Hostname: page-component-cd9895bd7-gxg78 Total loading time: 0 Render date: 2024-12-26T18:38:18.530Z Has data issue: false hasContentIssue false

Origin and evolution of West Nile virus lineage 1 in Italy

Published online by Cambridge University Press:  02 December 2024

Andrea Silverj*
Affiliation:
Centre Agriculture Food Environment, University of Trento, San Michele all’Adige, Trento, Italy Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy Department CIBIO, University of Trento, Trento, Italy
Giulia Mencattelli*
Affiliation:
Centre Agriculture Food Environment, University of Trento, San Michele all’Adige, Trento, Italy Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Federica Monaco
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Federica Iapaolo
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Liana Teodori
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Alessandra Leone
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Andrea Polci
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Valentina Curini
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Marco Di Domenico
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Barbara Secondini
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Valeria Di Lollo
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Massimo Ancora
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Annapia Di Gennaro
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Daniela Morelli
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
Maria Gabriella Perrotta
Affiliation:
Centro Nazionale di Lotta ed Emergenza Contro le Malattie Animali, Rome, Italy
Giovanni Marini
Affiliation:
Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy
Roberto Rosà
Affiliation:
Centre Agriculture Food Environment, University of Trento, San Michele all’Adige, Trento, Italy Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy
Nicola Segata
Affiliation:
Department CIBIO, University of Trento, Trento, Italy
Omar Rota-Stabelli
Affiliation:
Centre Agriculture Food Environment, University of Trento, San Michele all’Adige, Trento, Italy Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy Department CIBIO, University of Trento, Trento, Italy
Annapaola Rizzoli
Affiliation:
Research and Innovation Centre, Fondazione Edmund Mach, San Michele all’Adige, Trento, Italy National Biodiversity Future Center (NBFC), Palermo, Italy
Giovanni Savini
Affiliation:
Istituto Zooprofilattico Sperimentale dell’Abruzzo e del Molise, Teramo, Italy
West Nile Virus Working Group
Affiliation:
The members of the West Nile virus working group are listed in the group authorship list, located in supplementary materials
*
Corresponding authors: Andrea Silverj and Giulia Mencattelli; Emails: [email protected]; [email protected]
Corresponding authors: Andrea Silverj and Giulia Mencattelli; Emails: [email protected]; [email protected]
Rights & Permissions [Opens in a new window]

Abstract

West Nile virus (WNV) is a mosquito-borne pathogen that can infect humans, equids, and many bird species, posing a threat to their health. It consists of eight lineages, with Lineage 1 (L1) and Lineage 2 (L2) being the most prevalent and pathogenic. Italy is one of the hardest-hit European nations, with 330 neurological cases and 37 fatalities in humans in the 2021–2022 season, in which the L1 re-emerged after several years of low circulation. We assembled a database comprising all publicly available WNV genomes, along with 31 new Italian strains of WNV L1 sequenced in this study, to trace their evolutionary history using phylodynamics and phylogeography. Our analysis suggests that WNV L1 may have initially entered Italy from Northern Africa around 1985 and indicates a connection between European and Western Mediterranean countries, with two distinct strains circulating within Italy. Furthermore, we identified new genetic mutations that are typical of the Italian strains and that can be tested in future studies to assess their pathogenicity. Our research clarifies the dynamics of WNV L1 in Italy, provides a comprehensive dataset of genome sequences for future reference, and underscores the critical need for continuous and coordinated surveillance efforts between Europe and Africa.

Type
Original Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

Introduction

West Nile virus (WNV) is an arthropod-borne flavivirus member of a Japanese encephalitis serocomplex within the Flaviviridae family. In nature, it is maintained by several species of birds and competent mosquitoes mainly belonging to the Culex species. The former generally act as amplifying hosts, while the latter as vectors [Reference Rizzoli1, Reference Chancey2]. Humans and other animals are considered incidental dead-end hosts. In humans, most of the infections are mild (flu-like symptoms) or asymptomatic (West Nile fever, WNF) [Reference Kramer3]. In some cases (less than 1%), however, especially in older or immunocompromised people, they can cause severe and sometimes fatal neurological disease (WNND) [Reference Kramer3].

Since its first identification in Uganda in 1937 [Reference Smithburn4], WNV has progressively spread to Europe, America, Asia, and Oceania, becoming the most widespread arbovirus in the world [Reference Pierson and Diamond5]. Eight phylogenetic lineages are currently known: Lineage 1 (L1) to Lineage 8 (L8). Of them, WNV L1 and L2 are by far the most widespread and pathogenic. They are often associated with severe outbreaks in humans, horses, and birds [Reference Fall6]. WNV L1 cases started to be recorded in Europe in the middle 1950s [7], while phylogenetic studies show that the virus arrived in North America from the Middle East (Israel) in the 1990s (where it was first detected in 1999, but probably it had emerged the year before) [Reference Nash8Reference Beck10]. These and many of the subsequent introductions could be attributed to migratory birds or result from long-distance transport of infectious animals related to international trade [Reference Mancuso11, Reference Rappole12]. The circulation of WNV L1 has been observed in several European countries, and major clades have been characterized: (i) the Western Mediterranean clade, with strains first identified in Italy (1998) and France (2000) and (ii) the Eastern European clade that includes strains responsible for the outbreaks in Romania (1996) and Russia (1999) [Reference Beck10, Reference Sambri13, Reference Mencattelli14]. The 1998 WNV Italian strain was reported in the Tuscany region. It was able to cause severe neurological disease and deaths in horses, but no cases were recorded in humans [Reference Autorino15]. A re-emergence of WNV L1 strains belonging to the Western Mediterranean clade was reported 10 years later in North-Eastern Italian regions (Emilia-Romagna, Veneto, and Lombardy). They were responsible for severe illness in humans and horses. Numerous birds and mosquitoes were also found infected with WNV L1 [Reference Savini16] (https://westnile.izs.it/j6_wnd/wndItalia). In 2010, the circulation of WNV L1 was also observed in Southern Italy, and the outbreaks were reported in Sardinia in 2011 [Reference Spissu17, Reference Calistri18]. In the same year, the circulation of WNV L2 was detected for the first time in the Italian territory. It soon became the most prevalent lineage circulating in the country [Reference Savini19]; meanwhile, WNV L1 has been more and more sporadically reported. Evidence of its circulation was, however, observed in North-Eastern regions (2012–2014, 2017), Sardinia (2015–2016), and Campania (2020) [Reference Mencattelli14].

In 2021, unexpectedly, a new WNV L1 incursion re-occurred in the Northern part of Italy (https://westnile.izs.it/j6_wnd/docBolletItaPeriodico?annoDocumento=2021). The virus re-emerged in the same area in 2022, co-circulating with WNV L2 and causing nearly 300 WNND cases and 37 deaths in humans. WNV positivities were also reported in 45 horses and 349 wild and target birds. Similarly, numerous infected mosquito pools were found (https://westnile.izs.it/j6_wnd/docBolletItaPeriodico?annoDocumento=2022). The Veneto region was the major hotspot area for viral circulation, with 142 human WNND and 17 deaths [Reference Barzon20], (https://westnile.izs.it/j6_wnd/docBolletItaPeriodico?annoDocumento=2022).

The reappearance of WNV L1 in Italy, the pathogenic differences of the various circulating strains, the increasing incidence of WNND cases in humans, together with the gap of knowledge of several aspects of WNV biology, ecology, and genomics along with the lack of safe vaccines and specific therapies provide the motivation to improve research and deepen the understanding of the genetic features of the WNV L1 strains circulating in Italy. To this end, in this study, we have integrated a new genome dataset of WNV L1 strains circulating in Italy between 2008 and 2022 with a worldwide dataset of WNV L1 sequences publicly available in the NCBI. Furthermore, we have combined this consistent dataset with phylogenetic and phylogeographic analyses of L1 strains, displaying the dynamics of the viral circulation in the Italian territory and uncovering the spatial and temporal patterns of the virus between Western-Mediterranean countries.

Methods

Sample collection

WNV L1-positive samples were collected in Italy between 2008 and 2022 within the national surveillance plan, coordinated by the Ministry of Health, the Istituto Superiore di Sanità (epidemiology and national reference laboratory, human), and the Istituto Zooprofilattico of Abruzzo and Molise (IZS-Teramo) (epidemiology and national reference laboratory, animal/entomology) (https://westnile.izs.it/j6_wnd/home, https://www.epicentro.iss.it/westnile/). The plan is based on an integrated surveillance on birds, mosquitoes, horses, and humans, conducted on a regional scale in Italy [Reference Mencattelli21]. Any positive results from local surveillance activities are confirmed by the National Reference Centre for Foreign Animal Diseases (CESME) at IZS-Teramo. The official veterinary authorities are in charge of registering any notifications in the national information system for the notification of outbreaks in animals (SIMAN) [Reference Colangeli22].

Sample analysis

Tissue homogenization, RNA extraction, real-time RT-PCR

At IZS-Teramo, pools of mosquitoes, bird organs (heart, kidney, spleen, and brain), collected from either residential or wild birds, and horse samples (blood or brain) were homogenized in phosphate-buffered saline (PBS) with antibiotics. Viral RNA was extracted by using the MagMAX CORE Nucleic Acid Purification KIT (Applied Biosystem; Thermo Fisher Scientific, Life Technologies Corporation, TX, USA) according to the manufacturer’s instructions. The extracted viral RNA was amplified using two real-time reverse transcription polymerase chain reactions (qRT-PCR), one aiming for the simultaneous detection of WNV L1 and L2 [Reference Del Amo23], and the other aiming for the detection of all WNV lineages [Reference Vázquez24]. For Usutu virus, the qRT-PCR was performed as previously described by Cavrini et al. [Reference Cavrini25].

Sequence data preparation and retrieval

WNV-L1 positive samples were selected for Illumina and Sanger sequencing. Particularly, purified nucleic acids extracted from the WNV L1 viral strains obtained from samples collected in Italy between 2008 and 2012 were sequenced by using the Sanger methods [Reference Verma, Kulshrestha and Puri26]. Briefly, the total RNA was extracted from the collected sample using a High Pure Viral Nucleic Acid Kit (Roche Diagnostics GmbH; Roche Applied Science, Mannheim, Germany), according to the manufacturer instructions, and collected in 45 μL elution buffer prewarmed at 72 °C. The complete WNV-coding DNA sequences (cds) of the polyprotein precursor gene were amplified using 13 WNV primer pairs able to amplify 13 overlapping regions of the genome (the primer sequences can be found in Additional file 1: Table S1). Gel-based RT-PCR was performed using a Transcriptor One-Step RT-PCR kit (Roche Diagnostics Deutschland GmbH, Mannheim, Germany) according to the manufacturer’s instructions. The RT-PCR cycling conditions for the amplification were 50°C for 15 min and 94°C for 7 min, followed by 35 cycles of denaturation at 94°C for 10 s, annealing at 57.5°C for 30 s, and extension at 68°C for 4 min and 30 s, followed by 1 extension cycle performed at 68°C for 7 min. The gel-based RT-PCR amplicons were purified with a Qiaquick PCR Purification kit (Qiagen, Leipzig, Germany). The purified amplicons and the 13 WNV sequencing primers were sent to an external service, Eurofins Genomics (Eurofins Genomics, Germany GmbH, Ebersberg, Germany), to perform sequencing in both directions. The obtained sequences were analysed with SeqScape v3.0 (Thermo Fisher Scientific, Waltham, MA, USA). Furthermore, purified nucleic acids, obtained from the WNV L1-positive samples collected in Italy between 2020 and 2022, were sequenced by next-generation sequencing. In detail, each positive sample total RNA was subjected to Turbo DNase treatment (Thermo Fisher Scientific, Waltham, MA, USA) at 37°C for 20 min and then purified with an RNA Clean and Concentrator-5 Kit (Zymo Research, Irvine, CA, USA). The purified RNA was used for the assessment of the sequence-independent single primer amplification (SISPA) protocol [Reference Mencattelli14, Reference Marcacci27]. Following the SISPA protocol, a single-strand cDNA was obtained using reverse transcription (RT) in 20 μL reaction mixture with 5X SSIV buffer, 50 μM random hexamer FR26RV-N 50-GCCGGAGCTCTGCAGATATCNNNNNN-30, 10 mM dNTPs mix, 100 mM DTT, 200 units SuperScript IV Reverse Transcriptase (Thermo Fisher Scientific, Waltham, MA, USA), and 40 U RNAse OUT RNase inhibitor (Thermo Fisher Scientific, Waltham, MA, USA) following the manufacturers’ instructions. The reaction was incubated at 23°C for 10 min, 50°C for 50 min, and 80°C for 10 min. To convert the single-stranded cDNA into double-stranded (ds) cDNA, 1 μL (2.5 U) 3′-5′ Klenow Polymerase (New England Biolabs, Ipswich, MA, USA) was directly added to the reaction. The incubation was carried out at 37°C for 1 h and 75 °C for 10 min. Next, 5 μL of ds cDNA was amplified with a PCR master mix containing 5× Q5 reaction buffer, 10 mM dNTPs, 40 μM random primer FR20 Rv 50 -GCCGGAGCTCTGCAGATATC-30, 0.01 U/μL Q5 High Fidelity DNA polymerase (NEB, New England Biolabs, Ipswich, MA, USA), and 5× Q5 High Enhancer. The incubation was performed with the following thermal conditions: 98°C for 1 min, 40 cycles of 98°C for 10 s, 65°C for 30 s and 72°C for 3 min, and a final extension step of 72°C for 2 min. The PCR product was purified using Expin TM PCR SV (GeneAll Biotechnology CO., Seoul, Korea) and then quantified using a Qubit DNA HS Assay Kit (Thermo Fisher Scientific, Waltham, MA, USA). The sample was diluted to obtain a concentration of 100–500 ng, then used for library preparation with an Illumina DNA prep kit, and sequenced with a NextSeq 500 (Illumina Inc., San Diego, CA, USA) using a NextSeq 500/550 Mid Output Reagent Cartridge v2, 300 cycles, and standard 150 bp (base pairs) paired-end reads. After quality control with FastQC tool v0.11.5 (Bioinformatics Group, Babraham Institute, Cambridge, UK) [Reference Cito28, Reference Aprea29] and trimming with Trimmomatic v0.36 (Usadellab, Düsseldorf, Germany) [Reference Bolger, Lohse and Usadel30], raw reads (available upon request to the corresponding authors) were de novo assembled using SPADES v3.11.1 (Algorithmic Biology Lab, St Petersburg, Russia) [Reference Bankevich31]. The contigs obtained were analysed with BLASTn to identify the best match reference. We introduced snippy as aligner (version 4.5.1, default parameters; https://github.com/tseemann/snippy) and iVar (version 1.3.1, command consensus, -q 20, -m 1) to obtain consensus sequences [Reference Di Pasquale32]. This procedure was also integrated in a pipeline for the West Nile lineage prediction on GENPAT platform (https://genpat.izs.it/cmdbuild/ui/#login) of the “National Reference Centre for Whole Genome Sequencing of microbial pathogens: database and bioinformatic analysis.” Thirty-one WNV L1 consensus sequences were obtained at IZS-Teramo after trimmed reads were mapped to the WNV L1 reference sequence FJ483548 (Italy, 2008).

Metadata for the IZS-Teramo newly generated sequences were obtained from the Laboratory Information Management Systems (SILAB) at IZS-Teramo (https://www.izs.it/IZS/Engine/RAServePG.php/P/257610010719/L/1). A worldwide dataset of 130 WNV L1 sequences have been downloaded from the Supplementary Materials of [Reference Mencattelli21], including sequences obtained from NCBI on 01/03/2022. By using a custom R script for automatic sequence retrieval, a new search has been conducted on NCBI on 06/11/2022 and 28 newly published WNV L1 sequences >= 200 nt have been downloaded.

Sequence data cleaning and formatting

All sequences were quality-filtered and only data for sequences longer than 10 kb (i.e., sequences covering almost the entire WNV genome) were retained for subsequent steps, as described elsewhere [Reference Mencattelli21]. After this step, 2 sequences (OP734273, OU953898.1) were removed, as they contained a percentage of ambiguous bases above 10%. A total of 186 WNV L1 genomes, 55 of which coming from Italy (30 of which belonging to the IZS newly obtained dataset), were selected for further analysis. A table of sequence curated metadata can be found in Additional file 2: Table S2.

Alignment, recombination detection, and model selection

All the datasets were aligned individually using MAFFT v7 (https://mafft.cbrc.jp/alignment/server/) with the “--auto” option, and trimmed using trimAl v2 with the “-automated1” option [Reference Capella-Gutiérrez, Silla-Martínez and Gabaldón33]. The presence of recombinant sequences in the final dataset was checked by running the RDP4 program [Reference Martin34], with default options. The sequence AJ965626.2 from Portugal was indicated as suspect recombinant and was therefore excluded from the alignment. Model selection was carried out on all datasets using ModelFinder [Reference Kalyaanamoorthy35], implemented in IQTREE2 [Reference Minh36], using parameters “-T AUTO -m TESTONLY.” The best-fit model for both the worldwide and the Bayesian down-sampled datasets was GTR+F+I+G4, chosen according to both Akaike Information Criterion (AIC) and Bayesian Information Criterion (BIC).

Maximum likelihood phylogenies

A maximum likelihood phylogeny of the dataset including the 185 WNV L1 sequences was reconstructed by using RAxML v8.2.12 [Reference Stamatakis37], with commands “-p 1989 -m GTRGAMMAI -x 2483 -# 100 -f a -T 20”. Clades were annotated using the resulting topology when having bootstrap supports ≥90.

Molecular clock and phylogeographic analysis

We used the results obtained in the maximum likelihood analysis to subset our dataset for molecular clock analysis. In particular, we selected a highly supported group (see Figure 2), which includes clades 1–4 and all Italian genomes.

A full list of the sequence identifiers and the associated meta-data (retrieved from the NCBI or the literature) is provided in the Additional file 3: Table S3. When the exact position was not available, we approximated the location using the coordinates of the municipality from which the sample was collected. Phylogeography was reconstructed by using continuous traits (latitudinal and longitudinal coordinates for each sequence) in BEAST v1.10.4 [Reference Suchard38]. We divided the analysis into two different partitions (one for sequence data and the other for continuous coordinates) and used the relaxed random walk Cauchy diffusion model for our location partition, with bivariate traits representing latitude and longitude, adding random jitter to the tips (jitter window size: 0.01). For the location partition, we selected the option to reconstruct states at all ancestors. We employed an uncorrelated relaxed clock with a log-normal distribution and a Coalescent Bayesian Skyline tree prior. We ran four different analyses, setting for each run an MCMC length of 200 × 106 generations, sampling every 20000 steps. Convergence was assessed using Tracer v1.7.1 [Reference Rambaut39], to check the effective sample size (ESS) of each independent run and of the combined results, setting the burn-in values and making sure that all parameters were above a significance threshold of ESS (>200) for the combined run. We combined the four independent tree files using LogCombiner [Reference Suchard38], setting the burn-in percentages as follows: run1 = 30%, run2 = 50%, run3 = 45%, run4 = 45%. A maximum-clade credibility tree using median-heights was obtained with TreeAnnotator [Reference Suchard38], using the combined file with 460 × 106 states obtained with LogCombiner after the burn-in. Phylogenetic trees and phylogeographic diffusion patterns were represented using custom R scripts, employing the ggtree [Reference Yu40] and SERAPHIM [Reference Dellicour41] packages, respectively. A sensitivity analysis to check for the robustness of our results to differences in sampling was carried out by generating three datasets by randomly subsampling the dataset used for phylogeographic analysis with a custom Python script (available at https://github.com/andrea-silverj/WNV-Afr_Eur/tree/main/scripts/sequence_random-sampling). Specifically, we obtained 75% (52 genomes), 50% (35 genomes), and 30% (21 genomes) reduced datasets that were analysed with the same methods and models as described previously. To reach convergence (ESS > 200), we ran the Bayesian analyses setting the following parameters: 30% dataset: MCMC length of 250 × 106 generations, sampling every 25000 steps; 50% dataset: 8 runs with MCMC length of 500 × 106 generations, sampling every 50000 steps, that were then combined together; 75% dataset: MCMC length of 500 × 106 generations, sampling every 50000 steps. In all cases, we set a 10% burn-in.

Full-length polyprotein analysis

Using Ugene software v44.0 (https://ugene.net/download.html), a sequence comparison between the amino acid residues of the group of strains included in the phylogeographic analysis was performed. In particular, we compared the 2021–2022 group of Italian sequences and the strains: (i) OP850023, Italy, Campania region, 2022; (ii) MW627239, Italy, Campania region, 2020; (iii) Italy, 2008 group; (iv) Italy, 2011 group; (v) AF404757, Italy, Tuscany region, 1998; (vi) MT863559, France, 2015; (vii) OU953897, Spain, 2020; (viii) JF719069, Spain, 2010; (ix) AY701413, Morocco, 2003; (x) AJ965628, Portugal, 2004; and (xi) DQ786573, France, 2004. We also included in the sequence comparison the strain NC_009942/1999 because, although it is not part of the major Western Mediterranean cluster analysed here, it is representative of all virulent L1 strains from the United States.

Results

Genome sequence analysis

Consensus sequences were characterized by 31 WNV L1 genomes. Among them, 14 genomes were obtained from samples collected between 2008 and 2012 with Sanger sequencing, with a mean consensus length of 10263 bp. The genomes were submitted through BankIt (https://submit.ncbi.nlm.nih.gov/about/bankit/) and were released in the GenBank on 18 May 2021 under the accession numbers MW835351, MW835352, MW835353, MW835354, MW835355, MW835356, MW835357, MW835358, MW835359, MW835360, MW835361, MW835362, MW835363, and MW835364. The other 17 genomes were obtained with Illumina sequencing, with an average total number of 5096291 trimmed reads. The numbers of mapped reads (151 nucleotides [nt] in length) ranged from 159703 to 1333187, with coverage depth ranging from 2101.31× to 6381.67×. The mean consensus sequence length was 10966 bp. They were obtained from samples collected in Italy between 2020 and 2022 and were uploaded through BankIt and released in the GenBank on 26 October 2022 and 15 November 2022 under the accession numbers OP734262, OP734263, OP734264, OP734265, OP734266, OP734267, OP734268, OP734269, OP734270, OP734271, OP734272, OP734273, OP734274, OP850021, OP850022, OP850023, and MW627239. Run details for each of the 31 samples and negative controls can be found in Additional file 4: Table S4. A map of WNV L1 sequence geo-localization sites is shown in Figure 1.

Figure 1. Geo-localization of West Nile virus lineage 1 sequenced sample collection sites. The sampling date is indicated by a colour scale. The shape indicates the host from which the sample was extracted. For the map, we used region border shapes from GADM (which are freely available under a CC-BY license; https://gadm.org/license.html), retrieving them through the ‘getData’ function of the ‘raster’ R package (https://www.rdocumentation.org/packages/raster/versions/3.6-20/topics/getData).

Phylogenetic tree inferred with maximum likelihood analysis

Maximum likelihood phylogenetic tree of the WNV L1 genomes analysed in this study showed the presence of seven well-supported clusters. All Italian genomes were included in cluster 2, together with a sequence of a strain from Israel (Genbank HM152775) and one from Morocco (Genbank AY701412), as shown in Figure 2.

Figure 2. Maximum likelihood tree of West Nile virus lineage 1 sequences. Seven major clusters (C1–C7, bar on the right) can be identified. Mediterranean countries that are part of a major cluster that includes all Italian genomes (all part of the group that was used for the molecular clock and phylogeographic analysis, highlighted in a black square with rounded corners) are indicated by different colours. Inside this group, 4 Italian clades (Clade 1–4) can be defined based on the topology of the tree. Bootstrap support values (BS) >90 are not shown; bootstrap support values between 70 and 90 are represented by a full black circle; low bootstrap supports (BS < 50) are represented with a white circle. Sequences obtained in this study are indicated by the “*” symbol. A scale bar representing substitutions per site is shown at the bottom left part of the figure.

Sequences from Italy can be divided into 4 highly supported groups (clades 1–4, Figure 2). Clade 1 includes the 2008–2009 and the 2011 groups of Italian WNV L1 sequences. They are genetically similar to the Spanish sequences of 2020 (bootstrap support = 100), and all are closely related to the 2020 (GenBank MW627239) and 2022 (GenBank OP850023) Campanian sequences (bootstrap support > 80). Clade 2 includes the 2022 WNV L1 Italian strains, originating from Veneto, Emilia-Romagna, Friuli-Venezia Giulia, and Lombardy regions (Additional file 2: Table S2). They appear to be in the same monophyletic group of the sequences obtained in the Veneto region in 2021 and closely related to the WNV L1 sequence (GenBank MT863559) obtained from a horse with neurological signs in Southern France in 2015 (bootstrap support = 100). Clade 3 includes Spanish sequences and the first WNV L1 genome ever sampled in Italy (Genbank AF404757), while clade 4 comprises sequences sampled between 2011 and 2013 near Livenza. The two remaining Italian genomes (JN858069/Italy/2011/AN-1 and JQ928175/Italy/2011/Piave), which group together, can’t be placed confidently in the tree (the signal is too low to place them with sufficient precision).

The strains DQ786572/France/2004 and DQ786573/France/2004 appear to be at the root of clade 1 and 2 (bootstrap support = 68).

The two strains collected in the Campania region in 2020 and 2022 appear to be very closely related (bootstrap support = 100) and distant from the 2021–2022 North-Eastern Italian strains.

Phylogeographic analysis of the WNV L1 Western Mediterranean clade

We reconstructed a phylogeographic diffusion pattern for a subsample of cluster 2 (highlighted in Figure 2), including all Italian clades. We estimated the value of the evolutionary rate of the virus at a mean of 7.102 × 10−4 substitution/site/year (median = 7.085 × 10−4; 95% HPD: 4.854 × 10−4–9.333 × 10−4), which is in line with the rate of other +ssRNA viruses and with previous analyses carried out on WNV L1 [Reference Duffy, Shackelton and Holmes42, Reference Hadfield43]. The position inferred for the common ancestor of the sequences suggests its initial spread in the area of the Gulf of Lion, with a possible first introduction occurring around 1989 (median = 1989.34; mean = 1987.97; 95% HPD: 1977.15–1995.48) and later appeared in the Camargue region of France around 1993 (median = 1993.53; mean = 1993.13; 95% HPD: 1988.01–1997.21). From this region, the virus moved southwards towards Morocco (AY701412.1|Morocco|1996/08/31; AY701413.1|Morocco|31 August 2003), Spain (FJ766332.1|Spain|15 October 2007; FJ766331.1|Spain|15 October 2007; JF707789.1|Spain|31 August 2008; JF719069.1|Spain|15 August 2010), and Portugal (AJ965628.2|Portugal|15 July 2004) between 1995 and 2010 and eastwards towards Israel (HM152775.1|Israel|31 August 2000) and Italy, with four introductions (see Figure 3), the first around 1998 in Tuscany region, the second in 2002 (median = 2002.21; mean = 2001.45; 95% HPD: 1990.94–2010.51) near the Livenza river (between Veneto and Friuli-Venezia-Giulia regions), the third in 2004 (median = 2004.38; mean = 2003.92; 95% HPD: 1999.61–2007.31) in Emilia-Romagna region, and the last one around 2021 (median = 2020.91; mean = 2020.84; 95% HPD: 2020.08–2021.46) in the northeast of the country (Veneto region). Following these first introductions (in particular the third one; see Figure 3), the virus then spread to Emilia-Romagna (median = 2007.79; mean = 2007.72; 95% HPD: 2006.99–2008.32), Sardinia (median = 2011.26; mean = 2011.22; 95% HPD: 2010.76–2011.57), and later on to Campania (median = 2020.49; mean = 2020.3; 95% HPD: 2017.27–2020.75) regions and to Spain (median = 2020.18; mean = 2020.06; 95% HPD: 2017.25–2020.75) (Figure 3).

Figure 3. Phylogeographic analysis of the WNV L1 Western Mediterranean clade. (A) Molecular clock of the clade including all Italian sequences. Posterior probabilities of the nodes are indicated by different shapes. The bar at the top indicates the host from which the genome was isolated (human, bird, mosquito, or horse). The time scales on the right and left sides indicate, with a colour gradient, the sampling time (blue: older; red: more recent) of the genomes (see tips) or the median estimated age of their ancestors (see nodes). The four main introductions in Italy are indicated by numbers close to the nodes (1–4), while genomes from the recent outbreaks in Southern Italy (SI) and Northern Italy (NI) are highlighted by a dotted square and a full square, respectively. (B) Phylogeographic diffusion pattern in continuous scale of the virus in the Mediterranean basin. Genome sampling sites and reconstructed locations are plotted together on a map, showing the connections among all genomes and their inferred ancestors. Again, the same colour scale used in (A) indicates the sampling time of the genomes or the median estimated age of their ancestors. The coloured areas in the figure represent the 95% HPDs of the locations and times inferred by the model. For the map, we used region border shapes from GADM (which are freely available under a CC-BY license; https://gadm.org/license.html), retrieving them through the ‘getData’ function of the ‘raster’ R package (https://www.rdocumentation.org/packages/raster/versions/3.6-20/topics/getData).

Our phylogeographic analysis also indicates a clear genetic flow connecting Morocco, Spain, France, and Italy (Figure 3). This scenario was tested further by running a sensitivity analysis on three subset of the datasets, obtaining similar results (Additional file 6: Figure S1).

Amino acid sequence analysis

Pairwise alignment shows the presence of a new conservative amino acid substitution (R122H) in the NS2A region of all 2021–2022 Italian WNV L1 genomes. This mutation is not present on the strain OP850023 isolated in 2022 in the Campania region (Figure 4).

Figure 4. Amino acid sequence analysis. Amino acid sequence comparison between West Nile virus lineage 1 Italian strains and other Western-Mediterranean strains included in Cluster 2, other than the 1999 NC_009942 strain representative for the USA genome sequence group included in Cluster 4, is displayed. C, Capsid protein; prM, pre-membrane protein; M, Membrane protein; NS, non-structural proteins. Common point mutations (51-T, 99-T, 96-I) detected in the two Campanian strains OP850023/2022 and MW627239/2022 are displayed in pink; the point mutation 93-K of the Envelope protein detected in the 2021–2022 Northern-Eastern Italian strains and in the 2015 French strain MT863559 is displayed in light green; the 122H amino acid substitution of the NS2a protein detected in the 2021–2022 Northern-Eastern Italian strains is displayed in yellow; the 249-P point mutation is displayed in light blue.

Three conservative amino acid changes are observed in the E (A51T), NS2B (M99T), and NS4B (L96I) proteins of the OP850023/Campania/2022 and MW627239/Campania/2020 strains, but not in the other strains. The NS3T249P point mutation is also found in the recent 2021–2022 Italian genomes, in the group of 2008 Italian sequences, in the 2004 strain from Portugal, and in the 1999 American sequence. More details about the amino acid sequence comparison are shown in Additional file 5: Table S5.

Discussion

WNV is the most widespread mosquito-borne pathogen in the world [Reference Rizzoli44]. Currently, it is seriously affecting many European countries [Reference Young45]. In Italy, WNV L2 has been the most prevalent circulating lineage in the last decade. WNV L1 strains instead have been circulating sporadically, with only few evidence reported in North-Eastern regions (2012–2014, 2017), Sardinia (2015–2016), and Campania (2020) [Reference Mencattelli14] (IZS-Teramo Annual Epidemiological Bulletins, https://westnile.izs.it/j6_wnd/wndItalia, https://westnile.izs.it/j6_wnd/wndItaliaPeriodici). This was the reason why health authorities and scientists involved in the surveillance and control were greatly surprised when it reappeared in the Italian territory in the 2021–2022 epidemic seasons, even more as its emergence coincided with an increased incidence of severe WNND in humans [Reference Barzon20, Reference Barzon46]. This study investigated the possible origin and evolution of the WNV L1 strains circulating in Italy, which was severely impacted by the virus compared to other EU-neighbouring countries (https://www.ecdc.europa.eu/en/west-nile-fever/surveillance-and-disease-data/historical), using two complementary approaches (phylogenetics and phylogeography) and providing insights into the viral circulation dynamics of WNV L1 in the country since 1998.

The Bayesian and maximum likelihood trees reflect the results obtained by other authors [Reference Aguilera-Sepúlveda47], who also observed the presence of seven well-supported clusters (Figure 2). Among the seven identified clusters, cluster 2 contains most of the European strains included in this study. Inside this cluster, it is worth noting the close relationship between (i) the 1998 Italian strain and the 2007–2008 group of Spanish sequences (bootstrap support = 100) (clade 3), (ii) the 2008–2009 and 2011 Italian and the 2020 Southern Spain strains (clade 1) (Figure 3), and (iii) the two Italian strains isolated in 2020 (GenBank MW627239) and 2022 (GenBank OP850023) in the Campania region to the 2008–2011 group of Italian sequences, which are also (weakly) related to a sequence from Spain (Genbank JF719069) (clade 1) [Reference Mencattelli14, Reference Casimiro-Soriguer48]. While, on one hand, this situation implies the presence of a corridor connecting Italy and Spain, on the other hand, taking into account the frequent silent periods observed between outbreaks, it suggests that WNV may circulate quietly for extended durations within the Western Mediterranean region [Reference Mencattelli49Reference Monaco51].

This latter assumption is also supported by the presence of the 2008–2009 Italian group of strains, which re-emerged in Italy after 10 years of silence [Reference Monaco51, Reference Monaco52]. Unlike the 1998, the 2008 incursion involved a larger area (3 regions and 8 provinces) and caused disease not only in horses but also in humans. Birds and mosquitoes were found infected too [Reference Autorino15, Reference Savini16, Reference Monaco51]. The virus reappeared in the same areas in 2009 and later spread to Sardinia in 2011 [Reference Spissu17, Reference Calistri18]. Considering the possibility of unnoticed circulation of WNV in the area, it is not easy to figure out whether this group of Italian sequences derives from a new re-introductory event or is just the re-emergence of a strain already circulating in the past. Previous studies seem to lean towards the latter hypothesis [Reference Rambaut39]. Nevertheless, our phylogenetic tree shows a close genetic relationship between the 2004 French strains (DQ786573.1, DQ786572.1) and all Italian strains belonging to clade 1, suggesting a possible new re-introductory event from France to Italy (Figure 3). Such geographical connection is further supported by the close relatedness of the 2021–2022 North-Eastern Italian strains and the WNV L1 sequence MT863559, a strain isolated from the Natural Reserve of Camargue, in Southern France in 2015 [Reference Beck10], suggesting a new re-introduction to Northern Italy from Southern France. The existence of a corridor between France and Italy is also suggested by the two 2004 French strains DQ786573 and DQ786573, closely related and sister group (even if this association is not strongly supported in both maximum likelihood and clock analyses, see Figures 2 and 3) of the two major Italian clades 1 and 2 described in our phylogenetic tree.

Information on the shared ancestry assumed using phylogenetic methods has been obtained when using the phylogeographic reconstructions of WNV L1 diffusion, which provided valuable insight on the WNV L1 dynamics in a geographic and temporal context. According to our analysis, the first European WNV clade may have arisen around 1989 (median = 1989.34; mean = 1987.97; HPD 95%: 1977.15–1995.48) in a location around the Gulf of Lion, near the boundaries between France and Spain (Figure 3). All strains included in this analysis seem to be connected with this area, then spreading in two opposite directions: (i) to the Camargue region of France around 1993, moving then south-westwards (Morocco, South of Spain, and Portugal) and eastwards (Central and Northern Italy, with four independent introductions, and Southern and Central Spain) and (ii) possibly to Morocco around 1996 (see Results and Figure 3). These results must be interpreted with care and describe the entry of the clade defined by the strains included in our analysis, which is not necessarily coincident with the first entry of the virus in Europe. Having only a few samples from some European and African countries limits the power of the analysis. This is particularly evident in some situations, such as a genome of Israel (HM152775.1|Israel|31 August 2000) which is genetically associated with strains located in the Western part of the Mediterranean basin. Even though we cannot exclude a possible real close link between strains located in these two areas, the most likely scenario is an indirect connection and possible common origin of strains from sub-Saharan Africa, as suggested by a recent analysis that was oriented on clarifying the circulation of WNV between Africa and Europe [Reference Mencattelli53]. Therefore, the reconstruction presented in our analysis doesn’t have to be interpreted as a successive story of linked events, as it is inherently probabilistic and dependent on the quality and quantity of the available information. It reflects the overall direction of the spread but not necessarily the precise events that happened in the past. This is particularly true for the first phases of WNV-L1 evolution that we described, where the uncertainty is much higher than for more recent events.

Despite this uncertainty, which is strongly dependent on the lack of samples from Europe and Africa, our analysis robustly supports the existence of a constant circulation among Western Mediterranean European countries, with a genetic flow between Europe and Africa, and movements from the former to the latter continent, as recently shown elsewhere [Reference Mencattelli53], and also in the opposite direction. This general finding is also robust to perturbations in the dataset, as highlighted by the results of our sensitivity analysis (Additional file 6: Figure S1).

In the European continent, the great abundance of susceptible hosts and vectors that belong to the endemic fauna of the Western Mediterranean area might have indeed facilitated the rapid spread of WNV to the continent despite the diversity of habitat and climate characterizing the area [Reference De Ascentis54, Reference Federici55]. While the first introduction of WNV L1 in Europe can be attributed to long distance migratory birds along their Western African migration routes through Northern-western African countries, such as Senegal and Morocco [Reference Aguilera-Sepúlveda47, Reference Mencattelli53, Reference Ndione56, Reference Schuffenecker57], the virus likely established and spread within the Western Mediterranean countries through short distance migratory birds in their shift between breeding grounds and overwintering quarters [Reference Mancuso11, Reference Mencattelli14, Reference Mencattelli49, Reference Sotelo50, Reference Mancuso58]. In fact, it is well known that migratory birds play a significant role in the long-range spread of WNV, getting infected during their long-distance movements across continents, and spreading the virus through different stopovers at various locations to new geographic areas that were previously unaffected by the virus [Reference Mancuso11]. Once infected, some endemic bird species, more susceptible than others, might act as amplifying hosts, developing high levels of viraemia able to further infect mosquitoes, and increasing the likelihood of transmission to other animals or humans [Reference Vaughan, Newman and Turell59].

As shown by Pesko and Ebel [Reference Pesko and Ebel60], changes to the WNV genome could also affect the transmission routes, host immune responses, and virus pathogenesis [Reference Petit and Shah61]. In many epidemics, including those occurring in Italy, there were some WNV L1 strains capable of causing severe neurological cases in humans, horses, and birds [Reference Castillo-Olivares and Wood62] (https://westnile.izs.it/j6_wnd/home). Our pairwise alignment evidenced several point mutations, which can be screened in future studies to test their potential in increasing virulence (Figure 4 and Additional file 5: Table S5) [Reference Barzon20]. The ecological and climatic Italian conditions of the 2021–2022 seasons might have created a favourable environment for a strong viral circulation and many replication events and, in turn, the appearance of new genetic variants among the group of Italian strains [Reference Barzon20]. These studies might help to understand the mechanisms underlying clinical signs of WNV in humans, giving a possible explanation for the new WNV L1 North-Eastern Italian region’s pathogenic phenotype. Moreover, they may help the comprehension of the mechanisms at the basis of viral evolution, viral fitness, and host adaptation. Understanding how these factors shape WNV L1 diverse viral populations is crucial to uncover the virus evolution, epidemiology, and prevalence.

Interestingly, the genome sequence OP850023 obtained from a horse in the Campania region in October 2022 is not included in the group of North-Eastern Italian sequences, although coming from the same epidemic season. It is closely related (posterior probability = 1, Figure 3) to the sequence MW627239, obtained also in the Campania region from a Northern goshawk 2 years before (October 2020) [Reference Mencattelli14]. They do not carry the NS2AR122H point mutation reported in all 2021–2022 North-Eastern strains but show three common new amino acid substitutions (EA51T, NS2BM99T, and NS4BL96I) (Figure 4 and Additional file 5: Table S5). It is not simple to assess the exact source of the introduction of the virus in the Campania region. Even though our maximum likelihood tree placed the Campanian strains close to the 2008–2011 group of Italian sequences and to a 2010 Spanish sequence (Genbank JF719069) (clade 1), suggesting an introduction to the Campania region from Spain or Italy (Northern Italy/2008, Sardinia/2011), the latter also supported by our phylogeographic analysis (Figure 3). Once introduced in Campania, these strains might have found a new ecological niche, establishing an intra-endemic circulation [Reference Mencattelli14]. Another possibility is that the mutations observed may have accumulated before the invasion and could therefore be the result of other ecological processes and/or a founder effect.

Conclusion

In this study, the 31 new Italian genome sequences obtained by the IZS-Teramo from 2008 to 2022 were used to uncover the phylogenetic and phylogeographic relationships of WNV L1 Italian strains with other strains circulating worldwide. We clarified the spatial and temporal dynamics of WNV L1 in Italy in the last 24 years and identified the presence of two diverse strains currently circulating in the country (one in the North-East and one circulating intra-regionally in the Campania region). Furthermore, we found that WNV L1 spreads through an ideal line that connects Morocco, Spain, Southern France, and Northern Italy, with further diffusion between diverse regions of the Italian territory and possible events of endemization. These results help clarify the WNV L1 dynamics in Europe, provide a new consistent dataset that can be used as reference data for future WNV investigations, and highlight the strong need of constant and coordinated surveillance activities among European and African countries.

Due to gaps in genomic surveillance data for WNV in Europe and a limited number of newly generated genomes, our results stressed the importance to further intensify WNV surveillance in the Mediterranean countries and, in particular, in areas where the virus might find favourable conditions for its endemic circulation, either in wetlands inhabited by birds and mosquitoes (Southern France and Southern and Eastern Spain) [Reference Beck10, Reference Casimiro-Soriguer48] or in areas where the circulation might seem reduced or absent. This will contribute to building a more consistent and homogeneous European dataset that could tackle the uncertainties related to the WNV dynamics.

Supplementary material

The supplementary material for this article can be found at http://doi.org/10.1017/S0950268824001420.

Data availability statement

All scripts used to perform the analyses are available at the GitHub repository: https://github.com/andrea-silverj/WNV-L1_IT. All data that support the findings of the study are available from the material associated with Figshare data repository (Project n. 157871). Particularly, meta-data used for phylogenetic and phylogeographic analyses and Run details reported for each of the 31 IZS-samples and negative controls, and amino acid sequence comparison can be found at https://doi.org/10.6084/m9.figshare.21947681.v4; Tree files and sequence alignment used for phylogenetic analyses can be found at https://doi.org/10.6084/m9.figshare.21940898.v3; Phylogeographic inference tree files, geographic coordinates, and videos can be found at (i) https://doi.org/10.6084/m9.figshare.21940931.v2 (sub-selected dataset) and (ii) https://doi.org/10.6084/m9.figshare.23631735.v2 (complete dataset). This material is published under the Creative Commons Attribution 4.0 International (CC BY 4.0).

Acknowledgements

The authors acknowledge all personnel in the lab and in the field for their skill and technical support. Special thanks to the IZS-Teramo staff and, in particular, to Andrea Bucciacchio and Iolanda Mangone for their support in the sequencing analysis and genome assembly. A.R. acknowledges the support of NBFC to Fondazione Edmund Mach, funded by the Italian Ministry of University and Research, PNRR, Missione 4 Componente 2, “Dalla ricerca all’impresa”, Investimento 1.4, Project CN00000033. This manuscript reflects only the authors’ views and opinions; neither the European Union nor the European Commission can be considered responsible for them.

Authors contribution

Conceptualization, A.S., G.M., F.M., and G.S.; methodology, A.S., G.M., O.R.S. and G.S.; validation, F.M., A.S., G.M., O.R.S., A.R., and G.S.; formal analysis, A.S., and G.M.; investigation, L.T., A.L., V.C., M.D.D., A.P., M.A., B.S., V.D.L., A.D.G., and G.M.; resources, G.S.; data curation, F.I., A.S., V.C., and G.M.; writing G.M., A.S.; original draft preparation, G.M., A.S., and G.S.; writing—review and editing, F.M., G.Mar., R.R., N.S., A.R., O.R.S. and G.S.; visualization, A.S., G.M., F.M., F.I., L.T., A.L., A.P., V.C., M.D.D., B.S., V.D.L., M.A., A.D.G., D.M., M.G.P., G.Mar., R.R., N.S., O.R.S., A.R., and G.S.; supervision, O.R.S., F.M., A.R., and G.S.; project administration, G.S.; funding acquisition, G.S. All authors have read and agreed to the published version of the manuscript.

Funding statement

This research has been financially supported by an international PhD initiative including Fondazione Edmund Mach, University of Trento, and Istituto Zooprofilattico of Teramo. This study has been partially funded by the EU grant 874850 MOOD and is catalogued as MOOD 060. This work has been partially supported by Italy’s Ministry of Health (research grant no. MSRCTE 04/19).This study has been also partially funded by the National Biodiversity Future Centre and received funding from the European Union Next-Generation EU (PIANO NAZIONALE DI RIPRESA E RESILIENZA (PNRR) – MISSIONE 4 COMPONENTE 2, INVESTIMENTO 1.4 – D.D. 1034 17/06/2022, CN00000033).

Competing interest

The authors declare that they have no competing interests.

Ethics approval and consent to participate

The cases reported in this study were investigated with routine procedures according to the national surveillance plan for arbovirus infections. Therefore, no approval was required from the ethics committee.

Footnotes

These authors contributed equally to this work.

References

Rizzoli, A, et al. (2015) The challenge of West Nile virus in Europe: knowledge gaps and research priorities. Euro Surveillance 20, 21135.Google ScholarPubMed
Chancey, C, et al. (2015) The global ecology and epidemiology of West Nile virus. BioMed Research International 2015, 376230.CrossRefGoogle ScholarPubMed
Kramer, LD, et al. (2007). West Nile virus. The Lancet Neurology 6, 171181.CrossRefGoogle ScholarPubMed
Smithburn, KC, et al. (1940) A neurotropic virus isolated from the blood of a native of Uganda. American Journal of Tropical Medicine 20, 471472.CrossRefGoogle Scholar
Pierson, TC and Diamond, MS (2020) The continued threat of emerging flaviviruses. Nature Microbiology 5, 796812.CrossRefGoogle ScholarPubMed
Fall, G, et al. (2017) Biological and phylogenetic characteristics of West African lineages of West Nile virus. PLOS Neglected Tropical Diseases 11, e0006078.CrossRefGoogle ScholarPubMed
European Centre for Disease Prevention and Control/WHO Regional Office for Europe (2011). West Nile Virus Infection Outbreak in Humans in Romania. Accessed 28 December 2022. https://www.ecdc.europa.eu/sites/default/files/media/en/publications/Publications/1104MIR_West_Nile_outbreak_Romania.pdfGoogle Scholar
Nash, D, et al. (2001) The outbreak of West Nile virus infection in the New York City area in 1999. The New England Journal of Medicine 24, 18071814.CrossRefGoogle Scholar
Mencattelli, G, et al. (2022) Epidemiology of West Nile virus in Africa: an underestimated threat. PLOS Neglected Tropical Diseases 16, e0010075.CrossRefGoogle ScholarPubMed
Beck, C, et al. (2020) Contrasted epidemiological patterns of West Nile virus lineages 1 and 2 infections in France from 2015 to 2019. Pathogens 9, E908.CrossRefGoogle ScholarPubMed
Mancuso, E, et al. (2022) West Nile and Usutu virus introduction via migratory birds: a retrospective analysis in Italy. Viruses 14, 416.CrossRefGoogle ScholarPubMed
Rappole, J. (2000) Migratory birds and spread of West Nile virus in the Western Hemisphere. Emerging Infectious Diseases 6: 319328.CrossRefGoogle ScholarPubMed
Sambri, V, et al. (2013) West Nile virus in Europe: emergence, epidemiology, diagnosis, treatment, and prevention. Clinical Microbiology and Infection 19, 699704.CrossRefGoogle ScholarPubMed
Mencattelli, G, et al. (2022) West Nile virus lineage 1 in Italy: newly introduced or a re-occurrence of a previously circulating strain? Viruses 14, 64.CrossRefGoogle Scholar
Autorino, GL, et al. (2002) West Nile virus epidemic in horses, Tuscany region, Italy. Emerging Infectious Diseases 8, 13721378.CrossRefGoogle ScholarPubMed
Savini, G, et al. (2008) Phylogenetic analysis of West Nile virus isolated in Italy in 2008. Euro Surveillance 13, 19048.Google ScholarPubMed
Spissu, N, et al. (2013) West Nile virus outbreak in Sardinia, Italy, in 2011. Journal of Infection in Developing Countries 7, 69.CrossRefGoogle ScholarPubMed
Calistri, P, et al. (2010) Epidemiology of West Nile in Europe and in the Mediterranean Basin. The Open Virology Journal 4, 2937.Google ScholarPubMed
Savini, G, et al. (2012) Evidence of West Nile virus lineage 2 circulation in Northern Italy. Veterinary Microbiology 158, 267273.CrossRefGoogle ScholarPubMed
Barzon, L, et al. (2022) Rapid spread of a new West Nile virus lineage 1 associated with increased risk of neuroinvasive disease during a large outbreak in northern Italy, 2022: One Health analysis. Journal of Travel Medicine; 4, taac125.CrossRefGoogle Scholar
Mencattelli, G, et al. (2023) Epidemiological and evolutionary analysis of West Nile virus lineage 2 in Italy. Viruses 15, 35.CrossRefGoogle Scholar
Colangeli, P, et al. (2011) The national information system for the notification of animal diseases in Italy. Veterinaria Italiana 47: 303312, 291–301.Google ScholarPubMed
Del Amo, J, et al. (2013) A novel quantitative multiplex real-time RT-PCR for the simultaneous detection and differentiation of West Nile virus lineages 1 and 2, and of Usutu virus. Journal of Virological Methods 189, 321327.CrossRefGoogle ScholarPubMed
Vázquez, A, et al. (2016) Real time PCR assay for detection of all known lineages of West Nile virus. Journal of Virological Methods 236, 266270.CrossRefGoogle ScholarPubMed
Cavrini, F, et al. (2011) A rapid and specific real-time RT-PCR assay to identify Usutu virus in human plasma, serum, and cerebrospinal fluid. Journal of Clinical Virology 50, 221223.CrossRefGoogle Scholar
Verma, M, Kulshrestha, S and Puri, A. (2017) Genome sequencing. Methods in Molecular Biology 1525, 333.CrossRefGoogle ScholarPubMed
Marcacci, M, et al. (2016) Genome characterization of feline morbillivirus from Italy. Journal of Virological Methods 234, 160163.CrossRefGoogle ScholarPubMed
Cito, F, et al. (2018) The Italian information system for the collection and analysis of complete genome sequence of pathogens isolated from animal, food and environment. International Journal of Infectious Diseases 73, 296297.CrossRefGoogle Scholar
Aprea, G, et al. (2020) Whole genome sequencing characterization of HEV3-e and HEV3-f subtypes among the wild boar population in the Abruzzo Region, Italy: first report. Microorganisms 8, 1393.CrossRefGoogle ScholarPubMed
Bolger, AM, Lohse, M and Usadel, B. (2014) Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 21142120.CrossRefGoogle ScholarPubMed
Bankevich, A, et al. (2012) SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. Journal of Computational Biology 19, 455477.CrossRefGoogle ScholarPubMed
Di Pasquale, A, et al. (2021) SARS-CoV-2 surveillance in Italy through phylogenomic inferences based on Hamming distances derived from pan-SNPs, -MNPs and -InDels. BMC Genomics 22, 782.CrossRefGoogle ScholarPubMed
Capella-Gutiérrez, S, Silla-Martínez, JM and Gabaldón, T. (2009) trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 19721973.CrossRefGoogle ScholarPubMed
Martin, DP, et al. (2015) RDP4: Detection and analysis of recombination patterns in virus genomes. Virus Evolution 1, vev003.CrossRefGoogle ScholarPubMed
Kalyaanamoorthy, S, et al. (2017) ModelFinder: fast model selection for accurate phylogenetic estimates. Nature Methods 14, 587589.CrossRefGoogle ScholarPubMed
Minh, BQ, et al. (2020) IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Molecular Biology and Evolution 37, 15301534.CrossRefGoogle ScholarPubMed
Stamatakis, A (2014) RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 13121313.CrossRefGoogle ScholarPubMed
Suchard, MA, et al. (2018) Bayesian phylogenetic and phylodynamic data integration using BEAST 1.10. Virus Evolution 4, vey016. https://doi.org/10.1093/ve/vey016CrossRefGoogle ScholarPubMed
Rambaut, A, et al. (2018) Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Systematic Biology 67, 901904.CrossRefGoogle ScholarPubMed
Yu, G. (2020) Using ggtree to visualize data on tree-like structures. Current Protocols in Bioinformatics 69, e96. https://doi.org/10.1002/cpbi.96CrossRefGoogle ScholarPubMed
Dellicour, S, et al. (2016) SERAPHIM: studying environmental rasters and phylogenetically informed movements. Bioinformatics 32, 32043206.CrossRefGoogle ScholarPubMed
Duffy, S, Shackelton, LA and Holmes, EC. (2008) Rates of evolutionary change in viruses: patterns and determinants. Nature Reviews Genetics 9, 267276.CrossRefGoogle ScholarPubMed
Hadfield, J, et al. (2019) Twenty years of West Nile virus spread and evolution in the Americas visualized by Nextstrain. PLOS Pathogens 15, e1008042.CrossRefGoogle ScholarPubMed
Rizzoli, A, et al. (2015) Understanding West Nile virus ecology in Europe: Culex pipiens host feeding preference in a hotspot of virus emergence. Parasites & Vectors 8, 213.CrossRefGoogle Scholar
Young, JJ, et al. (2021) Epidemiology of human West Nile virus infections in the European Union and European Union enlargement countries, 2010 to 2018. Eurosurveillance 26, 2001095.CrossRefGoogle ScholarPubMed
Barzon, L, et al. (2022) Early start of seasonal transmission and co-circulation of West Nile virus lineage 2 and a newly introduced lineage 1 strain, northern Italy, June 2022. Eurosurveillance 27, 2200548.CrossRefGoogle Scholar
Aguilera-Sepúlveda, P, et al. (2022) A new cluster of West Nile virus lineage 1 isolated from a northern goshawk in Spain. Transboundary and Emerging Diseases 69, 31213127.CrossRefGoogle ScholarPubMed
Casimiro-Soriguer, CS, et al. (2021) Phylogenetic analysis of the 2020 West Nile Virus (WNV) outbreak in Andalusia (Spain). Viruses 13, 836.CrossRefGoogle ScholarPubMed
Mencattelli, G, et al. (2022) West Nile virus lineage 2 overwintering in Italy. Tropical Medicine and Infectious Disease 7, 160.CrossRefGoogle ScholarPubMed
Sotelo, E, et al. (2011) Phylogenetic relationships of Western Mediterranean West Nile virus strains (1996–2010) using full-length genome sequences: single or multiple introductions? Journal of General Virology 92, 25122522.CrossRefGoogle ScholarPubMed
Monaco, F, et al. (2010) Re-emergence of West Nile virus in Italy. Zoonoses and Public Health 57, 476486.CrossRefGoogle ScholarPubMed
Monaco, F, et al. (2011) 2009 West Nile disease epidemic in Italy: first evidence of overwintering in Western Europe? Research in Veterinary Science 91, 321326.CrossRefGoogle ScholarPubMed
Mencattelli, G, et al. (2023) Spatial and temporal dynamics of West Nile virus between Africa and Europe. Nature Communications 14, 6440.CrossRefGoogle ScholarPubMed
De Ascentis, M, et al. (2022) Species of mosquitoes present in Abruzzo and Molise and their possible role as vector of Usutu and West Nile viruses. Veterinaria Italiana 58, 10.Google ScholarPubMed
Federici, V, et al. (2016) Epizootic haemorrhagic disease in Italy: vector competence of indigenous Culicoides species and spatial multicriteria evaluation of vulnerability. Veterinaria Italiana 52, 271279.Google ScholarPubMed
Ndione, MHD, et al. (2022) Re-introduction of West Nile virus lineage 1 in Senegal from Europe and subsequent circulation in human and mosquito populations between 2012 and 2021. Viruses 14, 2720.CrossRefGoogle Scholar
Schuffenecker, I, et al. (2005) West Nile virus in Morocco, 2003. Emerging Infectious Diseases 11, 306309.CrossRefGoogle ScholarPubMed
Mancuso, E, et al. (2022) Direct and indirect role of migratory birds in spreading CCHFV and WNV: a multidisciplinary study on three stop-over islands in Italy. Pathogens 11, 1056.CrossRefGoogle Scholar
Vaughan, JA, Newman, RA and Turell, MJ. (2022) Bird species define the relationship between West Nile viremia and infectiousness to Culex pipiens mosquitoes. PLOS Neglected Tropical Diseases 16, e0010835.CrossRefGoogle ScholarPubMed
Pesko, KN and Ebel, GD. (2012) West Nile virus population genetics and evolution. Infection, Genetics and Evolution 12, 181190.CrossRefGoogle ScholarPubMed
Petit, MJ and Shah, PS. (2019) Mapping arbovirus-vector interactions using systems biology techniques. Frontiers in Cellular and Infection Microbiology 8, 440.CrossRefGoogle ScholarPubMed
Castillo-Olivares, J and Wood, J. (2004) West Nile virus infection of horses. Veterinary Research 35, 467483.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. Geo-localization of West Nile virus lineage 1 sequenced sample collection sites. The sampling date is indicated by a colour scale. The shape indicates the host from which the sample was extracted. For the map, we used region border shapes from GADM (which are freely available under a CC-BY license; https://gadm.org/license.html), retrieving them through the ‘getData’ function of the ‘raster’ R package (https://www.rdocumentation.org/packages/raster/versions/3.6-20/topics/getData).

Figure 1

Figure 2. Maximum likelihood tree of West Nile virus lineage 1 sequences. Seven major clusters (C1–C7, bar on the right) can be identified. Mediterranean countries that are part of a major cluster that includes all Italian genomes (all part of the group that was used for the molecular clock and phylogeographic analysis, highlighted in a black square with rounded corners) are indicated by different colours. Inside this group, 4 Italian clades (Clade 1–4) can be defined based on the topology of the tree. Bootstrap support values (BS) >90 are not shown; bootstrap support values between 70 and 90 are represented by a full black circle; low bootstrap supports (BS < 50) are represented with a white circle. Sequences obtained in this study are indicated by the “*” symbol. A scale bar representing substitutions per site is shown at the bottom left part of the figure.

Figure 2

Figure 3. Phylogeographic analysis of the WNV L1 Western Mediterranean clade. (A) Molecular clock of the clade including all Italian sequences. Posterior probabilities of the nodes are indicated by different shapes. The bar at the top indicates the host from which the genome was isolated (human, bird, mosquito, or horse). The time scales on the right and left sides indicate, with a colour gradient, the sampling time (blue: older; red: more recent) of the genomes (see tips) or the median estimated age of their ancestors (see nodes). The four main introductions in Italy are indicated by numbers close to the nodes (1–4), while genomes from the recent outbreaks in Southern Italy (SI) and Northern Italy (NI) are highlighted by a dotted square and a full square, respectively. (B) Phylogeographic diffusion pattern in continuous scale of the virus in the Mediterranean basin. Genome sampling sites and reconstructed locations are plotted together on a map, showing the connections among all genomes and their inferred ancestors. Again, the same colour scale used in (A) indicates the sampling time of the genomes or the median estimated age of their ancestors. The coloured areas in the figure represent the 95% HPDs of the locations and times inferred by the model. For the map, we used region border shapes from GADM (which are freely available under a CC-BY license; https://gadm.org/license.html), retrieving them through the ‘getData’ function of the ‘raster’ R package (https://www.rdocumentation.org/packages/raster/versions/3.6-20/topics/getData).

Figure 3

Figure 4. Amino acid sequence analysis. Amino acid sequence comparison between West Nile virus lineage 1 Italian strains and other Western-Mediterranean strains included in Cluster 2, other than the 1999 NC_009942 strain representative for the USA genome sequence group included in Cluster 4, is displayed. C, Capsid protein; prM, pre-membrane protein; M, Membrane protein; NS, non-structural proteins. Common point mutations (51-T, 99-T, 96-I) detected in the two Campanian strains OP850023/2022 and MW627239/2022 are displayed in pink; the point mutation 93-K of the Envelope protein detected in the 2021–2022 Northern-Eastern Italian strains and in the 2015 French strain MT863559 is displayed in light green; the 122H amino acid substitution of the NS2a protein detected in the 2021–2022 Northern-Eastern Italian strains is displayed in yellow; the 249-P point mutation is displayed in light blue.

Supplementary material: File

Silverj et al. supplementary material 1

Silverj et al. supplementary material
Download Silverj et al. supplementary material 1(File)
File 4.1 MB
Supplementary material: File

Silverj et al. supplementary material 2

Silverj et al. supplementary material
Download Silverj et al. supplementary material 2(File)
File 21.7 KB
Supplementary material: File

Silverj et al. supplementary material 3

Silverj et al. supplementary material
Download Silverj et al. supplementary material 3(File)
File 119.7 KB
Supplementary material: File

Silverj et al. supplementary material 4

Silverj et al. supplementary material
Download Silverj et al. supplementary material 4(File)
File 293.5 KB
Supplementary material: File

Silverj et al. supplementary material 5

Silverj et al. supplementary material
Download Silverj et al. supplementary material 5(File)
File 966.3 KB
Supplementary material: File

Silverj et al. supplementary material 6

Silverj et al. supplementary material
Download Silverj et al. supplementary material 6(File)
File 120.6 KB
Supplementary material: File

Silverj et al. supplementary material 7

Silverj et al. supplementary material
Download Silverj et al. supplementary material 7(File)
File 28.7 KB