Next Article in Journal
Interference of HTLV-1 Tax Protein with Cell Polarity Regulators: Defining the Subcellular Localization of the Tax-DLG1 Interaction
Next Article in Special Issue
Immunotherapeutic Potential of Oncolytic H-1 Parvovirus: Hints of Glioblastoma Microenvironment Conversion towards Immunogenicity
Previous Article in Journal
Pressure for Pattern-Specific Intertypic Recombination between Sabin Polioviruses: Evolutionary Implications
Previous Article in Special Issue
Protoparvovirus Interactions with the Cellular DNA Damage Response
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Human Protoparvoviruses

1
Department of Virology, University of Helsinki, Helsinki 00290, Finland
2
Helsinki University Hospital, Helsinki 00290, Finland
*
Author to whom correspondence should be addressed.
Submission received: 31 October 2017 / Revised: 17 November 2017 / Accepted: 19 November 2017 / Published: 22 November 2017
(This article belongs to the Special Issue Protoparvoviruses: Friends or Foes?)

Abstract

:
Next-generation sequencing and metagenomics have revolutionized the discovery of novel viruses. In recent years, three novel protoparvoviruses have been discovered in fecal samples of humans: bufavirus (BuV) in 2012, tusavirus (TuV) in 2014, and cutavirus (CuV) in 2016. BuV has since been studied the most, disclosing three genotypes that also represent serotypes. Besides one nasal sample, BuV DNA has been found exclusively in diarrheal feces, but not in non-diarrheal feces, suggesting a causal relationship. According to both geno- and seroprevalences, BuV appears to be the most common of the three novel protoparvoviruses, whereas TuV DNA has been found in only a single fecal sample, with antibody detection being equally rare. Moreover, the TuV sequence is closer to those of non-human protoparvoviruses, and so the evidence of TuV being a human virus is thus far insufficient. Interestingly, besides in feces, CuV has also been detected in skin biopsies of patients with cutaneous T-cell lymphoma and a patient with melanoma, while all other skin samples have tested PCR negative. Even if preliminary disease associations exist, the full etiological roles of these viruses in human disease are yet to be resolved.

Graphical Abstract

1. Introduction

Parvoviruses are small ssDNA viruses that infect a diverse group of animals; both vertebrates (subfamily Parvovirinae) and invertebrates (subfamily Densovirinae). Their non-enveloped capsid structure makes these viruses stable and very resistant to inactivation. Human parvoviruses are generally not easy to grow in standard cell culture systems, limiting the discovery of novel human parvoviruses in the past. However, in recent years great progress has been made in genetic sciences, where virion enrichment, host DNA depletion, sequence-independent amplification, next generation sequencing (NGS), and metagenomic analysis with diverse bioinformatic pipelines have revolutionized the discovery of novel viruses, including human parvoviruses.
For 30 years, human parvovirus B19 of the Erythroparvovirus genus was the only known indisputably human-pathogenic parvovirus, causing erythema infectiosum, arthropathies, anemias, and fetal death, while the highly prevalent adeno-associated viruses (AAV) of the Dependoparvovirus genus are apathogenic [1,2]. In 2005, two novel parvoviruses were discovered in humans: parvovirus 4 (PARV4) of the Tetraparvovirus genus in blood, and human bocavirus 1 (HBoV1) of the Bocaparvovirus genus in pediatric respiratory samples [3,4]. PARV4 is mainly detected in injecting drug users and hemophiliacs in the western world, but has not been associated with any specific symptoms, whereas HBoV1 causes mild to life-threatening respiratory tract infections in children [2]. Three more bocaviruses (HBoV2–4) have been discovered in stool, without clear clinical associations [5,6,7]. During the past six years, three novel parvoviruses have been described in humans, all belonging to the Protoparvovirus genus: bufavirus (BuV), tusavirus (TuV), and cutavirus (CuV) [8,9,10]. Although BuV has been associated with gastrointestinal symptoms and CuV is under study for its possible role in skin cancers, the etiological roles of these viruses in human disease are yet to be resolved. Here we review what is known about these new protoparvoviruses.

2. Bufavirus

2.1. Discovery and Taxonomy

Bufavirus was identified in 2012 in viral metagenomic analysis of fecal samples from diarrheic children in Burkina Faso [8]. The obtained sequence reads from one sample showed significant similarities to parvovirus sequences, and after filling the gaps between the NGS reads by PCR, a nearly full-length sequence of 4921 bp (GenBank # JQ918261) was constructed showing <39% and <31% identity of the NS1 (non-structural protein 1) and VP1 protein (viral capsid protein 1), respectively, to previously known protoparvoviruses. A unique middle open reading frame (ORF) of 130 amino acids was also identified that did not show similarity with any other parvovirus sequences (Figure 1a). Further PCR analysis of 98 rotavirus antigen-negative diarrheal fecal samples from Burkina Faso showed the presence of BuV DNA in 4/98 (4.1%) samples [8]. The complete protein-coding regions were determined for all BuV DNA isolates. Within the VP2 (viral capsid protein 2) region, one of the four isolates shared less than 73% amino-acid identity with the other isolates, whereas the non-structural NS1 regions were nearly identical. Subsequently, a second BuV genotype, bufavirus 2 (BuV2), was identified (GenBank # JX027297). The authors further studied 100 fecal samples from Chilean children with diarrhea and 63 samples from Tunisian children with non-polio acute flaccid paralysis, but only one BuV sequence was found among the latter [8].
Two years later, a third BuV genotype, BuV3 (GenBank # AB847987-9), was discovered in the diarrheal feces of 3/393 Bhutanese children [11]. All three current BuV genotypes contain an ATP- or GTP-binding Walker loop, phospholipase A2 (PLA2), two conserved replication initiator motifs, a glycine-rich sequence, and the same splice sites. Among the two major ORFs, the three genotypes are more divergent within the capsid proteins, sharing only 71–78% amino-acid identity within VP1 and 65–73% within VP2 [12]. In contrast, the NS1 regions of the three BuVs are nearly identical (94–96% amino acid identity). According to the International Committee on Taxonomy of Viruses (ICTV), parvoviruses in a single species encode NS1 proteins of >85% amino acid sequence identity [13]. Therefore, all three BuV genotypes can be phylogenetically classified into one species: Primate protoparvovirus 1, within the Protoparvovirus genus (Figure 1b).

2.2. PCR-Based Epidemiology

BuV DNA has been mainly searched for in feces, the sample type that the virus was originally identified in. BuV DNA has been detected in the diarrheal samples of children in Burkina Faso, Tunisia, Bhutan, Thailand, Turkey, China, and Finland, and of adults in Finland, the Netherlands, Thailand, and China (Table 1) [8,11,16,17,18,19,20,21,22]. In all studies, the BuV DNA prevalence in feces has been low, ranging from 0.3% to 4.1% (Table 1), and when reported, the viral titre was always low [16,22]. Recently, a partial sequence matching BuV3 was found by NGS in diarrheal feces from Peru in South America, expanding the geographic locations where BuV DNA has been detected (Table 1) [23]. Prevalence variation may occur even within one country; in China, the BuV DNA prevalence was 1.7% (9/520) among samples from the General Hospital of Beijing, whereas all samples (n = 1357) from the Children’s hospital of Chongqing were BuV DNA-negative [20]. So far, all BuV-DNA findings have been of genotypes 1 or 3, except the one original BuV2-DNA sequence from the single child in Burkina Faso [8,11,16,17,18,19,20,21,23], showing that BuV1 and 3 are geographically widespread.
In six studies, temporal clustering of BuV DNA detection to cold weather was observed as all or nearly all BuV DNA-positive samples were collected during September–April [11,18,19,20,21,22]. In four studies, negative controls were included; all fecal samples from non-diarrheic patients or healthy individuals (total n = 1855) were BuV DNA-negative (Table 1) [18,19,20,22]. However, in only two studies the non-diarrheal fecal samples were collected during the same time period as the BuV DNA-positive diarrheal samples [19,22]. This is of importance as in most studies BuV DNA was not detected throughout the year, and very often not at all during certain years.
Besides in feces, BuV DNA has been studied in nasal swabs of children and in cerebrospinal fluid (CSF) samples of adults and children [22,24]. Nasal swabs (n = 955) were collected in Finland from children with acute gastroenteritis (AGE, n = 172), acute respiratory tract infection (ARTI, n = 545), or both (n = 238), and BuV DNA was detected at low copy number in the nasal swab of one child with both AGE and ARTI (Table 1) [22]. However, the corresponding fecal sample of this child was BuV DNA-negative. In a study of central nervous system infections in Turkey, the CSF samples were collected from patients with a febrile disease and/or central nervous system infection with presumed viral etiology [24]. However, no BuV DNA sequences were detected by nested PCR in the CSF from these 93 Turkish children and 33 adults (Table 1).

2.3. Antibody-Based Epidemiology

A BuV enzyme immunoassay (EIA) has been developed to elucidate the prevalence of BuV1–3 IgG antibodies in various populations [22]. Humans have been shown to induce BuV IgG towards each of the three genotypes (Table 2). The BuV IgG prevalence was low in both Finnish-born adults (3.1%) and children (3.1%). Interestingly, among Asian-born university or hospital staff members working in Finland, BuV IgG was present in 5/12 of the tested subjects, suggesting that BuV is more prevalent in Asia [22]. The study moreover highlighted that the three BuV genotypes also represent serotypes, as no antibody cross-reactivity in EIA was observed. Further studies with larger cohorts from different parts of the world are ongoing to reveal the global BuV IgG prevalences in healthy populations and in different patient cohorts. In addition, no patients with serologically confirmed acute BuV infections have yet been reported. Such cases would be essential in order to broaden the knowledge of possible symptoms and diseases that BuV might cause.

2.4. Bufavirus-Like Animal Viruses

Since the discovery of human BuV, several animal species have been shown to have their own BuV-like viruses. The viruses phylogenetically closest to human BuVs have been detected in captive rhesus monkeys in the USA with a nucleotide identity of 77% in the VP1 gene, and in wild baboons in Zambia (Figure 1) [25,26]. In both primate species, BuV-like viruses were detected either in the blood or in the spleen of the animal, indicating that these BuV-like viruses can cause systemic infections. Besides in primates, BuV-like viruses have so far also been identified in shrews, bats, rats, swine, and fur seals [12,26,27,28,29,30].

3. Tusavirus

Tusavirus, the name corresponding to Tunisian stool-associated parvovirus, was identified in 2014 using NGS and metagenomics in the fecal sample of an 18-month-old child with unexplained diarrhea in Tunisia [10]. Only one pool showed a single read of a parvovirus-like sequence, which was most similar to the NS1 sequence of a rat protoparvovirus. A near-complete genome sequence of 4424 bp was obtained (GenBank # KJ495710), including the two major ORFs, NS1 and VP1, typical of parvoviruses. Two conserved replication initiator sites and the Walker loop were also identified within NS1, as well as the PLA2 motif and an unusual serine-rich sequence near the VP1 N-terminus. TuV was shown to be quite distant from other protoparvoviruses: NS1 and VP1 showing the closest identities of 44% and 39%, respectively, to the Kilham rat parvovirus (Figure 1). TuV was thus proposed to be the prototype member for a new species, Primate protoparvovirus 2, in the Protoparvovirus genus [10]. However, the final classification is still pending due to the scarcity of this virus. No other TuV sequences were found in these Tunisian fecal samples by a nested PCR targeting the NS1 gene, and thus the TuV DNA was present in only 1/180 (0.56%) diarrheal fecal samples. However, no other human-pathogenic viruses were detected in this sample.
No other studies of TuV DNA prevalence in any sample type have been reported, and only one other metagenomics study has disclosed sequences showing resemblance to TuV: a study of fur seals in Brazil described partial sequences ranging from 344 to 1519 nt in length, of both NS1 and VP regions with 39–82% similarity to TuV at the amino acid level [30]. Besides DNA, one serological study of TuV IgG in humans has been published, in which one child in Finland was shown by EIA to be barely positive for TuV IgG (1/228, 0.44%) [22].

4. Cutavirus

In 2016, yet another protoparvovirus, cutavirus, was detected by viral metagenomics in 2/245 diarrheal fecal samples from Brazilian children [9]. The near full-length 4456 bp sequence (GenBank # KT868811) of the CuV genome showed the typical major ORFs for NS1 and VP1 with 76% and 82% amino acid identity, respectively, to those of BuV2, the phylogenetically closest parvovirus, suggesting that CuV may be a distinct protoparvovirus species (Figure 1). The CuV NS1 gene was, however, shorter than that of BuV; the alignment revealing an 11 amino-acid deletion near the C-terminus. The CuV genome was shown to contain the typical NS1 helicase motif with NTP-binding domains, as well as the VP1 PLA2 motif and glycine-rich regions. In addition, the sequence included a BuV-like middle ORF of unknown function, 20 amino acids shorter than that of BuV, with a 45% identity (Figure 1).
By CuV VP-specific nested PCR, three more CuV sequences were detected, 2 more in Brazilian fecal samples and one in one hundred fecal samples from Botswana. In all, the CuV DNA fecal prevalence was 4/245, 1.6%, in Brazilian and 1/100, 1.0%, in Botswanan diarrheic children [9]. These CuV genomes were only partly sequenced. To search for other pathogens in the five CuV DNA-positive samples, NGS was performed on the individual samples. In one of the Brazilian samples, no other human viruses were detected [9]. In the other three Brazilian samples, one contained rotavirus A and AAV, one astrovirus and adenovirus, and one rotavirus A and anellovirus. The Botswanan sample harbored both picobirna- and anellovirus.
Interestingly, by further in silico screening of existing NGS libraries, CuV DNA was retrospectively detected in two skin biopsies of French cutaneous T-cell lymphoma (CTCL) patients, thereby resulting in the virus name [9]. By CuV-nested PCR, 15 additional CTCL skin samples were screened, of which two more (2/15, 13.3%) were found to be CuV DNA-positive. Skin samples of other cancers (n = 10) and non-malignant skin biopsies (n = 19) were all CuV PCR-negative (Table 3). Later on, CuV DNA was detected with NGS in the malignant skin lesions of one melanoma patient in Denmark [31], while 9 additional melanoma samples were CuV DNA-negative when using both NGS and PCR. The CuV nt sequence from Denmark was 96.5–93.7% identical to the previously published three sequences covering the entire NS1 and VP coding regions (Table 4, data calculated for this review, and Figure 1). When comparing all known CuV nt sequences (of 3980 nt in length), the BR-283 sequence from Brazilian feces was closer to the Danish and French sequences from skin than to the other Brazilian sequences (Table 4). These data indicate that the sequences of CuV strains found in skin and feces are similar.

5. Discussion

Three protoparvoviruses, BuV, TuV and CuV, are the newest parvoviruses identified among humans [8,9,10]. All three were originally discovered in diarrheal feces of children by using virion enrichment and nuclease treatment prior to NGS, suggesting that the discovered viral DNA in the samples might have been protected by a capsid structure. No cell culture experiments have been done. The subsequent studies have instead focused on finding the DNAs of these viruses, or antibodies towards them, in human samples. As parvoviruses generally infect host-specifically and human parvoviruses grow poorly in cell cultures, DNA and antibody detection are the most feasible ways to establish these viruses as human viruses. Studies and findings on TuV DNA or antibodies are, however, scarce; only one DNA isolate exists and only one child has been barely TuV IgG-positive [10,22]. Whether TuV truly is a human parvovirus and not of, e.g., dietary origin, requires further investigation.
Of these three protoparvoviruses, BuV was discovered first and has been studied the most. The fact that BuV DNA is found near-exclusively in the feces of patients with diarrhea, often without other pathogen coinfections, whereas non-diarrheal fecal samples all have been BuV DNA-negative, suggests that BuV might be causative for gastroenteritis. However, the low viral loads and infrequent detection of BuV DNA point to a less prominent role in the overall burden of gastroenteritis. Whether BuV is involved in other clinical manifestations is not yet known, as acute primary BuV infections have not yet been serologically identified. The presence of IgG in humans, however, indicates that BuV is a true, capsid-covered infectious virus. The remarkable difference in BuV seroprevalence between Finland and Asia is interesting and under ongoing investigation. The reasons for such diversity might be cultural, genetic, or demo-geographic.
The newest member of the human protoparvoviruses is CuV, the DNA of which has been detected both in the diarrheal feces of children and in 4/17 skin samples of patients with CTCL (mycosis fungoides), whereas all 31 non-CTCL skin biopsies were PCR negative [9]. In addition, one Danish melanoma skin biopsy was found to be CuV DNA-positive. That the virus has been detected both in feces and skin suggests that CuV is a human virus and may cause systemic infection. Further PCR and sero-epidemiological studies are ongoing to assess the cancer association and the global distribution of this novel human virus.
In conclusion, according to DNA and antibody data, BuV and CuV are indeed human viruses, whereas for TuV the evidence is inconclusive. Further DNA studies of diverse sample materials combined with serological analyses are needed to establish the pathogenicity, or the absence of it, regarding, in particular, gastrointestinal and neoplastic illnesses.

Acknowledgments

This work was supported by the Academy of Finland (grant #1257964), the Life and Health Medical Grant Association, the China Scholarship Council, the Sigrid Jusélius Foundation, the Jane and Aatos Erkko Foundation, the Helsinki University Hospital Research and Education Fund, the Medical Society of Finland (FLS), the Research Funds of the University of Helsinki, the Clinical Chemistry Foundation, the Biomedicum Helsinki Foundation, the Finnish Society for Study of Infectious Diseases, and the Jenny and Antti Wihuri Foundation.

Author Contributions

Elina Väisänen, Yu Fu, Klaus Hedman and Maria Söderlund-Venermo contributed to the writing of the paper. Yu Fu provided the phylogenetic trees.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cossart, Y.E.; Field, A.M.; Cant, B.; Widdows, D. Parvovirus-like particles in human sera. Lancet 1975, 1, 72–73. [Google Scholar] [CrossRef]
  2. Qiu, J.; Soderlund-Venermo, M.; Young, N.S. Human Parvoviruses. Clin. Microbiol. Rev. 2017, 30, 43–113. [Google Scholar] [CrossRef] [PubMed]
  3. Jones, M.S.; Kapoor, A.; Lukashov, V.V.; Simmonds, P.; Hecht, F.; Delwart, E. New DNA viruses identified in patients with acute viral infection syndrome. J. Virol. 2005, 79, 8230–8236. [Google Scholar] [CrossRef] [PubMed]
  4. Allander, T.; Tammi, M.T.; Eriksson, M.; Bjerkner, A.; Tiveljung-Lindell, A.; Andersson, B. Cloning of a human parvovirus by molecular screening of respiratory tract samples. Proc. Natl. Acad. Sci. USA 2005, 102, 12891–12896. [Google Scholar] [CrossRef] [PubMed]
  5. Arthur, J.L.; Higgins, G.D.; Davidson, G.P.; Givney, R.C.; Ratcliff, R.M. A novel bocavirus associated with acute gastroenteritis in Australian children. PLoS. Pathog. 2009, 5, e1000391. [Google Scholar] [CrossRef] [PubMed]
  6. Kapoor, A.; Slikas, E.; Simmonds, P.; Chieochansin, T.; Naeem, A.; Shaukat, S.; Alam, M.M.; Sharif, S.; Angez, M.; Zaidi, S.; et al. A newly identified bocavirus species in human stool. J. Infect. Dis. 2009, 199, 196–200. [Google Scholar] [CrossRef] [PubMed]
  7. Kapoor, A.; Simmonds, P.; Slikas, E.; Li, L.; Bodhidatta, L.; Sethabutr, O.; Triki, H.; Bahri, O.; Oderinde, B.S.; Baba, M.M.; et al. Human bocaviruses are highly diverse, dispersed, recombination prone, and prevalent in enteric infections. J. Infect. Dis. 2010, 201, 1633–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Phan, T.G.; Vo, N.P.; Bonkoungou, I.J.; Kapoor, A.; Barro, N.; O’Ryan, M.; Kapusinszky, B.; Wang, C.; Delwart, E. Acute diarrhea in West African children: Diverse enteric viruses and a novel parvovirus genus. J. Virol. 2012, 86, 11024–11030. [Google Scholar] [CrossRef] [PubMed]
  9. Phan, T.G.; Dreno, B.; da Costa, A.C.; Li, L.; Orlandi, P.; Deng, X.; Kapusinszky, B.; Siqueira, J.; Knol, A.C.; Halary, F.; et al. A new protoparvovirus in human fecal samples and cutaneous T cell lymphomas (mycosis fungoides). Virology 2016, 496, 299–305. [Google Scholar] [CrossRef] [PubMed]
  10. Phan, T.G.; Sdiri-Loulizi, K.; Aouni, M.; Ambert-Balay, K.; Pothier, P.; Deng, X.; Delwart, E. New parvovirus in child with unexplained diarrhea, Tunisia. Emerg. Infect. Dis. 2014, 20, 1911–1913. [Google Scholar] [CrossRef] [PubMed]
  11. Yahiro, T.; Wangchuk, S.; Tshering, K.; Bandhari, P.; Zangmo, S.; Dorji, T.; Tshering, K.; Matsumoto, T.; Nishizono, A.; Soderlund-Venermo, M.; et al. Novel human bufavirus genotype 3 in children with severe diarrhea, Bhutan. Emerg. Infect. Dis. 2014, 20, 1037–1039. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, S.; Liu, D.; Wang, Y.; Qu, F.; He, Y.; Sun, Z.; Shen, Q.; Li, W.; Fu, X.; Deng, X.; et al. Bufavirus Protoparvovirus in feces of wild rats in China. Virus Genes 2016, 52, 130–133. [Google Scholar] [CrossRef] [PubMed]
  13. Cotmore, S.F.; Agbandje-McKenna, M.; Chiorini, J.A.; Mukha, D.V.; Pintel, D.J.; Qiu, J.; Soderlund-Venermo, M.; Tattersall, P.; Tijssen, P.; Gatherer, D.; et al. The family Parvoviridae. Arch. Virol. 2014, 159, 1239–1247. [Google Scholar] [CrossRef] [PubMed]
  14. Thompson, J.D.; Higgins, D.G.; Gibson, T.J. CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994, 22, 4674–4680. [Google Scholar] [CrossRef]
  15. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed]
  16. Vaisanen, E.; Kuisma, I.; Phan, T.G.; Delwart, E.; Lappalainen, M.; Tarkka, E.; Hedman, K.; Soderlund-Venermo, M. Bufavirus in feces of patients with gastroenteritis, Finland. Emerg. Infect. Dis. 2014, 20, 1077–1080. [Google Scholar] [CrossRef] [PubMed]
  17. Smits, S.L.; Schapendonk, C.M.; van Beek, J.; Vennema, H.; Schurch, A.C.; Schipper, D.; Bodewes, R.; Haagmans, B.L.; Osterhaus, A.D.; Koopmans, M.P. New viruses in idiopathic human diarrhea cases, the Netherlands. Emerg. Infect. Dis. 2014, 20, 1218–1222. [Google Scholar] [CrossRef] [PubMed]
  18. Altay, A.; Yahiro, T.; Bozdayi, G.; Matsumoto, T.; Sahin, F.; Ozkan, S.; Nishizono, A.; Soderlund-Venermo, M.; Ahmed, K. Bufavirus genotype 3 in Turkish children with severe diarrhoea. Clin. Microbiol. Infect. 2015, 21, e1–e4. [Google Scholar] [CrossRef] [PubMed]
  19. Chieochansin, T.; Vutithanachot, V.; Theamboonlers, A.; Poovorawan, Y. Bufavirus in fecal specimens of patients with and without diarrhea in Thailand. Arch. Virol. 2015, 160, 1781–1784. [Google Scholar] [CrossRef] [PubMed]
  20. Huang, D.D.; Wang, W.; Lu, Q.B.; Zhao, J.; Guo, C.T.; Wang, H.Y.; Zhang, X.A.; Tong, Y.G.; Liu, W.; Cao, W.C. Identification of Bufavirus-1 and Bufavirus-3 in Feces of Patients with Acute Diarrhea, China. Sci. Rep. 2015, 5, 13272. [Google Scholar] [CrossRef] [PubMed]
  21. Ayouni, S.; Estienney, M.; Hammami, S.; Neji Guediche, M.; Pothier, P.; Aouni, M.; Belliot, G.; de Rougemont, A. Cosavirus, Salivirus and Bufavirus in Diarrheal Tunisian Infants. PLoS ONE 2016, 11, e0162255. [Google Scholar] [CrossRef] [PubMed]
  22. Vaisanen, E.; Paloniemi, M.; Kuisma, I.; Lithovius, V.; Kumar, A.; Franssila, R.; Ahmed, K.; Delwart, E.; Vesikari, T.; Hedman, K.; et al. Epidemiology of two human protoparvoviruses, bufavirus and tusavirus. Sci. Rep. 2016, 6, 39267. [Google Scholar] [CrossRef] [PubMed]
  23. Altan, E.; Del Valle Mendoza, J.; Deng, X.; Phan, T.G.; Sadeghi, M.; Delwart, E.L. Small Circular Rep-Encoding Single-Stranded DNA Genomes in Peruvian Diarrhea Virome. Genome Announc. 2017, 5. [Google Scholar] [CrossRef] [PubMed]
  24. Altay Kocak, A.; Ocal, M.; Polat, M.; Kanik Yuksek, S.; Aktas Tapisiz, A.; Tezer, H.; Ozkul, A.; Ergunay, K.; Bozdayi, G.; Ahmed, K. Multicenter investigation of bufavirus in the etiology of viral central nervous system infections of adults and children. Mikrobiyol. Bul. 2017, 51, 191–194. [Google Scholar] [CrossRef] [PubMed]
  25. Handley, S.A.; Thackray, L.B.; Zhao, G.; Presti, R.; Miller, A.D.; Droit, L.; Abbink, P.; Maxfield, L.F.; Kambal, A.; Duan, E.; et al. Pathogenic simian immunodeficiency virus infection is associated with expansion of the enteric virome. Cell 2012, 151, 253–266. [Google Scholar] [CrossRef] [PubMed]
  26. Sasaki, M.; Orba, Y.; Anindita, P.D.; Ishii, A.; Ueno, K.; Hang’ombe, B.M.; Mweene, A.S.; Ito, K.; Sawa, H. Distinct Lineages of Bufavirus in Wild Shrews and Nonhuman Primates. Emerg. Infect. Dis. 2015, 21, 1230–1233. [Google Scholar] [CrossRef] [PubMed]
  27. Kemenesi, G.; Dallos, B.; Gorfol, T.; Estok, P.; Boldogh, S.; Kurucz, K.; Oldal, M.; Marton, S.; Banyai, K.; Jakab, F. Genetic diversity and recombination within bufaviruses: Detection of a novel strain in Hungarian bats. Infect. Genet. Evol. 2015, 33, 288–292. [Google Scholar] [CrossRef] [PubMed]
  28. Hargitai, R.; Pankovics, P.; Kertesz, A.M.; Biro, H.; Boros, A.; Phan, T.G.; Delwart, E.; Reuter, G. Detection and genetic characterization of a novel parvovirus distantly related to human bufavirus in domestic pigs. Arch. Virol. 2016, 161, 1033–1037. [Google Scholar] [CrossRef] [PubMed]
  29. Liu, L.; Schwarz, L.; Ullman, K.; Ahola, H.; Qiu, Y.; Ma, Z.; Hennig-Pauka, I. Identification of a novel bufavirus in domestic pigs by a viral metagenomic approach. J. Gen. Virol. 2016, 97, 1592–1596. [Google Scholar] [CrossRef] [PubMed]
  30. Kluge, M.; Campos, F.S.; Tavares, M.; de Amorim, D.B.; Valdez, F.P.; Giongo, A.; Roehe, P.M.; Franco, A.C. Metagenomic Survey of Viral Diversity Obtained from Feces of Subantarctic and South American Fur Seals. PLoS ONE 2016, 11, e0151921. [Google Scholar] [CrossRef] [PubMed]
  31. Mollerup, S.; Fridholm, H.; Vinner, L.; Kjartansdottir, K.R.; Friis-Nielsen, J.; Asplund, M.; Herrera, J.A.; Steiniche, T.; Mourier, T.; Brunak, S.; et al. Cutavirus in Cutaneous Malignant Melanoma. Emerg. Infect. Dis. 2017, 23, 363–365. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of genomic structures and phylogenetic analyses of the sequences of BuVs, CuVs, TuV, and other members of Protoparvovirus. (a) Schematic illustration of BuV, TuV, CuV, and MVMp genomic structures. The hairpin regions are unknown for BuV, TuV, and CuV, and the hairpins of MVMp are represented as straight lines for this simplified scheme. Species, strain, GenBank no., and the length of genome are indicated; Phylogenetic analyses of (b) NS1 sequences; (c) VP2 sequences; (d) short middle ORF sequences (not available from TuV). The phylogenetic analysis was based on nucleic acid sequences. Sequences were aligned with ClustalX (Version 2.1) [14] using a protospacer adjacent motif (PAM) with a gap open penalty of 10 and gap extension penalty of 0.1. The neighbor-joining phylogenetic tree was generated with bootstrap values determined by 1000 replicates in the Molecular Evolutionary Genetics Analysis (MEGA)(Version 7.0) [15]. The evolutionary distances were calculated by the p-distance method. Bootstrap values are shown if >80%. NS1, non-structural protein 1; mORF, short middle ORF; VP1, viral capsid protein 1; VP2, viral capsid protein 2; NS2, non-structural protein 2.
Figure 1. Schematic illustration of genomic structures and phylogenetic analyses of the sequences of BuVs, CuVs, TuV, and other members of Protoparvovirus. (a) Schematic illustration of BuV, TuV, CuV, and MVMp genomic structures. The hairpin regions are unknown for BuV, TuV, and CuV, and the hairpins of MVMp are represented as straight lines for this simplified scheme. Species, strain, GenBank no., and the length of genome are indicated; Phylogenetic analyses of (b) NS1 sequences; (c) VP2 sequences; (d) short middle ORF sequences (not available from TuV). The phylogenetic analysis was based on nucleic acid sequences. Sequences were aligned with ClustalX (Version 2.1) [14] using a protospacer adjacent motif (PAM) with a gap open penalty of 10 and gap extension penalty of 0.1. The neighbor-joining phylogenetic tree was generated with bootstrap values determined by 1000 replicates in the Molecular Evolutionary Genetics Analysis (MEGA)(Version 7.0) [15]. The evolutionary distances were calculated by the p-distance method. Bootstrap values are shown if >80%. NS1, non-structural protein 1; mORF, short middle ORF; VP1, viral capsid protein 1; VP2, viral capsid protein 2; NS2, non-structural protein 2.
Viruses 09 00354 g001aViruses 09 00354 g001bViruses 09 00354 g001cViruses 09 00354 g001d
Table 1. BuV DNA studies of subjects with or without diarrhea.
Table 1. BuV DNA studies of subjects with or without diarrhea.
StudySample TypeCountryAge (Range) If KnownSampling TimenPositive (%)SymptomsGenotype
[8]FecesBurkina Faso<5 yrNovember 2008–February 2010984 (4.1%)GEThree BuV1, one BuV2
FecesTunisia“children”NA631 (1.6%)NPAFPunknown
FecesChile“children”NA1000 (0.0%)GE-
[11]FecesBhutan<5 yrFebruary 2010–January 20123933 (0.8%)GEAll BuV3
[16]FecesFinlandmedian 51.5 yr (0–99 yr)October 2012–March 2013; April–May 20136297 (1.1%)GESix BuV1, one unknown
[17]FecesThe Netherlandsmedian 47 yr (0–97 yr)2005–2009271 (3.7%)GEBuV3
[19]FecesThailandmedian NA (0–97 yr)January 2009–April 201414954 (0.3%)GEAll BuV1
FecesThailandmedian NA (0–39 yr)February 2010–July 20147410HFMD-
[18]FecesTurkeymean 19.5 mo (1–60 mo)September 2004–June 20115838 (1.4%)GESix BuV3, two unknown
FecesTurkeymean 17.3 mo (NA), age matchedFebruary–September 20131480healthy-
[20]FecesChina (General hospital, Beijing)median 35 yr (1 mo–85 yr)2010–20145209 (1.7%)GEFour BuV1, five BuV3
“children”NA760non-GE-
FecesChina (Children’s hospital, Chongqing)median 10 mo (1 day–14 yr)2010–201313570GE-
“children”NA3450non-GE-
[21]FecesTunisiamedian 7.0 mo (0.5–60 mo)October 2010–March 20122032 (1.0%)GEBoth BuV1
[22]FecesFinlandmedian 14 mo (6 days–15.6 yr)September 2009–August 20111722 (1.2%)GEUnknown 1
5450ARTI-
2381 (0.4%)GE & ARTIUnknown 1
nasal swabFinlandmedian 14 mo (6 days–15.6 yr)September 2009–August 20111720GE-
5450ARTI-
2381 (0.4%)GE & ARTIUnknown 1
[24]CSFTurkeymedian 32 yr (0–96 yr)October 2011–April 20151260febrile illness and/or CNS infection-
[23]FecesPeruUnknownNA300Unknown 2GEBuV3
1 Short sequence only from conserved NS1 region, cannot separate genotypes; 2 by NGS, only the presence of BuV3 sequence was reported, prevalence cannot be calculated; ARTI, acute respiratory tract infection; CNS, central nervous system; CSF, cerebrospinal fluid; GE, gastroenteritis; HFMD, hand, foot, and mouth disease; mo, month(s); NA, not available; non-GE, patients who had sought medical care from the same hospitals for reasons other than GE; NPAFP, non-polio acute flaccid paralysis; yr, year(s); n of the non-diarrheic control patients are bolded.
Table 2. BuV IgG-positive adults and children. The origins of the subjects are shown in the header.
Table 2. BuV IgG-positive adults and children. The origins of the subjects are shown in the header.
GenotypeAdultsChildren
Asia, n = 12Finland, n = 163Finland, n = 228
Middle EastIndiaChina
BuV11 13 12 2
BuV21 1 144 2
BuV31 1 2
Any BuV1 13157 2
1 One individual was positive for all three BuVs; 2 one individual was positive for both BuV1 and BuV2; Any BuV, sums of the BuV IgG-positive individuals.
Table 3. CuV DNA prevalences in fecal and skin samples.
Table 3. CuV DNA prevalences in fecal and skin samples.
StudySample TypeCountrynPositive (%)Symptoms/Diagnosis
[9]FecesBrazil2454 (1.6%)Diarrhea
[9]FecesBotswana1001 (1.0%)Diarrhea
[9]SkinFrance 1NA2CTCL, Mycosis fungoides
[9]SkinFrance152 (13.3%)CTCL, Mycosis fungoides
[9]SkinFrance100 (0%)Skin carcinoma
[9]SkinFrance 1NA0Parapsoriasis
[9]Skin2 France; 6 NA80 (0%)Parapsoriasis
[9]SkinNA80 (0%)Eczema or eczematoid dermatitis
[9]SkinNA30 (0%)Healthy
[31]SkinDenmark101 (10%)Melanoma
1 2 CuV DNA-positive skin samples from CTCL patients were detected with NGS and PCR. However, it is not known how many CTCL skin samples were tested by NGS, whereby these 2 cases are not included in the overall prevalence. The same holds for the 2 CuV DNA-negative parapsoriasis skin samples tested in NGS pools. NA, not available; CTCL, cutaneous T-cell lymphoma.
Table 4. Nucleotide identities between the currently available CuV sequences. A sequence of approximately 3980 nts, covering the partial NS1 and complete VP regions (nts 474–4456 of BR-337 [KT868811]) was used for calculations. Strain, sample type, and GenBank no. indicated.
Table 4. Nucleotide identities between the currently available CuV sequences. A sequence of approximately 3980 nts, covering the partial NS1 and complete VP regions (nts 474–4456 of BR-337 [KT868811]) was used for calculations. Strain, sample type, and GenBank no. indicated.
% Nucleotide Identity
BR-337, Feces [KT868811]BR-283, Feces [KT868810]BR-372, Feces [KT868809]BR-450, Feces [KT868812]BO-46, Feces [KT868813]FR-D, Skin [KT868813]FR-F, Skin [KT868813]CGG5-268, Skin [KX685945]
BR-337, feces [KT868811]100.0
BR-283, feces [KT868810]94.4100.0
BR-372, feces [KT868809]99.594.6100.0
BR-450, feces [KT868812]96.994.497.1100.0
BO-46, feces [KT868813]93.996.094.193.9100.0
FR-D, skin [KT868813]93.896.293.893.995.0100.0
FR-F, skin [KT868813]94.496.694.594.595.595.9100.0
CGG5-268, skin [KX685945]94.197.194.294.296.096.196.7100.0

Share and Cite

MDPI and ACS Style

Väisänen, E.; Fu, Y.; Hedman, K.; Söderlund-Venermo, M. Human Protoparvoviruses. Viruses 2017, 9, 354. https://0-doi-org.brum.beds.ac.uk/10.3390/v9110354

AMA Style

Väisänen E, Fu Y, Hedman K, Söderlund-Venermo M. Human Protoparvoviruses. Viruses. 2017; 9(11):354. https://0-doi-org.brum.beds.ac.uk/10.3390/v9110354

Chicago/Turabian Style

Väisänen, Elina, Yu Fu, Klaus Hedman, and Maria Söderlund-Venermo. 2017. "Human Protoparvoviruses" Viruses 9, no. 11: 354. https://0-doi-org.brum.beds.ac.uk/10.3390/v9110354

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

Article Metrics

Back to TopTop