Next Article in Journal
The Role of the CoREST/REST Repressor Complex in Herpes Simplex Virus 1 Productive Infection and in Latency
Next Article in Special Issue
Expanding Possibilities for Intervention against Small Ruminant Lentiviruses through Genetic Marker-Assisted Selective Breeding
Previous Article in Journal
Characterization of Clade 2.3.2.1 H5N1 Highly Pathogenic Avian Influenza Viruses Isolated from Wild Birds (Mandarin Duck and Eurasian Eagle Owl) in 2010 in Korea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Small Ruminant Lentiviruses: Genetic Variability, Tropism and Diagnosis

1
Laboratory of Virology, Genetics and Molecular Biology, FES-Cuautitlán, UNAM C-4 Veterinary, Cuautitlán Izcalli, State of Mexico 54714, Mexico
2
Institute of Agrobiotechnology, CSIC-UPNA-Government of Navarra, Ctra. Mutilva Baja s/n, Navarra 31192, Spain
*
Author to whom correspondence should be addressed.
Current address: Game and Fish Research Centre, University of Córdoba, Ctra Nacional IV-a Km 396, Córdoba 14071, Spain
Viruses 2013, 5(4), 1175-1207; https://doi.org/10.3390/v5041175
Submission received: 20 February 2013 / Revised: 9 April 2013 / Accepted: 12 April 2013 / Published: 23 April 2013
(This article belongs to the Special Issue Small Ruminant Lentiviruses)

Abstract

:
Small ruminant lentiviruses (SRLV) cause a multisystemic chronic disease affecting animal production and welfare. SRLV infections are spread across the world with the exception of Iceland. Success in controlling SRLV spread depends largely on the use of appropriate diagnostic tools, but the existence of a high genetic/antigenic variability among these viruses, the fluctuant levels of antibody against them and the low viral loads found in infected individuals hamper the diagnostic efficacy. SRLV have a marked in vivo tropism towards the monocyte/macrophage lineage and attempts have been made to identify the genome regions involved in tropism, with two main candidates, the LTR and env gene, since LTR contains primer binding sites for viral replication and the env-encoded protein (SU ENV), which mediates the binding of the virus to the host’s cell and has hypervariable regions to escape the humoral immune response. Once inside the host cell, innate immunity may interfere with SRLV replication, but the virus develops counteraction mechanisms to escape, multiply and survive, creating a quasi-species and undergoing compartmentalization events. So far, the mechanisms of organ tropism involved in the development of different disease forms (neurological, arthritic, pulmonary and mammary) are unknown, but different alternatives are proposed. This is an overview of the current state of knowledge on SRLV genetic variability and its implications in tropism as well as in the development of alternative diagnostic assays.

1. Introduction

Small ruminant lentiviruses (SRLV), which include caprine arthritis encephalitis virus (CAEV) and visna/maedi virus (VMV), also called ovine progressive pneumonia virus (OPPV), cause a progressive multisystemic chronic disease clinically characterized by wasting (visna); breathing difficulty (maedi) associated with pneumonia; encephalitis; arthritis and/or mastitis that affect considerably animal welfare and production. The economic effect of SRLV, often underestimated and still under study, depends on factors related to environment, breed, individual susceptibility, production system, farming practices and age of culling [1,2]. The productive impact is partly due to the premature removal of diseased animals and the consequent increase in the replacement rate [3]. In addition, SRLV infected sheep have shown in some studies decreased fertility and number of lambs per birth, as well as decreased birth weight and decreased weight gain from birth to weaning in progeny of seropositives [4,5]. However, there are studies not claiming any impact of this infection in prolificacy and weight at birth or weaning [6,7,8]. Differences in disease status of the animals and in severity of mammary gland lesions might explain the discrepancies. Animals with advanced disease present a significantly reduced body weight at slaughter, and their carcass may not be qualified for consumption [9]. Furthermore, production losses in adult females may result from decreased milk production accompanied by loss of lambs in the first week of life or a weight shortfall at weaning. SRLV infections represent a serious threat to production in small ruminants especially in intensive milk production systems, because SRLV infection occurs primarily in these systems [2]. The impact of SRLV-induced mastitis on production is also a controversial issue. Some studies do not detect differences in the quantity and quality of milk from infected and uninfected goats [10,11,12]. Other reports show a reduction in milk production (9% in goats of the Murciano-Grenadine breed) without altering milk quality [13]. Finally, there are reports showing a 15% decrease in milk production, associated with low quality by reducing the fat content [14,15]. In dairy sheep, a mean annual decline in milk production and milk fat percentage of 3.2 and 2%, respectively, has been observed [16]. In addition, SRLV infection in the small ruminants may negatively affect the quality of the milk, and it appears to trigger an increased number of somatic cells [15,17].
The intake of infected colostrum and milk by the offspring constitutes a major route of SRLV transmission, but the virus is also transmitted via the respiratory route, especially upon close contact, particularly under intensive housing or grazing conditions [2,18,19,20,21,22]. Both routes of transmission are not mutually exclusive, and the most probable scenario is that, after ingestion of infected colostrum, lambs are further exposed to the virus when raised together with infected adult animals (horizontal transmission). The horizontal transmission route has been widely accepted as responsible for SRLV spread between different geographic regions through programs involving export and exchange of goats and sheep [2]. In fact, the presence of visna/maedi was first described after importing Iceland 20 Karakul sheep from Halle, Germany in 1933 to improve Icelandic local breeds [23], and there are different documented examples that depict SRLV spread through the import of animals from Germany to Greece, Sweden to Finland, Denmark to Norway, Holland to France, Denmark to Scotland, Scotland to Canada, France to England, England to Hungary, Switzerland to USA, and USA to Mexico [2,24,25,26,27]. In the search for SRLV ancestors, recent findings reveal that, together with humans, domestic animals including goats or sheep have travelled from the Fertile Crescent to different countries through the Mediterranean Sea. Viruses infecting those animals could have been the original ancestors of SRLV [28,29,30].

2. Viral Genetic Variability Sources

As observed in other lentiviruses, the SRLV proviral genome consists in two identical positive-sense single-stranded RNA subunits (8.4–9.2 kb). Both contain structural (gag, pol and env) and regulatory (vpr-like, vif and rev) genes flanked by non-coding long terminal repeat regions (LTRs) [28]. Despite the high evolution rate of lentiviruses, many elements in the lentiviral genome are conserved over time. One of the most conserved regions of the lentiviral genome is the RNAtlys primer binding site (PBS-GAACAGGGACUUGAA), where the host lysine transfer RNA hybridizes to the viral RNA genome, serving as a primer for reverse transcription. The polypurine tract, Rev responsive element (RRE) and other elements involved in replication and packaging of the viral genome are also conserved at different degrees among lentiviruses [31]. The gag and pol genes and some regions of the env gene are relatively conserved, but others, such as those encoding Env surface protein sites able to bind antibodies, are highly variable. Genetic diversity displayed as viral quasi-species is one of the hallmarks of retroviral infection. The concept of viral quasi-species was first proposed by Manfred Eigen [32] and is defined as a set of viruses found in an infected individual [33]. Under certain circumstances of selective pressure such as that exerted by the immune system, the frequency of genetic forms in the viral population can shift. An “archive” of earlier forms of the virus is retained in proviral DNA and these forms may re-emerge. The extent of genetic diversity within a quasi-species depends on a complex set of factors, including high viral turnover, high mutation rates, retroviral recombination and selection by the host immune system until the limits of genetic and phenotypic constraints to variation [33,34,35,36].

2.1. Mutation

Mutations are the substrate for natural selection and underpin the ability of lentiviruses to evade the immune system. Like in other retroviruses, most SRLV mutations are introduced at the reverse transcription stage of the viral life cycle. The most prominent source of variation is attributed to the reverse transcriptase (RT) enzyme itself, which due to the lack of a proofreading capability leads to a high error rate (0.2–2 mutations per genome per cycle) [33,37]. This extremely low fidelity could explain the extremely high levels of genetic variation observed in vivo.
However, the genetic heterogeneity observed in vivo cannot be fully attributed to the low fidelity of RT. The minority of subpopulations in the mutant spectrum of the quasi-species viral variants that were dominant in vivo early in the evolutionary lineage of a virus can also influence the subsequent evolution of the quasi-species population [35]. In addition, early investigations into the mutation rate of human immunodeficiency virus (HIV) uncovered hypermutated retroviral genomes, where up to 40% of all available guanine bases are substituted by adenines [38]. It is now appreciated that this type of hypermutation is the result of cytosine deamination by members of the APOBEC family of nucleic acid editing enzymes [39,40]. APOBEC proteins are packaged into lentiviral virions and associate with the reverse transcription complex in the target cell, where they deaminate cytosine residues to uracyl in the single-stranded DNA minus strand, leading to G-to-A mutation in the plus strand. The cytosine deamination does not occur randomly, since APOBEC family members have distinct dinucleotide preferences. Furthermore, terminally differentiated cell types, such as macrophages, have imbalanced intracellular dNTP pools, with an excess of dUTP (uracyl) [41]. Uracyl is a natural base in RNA, but is not normally found in DNA. However, it can be incorporated into DNA due to the inability of RT to distinguish between dTTP and dUTP. Consistent with an important role for uracyl in the retroviral life cycle, many macrophage-tropic non-primate lentiviruses (such as SRLV) encode a deoxyuridine 5'-triphosphate nucleotidohydrolyase (dUTPase), which catalyzes the conversion of dUTP to dUMP, maintaining a low dUTP:dTTP ratio that ultimately prevents the misincorporation of dUTP by RT [42,43]. Inactivation of the dUTPase in CAEV and feline immunodeficiency virus (FIV) leads to an increase in the mutation rate with the accumulation of guanine to adenine mutations. Both, dUTPase and vpr-like deletions appear to be implicated in the RT fidelity [44]. The dUTPase defective recombinant viruses have a less efficient replication in macrophages and fibroblast-like cells. This may confer an advantage to the host in vivo leading to a decreased viral replication and pathogenesis, since dUTPase is apparently required to eventually develop lesions such as those involved in bilateral carpal arthritis [42,44]. The effect of dUTPase defect has been recently proposed also in infections with a field isolate (genotype E1)—whose genome naturally lacks the dUTPase encoding region and the vpr-like gene—which do not appear to ever reach clinical stages, including carpal arthritis [45]. Another genotype E variant (E2), also lacking dUTPase, however, showed certain pathogenic features in vitro and in vivo. Strikingly, neither E1 nor E2 genome exhibit high mutation rates [46]. Overall, these findings, together with recent reports describing alternative dUTPase functions may re-issue the study on the role of dUTPase in lentiviral replication [47]. Additional mechanisms altering dUTP:dTTP ratio in the cell, the docking of cellular DNA glycosylates, or the counteraction of host apolipoprotein (APOBEC3) by viral vif [48] may also result in a major source of viral heterogeneity and define the infection outcome [39].

2.2. Recombination

Mutation alone is unlikely to explain the adaptive flexibility of lentiviruses. Recombination may occur frequently in the viral genome, as shown in the env gene of VMV strain 1514 in vitro and in vivo [49,50]. This mechanism of genetic diversification can efficiently shuffle mutations within a quasi-species; can rapidly assemble beneficial genetic combinations that would be difficult to generate by mutation alone; and can also effectively remove deleterious mutations. In contrast with the slow and steady changes caused by mutation, recombination is a much more powerful evolutionary force. First, recombination facilitates the repair of viral genomes. This can be due to physical repair at genome or breaks accumulating deleterious mutations via a copy choice mechanism [36]. In the absence of recombination, organisms tend to accumulate deleterious mutations that reduce viral fitness in each replication cycle. Multiple recombination events take place, with estimates generally falling between three and nine recombination events per genome per replication cycle [51].
Efficient recombination in retrovirus arises as a result of the co-infection of two viruses in the same cell and co-packaging of two copies of the RNA genome into each virion [52]. When a cell becomes co-infected by two or more different viruses, the corresponding RNA genomes can become co-packaged into the viral progeny. During subsequent reverse transcription, the viral reverse transcriptase readily switches between these two templates [53,54], which leads to the production of recombinant cDNA. If these templates are identical, then template switching will be genetically neutral and recombination will not be detected. Conversely, if these two genomes are non-identical, template switching will lead to viral recombination and progeny virus will be genetically distinct from the parental strains [36]. In natural infections, SRLV recombinations between different genetic groups (CAEV (B)-VMV (A)) [55] and between genetic variants of the same group (B1 CAEV) in goats have been identified [25]. Recombinant viruses may contain recombinant env genes involved in SRLV tropism, so that they may exhibit a modified range of targets (cell, tissue, and host species).

3. Phylogeny

The existence of high genetic variability among SRLV has given rise to numerous studies on the phylogenetic relationships among sequences obtained in different countries [56]. The classification of viral genotypes into groups and subtypes proposed in the last decade [27] in studies involving two long segments of the SRLV genome (gag-pol segment 1.8 kb; and pol segment 1.2 kb) is widely accepted at present. Accordingly, SRLV are classified into five groups (A–E). Following HIV classification criteria [57], genotype groups differ by 25% to 37% in their nucleotide sequences. However, genotypes A, B and E may further be distributed into different subtypes, differing in 15% to 27% of their sequence. Genotype D has been only described in Swiss and Spanish sheep, and only regarding pol sequences. However, there are no other studies confirming the existence of this genotype. Rather, phylogenetic analysis on additional (gag) sequences of the same (group D) isolates, classify these sequences with genotype A, suggesting that genotype D is in fact genotype A, exhibiting divergence in pol gene.
Group A has so far 15 recognized subtypes, A1–A15, group B has three subtypes, B1–B3 and group E has only two subtypes, E1 and E2 (Table 1). Although SRLV infection in small ruminants is widely distributed in all continents [27,58], little information is available on the genetic variants circulating in different geographic regions (Table 1). SRLV complete genomes have been sequenced and are available in the GenBank derived from goat (CAEV-CO [59,60], 1GA [61,62], Gansu [63], Shanxi [64], FESC-752 [25], Seui [46] Roccaverano [45] and A4 [27] viruses) and sheep (Fonni [30], Volterra [30], 496 [65], SA-OMVV [66], KV1514 [67], KV1772 [68], LV1 [69], EV1 [70], P1OLV [71], 85/34 [72] and 697 [73] viruses) (Figure 1). In addition, partial sequences have been published in Brazil [74,75,76], Canada [77], Finland [78], France [27,79,80,81], Greece [82], Ireland [83], Japan [84], Netherlands [85], Poland [86,87], Russia [88], Slovenia [89], South Korea [90] and Turkey [29,30].
Table 1. Distribution of SRLV genotypes and subtypes which infect goats and sheep from different countries.
Table 1. Distribution of SRLV genotypes and subtypes which infect goats and sheep from different countries.
CountryGenotype AGenotype BGenotype CGenotype DGenotype EReferences
A1A2A3A4A5A6A7A8A9A10A11A12A13A14A15B1B2B3CDE1E2
BrazilG/S G [74,75,76]
Canada S G [77]
China G [63,64]
EnglandS [70]
FinlandSS [78]
FranceS G/S G/SG/S [27,79,80,81,91,92]
Greece S [82]
IcelandS [67,68,69]
IrelandG [83]
ItalyG GG/SGG/S G/SSG/S GG[30,45,46,89,93,94]
Japan G [84]
Mexico G/S [25]
NetherlandsS [85]
NorwayS G/S [61,62]
PolandG/S G/SG/S GS [86,87]
PortugalS [71]
Russia G [88]
Slovenia S GSG [89]
South AfricaS [66]
South Korea G [90]
SpainS S S G/S [65,73]
SwitzerlandS G/SG/SGG/SG GS S [27,95]
Turkey SS S S S S [29,30]
USA S G/S [59,60,72,96]
G = Goat; S = Sheep.
Figure 1. Phylogenetic tree involving SRLV complete sequences obtained from GenBank (accession numbers in bold and underlined) that include the name of isolate and country origin. Country abbreviations: CHI - China (Gansu AY900630 and Shanxi GU120138); ENG - England (EV1 S51392); ICE - Iceland (KV1514 M60610, LV1 M10608 and KV1772 L06906); ITA - Italy (Roccaverano EU293537, Seui GQ381130, Fonni JF502416 and Volterra JF502417); MEX - Mexico (FESC-752 HM210570); NOR - Norway (1GA AF322109); POR - Portugal (P1OLV AF479638); SOA - South Africa (SAOMVV M34193); SPA - Spain (496 FJ195346 and 697 HQ848062); SWI - Switzerland (A4 AY445885); USA - United States of America (CAEVCo M33677 and 85/34 AY101611, U64439). The SRLV genotypes and subtypes are indicated.
Figure 1. Phylogenetic tree involving SRLV complete sequences obtained from GenBank (accession numbers in bold and underlined) that include the name of isolate and country origin. Country abbreviations: CHI - China (Gansu AY900630 and Shanxi GU120138); ENG - England (EV1 S51392); ICE - Iceland (KV1514 M60610, LV1 M10608 and KV1772 L06906); ITA - Italy (Roccaverano EU293537, Seui GQ381130, Fonni JF502416 and Volterra JF502417); MEX - Mexico (FESC-752 HM210570); NOR - Norway (1GA AF322109); POR - Portugal (P1OLV AF479638); SOA - South Africa (SAOMVV M34193); SPA - Spain (496 FJ195346 and 697 HQ848062); SWI - Switzerland (A4 AY445885); USA - United States of America (CAEVCo M33677 and 85/34 AY101611, U64439). The SRLV genotypes and subtypes are indicated.
Viruses 05 01175 g001
According to the most recent phylogenetic information, mostly based on the gag gene, the types B, C and D and only nine of the 15 group subtypes A (A1, A3, A4, A5, A6, A9, A11, A12 and A13) infect both sheep and goats. Other groups and/or subtypes have only been described in one of the ruminant species: sheep (A2 and A15), or goats (A7, A8, A10 [93], A14, E1 and E2). However, as more information is generated, SRLV species restrictions are likely to dwindle (Table 1).

4. Tropism

SRLV tropism is linked to both, host genetics and viral genome heterogeneity, and can be studied according to differences in the targets addressed regarding: (a) host species (goats or sheep, or both); (b) tissue, differing according to the form of disease (mastitis, arthritis, encephalitis and/or pneumonia); and (c) cell type.

4.1. Host Species Restriction

Studies from several countries have confirmed that CAEV and VMV, originally established as specific pathogens in goats and sheep respectively, often cross the species barrier infecting the new host, persisting in it and spreading across the new host population. The first evidence of this transgression was obtained by experimentally infecting sheep and goats with CAEV or VMV [97,98,99,100,101,102]. Subsequently, evidence has been obtained from phylogenetic analysis of sequences derived from French sheep infected by viruses closer to CAEV (group B2) than to VMV (group A) [79,80,81,91,92]. On the other hand, molecular epidemiology studies have demonstrated direct virus passage from sheep infected with VMV-like strains (A4) to goats [95]. In line with this, the genotype B1 had been considered strictly caprine until the description of B1 infected sheep [103]. Thus, even if some genotypes might have been originally assigned to a single host species, the host species spectrum may be wider in nature.
Cell receptor usage explains, at least in part, different patterns in SRLV host restriction. As indicated above, the high variability of the ENV protein may be due to its interaction with the immune system and the generation of viral escape mutants [49,104]. Thus, changes in env may affect the ability of the virus to bind the putative cell receptor(s) and co-receptor(s). Differences in receptor usage have been observed between VMV and CAEV strains [105]. Also, differences in permissivity have been found in vitro between cell lines of heterologous origin (chicken, hamster, human, monkey and quail), which are permissive to SRLV infection and Chinese ovary hamster cells non-permissive to the infection [106,107]. Specific amino acid residues of the viral SU protein (see below) have been linked to cell receptor recognition [108]. Some of these residues may also be related to the presence of arthritis [104].
Another explanation for the host-specific viral restriction lies in the individual genetic background. Besides adaptive immunological correlates of protection according to immunization studies [109], cell molecules of the innate immunity may be involved in delimiting host restriction. Among these, APOBEC and TRIM5 are the most widely studied, as they act in different species directly by mutating viral genome and interacting with viral capsids, respectively, showing species specificity [110,111]. However, the virus has developed mechanism to counteract proteins of different host species. VMV studies have shown that the viral Vif protein can neutralize a broad number of A3Z3 proteins (APOBEC) irrespective of the species of origin (sheep, humans, macaques, cows and cats) [112].

4.2. Viral Genetics by Organ and Tissue

Amino acid motifs conserved at specific viral protein positions involved in the viral entry (“signature patterns”) [113] have been found in individuals presenting the same form of disease in HIV [114] and FIV [115] infections. Signature patterns that define organ tropism (disease form) have not been found so far in SRLV infections [116,117]. At the single individual level infected with a lentiviral quasi-species, each particular virus variant may be confined to one particular compartment, organ or tissue within the individual. In HIV infections it is known that these viral populations change to adapt to local media within the individual through genetic drift (founder principle) or selection pressure. However, when the flow of genes between viral subpopulations within the individual is significantly restricted, then each subpopulation can become genetically distinct. This phenomenon named compartmentalization [115,118,119] may be derived from micro-evolution and/or the presence of various but phylogenetically related genotypes [120,121,122]. High viral mutation rate in vivo can quickly enhance the genetic distance between subpopulations. Also, differences in selective pressures imposed by the immune system can result in divergent evolution of the virus, affecting cell tropism, phenotypic characteristics and/or pathogenesis, as shown in HIV infections [123,124,125,126].
Regarding the particular viral genetic region involved in tropism, LTR and env have been the most widely studied in lentiviral infections. LTR, being a non-coding region responsible for docking important cellular transcription factors, is essential for viral replication and may affect viral phenotype in cell culture [127,128]. This LTR function is achieved through LTR transcription factor binding sites within the U3 region, some of which are related to cell tropism [129]. Tissue tropism has been evaluated ex vivo by studying SRLV LTR sequences in different tissue samples but, again, no signature patterns have been found so far [130]. Alternatively, in other lentiviruses, genomic regions such as hypervariable regions of the env gene involved in compartmentalization have been found implicated in tropism. This is the case of the V3 hypervariable region of env in HIV infections, being determinant of cell tropism and replication efficiency as it relates to the fusion and virus adsorption to the cell in macrophage-tropic strains [113]. In SRLV genomes, five variable (V1 to V5) and four conserved (C1 to C4) regions in the envelope protein have been identified [108,131]. The SRLV env hypervariable region (V4) is structurally and functionally analogous to the V3 region of HIV [108,132,133]. To colonize different organs (lung, mammary gland, brain, joints) SRLV can undergo variations in the V4 region of the env gene during early infection giving rise to different viral subpopulations, as shown in other lentiviruses [132]. A study on SRLV compartmentalization in the mammary gland of goats and sheep showed the presence of different viral sequences in the V4 region of the env gene compared to blood-derived cells and colostrum. It was proposed that the mammary gland was colonized with a “founder virus” which possibly represented the most common variant circulating (dominant) in the blood at the time of infection. Alternatively, it was proposed that the initial mechanism could be totally different to compartmentalization and would reflect a selection pressure caused by a particular cellular tropism or related to the immune system action exerted against the virus in this compartment [116].
In another study, SRLV genotype A sequences from sheep suffering from neurological disease (visna) showed a clear tissue compartmentalization in the central nervous system (CNS) and other organs such as lung and mammary gland, related to horizontal and vertical transmission of SRLV infection, respectively. Bayesian approach inferences have suggested that proviruses from alveolar macrophages and peripheral blood mononuclear cells (PBMC) represent the most probable common ancestors (infecting viruses) in the animal. Likely, PBMC become infected after the intake of virus or infected cells/particles through respiratory and/or mammary secretions, including colostrum/milk. Neuroinvasion in the visna outbreak involved microevolution after initial infection with SRLV [117]. Of note, the findings on SRLV diversification, adaptation and/or compartmentalization in the brain may be independent of those observed in lymphoid tissues or other organs, where the virus is exposed to more abundant immune pressures affecting viral replication [134].

4.3 Cellular Tropism

The members of the lentivirus genus differ in cellular tropism and disease development, being distributed into two groups. One group includes HIV, FIV and simian immunodeficiency virus (SIV), all of which replicate in macrophages and lymphocytes causing acquired immunodeficiency syndrome and specific-organ disease affecting lungs, CNS and gastro-intestinal tract. The other group includes equine infectious anemia virus (EIAV) and SRLV, both of which replicate primarily in macrophages (lymphocytes are not infected) and cause a disease affecting specific organs, mainly lung, mammary gland, CNS and/or joints [127,135].
In the VMV epidemic in Iceland, the most common clinical signs corresponded to the maedi (respiratory form) and the visna (nervous form) was only reported in herds with high prevalence of maedi for several years. This was originally attributed to the cell tropism of the virus rather than the genetic variation of sheep, because there was only one sheep breed in Iceland [136,137]. Subsequently, it was shown that the cellular tropism of VMV strains from sheep with visna appeared to be different from those isolated from sheep with maedi. The virus isolated from brain cells replicated more rapidly in sheep choroid plexus compared to isolates from lung and this difference was related to differences in env and LTR regions [138].
In vivo SRLV have a marked tropism for cells of the monocyte/macrophage lineage and dendritic cells [139,140], the virus infects monocytes, so these cells are permissive for viral entry and genome integration (provirus), but the infection remains latent until the cell differentiates into macrophage [139,141,142]. This maturation enhances the expression of the transcription factors c-Fos and c-Jun, which bind to the AP-1 and AP-4 promoter binding sites of LTR triggering transcription, and consequently viral replication and productive infection [143]. Thus, SRLV replicate in mature macrophages or related tissues rather than in circulating monocytes, the latter acting as a “Trojan horse” [144]
SRLV are highly variable genetically and antigenically and also show phenotypic differences in vitro. Specifically, isolates may differ from each other in the ability to selectively and productively infect particular cell types and also in the capacity to induce a cytopathic effect. Accordingly, SRLV may be classified phenotypically as rapid/high or slow/low. The rapid/high strains replicate rapidly, inducing the formation of syncytia, cell lysis, and reaching high titers. In contrast, the slow/low viruses grow slowly and at a low titer. Frequently, sheep isolates belong to the first type; whereas isolates from goats show a slow/low phenotype. However, SRLV strains that show an intermediate phenotype between the most extreme phenotypes of VMV-like and CAEV-like viruses have been isolated from both sheep and goats [145,146].
Because SRLV show in vitro high viral production in permissive cells such as macrophages, these cells are used routinely for isolating the virus, even though they are terminally differentiated [147]. Fibro-epithelial synovial membrane cells and choroid plexus cells derived from goats or sheep are also routinely used for in vitro production of SRLV [96,148]. Other cell types permissive to SRLV infection are from lung explants [149], skin fibroblasts [140], spleen and corneal cells [150,151], testicle cells [152], goat endothelium [153,154], mammary gland epithelial cells (TIGMEC cell line) [155,156], granulose cells [157], oviduct cells [158], microglial cells [159,160], tubular epithelial cells [161], hepatocytes [162], cardiac myocytes [162] and the third eyelid cells [163]. However, the restricted viral antigen expression and the low number of infected cells found in some of these cells should be taken into account before interpreting the biological relevance of these cells in SRLV persistence [164]. Experiments employing pseudotyped vectors (vaccinia virus or murine leukemia virus-based) with SRLV ENV indicate that ENV-based interaction with cells renders them permissive to virus entry. This is the case of cell lines Hela, BHK-21 and others from different origins (human or other species) [106,107], suggesting that these cells have a receptor for ENV and that the virus enters the host cell by fusion of the membrane receptor to viral ENV SU protein gp135 [165]. However, this permissiveness at the entry step [166] may not be associated with a capability of productive infection or to a possible natural susceptibility. Although in previous SRLV studies restriction had been attributed to the lack of functional receptors [167], post-entry restriction factors such as those interacting with Gag [110] rather than Env proteins may have been responsible for the lack of productive infection.
Different attempts have been made to identify and characterize SRLV receptor(s). Candidate SRLV receptor molecules include: a membrane-associated proteoglycan substituted with one 30 kDa chain (s) chondroitin sulfate glycosaminoglycan [168]; molecules of the major histocompatibility complex class II (MHC class II), as preincubation with MHC class II antigens inhibits VMV infection (even though antibodies specific to antigens of MHC class II do not inhibit infection) [169]; CD4 and CXCR4 molecules, which have been proposed as optional auxiliary components for VMV receptor (or receptor complex) that facilitates membrane fusion events mediated by VMV [166,170]; and a complex comprising three membrane proteins of 15, 30 and 45 kDa, identified as VMV binding proteins [171]. However, none of these molecules has been established as the main/essential receptor. Analysis of somatic hybrid cell lines (CHO cell panel containing different mouse chromosomes) permissive to VMV entry has shown that the gene for cellular receptor is in the murine chromosome 2 or 4 (although involvement of chromosomes 6 and X is not excluded). On the other hand, it also showed using hybrid cell lines hamster × sheep and hamster × human [172] that the VMV receptor gene maps to sheep chromosome 3p (OAR-3p) and a region of human chromosome 2 (HAS-2p25 > q13) has retained synteny with sheep chromosome 3p. These regions do not include any gene of known receptor or co-receptor of lentiviruses, which would indicate that SRLV receptor uses a different molecule to infect human cells. Apparently, SRLV may use different cell receptors likely depending the cell it encounters [105]. Recently, the mannose receptor (MR, codified in mice by chromosome 2) has been described in sheep and identified as a potential SRLV receptor. Accordingly, three cell phenotypes have been proposed in small ruminants regarding the expression and usage of MR as SRLV receptor: cells using this molecule as the main receptor (synovial membrane cells), cells having this and at least another main SRLV receptor(s) (macrophages) and cells lacking MR and using other SRLV receptor(s) instead (skin fibroblasts) [173]. Interestingly, the presence of this receptor has been associated with the evolution of disease, as its expression increases in target organs exhibiting the most severe lesions in SRLV clinical infections [174].

5. In Vitro Diagnosis

Different laboratories have applied direct diagnostic methods such as viral isolation, a very laborious method whose performance can be hampered by the lack of permissive cell lines and limited viral production in cultured cells [96]. Isolation of virus attempted from leukocytes of infected animals often fail, indicating that a negative result is not always reliable [175]. Although this method is not suitable for large-scale studies, it can be performed by co-cultivation (commonly of blood monocyte derived macrophages with fibro-epithelial cells), whereby the detection of cytopathic effect or reverse transcriptase activity indicate the presence of virus [65]. Similarly, immunohistochemical methods, allowing the detection of viral protein antigens by specific antibodies in histological samples, smears or cytospin preparations, are useful in research and confirmatory studies, but are not commonly applied in differential routine diagnosis on live animals, because of the cost, low availability of reagents and samples and limited sensitivity of some of these techniques [175]. Also, in situ hybridization methods have been employed for histologic studies, but this laborious procedure is only used for research purposes [176,177]. Although PCR-based diagnosis is the most common of the molecular techniques applied in SRLV diagnosis in vitro (see below), indirect methods such as serological diagnosis to detect antibodies to SRLV are overall the most commonly applied approach to detect infection. However, there are problems inherent to serological diagnosis. These include the high genetic/antigenic variability existing in laboratory and field strains, the presence of maternal antibodies, the relatively long lag period between infection and antibody production (serological gap) and the intermittent seroconversion (by fluctuating titers throughout the animal’s life). Antibody production has been detected by techniques such as agar gel immunodiffusion (AGID), the enzyme-linked immunosorbent assay (ELISA), radioimmunoprecipitation (RIPA) and Western blot (WB) [178,179,180]. Both RIPA and WB are mainly used for confirmation [181]. AGID is highly specific but relatively insensitive [178] and often linked to subjective interpretation, inapplicability for the determination of antibodies in milk and lack of automation, motivating its replacement by ELISA methods of relatively low cost and easy implementation and interpretation [2,179,181].

5.1. ELISA Tests

In spite of the antigenic variability, cross-reacting antibodies between VMV-like and CAEV-like antigens have been described [182] and have enabled for years the control of SRLV-induced disease. Sheep and goats remain infected for life, but many seropositive animals may never show clinical signs of SRLV infection. Antibodies do not protect against the disease and are mainly indicators of infection. The virus escapes from the immune attack even in the presence of neutralizing antibodies, as proviral DNA integrates into host cell genome and viral mutations take place during viral replication [183]. On the other hand, a seronegative animal cannot be strictly considered free of infection, either because antibody titers are below detection level before seroconversion, or because fluctuation in antibody levels occurs throughout the animal’s life. Newborns infected at birth have maternal antibodies (through colostrum/milk intake) for at least two or three months. Thereafter, they are usually seronegative until they seroconvert between six and twelve months of age. Performing diagnostic tests immediately before pregnancy does not necessarily ensure that mothers will not seroconvert after parturition. Thus, infected seronegative animals constitute a potential source of infection through vertical and horizontal transmission [179,181].
To date, no single technique or test can be proposed as “gold standard” to determine the infection status of the animal [175]. Competitive ELISA methods (cELISA CAEV of VMRD Inc. Pullman, WA) [184,185] using monoclonal antibodies to viral envelope protein (ENV-SU, gp135) epitopes have been developed [186] and indirect ELISAs have been frequently applied [187,188], although few of them have been compared internationally for both sheep and goats [189,190,191]. Currently available indirect standard ELISAs are used in different formats and designs. Examples are those using either whole virus antigens such as the AG-CHEKIT (CAEV/MVV kit, IDEXX Switzerland AG, Liebefeld, Bern, Switzerland) [192] or recombinant proteins plus a peptide, specifically, the GAG p25 recombinant protein and a TM peptide derived from genotype A, as used in assays Elitest-MVV (HYPHEN Biomed, Neuville-sur-Oise, France) [193] and Pourquier (ELISA Maedi-Visna/CAEV serum verification Institut Pourquier, Montpellier, France). But in some cases, monostrain assays may not cover the whole SRLV antigenic spectrum and fail to detect genotype E [188], B2 [194] and A4 infections [195]. Besides the viral p25 protein included in the test, p14 and p17 proteins also present immunodominant epitopes and could be used for differential VMV/CAEV diagnosis [196,197]. Gag proteins (p25) have been used in early diagnosis and detection of cross-reacting antibodies for more than 20 years [179]. However, new genotypes have emerged [30,45], broadening diagnostic possibilities, leading to the development of new standard indirect ELISA tests, such as the one based on a mixture of gag and env peptides of three different SRLV genotypes (A, B and E) (IN3 diagnostic. Eradikit® SRLV indirect ELISA for Small Ruminant Lentivirus).
The ENV TM protein has an added value in diagnosis of long-term infections such as those present in diseased animals [198] and therefore is included in most subunit assays. SU peptides have also been widely used in serological diagnosis and, due to high variability of SU proteins, these peptides are being useful to serotype and perform strain-specific diagnosis [198]. Synthetic peptide ELISAs [199] have been evaluated by our group, outlining their utility in field sample diagnosis of animals from neurological [73] and arthritic [65] outbreaks.
Antibody detection technology is certainly a valuable tool for determining the infection status of a herd/flock, although in some areas with limited resources this technology may be somewhat expensive if performed at the individual level. Therefore, strategies for detecting antibodies and infection in the flock, involving the use of bulk milk [200,201,202] or pooled serum [203] and semen samples [191,204] are particularly useful in these areas.

5.2. Polymerase Chain Reaction (PCR)

Molecular techniques of use in SRLV diagnosis include the heteroduplex mobility assay (HMA) [80], which may help to a genotypic characterization of circulating strains within the flock, a loop-mediated isothermal amplification technique [64] and polymerase chain reaction (PCR) procedures, the most commonly used to directly assess the presence of viral nucleic acid. Conventional PCR for diagnostic purposes has been more extensively studied than the real-time PCR (rtPCR). The introduction of rtPCR represents a breakthrough in molecular diagnosis [205,206]. The PCR and the early detection of amplicons have been combined in this technique [206,207]. The rtPCR has been useful for detection and quantification of viral nucleic acids in different cells or tissues, although its use in SRLV routinary diagnosis is not widespread [207,208,209]. Primers designed in almost all SRLV genetic regions [205] have yielded reliable and fast results. The rtPCR diagnostic approach presents a decreased risk of cross-contamination and has a high sensitivity using SYBR Green I or TaqMan-based methods [207]. Furthermore, the use of oligonucleotide probes helps to increase rtPCR specificity [205]. However, this technique may work only on a restricted number of strains and its use in field diagnosis may present difficulties due to variation in the primer binding site.
The first conventional PCR protocols designed to detect SRLV were published in the early 1990s [210,211]. PCR was considered initially a successful tool in SRLV diagnosis, due to its ability to directly detect viral nucleic acids, either in the infected cell as a provirus (DNA) or in exudates that contain free virus particles (RNA; reverse transcription-PCR) [212]. In addition to conventional PCR, variants of this technique have appeared involving two (or more) amplification rounds (nested-PCR and seminested-PCR) in order to increase the sensitivity [213,214,215,216,217]. Different studies have shown the utility of conventional PCR to detect SRLV infections in different animal samples, namely PBMC [214,218,219,220], peripheral blood leukocytes (PBL) [213,221], milk or mammary secretions [212,214,218,219,220,222], semen [223,224], synovial fluid [218] and other tissues [214,222,225]. The efficiency of PCR depends mainly on the specificity of the primers designed, the choice of the amplified target viral region, and the sensitivity of the technique [35,127,226]. The low viral load existing in infected animals, the low proportion of permissive cells in some tissues (there is one infected monocyte in 104–105 PBMC) [206] and the high genetic heterogeneity may hamper PCR diagnosis of SRLV [2,226]. Furthermore, the cell type may affect PCR performance. However, PCR has the advantage over serological methods of detecting infection in animals with colostral antibodies. In milk samples from sheep and goats (small ruminants) a seminested-PCR method based on the pol region (pol-seminested-PCR) has yielded positive results in all the infected small ruminants tested (the seropositives determined previously by a serological test and a seronegative animal) [212]. Similar PCR sensitivities have also been found in PBMC, milk cells and synovial fluid [218]. Here, however, are also studies reporting a decreased sensitivity when using milk samples (23.6%) relative to blood samples (33.3%), using a gag-PCR [219] or a LTR-PCR procedure [222]. Target sequences for PCR primers design are widespread throughout the SRLV genome including LTR, gag, pol and env regions [164,212,222,224] and lead to different sensitivity and specificity values. In PCRs amplifying segments of the SRLV gag and pol regions, the gag-PCR was more sensitive than pol-PCR [211]. However, the LTR-PCR may be more sensitive than PCRs based on gag and pol regions [213,222,226]. In a study that examined four pairs of primers designed in LTR, gag (capsid and matrix) and env regions using DNA from cells infected with virus from various geographical areas, PCR primers targeting the LTR region, gag(capsid), env and gag (matrix) amplified 100%, 62.5%, 25% and 12.5% of the samples, respectively [227]. This is in contrast with another study showing that a PCR based on env sequences could be more efficient at detecting infected animals (93.3%) compared to pol-PCR which only detected 60% of animals infected. Despite its presumed high variability, the env region can be suitable for diagnosis by PCR (Table 2) [224].
In SRLV infections, PCR-positive reactions are often found amongst seronegative animals (false negatives) (Table 2) [219,228], which later seroconvert, thereby demonstrating a lack of sensitivity of the serological methods compared to PCR under particular conditions [195].
Table 2. Different PCRs techniques, target DNA samples, regions of the viral genome used for PCR design and sensitivity of the PCR method for detection of infection by SRLV.
Table 2. Different PCRs techniques, target DNA samples, regions of the viral genome used for PCR design and sensitivity of the PCR method for detection of infection by SRLV.
PCR typeDNA or cDNA sourcePrimers location into the viral genomeSensitivity testReferences
cPCRCell culturegag and polpol-PCR was more sensitive than gag-PCR[210]
snPCRPBLpol and LTRLTR-PCR was more sensitive than pol-PCR[212]
nPCRPBMCgag and polgag-PCR was more sensitive than pol-PCR[213]
snPCRPBMCgaggag-PCR was less sensitive than AGID[214]
snPCRCell culture and PBMCpolHigh sensitivity, by using degenerate primers[215]
nPCRPBMCgaggag-PCR was more sensitive than AGID in seronegative animals[216]
nPCRPBL and bloodgag and LTRgag-PCR was more sensitive than LTR-PCR[220]
cPCRSemenenv and polenv-PCR was more sensitive than gag-PCR[222]
cPCRPBMC, milk cells and tissuesLTRLTR-PCR had a sensitivity of 98% with regard to AGID and ELISA[223]
cPCRCell cultureLTR, gag and envLTR-PCR was more sensitive than gag-PCR and env-PCR[226]
cPCRPBMCgaggag-PCR was less sensitive than AGID[228]
cPCRCell culture and PBMCgaggag-PCR was more sensitive than the ELISA and WB[229]
nPCRPBLgaggag-PCR increases its sensitivity when used along with hybridization[230]
cPCR: conventional PCR; snPCR: seminested PCR; nPCR: nested PCR; PBMC: peripheral blood mononuclear cells; PBL: peripheral blood leukocytes; AGID: agar gel immunodiffusion; ELISA: enzyme-linked immunoassay; WB: western blot.
Studies based on gag-PCR diagnosis reveal that this method may detect infection in 88.3% of sheep and goats that are seropositive by AGID and in 11% of seronegatives (67% of animals eventually seroconverted) [229]. In line with this, a gag-seminested-PCR procedure allows the detection of 70% of the AGID positive sheep and 11% of seronegative animals [215]. Comparative studies on the sensitivity of the PCR with regard to serological tests (ELISA and AGID) have shown that LTR-PCR is less sensitive than the ELISA and AGID. However, PCR showed a specificity of 100% (like AGID), whereas ELISA was less specific (59%) [230]. LTR-PCR specificity is maintained (100%) in different tissues, but sensitivity values change according to the tissue under study [83.5% (peripheral blood leukocytes), 66.7% (milk cells) and 88% (tissues)], being less sensitive than the ELISA and AGID [222].
Since PCR assays may fail to detect virus when virus load is below the assay threshold, PCR sensitivity may increase upon co-cultivation of PBMC from infected animals with permissive fibroblasts in vitro [225,231,232] or by using probes in the Southern blot [214,219] or in situ hybridization [228,233] techniques. The use of strain-specific and sequence-degenerated primers could also improve sensitivity, as shown by Elthair and collaborators [216] using a pol-seminested-PCR on AGID-seropositive animals infected with two viral subtypes (genotypes A and B). The percentage of pol-PCR positives in this study was 62.2% and 30.5% for types A and B, respectively. In addition, the pol-PCR detected 38.2% of positive animals among AGID seronegatives [216].
Although different studies indicate that in general the PCR methods tend to be less sensitive than serological tests (particularly ELISA tests) (Table 2) [213,220], PCR testing can detect infected animals prior to seroconversion [213,219,228,234,235]. According to SRLV pathogenesis, provirus should be present prior to the antibody response to the virus during initial infection of adult animals, but this is not always detected. In studies comparing PCR with serological test (AGID and ELISA) diagnosis, agreement (positive concordance) ranged from 70% to 94.7%, whereas disagreement (negative concordance) ranged from 87.5% to 100% [181,218,219,222,224,229]. This suggests that a combination of serology and PCR might be optimal for detecting the infectious status of a population [217,230,236]. Furthermore, PCR products can be systematically sequenced [27,94], which may allow a better design of primers for new PCRs with a wide range of recognition. The development of a single PCR to detect all SRLV may still be utopian, although attempts to cover a large spectrum of genomes by including different sets of primers in the assay have been made [27].

6. Concluding Remarks

The basic mechanisms of mutation and recombination are complex biological processes, but they may help to understand the true role of viral diversity in SRLV pathogenesis. Virus and host genetic and microenvironmental factors involved in cell, tissue/organ and host tropism of these viruses are pivotal issues in current investigations. Genetic and antigenic variations of the virus represent challenges in SRLV diagnosis. The success in avoiding SRLV infection spread depends largely on early detection and culling of infected animals in the heard/flock. Detection of specific antibodies in serum or milk by ELISA, typing the circulating strains by local strain-based ELISAs, HMA or PCR-sequencing techniques, and use of PCR for confirmatory purposes are solid technological approaches of practical use for reducing the risk of misdiagnosis in different areas. Finally, diagnostic methods suitable to both sheep and goats should be implemented in joint programs for these species for successful SRLV control and eradication strategies.

Acknowledgments

The authors would like to thank Karen Janett Becerra for help in writing of the manuscript.

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Christodoulopoulos, G. Maedi-Visna: Clinical review and short reference on the disease status in Mediterranean countries. Small Rum. Res. 2006, 62, 47–53. [Google Scholar] [CrossRef]
  2. Peterhans, E.; Greenland, T.; Badiola, J.; Harkiss, G.; Bertoni, G.; Amorena, B.; Eliaszewicz, M.; Juste, R.A.; Krassnig, R.; Lafont, J.P.; et al. Routes of transmission and consequences of small ruminant lentiviruses (SRLVs) infection and eradication schemes. Vet. Res. 2004, 35, 257–274. [Google Scholar] [CrossRef]
  3. Anderson, B.C.; Bulgin, M.S.; Adams, S.; Duelke, B. Firm udder in periparturient ewes with lymphocytic accumulations, retrovirus infection, and milk unavailable at the teat. J. Am. Vet. Med. Assoc. 1985, 186, 391–393. [Google Scholar]
  4. Keen, J.E.; Hungerford, L.L.; Littledike, E.T.; Wittum, T.E.; Kwang, J. Effect of ewe ovine lentivirus infection on ewe and lamb productivity. Prev. Vet. Med. 1997, 30, 155–169. [Google Scholar] [CrossRef]
  5. Pekelder, J.J.; Veenink, G.J.; Akkermans, J.P.; van Eldik, P.; Elving, L.; Houwers, D.J. Ovine lentivirus induced indurative lymphocytic mastitis and its effect on the growth of lambs. Vet. Rec. 1994, 134, 348–350. [Google Scholar]
  6. Dungu, B.; Vorster, J.; Bath, G.F.; Verwoerd, D.W. The effect of a natural maedi-visna virus infection on the productivity of South African sheep. Onderstepoort. J. Vet. Res. 2000, 67, 87–96. [Google Scholar]
  7. Legrottaglie, R.; Martini, M.; Barsotti, G.; Agrimi, P. The effects of ovine lentivirus infection on some productive aspects in a Sardinian sheep flock from Italy. Vet. Res. Commun. 1999, 23, 123–131. [Google Scholar] [CrossRef]
  8. Snowder, G.D.; Gates, N.L.; Glimp, H.A.; Gorham, J.R. Prevalence and effect of subclinical ovine progressive pneumonia virus infection on ewe wool and lamb production. J. Am. Vet. Med. Assoc. 1990, 197, 475–479. [Google Scholar]
  9. Arsenault, J.; Dubreuil, P.; Girard, C.; Simard, C.; Belanger, D. Maedi-visna impact on productivity in Quebec sheep flocks (Canada). Prev. Vet. Med. 2003, 59, 125–137. [Google Scholar] [CrossRef]
  10. Leitner, G.; Krifucks, O.; Weisblit, L.; Lavi, Y.; Bernstein, S.; Merin, U. The effect of caprine arthritis encephalitis virus infection on production in goats. Vet. J. 2010, 183, 328–331. [Google Scholar] [CrossRef]
  11. Nord, K.; Adnoy, T. Effects of infection by caprine arthritis-encephalitis virus on milk production of goats. J. Dairy Sci. 1997, 80, 2391–2397. [Google Scholar] [CrossRef]
  12. Turin, L.; Pisoni, G.; Giannino, M.L.; Antonini, M.; Rosati, S.; Ruffo, G.; Moroni, P. Correlation between milk parameters in CAEV seropositive and negative primiparous goats during an eradication program in Italian farm. Small Rum. Res. 2005, 57, 73–79. [Google Scholar] [CrossRef]
  13. Martínez, N.B.; Peris, R.C.; Roche, J.M.L.; Caballero, G.C. Efectos del virus de la artritis-encefalitis caprina sobre la producción y composición de la leche en cabras murciano-granadinas. PR: Pequeños Rumiantes 2002, 3, 26–30. [Google Scholar]
  14. Ryan, D.P.; Greenwood, P.L.; Nicholls, P.J. Effect of caprine arthritis-encephalitis virus infection on milk cell count and N-acetyl-beta-glucosaminidase activity in dairy goats. J. Dairy Res. 1993, 60, 299–306. [Google Scholar] [CrossRef]
  15. Martinez-Navalon, B.; Peris, C.; Gomez, E.A.; Peris, B.; Roche, M.L.; Caballero, C.; Goyena, E.; Berriatua, E. Quantitative estimation of the impact of caprine arthritis encephalitis virus infection on milk production by dairy goats. Vet. J. 2013. [Google Scholar] [CrossRef]
  16. Christodoulopoulos, G. Milk Production and Milk Fat Content in Commercial Karagouniko Breed Flocks Infected by Maedi Disease in Greece. In Proceedings of the Sixth International Sheep Veterinary Congress, Hersonissos Crete, Greece, 17–21 June 2005; pp. 344–345.
  17. Leitner, G.; Silanikove, N.; Merin, U. Estimate of milk and curd yield loss of sheep and goats with intrammamary infection and its relation to somatic cell count. Small Rum. Res. 2008, 74, 221–225. [Google Scholar] [CrossRef]
  18. Broughton-Neiswanger, L.E.; White, S.N.; Knowles, D.P.; Mousel, M.R.; Lewis, G.S.; Herndon, D.R.; Herrmann-Hoesing, L.M. Non-maternal transmission is the major mode of ovine lentivirus transmission in a ewe flock: A molecular epidemiology study. Infect. Genet. Evol. 2010, 10, 998–1007. [Google Scholar] [CrossRef]
  19. Leginagoikoa, I.; Daltabuit-Test, M.; Alvarez, V.; Arranz, J.; Juste, R.A.; Amorena, B.; de Andres, D.; Lujan, L.L.; Badiola, J.J.; Berriatua, E. Horizontal Maedi-Visna virus (MVV) infection in adult dairy-sheep raised under varying MVV-infection pressures investigated by ELISA and PCR. Res. Vet. Sci. 2006, 80, 235–241. [Google Scholar] [CrossRef]
  20. Leginagoikoa, I.; Minguijon, E.; Juste, R.A.; Barandika, J.; Amorena, B.; de Andres, D.; Badiola, J.J.; Lujan, L.; Berriatua, E. Effects of housing on the incidence of visna/maedi virus infection in sheep flocks. Res. Vet. Sci. 2010, 88, 415–421. [Google Scholar] [CrossRef]
  21. Alvarez, V.; Arranz, J.; Daltabuit-Test, M.; Leginagoikoa, I.; Juste, R.A.; Amorena, B.; de Andres, D.; Lujan, L.L.; Badiola, J.J.; Berriatua, E. Relative contribution of colostrum from Maedi-Visna virus (MVV) infected ewes to MVV-seroprevalence in lambs. Res. Vet. Sci. 2005, 78, 237–243. [Google Scholar] [CrossRef]
  22. Leginagoikoa, I.; Juste, R.A.; Barandika, J.; Amorena, B.; De Andres, D.; Lujan, L.; Badiola, J.; Berriatua, E. Extensive rearing hinders Maedi-Visna Virus (MVV) infection in sheep. Vet. Res. 2006, 37, 767–778. [Google Scholar] [CrossRef]
  23. Sigurdsson, B.; Grimsson, H.; Palsson, P.A. Maedi, a chronic, progressive infection of sheep’s lungs. J. Infect. Dis. 1952, 90, 233–241. [Google Scholar] [CrossRef]
  24. Blacklaws, B.A.; Berriatua, E.; Torsteinsdottir, S.; Watt, N.J.; de Andres, D.; Klein, D.; Harkiss, G.D. Transmission of small ruminant lentiviruses. Vet. Microbiol. 2004, 101, 199–208. [Google Scholar] [CrossRef]
  25. Ramirez, H.; Glaria, I.; de Andres, X.; Martinez, H.A.; Hernandez, M.M.; Reina, R.; Iraizoz, E.; Crespo, H.; Berriatua, E.; Vazquez, J.; et al. Recombinant small ruminant lentivirus subtype B1 in goats and sheep of imported breeds in Mexico. Vet. J. 2011, 190, 169–172. [Google Scholar] [CrossRef]
  26. Straub, O.C. Maedi-Visna virus infection in sheep. History and present knowledge. Comp. Immunol. Microbiol. Infect. Dis. 2004, 27, 1–5. [Google Scholar] [CrossRef]
  27. Shah, C.; Boni, J.; Huder, J.B.; Vogt, H.R.; Muhlherr, J.; Zanoni, R.; Miserez, R.; Lutz, H.; Schupbach, J. Phylogenetic analysis and reclassification of caprine and ovine lentiviruses based on 104 new isolates: Evidence for regular sheep-to-goat transmission and worldwide propagation through livestock trade. Virology 2004, 319, 12–26. [Google Scholar] [CrossRef]
  28. Gifford, R.J. Viral evolution in deep time: Lentiviruses and mammals. Trends Genet. 2012, 28, 89–100. [Google Scholar] [CrossRef]
  29. Muz, D.; Oguzoglu, T.C.; Rosati, S.; Reina, R.; Bertolotti, L.; Burgu, I. First molecular characterization of visna/maedi viruses from naturally infected sheep in Turkey. Arch. Virol. 2012. [Google Scholar] [CrossRef]
  30. Bertolotti, L.; Mazzei, M.; Puggioni, G.; Carrozza, M.L.; Dei Giudici, S.; Muz, D.; Juganaru, M.; Patta, C.; Tolari, F.; Rosati, S. Characterization of new small ruminant lentivirus subtype B3 suggests animal trade within the Mediterranean Basin. J. Gen. Virol. 2011, 92, 1923–1929. [Google Scholar] [CrossRef]
  31. Foley, B.T. An Overview of the Molecular Phylogeny of Lentiviruses. In HIV Sequence Compedium; Kuiken, C., Foley, B., Freed, E., Hahn, B., Korber, B., Marx, P.A., McCutchan, F., Mellors, J.W., Mullins, J.I., Sodroski, J., Wolinksy, S., Eds.; Theoretical Biology and Biophysics Group, Los Alamos National Laboratory: Los Alamos, NM, USA, 2000; pp. 35–43. [Google Scholar]
  32. Eigen, M. Selforganization of matter and the evolution of biological macromolecules. Naturwissenschaften 1971, 58, 465–523. [Google Scholar] [CrossRef]
  33. Ojosnegros, S.; Perales, C.; Mas, A.; Domingo, E. Quasispecies as a matter of fact: Viruses and beyond. Virus Res. 2011, 162, 203–215. [Google Scholar] [CrossRef]
  34. Korber, B.; Gaschen, B.; Yusim, K.; Thakallapally, R.; Kesmir, C.; Detours, V. Evolutionary and immunological implications of contemporary HIV-1 variation. Br. Med. Bull. 2001, 58, 19–42. [Google Scholar] [CrossRef]
  35. Pasick, J. Maedi-visna virus and caprine arthritis-encephalitis virus: Distinct species or quasispecies and its implications for laboratory diagnosis. Can. J. Vet. Res. 1998, 62, 241–244. [Google Scholar]
  36. Smyth, R.P.; Davenport, M.P.; Mak, J. The origin of genetic diversity in HIV-1. Virus Res. 2012, 169, 415–429. [Google Scholar] [CrossRef]
  37. Roberts, J.D.; Bebenek, K.; Kunkel, T.A. The accuracy of reverse transcriptase from HIV-1. Science 1988, 242, 1171–1173. [Google Scholar]
  38. Vartanian, J.P.; Meyerhans, A.; Asjo, B.; Wain-Hobson, S. Selection, recombination, and G----A hypermutation of human immunodeficiency virus type 1 genomes. J. Virol. 1991, 65, 1779–1788. [Google Scholar]
  39. Bishop, K.N.; Holmes, R.K.; Sheehy, A.M.; Davidson, N.O.; Cho, S.J.; Malim, M.H. Cytidine deamination of retroviral DNA by diverse APOBEC proteins. Curr. Biol. 2004, 14, 1392–1396. [Google Scholar] [CrossRef]
  40. Dang, Y.; Wang, X.; Esselman, W.J.; Zheng, Y.H. Identification of APOBEC3DE as another antiretroviral factor from the human APOBEC family. J. Virol. 2006, 80, 10522–10533. [Google Scholar] [CrossRef]
  41. Terai, C.; Carson, D.A. Pyrimidine nucleotide and nucleic acid synthesis in human monocytes and macrophages. Exp. Cell. Res. 1991, 193, 375–381. [Google Scholar] [CrossRef]
  42. Turelli, P.; Guiguen, F.; Mornex, J.F.; Vigne, R.; Querat, G. dUTPase-minus caprine arthritis-encephalitis virus is attenuated for pathogenesis and accumulates G-to-A substitutions. J. Virol. 1997, 71, 4522–4530. [Google Scholar]
  43. Wain-Hobson, S.; Sonigo, P.; Guyader, M.; Gazit, A.; Henry, M. Erratic G→A hypermutation within a complete caprine arthritis-encephalitis virus (CAEV) provirus. Virology 1995, 209, 297–303. [Google Scholar] [CrossRef]
  44. Juganaru, M.; Reina, R.; Bertolotti, L.; Stella, M.C.; Profiti, M.; Armentano, M.; Bollo, E.; Amorena, B.; Rosati, S. In vitro properties of small ruminant lentivirus genotype E. Virology 2011, 410, 88–95. [Google Scholar] [CrossRef]
  45. Reina, R.; Grego, E.; Bertolotti, L.; De Meneghi, D.; Rosati, S. Genome analysis of small-ruminant lentivirus genotype E: A caprine lentivirus with natural deletions of the dUTPase subunit, vpr-like accessory gene, and 70-base-pair repeat of the U3 region. J. Virol. 2009, 83, 1152–1155. [Google Scholar] [CrossRef]
  46. Reina, R.; Bertolotti, L.; Dei Giudici, S.; Puggioni, G.; Ponti, N.; Profiti, M.; Patta, C.; Rosati, S. Small ruminant lentivirus genotype E is widespread in Sarda goat. Vet. Microbiol. 2010, 144, 24–31. [Google Scholar] [CrossRef] [Green Version]
  47. Tormo-Mas, M.A.; Mir, I.; Shrestha, A.; Tallent, S.M.; Campoy, S.; Lasa, I.; Barbe, J.; Novick, R.P.; Christie, G.E.; Penades, J.R. Moonlighting bacteriophage proteins derepress staphylococcal pathogenicity islands. Nature 2012, 465, 779–782. [Google Scholar]
  48. Sheehy, A.M.; Gaddis, N.C.; Choi, J.D.; Malim, M.H. Isolation of a human gene that inhibits HIV-1 infection and is suppressed by the viral Vif protein. Nature 2002, 418, 646–650. [Google Scholar] [CrossRef]
  49. Andresdottir, V.; Skraban, R.; Matthiasdottir, S.; Lutley, R.; Agnarsdottir, G.; Thorsteinsdottir, H. Selection of antigenic variants in maedi-visna virus infection. J. Gen. Virol. 2002, 83, 2543–2551. [Google Scholar]
  50. Andresdottir, V. Evidence for recombination in the envelope gene of maedi-visna virus. Virus Genes 2003, 27, 5–9. [Google Scholar] [CrossRef]
  51. Jetzt, A.E.; Yu, H.; Klarmann, G.J.; Ron, Y.; Preston, B.D.; Dougherty, J.P. High rate of recombination throughout the human immunodeficiency virus type 1 genome. J. Virol. 2000, 74, 1234–1240. [Google Scholar]
  52. Jolly, P.E.; Narayan, O. Evidence for interference, coinfections, and intertypic virus enhancement of infection by ovine-caprine lentiviruses. J. Virol. 1989, 63, 4682–4688. [Google Scholar]
  53. Hu, W.S.; Temin, H.M. Retroviral recombination and reverse transcription. Science 1990, 250, 1227–1233. [Google Scholar]
  54. Hu, W.S.; Temin, H.M. Genetic consequences of packaging two RNA genomes in one retroviral particle: Pseudodiploidy and high rate of genetic recombination. Proc. Natl. Acad. Sci. USA 1990, 87, 1556–1560. [Google Scholar] [CrossRef]
  55. Pisoni, G.; Bertoni, G.; Puricelli, M.; Maccalli, M.; Moroni, P. Demonstration of coinfection with and recombination by caprine arthritis-encephalitis virus and maedi-visna virus in naturally infected goats. J. Virol. 2007, 81, 4948–4955. [Google Scholar] [CrossRef]
  56. Zanoni, R.G. Phylogenetic analysis of small ruminant lentiviruses. J. Gen. Virol. 1998, 79, 1951–1961. [Google Scholar]
  57. Peeters, M.; Esu-Williams, E.; Vergne, L.; Montavon, C.; Mulanga-Kabeya, C.; Harry, T.; Ibironke, A.; Lesage, D.; Patrel, D.; Delaporte, E. Predominance of subtype A and G HIV type 1 in Nigeria, with geographical differences in their distribution. AIDS Res. Hum. Retroviruses 2000, 16, 315–325. [Google Scholar] [CrossRef]
  58. Adams, D.S.; Oliver, R.E.; Ameghino, E.; DeMartini, J.C.; Verwoerd, D.W.; Houwers, D.J.; Waghela, S.; Gorham, J.R.; Hyllseth, B.; Dawson, M.; et al. Global survey of serological evidence of caprine arthritis-encephalitis virus infection. Vet. Rec. 1984, 115, 493–495. [Google Scholar]
  59. Narayan, O.; Clements, J.E.; Strandberg, J.D.; Cork, L.C.; Griffin, D.E. Biological characterization of the virus causing leukoencephalitis and arthritis in goats. J. Gen. Virol. 1980, 50, 69–79. [Google Scholar] [CrossRef]
  60. Saltarelli, M.; Querat, G.; Konings, D.A.; Vigne, R.; Clements, J.E. Nucleotide sequence and transcriptional analysis of molecular clones of CAEV which generate infectious virus. Virology 1990, 179, 347–364. [Google Scholar] [CrossRef]
  61. Gjerset, B.; Jonassen, C.M.; Rimstad, E. Natural transmission and comparative analysis of small ruminant lentiviruses in the Norwegian sheep and goat populations. Virus Res. 2007, 125, 153–161. [Google Scholar] [CrossRef]
  62. Gjerset, B.; Storset, A.K.; Rimstad, E. Genetic diversity of small-ruminant lentiviruses: Characterization of Norwegian isolates of Caprine arthritis encephalitis virus. J. Gen. Virol. 2006, 87, 573–580. [Google Scholar] [CrossRef]
  63. Qu, J.; Guo, W.; Zhao, L.; Sheng, R.; Xiang, W. The gene cloning and sequence analysis of the whole genome of Caprine Arthritis Encephalitis virus (CAEV) GANSU strain. Chin. J. Virol. 2005, 21, 389–392. [Google Scholar]
  64. Huang, J.; Sun, Y.; Liu, Y.; Xiao, H.; Zhuang, S. Development of a loop-mediated isothermal amplification method for rapid detection of caprine arthritis-encephalitis virus proviral DNA. Arch. Virol. 2012, 157, 1463–1469. [Google Scholar] [CrossRef]
  65. Glaria, I.; Reina, R.; Crespo, H.; de Andres, X.; Ramirez, H.; Biescas, E.; Perez, M.M.; Badiola, J.; Lujan, L.; Amorena, B.; et al. Phylogenetic analysis of SRLV sequences from an arthritic sheep outbreak demonstrates the introduction of CAEV-like viruses among Spanish sheep. Vet. Microbiol. 2009, 138, 156–162. [Google Scholar] [CrossRef]
  66. Querat, G.; Audoly, G.; Sonigo, P.; Vigne, R. Nucleotide sequence analysis of SA-OMVV, a visna-related ovine lentivirus: Phylogenetic history of lentiviruses. Virology 1990, 175, 434–447. [Google Scholar] [CrossRef]
  67. Andresson, O.S.; Elser, J.E.; Tobin, G.J.; Greenwood, J.D.; Gonda, M.A.; Georgsson, G.; Andresdottir, V.; Benediktsdottir, E.; Carlsdottir, H.M.; Mantyla, E.O. Nucleotide sequence and biological properties of a pathogenic proviral molecular clone of neurovirulent visna virus. Virology 1993, 193, 89–105. [Google Scholar] [CrossRef]
  68. Sonigo, P.; Alizon, M.; Staskus, K.; Klatzmann, D.; Cole, S.; Danos, O.; Retzel, E.; Tiollais, P.; Haase, A.; Wain-Hobson, S. Nucleotide sequence of the visna lentivirus: Relationship to the AIDS virus. Cell 1985, 42, 369–382. [Google Scholar] [CrossRef]
  69. Staskus, K.A.; Retzel, E.F.; Lewis, E.D.; Silsby, J.L.; St Cyr, S.; Rank, J.M.; Wietgrefe, S.W.; Haase, A.T.; Cook, R.; Fast, D.; et al. Isolation of replication-competent molecular clones of visna virus. Virology 1991, 181, 228–240. [Google Scholar] [CrossRef]
  70. Sargan, D.R.; Bennet, I.D.; Cousens, C.; Roy, D.J.; Blacklaws, B.A.; Dalziel, R.G.; Watt, N.J.; McConnell, I. Nucleotide sequence of EV1, a British isolate of maedi-visna virus. J. Gen. Virol. 1991, 72, 1893–1903. [Google Scholar] [CrossRef]
  71. Fevereiro, M.T.; Barros, S.S. Caracterização biológica e molecular de um lentivírus de ovino isolado em Portugal. RPCV 2004, 99, 27–39. [Google Scholar]
  72. Karr, B.M.; Chebloune, Y.; Leung, K.; Narayan, O. Genetic characterization of two phenotypically distinct North American ovine lentiviruses and their possible origin from caprine arthritis-encephalitis virus. Virology 1996, 225, 1–10. [Google Scholar] [CrossRef]
  73. Glaria, I.; Reina, R.; Ramirez, H.; de Andres, X.; Crespo, H.; Jauregui, P.; Salazar, E.; Lujan, L.; Perez, M.M.; Benavides, J.; et al. Visna/Maedi virus genetic characterization and serological diagnosis of infection in sheep from a neurological outbreak. Vet. Microbiol. 2012, 155, 137–146. [Google Scholar] [CrossRef]
  74. Castro, R.S.; Greenland, T.; Leite, R.C.; Gouveia, A.; Mornex, J.F.; Cordier, G. Conserved sequence motifs involving the tat reading frame of Brazilian caprine lentiviruses indicate affiliations to both caprine arthritis-encephalitis virus and visna-maedi virus. J. Gen. Virol. 1999, 80, 1583–1589. [Google Scholar]
  75. Feitosaa, L.A.L.V.; de Silva Teixeira, M.F.; Rizaldo, P.R.; Silva da Cunhac, R.M.; Santos Limad, J.P.M.; Andriolib, A.; Medeiros, D.T.V.; Pessoa de Meloa, V.S.; Nunes Pinheiroa, D.C.S. Phylogenetic analysis of small ruminant lentiviruses from Northern Brazil. Small Rum. Res. 2010, 94, 205–209. [Google Scholar] [CrossRef]
  76. Ravazzolo, A.P.; Reischak, D.; Peterhans, E.; Zanoni, R. Phylogenetic analysis of small ruminant lentiviruses from Southern Brazil. Virus Res. 2001, 79, 117–123. [Google Scholar] [CrossRef]
  77. L'Homme, Y.; Ouardani, M.; Levesque, V.; Bertoni, G.; Simard, C.; Pisoni, G. Molecular characterization and phylogenetic analysis of small ruminant lentiviruses isolated from Canadian sheep and goats. Virol. J. 2011, 8. [Google Scholar] [CrossRef] [Green Version]
  78. Laamanen, I.; Jakava-Viljanen, M.; Sihvonen, L. Genetic characterization of maedi-visna virus (MVV) detected in Finland. Vet. Microbiol. 2007, 122, 357–365. [Google Scholar] [CrossRef]
  79. Leroux, C.; Chastang, J.; Greenland, T.; Mornex, J.F. Genomic heterogeneity of small ruminant lentiviruses: Existence of heterogeneous populations in sheep and of the same lentiviral genotypes in sheep and goats. Arch. Virol. 1997, 142, 1125–1137. [Google Scholar] [CrossRef]
  80. Germain, K.; Valas, S. Distribution and heterogeneity of small ruminant lentivirus envelope subtypes in naturally infected French sheep. Virus Res. 2006, 120, 156–162. [Google Scholar] [CrossRef]
  81. Valas, S.; Benoit, C.; Guionaud, C.; Perrin, G.; Mamoun, R.Z. North American and French caprine arthritis-encephalitis viruses emerge from ovine maedi-visna viruses. Virology 1997, 237, 307–318. [Google Scholar] [CrossRef]
  82. Angelopoulou, K.; Karanikolaou, K.; Papanastasopoulou, M.; Koumpati-Artopiou, M.; Vlemmas, I.; Papadopoulos, O.; Koptopoulos, G. First partial characterisation of small ruminant lentiviruses from Greece. Vet. Microbiol. 2005, 109, 1–9. [Google Scholar] [CrossRef]
  83. Rolland, M.; Mooney, J.; Valas, S.; Perrin, G.; Mamoun, R.Z. Characterisation of an Irish caprine lentivirus strain—SRLV phylogeny revisited. Virus Res. 2002, 85, 29–39. [Google Scholar] [CrossRef]
  84. Konishi, M.; Tsuduku, S.; Haritani, M.; Murakami, K.; Tsuboi, T.; Kobayashi, C.; Yoshikawa, K.; Kimura, K.M.; Sentsui, H. An epidemic of caprine arthritis encephalitis in Japan: Isolation of the virus. J. Vet. Med. Sci. 2004, 66, 911–917. [Google Scholar] [CrossRef]
  85. Zanoni, R.G.; Nauta, I.M.; Pauli, U.; Peterhans, E. Expression in Escherichia coli and sequencing of the coding region for the capsid protein of Dutch maedi-visna virus strain ZZV 1050: Application of recombinant protein in enzyme-linked immunosorbent assay for the detection of caprine and ovine lentiviruses. J. Clin. Microbiol. 1991, 29, 1290–1294. [Google Scholar]
  86. Kuzmak, J.; Rola, M.; Gallay, K.; Chebloune, Y. Molecular characterization of lentiviruses from goats from Poland based on gag gene sequence analysis. Comp. Immunol. Microbiol. Infect. Dis. 2007, 30, 211–223. [Google Scholar] [CrossRef]
  87. Olech, M.; Rachid, A.; Croise, B.; Kuzmak, J.; Valas, S. Genetic and antigenic characterization of small ruminant lentiviruses circulating in Poland. Virus Res. 2011, 163, 528–536. [Google Scholar]
  88. Baryshnikova, E.I.; Malogolovkin, A.S.; Kolbasova, O.L.; Tsybanov, S. Comparative characteristics of the biological properties of small ruminant lentiviruses. Vopr. Virusol. 2011, 56, 42–45. [Google Scholar]
  89. Kuhar, U.; Barlic-Maganja, D.; Grom, J. Phylogenetic analysis of small ruminant lentiviruses detected in Slovenia. Vet. Microbiol. 2012. [Google Scholar] [CrossRef]
  90. Park, J.E.; Son, S.Y.; Shin, H.J. Sequence comparison on gag gene of caprine arthritis encephalitis virus from Korea. Virus Genes 2010, 41, 99–101. [Google Scholar] [CrossRef]
  91. Leroux, C.; Vuillermoz, S.; Mornex, J.F.; Greenland, T. Genomic heterogeneity in the pol region of ovine lentiviruses obtained from bronchoalveolar cells of infected sheep from France. J. Gen. Virol. 1995, 76, 1533–1537. [Google Scholar] [CrossRef]
  92. Leroux, C.; Greenland, T.; Mornex, J.F. Molecular characterization of field isolates of lentiviruses of small ruminants. AIDS Res. Hum. Retroviruses 1996, 12, 427–429. [Google Scholar] [CrossRef]
  93. Pisoni, G.; Bertoni, G.; Manarolla, G.; Vogt, H.R.; Scaccabarozzi, L.; Locatelli, C.; Moroni, P. Genetic analysis of small ruminant lentiviruses following lactogenic transmission. Virology 2010, 407, 91–99. [Google Scholar] [CrossRef]
  94. Grego, E.; Bertolotti, L.; Quasso, A.; Profiti, M.; Lacerenza, D.; Muz, D.; Rosati, S. Genetic characterization of small ruminant lentivirus in Italian mixed flocks: Evidence for a novel genotype circulating in a local goat population. J. Gen. Virol. 2007, 88, 3423–3427. [Google Scholar] [CrossRef]
  95. Shah, C.; Huder, J.B.; Boni, J.; Schonmann, M.; Muhlherr, J.; Lutz, H.; Schupbach, J. Direct evidence for natural transmission of small-ruminant lentiviruses of subtype A4 from goats to sheep and vice versa. J. Virol. 2004, 78, 7518–7522. [Google Scholar] [CrossRef]
  96. Chebloune, Y.; Sheffer, D.; Karr, B.M.; Stephens, E.; Narayan, O. Restrictive type of replication of ovine/caprine lentiviruses in ovine fibroblast cell cultures. Virology 1996, 222, 21–30. [Google Scholar] [CrossRef]
  97. Banks, K.L.; Adams, D.S.; McGuire, T.C.; Carlson, J. Experimental infection of sheep by caprine arthritis-encephalitis virus and goats by progressive pneumonia virus. Am. J. Vet. Res. 1983, 44, 2307–2311. [Google Scholar]
  98. Dickson, J.; Ellis, T. Experimental caprine retrovirus infection in sheep. Vet. Rec. 1989, 125. [Google Scholar] [CrossRef]
  99. Oliver, R.; Cathcart, A.; McNiven, R.; Poole, W.; Robati, G. Transmission of caprine arthritis encephalitis virus to sheep. NZ Vet. J. 1984, 32, 199–200. [Google Scholar]
  100. Oliver, R.; Cathcart, A.; McNiven, R.; Poole, W.; Robati, G. Infection of lambs with caprine arthritis encephalitis virus by feeding milk from infected goats. Vet. Rec. 1985, 116, 83. [Google Scholar]
  101. Oliver, R.E.; McNiven, R.A.; Julian, A.F.; Poole, W.S. Experimental infection of sheep and goats with caprine arthritis-encephalitis virus. NZ Vet. J. 1982, 30, 158–159. [Google Scholar] [CrossRef]
  102. Smith, V.W.; Dickson, J.; Coackley, W.; Carman, H. Response of merino sheep to inoculation with a caprine retrovirus. Vet. Rec. 1985, 117, 61–63. [Google Scholar]
  103. Pisoni, G.; Quasso, A.; Moroni, P. Phylogenetic analysis of small-ruminant lentivirus subtype B1 in mixed flocks: Evidence for natural transmission from goats to sheep. Virology 2005, 339, 147–152. [Google Scholar] [CrossRef]
  104. Cheevers, W.P.; Knowles, D.P., Jr.; Norton, L.K. Neutralization-resistant antigenic variants of caprine arthritis-encephalitis lentivirus associated with progressive arthritis. J. Infect. Dis. 1991, 164, 679–685. [Google Scholar] [CrossRef]
  105. Hotzel, I.; Cheevers, W. Differential receptor usage of small ruminant lentiviruses in ovine and caprine cells: Host range but not cytopathic phenotype is determined by receptor usage. Virology 2002, 301, 21–31. [Google Scholar] [CrossRef]
  106. Bruett, L.; Clements, J.E. Functional murine leukemia virus vectors pseudotyped with the visna virus envelope show expanded visna virus cell tropism. J. Virol. 2001, 75, 11464–11473. [Google Scholar] [CrossRef]
  107. Hotzel, I.; Cheevers, W.P. Host range of small-ruminant lentivirus cytopathic variants determined with a selectable caprine arthritis-encephalitis virus pseudotype system. J. Virol. 2001, 75, 7384–7391. [Google Scholar] [CrossRef]
  108. Valas, S.; Benoit, C.; Baudry, C.; Perrin, G.; Mamoun, R.Z. Variability and immunogenicity of caprine arthritis-encephalitis virus surface glycoprotein. J. Virol. 2000, 74, 6178–6185. [Google Scholar] [CrossRef]
  109. Patel, J.R.; Heldens, J.G.; Bakonyi, T.; Rusvai, M. Important mammalian veterinary viral immunodiseases and their control. Vaccine 2012, 30, 1767–1781. [Google Scholar] [CrossRef]
  110. Bieniasz, P.D. An overview of intracellular interactions between immunodeficiency viruses and their hosts. AIDS 2012, 26, 1243–1254. [Google Scholar] [CrossRef]
  111. Jauregui, P.; Crespo, H.; Glaria, I.; Lujan, L.; Contreras, A.; Rosati, S.; de Andres, D.; Amorena, B.; Towers, G.J.; Reina, R. Ovine TRIM5alpha can restrict visna/maedi virus. J. Virol. 2012, 86, 9504–9509. [Google Scholar] [CrossRef]
  112. LaRue, R.S.; Lengyel, J.; Jonsson, S.R.; Andresdottir, V.; Harris, R.S. Lentiviral Vif degrades the APOBEC3Z3/APOBEC3H protein of its mammalian host and is capable of cross-species activity. J. Virol. 2010, 84, 8193–8201. [Google Scholar] [CrossRef]
  113. Korber, B.T.; Kunstman, K.J.; Patterson, B.K.; Furtado, M.; Mcevilly, M.M.; Levy, R.; Wolinsky, S.M. Genetic differences between blood- and brain-derived viral sequences from Human Immunodeficiency Virus Type-1 infected patients: Evidence of conserved elements in the V3 region of the Envelope protein of brain-derived sequences. J. Virol. 1994, 68, 7467–7481. [Google Scholar]
  114. Keys, B.; Karis, J.; Fadeel, B.; Valentin, A.; Norkrans, G.; Haqberg, L.; Chiodi, F. V3 sequences of paired HIV-1 isolates from blood and cerebrospinal fluid cluster according to host and show variation related to the clinical stage of disease. Virology 1993, 196, 475–483. [Google Scholar] [CrossRef]
  115. Liu, P.; Hudson, L.C.; Tompkins, M.B.; Vahlenkamp, T.W.; Meeker, R.B. Compartmentalization and evolution of feline immunodeficiency virus between the central nervous system and periphery following intracerebroventricular or systemic inoculation. J. Neurovirol. 2006, 12, 307–321. [Google Scholar]
  116. Pisoni, G.; Moroni, P.; Turin, L.; Bertoni, G. Compartmentalization of small ruminant lentivirus between blood and colostrum in infected goats. Virology 2007, 369, 119–130. [Google Scholar] [CrossRef]
  117. Ramirez, H.; Reina, R.; Bertolotti, L.; Cenoz, A.; Hernandez, M.M.; San Roman, B.; Glaria, I.; de Andres, X.; Crespo, H.; Jauregui, P.; et al. Study of compartmentalization in the visna clinical form of small ruminant lentivirus infection in sheep. BMC Vet. Res. 2012, 8, e8. [Google Scholar] [CrossRef] [Green Version]
  118. Chen, M.F.; Westmoreland, S.; Ryzhova, E.V.; Martín-García, J.; Soldan, S.S.; Lackner, A.; González-Scarano, F. Simian immunodeficiency virus envelope compartmentalizes in brain regions independent of neuropathology. J. Neurovirol. 2006, 12, 73–89. [Google Scholar] [CrossRef]
  119. Zárate, S.; Kosakovsky, P.S.L.; Shapshak, P.; Frost, S.D.W. Comparative study of methods for detecting sequence compartmentalization in human immunodeficiency virus type 1. J. Virol. 2007, 81, 6643–6651. [Google Scholar] [CrossRef]
  120. Abbate, I.; Cappiello, G.; Longo, R.; Ursitti, A.; Spano, A.; Calcaterra, S.; Dianzani, F.; Antinori, A.; Capobianchi, M.R. Cell membrane proteins and quasispecies compartmentalization of CSF and plasma HIV-1 from AIDS patients with neurological disorders. Infect. Genet. Evol. 2005, 5, 247–253. [Google Scholar] [CrossRef]
  121. Burkala, E.; He, J.; West, J.; Wood, C.; Petito, C. Compartmentalization of HIV-1 in the central nervous system: Role of the choroid plexus. AIDS 2005, 19, 675–684. [Google Scholar] [CrossRef]
  122. Philpott, S.; Burger, H.; Tsoukas, C.; Foley, B.; Anastos, K.; Kitchen, C.; Weiser, B. Human immunodeficiency virus type 1 genomic RNA sequences in the female genital tract and blood: Compartmentalization and intrapatient recombination. J. Virol. 2005, 79, 353–363. [Google Scholar]
  123. Saag, M.S.; Hahn, B.H.; Gibbons, J.; Li, Y. Extensive variation of human immunodeficiency virus type-1 in vivo. Nature 1988, 334, 440–444. [Google Scholar] [CrossRef]
  124. Shankarappa, R.; Margolick, J.B.; Gange, S.J.; Rodrigo, A.G.; Upchurch, D.; Farzadegan, H.; Gupta, P.; Rinaldo, C.R.; Learn, G.H.; He, X.; et al. Consistent viral evolutionary changes associated with the progression of human immunodeficiency virus type 1 infection. J. Virol. 1999, 73, 10489–10502. [Google Scholar]
  125. Si-Mohamed, A.; Kazatchkine, M.D.; Heard, I.; Goujon, C.; Prazuck, T.; Aymard, G.; Cessot, G.; Kuo, Y.H.; Bernard, M.C.; Diquet, B.; et al. Selection of drug-resistant variants in the female genital tract of human immunodeficiency virus type 1-infected women receiving antiretroviral therapy. J. Infect. Dis. 2000, 182, 112–122. [Google Scholar] [CrossRef]
  126. Smit, T.; Brew, B.; Tourtellotte, W.; Morgello, S.; Gelman, B.; Saksena., N. Independent evolution of human immunodeficiency virus (HIV) drug resistance mutations in diverse areas of the brain in HIV-infected patients, with and without dementia, on antiretroviral treatment. J. Virol. 2004, 78, 10133–10148. [Google Scholar] [CrossRef]
  127. Pepin, M.; Vitu, C.; Russo, P.; Mornex, J.F.; Peterhans, E. Maedi-visna virus infection in sheep: A review. Vet. Res. 1998, 29, 341–367. [Google Scholar]
  128. Barros, S.C.; Andresdottir, V.; Fevereiro, M. Cellular specificity and replication rate of Maedi Visna virus in vitro can be controlled by LTR sequences. Arch. Virol. 2005, 150, 201–213. [Google Scholar] [CrossRef]
  129. Oskarsson, T.; Hreggvidsdottir, H.S.; Agnarsdottir, G.; Matthiasdottir, S.; Ogmundsdottir, M.H.; Jonsson, S.R.; Georgsson, G.; Ingvarsson, S.; Andresson, O.S.; Andresdottir, V. Duplicated sequence motif in the long terminal repeat of maedi-visna virus extends cell tropism and is associated with neurovirulence. J. Virol. 2007, 81, 4052–4057. [Google Scholar] [CrossRef]
  130. Murphy, B.; McElliott, V.; Vapniarsky, N.; Oliver, A.; Rowe, J. Tissue tropism and promoter sequence variation in caprine arthritis encephalitis virus infected goats. Virus Res. 2010, 151, 177–184. [Google Scholar] [CrossRef]
  131. Bertoni, G.; Hertig, C.; Zahno, M.L.; Vogt, H.R.; Dufour, S.; Cordano, P.; Peterhans, E.; Cheevers, W.P.; Sonigo, P.; Pancino, G. B-cell epitopes of the envelope glycoprotein of caprine arthritis-encephalitis virus and antibody response in infected goats. J. Gen. Virol. 2000, 81, 2929–2940. [Google Scholar]
  132. Hotzel, I.; Cheevers, W.P. Sequence similarity between the envelope surface unit (SU) glycoproteins of primate and small ruminant lentiviruses. Virus Res. 2000, 69, 47–54. [Google Scholar] [CrossRef]
  133. Mwaengo, D.M.; Grant, R.F.; DeMartini, J.C.; Carlson, J.O. Envelope glycoprotein nucleotide sequence and genetic characterization of North American ovine lentiviruses. Virology 1997, 238, 135–144. [Google Scholar] [CrossRef]
  134. Blacklaws, B.; Bird, P.; McConnell, I. Early events in infection of lymphoid tissue by a lentivirus, maedi-visna. Trends Microbiol. 1995, 3, 434–440. [Google Scholar] [CrossRef]
  135. Leroux, C.; Mornex, J.F. Retroviral infections in sheep and the associated diseases. Small Rum. Res. 2008, 76, 68–76. [Google Scholar] [CrossRef]
  136. Palsson, P.A. Maedi and visna in sheep. Front. Biol. 1976, 44, 17–43. [Google Scholar]
  137. Torsteinsdottir, S.; Agnarsdottir, G.; Matthiasdottir, S.; Rafnar, B.; Andresdottir, V.; Andresson, O.S.; Staskus, K.; Petursson, G.; Palsson, P.A.; Georgsson, G. In vivo and in vitro infection with two different molecular clones of visna virus. Virology 1997, 229, 370–380. [Google Scholar] [CrossRef]
  138. Andresdottir, V.; Tang, X.; Agnarsdottir, G.; Andresson, O.S.; Georgsson, G.; Skraban, R.; Torsteinsdottir, S.; Rafnar, B.; Benediktsdottir, E.; Matthiasdottir, S.; et al. Biological and genetic differences between lung- and brain-derived isolates of maedi-visna virus. Virus Genes 1998, 16, 281–293. [Google Scholar] [CrossRef]
  139. Narayan, O.; Wolinsky, J.S.; Clements, J.E.; Strandberg, J.D.; Griffin, D.E.; Cork, L.C. Slow virus replication: The role of macrophages in the persistence and expression of visna viruses of sheep and goats. J. Gen. Virol. 1982, 59, 345–356. [Google Scholar] [CrossRef]
  140. Ryan, S.; Tiley, L.; McConnell, I.; Blacklaws, B. Infection of dendritic cells by the Maedi-Visna lentivirus. J. Virol. 2000, 74, 10096–10103. [Google Scholar] [CrossRef]
  141. Gendelman, H.E.; Narayan, O.; Kennedy-Stoskopf, S.; Kennedy, P.G.; Ghotbi, Z.; Clements, J.E.; Stanley, J.; Pezeshkpour, G. Tropism of sheep lentiviruses for monocytes: Susceptibility to infection and virus gene expression increase during maturation of monocytes to macrophages. J. Virol. 1986, 58, 67–74. [Google Scholar]
  142. Guiguen, F.; Asso, J. Inhibition of CAEV expression in goats experimentally and naturally infected. Dev. Biol. Stand. 1990, 72, 235–239. [Google Scholar]
  143. Narayan, O.; Kennedy-Stoskopf, S.; Sheffer, D.; Griffin, D.E.; Clements, J.E. Activation of caprine arthritis-encephalitis virus expression during maturation of monocytes to macrophages. Infect. Immun. 1983, 41, 67–73. [Google Scholar]
  144. Peluso, R.; Haase, A.T.; Stowring, L.; Edwards, M.; Ventura, P. A Trojan horse mechanism for the spread of visna virus in monocytes. Nucleic Acids Res. 1988, 16, 4116–4173. [Google Scholar]
  145. Barros, S.C.; Ramos, F.; Duarte, M.; Fagulha, T.; Cruz, B.; Fevereiro, M. Genomic characterization of a slow/low maedi visna virus. Virus Genes 2004, 29, 199–210. [Google Scholar] [CrossRef]
  146. Querat, G.; Barban, V.; Sauze, N.; Filippi, P.; Vigne, R.; Russo, P.; Vitu, C. Highly lytic and persistent lentiviruses naturally present in sheep with progressive pneumonia are genetically distinct. J. Virol. 1984, 52, 672–679. [Google Scholar]
  147. Singh, D.K.; Chebloune, Y.; Mselli-Lakhal, L.; Karr, B.M.; Narayan, O. Ovine lentivirus-infected macrophages mediate productive infection in cell types that are not susceptible to infection with cell-free virus. J. Gen. Virol. 1999, 80, 1437–1444. [Google Scholar]
  148. Sihvonen, L.; Veijalainen, P. Kinetics of maedi virus production in sheep choroid plexus cells. Vet. Microbiol. 1981, 6, 1–8. [Google Scholar] [CrossRef]
  149. Sigurdardottir, B.; Thormar, H. Isolation of a viral agent from the lungs of sheep affected with maedi. J. Infect. Dis. 1964, 114, 55–60. [Google Scholar] [CrossRef]
  150. Belov, L.; Whalley, J.M. Virus-specific polypeptides of caprine arthritis-encephalitis virus recognized by monoclonal antibodies to virion proteins p24 and p14. J. Gen. Virol. 1988, 69, 1097–1103. [Google Scholar] [CrossRef]
  151. Kirkland, P.D.; Batty, E.M. Caprine arthritis-encephalitis virus: An efficient method for the large scale production of serological antigens. J. Virol. Methods 1987, 16, 323–326. [Google Scholar] [CrossRef]
  152. Dahlberg, J.E.; Gaskin, J.M.; Perk, K. Morphological and immunological comparison of caprine arthritis encephalitis and ovine progressive pneumonia viruses. J. Virol. 1981, 39, 914–919. [Google Scholar]
  153. Lechat, E.; Milhau, N.; Brun, P.; Bellaton, C.; Greenland, T.; Mornex, J.F.; Le Jan, C. Goat endothelial cells may be infected in vitro by transmigration of caprine arthritis-encephalitis virus-infected leucocytes. Vet. Immunol. Immunopathol. 2005, 104, 257–263. [Google Scholar] [CrossRef]
  154. Milhau, N.; Bellaton, C.; Balleydier, S.; Gaonach, M.; Le Jan, C. In vitro infection of aortic endothelial cells by caprine arthritis encephalitis virus enhances in vitro transmigration of peripheral blood leukocytes and modulates their phenotypic expression. Vet. Res. 2003, 34, 273–284. [Google Scholar] [CrossRef]
  155. Mselli-Lakhal, L.; Guiguen, F.; Fornazero, C.; Favier, C.; Durand, J.; Grezel, D.; Moussa, A.; Mornex, J.F.; Chebloune, Y. Immortalized goat milk epithelial cell lines replicate CAEV at high level. Vet. Res. 2001, 32, 429–440. [Google Scholar] [CrossRef]
  156. Bolea, R.; Monleon, E.; Carrasco, L.; Vargas, A.; de Andres, D.; Amorena, B.; Badiola, J.J.; Lujan, L. Maedi-visna virus infection of ovine mammary epithelial cells. Vet. Res. 2006, 37, 133–144. [Google Scholar] [CrossRef]
  157. Lamara, A.; Fieni, F.; Mselli-Lakhal, L.; Tainturier, D.; Chebloune, Y. Efficient replication of caprine arthritis-encephalitis virus in goat granulosa cells. Virus Res. 2001, 79, 165–172. [Google Scholar] [CrossRef]
  158. Lamara, A.; Fieni, F.; Mselli-Lakhal, L.; Tainturier, D.; Chebloune, Y. Epithelial cells from goat oviduct are highly permissive for productive infection with caprine arthritis-encephalitis virus (CAEV). Virus Res. 2002, 87, 69–77. [Google Scholar] [CrossRef]
  159. Baszler, T.V.; Harwood, W.G.; Lester, K.L.; Davis, W.C.; Knowles, D.P. Characterization of caprine microglial cells and in vitro infection with caprine arthritis-encephalitis lentivirus. Lab. Invest. 1994, 70, 933–943. [Google Scholar]
  160. Adebayo, I.A.; Olaleye, O.D.; Awoniyi, T.A. Affinity (tropism) of caprine arthritis encephalitis virus for brain cells. Afr. J. Med. Med. Sci. 2008, 39, 89–93. [Google Scholar]
  161. Angelopoulou, K.; Brellou, G.D.; Vlemmas, I. Detection of maedi-visna virus in the kidneys of naturally infected sheep. J. Comp. Pathol. 2006, 134, 329–335. [Google Scholar] [CrossRef]
  162. Brellou, G.D.; Angelopoulou, K.; Poutahidis, T.; Vlemmas, I. Detection of maedi-visna virus in the liver and heart of naturally infected sheep. J. Comp. Pathol. 2007, 136, 27–35. [Google Scholar] [CrossRef]
  163. Capucchio, M.T.; Sanna, E.; Sanna, M.P.; Farigu, S.; Minelli, R.; Guarda, F. Maedi-visna virus detection in ovine third eyelids. J. Comp. Pathol. 2003, 129, 37–43. [Google Scholar] [CrossRef]
  164. Carrozza, M.L.; Mazzei, M.; Bandecchi, P.; Arispici, M.; Tolari, F. In situ PCR-associated immunohistochemistry identifies cell types harbouring the Maedi-Visna virus genome in tissue sections of sheep infected naturally. J. Virol. Methods 2003, 107, 121–127. [Google Scholar] [CrossRef]
  165. Sanchez, A.B.; Rodriguez, D.; Garzon, A.; Amorena, B.; Esteban, M.; Rodriguez, J.R. Visna/maedi virus Env protein expressed by a vaccinia virus recombinant induces cell-to-cell fusion in cells of different origins in the apparent absence of Env cleavage: Role of glycosylation and of proteoglycans. Arch. Virol. 2002, 147, 2377–2392. [Google Scholar] [CrossRef]
  166. Lyall, J.W.; Solanky, N.; Tiley, L.S. Restricted species tropism of maedi-visna virus strain EV-1 is not due to limited receptor distribution. J. Gen. Virol. 2000, 81, 2919–2927. [Google Scholar]
  167. Mselli-Lakhal, L.; Favier, C.; Leung, K.; Guiguen, F.; Grezel, D.; Miossec, P.; Mornex, J.F.; Narayan, O.; Querat, G.; Chebloune, Y. Lack of functional receptors is the only barrier that prevents caprine arthritis-encephalitis virus from infecting human cells. J. Virol. 2000, 74, 8343–8348. [Google Scholar] [CrossRef]
  168. Bruett, L.; Barber, S.A.; Clements, J.E. Characterization of a membrane-associated protein implicated in visna virus binding and infection. Virology 2000, 271, 132–141. [Google Scholar] [CrossRef]
  169. Dalziel, R.G.; Hopkins, J.; Watt, N.J.; Dutia, B.M.; Clarke, H.A.; McConnell, I. Identification of a putative cellular receptor for the lentivirus visna virus. J. Gen. Virol. 1991, 72, 1905–1911. [Google Scholar] [CrossRef]
  170. Hovden, A.O.; Sommerfelt, M.A. The influence of CD4 and CXCR4 on maedi-visna virus-induced syncytium formation. APMIS 2002, 110, 697–708. [Google Scholar] [CrossRef]
  171. Crane, S.E.; Buzy, J.; Clements, J.E. Identification of cell membrane proteins that bind visna virus. J. Virol. 1991, 65, 6137–6143. [Google Scholar]
  172. Hotzel, I.; Cheevers, W.P. A maedi-visna virus strain K1514 receptor gene is located in sheep chromosome 3p and the syntenic region of human chromosome 2. J. Gen. Virol. 2002, 83, 1759–1764. [Google Scholar]
  173. Crespo, H.; Reina, R.; Glaria, I.; Ramírez, H.; de Andrés, X.; Jáuregui, P.; Luján, L.; Martinez-Pomares, L.; Amorena, B.; de Andrés, D.F. Identification of the ovine mannose receptor and its possible role in Visna/Maedi virus infection. BMCVet. Res. 2011, 42, e28. [Google Scholar]
  174. Crespo, H.; Jáuregui, P.; Glaria, I.; Sanjose, L.; Polledo, L.; García-Marin, J.F.; Luján, L.; de Andrés, D.; Amorena, B.; Reina, R. Mannose receptor may be involved in small ruminant lentivirus pathogenesis. BMCVet. Res. 2012, 43, e43. [Google Scholar]
  175. Reina, R.; Berriatua, E.; Lujan, L.; Juste, R.; Sanchez, A.; de Andrés, D.; Amorena, B. Prevention strategies against small ruminant lentiviruses: An update. Vet. J. 2009, 182, 31–37. [Google Scholar] [CrossRef]
  176. Biescas, E.; Preziuso, S.; Bulgin, M.; DeMartini, J.C. Ovine lentivirus-associated leucomyelitis in naturally infected North American sheep. J. Comp. Pathol. 2005, 132, 107–116. [Google Scholar] [CrossRef]
  177. Storset, A.K.; Teig, A.; Rimstad, E. Detection of caprine arthritis-encephalitis virus RNA in macrophages by in situ hybridization using fluorescein-labelled single-stranded RNA probes. Vet. Microbiol. 1996, 52, 25–35. [Google Scholar] [CrossRef]
  178. Knowles, D.P., Jr.; Evermann, J.F.; Shropshire, C.; VanderSchalie, J.; Bradway, D.; Gezon, H.M.; Cheevers, W.P. Evaluation of agar gel immunodiffusion serology using caprine and ovine lentiviral antigens for detection of antibody to caprine arthritis-encephalitis virus. J. Clin. Microbiol. 1994, 32, 243–245. [Google Scholar]
  179. De Andres, D.; Klein, D.; Watt, N.J.; Berriatua, E.; Torsteinsdottir, S.; Blacklaws, B.A.; Harkiss, G.D. Diagnostic tests for small ruminant lentiviruses. Vet. Microbiol. 2005, 107, 49–62. [Google Scholar] [CrossRef]
  180. Knowles, D.P., Jr. Laboratory diagnostic tests for retrovirus infections of small ruminant. Vet. Clin. North. Am. Food Anim. Pract. 1997, 13, 1–11. [Google Scholar]
  181. Herrmann-Hoesing, L.M. Diagnostic assays used to control small ruminant lentiviruses. J. Vet. Diagn. Invest. 2010, 22, 843–855. [Google Scholar] [CrossRef]
  182. Gogolewski, R.P.; Adams, D.S.; McGuire, T.C.; Banks, K.L.; Cheevers, W.P. Antigenic cross-reactivity between caprine arthritis-encephalitis, visna and progressive pneumonia viruses involves all virion-associated proteins and glycoproteins. J. Gen. Virol. 1985, 66, 1233–1240. [Google Scholar] [CrossRef]
  183. Cheevers, W.P.; Cordery-Cotter, R.; McGuire, T.C.; DeMartini, J.C. Neutralizing antibody responses and evolution of antigenic variants in monozygotic twin lambs infected with phenotypically distinct ovine lentiviruses. Virology 1999, 258, 382–388. [Google Scholar] [CrossRef]
  184. Herrmann, L.M.; Cheevers, W.P.; Marshall, K.L.; McGuire, T.C.; Hutton, M.M.; Lewis, G.S.; Knowles, D.P. Detection of serum antibodies to ovine progressive pneumonia virus in sheep by using a caprine arthritis-encephalitis virus competitive-inhibition enzyme-linked immunosorbent assay. Clin. Diagn. Lab. Immunol. 2003, 10, 862–865. [Google Scholar]
  185. Herrmann, L.M.; Cheevers, W.P.; McGuire, T.C.; Adams, D.S.; Hutton, M.M.; Gavin, W.G.; Knowles, D.P. Competitive-inhibition enzyme-linked immunosorbent assay for detection of serum antibodies to caprine arthritis-encephalitis virus: Diagnostic tool for successful eradication. Clin. Diagn. Lab. Immunol. 2003, 10, 267–271. [Google Scholar]
  186. Herrmann-Hoesing, L.M.; Broughton-Neiswanger, L.E.; Gouine, K.C.; White, S.N.; Mousel, M.R.; Lewis, G.S.; Marshall, K.L.; Knowles, D.P. Evaluation of a caprine arthritis-encephalitis virus/maedi-visna virus indirect enzyme-linked immunosorbent assay in the serological diagnosis of ovine progressive pneumonia virus in U.S. sheep. Clin. Vaccine Immunol. 2010, 17, 307–310. [Google Scholar] [CrossRef]
  187. Carrozza, M.L.; Mazzei, M.; Lacerenza, D.; Del Chiaro, L.; Giammarioli, M.; Marini, C.; Rutili, D.; Rosati, S.; Tolari, F. Seroconversion against SU5 derived synthetic peptides in sheep experimentally infected with different SRLV genotypes. Vet. Microbiol. 2009, 137, 369–374. [Google Scholar] [CrossRef]
  188. Reina, R.; Grego, E.; Profiti, M.; Glaria, I.; Robino, P.; Quasso, A.; Amorena, B.; Rosati, S. Development of specific diagnostic test for small ruminant lentivirus genotype E. Vet. Microbiol. 2009, 138, 251–257. [Google Scholar] [CrossRef]
  189. Brinkhof, J.; van Maanen, C. Evaluation of five enzyme-linked immunosorbent assays and an agar gel immunodiffusion test for detection of antibodies to small ruminant lentiviruses. Clin. Vaccine Immunol. 2007, 14, 1210–1214. [Google Scholar] [CrossRef]
  190. Toft, N.; Akerstedt, J.; Tharaldsen, J.; Hopp, P. Evaluation of three serological tests for diagnosis of Maedi-Visna virus infection using latent class analysis. Vet. Microbiol. 2007, 120, 77–86. [Google Scholar] [CrossRef]
  191. Ramirez, H.; Roman, B.S.; Glaria, I.; Reina, R.; Hernandez, M.M.; de Andres, X.; Crespo, H.; Hichou, B.; Cianca, S.; Goni, C.; et al. Antibody-based diagnosis of small ruminant lentivirus infection in seminal fluid. Theriogenology 2009, 72, 1085–1096. [Google Scholar] [CrossRef]
  192. Zanoni, R.G.; Vogt, H.R.; Pohl, B.; Bottcher, J.; Bommeli, W.; Peterhans, E. An ELISA based on whole virus for the detection of antibodies to small-ruminant lentiviruses. Zentralbl. Veterinarmed. 1994, 41, 662–669. [Google Scholar]
  193. Saman, E.; Van Eynde, G.; Lujan, L.; Extramiana, B.; Harkiss, G.; Tolari, F.; Gonzalez, L.; Amorena, B.; Watt, N.; Badiola, J. A new sensitive serological assay for detection of lentivirus infections in small ruminants. Clin. Diagn. Lab. Immunol. 1999, 6, 734–740. [Google Scholar]
  194. Lacerenza, D.; Giammarioli, M.; Grego, E.; Marini, C.; Profiti, M.; Rutili, D.; Rosati, S. Antibody response in sheep experimentally infected with different small ruminant lentivirus genotypes. Vet. Immunol. Immunopathol. 2006, 112, 264–271. [Google Scholar] [CrossRef]
  195. Cardinaux, L.; Zahno, M.L.; Deubelbeiss, M.; Zanoni, R.; Vogt, H.R.; Bertoni, G. Virological and phylogenetic characterization of attenuated small ruminant lentivirus isolates eluding efficient serological detection. Vet. Microbiol. 2013, 162, 572–581. [Google Scholar] [CrossRef]
  196. Grego, E.; Bertolotti, L.; Carrozza, M.L.; Profiti, M.; Mazzei, M.; Tolari, F.; Rosati, S. Genetic and antigenic characterization of the matrix protein of two genetically distinct ovine lentiviruses. Vet. Microbiol. 2005, 106, 179–185. [Google Scholar] [CrossRef]
  197. Grego, E.; Profiti, M.; Giammarioli, M.; Giannino, L.; Rutili, D.; Woodall, C.; Rosati, S. Genetic heterogeneity of small ruminant lentiviruses involves immunodominant epitope of capsid antigen and affects sensitivity of single-strain-based immunoassay. Clin. Diagn. Lab. Immunol. 2002, 9, 828–832. [Google Scholar]
  198. Mordasini, F.; Vogt, H.R.; Zahno, M.L.; Maeschli, A.; Nenci, C.; Zanoni, R.; Peterhans, E.; Bertoni, G. Analysis of the antibody response to an immunodominant epitope of the envelope glycoprotein of a lentivirus and its diagnostic potential. J. Clin. Microbiol. 2006, 44, 981–991. [Google Scholar] [CrossRef]
  199. De Andrés, X.; Ramírez, H.; Bertolotti, L.; San Román, B.; Glaria, I.; Crespo, H.; Jáuregui, P.; Minguijón, E.; Juste, R.; Leginagoikoa, I.; et al. An insight into a combination of ELISA strategies to diagnose Small Ruminant Lentivirus infections. Vet. Immunol. Immunopathol. 2013, 152, 277–288. [Google Scholar] [CrossRef]
  200. Brinkhof, J.M.; Houwers, D.J.; Moll, L.; Dercksen, D.; van Maanen, C. Diagnostic performance of ELISA and PCR in identifying SRLV-infected sheep and goats using serum, plasma and milk samples and in early detection of infection in dairy flocks through bulk milk testing. Vet. Microbiol. 2010, 142, 193–198. [Google Scholar] [CrossRef]
  201. Motha, M.X.; Ralston, J.C. Evaluation of ELISA for detection of antibodies to CAEV in milk. Vet. Microbiol. 1994, 38, 359–367. [Google Scholar] [CrossRef]
  202. Plaza, M.; Sánchez, A.; Corrales, J.C.; De la Fe, C.; Contreras, A. Caprine arthritis encephalitis virus diagnosed by ELISA in lactating goats using milk samples. Small Rum. Res. 2009, 81, 189–192. [Google Scholar] [CrossRef]
  203. Brinkhof, J.M.A.; Houwers, D.J.; van Maanen, C. Development of a sample pooling strategy for the serodiagnosis of small ruminant lentiviral infections using the ELITEST-MVV ELISA. Small Rum. Res. 2007, 70, 194–199. [Google Scholar] [CrossRef]
  204. Reina, R.; Glaria, I.; Cianca, S.; Crespo, H.; Andres, X.D.; Goni, C.; Lasarte, J.M.; Lujan, L.; Amorena, B.; de Andres, D.F. Use of small ruminant lentivirus-infected rams for artificial insemination. Vet. J. 2010, 189, 106–107. [Google Scholar]
  205. Gudmundsson, B.; Bjarnadottir, H.; Kristjansdottir, S.; Jonsson, J.J. Quantitative assays for maedi-visna virus genetic sequences and mRNA’s based on RT-PCR with real-time FRET measurements. Virology 2003, 307, 135–142. [Google Scholar] [CrossRef]
  206. Zhang, Z.; Watt, N.J.; Hopkins, J.; Harkiss, G.; Woodall, C.J. Quantitative analysis of maedi-visna virus DNA load in peripheral blood monocytes and alveolar macrophages. J. Virol. Methods 2000, 86, 13–20. [Google Scholar] [CrossRef]
  207. Brinkhof, J.M.; van Maanen, C.; Wigger, R.; Peterson, K.; Houwers, D.J. Specific detection of small ruminant lentiviral nucleic acid sequences located in the proviral long terminal repeat and leader-gag regions using real-time polymerase chain reaction. J. Virol. Methods 2008, 147, 338–344. [Google Scholar] [CrossRef]
  208. Carrozza, M.L.; Mazzei, M.; Bandecchi, P.; Fraisier, C.; Perez, M.; Suzan-Monti, M.; de Andres, D.; Amorena, B.; Rosati, S.; Andresdottir, V.; et al. Development and comparison of strain specific gag and pol real-time PCR assays for the detection of Visna/maedi virus. J. Virol. Methods 2010, 165, 161–167. [Google Scholar] [CrossRef]
  209. Herrmann-Hoesing, L.M.; White, S.N.; Lewis, G.S.; Mousel, M.R.; Knowles, D.P. Development and validation of an ovine progressive pneumonia virus quantitative PCR. Clin. Vaccine Immunol. 2007, 14, 1274–1278. [Google Scholar] [CrossRef]
  210. Haase, A.T.; Retzel, E.F.; Staskus, K.A. Amplification and detection of lentiviral DNA inside cells. Proc. Natl. Acad. Sci. USA 1990, 87, 4971–4975. [Google Scholar] [CrossRef]
  211. Zanoni, R.; Pauli, U.; Peterhans, E. Detection of caprine arthritis-encephalitis- and maedi-visna viruses using the polymerase chain reaction. Experientia 1990, 46, 316–319. [Google Scholar] [CrossRef]
  212. Leroux, C.; Lerondelle, C.; Chastang, J.; Mornex, J.F. RT-PCR detection of lentiviruses in milk or mammary secretions of sheep or goats from infected flocks. Vet. Res. 1997, 28, 115–121. [Google Scholar]
  213. Alvarez, V.; Daltabuit-Test, M.; Arranz, J.; Leginagoikoa, I.; Juste, R.A.; Amorena, B.; de Andres, D.; Lujan, L.; Badiola, J.J.; Berriatua, E. PCR detection of colostrum-associated Maedi-Visna virus (MVV) infection and relationship with ELISA-antibody status in lambs. Res. Vet. Sci. 2006, 80, 226–234. [Google Scholar] [CrossRef]
  214. Barlough, J.; East, N.; Rowe, J.D.; Van Hoosear, K.; DeRock, E.; Bigornia, L.; Rimstad, E. Double-nested polymerase chain reaction for detection of caprine arthritis-encephalitis virus proviral DNA in blood, milk, and tissues of infected goats. J. Virol. Methods 1994, 50, 101–113. [Google Scholar] [CrossRef]
  215. Celer, V., Jr.; Celer, V.; Nejedla, E.; Bertoni, G.; Peterhans, E.; Zanoni, R.G. The detection of proviral DNA by semi-nested polymerase chain reaction and phylogenetic analysis of Czech Maedi-Visna isolates based on gag gene sequences. J. Vet. Med. B Infect. Dis. Vet. Public Health 2000, 47, 203–215. [Google Scholar] [CrossRef]
  216. Eltahir, Y.M.; Dovas, C.I.; Papanastassopoulou, M.; Koumbati, M.; Giadinis, N.; Verghese-Nikolakaki, S.; Koptopoulos, G. Development of a semi-nested PCR using degenerate primers for the generic detection of small ruminant lentivirus proviral DNA. J. Virol. Methods 2006, 135, 240–246. [Google Scholar] [CrossRef]
  217. Modolo, J.R.; Stachissini, A.V.M.; Padovani, C.R.; Araujo Júnior, J.P.; Castro, R.S.; Ravazzolo, A.P.; Leite, B.L.S. PCR associated with agar gel immunodiffusion assay improve caprine arthritis-encephalitis (CAEV) control. Small Rum. Res. 2009, 81, 18–20. [Google Scholar] [CrossRef]
  218. Reddy, P.G.; Sapp, W.J.; Heneine, W. Detection of caprine arthritis-encephalitis virus by polymerase chain reaction. J. Clin. Microbiol. 1993, 31, 3042–3043. [Google Scholar]
  219. Rimstad, E.; East, N.E.; Torten, M.; Higgins, J.; DeRock, E.; Pedersen, N.C. Delayed seroconversion following naturally acquired caprine arthritis-encephalitis virus infection in goats. Am. J. Vet. Res. 1993, 54, 1858–1862. [Google Scholar]
  220. Zanoni, R.G.; Cordano, P.; Nauta, I.M.; Peterhans, E. PCR for the detection of lentiviruses from small ruminants. Schweiz Arch. Tierheilkd. 1996, 138, 93–98. [Google Scholar]
  221. Leginagoikoa, I.; Minguijon, E.; Berriatua, E.; Juste, R.A. Improvements in the detection of small ruminant lentivirus infection in the blood of sheep by PCR. J. Virol. Methods 2009, 156, 145–149. [Google Scholar] [CrossRef]
  222. Extramiana, A.B.; González, L.; Cortabarría, N.; García, M.; Juste, R.A. Evaluation of a PCR technique for the detection of Maedi-Visna proviral DNA in blood, milk and tissue samples of naturally infected sheep. Small Rum. Res. 2002, 44, 109–118. [Google Scholar] [CrossRef]
  223. Ali Al Ahmad, M.Z.; Fieni, F.; Pellerin, J.L.; Guiguen, F.; Cherel, Y.; Chatagnon, G.; Bouzar, A.B.; Chebloune, Y. Detection of viral genomes of caprine arthritis-encephalitis virus (CAEV) in semen and in genital tract tissues of male goat. Theriogenology 2008, 69, 473–480. [Google Scholar] [CrossRef]
  224. Travassos, C.; Benoit, C.; Valas, S.; da Silva, A.; Perrin, G. Caprine artritis-encephalitis virus in semen of naturally infected bucks. Small Rum. Res. 1999, 32, 101–106. [Google Scholar] [CrossRef]
  225. Brodie, S.J.; Pearson, L.D.; Snowder, G.D.; DeMartini, J.C. Host-virus interaction as defined byamplification of viral DNA and serology in lentivirus-infected sheep. Arch. Virol. 1993, 130, 413–428. [Google Scholar] [CrossRef]
  226. Zanoni, R.; Nauta, I.M.; Kuhnert, P.; Pauli, U.; Pohl, B.; Peterhans, E. Genomic heterogeneity of small ruminant lentiviruses detected by PCR. Vet. Microbiol. 1992, 33, 241–351. [Google Scholar]
  227. Rosati, S.; Kwang, J.; Keen, J.E. Genome analysis of North American small ruminant lentiviruses by polymerase chain reaction and restriction enzyme analysis. J. Vet. Diagn. Invest. 1995, 7, 437–443. [Google Scholar] [CrossRef]
  228. Johnson, L.K.; Meyer, A.L.; Zink, M.C. Detection of ovine lentivirus in seronegative sheep by in situ hybridization, PCR, and cocultivation with susceptible cells. Clin. Immunol. Immunopathol. 1992, 65, 254–260. [Google Scholar] [CrossRef]
  229. Wagter, L.H.; Jansen, A.; Bleumink-Pluym, N.M.; Lenstra, J.A.; Houwers, D.J. PCR detection of lentiviral GAG segment DNA in the white blood cells of sheep and goats. Vet. Res. Commun. 1998, 22, 355–362. [Google Scholar] [CrossRef]
  230. Karanikolaou, K.; Angelopoulou, K.; Papanastasopoulou, M.; Koumpati-Artopiou, M.; Papadopoulos, O.; Koptopoulos, G. Detection of small ruminant lentiviruses by PCR and serology tests in field samples of animals from Greece. Small Rum. Res. 2005, 58, 181–187. [Google Scholar] [CrossRef]
  231. Brodie, S.J.; Marcom, K.A.; Pearson, L.D.; Anderson, B.C.; de la Concha-Bermejillo, A.; Ellis, J.A.; DeMartini, J.C. Effects of virus load in the pathogenesis of lentivirus-induced lymphoid interstitial pneumonia. J. Infect. Dis. 1992, 166, 531–541. [Google Scholar] [CrossRef]
  232. Clavijo, A.; Thorsen, J. Application of polymerase chain reaction for the diagnosis of caprine arthritis-encephalitis. Small Rum. Res. 1996, 22, 69–77. [Google Scholar] [CrossRef]
  233. Kuzmak, J.; Kedziora, A.; Rola, M.; Kozaczynska, B.; Chebloune, Y.; Gallay, K. Evaluation of PCR and PCR/hybridation method for the detection of caprine arthritis-encephalitis virus infection in goats. Bull. Vet. Inst. Pulawy 2003, 47, 293–300. [Google Scholar]
  234. Herrmann-Hoesing, L.M.; Palmer, G.H.; Knowles, D.P. Evidence of proviral clearance following postpartum transmission of an ovine lentivirus. Virology 2007, 362, 226–234. [Google Scholar] [CrossRef]
  235. Barquero, N.; Arjona, A.; Domenech, A.; Toural, C.; de las Heras, A.; Fernandez-Garayzabal, J.F.; Ruiz-Santa Quiteria, J.A.; Gomez-Lucia, E. Diagnostic performance of PCR and ELISA on blood and milk samples and serological survey for small ruminant lentiviruses in central Spain. Vet. Rec. 2011, 168. [Google Scholar] [CrossRef]
  236. Brinkhof, J.M.; Moll, L.; van Maanen, C.; Houwers, D.J. Use of serology and polymerase chain reaction for the rapid eradication of small ruminant lentivirus infections from a sheep flock: A case report. Res. Vet. Sci. 2010, 88, 41–43. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Ramírez, H.; Reina, R.; Amorena, B.; Andrés, D.D.; Martínez, H.A. Small Ruminant Lentiviruses: Genetic Variability, Tropism and Diagnosis. Viruses 2013, 5, 1175-1207. https://doi.org/10.3390/v5041175

AMA Style

Ramírez H, Reina R, Amorena B, Andrés DD, Martínez HA. Small Ruminant Lentiviruses: Genetic Variability, Tropism and Diagnosis. Viruses. 2013; 5(4):1175-1207. https://doi.org/10.3390/v5041175

Chicago/Turabian Style

Ramírez, Hugo, Ramsés Reina, Beatriz Amorena, Damián De Andrés, and Humberto A. Martínez. 2013. "Small Ruminant Lentiviruses: Genetic Variability, Tropism and Diagnosis" Viruses 5, no. 4: 1175-1207. https://doi.org/10.3390/v5041175

Article Metrics

Back to TopTop