Skip to main content
Advertisement
  • Loading metrics

Dipeptidyl peptidase III as a DNA marker to investigate epidemiology and taxonomy of Old World Leishmania species

  • Insaf Bel Hadj Ali,

    Roles Conceptualization, Investigation, Methodology, Project administration, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia

  • Hamed Chouaieb ,

    Contributed equally to this work with: Hamed Chouaieb, Yusr Saadi Ben Aoun

    Roles Investigation, Methodology, Resources

    Affiliations Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia, Service de parasitologie, EPS Farhat Hached, Faculté de Médecine de Sousse, Université de Sousse, Sousse, Tunisia

  • Yusr Saadi Ben Aoun ,

    Contributed equally to this work with: Hamed Chouaieb, Yusr Saadi Ben Aoun

    Roles Methodology, Validation, Visualization

    Affiliation Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia

  • Emna Harigua-Souiai ,

    Roles Data curation, Formal analysis, Visualization

    ‡ EHS and HS also contributed equally to this work.

    Affiliation Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia

  • Hejer Souguir ,

    Roles Validation

    ‡ EHS and HS also contributed equally to this work.

    Affiliation Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia

  • Alia Yaacoub,

    Roles Investigation, Methodology, Resources

    Affiliations Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia, Service de parasitologie, EPS Farhat Hached, Faculté de Médecine de Sousse, Université de Sousse, Sousse, Tunisia

  • Oussaïma El Dbouni,

    Roles Investigation

    Affiliation Department of Infectious Diseases, Rafik Hariri Hospital, Beirut, Lebanon

  • Zoubir Harrat,

    Roles Resources

    Affiliation Laboratoire d’Eco-épidémiologie Parasitaire et Génétique des Populations, Institut Pasteur d’Algérie, Algiers, Algeria

  • Maowia M. Mukhtar,

    Roles Conceptualization, Resources

    Affiliation Bioscience Research Institute, Ibn Sina University, Khartoum, Sudan

  • Moncef Ben Said,

    Roles Conceptualization, Resources

    Affiliation Service de parasitologie, EPS Farhat Hached, Faculté de Médecine de Sousse, Université de Sousse, Sousse, Tunisia

  • Nabil Haddad,

    Roles Investigation, Methodology

    Affiliation Laboratory of Immunology and Vector-Borne Diseases, Faculty of Public Health Lebanese University, Hadath, Lebanon

  • Akila Fathallah-Mili,

    Roles Conceptualization, Investigation, Methodology, Resources

    Affiliations Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia, Service de parasitologie, EPS Farhat Hached, Faculté de Médecine de Sousse, Université de Sousse, Sousse, Tunisia

  • Ikram Guizani

    Roles Conceptualization, Funding acquisition, Methodology, Supervision, Writing – review & editing

    ikram.Guizani@pasteur.tn

    Affiliation Laboratory of Molecular Epidemiology and Experimental Pathology, Institut Pasteur de Tunis, Université de Tunis El Manar, Tunisia

Abstract

Background

Dipeptidyl peptidase III (DPPIII) member of M49 peptidase family is a zinc-dependent metallopeptidase that cleaves dipeptides sequentially from the N-terminus of its substrates. In Leishmania, DPPIII, was reported with other peptidases to play a significant role in parasites’ growth and survival. In a previous study, we used a coding sequence annotated as DPPIII to develop and evaluate a PCR assay that is specific to dermotropic Old World (OW) Leishmania species. Thus, our objective was to further assess use of this gene for Leishmania species identification and for phylogeny, and thus for diagnostic and molecular epidemiology studies of Old World Leishmania species.

Methodology

Orthologous DDPIII genes were searched in all Leishmania genomes and aligned to design PCR primers and identify relevant restriction enzymes. A PCR assays was developed and seventy-two Leishmania fragment sequences were analyzed using MEGA X genetics software to infer evolution and phylogenetic relationships of studied species and strains. A PCR-RFLP scheme was also designed and tested on 58 OW Leishmania strains belonging to 8 Leishmania species and evaluated on 75 human clinical skin samples.

Findings

Sequence analysis showed 478 variable sites (302 being parsimony informative). Test of natural selection (dN-dS) (-0.164, SE = 0.013) inferred a negative selection, characteristic of essential genes, corroborating the DPPIII importance for parasite survival. Inter- and intra-specific genetic diversity was used to develop universal amplification of a 662bp fragment. Sequence analyses and phylogenies confirmed occurrence of 6 clusters congruent to L. major, L. tropica, L. aethiopica, L. arabica, L. turanica, L. tarentolae species, and one to the L. infantum and L. donovani species complex.

A PCR-RFLP algorithm for Leishmania species identification was designed using double digestions with HaeIII and KpnI and with SacI and PvuII endonucleases. Overall, this PCR-RFLP yielded distinct profiles for each of the species L. major, L. tropica, L. aethiopica, L. arabica and L. turanica and the L. (Sauroleishmania) L. tarentolae. The species L. donovani, and L. infantum shared the same profile except for strains of Indian origin. When tested on clinical samples, the DPPIII PCR showed sensitivities of 82.22% when compared to direct examination and was able to identify 84.78% of the positive samples.

Conclusion

The study demonstrates that DPPIII gene is suitable to detect and identify Leishmania species and to complement other molecular methods for leishmaniases diagnosis and epidemiology. Thus, it can contribute to evidence-based disease control and surveillance.

Author summary

Leishmaniases correspond to a group of neglected tropical diseases caused by protozoan parasites members of more than 20 Leishmania species. Leishmaniases are epidemiologically complex diseases. Their complexity relates to the diversity of species and transmission cycle components, often poorly elucidated. Risk factors for changing eco-epidemiological profiles include environmental and climate changes, migrations and conflicts, contributing to leishmaniases emergence in disease-free areas, leading to co-sympatry of the major pathogens in the same transmission areas and even foci. Taxonomical identification of the parasites is a central issue for patient management, molecular epidemiology and thus disease control and surveillance. Thus, development of adequate specific and sensitive DNA tools to identify and classify Leishmania parasites for surveillance and control remains a research priority. The DPPIII gene is an essential gene conserved in all Leishmania species. By comparative sequence analyses of the DPPIII gene fragments and the development of DPPIII-based PCR RFLP assays, we showed its potential for molecular diagnosis, epidemiology and taxonomy. We conclude that this molecular marker can be used to complement other molecular methods for OW Leishmania species identification and discrimination and can contribute to evidence- based disease control and surveillance.

Introduction

Leishmaniases correspond to a group of diseases caused by more than 20 protozoan Leishmania species parasites. They are transmitted to humans by the bites of infected female phlebotomine sandflies. Three main disease forms are encountered: cutaneous, visceral and mucocutaneous leishmaniases. An estimated 1 to 1.5 million cutaneous leishmaniasis (CL) and 500000 visceral leishmaniasis (VL) cases are annually reported worldwide [1,2]. Species identification of Leishmania is crucial for selecting the most appropriate therapy to be administrated to patients [35]. Parasite classification was initially based on eco-biological criteria including vectors, geographical distribution, clinical manifestation and disease tropism [69]. But nowadays, it is based on molecular technologies [10]. Leishmaniases are considered as taxonomically and epidemiologically complex diseases. Their complexity relates to the co-endemicity of the major pathogens [11,12], the diversity of transmission cycle components, often poorly elucidated, environmental and climate changes [13,14], migrations and conflicts [15], all contributing to leishmaniases emergence in disease-free areas [16] and potential involvement of novel hosts [17,18].

In Tunisia and worldwide, these diseases show emerging trends in their epidemiology resulting in changes in their distribution illustrated by the spread of the species from their classical foci to neighboring areas [1924]. Identification of these emerging foci is important for disease control. This is mainly based on the use of adequate DNA tools for identifying and classifying Leishmania parasites. Advances in systematics and taxonomy methods have improved knowledge about parasite evolution, phylogeny, and molecular species identification. Leishmania species discrimination and evolutionary relationships have been inferred using a range of assays targeting molecular markers including coding regions such as for heat shock proteins (hsp 20 [25] or hsp70 [26]), glycoprotein 63 (gp63) [2729], cytochrome oxidase II [30,31], cysteine protease B [32,33], Tryparedoxine peroxidase [34], and non-coding regions such as the internal transcribed spacer (ITS) 1 and 2 [3537] and some repetitive regions [38]. Some assays are specific to certain taxonomic groups [30], others lack sensitivity [39] or specificity [34], or are only suitable for intra-specific analyses using for instance microsatellites markers [15,4042]. Increasing knowledge about evolutionary relationships is generated through genome projects [43,44] but also by resorting to well characterized markers that provide thorough supplementary data about inter and intra-species variations in clinical isolates. Such knowledge improves debated phylogeny and taxonomy [43,4547] and strengthens capacity for molecular epidemiology investigations.

Metallopeptidases constitute a diverse group of distinct evolutionary families. Dipeptidyl peptidase III (DPPIII) belongs to the M49 peptidase family of zinc-dependent enzymes that sequentially cleave dipeptides from the N-terminus of its substrates [48]. Studies suggested the involvement of the human DPPIII protein in protein turn over, oxidative stress and pain modulation and inflammation [49]. In Leishmania, the protein was characterized in L. braziliensis demonstrating its role in parasite survival in the vector, or host environments through peptide degradation and thus use of amino acids for energy production in stressful environment [50]. Presence of this protein was demonstrated in the secretome of L. donovani, which classified it as a candidate virulence factor that might be part of a stress response of the parasite [51]. A 2040bp DPPIII gene is present in all Leishmania and other trypanosomatids sequenced genomes [50]. However, the gene seemed absent in the Trypanosoma genus including T. brucei and T. cruzi species [50,52]. The L. braziliensis DPPIII translated sequence has a 65–80% similarity rate to other microorganisms belonging to Endotrypanum, Blechomonas, Crithidia and Leptomonas genera [50]. Within Leishmania, the gene is conserved in species infecting humans and other mammals, with predicted 88–89% amino acids sequence identity to the L. braziliensis enzyme. In a previous study, primers targeting one part of the gene were used to amplify L. major, L. tropica, L. arabica and L. aethiopica DNAs but the reaction did not yield any product with the viscerotropic L. infantum and L. donovani species DNAs nor with L. tarentolae [53]. In this study, we aimed at extending the validation of this gene as a potential target for DNA diagnosis of Old World Leishmania species including in countries which are also endemic for viscerotropic species. Therefore, we further investigated genetic diversity of the DPPIII gene and products of a new PCR in a panel of strains belonging to a range of pathogenic and non- pathogenic endemic species in the Old World to assess its relevance as a molecular marker for phylogeny and species identification, and thus to complement others for molecular epidemiology studies of Old World Leishmania species and disease surveillance.

Methods

Ethical statement

This work complies with the ethical standards as required by the ethical committee of Institut Pasteur de Tunis, lead institution in the studies (Ref. 2016/24/I/LRIPT04; 2016/13D/I/CIC).

The samples correspond to scrapings done in the frame of routine diagnosis of suspected CL patients with consent of patients in prospective studies (samples 1–13). In case of samples 14 to 75, they were also collected in the frame of routine diagnosis of suspected patients, the preserved samples’ remains were used retrospectively with the consent of the ethics committee of IPT with respect of patient anonymity.

Parasite strains and DNAs

In total, we analyzed 58 Leishmania DNA belonging to 8 Leishmania species isolated from a range of hosts: human cases, reservoirs or vectors, in various geographical origins, to cover a maximum of Old-World parasite species, available in our local DNA bank as detailed in S1 Table. They correspond to already characterized strains or DNAs by isoenzyme, and genomic RFLP [54] or PCR-RFLP ITS1 [55], obtained from reference centers in Montpellier or Rome, and in case of clinical isolates from health centers in Tunisia, Sudan and Algeria (S1 Table). The promastigote DNAs were extracted by phenol-chloroform procedure as described [54].

Human CL samples

Seventy-five DNAs extracted from clinical samples using DNA purification kits as recommended by the supplier (Qiagen QIAamp DNA Mini Kit) were used (S2 Table). The samples correspond to scrapings done in the frame of routine diagnosis of suspected CL patients with consent of patients in prospective studies (samples 1–13) in clinical departments in Tunisia (N = 10; University Hospital Farhat Hached, Sousse) and Lebanon (N = 3; Rafik Hariri Hospital, Beirut). In case of remains of lesion scrapings made during diagnosis at the clinical parasitology department of the Farhat Hached University hospital in Sousse, Tunisia (N = 62; samples 14–75), they were used with the consent of the ethics committee of IPT. This sampling includes CL cases and patients having other cutaneous/ dermal diseases.

CL was confirmed for 45 patients by direct examination (DE) under microscope (x1000) of Giemsa- stained smears and subsequent observation of amastigotes. Twenty- six were DE negative, and in case of 4 CL patients, the direct examination was not done or not reported on the data collection file. For molecular identification purposes, ITS1-PCR followed by HaeIII RFLP was used as an identification technique as previously described [55]. No- template reactions were the internal negative controls of the assays; positive reaction controls contained known Leishmania DNAs inputs that were at the limit of detection.

Performances of ITS1 PCR and DPPIII PCR assays were computed using DE as gold standard. Besides sensitivity and specificity, we generated the receiver operating curves (ROC) using the scikit-learn library under Python3. The area under curve (AUC) was calculated to estimate the ability of a given diagnosis test to correctly differentiate positive versus negative cases.

PCR design and analyses of the products

Available dipeptidyl peptidase III gene sequences of different Leishmania species (S3 Table) were retrieved from TriTrypDB Release 26 (14 Oct 15) and aligned using Geneious v3.6.2 computer program to investigate sequence polymorphisms including in restriction sites and design PCR based assays. Restriction sites having minimum effective recognition sequence length of 4 nucleotides were searched within the Restriction Enzyme Database (REBASE, http://rebase.neb.com/). Two generic primers were manually designed in a gene region selected upon this multiple alignment analysis in conserved parts of the gene (Fig 1) flanking a region showing extensive inter-species polymorphisms. Then to verify adequacy of this design, the primers were further analyzed and their sequence was adjusted for primer melting temperature, primer secondary structures including hairpins, self-dimers, and occurrence of cross-dimers in the primers pair using NetPrimer, an online primer analysis software (http://www.premierbiosoft.com/netprimer/). NCBI primer-blast analysis did not find a target template in Homo sapiens genome database. Finally, the retained primers cover the region of the gene defined by the positions 832 to 1494 and have the following sequences: DPPIII-F 5’-AGGAGTGGGTGAAGGATGTG-3’ and DPPIII-R 5’-CAGCAAGCAGAGGTACAGC-3’.

thumbnail
Fig 1. Schematic representation of the DPPIII gene and PCR primers positions.

1 & 2: Primers used by [53] (DPP F/DPP R). 3 & 4: Primers designed in this study (DPPIII-F/DPPIII-R) amplify a 662bp fragment.

https://doi.org/10.1371/journal.pntd.0009530.g001

The reaction was set up to amplify 0.2 ng of DNA in 25μL using a mix that contained 1x standard PCR buffer, 1.5 mM MgCl2, 200 μM of each 2’- deoxy-nucleoside 5’-triphosphate, 1U Taq Platinium DNA polymerase (Invitrogen), 0.3 μM of each primer. No template reactions were the negative PCR controls. The thermal cycling parameters of the assay were as follows: initial denaturation at 94°C for 3 min followed by 35 cycles consisting of 94°C for 1 min, 62°C for 2 min, and 72°C for 2 min, and a final extension step of 5 min at 72°C. Amplicons were visualized on a 1.5% agarose gel under UV light in presence of ethidium bromide.

Analytical detection limit of the assay was assessed testing serial ten-fold DNA dilutions, ranging from 20ng/μl to 0.002fg/μl, whereby 1μl was used for PCR.

The DPPIII F/R PCR products were also sequenced or digested using selected enzymes.

Direct sequencing of PCR products was performed on both strands using the DPPIII-F and DPPIII-R oligonucleotides, the BigDye Terminator v3.1 Cycle sequencing Kit (Applied Biosystem), and an ABI 3500 sequencer. The chromatograms were visualized and manually adjusted using DNA Baser sequence assembler v4 program (2013) (Heracle Biosoft, www.DnaBaser.com).

In case of RFLP analysis, the enzymes were first selected for their ability to differentiate OW Leishmania species with a minimum number of cuts and without generating too small fragments (<50bp) using predicted RFLP patterns visualized with the SnapGene 2.8.3 program from GSL Biotech, available at snapgene.com. Double digestions were done in 20μl reactions using 1 unit of each restriction enzyme, HaeIII and KpnI or SacI and PvuII, as recommended by the manufacturer (New England Biolabs Inc.) using 1x CutSmart buffer, and 10μl of reaction product. Incubations were done at 37°C for 2h. Then the restriction fragments were visualized on a 3% agarose gel.

Phylogenetic analyses

Phylogenetic analysis was foremost carried out with previously published sequences (S3 Table) with the software package MEGAX (Molecular Evolutionary Genetic Analysis across computing platforms [56]). Nineteen sequences corresponding to 16 Leishmania (representing 13 Leishmania species), one L. (Sauroleishmania) and two Leptomonas species (Leptomonas pyrrhocoris and Leptomonas seymouri) were aligned using ClustalW program, and MEGAX software was used to build phylogenetic trees. Then, sequenced DPPIII fragments of a selection of DNAs (S1 Table) were aligned with the published sequences for further phylogenetic analyses of the amplified fragment.

Evolutionary distances were computed with the substitution model Tamura 3-parameter method [57] with Gamma distribution variation rate [58] (shape parameter = 0.65) chosen using the Bayesian information criterion (BIC), as implemented in the MEGAX software. Monophyletic groups were supported by 1000 bootstrap resampling method [59]. UPGMA [60], Neighbor-Joining (NJ) [61], Minimum Evolution (ME) [62], and Maximum likelihood (ML) [63] phylogenetic trees were constructed as implemented in MEGAX software.

We also used MEGAX to analyze the extent of sequence variation by calculating the number of polymorphic and parsimony-informative sites. The number of synonymous substitutions by synonymous sites and non-synonymous substitutions by non-synonymous sites were calculated from averaging the distance estimation of the codon-based evolutionary divergence over all sequence pairs using Nei-Gojobori model [64] from which the subsequent difference between dN and dS was deduced. Standard errors were obtained by a bootstrap procedure (1000 replicates). We used codon-based Z-test of neutral evolution to assess if the gene is under a selective pressure. Average sequence composition and the average percentage of pairwise similarity over the alignment were determined using the Geneious 3.6.2 program.

Results

In silico analysis of the entire gene confirms taxonomical potential of DPPIII gene

We used phylogenetic analysis to assess taxonomical potential of the DPPIII gene and to detect species- specific single nucleotide polymorphisms (SNPs) and indels. We did multiple sequence alignments using the entire DPPIII gene (2040bp) of 19 published database sequences, covering 13 Leishmania, 1 L. (Sauroleishmania) and 2 Leptomonas species (S3 Table). Sequence analysis revealed 478 variable sites of which 302 (63%) were parsimony informative (S4 Table). To determine the selective evolutionary pressure on the predicted DPPIII protein, codon-based evolutionary divergence from averaging all sequence pairs was used to estimate the number of synonymous substitutions per synonymous site (dS = 0.199; SE = 0.003) and the number of non-synonymous substitutions per non-synonymous sites (dN = 0.035; SE = 0.012). The deduced difference between the non-synonymous and synonymous distances per site from averaging over all sequence pairs dN-dS is -0.164 (SE = 0.013). The probability of rejection of the null hypothesis of strict neutrality (dN = dS) using the codon- based Z test of neutrality was in favor of the deviation of the DPPIII gene from neutrality and tendency for a negative selection (dN<dS; p<0.05).

The topology of the Neighbor joining phylogenetic tree of these sequences matched taxonomy and known evolutionary relationships of the pertaining species. This corroborated the hypothesis regarding the phylogenetic and taxonomic potential of the DPPIII gene (Fig 2).

thumbnail
Fig 2. Maximum Likelihood tree constructed using published database kinetoplastid sequences of the entire DPPIII gene.

The evolutionary history of Leishmania parasites was inferred using the Maximum Likelihood method and Tamura-Nei model. The tree with the highest log likelihood (-8661.86) is shown. Initial tree(s) for the heuristic search were obtained automatically by applying Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using the Tamura-Nei model, and then selecting the topology with superior log likelihood value. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.4323)). The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. This analysis involved 19 nucleotide sequences. Codon positions included were 1st+2nd+3rd+Noncoding. There were a total of 2046 positions in the final dataset. Evolutionary analyses were conducted in MEGA X.

https://doi.org/10.1371/journal.pntd.0009530.g002

A valid generic PCR test and its analytic limit of detection

Then, to determine a smaller similarly informative part of the gene that could be used in molecular assays development, we looked at the mutations and restriction sites distribution as markers of species divergence. We selected a variable part of the gene showing 32.2% single nucleotide polymorphisms (SNPs) including variable restriction sites, that also encodes for the active site motives, HELLGH and EECRAE, which are involved in zinc binding. This region was flanked by interspecies conserved sequences that we used to design a novel primers pair, DPPIII-F/DPPIII-R, for the generic amplification of this part of the gene (Fig 1). It is where the most discriminatory restriction sites were located.

The primers amplified the expected 662bp fragment as tested on a representative DNA panel corresponding to OW Leishmania and L. (Sauroleishmania) species, having different geographical origins and hosts (S1 Table), which allowed their validation as generic primers (S1 Fig). In order to assess the analytical sensitivity of this PCR, we tested ten-fold serial DNAs dilutions ranging from 20ng/μl to 0.2fg/μl of three strains (IPT1, Ron44 and Bag17) representing L. infantum, L. major and L. tropica species respectively. The Fig 3 illustrates that depending on the species studied, analytical detection limit varied according to the species/ strain in the 200 – 2fg range. Assuming an average diploid genome mass of 80fg [65], the range of the analytic detection limit was 0.025–2,5 parasites.

thumbnail
Fig 3. PCR-DPPIII detection limit based on ten-fold serial DNA dilutions.

1: L. major (Ron44); 2: L. infantum (IPT1); 3: L. tropica (Bag17). For each DNA, ten- fold serial DNA dilutions starting from 0.02ng were input into the reaction.

https://doi.org/10.1371/journal.pntd.0009530.g003

Phylogenetic validation of the amplified 662bp DPPIII gene fragment

To further validate taxonomical potential of the selected 662bp DPPIII gene fragment and to possibly deduce species- specific DNA signatures, we successfully sequenced the amplicons of 53 Leishmania DNA representing different species and host or geographical origins (S1 Table). In addition, 16 Leishmania, 2 Leptomonas and 1 L. (Sauroleishmania) DPPIII sequences were retrieved from the TritrypDB database and included in the analysis (S3 Table). This amounts the sample used in this analysis to a total of 72 sequences of different genera, species and strains. Sequence alignment illustrating the variable sites and the associated haplotypes of all these DNA fragments are shown in S4 and S5 Tables. The sequence analyses of the studied DPPIII region among Leishmania species revealed the average nucleotide frequencies: T (18.1%), C (28%), A (20.7%), G (33.2%); indels (0.5%); GC content (61.3%) and the average pairwise similarity (96.3%). The selective evolutionary pressure on the 662 bp DPPIII fragment was determined, and it showed that as for the entire gene, the studied fragment is under negative selection (dN-dS = -0.178; SE = 0.014).

Twenty haplotypes were identified based on sequence polymorphisms; to each species tested corresponded at least one haplotype. We observed that some SNPs were only seen in species- or group of species- specific haplotypes, which could correspond to DNA signatures of relevance to taxonomy. In addition, in the case of L. tropica and L. donovani, some SNPs appeared as specific to some of their strains, may be linked to the geographical origins (S5 Table).

The Maximum Likelihood based on the amplified DPPIII sequence alignment (Fig 4) showed a separation between the New World (L. mexicana, L. braziliensis, L. panamensis and L. enriettii) and the Old- World species (L. major, L. tropica, L. aethiopica, L. turanica, L. gerbilli, L. arabica and L. infantum/ L. donovani). Among the Old- World species, 6 clusters were observed that are congruent to L. major, L. tropica, L. aethiopica, L. arabica, L. turanica, L. tarentolae species, and another one to the L. infantum and L. donovani species complex. This is in accordance with the phylogeny constructed using the entire DPPIII gene.

All the groups mentioned above were also observed using UPGMA, NJ and ME phylogenies (S2S5 Figs) indicating that the derived groups are robust and do not depend on the evolutionary method used.

thumbnail
Fig 4. Evolutionary relationships of taxa.

The evolutionary history of Leishmania parasites was inferred from the 662bp fragment of the DPPIII gene using the Maximum Likelihood method and Tamura 3-parameter model. The tree with the highest log likelihood (-2620.35) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. Initial tree(s) for the heuristic search were obtained automatically by applying Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using the Tamura 3 parameter model, and then selecting the topology with superior log likelihood value. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.3096)). The tree is drawn to scale with branch lengths measured in the number of substitutions per site. This analysis involved 72 nucleotide sequences. Codon positions included were 1st+2nd+3rd+Noncoding. There was a total of 662 positions in the final dataset. Evolutionary analyses were conducted in MEGA X.

https://doi.org/10.1371/journal.pntd.0009530.g004

At the intra-specific level, polymorphisms were observed for different taxa. L. tropica showed two groups according to their geographical origins. The first group gathers the Mediterranean strains LA28, AM, CJ and Leep0920, and the second, more heterogeneous, includes strains from the Middle East and India. The SNPs that differentiate these subgroups are situated at the positions 1131 and 1350 of the DPPIII gene (S5 Table). In the second case, 2 SNPs (1103 and 1187) separated the Indian from the African and Middle Eastern L. donovani strains that were grouped with the L. infantum strains (S5 Table).

PCR-RFLP analysis

In parallel, we took profit of the variable restriction sites to develop PCR-RFLP based identification as an alternative to the sequence analysis. Indeed, to develop the species identification scheme, we used published sequences of Old World Leishmania species (L. major, L. tropica, L. aethiopica, L. infantum, L. donovani, L. arabica and L. turanica) and L. (Sauroleishmania) tarentolae to identify within the 662bp fragment, the variable restriction sites, which are specific to species- or group of species. We selected four restriction enzymes (RE) (HaeIII, KpnI, PvuII and SacI) that distinguish between the different Leishmania species/group of species according to predicted double- digestion restriction fragments length polymorphisms, as detailed in Table 1 and visualized using the SnapGene 2.8.3 program on S6 Fig. Leishmania species identification is possible using the scheme on Fig 5 where up to 2 steps could be taken to identify the Leishmania species.

thumbnail
Fig 5. PCR-RFLP algorithm for identification of Old World Leishmania species using DPPIII gene as target and DPPIII F/R primers pair.

A: African, ME: Middle Eastern.

https://doi.org/10.1371/journal.pntd.0009530.g005

thumbnail
Table 1. In silico predicted restriction fragments of the DPPIII F/R PCR products of the TritrypDB database sequences.

https://doi.org/10.1371/journal.pntd.0009530.t001

In the first step, the double digestion with HaeIII and KpnI RE would identify L. major, L. tropica, L. aethiopica, L. tarentolae, L. donovani but would not distinguish within the group made of L. infantum, L. turanica and L. arabica (S5A Fig). In a second step, the unresolved DNAs would be digested with SacI and PvuII to differentiate and identify L. infantum, L. turanica and L. arabica (S5B Fig). L. turanica and L. gerbilli share the same RFLP patterns.

We validated experimentally this scheme on the selection of DNAs presented on S1 Table corresponding to well- characterized Leishmania strains (Table 2). The analysis showed that as expected the double HaeIII and KpnI digestion allowed the differentiation between L. major, L. tropica and L. aethiopica and their identification while L. infantum, L. arabica and L. turanica species showed the same restriction pattern. Unexpectedly, this pattern was also shared with the African and Middle Eastern L. donovani strains, and as predicted by the in silico analysis, the unique strain tested from India, MHOM/IN/00/DEVI, had a distinctive RFLP pattern (Figs 6A and S5A). This is in accordance with the sequence analyses of the corresponding fragments. The unique L. tarentolae tested DNA showed a unique profile, different from species of the Leishmania subgenus (Fig 6A). Except for the case of L. donovani, for each species, all tested strains had consistently the same RFLP pattern whatever were their geographical origin, their host or their isolation date, which allows considering these patterns as species-specific (Table 2). The L. infantum, L. donovani (African and Middle Eastern), L. arabica and L. turanica species DNA products were used in a PvuII and SacI double digestion. We obtained a different and unique restriction pattern for each of L. arabica, L. turanica and L. infantum species DNA as predicted by the in silico analysis; the African and Middle Eastern L. donovani and L. infantum DNAs shared the same profile (Fig 6B).

thumbnail
Fig 6.

(A) RFLP patterns of the HaeIII/KpnI double digested, amplified DPPIII fragments. L. major, L. tropica, L. aethiopica, L. tarentolae, and Indian L. donovani presented distinctive profiles while African and Middle Eastern L. donovani, L. infantum, L. turanica and L. arabica shared the same patterns. (B) RFLP patterns of the PvuII/SacI double digested, amplified DPPIII fragments. The L. infantum, L. arabica and L. turanica DNAs were distinguished by their PCR- RFLP profiles. The African and Middle Eastern L. donovani and L. infantum DNAs presented the same patterns. MW: 50bp Molecular Weight. L. m: L. major; L. t: L. tropica; L.ae: L. aethiopica; L. i: L. infantum; *Brazilian L. infantum strain (also known as L. chagasi); L. d: L. donovani; *East African L. donovani strain (also known as L. archibaldi) ; L. tu: L. turanica; L. ar: L. arabica; L. tar: L. tarentolae.

https://doi.org/10.1371/journal.pntd.0009530.g006

PCR-RFLP identification of parasites within suspected human cutaneous leishmaniasis samples

To validate use of our DPPIII PCR and RFLP assays for Leishmania parasites detection and identification in clinical samples, we tested 75 DNAs of cutaneous samples taken from patients referred to clinicians for CL diagnosis during the 2010–2018 period. We also used ITS1 PCR-RFLP on the same sample and compared the results to direct smear examination (DE) as gold standard. The S2 Table gathers all the results obtained with these samples. As shown on Fig 7A, out of the 75 samples, 48 were ITS1 PCR+ and 46 DPPIII PCR+. The distribution of positive results across the 3 tests on all the samples is shown on Fig 7B. Six samples that were ITS1 PCR- were DPPIII PCR+. However, 8 ITS1 PCR+ samples were DPPIII PCR-. Two DE+ samples were negative by both PCR assays. Measured sensitivity using DE as gold standard was 82.22% for DPPIII and 88.8% for ITS1. Specificity was estimated to be 76.92% for both methods. Computed areas under the curve (AUC_ROC) were 0.80 and 0.83 respectively (Fig 7C).

thumbnail
Fig 7.

(A) Results of CL samples. (B)Venn Diagram showing concordance of detected Leishmania parasites on 71 CL samples between DE, PCR-ITS1 and PCR-DPPIII. (C) Performances measuring of DPPIII and ITS1 PCRs versus Direct examination using ROC curves. ND: Not determined.

https://doi.org/10.1371/journal.pntd.0009530.g007

The RFLP analysis of the amplicons identified 39 samples out of the 46 DPPIII positives and 47 among the 48 ITS1 positives. S7 Fig illustrates some of the identification results made by the DPPIII PCR HaeIII/KpnI RFLP. The two molecular identification methods had 100% agreement as regards Leishmania species identification (S2 Table).

Discussion

There is need for novel molecular tools to respond to challenges associated to leishmaniases and continuous changes in their epidemiology at local and global levels [12,20,21,66]. Indeed, molecular tools are instrumental to define strategies for disease control as they allow identification and characterization of transmitted strains or species and use of genetic diversity for spatial and temporal tracking and comparison of strains. The use of DNA markers is the alternative of choice to laborious, costly and time consuming MLEE methods [25,67,68]. During the last decade, this technique was increasingly replaced by other efficient molecular approaches such as MLST [6972] and MLMT [4042] and whole genome sequences [73,74] to study genetic diversity, epidemiology, population genetics and evolution of Leishmania parasites. However, these methods remain laborious and expensive. So, referring to individual markers for Leishmania molecular characterization, epidemiology and phylogenetic studies remains an alternative of choice despite the lower resolution for population genetic studies. When single markers evolution reflects Leishmania parasites phylogeny, their sequence analysis or use are effective in a congruent species assignment to MLST or MLEE identification [25,39,75].

Dipeptidyl Peptidase III protein coding sequence has been identified as a potential target for DNA parasite identification. The gene is present in the genome of Leishmania, Endotrypanum, Crithidia, Leptomonas and Blechomonas [50] but is lost in the Trypanosoma genus including T. brucei and T. cruzi human pathogens [52]. Interestingly, although the DNA sequence is present in the L. (Sauroleishmania) subgenus represented by L. tarentolae species and annotated as a putative DPPIII protein, the CDS is initiated much downstream from the ATG site in other Leishmania genes, resulting in a shorter coding region (470bp vs 2040bp).

We previously used DPPIII gene as a target to amplify a 664bp fragment from L. major, L. tropica, L. arabica and L. aethiopica, while the PCR was negative for strains of L. infantum, L. donovani, and L. tarentolae species [53]. Thus, the main objective of this study was to further characterize and validate the DPPIII gene as a molecular marker for Old World Leishmania taxonomy and molecular epidemiology. For this purpose, we used sequences or strains belonging to a range of pathogenic, non- pathogenic (e.g. L. turanica, L. arabica, L. gerbilli, L. tarentolae) Leishmania species to humans, and monoxenous kinetoplastids (Leptomonas pyrrocoris, Leptomonas seymouri) that also could be opportunistic to Leishmania species in humans [7678]. We attempted to cover different Leishmania taxa, geographical origins, hosts and diseases resulting in a representative set of DNAs of pathogenic Leishmania species in the OW; L. aethiopica and non-pathogenic Leishmania were represented each by a unique DNA.

Leishmania genus evolution is mainly marked by the development of a digenetic life cycle and the introduction of the intracellular amastigote stage in the development cycle [79]. In this study, the phylogenetic analysis of the entire gene has shown a congruent grouping of the parasites to the genus and subgenus level [79], or Leishmania MLSA groups [80] and the reported species assignment of the parasites. The sequence analysis of the 662bp fragment of the DPPIII gene that contains the zinc binding motifs of the active site, in 69 Leishmania DNAs and 19 Database sequences, also showed a clear separation between the New World (L. mexicana, L. braziliensis, L. panamensis and L.enriettii) and the Old World species (L. major, L. tropica, L. aethiopica, L. turanica, L. gerbilli, L. arabica and L. infantum/ L. donovani). L. tarentolae was grouped with Old World species which is in accordance with the most recent classification [10].

Interestingly, the analysis also suggested that the gene is under negative selection, a known feature of essential genes [81] that corroborates the physiological importance of the protein [50]. This feature, also observed for the 662bp sequenced fragments, thus infers a potential stability of the observed haplotypes. Among the twenty haplotypes observed, some species had more than one sequence type, some mutations seemed to be species specific and may constitute potential bioinformatics species signatures.

The two sequence types that were observed for L. major were different from those of the closely related, supposedly non- pathogenic to human, L. arabica, L. turanica and L. gerbilli. These species, known to be endemic in Middle East and Central Asia, share the same transmission areas and even cycle components as L. major; mixed infections were even described [8286]. The sequences of these 3 species were also different from one another.

Two sequences types were also observed for L. infantum, and L. donovani which is a closely related species. These parasites have a still debated taxonomy. While some studies described them as two distinct groups [87,88], others considered that they should be considered as one species complex (L. donovani complex) due to the low genetic diversity between them [43,47]. In addition, in a genome-wide global study that included Indian, East African and Middle Eastern strains, the phylogenetic reconstruction of the L. donovani complex based on SNP variations clearly separated these parasites into five major groups that coincided with their geographic origin [74]. Microsatellites and enzyme coding genes analyses have also described the Indian strains of L. donovani as a distinct genetic sub-group from East African strains [89,90]. In another study based on cpB sequence analysis, Sudanese and Ethiopian L. donovani strains constituted a distinct subgroup from other L. donovani complex strains that included L. infantum from Tunisia and L. donovani from Kenya, India, and Iran [33]. In this study, two specific SNPs differentiated the Indian L. donovani strains from the North (L. infantum) and East African (L. donovani) ones, corroborating close relatedness of the African parasites of this species complex, not allowing to identify an L. infantum sequence type.

The DPPIII fragment sequence analyses also showed that, among the Old World studied species, L. tropica was the most polymorphic. Indeed, we noticed the presence of 5 haplotypes corresponding to 3 different geographical origins. The first haplotype was observed for the Mediterranean strains from Tunisia and Greece; three other haplotypes corresponded to Middle Eastern parasites, and the last one was shared between the unique Indian strain and a Middle Eastern one. Geographical subdivision within L. tropica was also observed in other studies using more sophisticated methods such as microsatellites [91] and genome sequence analyses [92].

We tested only one DNA in case of L. aethiopica and non- pathogenic Leishmania species. Our findings would need to be validated using additional strains. However, we have observed that each of their sequences were sufficiently different from the other studied ones to be distinguished. L. (Sauroleishmania), including Leishmania tarentolae species, are reptile parasites, described in different evolution studies as deriving from the mammalian forms of Leishmania [89,90,93], and recently recognized as a subgenus of the Leishmania genus [10]. L. tarentolae was detected and identified in hard and soft tissues of a 300 years old Brazilian mummy, emphasizing the fact that these parasites might infect and visceralize in humans [94]. This species is often overlooked in epidemiological investigations and little work describes the impact of its presence in transmission foci. Ability to detect and identify non-pathogenic species is cornerstone to gain insights into established Leishmaniases transmission cycles or in epidemiological investigations.

In this study, we further validated DPPIII gene as a Leishmania taxonomy marker by developing a new DPPIII PCR-RFLP based parasite identification where one or two steps would be needed for taxonomical assignments of the parasites in the Old World. Thereby, HaeIII/KpnI double digestion allows successful identification of L. major, L. tropica, L. aethiopica, the group representing L. infantum and L. donovani species, and even the species L. tarentolae of the subgenus L. (Sauroleishmania). Depending on the studied geographical region, this first analysis step could be sufficient to identify the endemic species making the assays cheaper and quicker. However, exceptions will be if imported species are observed and studied, or, in case of Asian and Middle Eastern regions, where L. turanica and L. arabica are respectively co-sympatric to L. major, and co-endemic with L. infantum or L. donovani. A second SacI/PvuII digestion step will precise the species identity that could not be ascertained during the first step. Although the L. gerbilli and L. turanica sequences are different, the selected enzymes did not generate discriminating RFLP patterns. Therefore, for these parasite species the sequence analysis is more appropriate for their identification.

In previous studies, PCR-RFLP analysis using a single marker like gp63 [27], cytB [95], and hsp70 coding sequences [26] could not differentiate between L. tropica and L. aethiopica species. However it was possible to differentiate them using species- specific PCR and MLMT assays [96,97]. In another study, discrimination of these two species required the concomitant analysis of two markers, hsp20 and hsp70 sequences [25]. Here, we were able to distinguish between the two species by a double digestion (HaeIII & KpnI) of a single maker, and we recommend the validation of its use for the identification of L. aethiopica. Using ITS1 PCR-HaeIII RFLP method [55], L. aethiopica digestion profile looked remarkably like the L. infantum/L.donovani species profile. It needed the use of metaphor agarose gel electrophoresis or a second digestion step (using CfoI) to differentiate these species. In case of DPPIII-RFLP assay, we recommend use of 2% - 3% agarose gel electrophoresis to achieve optimal resolution of the L. aethiopica RFLP pattern so it can be clearly distinguished from the L. donovani one (S8 Fig).

In addition, as expected from the sequence analyses, L. infantum and East African L. donovani strains shared the same RFLP pattern. Interestingly, RFLP of single markers like HSP70 did not differentiate these two species [25,98]. However, using PCR-RFLP ITS1-HaeIII, a slight size difference was observed in RFLP patterns between L. infantum and L. donovani when a prolonged 2% metaphor agarose gel electrophoresis was performed [55]. Differences in RFLP patterns between the two species was also observed when intragenic gp63 PCR was followed by an MscI restriction, and when SalI was used the 3 species L. infantum, L. donovani and L. archibaldi (Sudanese and Ethiopian L. donovani) presented distinct profiles. But Southern blotting and radioactive hybridization were required to allow full discriminative resolution of the digestion profiles, notably to detect the small sized bands that had faint intensity. An alternative for detection of these bands was the use of specific conditions for the electrophoresis (time, %, agarose brand) [27].

The application of DPPIII PCR on clinical samples demonstrated its ability to detect Leishmania parasites in DE+ human samples without reacting with human DNA as shown in Leishmania negative samples. The performances of this method were shown close to those of ITS1 PCR using direct examination diagnosis as gold standard, although sensitivity of ITS1 was higher. Their AUC ROC also pointed to the ability of these tests to distinguish True Positive from False Positive cases. The performance difference could be due to copy number of the targets, here a single copy DPPIII gene versus multi-copy ITS1, or also by the fact that our target sequence (662bp) is longer than the ITS1 one (300-350bp). In a similar context, reducing the target length was shown to greatly improve the sensitivity of PCR assays [98]. Interestingly, the DPPIII PCR was able to detect 3 DE+ samples that were not detected by ITS1 PCR despite the repeated nature of this target. Multicopy or repetitive sequences, notably intergenic and spacer regions, could be prone to sequence variations and more rapid accumulation of mutations than in coding regions [99]. Interestingly, hypothesis of occurrence of mutations in priming sites of ITS2 sequences was proposed to explain lack of amplification in infected samples [100]. Heterogeneity of ITS1 sequences was the cause of unreadable sequences [101]. RFLP patterns of multicopy sequences (e.g.Cpb) could also be difficult to analyze due to complexity created by isogene variations [39]. In case of the samples here tested, RFLP profiles obtained with both DPPIII and ITS1 amplified products were relatively easy to analyze, although in other ongoing studies it was not uncommon for us to observe difficult to interpret ITS1 RFLP profiles hampering Leishmania identification. We could identify the Leishmania species using one RFLP step in case of 39 DPPIII+ samples, while in case of 7 the amplicons were in too low amount to be further processed.

In the present study, we demonstrated by a phylogenetic approach that the DPPIII gene is a suitable marker for taxonomy and demonstrated ability of the developed DPPIII PCR to detect and identify parasites within human samples. So it could be used for diagnosis and molecular epidemiology studies to complement other methods including the ITS1 PCR assay. Indeed, it is recommended to use a combination of markers and techniques to provide an accurate diagnosis [102] and to achieve the aimed taxonomy resolution [10,102] in case of migration, travel or emergence.

Conclusion

Application of molecular identification methods for robust Leishmania taxonomical identification constitutes a key step to conduct molecular epidemiology investigations, to monitor disease outbreak and spread, and to implement rapid and adequate preventive measures. Molecular analyses of the DPPIII gene demonstrate it to be a plausible complement / alternative for such purposes in North Africa. Further investigations need be done to assess gene diversity in the under-represented species of this study, especially L. donovani in the Indian subcontinent, L. aethiopica in Africa, and L. turanica, L. gerbilli and L. arabica in Asian and Middle Eastern foci. This marker may contribute to advance in the fight against leishmaniases.

Data bases

  1. TritrypDB (https://tritrypdb.org/tritrypdb/)
  2. The Restriction Enzyme Database (REBASE) (http://rebase.neb.com/)

Softwares

  1. Geneious v3.6.2 (www.geneious.com)
  2. NetPrimer (http://www.premierbiosoft.com/netprimer/)
  3. MEGA 6 (https://www.megasoftware.net/)
  4. SnapGene (www.snapgene.com)
  5. DNA Baser sequence assembler v4 program (www.DnaBaser.com)

Supporting information

S1 Fig. Generic amplification of a 662bp DPPIII fragment from the studied Leishmania species.

MW: 100bp Molecular Weight. L.m: L. major; L.t: L. tropica; L.ae: L. aethiopica; L.i: L. infantum; *Brazilian L. infantum strain(also known as L. chagasi); L.d: L. donovani; **East African L. donovani strain (also known as L. archibaldi); L.tu: L. turanica; L.ar: L. arabica; L.tar: L. tarentolae.

https://doi.org/10.1371/journal.pntd.0009530.s001

(TIF)

S2 Fig. Evolutionary analysis of Leishmania parasites using the Maximum Likelihood method.

The evolutionary history of Leishmania parasites was inferred by using the Maximum Likelihood method and Tamura 3-parameter model. The bootstrap consensus tree inferred from 1000 replicates is taken to represent the evolutionary history of the taxa analyzed. Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches. Initial tree(s) for the heuristic search were obtained automatically by applying Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using the Tamura 3 parameter model, and then selecting the topology with superior log likelihood value. A discrete Gamma distribution was used to model evolutionary rate differences among sites (5 categories (+G, parameter = 0.3096)). This analysis involved 72 nucleotide sequences. Codon positions included were 1st+2nd+3rd+Noncoding. There were a total of 662 positions in the final dataset. Evolutionary analyses were conducted in MEGA X.

https://doi.org/10.1371/journal.pntd.0009530.s002

(TIF)

S3 Fig. Evolutionary relationships of Leishmania parasites using the Neighbor-Joining method.

The bootstrap consensus tree inferred from 1000 replicates is taken to represent the evolutionary history of the taxa analyzed. Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches. The evolutionary distances were computed using the Tamura 3-parameter method and are in the units of the number of base substitutions per site. The rate variation among sites was modeled with a gamma distribution (shape parameter = 0.65). This analysis involved 72 nucleotide sequences. Codon positions included were 1st+2nd+3rd+Noncoding. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There were a total of 662 positions in the final dataset. Evolutionary analyses were conducted in MEGA X.

https://doi.org/10.1371/journal.pntd.0009530.s003

(TIF)

S4 Fig. Evolutionary relationships of Leishmania parasites using the Minimum Evolution method.

The bootstrap consensus tree inferred from 1000 replicates is taken to represent the evolutionary history of the taxa analyzed. Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches. The evolutionary distances were computed using the Tamura 3-parameter method and are in the units of the number of base substitutions per site. The rate variation among sites was modeled with a gamma distribution (shape parameter = 0.65). The ME tree was searched using the Close-Neighbor-Interchange (CNI) algorithm at a search level of 1. The Neighbor-joining algorithm was used to generate the initial tree. This analysis involved 72 nucleotide sequences. Codon positions included were 1st+2nd+3rd+Noncoding. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There were a total of 662 positions in the final dataset. Evolutionary analyses were conducted in MEGA X.

https://doi.org/10.1371/journal.pntd.0009530.s004

(TIF)

S5 Fig. Evolutionary relationships of Leishmania parasites using the UPGMA method.

The bootstrap consensus tree inferred from 1000 replicates is taken to represent the evolutionary history of the taxa analyzed. Branches corresponding to partitions reproduced in less than 50% bootstrap replicates are collapsed. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (1000 replicates) are shown next to the branches. The evolutionary distances were computed using the Tamura 3-parameter method and are in the units of the number of base substitutions per site. The rate variation among sites was modeled with a gamma distribution (shape parameter = 0.65). This analysis involved 72 nucleotide sequences. Codon positions included were 1st+2nd+3rd+Noncoding. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There were a total of 662 positions in the final dataset. Evolutionary analyses were conducted in MEGA X.

https://doi.org/10.1371/journal.pntd.0009530.s005

(TIF)

S6 Fig. In silico predicted DPPII PCR-RFLP patterns.

A. After double digestion with HaeIII&KpnI endonucleases. B. After double digestion with SacI and PvuII of the PCR products. The parasites that could not be distinguished by HaeIII/KpnI digestion are differentiated by the SacI/PvuII restriction. MW: 100 bp Molecular Weight. L.m: L. major, L.i: L. infantum, L.t: L. tropica, L.d: L. donovani, L.ae: L. aethiopica, L.ar: L. arabica, L.tu: L. turanica, L.ta: L. tarentolae.

https://doi.org/10.1371/journal.pntd.0009530.s006

(TIF)

S7 Fig. Identification of Leishmania species in clinical material using DPPII-PCR and restriction enzyme analysis.

DNA was isolated directly from clinical samples, amplified by DPPIII-PCR and double digested with HaeIII-KpnI. MW: 50bp Molecular Weight. L. m: L. major (IL24), L. i: L. infantum (Drep 5) and L. t: L. tropica (LA28). Lanes 1–13: Clinical samples.

https://doi.org/10.1371/journal.pntd.0009530.s007

(TIF)

S8 Fig.

(A) Second round digestion of L. aethiopica strain. MW: 50bp Molecular weight; L. i: L. infantum (LV08); L. d: L. donovani (L1005); L. tu: L. turanica (95A); L. ara: L. arabica (J238); L. aeth: L. aethiopica (L100). (B) Two round digestion of L. aethiopica strain visualized on a 2% agarose gel. Lane 1: HaeII-KpnI double digestion, Lane 2: SacI-PvuII double digestion.

https://doi.org/10.1371/journal.pntd.0009530.s008

(TIF)

S3 Table. Selection of 19 gene sequences taken from the public TritrypDB databases and used for the typing scheme elaboration.

https://doi.org/10.1371/journal.pntd.0009530.s011

(DOCX)

S4 Table. Variable sites in the DPPIII gene between Leishmania species.

https://doi.org/10.1371/journal.pntd.0009530.s012

(XLSX)

S5 Table. Variable sites in the 662bp fragment and the associated haplotypes of the studied Leishmania species.

https://doi.org/10.1371/journal.pntd.0009530.s013

(XLSX)

References

  1. 1. Alvar J, Vélez ID, Bern C, Herrero M, Desjeux P, Cano J, et al. Leishmaniasis Worldwide and Global Estimates of Its Incidence. PLoS ONE. 2012;7. pmid:22693548
  2. 2. WHO. WHO Report on Global Surveillance of Epidemic-prone Infectious Diseases—Leishmaniasis. 2019. Available: https://www.who.int/csr/resources/publications/CSR_ISR_2000_1leish/en/
  3. 3. Arevalo J, Ramirez L, Adaui V, Zimic M, Tulliano G, Miranda-Verástegui C, et al. Influence of Leishmania (Viannia) species on the response to antimonial treatment in patients with American tegumentary leishmaniasis. J Infect Dis. 2007;195: 1846–1851. pmid:17492601
  4. 4. Mosimann V, Neumayr A, Hatz C, Blum JA. Cutaneous leishmaniasis in Switzerland: first experience with species-specific treatment. Infection. 2013;41: 1177–1182. pmid:23835701
  5. 5. de Menezes JPB, Guedes CES, Petersen AL de OA, Fraga DBM, Veras PST. Advances in Development of New Treatment for Leishmaniasis. BioMed Res Int. 2015;2015: 1–11. pmid:26078965
  6. 6. Bray RS, Ashford RW, Bray MA. The parasite causing cutaneous leishmaniasis in Ethiopia. Trans R Soc Trop Med Hyg. 1973;67: 345–348. pmid:4778189
  7. 7. Lumsden WH. Letter: Biochemical taxonomy of Leishmania. Trans R Soc Trop Med Hyg. 1974;68: 74–75. pmid:4818670
  8. 8. Pratt DM, David JR. Monoclonal antibodies that distinguish between New World species of Leishmania. Nature. 1981;291: 581–583. pmid:6787433
  9. 9. Lainson R, Shaw JJ. Evolution, classification and geographical distribution. Peters, W., Killick-Kendrick, R. The Leishmaniases in Biological and Medicine. Peters, W., Killick-Kendrick, R. London; 1987. pp. 1–120.
  10. 10. Akhoundi M, Downing T, Votýpka J, Kuhls K, Lukeš J, Cannet A, et al. Leishmania infections: Molecular targets and diagnosis. Mol Aspects Med. 2017;57: 1–29. pmid:28159546
  11. 11. Hakkour M, Hmamouch A, El Alem MM, Rhalem A, Amarir F, Touzani M, et al. New epidemiological aspects of visceral and cutaneous leishmaniasis in Taza, Morocco. Parasit Vectors. 2016;9. pmid:26732186
  12. 12. Jaouadi K, Bettaieb J, Bennour A, Salem S, Rjeibi MR, Chaabane S, et al. First Report on Natural Infection of Phlebotomus sergenti with Leishmania tropica in a Classical Focus of Leishmania major in Tunisia. Am J Trop Med Hyg. 2017;97: 291–294. pmid:28719307
  13. 13. Kholoud K, Denis S, Lahouari B, El Hidan MA, Souad B. Management of Leishmaniases in the Era of Climate Change in Morocco. Int J Environ Res Public Health. 2018;15. pmid:30037049
  14. 14. Moradi-Asl E, Rassi Y, Hanafi-Bojd AA, Vatandoost H, Saghafipour A, Adham D, et al. The Relationship between Climatic Factors and the Prevalence of Visceral Leishmaniasis in North West of Iran. Int J Pediatr. 2018;6.
  15. 15. Karakuş M, Çizmeci Z, Karabela ŞN, Erdoğan B, Güleç N. The impact of refugees on leishmaniasis in Turkey: a new Syrian/Turkish Leishmania tropica population structure described by multilocus microsatellite typing (MLMT). Parasitol Res. 2019;118: 2679–2687. pmid:31289943
  16. 16. Alawieh A, Musharrafieh U, Jaber A, Berry A, Ghosn N, Bizri AR. Revisiting leishmaniasis in the time of war: the Syrian conflict and the Lebanese outbreak. Int J Infect Dis IJID Off Publ Int Soc Infect Dis. 2014;29: 115–119. pmid:25449245
  17. 17. Ben Othman S, Ghawar W, Chaouch M, Ayari C, Chemkhi J, Cancino-Faure B, et al. First detection of Leishmania DNA in Psammomys obesus and Psammomys vexillaris: Their potential involvement in the epidemiology of leishmaniasis in Tunisia. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2018;59: 7–15. pmid:29413886
  18. 18. Souguir-Omrani H, Chemkhi J, Fathallah-Mili A, Saadi-BenAoun Y, BelHadjAli I, Guizani I, et al. Paraechinus aethiopicus (Ehrenberg 1832) and Atelerix algirus (Lereboullet 1842) hedgehogs: Possible reservoirs of endemic leishmaniases in Tunisia. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2018;63: 219–230. pmid:29860099
  19. 19. Rhajaoui M, Nasereddin A, Fellah H, Azmi K, Amarir F, Al-Jawabreh A, et al. New clinico-epidemiologic profile of cutaneous leishmaniasis, Morocco. Emerg Infect Dis. 2007;13: 1358–1360. pmid:18252108
  20. 20. Fathallah-Mili A, Saghrouni F, BenSaid Z, BenAoun YS-, Guizani I, BenSaid M. Retrospective Analysis of Leishmaniasis in Central Tunisia: An Update on Emerging Epidemiological Trends. Curr Top Trop Med. 2012; 27.
  21. 21. Garni R, Tran A, Guis H, Baldet T, Benallal K, Boubidi S, et al. Remote sensing, land cover changes, and vector-borne diseases: use of high spatial resolution satellite imagery to map the risk of occurrence of cutaneous leishmaniasis in Ghardaïa, Algeria. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2014;28: 725–734. pmid:25305006
  22. 22. Haddad N, Saliba H, Altawil A, Villinsky J, Al-Nahhas S. Cutaneous leishmaniasis in the central provinces of Hama and Edlib in Syria: Vector identification and parasite typing. Parasit Vectors. 2015;8: 524. pmid:26459055
  23. 23. Bettaieb J, Nouira M. Epidemiology of Cutaneous Leishmaniasis in Tunisia. Epidemiol Ecol Leishmaniasis. 2017 [cited 18 Nov 2019].
  24. 24. Alves Souza N, Souza Leite R, de Oliveira Silva S, Groenner Penna M, Figueiredo Felicori Vilela L, Melo MN, et al. Detection of mixed Leishmania infections in dogs from an endemic area in southeastern Brazil. Acta Trop. 2019;193: 12–17. pmid:30772331
  25. 25. Fraga J, Montalvo AM, llse Maes, Dujardin J-C, Van der Auwera G. HindII and SduI digests of heat-shock protein 70 PCR for Leishmania typing. Diagn Microbiol Infect Dis. 2013;77: 245–247. pmid:24050933
  26. 26. Fraga J, Montalvo AM, De Doncker S, Dujardin J-C, Van der Auwera G. Phylogeny of Leishmania species based on the heat-shock protein 70 gene. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2010;10: 238–245. pmid:19913110
  27. 27. Guerbouj S, Victoir K, Guizani I, Seridi N, Nuwayri-Salti N, Belkaid M, et al. Gp63 gene polymorphism and population structure of Leishmania donovani complex: influence of the host selection pressure? Parasitology. 2001;122: 25–35. pmid:11197761
  28. 28. Guerbouj S, Djilani F, Bettaieb J, Lambson B, Diouani MF, Ben Salah A, et al. Evaluation of a gp63-PCR based assay as a molecular diagnosis tool in canine leishmaniasis in Tunisia. PloS One. 2014;9: e105419. pmid:25153833
  29. 29. Mauricio IL, Gaunt MW, Stothard JR, Miles MA. Glycoprotein 63 (gp63) genes show gene conversion and reveal the evolution of Old World Leishmania. Int J Parasitol. 2007;37: 565–576. pmid:17280675
  30. 30. Ibrahim ME, Barker DC. The origin and evolution of the Leishmania donovani complex as inferred from a mitochondrial cytochrome oxidase II gene sequence. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2001;1: 61–68. pmid:12798051
  31. 31. Cao D-P, Guo X-G, Chen D-L, Chen J-P. Species delimitation and phylogenetic relationships of Chinese Leishmania isolates reexamined using kinetoplast cytochrome oxidase II gene sequences. Parasitol Res. 2011;109: 163–173. pmid:21221640
  32. 32. Hide M, Bras-Gonçalves R, Bañuls AL. Specific cpb copies within the Leishmania donovani complex: evolutionary interpretations and potential clinical implications in humans. Parasitology. 2007;134: 379–389. pmid:17129395
  33. 33. Chaouch M, Fathallah-Mili A, Driss M, Lahmadi R, Ayari C, Guizani I, et al. Identification of Tunisian Leishmania spp. by PCR amplification of cysteine proteinase B (cpb) genes and phylogenetic analysis. Acta Trop. 2013;125: 357–365. pmid:23228525
  34. 34. Khosravi S, Hejazi SH, Hashemzadeh M, Eslami G, Darani HY. Molecular diagnosis of Old World leishmaniasis: real-time PCR based on tryparedoxin peroxidase gene for the detection and identification of Leishmania spp. J Vector Borne Dis. 2012;49: 15–18. pmid:22585237
  35. 35. Cupolillo E, Grimaldi Júnior G, Momen H, Beverley SM. Intergenic region typing (IRT): a rapid molecular approach to the characterization and evolution of Leishmania. Mol Biochem Parasitol. 1995;73: 145–155. pmid:8577322
  36. 36. Dávila AM, Momen H. Internal-transcribed-spacer (ITS) sequences used to explore phylogenetic relationships within Leishmania. Ann Trop Med Parasitol. 2000;94: 651–654. pmid:11064767
  37. 37. Kuhls K, Mauricio IL, Pratlong F, Presber W, Schönian G. Analysis of ribosomal DNA internal transcribed spacer sequences of the Leishmania donovani complex. Microbes Infect. 2005;7: 1224–1234. pmid:16002315
  38. 38. Piarroux R, Fontes M, Perasso R, Gambarelli F, Joblet C, Dumon H, et al. Phylogenetic relationships between Old World Leishmania strains revealed by analysis of a repetitive DNA sequence. Mol Biochem Parasitol. 1995;73: 249–252. pmid:8577334
  39. 39. Van der Auwera G, Dujardin J-C. Species Typing in Dermal Leishmaniasis. Clin Microbiol Rev. 2015;28: 265–294. pmid:25672782
  40. 40. Bulle B, Millon L, Bart J-M, Gállego M, Gambarelli F, Portús M, et al. Practical approach for typing strains of Leishmania infantum by microsatellite analysis. J Clin Microbiol. 2002;40: 3391–3397. pmid:12202583
  41. 41. Al-Jawabreh A, Diezmann S, Müller M, Wirth T, Schnur LF, Strelkova MV, et al. Identification of geographically distributed sub-populations of Leishmania (Leishmania) major by microsatellite analysis. BMC Evol Biol. 2008;8: 183. pmid:18577226
  42. 42. Alam MZ, Kuhls K, Schweynoch C, Sundar S, Rijal S, Shamsuzzaman AKM, et al. Multilocus microsatellite typing (MLMT) reveals genetic homogeneity of Leishmania donovani strains in the Indian subcontinent. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2009;9: 24–31. pmid:18957333
  43. 43. Fernández-Arévalo A, El Baidouri F, Ravel C, Ballart C, Abras A, Lachaud L, et al. The Leishmania donovani species complex: A new insight into taxonomy☆. Int J Parasitol. 2020;50: 1079–1088. pmid:32889062
  44. 44. Cotton JA, Durrant C, Franssen SU, Gelanew T, Hailu A, Mateus D, et al. Genomic analysis of natural intra-specific hybrids among Ethiopian isolates of Leishmania donovani. PLoS Negl Trop Dis. 2020;14: e0007143. pmid:32310945
  45. 45. Bañuls A-L, Hide M, Prugnolle F. Leishmania and the Leishmaniases: A Parasite Genetic Update and Advances in Taxonomy, Epidemiology and Pathogenicity in Humans. In: Baker JR, Muller R, Rollinson D, editors. Advances in Parasitology. Academic Press; 2007. pp. 1–458. https://doi.org/10.1016/S0065-308X(06)64001-3
  46. 46. Patwardhan A, Ray S, Roy A. Molecular Markers in Phylogenetic Studies-A Review. J Phylogenetics Evol Biol. 2014;02.
  47. 47. Maurício IL. Leishmania Taxonomy. In: Bruschi F, Gradoni L, editors. The Leishmaniases: Old Neglected Tropical Diseases. Cham: Springer International Publishing; 2018. pp. 15–30. https://doi.org/10.1007/978-3-319-72386-0_2
  48. 48. Baral PK, Jajčanin-Jozić N, Deller S, Macheroux P, Abramić M, Gruber K. The First Structure of Dipeptidyl-peptidase III Provides Insight into the Catalytic Mechanism and Mode of Substrate Binding. J Biol Chem. 2008;283: 22316–22324. pmid:18550518
  49. 49. Prajapati SC, Chauhan SS. Dipeptidyl peptidase III: a multifaceted oligopeptide N-end cutter: Dipeptidyl peptidase III. FEBS J. 2011;278: 3256–3276. pmid:21794094
  50. 50. Diaz JR, Ramírez CA, Nocua PA, Guzman F, Requena JM, Puerta CJ. Dipeptidyl peptidase 3, a novel protease from Leishmania braziliensis. Oberer M, editor. PLOS ONE. 2018;13: e0190618. pmid:29304092
  51. 51. Silverman JM, Chan SK, Robinson DP, Dwyer DM, Nandan D, Foster LJ, et al. Proteomic analysis of the secretome of Leishmania donovani. Genome Biol. 2008;9: R35. pmid:18282296
  52. 52. Besteiro S, Williams RAM, Coombs GH, Mottram JC. Protein turnover and differentiation in Leishmania. Int J Parasitol. 2007;37: 1063–1075. pmid:17493624
  53. 53. Kbaier-Hachemi H, Barhoumi M, Chakroun AS, Fadhel MB, Guizani EI. Differenciation des especes responsables de leishmaniose cutanée par une amplification PCR du gene codant pour la dipeptidyl peptidase III. Archs Inst Pasteur Tunis. 2008;85: 45–54.
  54. 54. Guizani I, Eys GJJMV, Ismail RB, Dellagi K. Use of Recombinant DNA Probes for Species Identification of Old World Leishmania Isolates. Am J Trop Med Hyg. 1994;50: 632–640. pmid:8203714
  55. 55. Schönian G, Nasereddin A, Dinse N, Schweynoch C, Schallig HDFH, Presber W, et al. PCR diagnosis and characterization of Leishmania in local and imported clinical samples. Diagn Microbiol Infect Dis. 2003;47: 349–358. pmid:12967749
  56. 56. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol Biol Evol. 2018;35: 1547–1549. pmid:29722887
  57. 57. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol. 2013;30: 2725–2729. pmid:24132122
  58. 58. Yang Z. Maximum likelihood phylogenetic estimation from DNA sequences with variable rates over sites: approximate methods. J Mol Evol. 1994;39: 306–314. pmid:7932792
  59. 59. Felsenstein J. Confidence limits on phylogenies: an approach using the bootstrap. Evol Int J Org Evol. 1985;39: 783–791. pmid:28561359
  60. 60. Sokal R, Michener C. A statistical method for evaluating systematic relationships. Sci Bull. 1958;38: 1409–1438.
  61. 61. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4: 406–425. pmid:3447015
  62. 62. Rzhetsky A, Nei M. Statistical properties of the ordinary least-squares, generalized least-squares, and minimum-evolution methods of phylogenetic inference. J Mol Evol. 1992;35: 367–375. pmid:1404422
  63. 63. Felsenstein J. Evolutionary trees from DNA sequences: a maximum likelihood approach. J Mol Evol. 1981;17: 368–376. pmid:7288891
  64. 64. Nei M, Gojobori T. Simple methods for estimating the numbers of synonymous and nonsynonymous nucleotide substitutions. Mol Biol Evol. 1986;3: 418–426. pmid:3444411
  65. 65. Tupperwar N, Vineeth V, Rath S, Vaidya T. Development of a real-time polymerase chain reaction assay for the quantification of Leishmania species and the monitoring of systemic distribution of the pathogen. Diagn Microbiol Infect Dis. 2008;61: 23–30. pmid:18255247
  66. 66. Herrador Z, Gherasim A, Jimenez BC, Granados M, San Martín JV, Aparicio P. Epidemiological Changes in Leishmaniasis in Spain According to Hospitalization-Based Records, 1997–2011: Raising Awareness towards Leishmaniasis in Non-HIV Patients. Carvalho EM, editor. PLoS Negl Trop Dis. 2015;9: e0003594. pmid:25756785
  67. 67. da Silva LA, de Sousa CDS, da Graça GC, Porrozzi R, Cupolillo E. Sequence analysis and PCR-RFLP profiling of the hsp70 gene as a valuable tool for identifying Leishmania species associated with human leishmaniasis in Brazil. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2010;10: 77–83. pmid:19913112
  68. 68. Requena JM, Chicharro C, García L, Parrado R, Puerta CJ, Cañavate C. Sequence analysis of the 3’-untranslated region of HSP70 (type I) genes in the genus Leishmania: its usefulness as a molecular marker for species identification. Parasit Vectors. 2012;5: 87. pmid:22541251
  69. 69. Zhang C-Y, Lu X-J, Du X-Q, Jian J, Shu L, Ma Y. Phylogenetic and evolutionary analysis of Chinese Leishmania isolates based on multilocus sequence typing. PloS One. 2013;8: e63124. pmid:23646184
  70. 70. Marlow MA, Boité MC, Ferreira GEM, Steindel M, Cupolillo E. Multilocus sequence analysis for Leishmania braziliensis outbreak investigation. PLoS Negl Trop Dis. 2014;8: e2695. pmid:24551258
  71. 71. Marco JD, Barroso PA, Locatelli FM, Cajal SP, Hoyos CL, Nevot MC, et al. Multilocus sequence typing approach for a broader range of species of Leishmania genus: describing parasite diversity in Argentina. Infect Genet Evol J Mol Epidemiol Evol Genet Infect Dis. 2015;30: 308–317. pmid:25558029
  72. 72. Herrera G, Hernández C, Ayala MS, Flórez C, Teherán AA, Ramírez JD. Evaluation of a Multilocus Sequence Typing (MLST) scheme for Leishmania (Viannia) braziliensis and Leishmania (Viannia) panamensis in Colombia. Parasit Vectors. 2017;10: 236. pmid:28499458
  73. 73. Teixeira DG, Monteiro GRG, Martins DRA, Fernandes MZ, Macedo-Silva V, Ansaldi M, et al. Comparative analyses of whole genome sequences of Leishmania infantum isolates from humans and dogs in northeastern Brazil. Int J Parasitol. 2017;47: 655–665. pmid:28606698
  74. 74. Franssen SU, Durrant C, Stark O, Moser B, Downing T, Imamura H, et al. Global genome diversity of the Leishmania donovani complex. Genomics; 2019.
  75. 75. Van der Auwera G, Ravel C, Verweij JJ, Bart A, Schonian G, Felger I. Evaluation of Four Single-Locus Markers for Leishmania Species Discrimination by Sequencing. J Clin Microbiol. 2014;52: 1098–1104. pmid:24452158
  76. 76. Ghosh S, Banerjee P, Sarkar A, Datta S, Chatterjee M. Coinfection of Leptomonas seymouri and Leishmania donovani in Indian Leishmaniasis. J Clin Microbiol. 2012;50: 2774–2778. pmid:22622439
  77. 77. Kraeva N, Butenko A, Hlaváčová J, Kostygov A, Myškova J, Grybchuk D, et al. Leptomonas seymouri: Adaptations to the Dixenous Life Cycle Analyzed by Genome Sequencing, Transcriptome Profiling and Co-infection with Leishmania donovani. PLoS Pathog. 2015;11: e1005127. pmid:26317207
  78. 78. Thakur L, Kushwaha HR, Negi A, Jain A, Jain M. Leptomonas seymouri Co-infection in Cutaneous Leishmaniasis Cases Caused by Leishmania donovani From Himachal Pradesh, India. Front Cell Infect Microbiol. 2020;10: 345. pmid:32760679
  79. 79. Akhoundi M, Kuhls K, Cannet A, Votýpka J, Marty P, Delaunay P, et al. A Historical Overview of the Classification, Evolution, and Dispersion of Leishmania Parasites and Sandflies. PLoS Negl Trop Dis. 2016;10: e0004349. pmid:26937644
  80. 80. El Baidouri F, Diancourt L, Berry V, Chevenet F, Pratlong F, Marty P, et al. Genetic Structure and Evolution of the Leishmania Genus in Africa and Eurasia: What Does MLSA Tell Us. Schönian G, editor. PLoS Negl Trop Dis. 2013;7: e2255. pmid:23785530
  81. 81. Luo H, Gao F, Lin Y. Evolutionary conservation analysis between the essential and nonessential genes in bacterial genomes. Sci Rep. 2015;5: 13210. pmid:26272053
  82. 82. Peters W, Elbihari S, Evans DA. Leishmania infecting man and wild animals in Saudi Arabia. 2. Leishmania arabica n. sp. Trans R Soc Trop Med Hyg. 1986;80: 497–502. pmid:3544354
  83. 83. Akhavan AA, Yaghoobi-Ershadi MR, Khamesipour A, Mirhendi H, Alimohammadian MH, Rassi Y, et al. Dynamics of Leishmania infection rates in Rhombomys opimus (Rodentia: Gerbillinae) population of an endemic focus of zoonotic cutaneous leishmaniasis in Iran. Bull Soc Pathol Exot 1990. 2010;103: 84–89. pmid:20390397
  84. 84. Akhoundi M, Mohebali M, Asadi M, Mahmodi MR, Amraei K, Mirzaei A. Molecular characterization of Leishmania spp. in reservoir hosts in endemic foci of zoonotic cutaneous leishmaniasis in Iran. Folia Parasitol (Praha). 2013;60: 218–224. pmid:23951928
  85. 85. Darvishi M, Yaghoobi-Ershadi MR, Shahbazi F, Akhavan AA, Jafari R, Soleimani H, et al. Epidemiological study on sand flies in an endemic focus of cutaneous leishmaniasis, bushehr city, southwestern iran. Front Public Health. 2015;3: 14. pmid:25699245
  86. 86. Rafizadeh S, Saraei M, Abaei MR, Oshaghi MA, Mohebali M, Peymani A, et al. Molecular Detection of Leishmania major and L. turanica in Phlebotomus papatasi and First Natural Infection of P. salehi to L. major in North-East of Iran. J Arthropod-Borne Dis. 2016;10: 141–147. pmid:27308272
  87. 87. Rioux JA, Lanotte G, Serres E, Pratlong F, Bastien P, Perieres J. Taxonomy of Leishmania. Use of isoenzymes. Suggestions for a new classification. Ann Parasitol Hum Comp. 1990;65: 111–125. pmid:2080829
  88. 88. Lukes J, Mauricio IL, Schonian G, Dujardin J-C, Soteriadou K, Dedet J-P, et al. Evolutionary and geographical history of the Leishmania donovani complex with a revision of current taxonomy. Proc Natl Acad Sci. 2007;104: 9375–9380. pmid:17517634
  89. 89. Croan DG, Morrison DA, Ellis JT. Evolution of the genus Leishmania revealed by comparison of DNA and RNA polymerase gene sequences. Mol Biochem Parasitol. 1997;89: 149–159. pmid:9364962
  90. 90. Noyes HA, Arana BA, Chance ML, Maingon R. The Leishmania hertigi (Kinetoplastida; Trypanosomatidae) complex and the lizard Leishmania: their classification and evidence for a neotropical origin of the Leishmania-Endotrypanum clade. J Eukaryot Microbiol. 1997;44: 511–517. pmid:9304821
  91. 91. Schwenkenbecher JM, Wirth T, Schnur LF, Jaffe CL, Schallig H, Al-Jawabreh A, et al. Microsatellite analysis reveals genetic structure of Leishmania tropica. Int J Parasitol. 2006;36: 237–246. pmid:16307745
  92. 92. Hamouchi AE, Ajaoud M, Arroub H, Charrel R, Lemrani M. Genetic diversity of Leishmania tropica in Morocco: does the dominance of one haplotype signify its fitness in both predominantly anthropophilic Phlebotomus sergenti and human beings? Transbound Emerg Dis. 2019;66: 373–380. pmid:30281944
  93. 93. Klatt S, Simpson L, Maslov DA, Konthur Z. Leishmania tarentolae: Taxonomic classification and its application as a promising biotechnological expression host. Guizani I, editor. PLoS Negl Trop Dis. 2019;13: e0007424. pmid:31344033
  94. 94. Novo S, Leles D, Bianucci R, Araujo A. Leishmania tarentolae molecular signatures in a 300 hundred-years-old human Brazilian mummy. Parasit Vectors. 2015;8: 72. pmid:25649153
  95. 95. Luyo-Acero GE, Uezato H, Oshiro M, Takei K, Kariya K, Katakura K, et al. Sequence variation of the cytochrome b gene of various human infecting members of the genus Leishmania and their phylogeny. Parasitology. 2004;128: 483–491. pmid:15180316
  96. 96. Kuru T, Janusz N, Gadisa E, Gedamu L, Aseffa A. Leishmania aethiopica: Development of specific and sensitive PCR diagnostic test. Exp Parasitol. 2011;128: 391–395. pmid:21616071
  97. 97. Kebede N, Oghumu S, Worku A, Hailu A, Varikuti S, Satoskar AR. Multilocus microsatellite signature and identification of specific molecular markers for Leishmania aethiopica. Parasit Vectors. 2013;6: 160. pmid:23734874
  98. 98. Montalvo AM, Fraga J, Maes I, Dujardin J-C, Van der Auwera G. Three new sensitive and specific heat-shock protein 70 PCRs for global Leishmania species identification. Eur J Clin Microbiol Infect Dis Off Publ Eur Soc Clin Microbiol. 2012;31: 1453–1461. pmid:22083340
  99. 99. María Requena J, Soto M, Quijada L, Alonso C. Genes and Chromosomes of Leishmania infantum. Mem Inst Oswaldo Cruz. 1997;92: 853–858. pmid:9566218
  100. 100. Parvizi P, Ready PD. Nested PCRs and sequencing of nuclear ITS-rDNA fragments detect three Leishmania species of gerbils in sandflies from Iranian foci of zoonotic cutaneous leishmaniasis. Trop Med Int Health TM IH. 2008;13: 1159–1171. pmid:18631311
  101. 101. Mirzaei A, Rouhani S, Kazerooni P, Farahmand M, Parvizi P. Molecular detection and conventional identification of leishmania species in reservoir hosts of zoonotic cutaneous leishmaniasis in fars province, South of iran. Iran J Parasitol. 2013;8: 280–288. pmid:23914242
  102. 102. Hijawi KJF, Hijjawi NS, Ibbini JH. Detection, genotyping, and phylogenetic analysis of Leishmania isolates collected from infected Jordanian residents and Syrian refugees who suffered from cutaneous leishmaniasis. Parasitol Res. 2019;118: 793–805. pmid:30729301