (Arguello-Astorga et al., 2004; Laufs et al., 1995; Orozco and Hanley-Bowdoin, 1998) . A third sequence, motif III, is the catalytic site that is required for DNA cleavage and ligation (Orozco et al., 1997) , thus providing the necessary 3′-OH for priming nascent viral DNA synthesis by host DNA polymerases. Recently, Nash et al. (2011) have identified a conserved motif, named the geminivirus Rep sequence (GRS) (Fig. 2a) , and have shown that it is required for the initiation of virus replication. An artificial zincfinger protein designed to bind to the Tomato yellow leaf curl virus (TYLCV) Rep binding site with high affinity has been shown recently to inhibit Rep binding and TYLCV replication (Mori et al., 2012) . Several Rep-interacting proteins that are important in replication have been identified. Thus, a family of proteins involved in the negative regulation of the cell cycle related to retinoblastoma, designated retinoblastoma-related proteins (RBR), interacts with Reps of several geminiviruses (Ach et al., 1997; Kong et al., 2000; Xie et al., 1996) . RBR represses cell cycle progression through its interactions with E2F (Arguello- Astorga et al., 2004) . Accordingly, interaction between Rep and RBR interferes with the interaction between RBR and E2F and frees the activator class of E2F to activate the expression of genes required for DNA replication, including proliferating cell nuclear antigen (PCNA) (Arguello-Astorga et al., 2004; Egelkrout et al., 2001; Kong et al., 2000) , which is permissive for viral DNA replication. Also critical to viral DNA replication is the interaction between Rep and replication factor C (RFC), which catalyses the loading of PCNA, a sliding clamp for DNA polymerase at the primer-template junction (Pavlov et al., 2004) . Interaction between Rep and RFC is said to recruit RFC into the replication complex, where it facilitates the assembly of other replication factors (Luque et al., 2002) . Rep also binds PCNA, which starts to accumulate in the G1 phase of the cell cycle, reaching the highest level during the S phase and decreasing during the G2 and M phases (Gutierrez, 2000) . Furthermore, TGMV Rep has been shown recently to interact with small ubiquitin-related modifier (SUMO)-conjugating enzyme (SCE1) to presumably modulate the SUMOylation level of host targets, thereby creating a suitable environment for virus replication (Castillo et al., 2004; Sánchez-Durán et al., 2011) . Although SUMOylation plays an important role in human DNA virus infection by altering the molecular interactions of the modified substrate, changing the localization, stability and/or activity (reviewed in Wimmer et al., 2012) , definitive conclusions on the role of SUMO modification during plant virus infection are yet to be determined. Rep also interacts with the replication protein A (RPA) (Singh et al., 2007 (Singh et al., , 2008 , a key protein in the recruitment of important proteins of the replication apparatus, such as DNA polymerase a, RFC and PCNA (Loor et al., 1997) . It is important to indicate that RPA is critical to other viral replication initiator proteins, including SV40 large T antigen (Weisshart et al., 1998) , EBNA1 proteins of Epstein-Barr virus (Zhang et al., 1998) , E1 proteins of papilloma virus (Loo and Melendy, 2004) and NS1 proteins of parvovirus (Christensen and Tattersall, 2002) . Furthermore, Rep interacts with a kinesin-like motor protein (Kong and Hanley-Bowdoin, 2002) to presumably prevent cell cycle progression through M and favour the occurrence of endocycles (Desvoyes et al., 2006; Jordan et al., 2007) , thus favouring viral DNA replication. We have also found that Rep interacts with b-tubulin6 (TUB6) (V. N. Fondong et al., unpublished data) , suggesting that Rep movement may be microtubule dependent. In addition to replication, some geminivirus Reps have been shown to be involved in viral gene regulation by repressing complementary-sense gene expression of mastreviruses (Collin et al., 1996; Hefferon et al., 2006) and begomoviruses (Eagle et al., 1994; Gröning et al., 1994; Haley et al., 1992; Shivaprasad et al., 2005) , presumably after binding to Rep binding sites in the IR (Fontes et al., 1992; Frey et al., 2001) . However, Rep appears not to repress virion-sense gene expression, as elucidated for Mungbean yellow mosaic virus (MYMV) Rep . Together, these findings have contributed significantly to our understanding of the role of Rep in geminivirus replication and the mechanism of RCR; they have revealed the functional relationship between the geminivirus Rep and replication initiation proteins of other ssDNA viruses, as well as plasmids, and have supported the view that geminiviruses evolved from plasmids, even though, in a recent study based on sequence analysis, Saccardo et al. (2011) rejected the hypothesis that geminiviruses originated from plasmids of phytoplasma. However, still to be fully determined are the mechanism of cellular factor recruitment and the biochemical events leading to Rep binding to viral DNA to initiate genome replication. In addition, it is not yet known whether a ssDNA binding protein coats and stabilizes the geminivirus genomic ssDNA prior to replicative form synthesis, as has long been known for ssDNA phages that replicate via the RCR mechanism (Arai and Kornberg, 1979; Geider et al., 1978) . The replication-associated protein A (RepA), the protein product of ORF C1, and Rep are the two complementary-sense proteins of the mastreviruses (Collin et al., 1996; Gutierrez et al., 2004; Hofer et al., 1992; Liu et al., 1997) , and are both required for virus replication. The RepA mRNA accounts for most of the complementary-sense transcripts in monocot-infecting mastreviruses, including MSV, WDV and Digitaria streak virus (DSV) (Dekker et al., 1991; Mullineaux et al., 1990; Wright et al., 1997) , as well as in the dicot-infecting mastreviruses (Gutierrez, 2000; Hefferon et al., 2006; Munoz-Martin et al., 2003) . RepA and Rep have identical primary sequences in their approximately 200 N-terminal amino acid residues (Fig. 2b) . Like Rep, RepA is multifunctional and, in several cases, performs the functions of bipartite geminivirus Reps, including the transactivation of virus-sense ORFs, as has been elucidated for MSV, WDV and Bean yellow dwarf virus (BeYDV) (Hefferon et al., 2006; Munoz-Martin et al., 2003) . RepA also represses its own expression and the expression of Rep, as elucidated in BeYDV (Hefferon et al., 2006) , by binding to a site in the LIR. RepA interaction with RBR has been mapped to the LXCXE motif in the mastreviruses MSV, BeYDV and WDV (Horvath et al., 1998; Liu et al., 1999; Xie et al., 1995) . However, the LXCXE motif is absent in some mastrevirus RepAs, including Sugarcane streak virus (SSV), Miscanthus streak virus (MiSV) (Hughes et al., 1993) and BeYDV (Hefferon et al., 2006) , suggesting that it is not an absolute requirement for mastreviruses. Furthermore, LXCXE is absent in begomovirus Reps, which bind to RBR through a novel, conserved a-helical region (Arguello-Astorga et al., 2004; Hanley-Bowdoin et al., 2004; Kong et al., 2000) . Other RepA interacting partners include a host transcription factor, designated GRAB (geminivirus RepA-binding), in the NAC family [acronym for NAM (for No Apical Meristem), ATAF1-2 and CUC2 (for Cup-Shaped Cotyledon)] (Gutierrez et al., 2004; Xie et al., 1999) . GRAB proteins, which are involved in plant development and senescence, impair geminivirus replication, suggesting that they modulate the viral replication cycle (Lozano-Duran et al., 2011; Xie et al., 1999) . Curiously, circoviruses, another group of circular ssDNA viruses, which infect animals and replicate using the RCR mechanism, also have two replication initiation proteins, designated Rep and Rep'. Rep and Rep' (a spliced isoform of Rep) are indispensable for circovirus replication (Mankertz and Hillenbrand, 2001) and have structural and functional similarities to Rep/RepA, suggesting a close evolutionary relationship between the two virus groups. However, whereas, in circoviruses, replication is terminated by a Rep/Rep'-catalysed nucleotidyl transfer reaction to cleave monomer units (Steinfeldt et al., 2006) , geminivirus replication produces monomers and multimers, suggesting an inefficient Repmediated monomer cleavage during RCR. The transcriptional activator protein (TrAP) is encoded by the AC2 (or AL2) ORF in bipartite begomoviruses and by C2 in monopartite begomoviruses. TrAP is a positional analogue of C2 (or L2) protein in curtoviruses and topocuviruses. AC2, together with AC3, is expressed from a dicistronic transcript driven by a strong monodirectional promoter located within the AC1 coding sequence . TrAP, which has been well characterized in begomoviruses, is a multifunctional protein. It is involved in gene activation (Haley et al., 1992; Shivaprasad et al., 2005; Sunter and Bisaro, 1992) , virus pathogenicity (Hong et al., 1996) and suppression of gene silencing (Chowda-Reddy et al., 2009; Trinks et al., 2005; Vanitharani et al., 2005; Voinnet et al., 1999; Wang et al., 2005) . TrAP activates the transcription of the genes coding for coat protein (CP) and movement protein (MP) in a non-virus-specific manner (Hanley-Bowdoin et al., 1999; Saunders and Stanley, 1995; Sunter et al., 1994) , and transactivation is dependent on its zinc-finger and C-terminal acidic domains ( Fig. 3 ) (Hartitz et al., 1999) , as well as on its ability to form multimers (Yang et al., 2007) . TrAP regulates the tissue-specific expression of the CP promoter by overriding a putative host repressor in phloem tissues (Sunter and Bisaro, 1997) . CP promoter modulation is thought to occur through TrAP interaction with different components of the cellular transcriptional machinery which binds the viral sequences required for the regulation of CP promoters (Lacatus and Sunter, 2009 ). The Arabidopsis transcription factor, PEAPOD2, specifically binds to activating sequences in the CP promoter of TGMV and Cabbage leaf curl virus (CaLCV) in mesophyll tissues, but not to sequences required for TrAP-mediated derepression in phloem tissues (Lacatus and Sunter, 2009 ). TrAP does not specifically bind to double-stranded DNA (dsDNA) and, instead, is thought to be directed to responsive promoters through interactions with cellular proteins (Hartitz et al., 1999; Lacatus and Sunter, 2009 ). Thus, TrAP is similar to the Adenovirus E1A protein, a transactivational regulatory factor, which does not bind DNA, but is necessary for transcriptional activation (Avvakumov et al., 2002; Zu et al., 1992) . Conclusive evidence of transcriptional activation activity by a monopartite begomovirus C2 protein was reported for Tomato leaf curl virus (ToLCV) (Dry et al., 2000) , and it is likely that other monopartite begomovirus C2 proteins are also transcriptional activators. In contrast, the curtovirus C2 does not activate transcription (Hormuzdi and Bisaro, 1995) . A possible mechanism of TrAP nuclear import for transcriptional activation is provided by the reported interaction between C2 of the begomovirus Bhendi yellow vein mosaic virus and karyopherin a (Chandran et al., 2012), a soluble receptor, which probably transports C2 through nuclear pore complexes into the nucleoplasm. TrAP is also a pathogenicity factor, and may counter a hypersensitive response (HR) in infected cells. The HR, a form of programmed cell death (PCD) associated with resistance to pathogens, is induced at infected sites or within defined areas surrounding the infected sites and limits pathogen growth (a review is provided in Postel and Kemmerling, 2009 and in Coll et al., 2011) . However, the TrAP of the bipartite begomovirus Tomato leaf curl New Delhi virus (ToLCNDV) can overcome nuclear shuttle protein (NSP)-induced HR (discussed below) (Hussain et al., 2007) . TrAP compromises the ability of COP9 signalosome (CSN), a protein complex that functions in the ubiquitinproteasome pathway, to bind to Cullin-1 (CUL1), which is an essential component of the SCF (SKP1, CUL1/CDC53, F box proteins) ubiquitin E3 ligase complex. TrAP-COP9 interaction alters the cellular processes regulated by SCF complexes, including jasmonate signalling (Lozano-Duran et al., 2011) , thereby regulating host response to infection. Most plant viruses contain suppressors of RNA silencing, a phenomenon that regulates gene expression and protects plants from transposable elements and viruses (reviews are available in Pantaleo, 2011 and in Wang et al., 2012) . The bipartite geminivirus TrAP and C2 from monopartite geminiviruses have been shown to suppress RNA silencing (Chowda-Reddy et al., 2009; Dong et al., 2003; Hamilton et al., 2002; Trinks et al., 2005; Vanitharani et al., 2004; Voinnet et al., 1999; Wang et al., 2005) . Correspondingly, Beet severe curly top virus (BSCTV) C2 has been shown to decrease DNA methylation, thus affecting the production of small interfering RNAs (siRNAs) which are required for the targeting and reinforcing of RNA-directed DNA methylation and RNA silencing (Yang L.P. et al., 2012) . Consistent with these findings, the transient expression of TGMV TrAP and BCTV C2 increases the susceptibility of tobacco to these geminiviruses (Hao et al., 2003; Sunter et al., 2001) . Taken together, these findings establish the involvement of TrAP and C2 in gene regulation and virus pathogenicity, and emphasize the multifunctional nature of geminivirus proteins, which, together with overlapping genes, contribute to viral genetic economy. The replication enhancer protein (REn), which is absent in mastreviruses, is encoded by the AC3 (or AL3) ORF in bipartite begomoviruses and by the C3 (or L3) ORF in curtoviruses and monopartite begomoviruses. Like AC2, AC3 expression is driven by a strong monodirectional promoter located within the AC1 coding sequence . Although not essential for virus replication, REn enhances viral DNA accumulation and symptom development in plants infected by begomoviruses (Sung and Coutts, 1995; Sunter et al., 1990) and curtoviruses . The role of REn in viral DNA replication involves its interactions with Rep and PCNA. Regions of REn responsible for these interactions were elucidated in two bipartite begomoviruses, TYLCV and Tomato yellow leaf curl Sardinia virus (TYLCSV) (Castillo et al., 2003; Settlage et al., 2005) (Fig. 4) . Like Rep, REn does not contain an LXCXE motif, which mediates the binding of several mastrevirus RepAs to RBR, suggesting the existence of an alternative RBR interaction domain in REn (Settlage et al., 2001) . Like RepA, REn also interacts with SINAC1 (Solanum lycopersicum NAC1) (Selth et al., 2005) . SINAC1 levels were shown to be higher in ToLCV-infected cells, suggesting that NAC1 is involved in viral DNA replication (Selth et al., 2005) , possibly through an interaction with REn. Recently, the Tomato leaf curl Kerala virus REn was shown to interact with Rep and enhance the Repmediated ATPase activity (Pasumarthy et al., 2010) , thus confirming a role for REn in viral DNA replication. The ORFs AC4 (or AL4) in bipartite geminiviruses and C4 (or L4) in monopartite geminiviruses (except mastreviruses, where C4 is absent) are contained entirely within the AC1 ORF, but in a different frame, and are the least conserved of all geminivirus proteins, with divergent biological functions in monopartite and bipartite geminiviruses. Disruption of the C4 ORF of ToLCV (Rigden et al., 1994) , TYLCV (Jupin et al., 1994) and BSCTV (Teng et al., 2010) caused a reduction in viral DNA levels and symptom development, suggesting its involvement in symptom development and virus movement. C4 is also the major determinant of the characteristic vein swelling phenotype observed in plants infected by BCTV (Mills-Lujan and Deom, 2010; . Expression of the BCTV and BSCTV C4 proteins in transgenic Arabidopsis results in phenotypes that mimic the symptoms seen during viral infection (Mills-Lujan and Deom, 2010) . Similarly, the expression of BCTV C4 in transgenic N. benthamiana results in ectopic cell division (Latham et al., 1997; Piroux et al., 2007) , implying that the vein swelling associated with BCTV C4 is caused by abnormal cell division. Consistent with this finding, Lai et al. (2009) reported that BSCTV C4 induced the expression of RKP, which may be involved in cell cycle regulation. According to a study by Park et al. (2011) , infection by BSCTV, as well as overexpression of BSCTV C4, leads to the induction of expression of ATHB7 and ATHB12 in Arabidopsis thaliana. These observations implicate C4 in infected cell development and differentiation. TGMV AC4 has been reported to be involved in virus movement , but is not critical for virus infection (Elmer et al., 1988) , and disruption of the AC4 ORF in African cassava mosaic virus (ACMV) and East African cassava mosaic Zanzibar virus (EACMZV) has no effect on the infection phenotype in N. benthamiana (Bull et al., 2007; Etessami et al., 1991) . Similar observations were reported for AC4 encoded by BGMV (Hoogstraten et al., 1996) and Potato yellow mosaic virus (Sung and Coutts, 1995) . The ability of AC4 and C4 to suppress RNA silencing is conserved for several bipartite and monopartite geminiviruses. For example, AC4 proteins of ACMV and East African cassava mosaic virus (EACMV)-like viruses have been shown to suppress RNA silencing (Fondong et al., 2007; Vanitharani et al., 2004) by binding microRNAs (miRNAs) and siRNAs (Chellappan et al., 2004) . The ability of East African cassava mosaic Zanzibar virus (EACMZV) AC4 to suppress silencing depends on its N-terminal myristoylation sequence, in which acetylation occurs as a result of the attachment of a myristate and a palmitate in the N-terminus (Fondong et al., 2007) . Furthermore, ToLCV-Au C4 suppresses silencing by binding to shaggy-like kinase (SlSK) through its C-terminus (Dogra et al., 2009) . The geminivirus CP is a late gene and is the only structural protein of geminivirus particles (Stanley and Gay, 1983) . It is encoded by the AV1 (also AR1) ORF in bipartite geminiviruses and by the V1 ORF in monopartite geminiviruses (Zhang W. et al., 2001) . In addition to virus genome packaging, the CP has been associated with several other functions, including insect transmission (Boulton et al., 1993; Briddon et al., 1990; Mullineaux et al., 1984) , shuttling of viral DNA into and out of the nucleus in monopartite geminiviruses (Liu et al., 1999) , and cell-to-cell and systemic spread of virus (Boulton et al., 1989; Liu et al., 1997; Pitaksutheepong, 1999) . The CP binds ss-and dsDNA (Liu et al., 1997) and may participate indirectly in viral DNA replication during RCR by binding and sequestering ssDNA from the replication cycle (Saunders et al., 1991; Stenger et al., 1991) . Although all begomovirus CPs contain sequences that are highly conserved, they also contain regions that are variable and can be used to correlate phylogenetic inferences with biotic and geographic characteristics (Padidam et al., 1995) . Evidence that the CP plays a critical role in insect virus transmission was elucidated using the whitefly-transmitted ACMV and the leafhopper-transmitted BCTV. An ACMV chimeric virus containing the BCTV CP gene is transmitted by leafhoppers, which, in nature, transmit BCTV but not ACMV (Briddon et al., 1990) . Sequences important for insect transmission have been mapped to the central part of CP (Hohnle et al., 2001; Kheyr-Pour et al., 2000; Liu et al., 2001; Noris et al., 1998; Unseld et al., 2001) and may play a role in CP multimerization, which is necessary for virus capsid assembly and insect transmission (Hallan and Gafni, 2001; Noris et al., 1998; Zhang W. et al., 2001) . Some monopartite geminivirus CPs shuttle viral DNA between the nucleus and the cytoplasm, as has been elucidated for mastreviruses (Liu et al., 1999 (Liu et al., , 2001 and for TYLCV (Kunik et al., 1998; Rojas et al., 2001) . This is consistent with the fact that the CP of Squash leaf curl virus (SqLCV), a bipartite begomovirus, can functionally replace NSP (encoded by BV1) (Ingham et al., 1995; Qin et al., 1998) . Nuclear import and export appear to be dependent on ssDNA binding to the N-terminal domain of the CP (Pitaksutheepong et al., 2007) . The CP nuclear localization signal (NLS) has been localized to the N-and C-termini and the central region of the protein (Unseld et al., 2001) (Fig. 5) . The bipartite begomovirus MYMV CP interacts with importin a, a component of the nuclear pore-targeting complex (Guerra-Peraza et al., 2005) , and the N-terminal NLS of TYLCV CP binds to karyopherin a1 (Kunik et al., 1999) for nucleocytoplasmic trafficking. The N-terminal NLS also binds to the GroEL protein of Arsenophonus, a bacterial endosymbiont found in the midgut of Bemisia tabaci, presumably to protect the virions in the insect vector haemolymph (Kunik et al., 1999; Morin et al., 2000; Rana et al., 2012; Yaakov et al., 2011) . A leucine-rich nuclear export signal (NES) that mediates trafficking from the nucleus to the cytoplasm was mapped to the CP central region (Ward and Lazarowitz, 1999) . It is important to note that bipartite begomoviruses with mutated or replaced CP are infectious, albeit with delayed and attenuated symptoms (Brough et al., 1988; Sudarshana et al., 1998) , suggesting that other host proteins functionally replace the CP and import the genome to the nucleus on initial viral infection and prior to the synthesis of NSP. Interestingly, geminivirus CPs accumulate preferentially in the nucleolus, as exemplified by CPs of the monopartite begomoviruses Tomato leaf curl Java virus (ToLCJV) and TYLCV (Rojas et al., 2001) , and the bipartite begomovirus EACMCV (V. N. Fondong et al., unpublished data) . It is worth noting that increasing numbers of key proteins from both animal RNA and DNA viruses have been shown to localize to the nucleolus and to play important roles in the virus infection cycle, including herpesvirus saimiri ORF57 protein (Boyne and Whitehouse, 2006), coronavirus protein N (Hiscox et al., 2001) , Groundnut rosette virus ORF3 (Kim et al., 2007a) , Potato leaf roll virus CP (Kim et al., 2007a) and Potato mop top virus triple gene block 1 protein (TGB1) (Kim et al., 2007b) . These observations implicate the nucleolus in the virus infection cycle. Indeed, it has been shown recently that nucleolar localization is required for Alfalfa mosaic virus virion formation and systemic movement (Herranz et al., 2012) . In addition to nuclear shuttling, CPs of mastreviruses (Liu et al., 2001) and monopartite begomoviruses (Briddon et al., 1989; Padidam et al., 1996; Rigden et al., 1993) are involved in cell-to-cell movement and systemic spread of viral DNA. Movement is effected presumably through interaction with the MP (V2 ORF), as demonstrated for TYLCV (Poornima et al., 2011) and Cotton leaf curl Kokhran virus (Poornima et al., 2011) . Although bipartite begomovirus CPs may not be required for infection, they have been shown to enhance cell-to-cell and/or systemic spread of several viruses, including SqLCV (Ingham et al., 1995) , ACMV (Stanley and Townsend, 1986) , Bean dwarf mosaic virus (BDMV) (Seo et al., 2004) , Pepper huasteco virus (Guevara-González et al., 1999) , BGMV ), CaLCV (Carvalho et al., 2008b and TGMV (Gillette et al., 1998; . These observations suggest that the bipartite begomovirus CPs have not lost these functions during evolution from their monopartite progenitors. Although not essential in virus replication, the absence of CPs results in reduced levels of viral ssDNA with or without a reduction in the level of dsDNA of TYLCV (Padidam et al., 1996) , TGMV (Brough et al., 1988; Sunter et al., 1990) , BCTV (Briddon et al., (Padidam et al., 1999) , SqLCV (Qin et al., 1998) and BeYDV (Hefferon and Dugdale, 2003) . Together, these findings provide further evidence that the evolution of multifunctional proteins compensates for small-genome viruses and that multifunctional proteins highlight the plasticity through which evolution brings together functional domains into a single polypeptide chain (Walsh and Mohr, 2006) . The AV2 ORF (V2 in monopartite geminiviruses), whose start codon precedes that of the AV1 ORF, is found in 'Old World', but not in 'New World', bipartite begomoviruses (Bull et al., 2007; Padidam et al., 1996; Rojas et al., 2001; Selth et al., 2004) , and its exact function is uncertain. Although ACMV clones containing mutations in AV2 are infectious in N. benthamiana (Etessami et al., 1989) , recent findings showing that Mungbean yellow mosaic India virus (MYMIV) clones containing mutations in AV2 reverted to wild-type (Rouhibakhsh et al., 2011) raise the possibility that the progeny from mutant ACMV clones might have reverted to wild-type, thereby masking the effect of the mutation. This is supported by evidence that ToLCNDV containing mutations in AV2 accumulates low levels of viral DNA in infected plant tissues (Padidam et al., 1996) . Correspondingly, V2 ORFs of the monopartite begomovirus TYLCV (Wartig et al., 1997) and the curtovirus BCTV (Hormuzdi and Bisaro, 1995; appear to be pathogenicity determinants, and mastrevirus and monopartite begomovirus V2 ORFs encode MPs, as discussed in the section on Movement proteins below. Plants infected with ToLCV and ToLCNDV containing mutations in V2 and AV2, respectively, resulted in low levels of mutant viral DNA in infected plant tissues (Padidam et al., 1996; Rigden et al., 1993; Selth et al., 2004) . Similar results have been reported recently for Papaya leaf curl virus and Cotton leaf curl Kokhran virus (Mubin et al., 2010) . These findings were subsequently supported by studies showing that both AV2 and V2 suppress RNA silencing, as illustrated for EACMCV AV2 (Chowda-Reddy et al., 2008) and V2 of the monopartite begomovirus Ageratum yellow vein virus . Similar observations have been made for TYLCV V2, most probably as a result of its interaction with the suppressor of gene silencing3 protein (SGS3) (Glick et al., 2008) . BV1 (also BR1) is one of the two ORFs found in the B component of bipartite begomoviruses. BV1 encodes the NSP (Fig. 6 ) required for trafficking viral ssDNA between the nucleus and the cytoplasm (Noueiry et al., 1994) in the form of a viral DNA-NSP complex (Ward and Lazarowitz, 1999) . Evidence that NSP binds to DNA in a sequence non-specific manner has been illustrated for SqLCV (Pascal et al., 1994) , BDMV (Rojas et al., 1998) and AbMV (Hehnle et al., 2004) . To achieve cell-to-cell and long-distance movement, the NSP-viral DNA complex is trapped by the MP in the cytoplasm and redirected to and across the cell wall into adjacent cells, where NSP directs the viral genome to the nucleus for new cycles of replication (Lazarowitz and Beachy, 1999; Noueiry et al., 1994; Sanderfoot and Lazarowitz, 1995; Sanderfoot et al., 1996; Ward et al., 1997) . It is important to note that NSP is consistently being shown not to be required for virus infectivity, presumably because the CP exhibits nuclear import and export functions as discussed above. This has led to the suggestion that both proteins have a common evolutionary origin and share redundant functions (Zhou et al., 2007) , including the nucleocytoplasmic transport of viral DNA (Liu et al., 1997; Rigden et al., 1993; Saeed et al., 2007) . Accordingly, both proteins localize in the nucleus, preferentially to the nucleolus (Fig. 6) (Rojas et al., 2001; Sharma and Ikegami, 2009; Zhou et al., 2011) . The mechanism by which NSP shuttles viral DNA between the nucleus and the cytoplasm has been elucidated using CaLCV NSP, which interacts with an Arabidopsis protein, nuclear shuttle protein interactor (AtNSI) (McGarry et al., 2003) . This interaction presumably results in the acetylation of NSP and the segregation of ssDNA for movement and encapsidation. Several geminivirus NSPs are also known to interact with a cytoplasmic GTPase to export the viral DNA-NSP complex from the nucleus to the cytoplasm, where it is redirected to the cell surface to interact with the viral MP (Carvalho et al., 2008a) . Evidence from SqLCV NSP shows that the nuclear export of nascent viral DNA from the nucleus to the cytoplasm is mediated by a leucine-rich NES. In addition to virus trafficking, NSP is an avirulence determinant in some hosts. Thus, BDMV NSP has been reported to induce an HR in Phaseolus vulgaris (Garrido-Ramirez et al., 2000; Zhou et al., 2007) , as does ToLCNDV NSP in N. tabacum and tomato (Solanum lycopersicum) (Hussain et al., 2005 (Hussain et al., , 2007 . The involvement of NSP in virus pathogenicity is supported by its interactions with membrane receptor kinases, including NSP-interacting kinase (NIK), belonging to the receptor-like group of kinases implicated in a wide range of signal transduction pathways (Fontes et al., 2004; Rocha et al., 2008; Sakamoto et al., 2012; Santos et al., 2009) . NSP also interacts with proline-rich extensin-like receptor protein Fig. 6 The nuclear shuttle protein (NSP). The NSP N-terminal half contains two nuclear localization signals (NLS-A and NLS-B), DNA binding and sequences required for the induction of the hypersensitive response, whereas the NSP C-terminus contains the nuclear export signal (NES), movement protein (MP) binding site and the Arabidopsis nuclear shuttle protein interactor (AtNSI). kinase (PERK), probably to phosphorylate and regulate NSP function (Florentino et al., 2006; Fontes et al., 2004) . Furthermore, a recent study has shown that a host nucleoprotein, histone H3, interacts with BDMV NSP and MP, forming a complex of histone H3, NSP, MP and viral DNA (Zhou et al., 2011) , to presumably facilitate the export of nascent viral DNA from the nucleus to the cytoplasm for cell-to-cell and long-distance movement. The MP is encoded by the BC1 (or BL1) ORF on the DNA-B components of bipartite begomoviruses. In monopartite begomoviruses, it is encoded by the V2 ORF, which has no sequence identity with BV1. The MP is required for virus cell-to-cell and long-distance movement, through its cooperative interaction with the NSP in bipartite begomoviruses (Noueiry et al., 1994; Ward et al., 1997) . Recent data have suggested that MP-NSP cooperation, leading to nascent viral DNA export from the nucleus, follows a 'couple-skating' transport model (Frischmuth et al., 2007; Hehnle et al., 2004; Zhang S.C. et al., 2001) , in which MP binds to NSP-DNA complexes at the cytoplasmic face of plasma membranes or microsomal vesicles and enables transfer to adjacent cells (Kleinow et al., 2009) . The central region of MP is required for the formation of the MP-DNA-NSP complex (Fig. 7) (Frischmuth et al., 2007; Sanderfoot and Lazarowitz, 1995; Zhang et al., 2002) . Unlike NSP, some geminivirus MPs do not appear to bind to DNA, as elucidated for MPs of SqLCV (Pascal et al., 1994) and AbMV (Hehnle et al., 2004) . In contrast, MPs of BDMV (Rojas et al., 1998) and MYMIV (Radhakrishnan et al., 2008) bind both ss-and dsDNA with high affinity. In monopartite begomoviruses and mastreviruses, the MP is encoded by the V2 ORF (Boulton et al., 1989 (Boulton et al., , 1993 Mullineaux et al., 1988; Wartig et al., 1997) , whereas CP (encoded by V1) shuttles viral DNA between the nucleus and cytoplasm, functionally replacing NSP. Although the monopartite geminivirus CP is required for viral movement (Boulton et al., 1989) , V1 and V2 share no significant sequence identity with BC1 or BV1 (Rojas et al., 2001) . Furthermore, there are functional differences between monopartite begomovirus MPs and MPs of mastreviruses. For instance, similar to bipartite begomovirus MPs, monopartite begomovirus MPs bind viral DNA, whereas mastrevirus MPs do not. Furthermore, cell-to-cell movement in monopartite begomoviruses is mediated by C4 (Krichevsky et al., 2006; Rojas et al., 1998 Rojas et al., , 2001 ; CP, MP and C4 are involved in the delivery of TYLCV DNA, as virions or as nucleoprotein complexes, to the plant cell periphery (Rojas et al., 2001) . To move the complex from the cell periphery for cell-to-cell transport, MP interacts with plasmodesmata (Rojas et al., 2001) and, because mastrevirus MPs do not bind viral DNA, virus movement is effected through interaction with the CP (Kotlitzky et al., 2000; Liu et al., 2001) . Monopartite geminivirus MPs presumably enhance CP-mediated nuclear export of viral DNA and a NES was mapped to the N-terminus of ToLCJV V2 . A model has been proposed, whereby MP diverts the CP-DNA complex from the nucleus to the cell periphery, where it mediates movement through the PD (Liu et al., 2001) . Curtovirus movement has been much less studied compared with that of other geminiviruses. A genetic analysis of BCTV genes showed that the V3 ORF encodes the MP (Hormuzdi and Bisaro, 1995), which forms vesicle-like structures that co-localize with the endoplasmic reticulum (ER) and are trafficked intracellularly from the nucleus to the cell periphery (Soto, 2001) . The significance of these vesicles remains to be established. The ER appears to play important, but differential, roles in MP intracellular trafficking. For instance, the MP of AbMV, a phloemlimited virus, exploits the cellular membrane flow from the ER to the plasma membrane (Zhang et al., 2002) . Conversely, in virusinfected pumpkin sink leaves, MPs of SqLCV (a phloem-limited virus) and CaLCV (which infects all cell types) predominantly associate with ER-derived tubules that penetrate the cell wall and are found only in undifferentiated vascular tissues (procambium) and in lesser amounts within the plasma membrane (Ward et al., 1997) . Furthermore, microtubules appear to play a role in MP-viral genome movement, as has been shown for Euphorbia mosaic virus (Kim and Lee, 1992) and Indian cassava mosaic virus (Roberts, 1989) . In these cases, the MP-containing tubules appear to serve as conduits for the intercellular transport of viral complexes. Several host proteins that interact with geminivirus MPs have been identified, including an Arabidopsis regulator of endocytosis, synaptotagmin (SYTA), which interacts with CaLCV MP at the plasma membrane and directs the MP onto early endosomes (Lewis and Lazarowitz, 2010) . These endosomes are thought to traffic the complex via a recapture pathway to dock at the PD for intercellular transport. A similar interaction occurs between SYTA and the MP of Tobacco mosaic virus, an RNA virus, suggesting a convergent evolutionary lineage between these MPs, as there is no evidence that they share a common ancestry. Synaptotagmins are calcium sensors, and their ability to regulate synaptic vesicle exo-/endocytosis has long been characterized in animals (Fukuda, 2003) and, more recently, in plants (Chapman, 2008) . Recently, AbMV MP has been reported to form a complex with the heat Fig. 7 The movement protein (MP). The BC1 encoded MP contains a pilot domain required for localization to the cell periphery, a central anchor domain that also targets the protein to the cell periphery and a C-terminal domain that facilitates oligomerization. The nuclear shuttle protein (NSP) binding site is in the centre of MP. shock cognate 70-kDa protein to facilitate viral movement along cellular plastids and stromules into a neighbouring cell (Krenz et al., 2010 (Krenz et al., , 2012 . These data suggest that geminivirus movement may be tubule guided and non-tubule guided. It is clear from these findings that geminiviruses, unlike most other virus groups, use different mechanisms to move within and between cells. There have been an increasing number of reports of monopartite begomoviruses occurring in association with circular ssDNA satellites, referred to as alphasatellites (Briddon et al., 2004; Paprotka et al., 2010; Romay et al., 2010) and betasatellites Yang X. et al., 2012) . Furthermore, a new class of satellite DNAs has been found to occur in association with the bipartite Sida golden yellow vein virus (Fiallo-Olivé et al., 2012) . These satellite DNAs depend on the helper virus for encapsidation and systemic infection (Briddon and Stanley, 2006) . Alphasatellites encode a Rep and are capable of autonomous replication in host plant cells . Betasatellites, which encode the bC1 protein, depend on the helper virus for replication (Cui et al., 2005; Saunders et al., 2008) , and may augment the accumulation and symptoms of their helper viruses (Briddon et al., 2003; Jyothsna et al., 2013; Patil and Fauquet, 2010) as a result of the interaction of bC1 with SNF1-related kinase and with S-adenosyl homocysteine hydrolase to suppress RNA silencing (Yang X. et al., 2012) . Because of their very small genomes, geminiviruses have only four to eight proteins, some of which, most notably Rep/RepA, TrAP, and CP have evolved into multifunctional proteins to compensate for the small genomes. Protein multifunctionality is evidence of the plasticity through which evolution brings together functional domains into a single polypeptide chain; together with gene overlap it contributes to genetic economy exhibited by geminiviruses. In this review, current knowledge on the function of these proteins and their interactions with host proteins have been discussed. While many of these interactions and their biological functions have been elucidated, many are still poorly understood and more information will continue to be discovered so as to provide new opportunities in the design of virus control strategies. with the geminivirus nuclear shuttle protein and potentiates viral infection The consensus N-myristoylation motif of a geminivirus AC4 protein is required for membrane binding and pathogenicity The geminivirus nuclear shuttle protein is a virulence factor that suppresses transmembrane receptor kinase activity A geminivirus replication protein is a sequence-specific DNA binding protein Interaction between a geminivirus replication protein and origin DNA is essential for viral replication Simultaneous analysis of the bidirectional African cassava mosaic virus promoter activity using two different luciferase genes The movement protein BC1 promotes redirection of the nuclear shuttle protein BV1 of Abutilon mosaic geminivirus to the plasma membrane in fission yeast Molecular cloning, expression, and characterization of a novel class of synaptotagmin (Syt XIV) conserved from Drosophila to humans Bean dwarf mosaic virus BV1 protein is a determinant of the hypersensitive response and avirulence in Phaseolus vulgaris An RNA transcribed from DNA at the origin of phage fd single strand to replicative form conversion Potyviruses and the digital revolution Genetic determinants of host-specificity in bipartite geminivirus DNA A components Interaction with host SGS3 is required for suppression of RNA silencing by Tomato yellow leaf curl virus V2 protein Simultaneous regulation of Tomato golden mosaic virus coat protein and AL1 gene expression: expression of the AL4 gene may contribute to suppression of the AL1 gene Coat proteins of Rice tungro bacilliform virus and Mungbean yellow mosaic virus contain multiple nuclear-localization signals and interact with importin alpha Complementation of coat protein mutants of pepper huasteco geminivirus in transgenic tobacco plants DNA replication and cell cycle in plants: learning from geminiviruses Geminivirus DNA replication and cell cycle interactions Molecular characterisation of dicot-infecting mastreviruses from Australia Regulation of the activities of African cassava mosaic virus promoters by the AC1, AC2, and AC3 gene products Tomato yellow leaf curl virus (TYLCV) capsid protein (CP) subunit interactions: implications for viral assembly Two classes of short interfering RNA in RNA silencing Geminiviruses: models for plant DNA replication, transcription, and cell cycle regulation gene results in cell death and altered development Movement of tomato yellow leaf curl geminivirus (TYLCV): involvement of the protein encoded by ORF C4 Infection of Tomato leaf curl New Delhi virus (ToLCNDV), a bipartite begomovirus with betasatellites, results in enhanced level of helper virus components and antagonistic interaction between DNA B and betasatellites DNA replication of wheat dwarf virus, a geminivirus, requires two cis-acting signals Watermelon chlorotic stunt virus from the Sudan and Iran: sequence comparisons and identification of a whitefly-transmission determinant Geminivirus-induced macrotubules and their suggested role in cell-to-cell movement Interaction of a plant virusencoded protein with the major nucleolar protein fibrillarin is required for systemic virus infection Cajal bodies and the nucleolus are required for a plant virus systemic infection Expression dynamics and ultrastructural localization of epitope-tagged Abutilon mosaic virus nuclear shuttle and movement proteins in Nicotiana benthamiana cells A geminivirus replication protein interacts with a protein kinase and a motor protein that display different expression patterns during plant development and infection A geminivirus replication protein interacts with the retinoblastoma protein through a novel domain to determine symptoms and tissue specificity of infection in plants Intracellular and intercellular movement of maize streak geminivirus V1 and V2 proteins transiently expressed as green fluorescent protein fusions Cell-free construction of disarmed Abutilon mosaic virus-based gene silencing vectors Self-interaction of Abutilon mosaic virus replication initiator protein (Rep) in plant cell nuclei The induction of stromule formation by a plant DNA-virus in epidermal leaf tissues suggests a novel intra-and intercellular macromolecular trafficking route Complete viral genome sequence and discovery of novel viruses by deep sequencing of small RNAs: a generic method for diagnosis, discovery and sequencing of viruses Nuclear import and export of plant virus proteins and genomes Geminiviruses: a tale of a plasmid becoming a virus Nuclear import of the capsid protein of Tomato yellow leaf curl virus (TYLCV) in plant and insect cells Characterization of a tomato karyopherin alpha that interacts with the Tomato yellow leaf curl virus (TYLCV) capsid protein The Arabidopsis PEAPOD2 transcription factor interacts with geminivirus AL2 protein and the coat protein promoter RKP, a RING finger E3 ligase induced by BSCTV C4 protein, affects geminivirus infection by regulation of the plant cell cycle Induction of plant cell division by Beet curly top virus gene C4 Geminivirus replication: genetic and biochemical characterization of rep protein function, a review Viral movement proteins as probes for intracellular and intercellular trafficking in plants Sequence-specific interaction with the viral AL1 protein identifies a geminivirus DNA replication origin Arabidopsis synaptotagmin SYTA regulates endocytosis and virus movement protein cell-to-cell transport Maize streak virus coat protein binds single-and double-stranded DNA in vitro A single amino acid change in the coat protein of Maize streak virus abolishes systemic infection, but not interaction with viral DNA or movement protein Bean yellow dwarf virus RepA, but not Rep, binds to maize retinoblastoma protein, and the virus tolerates mutations in the consensus binding motif Recruitment of Replication protein A by the papillomavirus E1 protein and modulation by single-stranded DNA Identification of DNA replication and cell cycle proteins that interact with PCNA Geminiviruses subvert ubiquitination by altering CSN-mediated derubylation of SCF E3 ligase complexes and inhibit jasmonate signaling in Arabidopsis thaliana Interaction of geminivirus Rep protein with replication factor C and its potential role during geminivirus DNA replication Replication of porcine circovirus type 1 requires two proteins encoded by the viral rep gene A novel Arabidopsis acetyltransferase interacts with the geminivirus movement protein NSP Geminivirus C4 protein alters Arabidopsis development Inhibition of binding of Tomato yellow leaf curl virus Rep to its replication origin by artificial zinc-finger protein The GroEL protein of the whitefly Bemisia tabaci interacts with the coat protein of transmissible and nontransmissible begomoviruses in the yeast two-hybrid system The hypersensitive response induced by the V2 protein of a monopartite begomovirus is countered by the C2 protein The nucleotide sequence of Maize streak virus DNA Detection of a non-structural protein of MW11000 encoded by the virion DNA of Maize streak virus Processing of complementary sense RNAs of Digitaria streak virus in its host and in transgenic tobacco Regulation of MSV and WDV virion-sense promoters by WDV nonstructural proteins: a role for their retinoblastoma protein-binding motifs Two dicot-infecting mastreviruses (family Geminiviridae) occur in Pakistan Functional analysis of a novel motif conserved across geminivirus Rep proteins Amino acids in the capsid protein of Tomato yellow leaf curl virus that are crucial for systemic infection, particle formation, and insect transmission Two proteins of a plant DNA virus coordinate nuclear and plasmodesmal transport Conserved sequence and structural motifs contribute to the DNA binding and cleavage activities of a geminivirus replication protein Functional domains of a geminivirus replication protein A plasmid of phytoplasma encodes a unique replication protein having both plasmid-and virus-like domains: clue to viral ancestry or result of virus/plasmid recombination? Virology Tomato leaf curl geminivirus from India has a bipartite genome and coat protein is not essential for infectivity The role of AV2 ('precoat') and coat protein in viral replication and movement in tomato leaf curl geminivirus A phage single-stranded DNA (ssDNA) binding protein complements ssDNA accumulation of a geminivirus and interferes with viral movement Plant RNA silencing in viral defence The first DNA 1-like a satellites in association with New World begomoviruses in natural infections The Arabidopsis thaliana homeobox gene ATHB12 is involved in symptom development caused by geminivirus infection The geminivirus BR1 movement protein binds single-stranded DNA and localizes to the cell nucleus Tomato leaf curl Kerala virus (ToLCKeV) AC3 protein forms a higher order oligomer and enhances ATPase activity of replication initiator protein (Rep/AC1) Differential interaction between cassava mosaic geminiviruses and geminivirus satellites Evidence for interplay among yeast replicative DNA polymerases a, d and e from studies of exonuclease and polymerase active site mutations Geminivirus pathogenicity protein C4 interacts with Arabidopsis thaliana shaggy-related protein kinase AtSKeta, a component of the brassinosteroid signalling pathway The N-terminal 62 amino acid residues of the coat protein of Tomato yellow leaf curl Thailand virus are responsible for DNA binding Tomato golden mosaic virus open reading frame AL4 is genetically distinct from its C4 analogue in monopartite geminiviruses Host and viral factors determine the dispensability of coat protein for bipartite geminivirus systemic movement Functional characterization of coat protein and V2 involved in cell to cell movement of Cotton leaf curl Kokhran virus-Dabawali Plant systems for recognition of pathogenassociated molecular patterns The bipartite geminivirus coat protein aids BR1 function in viral movement by affecting the accumulation of viral singlestranded DNA DNA recognition properties of the cell-to-cell movement protein (MP) of soybean isolate of Mungbean yellow mosaic India virus (MYMIV-Sb) Arsenophonus GroEL interacts with CLCuV and is localized in midgut and salivary gland of whitefly B. tabaci Mutagenesis of the virion-sense open reading frames of tomato leaf curl virus ORF C4 of tomato leaf curl geminivirus is a determinant of symptom severity Indian cassava mosaic virus: ultrastructure of infected cells The ribosomal protein L10/QM-like protein is a component of the NIK-mediated antiviral signaling Bean dwarf mosaic geminivirus movement proteins recognize DNA in a form-and size-specific manner Functional analysis of proteins involved in movement of the monopartite begomovirus, Tomato yellow leaf curl virus Association of an atypical alphasatellite with a bipartite New World begomovirus Discovery of a novel mastrevirus and alphasatellitelike circular DNA in dragonflies (Epiprocta) from Puerto Rico Mutagenesis in ORF AV2 affects viral replication in Mungbean yellow mosaic India virus On the alleged origin of geminiviruses from extrachromosomal DNAs of phytoplasmas A monopartite begomovirus-associated DNA beta satellite substitutes for the DNA B of a bipartite begomovirus to permit systemic infection The tomato RLK superfamily: phylogeny and functional predictions about the role of the LRRII-RLK subfamily in antiviral defense Cooperation in viral movement: the geminivirus BL1 movement protein interacts with BR1 and redirects it from the nucleus to the cell periphery A viral movement protein as a nuclear shuttle-The geminivirus BR1 movement protein contains domains essential for interaction with BL1 and nuclear localization Conserved threonine residues within the A-loop of the receptor NIK differentially regulate the kinase function required for antiviral signaling Complementation of African cassava mosaic virus AC2 gene function in a mixed bipartite geminivirus infection A nanovirus-like DNA component associated with yellow vein disease of Ageratum conyzoides: evidence for interfamilial recombination between plant DNA viruses DNA forms of the geminivirus African cassava mosaic virus consistent with a rolling circle mechanism of replication Replication promiscuity of DNA-b satellites associated with monopartite begomoviruses; deletion mutagenesis of the Ageratum yellow vein virus DNA-b satellite localizes sequences involved in replication Top 10 plant viruses in molecular plant pathology Host responses to transient expression of individual genes encoded by Tomato leaf curl virus A NAC domain protein interacts with Tomato leaf curl virus replication accessory protein and enhances viral replication Genetics of resistance to the geminivirus, Bean dwarf mosaic virus, and the role of the hypersensitive response in common bean Dual interaction of a geminivirus replication accessory factor with a viral replication protein and a plant cell cycle regulator Geminivirus C3 protein: replication enhancement and protein interactions Characterization of signals that dictate nuclear/ nucleolar and cytoplasmic shuttling of the capsid protein of Tomato leaf curl java virus associated with DNA beta satellite Tomato leaf curl Java virus V2 protein is a determinant of virulence, hypersensitive response and suppression of posttranscriptional gene silencing Identification of the virulence factors and suppressors of posttranscriptional gene silencing encoded by Ageratum yellow vein virus, a monopartite begomovirus A plant kinase plays roles in defense response against geminivirus by phosphorylation of a viral pathogenesis protein Mutational analysis of the small intergenic region of Maize streak virus ) Promoters, transcripts, and regulatory proteins of Mungbean yellow mosaic geminivirus The 32 kDa subunit of replication protein A (RPA) participates in the DNA replication of Mung bean yellow mosaic India virus (MYMIV) by interacting with the viral Rep protein MYMIV replication initiator protein (Rep): roles at the initiation and elongation steps of MYMIV DNA replication Intracellular distribution of the three virion-sense encoded proteins of Beet curly top virus Subviral DNAs associated with geminivirus disease complexes Nucleotide sequence of Cassava latent virus DNA A symptom variant of beet curly top geminivirus produced by mutation of open reading frame C4 Infectious mutants of Cassava latent virus generated in vivo from intact recombinant DNA clones containing single copies of the genome The nucleotide sequence of an infectious clone of the geminivirus Beet curly top virus Mutational analysis of the monopartite geminivirus Beet curly top virus Demonstration of nicking/ joining activity at the origin of DNA replication associated with the Rep and Rep' proteins of porcine circovirus type 1 Replicational release of geminivirus genomes from tandemly repeated copies: evidence for rolling-circle replication of a plant viral DNA Dynamics of bean dwarf mosaic geminivirus cell-to-cell and long distance movement in Phaseolus vulgaris revealed, using the green fluorescent protein Mutational analysis of potato yellow mosaic geminivirus Transactivation of geminivirus AR1 and BR1 gene expression by the viral AL2 gene product occurs at the level of transcription Regulation of a geminivirus coat protein promoter by AL2 protein (TrAP): evidence for activation and derepression mechanisms Genetic analysis of tomato golden mosaic virus: ORF AL2 is required for coat protein accumulation while ORF AL3 is necessary for efficient DNA replication Heterologous complementation by geminivirus AL2 and AL3 genes Plants expressing Tomato golden mosaic virus AL2 or Beet curly top virus L2 transgenes show enhanced susceptibility to infection by DNA and RNA viruses Involvement of C4 protein of Beet severe curly top virus (family Geminiviridae) in virus movement Suppression of RNA silencing by a geminivirus nuclear protein, AC2, correlates with transactivation of host genes Subcellular targeting of the coat protein of African cassava mosaic geminivirus Differential roles of AC2 and AC4 of cassava geminiviruses in mediating synergism and suppression of posttranscriptional gene silencing Geminiviruses and RNA silencing Suppression of gene silencing: a general strategy used by diverse DNA and RNA viruses of plants Assembly of an active translation initiation factor complex by a viral protein Adenosine kinase inhibition and suppression of RNA silencing by geminivirus AL2 and L2 proteins RNA silencing and plant viral diseases Nuclear export in plants: use of geminivirus movement proteins for a cell-based export assay The geminivirus BL1 movement protein is associated with endoplasmic reticulum-derived tubules in developing phloem cells Genetic analysis of the monopartite tomato yellow leaf curl geminivirus: roles of V1, V2, and C2 ORFs in viral pathogenesis T antigen and its role in the initial steps of SV40 DNA replication Human pathogens and the host cell SUMOylation system Splicing features in Maize streak virus virion-and complementary-sense gene expression Identification and analysis of a retinoblastoma binding motif in the replication protein of a plant DNA virus: requirement for efficient viral DNA replication Plant cells contain a novel member of the retinoblastoma family of growth regulatory proteins GRAB proteins, novel members of the NAC domain family, isolated by their interaction with a geminivirus protein Effect of a single amino acid substitution in the NLS domain of Tomato yellow leaf curl virus-Israel (TYLCV-IL) capsid protein (CP) on its activity and on the virus life cycle C2-mediated decrease in DNA methylation, accumulation of siRNAs, and increase in expression for genes involved in defense pathways in plants infected with beet severe curly top virus Functional modulation of the geminivirus AL2 transcription factor and silencing suppressor by self-interaction Suppression of methylation-mediated transcriptional gene silencing by bC1-SAHH protein interaction during geminivirus-betasatellite infection Human RPA (hSSB) interacts with EBNA1, the latent origin binding protein of Epstein-Barr virus Movement proteins (BC1 and BV1) of Abutilon mosaic geminivirus are cotransported in and between cells of sink but not of source leaves as detected by green fluorescent protein tagging Subcellular targeting domains of Abutilon mosaic geminivirus movement protein BC1 Structure of the Maize streak virus geminate particle Histone H3 interacts and colocalizes with the nuclear shuttle protein and the movement protein of a geminivirus The N-terminus of the Begomovirus nuclear shuttle protein (BV1) determines virulence or avirulence in Phaseolus vulgaris Transcriptional regulation by a point mutant of adenovirus-2 E1a product lacking DNA binding activity I thank Linda Hanley-Bowdoin for critical reading of the manuscript. This work was supported by the National Science Foundation award #0724083. Geminivirus-induced gene silencing of the tobacco retinoblastoma-related