Next Article in Journal
Unrevealing the Mystery of Latent Leishmaniasis: What Cells Can Host Leishmania?
Next Article in Special Issue
Occurrence, Distribution, and Management of Aphid-Transmitted Viruses in Cucurbits in Spain
Previous Article in Journal
SARS-CoV-2-Specific T Cell Responses in Immunocompromised Individuals with Cancer, HIV or Solid Organ Transplants
Previous Article in Special Issue
A Simplified Multiplex PCR Assay for Simultaneous Detection of Six Viruses Infecting Diverse Chilli Species in India and Its Application in Field Diagnosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging and Re-Emerging Diseases Caused by Badnaviruses

by
Alangar Ishwara Bhat
1,*,†,
Ramasamy Selvarajan
2,† and
Velusamy Balasubramanian
2
1
Division of Crop Protection, ICAR-Indian Institute of Spices Research, Kozhikode 673012, Kerala, India
2
Division of Crop Protection, ICAR-National Research Centre for Banana, Trichy 620102, Tamil Nadu, India
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pathogens 2023, 12(2), 245; https://doi.org/10.3390/pathogens12020245
Submission received: 23 December 2022 / Revised: 27 January 2023 / Accepted: 31 January 2023 / Published: 3 February 2023
(This article belongs to the Special Issue Emerging and Re-emerging Plant Viruses in a Context of Global Change)

Abstract

:
New and emerging plant diseases are caused by different pathogens including viruses that often cause significant crop losses. Badnaviruses are pararetroviruses that contain a single molecule of ds DNA genome of 7 to 9 kb in size and infect a large number of economically important crops such as banana and plantains, black pepper, cacao, citrus, grapevine, pineapple, sugarcane, sweet potato, taro, and yam, causing significant yield losses. Many of the species in the genus have a restricted host range and several of them are known to infect a single crop. Combined infections of different virus species and strains offer conditions that favor the development of new strains via recombination, especially in vegetatively propagated crops. The primary spread of badnaviruses is through vegetative propagating materials while for the secondary spread, they depend on insects such as mealybugs and aphids. Disease emerges as a consequence of the interactions between host and pathogens under favorable environmental conditions. The viral genome of the pararetroviruses is known to be integrated into the chromosome of the host and a few plants with integrants when subjected to different kinds of abiotic stress will give rise to episomal forms of the virus and cause disease. Attempts have been made to develop management strategies for badnaviruses both conventionally and using precision breeding techniques such as genome editing. Until 2016 only 32 badnavirus species infecting different crops were known, but in a span of six years, this number has gone up to 68. The current review highlights the emerging disease problems and management options for badnaviruses infecting economically important crops.

1. Introduction

Badnaviruses are plant pararetroviruses that belong to the family Caulimoviridae and contain 68 species demarcated based on the sequence identity in the conserved reverse transcriptase (RT)/ribonuclease H (RNase H) coding region [1,2]. The family Caulimoviridae consists of 11 genera including Badnavirus, namely, Caulimovirus (12 species), Cavemovirus (three species), Dioscovirus (one species), Petuvirus (one species), Rosadnavirus (one species), Ruflodivirus (one species), Solendovirus (two species), Soymovirus (one species), Tungrovirus (one species), and Vaccinivirus (one species) [2]. Of these, Dioscovirus and Ruflodivirus have no particle morphology while badnaviruses and tungrovirus have bacilliform particles and the rest of the viruses have isometric particles. Genomes of all pararetroviruses comprise a double-stranded DNA and they replicate through an RNA intermediate. However, in contrast to retroviruses, integration of the viral genome is not mandatory for the replication of pararetroviruses. Instead, they accumulate as minichromosomes in the host nucleus. Illegitimate and generally fragmented integration occurs once in every million years [3].
Particles of badnaviruses are approximately 30 nm × 120–150 nm in size and their genome consists of a circular, double-stranded DNA genome of 7.2–9.2 kb coding for a minimum of three open reading frames (ORFs) [4,5]. Replication of these viruses occurs via reverse transcription of a greater-than-genome length RNA that serves as a template both for the translation of viral proteins and for reverse transcription to replicate the genome [4,6]. Further, genomes of some badnaviruses integrate into the chromosomes of their hosts through illegitimate recombination, and these linear forms of integrants are known as endogenous badnaviruses [5,7,8]. A few of such endogenous badnaviruses were shown to give rise to the episomal form of the virus (upon exposure to abiotic stress) that cause systemic infection and spread from plant to plant through insect vectors [5,9,10,11].
Badnaviruses are one of the emerging groups of viruses infecting a large number of economically important crop plants, especially in the tropical and subtropical regions of the world [6]. The number of ICTV-recognized badnavirus species in 2016 was just 32 [5], while this number has now increased to 68 species in a span of just six years [2] (Table 1), indicating their importance. Combined infections of different virus species and strains offer conditions that favor the development of new strains via recombination, especially in vegetatively propagated crops. These viruses are reported to cause negligible to significant yield loss varying between 10 and 90% in different crops.
Although badnaviruses occur throughout the world both in monocots and dicots, they cause economically important diseases in tropical and subtropical regions on perennial crops such as banana, black pepper, citrus, cocoa, grapevine, pineapple, sugarcane, sweet potato, taro, and yam. Many of the species in the genus have a restricted host range and several species are known to infect a single crop (cocoa, banana, yam, sugarcane, and grapevine are infected by ten, nine, eight, four, and three distinct species respectively, while pineapple and taro are infected by two species each) (Table 1) (Figure 1). This kind of mixed infection of a single plant by different virus species and strains might lead to the development of new strains/species through recombination, especially in vegetatively propagated crops.
The viral genome of a few badnaviruses and other pararetroviruses is known to be integrated into the chromosome of the host through illegitimate recombination. These sequences are known as endogenous pararetroviruses (EPRVs) [7,12,13,14]. In most cases, EPRVs denote only partial, linear forms of the original circular viral DNA, do not cause any symptoms or disease in the plants, and may provide viral immunity [15]. Since there is sequence similarity between the endogenous and episomal virus sequence, PCR assay may not be able to differentiate them instead, methods such as Southern or in situ hybridization of the chromosome with the viral probe, immunocapture (IC)-PCR, rolling circle amplification (RCA), reverse transcription-PCR, electron microscopy, and host genome sequencing are needed to distinguish episomal and integrated forms of the virus. The integrants of three badnaviruses infecting bananas (banana streak OL virus, BSOLV; banana streak GF virus, BSGFV and banana streak IM virus, BSIMV) give rise to episomal forms in some hybrid plants under abiotic stress conditions such as temperature and tissue culture [9,11,13,16]. The episomal forms of these viruses can cause disease in plants and spread to other plants through mealybug vectors [5].
Badnaviruses cause variable symptoms ranging from asymptomatic, mild, and moderate to severe in the hosts that are influenced by abiotic factors, especially environmental conditions and the nutrient status of the crop (Figure 2). Hence, symptoms are not reliable criteria for the detection of badnavirus infection on a plant. Serological assays are also not reliable due to the low titer and heterogeneity of virus isolates infecting a plant. PCR is the most commonly used assay for the detection of the virus. However, PCR may sometimes detect both endogenous and episomal viruses. Thus, to detect the presence of the episomal form of the virus, RT-PCR or rolling circle amplification (RCA) may be used.

2. Emerging Disease Problems in Plants

2.1. Banana

Banana streak virus disease was first reported in Morocco in the year 1974 by the Lockhart group from the University of Minnesota. Later it was reported to occur across the world including many countries in Africa, Asia, Central, and South America, and the Pacific. It is now known that wherever bananas are grown the streak virus also exists. Banana streak virus induces typical chlorotic or necrotic streak symptoms not only on the leaf lamina but also on the midrib, leaf sheaths, and peduncle. The chlorotic streaks are narrow, small, discontinuous, and parallel to the veins and subsequently, these streaks are turned into necrotic, sometimes splitting of the pseudostem is observed to be due to the virus infection. Internal necrosis of the pseudostem, abnormal bunch emergence by piercing through the pseudostem, and necrotic streaks on fruits have also been recorded. Peel splitting and necrotic spots on fruits are reported to occur. The symptoms are not uniformly distributed on all leaves of the infected tree, and the temperature plays a major role in the type and severity of the of symptoms. The virus titer and the symptoms keep changing due to the temperature and often there will be symptomless growth for a long period, and it reappears suddenly on a new leaf. At 22 °C plants tend to express the symptoms of streak conspicuously whereas above 28 °C to 35 °C, the symptoms are invisible and if the plants are moved to a growth chamber at 22 °C the symptoms are significantly increased and become severe [18,19]. Plants of Pome banana variety Virupakshi (AAB) when subjected to 22 °C for six months to a year are characterized by the expression of streak symptoms, and the virus associated was identified as BSGFV based on the PCR and sequencing of RCA amplified products. Daniells et al. [20] reported a yield reduction of 6% in Cavendish besides an 18 days delay in harvest under Australian conditions. Dallol et al. [9] reported that the process of micropropagation and the hybrids and interspecific natural hybrids of Musa accuminata and M. balbisiana express severe symptoms of streak upon tissue culturing and they assumed to be due to the expression of virus from the endogenous sequence.
BSVs are primarily transmitted through infected planting materials such as corms, corm bits, and infected plants mass propagated through tissue culture and the secondary transmission is through sap-feeding mealy bugs such as Planococcus citri, which transfers the virus in a semi-persistent manner. Lockhart et al. [21] reported BSV initially but later many distinct species of BSV infecting bananas were reported based on the sequence data of the whole genome or RT/RNaseH region. The very first isolate of BSV was later named BSOLV which was from Nigeria and the complete genome of this isolate had 7389 bp with three Open Reading Frames (ORFs) coding proteins of 20.8. 14.5 and 208 KDa of which the polyprotein contained putative movement protein, aspartic protease, RT/RNaseH, and coat protein. BSOLV showed high similarity with another badnavirus infecting sugarcane namely, sugarcane bacilliform virus (SCBV). Subsequently, Geering et al. [22,23] reported two additional new species from Australia and named them banana streak Mysore virus (BSMYV) and banana streak Gold finger virus (BSGFV); later, Harper et al. [24,25] described 13 new BSV species based on the extreme variability in RT/RNaseH sequences from Uganda. Further, Geering et al. [26] reported a complete genome sequence of the banana streak IM virus (BSIMV), and Lheureux et al. [27] reported two distinct complete genome sequences, namely banana streak VN virus (BSVNV) and banana streak Yunnan virus. Currently, nine BSV species are recognized by the International Committee on Taxonomy of Viruses (ICTV), namely, banana streak GF virus (BSGFV), banana streak IM virus (BSIMV), banana streak MY virus (BSMYV), banana streak OL virus (BSOLV), banana streak UA virus (BSUAV), banana streak UI virus (BSUIV), banana streak UL virus (BSULV), banana streak UM virus (BSUMV), and banana streak VN virus (BSVNV) (Figure 3). There are also other tentative species, such as banana streak CA virus (BSCAV), banana streak Peru virus (BSPEV), banana streak UJ virus (BSUJV), banana streak UK virus (BSUKV), and banana streak acuminata Yunnan virus (BSAcYNY) awaiting ICTV recognition (Figure 3).
In bananas, being a clonally propagated fruit crop, detection of viruses in the suckers or corms or corm bits is essential to ensure virus-free planting materials to the growers. Sensitive detection methods are the key to certifying the germplasm and mother plants used for mass propagation free of viruses. BSVs are serologically heterogeneous, and hence serological-based detection is cumbersome; however, IC-RT-PCR is applied using antiserum raised against a mixture of BSVs and SCMV [21]. Moreover, these pararetroviruses have their genomes integrated into the chromosomes, affecting the PCR-based detection of BSV as false positives can be possible from eBSVs. Bananas with a ‘B’ genome always have eBSVs whereas ‘A’ genome varieties are devoid of such integrants. Activatable eBSVs such as BSOLV, BSIMV, and BSGFV and episomal BSVs could be differentiated using RCA and PCR-based approaches, especially for banana plants that contain the ‘B’ genome, while for banana plants with ‘A’ genomes, RCA is the only option to distinguish between endogenous and free viral genomes [5]. Different species of BSVs (including the three species that harbour activable endogenous viruses) causing the streak disease were shown to belong to clade 1 of badnaviruses while clade 2 consisted only of endogenous BSVs that lack episomal counterparts [28]. Clade 3 consisted of species of BSVs from East African countries that also cause streak disease but lack endogenous counterparts. Based on the three abovementioned clades, a model for the co-evolution of virus and banana was proposed and it was suggested that endogenous badnaviral sequences can be used as markers to characterize the Musa phylogeny [28,29]. Recombination analysis of BSV and SCBV sequences detected a total of 32 recombinants between BSV and SCBV, besides many intra-BSV and intra-SCBV recombinants [30]. The breakpoints in the recombinants are mainly located in the intergenic region and C-terminal of ORF3 in both species of the viruses. The unique fragments detected in the analysis suggest the possibility that BSV and SCBV or their ancestors infected the same host before making the host shift.
A real-time PCR was reported for the rapid detection of episomal BSOLV and BSMYV [31], immunocapture PCR (IC-PCR), multiplex immunocapture PCR (M-IC-PCR), and rolling circle amplification (RCA) have been developed for the detection of episomal circular BSV genomes [32,33,34,35]. Reverse transcription PCR (RT-PCR) was successfully used to target the transcripts of episomal BSMYV [36,37]. James et al. [33] developed the RCA method using BSV-specific primers and demonstrated the ability of this method to differentiate between episomal and integrated viral genomic sequences. Serological-based specific detection was reported for BSMYV using the antibodies produced against the bacterially expressed virus-associated protein encoded by ORF II [38]. Zhang et al. [39] developed an RT-LAMP multiplex assay to detect three different DNA and RNA viruses in bananas. The reverse transcription-recombinase polymerase amplification assay (RT-RPA) has been developed for the detection of episomal BSMYV [40]. Recently, a rapid, sensitive, and simultaneous detection method was developed for BSV, BBTV, and CMV using a bead-based multiplex assay has been used for diagnosis [41].
Elimination of virus from the infected propagule using cryotherapy has become successful for many plant viruses. Helliot et al. [42] reported that 90% of virus elimination could be achieved in cv. Williams (AAA) of the Cavendish subgroup using cryopreservation of meristem through vitrification by plant vitrification solution 2 (PVS-2) solution whereas only 76% elimination was achieved by meristem tip culture. Cryopreservation followed by apical meristem culture has a high rate of success in eliminating episomal BSV infections in cv. Williams, a banana genotype free of eBSV [42]. Two anti-retroviral and anti-hepadnavirus molecules have also been effective in eradicating the episomal form of BSV from banana plants [43]. Mutating the eBSV sequences of bananas shall stop the spontaneous production of episomal virus and may stop the expression of the symptoms. Genome editing is widely used in agriculture to add new traits or to control viral pathogens by manipulating the host and virus protein interactions. Tripathi et al. [44] applied a CRISPR/Cas9-based genome-editing approach for banana and demonstrated by inactivation of the endogenous eBSV sequence integrated into the ‘B’ genome of plantain using multiple gRNAs.

2.2. Black Pepper

The yellow mottle disease (also known as stunt disease) of black pepper (Piper nigrum) caused by the badnavirus, Piper yellow mottle virus (PYMoV) is reported in Brazil, China, India, Indonesia, Malaysia, Philippines, Sri Lanka, Thailand, and Vietnam [45,46,47,48,49]. The disease is characterized by chlorotic flecks, mild to severe mottles, and brittle, leathery, deformed, and reduced leaf size (Figure 2). The length between the nodes of infected plants is reduced at the later stage, leading to the stunting of the entire vine, hence the name stunt disease [47]. The yield loss due to the disease may vary from negligible to up to 80% depending on the severity of the disease. Symptom expression and disease severity are positively correlated with high temperature and high relative humidity [50]. Severe diseases also develop under poor nutrition of the plants. Thus, soil and plant health play an important role in the symptom expression and the severity of the disease. Lockhart et al. [45] for the first time reported the association of a distinct badnavirus with yellow mottle disease affecting black pepper from Indonesia, Malaysia, Philippines, and Thailand and named it Piper yellow mottle virus (PYMoV) based on bacilliform virus particles, transmission by mealybug, its weak serological relationship with other badnaviruses, and its partial nucleotide sequence of the reverse transcriptase (RT)/RNase H region of the virus. Later association of PYMoV with black pepper was reported in Brazil [51], Sri Lanka [46], India [47], Vietnam, and China [48]. In addition to black pepper, PYMoV also infects other economically important Piper species such as P. betle (betelvine), P. longum (Indian long pepper), and many other wild Piper species [52,53]. The first complete genome sequence of the PYMoV of an Indian isolate was 7622 nucleotides long, potentially coding for four ORFs of proteins with 16, 17, 219, and 17 kDa, respectively [54]. Currently, the complete genome sequence of six isolates of PYMoV infecting black pepper, betelvine, and Indian long pepper originating from India and China showed that the length of the genome varied from 7557 to 7584 nucleotides and all isolates possessed four ORFs [48,55]. Of this, lengths of ORF 2 and ORF 4 showed country-specific characteristics. Further, the complete genome phylogenetic tree also showed country-specific clustering of PYMoV isolates [56]. The ORF III of the PYMoV codes a polyprotein with homology to different proteins such as viral movement, trimeric dUTPase, zinc finger, aspartic protease, reverse transcriptase, and RNase H, while proteins coded by the other three ORFs (I, II, and IV) do not show homology to any known proteins. PYMoV showed 39% to 56% identity with other badnaviruses and phylogenetic analysis using either the complete genome or conserved region of the RT/RNaseH showed grouping of all PYMoV isolates into a single cluster separated from other badnaviruses. Black pepper is a perennial and propagated by vegetative means; the primary spread of the virus occurs through vegetative means while secondary spread in the field happens through semi-persistent transmission by mealybugs (Ferrisia virgata and Planococcus citri) and black pepper lace bug [45,46,47]. Methods such as PCR, real-time PCR, loop-mediated isothermal amplification (LAMP), and recombinase polymerase amplification (RPA) are reported for the identification of virus-free mother plants for propagation [57,58,59,60]. Further, the elimination of PYMoV through somatic embryogenesis combined with pre-treatment with an antiviral agent, ribavirin @20 mg/mL, and meristem tip culture combined with chemotherapy was reported [61]. Production and use of virus-free planting materials should be the main focus for the management of the virus. Under field conditions, controlling the mealybug vector, correcting the soil acidity and NPK application (based on the soil test), spraying with a micronutrient mixture, and application of farm yard manure and plant growth-promoting rhizobacteria consortia are recommended [62].

2.3. Citrus

A badnavirus associated with the yellow mosaic disease in citrus from India was reported based on the morphology of virus particles, serological reaction, and the partial genome sequence by Ahlawat et al. [63], and the causative virus was named citrus yellow mosaic virus (CYMV). The typical symptoms of the disease are vein flecking and yellow mosaic on the leaves of sweet oranges with an incidence as high as 10–70% and a yield loss of 77% with lesser juice and ascorbic acid content. Dodder is known to transmit CYMV to different citrus species such as sweet and sour oranges, rangpur lime, Pummelo, and Volkamer lemon. Interestingly the CYMV is also transmitted by mechanical sap inoculation to sweet orange, pummelo, and Citrus decumana. As with all the other badnaviruses, CYMV is also a non-enveloped bacilliform virus having a dimension of 150 × 30 nm in size with a 7559 bp length circular dsDNA with some discontinuous in its genome [63,64]. An infectious clone of CYMV was shown to infect sweet oranges through agro-infection [64]. Vadlamudi et al. [65] reported that ORFI of CYMV possesses RNA-silencing suppression activity. The complete genome sequence of CYMV from sweet orange, rangpur lime, acid lime, and pummelo shared >90% identity [66,67]. Endogenous florendovirus has also been reported to occur in citrus [68]. An in vitro-expressed virion-associated protein of the virus was reported to be useful for the detection of the virus [69]. Dot blot hybridization, RT-PCR, real-time PCR, and RPA-based methods have been reported for the detection of the virus in plants [69,70,71,72].

2.4. Cacao

Cacao swollen shoot disease (CSSD) of cacao (Theobroma cacao L.) was first discovered in Ghana in 1936 and later in other West African countries such as Ivory Coast, Liberia, Nigeria, Senegal, Sierra Leone, and Togo [73]. It is also reported in other countries such as Brazil [74], Indonesia [75], Sri Lanka [76], and Trinidad [77,78]. The typical symptom of the disease is a swollen shoot, which may be rounded on the end, and necrosis and death [79]. The foliar symptoms include vein-clearing, mosaic/mottling, red vein-banding, fern-like patterns, and leaf necrosis [80]. Both virulent and mild strains of the virus are known to occur with severe strains, causing the rapid decline and sudden death of plants [81]. The diseased plants produce a few smaller and discolored pods with poor bean quality. Agro-inoculation of the viral genome into the cacao plant showed the presence of virions in the cytoplasm of phloem companion cells and a few xylem parenchyma cells [82]. Further, stem swelling occurs due to the proliferation of the xylem, phloem, and cortex cells. The CSSD-associated viruses are transmitted semi-persistently by Planococcoides njalensis and Planococus citri and at least other 12 species of mealybugs and also through seeds and grafting [83,84,85].
Initially, the disease was thought to be caused by different strains of the virus that cause mild, moderate, and severe symptoms. Later, viral genomic sequence data indicated that many distinct badnavirus species are associated with the disease. Currently, ICTV lists ten different cacao swollen shoot virus species associated with the disease in different cacao-growing regions. They include cacao bacilliform Sri Lanka virus (CBSLV), cacao mild mosaic virus (CaMMV), cacao swollen shoot CD virus (CSSCDV), cacao swollen shoot CE virus (CSSCEV), cacao swollen shoot Ghana M virus (CSSGMV), cacao swollen shoot Ghana N virus (CSSGNV), cacao swollen shoot Ghana Q virus (CSSGQV), cacao swollen shoot Togo A virus (CSSTAV), cacao swollen shoot Togo B virus CSSTBV, and cacao yellow vein banding virus (CYVBV) [2,86]. The different species associated with CSSD encode four to six ORFs named ORFs1-4, X, and Y and an intergenic region [87]. Of these, ORF 2 encodes a protein of about 14–15 kDa with homology to nucleic acid binding activity while ORF 3 encodes a polyprotein of about 210 kDa containing domains for the movement protein, coat protein, aspartic protease, and reverse transcriptase (RT) and ribonuclease H (RNaseH) proteins [74,75,77,78,86,87,88,89,90,91,92]. The functions of the remaining ORFs (ORF I, X, and Y) are unknown. Integrated badnaviral sequences in the genome of many cacao genotypes named endogenous T. cacao bacilliform virus (eTcBV) were reported [93]. Mixed infections, sequence variation, and recombination resulted in the evolution of new badnavirus species in cacao. Among various badnaviruses, CSSTBV showed the least nucleotide diversity while CSSCEV exhibited the highest variability [94]. Recombination analysis of 82 badnaviral genomes showed eight recombination events located between the intergenic region and ORF 2 in the West African species CSSTBV, CSSCDV, and CSSTAV. Similarly, a strain of the CaMMV (CaMMV-PR3) was identified as a possible recombinant, with CaMMV-BR as the major parent [94].
Serological and PCR amplification assays, though developed, failed to detect the virus in all symptomatic plants owing to the association of many badnaviral species in the disease. In a study using 189 samples, the detection potential of the primers varied from 4% to 23% with the four best primers [95] indicating the need for new approaches for the detection and diagnosis of different badnaviruses. Based on PCR using primers specific for different viruses, it was inferred that CSSTBV is the predominant species associated with diseased cacao in Côte d’Ivoire [74]. Management of the disease mainly focused on the eradication of virus-infected plants, resistance breeding, mild strain cross-protection, and control of mealybug vectors. Rogueing infected trees and replanting with tolerant genotypes showed only limited success in the management of the disease. Germplasms introduced from South America to west Africa were also susceptible to the disease. The mild strain CSSTBV-N1 protected the plants against the severe strain of the virus for 11–14 years thereafter additional field inoculation with the mild strain was needed for achieving continuous protection against the severe strain [96]. Barrier cropping is another effective way to reduce the disease spread while providing shade to virus-infected plants reduces symptom severity [97]. Production of virus-free cacao plants by eliminating the virus from the virus-infected plants through somatic embryogenesis was reported [98].

2.5. Dioscorea

Dioscorea bacilliform virus (DBV) infects several species of yam and negatively impacts production in all the regions of the world where the crop is grown. The common symptoms are chlorosis on veins and necrosis, and leaf distortions such as crinkling and puckering have been noticed. DBVs are primarily spread through vegetatively propagated tubers of yam which helps the virus to perpetuate, and it is also transmitted mechanically through the sap; secondarily, they are transmitted by multiple species of mealybugs of the family Pseudococcidae. Turaki et al. [99] reported that there are several species of badnaviruses infecting the germplasm of yam in West Africa. Besides the episomal virus, a diverse range of endogenous sequences of DBV (eDBV) has been reported to occur on the yams cultivated around the world and more particularly in West Africa. However, so far, no eDBVs are reported to form into episomal viruses through homologous recombination as reported in bananas.
Both DBVs and eDBVs have been characterized by genome sequencing. The first complete genome sequence of DBV named DBALV is from Nigeria and infects Dioscorea alata. The complete genome sequence of two isolates of DBALV had 7413 and 7415 nt with three potential ORFs coding 29, 25, and 228 KDa proteins. Subsequently, two more isolates from Benin with 7262 and 7276 nt that showed 62% similarity with DBALV were named Dioscorea sansibarensis bacilliform virus (DBSNV). Later, another six distinct DBV genomes classified to be different species were completely sequenced; the six DBVs are, viz., dioscorea bacilliform AL virus 2 (DBALV2), dioscorea bacilliform ES virus (DBESV), dioscorea bacilliform RT virus 1 (DBRTV1), dioscorea bacilliform RT virus 2 (DBRTV2), dioscorea bacilliform RT virus 3 (DBRTV3) and dioscorea bacilliform TR virus (DBTRV) [100,101,102,103]. All the above-mentioned eight species are recognized by ICTV while seven more isolates are awaiting recognition [104] (Figure 4). The yam germplasm of the Pacific region contained four species (DBALV, DBALV2, DBRTV2, and DBRTV2). In Vanuatu and Tonga, DBALV was the most common species found whereas DBALV2, DBESV, and DBRTV2 were present in Papua New Guinea, Fiji, and Samoa. Recombination analysis of different DBV genomes revealed the occurrence of a unique recombination event located between the intergenic region and 5′ end of ORF 1 in the DBRTV3 with DBSNV and DBALV as major and minor parents, respectively [104].
DBV genomes are known to be integrated into the genomes of yam, which complicates the detection of the virus [102,105]. High heterogeneity among the episomal DBVs and lack of specific polyclonal and monoclonal antisera also hamper the detection of DBV by serological methods. Bomer et al. [100] developed an effective RCA protocol to identify, detect, and characterize episomal DBVs infecting yam. Turaki et al. [99] developed a PCR-dependent denaturing gradient gel electrophoresis (PCR-DGGE) for a quick and effective determination of complex mixtures of potentially episomal and endogenous DBV sequences. Onsite diagnostics using nanopore-based high-throughput sequencing for the discovery and reliable detection and sequencing of full-length genomes of yam viruses was reported [106]. The use of virus-free clean planting material, surveillance, vector management, and removal of alternate hosts of DBVs need to be integrated to manage the disease in the field.

2.6. Grapevine

Grapevine is infected by three badnaviruses of which the grapevine vein-clearing virus (GVCV) is one of the most important emerging viruses whose distribution is, so far, limited to the USA. GVCV causes symptoms such as vein-clearing of leaves, reduced internode length, and vine decline syndrome and it is transmissible by the aphid Aphis illinoisensis [107,108]. Grapevine Roditis leaf discoloration-associated virus (GRLDaV) causes yellow and reddish discoloration, deformity, and reduction in the leaf size and sugar content of the berries of grapevine, cv. “Roditis” in Greece, Italy, South Africa, and Turkey [109,110,111,112,113,114], while grapevine badnavirus 1 (GBV-1) is reported from Croatia [115]. Besides grafting, both GRLDDaV and GBV-1 are transmissible through the vine mealybug (Planococcus ficus) [114,115,116]. The complete genome of GVCV and GBV-1 contained 7753 and 7145 nucleotides, respectively, potentially coding for three typical ORFs while that of GRLDaV contained 6988–7090 nucleotides with four ORFs.

2.7. Pineapple

Virions of pineapple bacilliform virus (PBV) were first detected in pineapples [117], and later occurrence of a badnavirus-like sequence was detected in a large number of pineapple plants [118]. A degenerate PCR and immune-capture PCR were used to determine the diversity of badnaviruses in pineapples grown in Australia that indicated the occurrence of two new badnaviruses, the pineapple bacilliform comosus virus (PBCOV) and pineapple bacilliform erectifolius virus (PBERV), and an endogenous badnavirus, Endogenous pineapple pararetrovirus-1 (ePPPV-1), and a retrotransposon, Ananas metavirus (AMtV) [119]. Hawaiian pineapples also showed an occurrence of badnaviral sequences and were designated as pineapple bacilliform comosus virus-HI1 (PBCOV-HI1) along with nine genomic variants (A to H) [120]. The RT and RNaseH regions of PBCoV-HI1 showed 98% nucleotide similarity with PBCOV from China and Australia, suggesting that they are strains of a single species [121], whereas the amino acid level and the identity for ORF1, ORF2, and ORF3 were only 47% to 80%. The complete genome sequences of PBCOV isolates from China and Hawaii comprised 7543 and 7451 bp with the three ORFs [120,121]. Among the two, PBCOV is transmitted by mealybugs (Dysmicoccus neobrevipes, D. neobrevipes, and Planococcus citri) [117,119,120].

2.8. Sugarcane

The first recorded report of sugarcane bacilliform virus (SCBV) was from Cuba in 1985 followed by many other sugarcane-growing countries of the world through germplasm exchange due to lack of characteristic symptoms and vegetative propagation of the sugarcane [122,123]. Though the disease is characterized by symptoms such as chlorosis, mottle, and leaf freckle, many of the infected plants remain asymptomatic. However, symptoms are prominent when there is a mixed infection of SCBV with other viruses such as the sugarcane mosaic virus and sugarcane mild mosaic virus [123]. In China, SCBV-infected sugarcane is reported to decrease juice sucrose content, gravity purity, and stalk weight [124]. It is suggested that SCBV originated at the center of sugarcane origin in Papua New Guinea and spread to other countries. The secondary spread of the virus occurs semipersistently through mealybug vectors such as Sacharicoccus sachhari and Dysmicoccus boninsis [125,126,127]. Recently, true seed transmission of SCBV was shown in 18 crosses and their parents through PCR assays in India [128]. Even the pluff samples of parents used in the hybridization also tested positive for the virus, suggesting that the seed acquires the virus from the parents. Initial studies showed that SCBV and BSV were indistinguishable as SCBV-transmitted bananas showed symptoms similar to those described for BSV. Similarly, a Moroccan isolate of BSV and SCBV were serologically indistinguishable and a few isolates of BSV and SCBV could be transmitted mechanically to healthy sugarcane and banana, respectively [125]. Unlike many badnaviruses, SCBV has a relatively broad host range infecting members of the Poaceae and Musaceae, Chenopdoium quinoa, and tobacco [6]. SCBV is reported from noble canes, hybrids, and clones within the ‘Saccharum complex’ including S. robustum, S. spontaneum, S. barberi, Erianthus arundinaceus, and E. ravennae [129].
Currently, ICTV recognizes four species, namely, sugarcane bacilliform Guadeloupe A virus (SCBGAV), sugarcane bacilliform Guadeloupe D virus (SCBGDV), sugarcane bacilliform IM virus (SCBIMV), and sugarcane bacilliform MO virus (SCBMOV) while other isolates are still considered only as tentative species. Of these, the SCBMOV was the first to be completely sequenced, followed by SCBIMV [127,130]. The complete genome of the SCBMOV from Morocco contained 7568 nucleotides coding for typical three ORFs and an intergenic region [127]. Infectious clones of this virus when inoculated through agro-inoculation caused an infection in rice and banana plants. Complete genome sequencing of the SCBIMV from Australia (Ireng Maleng isolate) contained 7687 nucleotides that shared only 72% nucleotide sequence identity with SCBMOV. Later, based on the high sequence variability in the RT/RNaseH region among 35 isolates from Guadeloupe, seven phylogenic groups (A to G) including SCBMOV (group E) and SCBIMV (group F) were reported [131]. Based on the complete genome of the isolates from groups A and D, two more new species, namely, sugarcane bacilliform Guadeloupe A virus (SCBGAV) and sugarcane bacilliform Guadeloupe D virus (SCBGDV) were recognized by ICTV [131]. Similarly, Karuppaiah et al. [132] reported sequence variability ranging from 15 to 31% in the complete genome of five isolates of SCBV from India that showed 70–82% identity with SCBIMV and SCBMOV and suggested the creation of three more new species for which the names sugarcane bacilliform black Reunion virus (SCBBRV), sugarcane bacilliform BO virus (SCBBOV), and sugarcane bacilliform BB virus (SCBBBV) were proposed, which are yet to attain ICTV recognition. The complete genome sequence of two divergent strains of SCBV infecting sugarcane in China that aligned with SCBV-H and SCBV-G clades was reported [133]. Similarly, based on the RT/RNaseH region sequence analysis of 63 SCBV-infected sugarcanes from different regions of China, they were grouped into nine SCBV phylogroups of which phylogroups SCBV-S and SCBV-T were newly proposed [124]. Later, Janiga et al. [134] added five more phylogroups (SCBV-U to SCBV-Y) based on the RT/RNaseH region sequencing of 104 SCBV-infected sugarcane germplasms and Saccharum hybrids from India. Recently the occurrence of SCBV and its variability was reported in Ethiopia [135]. High genetic variability observed among SCBV isolates infecting sugarcane around the world clearly indicates the co-evolution of the virus and its host. Until now, complete genome sequences of 12 isolates of SCBV are available in the public domain. The occurrence of recombinant strains within SCBV and between SCBV and BSV strains was reported [133,136]. The putative species SCBV-U, SCBV-W, and SCBV-X and six other species were found to be recombinants, indicating the genetic exchange occurring between SCBV species over time, leading to the evolution of new strains/species [134]. The intergenic region was considered to be the hotspot for recombination followed by ORF1. Geijskes et al. [137] reported non-integration of the SCBV genome in the sugarcane chromosomes; however, based on PCR amplification and Southern hybridization, Cai et al. [138] reported integration of SCBV DNA fragments into the chromosomes of Saccharum inter-specific hybrids grown in China.
The variable region from SCBV was shown to serve as a promoter for high-level expression of foreign genes in both monocot and dicot transgenic plants [139,140]. miRNAs for silencing of the SCBV genome using in silico algorithms identified 14 potential candidate miRNAs from sugarcane, of which sof-miR159e was the top candidate that is capable of targeting ORF 3 of SCBV that can be exploited for the development of SCBV-resistant sugarcane plants [141].

2.9. Sweet Potato

Sweet potato pakakuy virus (SPPV) is a collective name coined for two distinct badnaviruses, namely, sweet potato badnavirus a (SPVa) and sweet potato badnavirus b (SPVb) obtained through siRNA sequencing of a virus-infected sweet potato landrace from Peru [142]. The complete genome of SPVa contained 8082 nt while that of SPVb contained 7961 nt, and both had four ORFs. The occurrence of both these viruses was also reported in Honduras, Guatemala, and Tanzania through small-RNA sequencing of symptomatic sweet potato plants [143,144]. The occurrence of an additional virus species based on the partial sequence of 3065 nt encoding putative movement and structural coat proteins, showing 86.3% and 73.1% identity to SPVb and SPVa, respectively, was also reported. PCR-based diagnostics and virus elimination through heat treatment combined with meristem tip culture were reported in Ethiopia for the management of the virus [145]. SPPV was omnipresent in very low titers in sweet potato germplasm and cultivars grown throughout the world and it does not occur in endogenous form. It is transmissible through seeds and cuttings and does not induce apparent symptoms [146]. Considerable variation was observed in the nucleotide sequences among isolates of the virus and it showed only a limited synergistic increase in the titers in plants infected with other viruses. Based on this it was concluded that these viruses persist and may not be a production constraint in sweet potatoes.

2.10. Taro

Taro (Colocasia esculenta) is a perennial aroid crop grown for its edible corms in Southeast Asia and the Pacific Island countries is infected by two badnaviruses, namely, taro bacilliform virus (TaBV) and taro bacilliform CH virus (TaBCHV) [2]. The disease is characterized by vein clearing, stunting, and downward curling of leaf blades, though sometimes infected plants do not show any external symptoms. TaBV and TaBCHV are transmitted by vegetation, seeds or pollen, and mealybugs (Pseudococcus solomonensis) [147,148]. TaBV and TaBCHV differ in their genome sequence and organization-TaBV containing four ORFs while TaBCHV contained six ORFs. Currently, the complete genome sequence of six isolates of TaBV representing Papua New Guinea, Australia, and East Africa and eight isolates of TaBCHV representing China, East Africa, and Hawaii are available. The complete nucleotide sequence of TaBV varied from 7458 to 7805 nucleotides in different geographical isolates with genome organization of ORFs I to III similar to the typical badnaviruses, while the ORFs IV and X are similar to those of ORF IV of CYMV and ORF X of CSSV and in the phylogenetic analysis, TaBV showed a close relationship with DBV [148]. Analysis of the coding region sequence of RT/RNaseH of different isolates representing different geographical regions also showed variability up to 23% and 33% in the coat protein region [149]. TaBCHV was first identified based on small-RNA sequencing of symptomatic taro plants from China, the complete genome of which showed 7641 nucleotides (identity with TaBV was only 55.8%) with the potential to code for six ORFs [150]. Currently, the complete genome of eight isolates of TaBCHV ranging in nucleotides from 7389 to 7647 with six ORFs encoding proteins of 16.8, 14.1, 206.4, 13.2, 11.9 and 12.4 kD is reported [2]. Yang et al. [151] reported a PCR assay for the detection of TaBV infecting taro and also suggested the possibility of the integration of TaBV sequences in the taro chromosome. A region of TaBV genome was successfully used as a promoter for the expression of foreign genes in bananas and tobacco, suggesting the possibility of using the same in plant expression vectors for high-level constitutive expression of transgenes both in dicotyledonous or monocotyledonous plants. The approaches suggested for the management include adhering to the international guideline for the taro movement including screening for viruses and developing virus-free healthy taro for farmers [152,153], besides the removal of infected plants and insecticide spray to control the vectors.

2.11. Badnaviruses Infecting Other Plants

Badnaviruses are also reported to infect several ornamentals, medicinal herbs, fruits, trees, and weeds [154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189]. Birch leaf roll-associated virus, pagoda yellow mosaic-associated virus, and mulberry badnavirus 1 are known to infect different tree crops (Table 1) [156,157,158]. Similarly, aglaonema bacilliform virus, aucuba ringspot virus, bougainvillea spectabilis chlorotic vein-banding virus, camellia lemon glow virus, canna yellow mottle-associated virus, canna yellow mottle virus, codonopsis vein-clearing virus, epiphyllum mottle-associated virus, ivy ringspot-associated virus, kalanchoë top-spotting virus, spiraea yellow leaf spot virus, wisteria badnavirus 1, and rubus yellow net virus are reported to infect different ornamentals (Table 1) [154,159,160,161,162,163,164,165,166,167,168,170,171,172,174,175,176], while cycad leaf necrosis virus infects the gymnosperm Zamia fischeri [177]. Viruses such as blackberry virus F, chestnut mosaic virus, fig badnavirus 1, gooseberry vein banding-associated virus, and jujube mosaic-associated virus infect different fruit crops [155,178,179,180,181,182,183,184,185], while polyscias mosaic virus and green Sichuan pepper vein clearing-associated virus infect medicinal and spice crops [186,187]. The yacon necrotic mottle virus infects the tuber crop, Smallanthus sonchifolius while commelina yellow mottle virus, the type species of the genus, Badnavirus, infects the weed plant (Commelina diffusa) [188,189]. Of these, yield loss data are not available for many viruses except for the rubus yellow net virus, which infects red and blackberry in North America and Europe, causing a loss of 30–75% in the first year in mixed infection with black raspberry necrosis virus [154]. The majority of these viruses cause vein clearing, mosaic, mottling, chlorosis, necrosis, leaf malformation, and stunting, and are transmitted by vegetative propagules, seeds, aphids, and mealybugs.

3. Recombination among Badnaviruses

Recombination analysis provides important information on the evolution of viruses. It is considered a major factor contributing to the genomic variability and diversity among badnaviruses. We subjected complete genome sequences of 215 badnaviruses available in the GenBank for recombination analysis using RDP, Geneconv, Boot-scan, Maximum Chi Square, Chimaera, SisterScan, and 3Seq implemented in the Recombination Detection Program (RDP) v.4 with default settings [190]. The statistical significance was inferred by a p-value lower than a Bonferroni-corrected cut-off of 0.05. Only recombination events detected by five or more of the different methods were considered to yield a statistically reliable prediction. Recombination events were considered credible when an event was identified by at least three detection methods with an associated p value < 0.05. The present analysis detected a total of 34 independent recombination events, the majority of which are located between the intergenic region and ORF 2 (Table S1). Of these, recombinants such as BSMYV-TRY, BSMYV-IN2, and BSMYV-IN9 were detected using seven RDP detection methods, followed by recombinants such as TaBV-Ke52, CYMV-ROL, and CYMV-Nagri, detected using six RDP detection methods, and recombinants such as JuMaV-HZ/AKS-6, CSSCDV-CI152-09, TaBCHV-PNG-K, BLRaV-BpenGer407526, detected using five RDP detection methods (Table S1). Among all, maximum recombinants were detected in the BSMYV-TRY isolate with five different break vent points followed by BSMYV-IN9 isolate with four predicted recombinants. For different recombinants of BSV species (except for the BSGF-Yunnan), either isolates of the same or different BSV species served as major and minor parents. For BSGF-Yunnan, BSOLV-5-S68 acted as the major parent, and SCBGAV1 as the minor parent. Similarly, for different recombinants of CYMV, in addition to CYMV isolates, an isolate of CSSGNV served as the major parent (Table S1). For recombinants JuMaV, in addition to isolates of JuMaV, CSSCDV, and PYMoV served as parents. Similarly, the four predicted recombinants of TaBV, in addition to TaBV, JuMaV, BSOLV, BSUMV, and ComYMV served as parents. For TaBCHV, DBALV2, YNMoV, CSSCEV, GRLDAV, DBALV2, and CYVBV served as major and minor parents. Some of the recombinant events in the present study were detected using unknown parents, suggesting the existence of undetermined diversity in badnaviruses. Mixed infection of a crop with more than one virus isolate/species might lead to the emergence of a new variant/species. The majority of badnaviruses do not cause any visible external symptoms; cultivation of such plants may facilitate the transmission of viruses between cultivars/varieties/germplasm, leading to the emergence of more virulent recombinants.

4. Conclusions and Future Research Needs

Badnaviruses are emerging viruses that cause significant economic loss in yield and quality, especially in the tropical region of the world in crops such as banana, black pepper, citrus, cacao, grapevine, sugarcane, taro, and yam. ICTV listed 68 species in the genus with many hosts (banana, cacao, grapevine, pineapple, sugarcane, sweet potato, taro, and yam) infected by more than one species. In general, viruses in the genus have a restricted host range, infecting a limited number of hosts. The majority of the badnaviruses are transmitted by different species of mealybugs in a semipersistent manner. Wild and uncultivated plants infected with badnaviruses usually remain symptomless or show only mild symptoms. The severity of symptoms of badnavirus-infected cultivated plants generally depends on abiotic factors such as temperature, relative humidity, light intensity, and nutrition of the plants. Hence, more studies are needed on the emerging problems of badnaviruses in the light of changing climatic conditions to identify the factors that trigger symptom development and the severity of the diseases in different hosts. A few studies have also indicated the role of soil pH and nutrients, including micronutrients in symptom development, also need detailed investigation. Similarly, the effect of beneficial and pathogenic organisms including fungal, bacterial, and viral on the badnavirus-infected plants also need to be studied. The information generated from the abovementioned studies would help to develop a strategy to manage the disease.
Currently, the complete genome of many badnaviruses has been determined and many useful assays based on PCR, LAMP, and RPA combined with the lateral flow have been developed for the quick and sensitive detection of viruses that help in the production of virus-free plants. Similarly, methods such as RCA, in situ hybridization, and RT-PCR and IC-PCR are available to differentiate plants infected with episomal and endogenous badnaviruses. The recombinant analysis predicted the occurrence of recombinants among badnaviruses and also suggested the existence of hitherto undetermined diversity in badnaviruses. The intergenic regions of a few badnaviruses have proven that they can be used as promoters for the expression of foreign genes in different hosts. However, the role of different genomic components of the virus on disease development is still unknown.
Methods such as thermo- and chemo-therapies combined with meristem tip culture were reported for the production of virus-free plants of a few badnaviruses. This should be expanded to more hosts, especially the germplasm accessions that are used to breed crops resistant to different pathogens. So far, no varieties resistant to badnaviruses are reported. Hence, more research efforts are needed to develop varieties resistant to badnaviruses using both conventional and biotechnological including gene editing approaches. Recently, endogenous badnaviral sequence-edited banana plants lost their ability to give rise to the episomal form. The role of integrated badnavirus sequences that might impart immunity to their hosts requires further study.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pathogens12020245/s1, Table S1: Predicted recombination events detected with badnavirus isolates based on complete genome sequences.

Author Contributions

A.I.B. and R.S. collected all literature and prepared the manuscript, V.B. prepared all figures. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The senior author is thankful to the Science and Engineering Research Board (SERB), Government of India for supporting the project (CRG/2021/000292). We are very thankful Dilip Ghosh, ICAR-CCRI, Nagpur, Makesh Kumar, ICAR-CTCRI, Thiruvananthapuram, R. Viswanthan, ICAR-SBI, Coimbatore, India for providing the pictures presented in Figure 2.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. King, A.M.Q.; Adams, M.J.; Lefkowitz, E.J.; Carstens, E.B. Virus Taxonomy: Ninth Report of the International Committee on Taxonomy of Viruses; Academic Press: San Diego, CA, USA, 2012; 1463p. [Google Scholar]
  2. International Committee on Taxonomy of Viruses (ICTV). Virus Taxonomy: 2021 Release. Available online: http://www.ictvonline.org/virusTaxonomy.asp (accessed on 28 November 2022).
  3. Chabannes, M.; Iskra-Caruana, M.-L. Endogenous pararetroviruses—A reservoir of virus infection in plants. Curr. Opin. Virol. 2013, 3, 615–620. [Google Scholar] [CrossRef] [PubMed]
  4. Geering, A.D.W.; Hull, R. Family Caulimoviridae. In Virus Taxonomy: Ninth Report of the International Committee on Taxonomy of Viruses; King, A.M.Q., Adams, M.J., Carstens, E.B., Lefkowitz, E.J., Eds.; Academic Press: Boston, MA, USA, 2012; pp. 429–443. [Google Scholar]
  5. Bhat, A.I.; Hohn, T.; Selvarajan, R. Badnaviruses: The current global scenario. Viruses 2016, 8, 177. [Google Scholar] [CrossRef] [PubMed]
  6. Borah, B.K.; Sharma, S.; Kant, R.; Anthony-Johnson, A.M.; Saigopal, D.V.R.; Dasgupta, I. Bacilliform DNA-containing plant viruses in the tropics: Commonalities within a genetically diverse group. Mol. Plant Pathol. 2013, 14, 759–771. [Google Scholar] [CrossRef] [PubMed]
  7. Hohn, T.; Richert-Poggeler, K.R.; Harper, G.; Schawarzacher, T.; Teo, C.; Teycheney, P.-Y.; Iskra-Caruana, M.-L.; Hull, R. Evolution of integrated plant viruses. In Plant Virus Evolution; Roosinck, M., Ed.; Springer: Berlin/Heidelberg, Germany, 2008; pp. 54–76. [Google Scholar]
  8. Staginnus, C.; Iskra-Caruana, M.; Lockhart, B.; Hohn, T.; Richert-Pöggeler, K.R. Suggestions for a nomenclature of endogenous pararetroviral sequences in plants. Arch. Virol. 2009, 154, 1189–1193. [Google Scholar] [CrossRef] [PubMed]
  9. Dallot, S.; Accuna, P.; Rivera, C.; Ramirez, P.; Cote, F.; Lockhart, B.E.L.; Caruana, M.-L. Evidence that the proliferation stage of micropropagation procedure is determinant in the expression of Banana streak virus integrated into the genome of the FHIA21 hybrid (Musa AAAB). Arch. Virol. 2001, 146, 2179–2190. [Google Scholar] [CrossRef]
  10. Lheureux, F.; Carreel, F.; Jenny, C.; Lockhart, B.E.L.; Iskra-Caruana, M.-L. Identification of genetic markers linked to banana streak disease expression in inter-specific Musa hybrids. Theor. Appl. Genet. 2003, 106, 594–598. [Google Scholar] [CrossRef]
  11. Côte, F.X.; Galzi, S.; Folliot, M.; Lamagnère, Y.; Teycheney, P.-Y.; Iskra-Caruana, M.-L. Micropropagation by tissue culture triggers differential expression of infectious endogenous Banana streak virus sequences (eBSV) present in the B genome of natural and synthetic interspecific banana plantains. Mol. Plant Pathol. 2010, 11, 137–144. [Google Scholar] [CrossRef]
  12. Harper, G.; Osuji, J.O.; Heslop-Harrison, J.S.; Hull, R. Integration of banana streak badnavirus into the Musa genome: Molecular and cytogenetic evidence. Virology 1999, 255, 207–213. [Google Scholar] [CrossRef]
  13. Harper, G.; Hull, R.; Lockhart, B.E.N.; Olsweski, N. Viral sequences integrated in to plant genomes. Ann. Rev. Phytopathol. 2002, 40, 119–136. [Google Scholar] [CrossRef]
  14. Staginnus, C.; Richert-Poggeler, K.R. Endogenous pararetroviruses: Two-faced travelers in the plant genome. Trends Plant Sci. 2006, 11, 485–491. [Google Scholar] [CrossRef]
  15. Mette, M.F.; Kanno, T.; Aufsatz, W.; Jakowitsch, J.; Van der Winden, J.; Matzke, M.A.; Matzke, A.J.M. Endogenous viral sequences and their potential contribution to heritable virus resistance in plants. EMBO J. 2002, 21, 461–469. [Google Scholar] [CrossRef]
  16. Ndowora, T.; Ganesh, D.; Daryl, L.F.; Harper, G.; Hull, R.; Neil, E.O.; Lockhart, B. Evidence that Badnavirus infection in Musa can originate from integrated Pararetroviral sequences. Virology 1999, 225, 214–220. [Google Scholar] [CrossRef] [PubMed]
  17. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed]
  18. Dahal, G.; Hughes, J.D.A.; Thottappilly, G.; Lockhart, B.E.L. Effect of temperature on symptom expression and reliability of banana streak badnavirus detection in naturally infected plantain and banana (Musa spp.). Plant Dis. 1998, 82, 16–21. [Google Scholar] [CrossRef] [PubMed]
  19. Dahal, G.; Ortiz, R.; Tenkouano, A.; Hughes, J.D.A.; Thottappilly, G.; Vuylsteke, D.; Lockhart, B.E.L. Relationship between natural occurrence of Banana streak badnavirus and symptom expression, relative concentration of viral antigen, and yield characteristics of some micropropagated Musa spp. Plant Pathol. 2000, 49, 68–79. [Google Scholar] [CrossRef]
  20. Daniells, J.W.; Geering, A.D.W.; Bryde, N.J.; Thomas, J.E. The effect of Banana streak virus on the growth and yield of dessert bananas in tropical Australia. Ann. Appl. Biol. 2001, 139, 51–60. [Google Scholar] [CrossRef]
  21. Lockhart, B.E.L. Purification and serology of a bacilliform virus associated with banana streak disease. Phytopathology 1986, 76, 995–999. [Google Scholar] [CrossRef]
  22. Geering, A.D.W.; McMichael, L.A.; Dietzgen, R.G.; Thomas, J.E. Genetic diversity among banana streak virus isolates from Australia. Phytopathology 2000, 90, 921–927. [Google Scholar] [CrossRef] [PubMed]
  23. Geering, A.D.W.; Olszewski, N.E.; Harper, G.; Lockhart, B.E.L.; Hull, R.; Thomas, J.E. Banana contains a diverse array of endogenous badnaviruses. J. Gen. Virol. 2005, 86, 511–520. [Google Scholar] [CrossRef]
  24. Harper, G.; Hart, D.; Moult, S.; Hull, R.; Thresh, J.M.; Jones, R.A.C.; Kuehne, T. Banana streak virus is very diverse in Uganda. Virus Res. 2004, 100, 51–56. [Google Scholar] [CrossRef]
  25. Harper, G.; Hart, D.; Moult, S.; Hull, R.; Geering, A.; Thomas, J. The diversity of Banana streak virus isolates in Uganda. Arch. Virol. 2005, 150, 2407–2420. [Google Scholar] [CrossRef]
  26. Geering, A.D.; Parry, J.N.; Thomas, J.E. Complete genome sequence of a novel badnavirus, Banana streak IM virus. Arch. Virol. 2011, 156, 733–737. [Google Scholar] [CrossRef] [PubMed]
  27. Lheureux, F.; Laboureau, N.; Muller, E.; Lockhart, B.E.; Iskra-Caruana, M.L. Molecular characterization of banana streak acuminata Vietnam virus isolated from Musa acuminata siamea (banana cultivar). Arch. Virol. 2007, 152, 1409–1416. [Google Scholar] [CrossRef] [PubMed]
  28. Iskra-Caruana, M.L.; Duroy, P.O.; Chabannes, M.; Muller, E. The common evolutionary history of badnaviruses and banana. Infect. Genet. Evol. 2014, 21, 83–89. [Google Scholar] [CrossRef]
  29. Chabannes, M.; Gabriel, M.; Aksa, A.; Galzi, S.; Dufayard, J.F.; Iskra-Caruana, M.L.; Muller, E. Badnaviruses and banana genomes: A long association sheds light on Musa phylogeny and origin. Mol. Plant Pathol. 2021, 22, 216–230. [Google Scholar] [CrossRef] [PubMed]
  30. Sharma, S.K.; Vignesh Kumar, P.; Geetanjali, A.S.; Pun, K.B.; Baranwal, V.K. Subpopulation level variation of banana streak viruses in India and common evolution of banana and sugarcane badnaviruses. Virus Genes 2015, 50, 450–465. [Google Scholar] [CrossRef]
  31. Delanoy, M.; Salmon, M.; Kummert, J.; Frison, E.; Lepoivre, P. Development of real-time PCR for the rapid detection of episomal Banana streak virus (BSV). Plant Dis. 2003, 87, 33–38. [Google Scholar] [CrossRef]
  32. Le Provost, G.; Iskra-Caruana, M.L.; Acina, I.; Teycheney, P.Y. Improved detection of episomal banana streak viruses by multiplex immunocapture PCR. J. Virol. Methods 2006, 137, 7–13. [Google Scholar] [CrossRef]
  33. James, A.P.; Geijskes, R.J.; Dale, J.L.; Harding, R.M. Development of a novel rolling-circle amplification technique to detect Banana streak virus which also discriminates between integrated and episomal virus sequences. Plant Dis. 2011, 95, 57–62. [Google Scholar] [CrossRef]
  34. Javer-Higginson, E.; Acina-Mambole, I.; González, J.E.; Font, C.; González, G.; Echemendía, A.L.; Muller, E.; Teycheney, P.-Y. Occurrence, prevalence and molecular diversity of banana streak viruses in Cuba. Eur. J. Plant Pathol. 2013, 138, 157–166. [Google Scholar] [CrossRef]
  35. Sharma, S.K.; Vignesh Kumar, P.; Baranwal, V.K. Immunodiagnosis of episomal Banana streak MY virus using polyclonal antibodies to an expressed putative coat protein. J. Virol. Methods 2014, 207, 86–94. [Google Scholar] [CrossRef]
  36. Selvarajan, R.; Sheeba, M.M.; Balasubramanian, V. Simultaneous detection of episomal Banana streak Mysore virus and Banana bunchy top virus using multiplex RT-PCR. Curr. Sci. 2011, 100, 31–34. [Google Scholar]
  37. Liu, F.X.; Feng, L.X.; Chen, X.; Han, Y.C.; Li, W.D.; Xu, W.; Cai, B.; Lin, M.G. Simultaneous detection of four banana viruses by multiplex PCR. J. Phytopathol. 2012, 160, 622–627. [Google Scholar] [CrossRef]
  38. Selvarajan, R.; Balasubramanian, V.; Gayathrie, T. Highly efficient immunodiagnosis of episomal Banana streak MY virus (BSMYV) using polyclonal antibodies raised against recombinant viral associated protein. J. Phytopathol. 2016, 207, 86–94. [Google Scholar]
  39. Zhang, J.; Borth, W.; Lin, B.; Melzer, M.; Shen, H.; Pu, X.; Sun, D.; Nelson, S.; Hu, J. Multiplex detection of three banana viruses by reverse transcription loop-mediated isothermal amplification (RT-LAMP). Trop. Plant Pathol. 2018, 43, 543–551. [Google Scholar] [CrossRef]
  40. Kapoor, R.; Srivastava, N.; Kumar, R.; Sharma, S.K.; Rai, R.; Kumar, S.; Baranwal, V.K. Detection of episomal banana streak Mysore virus by reverse transcription-recombinase polymerase amplification assay. J. Plant Pathol. 2020, 102, 499–503. [Google Scholar] [CrossRef]
  41. Kuan, C.P.; Tsai, C.H.; Tseng, C.S.; Yang, T.C. Development of a bead-based assay for detection of three banana-infecting viruses. Peer J. 2022, 26, e13409. [Google Scholar] [CrossRef]
  42. Helliot, B.; Panis, B.; Poumay, Y.; Swennen, R.; Lepoivre, P.; Frison, E. Cryopreservation for the elimination of cucumber mosaic and banana streak viruses from banana (Musa spp.). Plant Cell Rep. 2002, 20, 1117–1122. [Google Scholar] [CrossRef]
  43. Helliot, B.; Panis, B.; Frison, E.; De Clercq, E.; Swennen, R.; Lepoivre, P.; Neyts, J. The acyclic nucleoside phosphonate analogues, adefovir, tenofovir and PMEDAP, efficiently eliminate banana streak virus from banana (Musa) sp. Antivir. Res. 2003, 59, 121–126. [Google Scholar] [CrossRef]
  44. Tripathi, J.N.; Ntui, V.O.; Ron, M.; Muiruri, S.K.; Britt, A.; Tripathi, L. CRISPR/Cas9 editing of endogenous banana streak virus in the B genome of Musa spp. overcomes a major challenge in banana breeding. Commun. Biol. 2019, 2, 46. [Google Scholar] [CrossRef]
  45. Lockhart, B.E.L.; Kirtisak, K.A.; Jones, P.; Padmini, D.D.; Olszieewski, N.E.; Lockhart, N.; Nuarchan, D.; Sangalang, J. Identification of Piper yellow mottle virus, a mealy bug transmited badnavirus infecting Piper spp. in south East Asia. Eur. J. Plant Pathol. 1997, 103, 303–311. [Google Scholar] [CrossRef]
  46. De Silva, D.P.; Jones, P.P.; Shaw, M.W. Identification and transmission of Piper yellow mottle virus and Cucumber mosaic virus infecting black pepper (Piper nigrum) in Sri Lanka. Plant Pathol. 2002, 51, 537–545. [Google Scholar] [CrossRef] [Green Version]
  47. Bhat, A.I.; Devasahayam, S.; Sarma, Y.R.; Pant, R.P. Association of a Badnavirus in black pepper (Piper nigrum L.) transmitted by mealybug (Ferrisia virgata) in India. Curr. Sci. 2003, 84, 1547–1550. [Google Scholar]
  48. Che, H.; Cao, X.; Liu, P. Occurrence and characterization of virus species associated with black pepper (piper nigrum l.) virus diseases in Hainan province, China. J. Phytopathol. 2020, 169, 247–252. [Google Scholar] [CrossRef]
  49. Bhat, A.I.; Biju, C.N.; Srinivasan, V.; Ankegowda, S.J.; Krishnamurthy, K.S. Current status of viral diseases affecting black pepper and cardamom. J. Spices Aromat. Crops 2018, 27, 1–16. [Google Scholar]
  50. Ahamedemujtaba, V.; Atheena, P.V.; Bhat, A.I.; Krishnamurthy, K.S.; Srinivasan, V. Symptoms of piper yellow mottle virus in black pepper as influenced by temperature and relative humidity. Virus Dis. 2021, 32, 305–313. [Google Scholar] [CrossRef]
  51. Duarte, M.L.R.; Albuquerque, F.C.; Poltranieri, L.S.; Trindade, D.R.; Kitajima, E.W.; Brioso, P.S.T. Mosqueado Amarelo da Pimenta-do-Reino; Embrapa Amazônia Oriental: Belém, Brazil, 2001; p. 30. [Google Scholar]
  52. Siju, S.; Bhat, A.I.; Hareesh, P.S. Identification and characterization of a Badnavirus infecting betel vine and Indian long pepper. J. Plant Biochem. Biotechnol. 2008, 17, 73–76. [Google Scholar] [CrossRef]
  53. Bhat, A.I.; Sasi, S.; Revathy, K.A.; Deeshma, K.P.; Saji, K.V. Sequence diversity among badnavirus isolates infecting black pepper and related species in India. Virus Dis. 2014, 25, 402–407. [Google Scholar] [CrossRef] [PubMed]
  54. Hany, U.; Adams, I.P.; Glover, R.; Bhat, A.I.; Boonham, N. The complete genome sequence of Piper yellow mottle virus (PYMoV). Arch. Virol. 2014, 159, 385–388. [Google Scholar] [CrossRef]
  55. Deeshma, K.P.; Bhat, A.I. Complete genome sequencing of Piper yellow mottle virus infecting black pepper, betelvine and Indian long pepper. Virus Genes 2015, 50, 172–175. [Google Scholar] [CrossRef] [PubMed]
  56. Lim, L.W.T.; Liew, J.X.; Chung, H.H. Piper yellow mottle virus: A deep dive into the genome. Gene Rep. 2022, 29, 101680. [Google Scholar] [CrossRef]
  57. Hareesh, P.S.; Bhat, A.I. Detection and partial nucleotide sequence analysis of Piper yellow mottle virus infecting black pepper in India. Indian J. Virol. 2008, 19, 160–167. [Google Scholar]
  58. Bhat, A.I.; Siljo, A.; Deeshma, K.P. Rapid detection of Piper yellow mottle virus and Cucumber mosaic virus infecting black pepper (Piper nigrum) by loop-mediated isothermal amplification (LAMP). J. Virol. Methods 2013, 193, 190–196. [Google Scholar] [CrossRef] [PubMed]
  59. Bhat, A.I.; Siljo, A. Detection of viruses infecting black pepper by SYBR Green-based real-time PCR assay. J. Plant Pathol. 2014, 96, 105–109. [Google Scholar]
  60. Mohandas, A.; Bhat, A.I. Recombinase polymerase amplification assay for the detection of piper yellow mottle virus infecting black pepper. Virus Dis. 2020, 31, 38–44. [Google Scholar] [CrossRef]
  61. Sasi, S.; Bhat, A.I. In vitro elimination of Piper yellow mottle virus from infected black pepper through somatic embryogenesis and meristem-tip culture. Crop Prot. 2018, 103, 39–45. [Google Scholar] [CrossRef]
  62. Srinivasan, V.; Bhat, A.I.; Dinesh, R.; Ankegowda, S.J.; Hamza, S.; Biju, C.N.; Krishnamurthy, K.S. Rejuvenation of virus affected black pepper plantations through soil and plant health management. Indian J. Arecanut Spices Med. Plants 2017, 19, 24–27. [Google Scholar]
  63. Ahlawat, Y.S.; Pant, R.P.; Lockhart, B.E.L.; Srivastava, M.; Chakraborty, N.K.; Varma, A. Association of a badnavirus with citrus mosaic disease in India. Plant Dis. 1996, 80, 590–592. [Google Scholar] [CrossRef]
  64. Huang, Q.; Hartung, J.S. Cloning and sequence analysis of an infectious clone of Citrus yellow mosaic virus that can infect sweet orange via Agrobacterium-mediated inoculation. J. Gen. Virol. 2001, 82, 2549–2558. [Google Scholar] [CrossRef]
  65. Vadlamudi, T.; Kaldis, A.; Divi, V.S.G.; Patil, B.L.; Voloudakis, A.E. The Citrus yellow mosaic badnavirus ORFI functions as a RNA-silencing suppressor. Virus Genes 2021, 57, 469–473. [Google Scholar] [CrossRef]
  66. Borah, B.K.; Johnson, A.M.A.; Sai-Gopal, D.V.R.; Dasgupta, I. Sequencing and computational analysis of complete genome sequences of Citrus yellow mosaic badnavirus from acid lime and pummelo. Virus Genes 2009, 39, 137–140. [Google Scholar] [CrossRef] [PubMed]
  67. Anthony-Johnson, A.M.; Borah, B.K.; Sai Gopal, D.V.; Dasgupta, I. Analysis of full-length sequences of two Citrus yellow mosaic badnavirus isolates infecting Citrus jambhiri (Rough Lemon) and Citrus sinensis L. Osbeck (Sweet Orange) from a nursery in India. Virus Genes 2012, 45, 600–605. [Google Scholar] [CrossRef]
  68. Geering, A.D.W.; Maumus, F.; Copetti, D.; Choisne, N.; Zwickl, D.J.; Zytnicki, M.; McTaggart, A.R.; Scalabrin, S.; Vezzulli, S.; Wing, R.A.; et al. Endogenous florendoviruses are major components of plant genomes and hallmarks of virus evolution. Nat. Commun. 2014, 5, 5269. [Google Scholar] [CrossRef]
  69. Kumar, P.V.; Sharma, S.K.; Rishi, N.; Ghosh, D.K.; Baranwal, V.K. An isothermal based recombinase polymerase amplification assay for rapid, sensitive, and robust indexing of citrus yellow mosaic virus. Acta Virol. 2018, 62, 104–108. [Google Scholar] [CrossRef]
  70. Ghosh, D.K.; Aglave, B.; Bhanare, K.; Baranwal, V.K. PCR based detection of Citrus yellow mosaic disease from Vidarbha region of Maharastra. Indian Phytopathol. 2007, 60, 520–526. [Google Scholar]
  71. Gopi, V.; Gopal, K.; Gourisankar, T.; Palanivel, S. Detection of citrus yellow mosaic virus by PCR and nucleic acid spot hybridisation using non-radioactive probes in commercial citrus species. Arch. Phytopathol. Plant Prot. 2010, 43, 892–899. [Google Scholar] [CrossRef]
  72. Motghare, M.; Dhar, A.K.; Kokane, A.; Warghane, A.; Kokane, S.; Sharma, A.K.; Reddy, M.K.; Ghosh, D.K. Quantitative Distribution of Citrus yellow mosaic badnavirus in Sweet orange (Citrus sinensis) and its Implication in Developing Disease Diagnostics. J. Virol. Methods 2018, 259, 25–31. [Google Scholar] [CrossRef] [PubMed]
  73. Thresh, M.J. The origin and epidemiology of some important plant virus diseases. Appl. Biol. 1980, 5, 1–65. [Google Scholar]
  74. Ramos-Sobrinho, R.; Kouakou, K.; Bi, A.B.; Keith, C.V.; Diby, L.; Kouame, C.; Aka, R.A.; Marelli, J.-P.; Brown, J.K. Molecular detection of cacao swollen shoot badnavirus species by amplification with four PCR primer pairs, and evidence that Cacao swollen shoot Togo B virus-like isolates are highly prevalent in Côte d’Ivoire. Eur. J. Plant Pathol. 2021, 159, 941–947. [Google Scholar] [CrossRef]
  75. Kandito, A.; Hartono, S.; Trisyono, Y.A.; Somowiyarjo, S. First report of Cacao mild mosaic virus associated with cacao mosaic disease in Indonesia. New Dis. Rep. 2022, 45, e12071. [Google Scholar] [CrossRef]
  76. Muller, E.; Ravel, S.; Agret, C.; Abrokwah, F.; Dzahini-Obiatey, H.; Galyuon, I.; Kouakou, K.; Jeyaseelan, E.C.; Allainguillaume, J.; Wetten, A. Next generation sequencing elucidates cacao badnavirus diversity and reveals the existence of more than ten viral species. Virus Res. 2018, 244, 235–251. [Google Scholar] [CrossRef] [PubMed]
  77. Chingandu, N.; Zia-ur-rehman, M.; Sreenivasan, T.N.; Surujdeo-Maharaj, S.; Umaharan, P.; Gutierrez, O.A.; Brown, J.K. Molecular characterization of previously elusive badnaviruses associated with symptomatic cacao in the New World. Arch. Virol. 2017, 162, 1363–1371. [Google Scholar] [CrossRef] [PubMed]
  78. Ullah, I.; Daymond, A.J.; Hadley, P.; End, M.J.; Umaharan, P.; Dunwell, J.M. Identification of Cacao Mild Mosaic Virus (CaMMV) and Cacao Yellow Vein-Banding Virus (CYVBV) in Cocoa (Theobroma cacao) Germplasm. Viruses 2021, 13, 2152. [Google Scholar] [CrossRef] [PubMed]
  79. Posnette, A.F. Virus diseases of cacao in West Africa VII: Virus transmission by different vector species. Ann. Appl. Biol. 1950, 37, 378–384. [Google Scholar] [CrossRef]
  80. Marelli, J.P.; Guest, D.I.; Bailey, B.A.; Evans, H.C.; Brown, J.K.; Junaid, M.; Barreto, R.W.; Lisboa, D.O.; Puig, A.S. Chocolate under threat from old and new cacao diseases. Phytopathology 2019, 109, 1331–1343. [Google Scholar] [CrossRef]
  81. Posnette, A.F. Virus diseases of cacao in West Africa. 1. Cacao viruses 1A, 1B, 1C and 1D. Ann. Appl. Biol. 1947, 34, 388–402. [Google Scholar] [CrossRef]
  82. Jacquot, E.; Hagen, L.S.; Michler, P.; Rohfritsch, O.; Stussi-Garaud, C.; Keller, M.; Jacquemond, M.; Yot, P. In situ localisation of cacao swollen shoot virus in agroinoculated Theobroma cacao. Arch. Virol. 1999, 144, 259–271. [Google Scholar] [CrossRef]
  83. Roivainen, O. Transmission of cocoa viruses by mealy bugs (Homoptera: Pseudococcidae). J. Sci. Agric. Soc. 1976, 48, 433–453. [Google Scholar]
  84. Quainoo, A.K.; Wetten, A.C.; Allainguillaume, J. Transmission of cocoa swollen shoot virus by seeds. J. Virol. Methods 2008, 150, 45–49. [Google Scholar] [CrossRef]
  85. Ameyaw, G.A.; Wetten, A.; Dzahini-Obiatey, H.; Domfeh, O.; Allainguillaume, J. Investigation on Cacao swollen shoot virus (CSSV) pollen transmission through cross-pollination. Plant Pathol. 2013, 62, 421–427. [Google Scholar] [CrossRef]
  86. Abrokwah, F.K.; Dabie, K.; Adulley, F. Genome variability, species diversity, phylogenetic relationships, origin, and geographical distribution of badnaviruses involved in the cacao swollen shoot disease: The case of West Africa. Trop. Plant Pathol. 2022, 47, 201–213. [Google Scholar] [CrossRef]
  87. Hagen, L.S.; Jacquemond, M.; Lepingle, A.; Lot, H.; Tepfer, M. Nucleotide sequence and genomic organization of Cacao swollen shoot virus. Virology 1993, 196, 619–628. [Google Scholar] [CrossRef]
  88. Jacquot, E.; Hagen, L.S.; Jacquemond, M.; Yot, P. The open reading frame 2 product of cacao swollen shoot badnavirus is a nucleic acid-binding protein. Virology 1996, 225, 191–195. [Google Scholar] [CrossRef]
  89. Chingandu, N.; Kouakou, K.; Aka, R.; Ameyaw, G.; Gutierrez, O.A.; Herrmann, H.; Brown, J.K. The proposed new species, Cacao red vein virus, and three previously recognized badnavirus species are associated with cacao swollen shoot disease. Virol. J. 2017, 14, 199. [Google Scholar] [CrossRef]
  90. Chingandu, N.; Kouakou, K.; Aka, R.; Gutierrez, O.A.; Brown, J.K. Unexpected genome variability at multiple loci suggests Cacao swollen shoot virus comprises multiple, divergent molecular variants. J. Emerg. Dis. Virol. 2017, 3, 2473-1846. [Google Scholar]
  91. Chingandu, N.; Dongo, L.; Gutierrez, O.A.; Brown, J.K. The previously unidentified, divergent badnavirus species Cacao red vein-banding virus is associated with cacao swollen shoot disease in Nigeria. Plant Dis. 2019, 103, 1302–1308. [Google Scholar] [CrossRef]
  92. Ramos-Sobrinho, R.; Chingandu, N.; Gutierrez, O.A.; Marelli, J.-P.; Brown, J.K. A Complex of Badnavirus Species Infecting Cacao Reveals Mixed Infections, Extensive Genomic Variability, and Interspecific Recombination. Viruses 2020, 12, 443. [Google Scholar] [CrossRef] [PubMed]
  93. Muller, E.; Ullah, I.; Dunwell, J.M.; Richardson, M.; Allainguillaume, J.; Wetten, A. Identification and distribution of novel badnaviral sequences integrated in the genome of cacao (Theobroma cacao). Sci. Rep. 2021, 11, 8270. [Google Scholar] [CrossRef] [PubMed]
  94. Ramos-Sobrinho, R.; Ferro, M.M.M.; Nagata, T.; Puig, A.S.; Keith, C.V.; Britto, D.S.; Gutierrez, O.A.; Mareli, J.P.; Brown, J.K. Complete genome sequences of three newly discovered cacao mild mosaic virus isolates from Theobroma cacao L. in Brazil and Puerto Rico and evidence for recombination. Arch. Virol. 2021, 166, 2027–2031. [Google Scholar] [CrossRef] [PubMed]
  95. Ameyaw, G.A.; Domfeh, O.; Armooh, B.; Boakye, A.Y.; Arjarquah, A. Inconsistent PCR detection of Cacao swollen shoot virus (CSSV) is linked to the occurrence of different variants across the cocoa regions of Ghana. J. Virol. Methods 2022, 300, 114400. [Google Scholar] [CrossRef] [PubMed]
  96. Domfeh, O.; Ameyaw, G.A.; Gyamera, E.A. Mild strain cross-protection in cacao: Evaluation of symptom suppression in three progenies exposed to natural spread of the New Juabeng strain of the Cacao swollen shoot Togo B virus (CSSTBV). Trop. Plant Pathol. 2022, 47, 562–568. [Google Scholar] [CrossRef]
  97. Andres, C.; Gattinger, A.; Dzahini-Obiatey, H.K.; Blaser, W.J.; Offei, S.K.; Six, J. Combatting Cocoa Swollen Shoot Virus Disease: What do we know? Crop Prot. 2017, 98, 76–84. [Google Scholar] [CrossRef]
  98. Quainoo, A.K.; Wetten, A.C.; Allainguillaume, J. The effectiveness of somatic embryogenesis in eliminating the cocoa swollen shoot virus from infected cocoa trees. J. Virol. Methods 2008, 149, 91–96. [Google Scholar] [CrossRef] [PubMed]
  99. Turaki, A.A.; Bömer, M.; Silva, G.; Lava Kumar, P.; Seal, S.E. PCR-DGGE analysis: Unravelling complex mixtures of badnavirus sequences present in yam germplasm. Viruses 2017, 9, 181. [Google Scholar] [CrossRef]
  100. Bömer, M.; Rathnayake, A.I.; Visendi, P.; Silva, G.; Seal, S.E. Complete genome sequence of a new member of the genus Badnavirus, Dioscorea bacilliform RT virus 3, reveals the first evidence of recombination in yam badnaviruses. Arch. Virol. 2018, 163, 533–538. [Google Scholar] [CrossRef]
  101. Bömer, M.; Turaki, A.; Silva, G.; Kumar, P.; Seal, S. A sequence-independent strategy for amplification and characterization of episomal badnavirus sequences reveals three previously uncharacterized yam badnaviruses. Viruses 2016, 8, 188. [Google Scholar] [CrossRef]
  102. Umber, M.; Gomez, R.; Gelabale, S.; Bonheur, L.; Pavis, C.; Teycheney, P. The genome sequence of Dioscorea bacilliform TR virus, a member of the genus Badnavirus infecting Dioscorea spp., sheds light on the possible function of endogenous Dioscorea bacilliform viruses. Arch. Virol. 2016, 162, 517–521. [Google Scholar] [CrossRef]
  103. Sukal, A.; Kidanemariam, D.; Dale, J.; James, A.; Harding, R. Characterization of badnaviruses infecting Dioscorea spp. in the Pacific reveals two putative novel species and the first report of Dioscorea bacilliform RT virus 2. Virus Res. 2017, 238, 29–34. [Google Scholar] [CrossRef]
  104. Diouf, M.B.; Festus, R.; Silva, G.; Guyader, S.; Umber, M.; Seal, S.; Teycheney, P.Y. Viruses of Yams (Dioscorea spp.): Current Gaps in Knowledge and Future Research Directions to Improve Disease Management. Viruses 2022, 14, 1884. [Google Scholar] [CrossRef]
  105. Umber, M.; Filloux, D.; Muller, E.; Laboureau, N.; Galzi, S.; Roumagnac, P.; Iskra-Caruana, M.-L.; Pavis, C.; Teycheney, P.-Y.; Seal, S. The genome of African yam (Dioscorea cayenensis-rotundata complex) hosts endogenous sequences from four distinct badnavirus species. Mol. Plant Pathol. 2014, 15, 790–801. [Google Scholar] [CrossRef]
  106. Filloux, D.; Fernandez, E.; Loire, E.; Claude, L.; Galzi, S.; Candresse, T.; Winter, S.; Jeeva, M.L.; Makeshkumar, T.; Martin. D.P.; et al. Nanopore-Based Detection and Characterization of Yam Viruses. Sci. Rep. 2018, 148, 17879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Zhang, Y.; Singh, K.; Kaur, R.; Qiu, W. Association of a novel DNA virus with the grapevine vein-clearing and vine decline syndrome. Phytopathology 2011, 101, 1081–1090. [Google Scholar] [CrossRef]
  108. Cieniewicz, E.J.; Qiu, W.; Saldarelli, P.; Fuchs, M. Believing is seeing: Lessons from emerging viruses in grapevine. J. Plant Pathol. 2020, 102, 619–632. [Google Scholar] [CrossRef]
  109. Rumbos, I.C.; Avgelis, A.D. Roditis leaf discoloration—A new virus disease of grapevine: Symptomatology and transmission to indicators plants. Phytopathol. Z. 1989, 152, 274–278. [Google Scholar] [CrossRef]
  110. Maliogka, V.I.; Olmos, A.; Pappi, P.G.; Lotos, L.; Efthimiou, K.; Grammatikaki, G.; Candresse, T.; Katis, N.I.; Avgeli, A.D. A novel grapevine badnavirus is associated with the Roditis leaf discoloration disease. Virus Res. 2015, 203, 47–55. [Google Scholar] [CrossRef]
  111. Chiumenti, M.; Morelli, M.; Giampetruzzi, A.; Palmisano, F.; Savino, V.N.; La Notte, P.; Martelli, G.P.; Saldarelli, P. First report of grapevine Roditis leaf discoloration-associated virus in Italy. J. Plant Pathol. 2015, 97, 551. [Google Scholar]
  112. Uluba, S.; Serçe, Ç.; Altan, B.; Bolat, V.; Ayyaz, M.; Çifçi, O.; Önder, S.; Öztürk Gökçe, Z.N.; Maliogka, V.I. First Report of grapevine Roditis leaf discoloration-associated virus infecting grapevine (Vitis vinifera) in Turkey. Plant Dis. 2018, 2, 256. [Google Scholar]
  113. Bester, R.; Lotos, L.; Vermeulen, A.; Pietersen, G.; Maliogka, V.I.; Maree, H.J. Complete genome sequence of a grapevine Roditis leaf discoloration-associated virus (GRLDaV) variant from South Africa. Arch. Virol. 2021, 166, 2041–2044. [Google Scholar] [CrossRef]
  114. Ekemen, M. Investigation of Etiology on Grapevine Roditis Leaf Discoloration-Associated Virus. Master’s Thesis, Nigde Omer Halisdemir University, Niğde, Turkey, 2021. Available online: http://acikerisim.ohu.edu.tr/xmlui/handle/11480/8548 (accessed on 12 January 2023).
  115. Jagunic, M.; De Stradis, A.; Preiner, D.; La Notte, P.; Al Rwahnih, M.; Almeida, R.P.P.; Voncina, D. Biology and ultrastructural characterization of Grapevine Badnavirus 1 and Grapevine Virus G. Viruses 2022, 14, 2695. [Google Scholar] [CrossRef] [PubMed]
  116. Vončina, D.; Almeida, R.P.P. Screening of some Croatian autochthonous grapevine varieties reveals a multitude of viruses, including novel ones. Arch. Virol. 2018, 163, 2239–2243. [Google Scholar] [CrossRef]
  117. Wakman, W.D.; Teakle, D.S.; Thomas, J.E.; Dietzgen, R.G. Presence of clostero-like virus and a bacilliform virus in pineapple plants in Queensland. Aust. J. Agric. Res. 1995, 46, 947–958. [Google Scholar] [CrossRef]
  118. Thomson, K.G.; Dietzgen, R.G.; Thomas, J.E.; Teakle, D.S. Detection of pineapple bacilliform virus using the polymerase chain reaction. Ann. Appl. Biol. 1996, 129, 57–69. [Google Scholar] [CrossRef]
  119. Gambley, C.F.; Geering, A.D.; Steele, V.; Thomas, J.E. Identification of viral and non-viral reverse transcribing elements in pineapple (Ananas comosus), including members of two new badnavirus species. Arch. Virol. 2008, 153, 1599–1604. [Google Scholar] [CrossRef] [PubMed]
  120. Sether, D.M.; Melzer, M.J.; Borth, W.B.; Hu, J.S. Pineapple bacilliform CO virus: Diversity, detection, distribution, and transmission. Plant Dis. 2012, 96, 1798–1804. [Google Scholar] [CrossRef]
  121. Liting, W.; Xiaolei, R.; Wenjin, S. Sequencing and analysis of the complete genomic sequence of pineapple bacilliform comosus virus. China Agric. Sci. 2010, 43, 1969–1976. [Google Scholar]
  122. Autrey, L.J.C.; Boolell, S.; Jones, P. Distribution of sugarcane bacilliform virus in various geographical regions. In Proceedings of the XXI Congress of the International Society of Sugar Cane Technologists, Bangkok, Thailand, 5–14 March 1992; Kasetsart University: Bangkok, Thailand, 1995. [Google Scholar]
  123. Rao, G.P.; Sharma, S.K.; Kumar, P.V. Sugarcane Bacilliform Viruses: Present Status. In Plant Viruses; CRC Press: Boca Raton, FL, USA, 2018; pp. 117–130. [Google Scholar]
  124. Ahmad, K.; Sun, S.R.; Chen, J.L.; Huang, M.T.; Fu, H.Y.; Gao, S.J. Presence of diverse sugarcane bacilliform viruses infecting sugarcane in China revealed by pairwise sequence comparisons and phylogenetic analysis. Plant Pathol. J. 2019, 35, 41–50. [Google Scholar] [CrossRef]
  125. Lockhart, B.E.L.; Autrey, L.J.C. Occurrence in sugarcane of a bacilliform virus related serologically to banana streak virus. Plant Dis. 1988, 72, 230–233. [Google Scholar] [CrossRef]
  126. Lockhart, B.E.L.; Irey, M.J.; Comstock, J.C. Sugarcane bacilliform virus, Sugarcane mild mosaic virus, and sugarcane yellow leaf syndrome. In Sugarcane Germplasm Conservation and Exchange; Croft, B.J., Piggin, C.T., Wallis, E.S., Hogarth, D.M., Eds.; Australian Centre for International Agricultural Research (ACIAR): Brisbane, Australia, 1996; pp. 108–112. [Google Scholar]
  127. Bouhida, M.; Lockhart, B.E.; Olszewski, N.E. An analysis of the complete sequence of a sugarcane bacilliform virus genome infectious to banana and rice. J. Gen. Virol. 1993, 74, 15–22. [Google Scholar] [CrossRef] [PubMed]
  128. Balan, S.; Nithya, K.; Cherian, K.A.; Viswanathan, R. True seed transmission of Sugarcane bacilliform virus (SCBV) in sugarcane. Sugar Tech 2022, 24, 513–521. [Google Scholar] [CrossRef]
  129. Braithwaite, K.S.; Egeskov, N.M.; Smith, G.R. Detection of Sugarcane bacilliform virus using the polymerase chain reaction. Plant Dis. 1995, 79, 792–796. [Google Scholar] [CrossRef]
  130. Geijskes, R.J.; Braithwaite, K.S.; Dale, J.L. Sequence analysis of an Australian isolate of sugarcane bacilliform badnavirus. Arch. Virol. 2002, 147, 2393–2404. [Google Scholar] [CrossRef]
  131. Muller, E.; Dupuy, V.; Blondin, L.; Bauffe, F.; Daugrois, J.H.; Nathalie, L.; Iskra-Caruana, M.L. High molecular variability of sugarcane bacilliform viruses in Guadeloupe implying the existence of at least three new species. Virus Res. 2011, 160, 414–419. [Google Scholar] [CrossRef] [PubMed]
  132. Karuppaiah, R.; Viswanathan, R.; Kumar, V.G. Genetic diversity of Sugarcane bacilliform virus isolates infecting Saccharum spp. in India. Virus Genes 2013, 46, 505–516. [Google Scholar] [CrossRef] [PubMed]
  133. Sun, S.R.; Damaj, M.B.; Alabi, O.J.; Wu, X.B.; Erik Mirkov, T.; Fu, H.Y.; Chen, R.K.; Gao, S.J. Molecular characterization of two divergent variants of sugarcane bacilliform viruses infecting sugarcane in China. Eur. J. Plant Pathol. 2016, 145, 375–384. [Google Scholar] [CrossRef]
  134. Janiga, P.K.; Nithya, K.; Viswanathan, R. Dynamics of Genetic Diversity Among Indian Sugarcane Bacilliform Virus Species and Implications of Associated Recombination Events in the Virus. Sugar Tech 2022. [Google Scholar] [CrossRef]
  135. Haregu, S.; Kidanemariam, D.; Abraham, A. Molecular identification and characterization of badnaviruses infecting sugarcane in Ethiopia. Acta Virol. 2022, 66, 3–10. [Google Scholar] [CrossRef]
  136. Rao, G.P.; Sharma, S.K.; Singh, D.; Arya, M.; Singh, P.; Baranwal, V.K. Genetically diverse variants of Sugarcane bacilliform virus infecting sugarcane in India and evidence of a novel recombinant Badnavirus variant. J. Phytopathol. 2014, 162, 779–787. [Google Scholar] [CrossRef]
  137. Geijskes, R.J.; Braithwaite, K.S.; Smith, G.R.; Dale, J.L.; Harding, R.M. Sugarcane bacilliform virus encapsidates genome concatamers and does not appear to integrate into the Saccharum officinarum genome. Arch. Virol. 2004, 149, 791–798. [Google Scholar] [CrossRef]
  138. Cai, Y.; Xu, D.; Zhou, G. Evidence of sugarcane bacilliform virus DNA fragment integrated into the saccharum inter-specific hybrids genome. J. South China Agric. Univ. 2009, 30, 19–23. [Google Scholar]
  139. Schenk, P.M.; Sagi, L.; Remans, T.; Dietzgen, R.G.; Bernard, M.J.; Graham, M.W.; Manners, J.M. A promoter from sugarcane bacilliform badnavirus drives transgene expression in banana and other monocot and dicot plants. Plant Mol. Biol. 1999, 39, 1221–1230. [Google Scholar] [CrossRef]
  140. Braithwaite, K.S.; Geijskes, R.J.; Smith, G.R. A variable region of the sugarcane bacilliform virus (SCBV) genome can be used to generate promoters for transgene expression in sugarcane. Plant Cell Rep. 2004, 23, 319–326. [Google Scholar] [CrossRef] [PubMed]
  141. Ashraf, M.A.; Feng, X.; Hu, X.; Ashraf, F.; Shen, L.; Iqbal, M.S.; Ashraf, F.; Shen, L.; Iqbal, M.S.; Zhang, S. In silico identification of sugarcane (Saccharum officinarum L.) genome encoded microRNAs targeting sugarcane bacilliform virus. PLoS ONE 2022, 17, e0261807. [Google Scholar] [CrossRef] [PubMed]
  142. Kreuze, J.F.; Perez, A.; Untiveros, M.; Quispe, D.; Fuentes, S.; Barker, I.; Simon, R. 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. Virology 2009, 388, 1–7. [Google Scholar] [CrossRef] [PubMed]
  143. Kashif, M.; Pietilä, S.; Artola, K.; Jones, R.A.C.; Tugume, A.K.; Mäkinen, V.; Valkonen, J.P.T. Detection of viruses in sweetpotato from Honduras and Guatemala augmented by deep-sequencing of small-RNAs. Plant Dis. 2012, 96, 1430–1437. [Google Scholar] [CrossRef]
  144. Mbanzibwa, D.R.; Tugume, A.K.; Chiunga, E.; Mark, D.; Tairo, F.D. Small RNA deep sequencing-based detection and further evidence of DNA viruses infecting sweetpotato plants in Tanzania. Ann. Appl. Biol. 2014, 165, 329–339. [Google Scholar] [CrossRef]
  145. Buko, D.; Spetz, C.; Hvoslef-Eide, T. Next generation sequencing as a method to verify virus elimination using heat treatment and meristem tip culture in the five most widely used sweet potato varieties in Ethiopia. Afr. J. Biotechnol. 2020, 19, 458–463. [Google Scholar]
  146. Kreuze, J.F.; Perez, A.; Gargurevich, M.G.; Cuellar, W.J. Badnaviruses of sweet potato: Symptomless coinhabitants on a global scale. Front Plant Sci. 2020, 11, 313. [Google Scholar] [CrossRef] [PubMed]
  147. Devitt, L.; Ebenebe, A.; Gregory, H.; Harding, R.; Hunter, D.; Macanawai, A. Investigations into the seed and mealybug transmission of Taro bacilliform virus. Aust. Plant Pathol. 2005, 34, 73–76. [Google Scholar]
  148. Yang, I.C.; Hafner, G.J.; Dale, J.L.; Harding, R.M. Genomic characterisation of taro bacilliform virus. Arch. Virol. 2003, 148, 937–949. [Google Scholar] [CrossRef]
  149. Yang, I.C.; Hafner, G.J.; Revill, P.A.; Dale, J.L.; Harding, R.M. Sequence diversity of South Pacific isolates of Taro bacilliform virus and the development of a PCR-based diagnostic test. Arch. Virol. 2003, 148, 1957–1966. [Google Scholar] [CrossRef]
  150. Kazmi, S.A.; Yang, Z.; Hong, N.; Wang, G.; Wang, Y. Characterization by small RNA sequencing of Taro Bacilliform CH Virus (TaBCHV), a novel Badnavirus. PLoS ONE 2015, 10, e0134147. [Google Scholar] [CrossRef] [PubMed]
  151. Yang, I.C.; Iommarini, J.P.; Becker, D.K.; Hafner, G.J.; Dale, J.L.; Harding, R.M. A promoter derived from taro bacilliform badnavirus drives strong expression in transgenic banana and tobacco plants. Plant Cell Rep. 2003, 21, 1199–1206. [Google Scholar] [CrossRef]
  152. Li, Y.; Xu, C.; Chen, J. Establishment of virus-free Taro (Colocasia esulenta cv. Fenghuayunaitou) by meristem-tip culture combined with thermotherapy. Pak. J. Plant Pathol. 2002, 1, 40–43. [Google Scholar] [CrossRef]
  153. Zhao, J.-P.; Jiang, X.-M.; Bai, X.-F.; Zhang, P.; Yong, J.-P. Several Physiological Factors Influencing Virus-free and Rapid Propagation in Shoot-tip Culture of Colocasia esculenta L. Schott. Plant Phisiol. J. 2009, 45, 133–136. [Google Scholar]
  154. Stace-Smith, R.; Jones, A.T. Virus Diseases of Small Fruits; Converse, R.H., Ed.; United States Department of Agriculture: Washington, DC, USA, 1987; Section 4; p. 175. [Google Scholar]
  155. Marais, A.; Murolo, S.; Faure, C.; Brans, Y.; Larue, C.; Maclot, F.; Massart, S.; Chiumenti, M.; Minafra, A.; Romanazzi, G.; et al. Sixty Years from the First Disease Description, a Novel Badnavirus Associated with Chestnut Mosaic Disease. Phytopathology 2021, 111, 1051–1058. [Google Scholar] [CrossRef] [PubMed]
  156. Rumbou, A.; Candresse, T.; Marais, A.; Theil, S.; Langer, J.; Jalkanen, R.; Büttner, C. A novel badnavirus discovered from Betula sp. affected by birch leaf-roll disease. PLoS ONE 2018, 13, e0193888. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, Y.; Cheng, X.; Wu, X.; Wang, A.; Wu, X. Characterization of complete genome and small RNA profile of Pagoda yellow mosaic associated virus, a novel Badnavirus in China. Virus Res. 2014, 188, 103–108. [Google Scholar] [CrossRef]
  158. Chiumenti, M.; Morelli, M.; Elbeaino, T.; Stavolone, L.; De Stradis, A.; Digiaro, M.; Minafra, A.; Martelli, G.P. Sequencing an unconventional virus genome: The mulberry badnavirus 1 case. J. Plant Pathol. 2014, 96, 36–37. [Google Scholar]
  159. Alvarez-Quinto, R.A.; Lockhart, B.E.L.; Moreno-Martinez, J.M.; Olszewski, N.E. Complete genome sequence of aglaonema bacilliform virus (ABV). Arch. Virol. 2020, 165, 237–239. [Google Scholar] [CrossRef]
  160. Uke, A.; Pinili, M.S.; Natsuaki, K.T.; Geering, A.D.W. Complete genome sequence of aucuba ringspot virus. Arch. Virol. 2021, 166, 1227–1230. [Google Scholar] [CrossRef]
  161. Tsai, C.H.; Su, H.J.; Wu, M.-L.; Feng, Y.C.; Hung, T.H. Identification and detection of Bougainvillea spectabilis chlorotic vein-banding virus in different bougainvillea cultivars in Taiwan. Ann. Appl. Biol. 2008, 53, 187–193. [Google Scholar] [CrossRef]
  162. Baranwal, V.K.; Meenakshi, A.; Singh, J. First report of two distinct badnaviruses associated with Bougainvillea spectabilis in India. J. Gen. Plant Pathol. 2010, 76, 236–239. [Google Scholar] [CrossRef]
  163. Yusop, M.S.M.; Ersoy, R.; Akbar, M.A.; Saad, M.F.M.; Goh, H.H.; Baharum, S.N.; Bunawan, H. First report of Bougainvillea chlorotic vein-banding virus causing Chlorosis and vein banding of Bougainvillea spectabilis in Malaysia. Plant Dis. 2019, 103, 2974. [Google Scholar] [CrossRef]
  164. Naito, F.Y.B.; De Nazaré Almeida dos Reis, L.; Batista, J.G.; Nery, F.M.B.; Rossato, M.; Melo, F.L.; De Cássia Pereira-Carvalho, R. Complete genome sequence of Bougainvillea chlorotic vein banding virus in Bougainvillea spectabilis from Brazil. Trop. Plant Pathol. 2020, 45, 159–162. [Google Scholar] [CrossRef]
  165. Li, R.; Zheng, L.; Cao, M.; Wu, L.; Normandy, P.; Liu, H. First identification and molecular characterization of a new badnavirus infecting camellia. Arch. Virol. 2020, 165, 2115–2118. [Google Scholar] [CrossRef] [PubMed]
  166. Wijayasekara, D.; Hoyt, P.; Gimondo, A.; Dunn, B.; Thapa, A.; Jones, H.; Verchot, J. Molecular characterization of two badnavirus genomes associated with Canna yellow mottle disease. Virus Res. 2018, 2, 19–24. [Google Scholar] [CrossRef] [PubMed]
  167. Zhang, J.; Dey, K.K.; Lin, B.; Borth, W.B.; Melzer, M.J.; Sether, D.; Wang, Y.; Wang, I.C.; Shen, H.; Pu, X.; et al. Characterization of Canna yellow mottle virus in a New Host, Alpinia purpurata, in Hawaii. Phytopathology 2017, 107, 791–799. [Google Scholar] [CrossRef]
  168. Lim, S.; Park, J.M.; Kwon, S.Y.; Cho, H.S.; Kim, H.S.; Lee, S.H.; Moon, J.S. Complete genome sequence of a tentative new member of the genus Badnavirus identified in Codonopsis lanceolata. Arch. Virol. 2019, 164, 1733–1737. [Google Scholar] [CrossRef]
  169. Su, L.; Gao, S.; Huang, Y.; Ji, C.; Wang, D.; Ma, Y.; Fang, R.; Chen, X. Complete genomic sequence of Dracaena mottle virus, a distinct badnavirus. Virus Genes 2007, 35, 423–430. [Google Scholar] [CrossRef]
  170. Lan, P.X.; Tian, T.Y.; Pu, L.L.; Rao, W.L.; Li, F.; Li, R.H. Characterization and detection of a new badnavirus infecting Epiphyllum spp. Arch. Virol. 2019, 164, 1837–1841. [Google Scholar] [CrossRef]
  171. Bester, R.; Burger, J.T.; Maree, H.J. Genomic characterisation of a newly identified badnavirus infecting ivy (Hedera helix). Arch. Virol. 2020, 165, 1511–1514. [Google Scholar] [CrossRef] [PubMed]
  172. Yang, Z.; Nicotaisen, M.; Olszewski, N.E.; Lockhart, B.E. Sequencing, improved detection, and a novel form of Kalanchoë top-spotting virus. Plant Dis. 2005, 89, 298–302. [Google Scholar] [CrossRef]
  173. Diaz-Lara, A.; Mosier, N.J.; Keller, K.E.; Martin, R.R. A variant of Rubus yellow net virus with altered genomic organization. Virus Genes 2015, 50, 104–110. [Google Scholar] [CrossRef] [PubMed]
  174. Alvarez-Quinto, R.A.; Serrano, J.; Olszewski, N.; Grinstead, S.; Mollov, D.; Lockhart, B.E.L. Complete genome sequences of two isolates of spiraea yellow leafspot virus (genus Badnavirus) from Spiraea x bumalda ‘Anthony Waterer’. Arch. Virol. 2022, 167, 631–634. [Google Scholar] [CrossRef] [PubMed]
  175. Li, Y.; Deng, C.; Qiao, Y.; Zhao, X.; Zhou, Q. Characterization of a new badnavirus from Wisteria sinensis. Arch. Virol. 2017, 162, 2125–2129. [Google Scholar] [CrossRef]
  176. Mehrvar, M.; Moradi, Z.; Al-Jaberi, M.S. First report of Wisteria badnavirus 1 infecting Wisteria sinensisin Iran. New Dis. Rep. 2022, 46, e12133. [Google Scholar] [CrossRef]
  177. Alvarez-Quinto, R.A.; Lockhart, B.E.L.; Fetzer, J.L.; Olszewski, N.E. Genomic characterization of cycad leaf necrosis virus, the first badnavirus identified in a gymnosperm. Arch Virol. 2020, 165, 1671–1673. [Google Scholar] [CrossRef]
  178. Shahid, M.S.; Aboughanem-Sabanadzovic, N.; Sabanadzovic, S.; Tzanetakis, I.E. Genomic Characterization and Population Structure of a Badnavirus Infecting Blackberry. Plant Dis. 2017, 101, 110–115. [Google Scholar] [CrossRef]
  179. Laney, A.G.; Hassan, M.; Tzanetakis, I.E. An integrated badnavirus is prevalent in fig germplasm. Phytopathology 2012, 102, 1182–1189. [Google Scholar] [CrossRef] [Green Version]
  180. Alishiri, A.; Rakhshandehroo, F.; Jouzani, G.S.; Shams-bakhsh, M. Exploring the genetic diversity and molecular evolution of fig badnavirus-1 from Iran. J. Plant Pathol. 2018, 100, 287–299. [Google Scholar] [CrossRef]
  181. Xu, D.; Mock, R.; Kinard, G.; Li, R. Molecular analysis of the complete genomic sequences of four isolates of Gooseberry vein banding associated virus. Virus Genes 2011, 43, 130–137. [Google Scholar] [CrossRef]
  182. Petrzik, K.; Přibylová, J.; Spak, J. Molecular analysis of gooseberry vein banding associated virus. Acta Virol. 2012, 56, 119–124. [Google Scholar] [CrossRef] [PubMed]
  183. Zuļģe, N.; Stalažs, A.; Moročko-Bičevska, I.; Namniece, S.; Drevinska, K.; Konavko, D. Gooseberry vein banding associated virus on Ribes in Latvia: Occurrence, host plants and vectors. Plant Pathol. 2022, 71, 1910–1920. [Google Scholar] [CrossRef]
  184. Du, K.; Liu, S.; Chen, Z.; Fan, Z.; Wang, H.; Tian, G.; Zhou, T. Full genome sequence of jujube mosaic-associated virus, a new member of the family Caulimoviridae. Arch. Virol. 2017, 162, 3221–3224. [Google Scholar] [CrossRef]
  185. Liu, B.; Zhang, G.; Song, D.; Wang, Q.; Li, H.; Gu, A.; Bai, J. Complete genome sequence of a novel virus belonging to the genus Badnavirus in jujube (Ziziphus jujuba Mill.) in China. Arch. Virol. 2022, 167, 1885–1888. [Google Scholar] [CrossRef] [PubMed]
  186. Alvarez-Quinto, R.A.; Lockhart, B.E.L.; Olszewski, N. Complete genome sequence of a previously undescribed badnavirus occurring in Polyscias fruticosa L. (Ming aralia). Arch Virol. 2019, 164, 2371–2374. [Google Scholar] [CrossRef]
  187. Xu, M.; Zhang, S.; Xuan, Z.; Wu, J.; Dong, P.; Zhou, Y.; Li, R.; Cao, M. Molecular characterization of a new badnavirus infecting green Sichuan pepper (Zanthoxylum schinifolium). Arch. Virol. 2019, 164, 2613–2616. [Google Scholar] [CrossRef] [PubMed]
  188. Lee, Y.J.; Kwak, H.R.; Lee, Y.K.; Kim, M.K.; Choi, H.S.; Seo, J.K. Complete genome sequence of yacon necrotic mottle virus, a novel putative member of the genus Badnavirus. Arch. Virol. 2015, 160, 1139–1142. [Google Scholar] [CrossRef]
  189. Medberry, S.L.; Lockhart, B.E.L.; Olszewski, N.E. The Commelinayellow m ottlevirus promoter is a strong promoter in vascular and reproductive tissues. Plant Cell 1992, 4, 185–192. [Google Scholar]
  190. Martin, D.P.; Murrell, B.; Golden, M.; Khoosal, A.; Muhire, B. RDP4: Detection and analysis of recombination patterns in virus genomes. Virus Evol. 2015, 1, vev003. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Phylogenetic tree constructed using full-length genome sequences of badnavirus approved by the ICTV. The evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model conducted in MEGA7 [17]. Bootstrap values for 1000 replicates are given when above 50%. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The GenBank accession numbers and acronyms of viruses are shown while the details of acronyms used are provided in Table 1.
Figure 1. Phylogenetic tree constructed using full-length genome sequences of badnavirus approved by the ICTV. The evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model conducted in MEGA7 [17]. Bootstrap values for 1000 replicates are given when above 50%. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The GenBank accession numbers and acronyms of viruses are shown while the details of acronyms used are provided in Table 1.
Pathogens 12 00245 g001
Figure 2. Symptoms caused by badnaviruses on different hosts: (A) banana streak virus on banana, (B) piper yellow mottle virus on black pepper, (C) citrus yellow mosaic virus on citrus (source: Dilip Ghosh, ICAR-Central Citrus Research Institute, Nagpur, India), (D) Dioscorea bacilliform virus on yam (source: Makeshkumar, ICAR-Central Tuber Crops Research Institute, Thiruvananthapuram, India), (E) sugarcane bacilliform virus on sugarcane (source: R. Viswanathan, ICAR-Indian Institute of Sugarcane Research, Lucknow, India), and (F) taro bacilliform virus on taro (source: T. Makeshkumar, ICAR-Central Tuber Crops Research Institute, Thiruvananthapuram, India).
Figure 2. Symptoms caused by badnaviruses on different hosts: (A) banana streak virus on banana, (B) piper yellow mottle virus on black pepper, (C) citrus yellow mosaic virus on citrus (source: Dilip Ghosh, ICAR-Central Citrus Research Institute, Nagpur, India), (D) Dioscorea bacilliform virus on yam (source: Makeshkumar, ICAR-Central Tuber Crops Research Institute, Thiruvananthapuram, India), (E) sugarcane bacilliform virus on sugarcane (source: R. Viswanathan, ICAR-Indian Institute of Sugarcane Research, Lucknow, India), and (F) taro bacilliform virus on taro (source: T. Makeshkumar, ICAR-Central Tuber Crops Research Institute, Thiruvananthapuram, India).
Pathogens 12 00245 g002
Figure 3. Phylogenetic tree constructed based on the amino acid sequences of open reading frame III from isolates belonging to different BSV species available in the GenBank. The evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model conducted in MEGA7 [17]. Bootstrap values for 1000 replicates are given when above 50%. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The GenBank accession numbers and acronyms of viruses are shown while the details of acronyms used are provided in Table 1. The tentative species are highlighted in green.
Figure 3. Phylogenetic tree constructed based on the amino acid sequences of open reading frame III from isolates belonging to different BSV species available in the GenBank. The evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model conducted in MEGA7 [17]. Bootstrap values for 1000 replicates are given when above 50%. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The GenBank accession numbers and acronyms of viruses are shown while the details of acronyms used are provided in Table 1. The tentative species are highlighted in green.
Pathogens 12 00245 g003
Figure 4. The phylogenetic tree was constructed based on the amino acid sequences of open reading frame III from isolates belonging to different dioscorea bacilliform virus species available in the GenBank. The evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model conducted in MEGA7 [17]. Bootstrap values for 1000 replicates are given when above 50%. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The GenBank accession numbers and acronyms of viruses are shown while the details of acronyms used are provided in Table 1.
Figure 4. The phylogenetic tree was constructed based on the amino acid sequences of open reading frame III from isolates belonging to different dioscorea bacilliform virus species available in the GenBank. The evolutionary history was inferred by using the Maximum Likelihood method based on the JTT matrix-based model conducted in MEGA7 [17]. Bootstrap values for 1000 replicates are given when above 50%. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. The GenBank accession numbers and acronyms of viruses are shown while the details of acronyms used are provided in Table 1.
Pathogens 12 00245 g004
Table 1. Details of ICTV-approved species under the genus Badnavirus.
Table 1. Details of ICTV-approved species under the genus Badnavirus.
Sl. NoSpecies HostAcronymGenBank Accession Number No. of ORFsGenome Length (Nucleotide)
1.aglaonema bacilliform virusAglaonema modestumABVMH38483737178
2.Badnavirus aucubae (aucuba ringspot virus)Aucuba japonicaAuRVLC48741139092
3.Badnavirus castaneae (chestnut mosaic virus)Castanea sativaChMDMT26985337160
4.banana streak GF virusPlantainBSGFVAY49350937263
5.banana streak IM virusMusa sp. cv. MshuleBSIMVHQ59311237769
6.banana streak MY virusMusa sp.BSMYVAY80507437650
7.banana streak OL virusMusa sp.BSOLVAJ00223437389
8.banana streak UA virusMusa sp. cv. LikhakoBSUAV HQ59310737519
9.banana streak UI virusMusa sp. cv. KisansaBAUIVHQ59310837458
10.banana streak UL virusMusa sp. cv. KibuziBSULVHQ59310937401
11.banana streak UM virusMusa sp. cv. MbwazirumeBSUMVHQ59311037532
12.banana streak VN virusPlantainBAVNVAY75015537801
13.birch leaf roll-associated virusBetula pubescensBLRaVMG68641947864
14.blackberry virus FRubus sp.BVFKJ41325237663
15.bougainvillea chlorotic vein banding virusBougainvillea sp.BCVBVEU03453948759
16.cacao bacilliform Sri Lanka virusTheobroma cacaoCBSLVMF64273631772
17.cacao mild mosaic virusTheobroma cacaoCaMMVKX27664047533
18.cacao swollen shoot CD virusTheobroma cacaoCSSCDVJN60611047203
19.cacao swollen shoot CE virusTheobroma cacaoCSSCEVMF64272047131
20.cacao swollen shoot Ghana M virusTheobroma cacaoCSSGMVMF64272447009
21.cacao swollen shoot Ghana N virusTheobroma cacaoCSSGNVMF64272547173
22.cacao swollen shoot Ghana Q virusTheobroma cacaoCSSGQVMF64273447102
23.cacao swollen shoot Togo A virusTheobroma cacaoCSSTAVAJ78100397297
24.cacao swollen shoot Togo B virusTheobroma cacaoCSSTBVL1454637161
25.cacao yellow vein banding virusTheobroma cacaoCYVBVKX27664131958
26.camellia lemon glow virusCamellia japonica cv. Lemon GlowCLGVMN54241738203
27.canna yellow mottle associated virusCanna indica cv. Striped Beauty leafCaYMAVKX06602036966
28.canna yellow mottle virusAlpinia purpurataCaYMVKU16831237120
29.citrus yellow mosaic virusCitrussinensisCYMVAF34769567559
30.codonopsis vein clearing virusCodonopsis lanceolataCoVCVMK04482138112
31.commelina yellow mottle virusCommelina sp.ComYMVX5293837489
32.cycad leaf necrosisZamia fischeriCLNVEU85370939205
33.dioscorea bacilliform AL virusDioscorea rotundataDBALVKX00857337609
34.dioscorea bacilliform AL virus 2Dioscorea alataDBALV2KY82739537871
35. dioscorea bacilliform ES virusDioscorea esculentaDBESVKY82739438106
36.dioscorea bacilliform RT virus 1Dioscorea rotundataDBRTV1KX00857437702
37.dioscorea bacilliform RT virus 2Dioscorea rotundataDBRTV2KX00857737438
38.dioscorea bacilliform RT virus 3Dioscorea rotundataDBRTV3MF47684537506
39. dioscorea bacilliform SN virusDioscorea sansibarensisDBSNVDQ82207337261
40. dioscorea bacilliform TR virusDioscorea trifida var. Borelli voucher DBTRVKX43025737333
41.dracaena mottle virusDracaena sanderianaDrMVDQ47347877531
42.epiphyllum mottle-associated virusHylocereus polyrhizusEpMoaVMH39644047837
43.fig badnavirus 1Ficus caricaFBV -1JF41198937140
44.gooseberry vein banding associated virusRibes rubrum cv. Holandsky cervenyGVBAVJQ31611437659
45.grapevine badnavirus 1Vitis vinifera var. LjutunGBV-1MF78108237145
46.grapevine Roditis leaf discoloration-associated virusVitis viniferaGRLDaVHG94050346988
47.grapevine vein clearing virusVitis viniferaGVCVJF30166937753
48.green Sichuan pepper vein clearing-associated virusZanthoxylumGSPVCaVMK37135338070
49.ivy ringspot-associated virusHedera helixIRSaVMN85049038885
50.jujube mosaic-associated virusZiziphus jujubeJuMaVKX85247647194
51.kalanchoe top-spotting virusKalanchoe blossfeldianaKTSVAY18013737591
52. mulberry badnavirus 1Morus albaMBV-1LN65125826945
53.pagoda yellow mosaic associated virusStyphnolobium japonicumPYMaVKJ01330257424
54.pineapple bacilliform CO virusAnanas comosusPBCOVGU12167637543
55.pineapple bacilliform ER virusAnanas comosus var. erectifoliusPBERVEU377673 31510
56.piper yellow mottle virusPiper nigrumPYMoVKC80871247562
57.polyscias mosaic virusPolyscias fruticosa cv. Ming araliaPoMVMH47591837592
58.rubus yellow net virusRubus sp.RYNVKM07803467836
59.spiraea yellow leafspot virusSpiraea (Spiraea × bumalda)SYLSVMW08036938017
60.sugarcane bacilliform Guadeloupe A virusSaccharum officinarum var. R570 SCBGAVFJ824813 37444
61.sugarcane bacilliform Guadeloupe D virusSaccharum officinarum cv.BataviaSCBGDVFJ439817 37317
62.sugarcane bacilliform IM virusSaccharum officinarumSCBIMVAJ277091 37687
63.sugarcane bacilliform MO virusSaccharum officinarumSCBMOVM89923 37568
64.sweet potato pakakuy virusIpomoea batatasSPPVFJ560943 48082
65.taro bacilliform CH virusColocasia esculentaTaBCHVKP71017867641
66.taro bacilliform virusColocasia esculentaTaBVAF357836 37458
67.wisteria badnavirus 1Wisteria sinensisWBV1KX168422 47326
68.yacon necrotic mottle virusSmallanthus sonchifoliusYNMoVKM229702 47661
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ishwara Bhat, A.; Selvarajan, R.; Balasubramanian, V. Emerging and Re-Emerging Diseases Caused by Badnaviruses. Pathogens 2023, 12, 245. https://doi.org/10.3390/pathogens12020245

AMA Style

Ishwara Bhat A, Selvarajan R, Balasubramanian V. Emerging and Re-Emerging Diseases Caused by Badnaviruses. Pathogens. 2023; 12(2):245. https://doi.org/10.3390/pathogens12020245

Chicago/Turabian Style

Ishwara Bhat, Alangar, Ramasamy Selvarajan, and Velusamy Balasubramanian. 2023. "Emerging and Re-Emerging Diseases Caused by Badnaviruses" Pathogens 12, no. 2: 245. https://doi.org/10.3390/pathogens12020245

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop