Skip to main content

Genetic differentiation of Plasmodium vivax duffy binding protein in Ethiopia and comparison with other geographical isolates

Abstract

Background

Plasmodium vivax Duffy binding protein (PvDBP) is a merozoite surface protein located in the micronemes of P. vivax. The invasion of human reticulocytes by P. vivax merozoites depends on the parasite DBP binding domain engaging Duffy Antigen Receptor for Chemokine (DARC) on these red blood cells (RBCs). PvDBPII shows high genetic diversity which is a major challenge to its use in the development of a vaccine against vivax malaria.

Methods

A cross-sectional study was conducted from February 2021 to September 2022 in five study sites across Ethiopia. A total of 58 blood samples confirmed positive for P. vivax by polymerase chain reaction (PCR) were included in the study to determine PvDBPII genetic diversity. PvDBPII were amplified using primers designed from reference sequence of P. vivax Sal I strain. Assembling of sequences was done using Geneious Prime version 2023.2.1. Alignment and phylogenetic tree constructions using MEGA version 10.1.1. Nucleotide diversity and haplotype diversity were analysed using DnaSP version 6.12.03, and haplotype network was generated with PopART version 1.7.

Results

The mean age of the participants was 25 years, 5 (8.6%) participants were Duffy negatives. From the 58 PvDBPII sequences, seven haplotypes based on nucleotide differences at 8 positions were identified. Nucleotide diversity and haplotype diversity were 0.00267 ± 0.00023 and 0.731 ± 0.036, respectively. Among the five study sites, the highest numbers of haplotypes were identified in Arbaminch with six different haplotypes while only two haplotypes were identified in Gambella. The phylogenetic tree based on PvDBPII revealed that parasites of different study sites shared similar genetic clusters with few exceptions. Globally, a total of 39 haplotypes were identified from 223 PvDBPII sequences representing different geographical isolates obtained from NCBI archive. The nucleotide and haplotype diversity were 0.00373 and 0.845 ± 0.015, respectively. The haplotype prevalence ranged from 0.45% to 27.3%. Two haplotypes were shared among isolates from all geographical areas of the globe.

Conclusions

PvDBPII of the Ethiopian P. vivax isolates showed low nucleotide but high haplotype diversity, this pattern of genetic variability suggests that the population may have undergone a recent expansion. Among the Ethiopian P. vivax isolates, almost half of the sequences were identical to the Sal-I reference sequence. However, there were unique haplotypes observed in the Ethiopian isolates, which does not share with isolates from other geographical areas. There were two haplotypes that were common among populations across the globe. Categorizing population haplotype frequency can help to determine common haplotypes for designing an effective blood-stage vaccine which will have a significant role for the control and elimination of P. vivax.

Background

In 2022, there were an estimated of 6.9 million cases of Plasmodium vivax globally [1]. Plasmodium vivax malaria is recognized as a cause of severe morbidity and mortality, with a significant negative impact on health in endemic countries [2,3,4]. The unique biological features of P. vivax, such as hypnozoite formation and early gametocytogenesis, pose a challenge to global malaria elimination efforts as traditional control measures may not be as effective. A P. vivax vaccine would provide a powerful tool for malaria control [5].

Despite its African origins, the presence of P. vivax on the continent is unevenly distributed, with East Africa, particularly the horn of Africa experiencing endemicity and significant clinical disease [6]. In Ethiopia, the impact of P. vivax on public health is substantial, given the country’s elevated malaria burden. P. vivax constitutes a significant portion of malaria cases, and its transmission occurs seasonally in highland areas where the Anopheles mosquito, the primary vector, thrives [7]. This distribution pattern is likely linked to the prevalence of Duffy positive individuals, as P. vivax is believed to rely on the Duffy receptor for invasion of reticulocytes and the manifestation of clinical disease. These findings underscore the significance of studying P. vivax in Ethiopia, emphasizing its prevalence and the need for further research in this context.

The invasion pathway of P. vivax into reticulocytes was previously believed to primarily depend on the interaction between the Duffy binding protein (PvDBP) and its receptor, the Duffy Antigen Receptor for Chemokines (DARC) [8,9,10]. P. vivax Duffy binding protein (PvDBP) is a merozoite surface protein of 140 kDa located in the micronemes of P. vivax, and is used to invade RBCs [11]. The erythrocyte-binding motif of DBP resides in the N-terminal conserved cysteine-rich domain (C1–C12) known as DBP region II [12,13,14]. Region II of PvDBP1 (PvDBPII) is 330 amino acids in length and contains 12 cysteines that are conserved among different DBL domains. Since it has conserved cysteine-rich domain, PvDBPII is one of the most promising vaccine targets [15, 16]. The central stretch of P. vivax region II (cysteines at positions 4–7) is important for receptor recognition and critical for binding to DARCs on human RBCs [15, 17].

Genetic variation analysis indicated that PvDBPII shows high genetic diversity [18] and has a 4 times higher substitution rate compared to the rest of the gene [19, 20]. The region involved in RBC binding is highly polymorphic and there is evidence of strain specific protection [21]. The current vaccine development is based on the DBPII haplotype of the P. vivax laboratory-adapted strain Sal-1 [11]. So identifying the genetic diversity in PvDBPII and its dominant haplotypes is important for the rational design of DBP-based vaccine against vivax malaria. This study aims to assess the genetic diversity of PvDBPII across various study sites in Ethiopia and comparison with isolates from other regions of the world.

Methods

Study sites and sample collection

A health-facility based cross-sectional study was conducted from February 2021 to September 2022 at five hospitals: Arbaminch, Dubti, Gambella, Metehara, and Shewarobit hospitals (Fig. 1). Finger prick blood sample was collected from each malaria suspected febrile individuals for the purpose of doing malaria laboratory diagnosis using microscopy and Rapid Diagnostic Tests (RDTs). In order to maximize the accuracy of research participant inclusion, patients who were diagnosed positive for P. vivax by both RDT and microscopy were included in the study. Parasite densities were determined on thick blood film by malaria microscopists by considering a minimum of 200 WBCs, the reported number of parasites per microliter of blood, and an assumed total white blood cell count of 8000/μL. Venous blood samples were collected and spotted onto Whatman 903 Protein Saver Card, US (Dried Blood Spot, DBS). Genomic DNA (parasite and human DNA) was extracted from a 6 mm diameter spot of a DBS sample and eluted in 100 µl of TE Buffer (Tris–EDTA) using the QIAamp DNA Extraction Kit according to the manufacturer’s instructions. To avoid contamination, the scissor blade (puncher) and forceps were immersed in the RNASE AWAY solution for 1 min and then cleaned with distilled water in between each sample. In addition to the microscopy and RDT, malaria parasites identification was confirmed by Bio-Rad CFX96 Real-Time PCR [22].

Fig. 1
figure 1

A map of the study sites, Ethiopia

Amplifications and sequencing of PvDBPII

P. vivax samples collected at five hospitals were evaluated for PvDBPII polymorphism. The amplification of PvDBP sequences utilized primers targeting the DBP1 region II, (forward: 5′-GATATTGATCATAAGAAAACGATCTCTAGT- 3′; reverse: 5′-TGTCACAACTTCCTGAGTATTTTTTTTAGCCTC- 3′) [23], designed based on the reference sequence P. vivax PVX_110810 from of Sal I (Reference assembly: NC_009911.1). The amplification was carried out in a 20-μL reaction mixture containing 2 μL of genomic DNA, 10 μL of 2 × DreamTaq Green PCR Master Mix (Fermentas), and 0.5 μM primer. Amplification reactions were performed in a Bio-Rad T100 thermal cycler, with an initial denaturation at 94 °C for 2 min, followed by 35 cycles at 94 °C for 30 s, 58 °C for 30 s, and 72 °C for 90 s with a final 5-min extension at 72 °C. PCR products were purified prior to Sanger sequencing using Sephadex G-50 (Merck, Darmstadt, Germany) and purified amplicons were then used as templates for cycle sequencing reaction, using the Applied Biosystems BigDye Terminator v3.1 Cycle Sequencing Kit, according to manufacturers’ protocols. Sequences were detected on the SeqStudio genetic analyzer (Applied Biosystems, California, USA) and the data were generated as Fasta format for analysis. In order to ensure quality, positive and negative controls were used in each PCR run, and gel electrophoresis was performed to visualize the amplified product.

Genetic diversity and statistical analysis

Sequences of PvDBPII genes were trimmed and assembled using De Novo assembler of Geneious Prime version 2023.2.1 [24]. Pairwise and multiple aliment was done using ClustalW alignment algorithm in Diversity Molecular evolutionary Genetic Analysis (MEGA X). Missing values across the column of the total sequences were removed using Jalview version 2.11.2 (a multiple sequence alignment editor and analysis workbench) [25]. All ambiguous positions were removed for each sequence pair (pairwise deletion option). Nucleotide substitution (r) pattern and rates were estimated under the Tamura-Nei model, and phylogenetic tree was built to see evolutionary divergence between Sequences using MEGA version 10.1.1 software [26]. Nucleotide diversity (π), number of segregating sites (S), the number of haplotypes (H), haplotype diversity (Hd), the total number of mutations (η), and the average number of nucleotide differences (k) were analysed using DnaSP version 6.12.03 software [27]. Frequencies were analysed using Statistical Package for Social Sciences (SPSS, version 25). P-values less than 0.05 were considered as statistically significant.

Genetic differentiation and haplotype network analysis

Gene flow and genetic differentiation estimates (Fst) were analysed using DnaSP version 6.12.03 software [27] and haplotypes network was constructed using PopART version 1.7 software [28]. The level of genetic differentiation of PvDBPII was estimated by FST values. These values were interpreted based on established criteria: 0–0.05 indicating minimal genetic differentiation, 0.05–0.15 denoting moderate differentiation, 0.15–0.25 representing substantial differentiation, and values exceeding 0.25 signifying significant genetic divergence [29]. To estimate the degree of genetic differentiation of the PvDBPII in global isolates, a total of 165 global PvDBPII sequences from 5 P. vivax endemic countries of different continents were retrieved from the National Centre for Biotechnology Information (NCBI) archive. These included; isolates from South America (n = 34): Brazil (EU812840.1-EU812873.1) [30], isolates from Asia (n = 9): Iran (EU860429-EU860437) [31], isolates from North America (n = 35): Mexico (KP759780-KP759814) [32], isolates from Oceania (n = 45): Papua New Guinea (PNG) (AF469550:AF469594) [33], and 42 isolates from Africa: Sudan (MG805616.1-MG805657.1) [13]. The reference sequence from Sal-1 (NC_009911.1)) was included in the final analysis.

Results

Characteristics of the study participants

A total of 80 samples with 16 samples per study site were sequenced to assess genetic diversity of PvDBP. Of these, 58 isolates which were effectively sequenced and had good quality were considered for analysis. The mean parasite density of the asexual stage was 9735 parasite/µl. Out of the 58 participants, 38 (65.5%) were male, the mean age was 25 years, and the study participants aged between 20 and 40 years [18] accounted for 70.7% (41)(Table 1). The frequency of male participants from ArbaMinch, Shewarobit, Metehara, Gambella and Dubti was 9/11 (81.8%), 4/7 (57.1%), 9/14 (64.3%), 6/12 (50.0%), and 10/14 (71.4%), respectively.

Table 1 Demographic information of study participants

Nucleotides and amino acid frequencies of PvDBPII gene sequences

Analysis was done using the consensus sequence of 736 bps which corresponds to nucleotide positions from 838 to 1573 of the Salvador I reference. The nucleotide frequencies of the 58 sequences were 40.36% (A), 26.06% (T), 11.66% (C), and 21.92% (G). The mean percent of GC in assembled sequence was 33.5%. The amino acid with highest frequency was leucine (L) (9.11) followed by alanine (A) 97.69%, whereas the frequency of tryptophan (W) (1.43%) was the least (Table 2).

Table 2 The amino acid frequencies among the 58 sequences

Single nucleotide polymorphism (SNP) of PvDBPII among the Ethiopian isolates

The number of segregating sites were 8, with a nucleotide diversity of 0.00267 ± 0.00023 and an average of 1.791 pairwise nucleotide differences (k). The nucleotide diversity per study site was 0.00295, 0.00177, 0.00216, 0.00363, and 0.0022 among the isolates from Metehara, Gambella, Dubti, ArbaMinch, and ShewaRobit, respectively (Table 3). A total of 8 mutations were observed, consisting of 4 singleton mutations (present in only one sequence) at positions 12, 43, 149, and 164, and 4 parsimony informative sites (with at least two types of nucleotides or occurring with a minimum frequency of two) at positions 274, 316, 414, and 433. From the 58 PvDBPII sequences, 7 haplotypes were identified (GenBank accession numbers: PP141174-PP141180), yielding a haplotype diversity of 0.731 ± 0.036. Among the 58 P. vivax isolates from Ethiopia, 41.38% (24 sequences) matched the Sal-1 reference sequence, with the majority originating from Gambella (37.5% of the 24). All identified haplotypes exhibited dimorphism, except for the predominant haplotype 1, which accounted for 41.38% (24 out of 58) of the Ethiopian isolates (Fig. 2). Of the total sequences, the Duffy status of five isolates were negative, and of these most of them (80% of the 5) were identified as haplotype 1 and the other one was haplotype 4 (see Additional file 1).

Table 3 Nucleotide and haplotype diversity of PvDBPII gene and number of isolates per haplotypes among different Ethiopian populations
Fig. 2
figure 2

Single nucleotides polymorphism of PvDBPII among Ethiopian isolates. A Position of nucleotide changes. Polymorphic nucleotides are listed for each haplotype. Nucleotides identical to those of the reference sequence, Sal-I (NC_009911.1), are marked in green. The dimorphic nucleotide changes are marked in red color. Y-axis represent haplotypes list in the left while the total number of sequences for each haplotype at right panel (D: + ve; Duffy positive, and D:-ve; Duffy negative). X-axis represent nucleotides position. B Frequencies (%) of nucleotide changes found in PvDBPII among 58 Ethiopian isolates. Frequencies in percent denoted on the Y-axis, and nucleotide changes on the X-axis

Diversity and evolutionary history among isolates from different study sites

Within the five study sites, Arbaminch exhibited the highest diversity, housing 6 distinct haplotypes, while Gambella had only 2. Haplotype 2 was exclusive to Metehara, while Arbaminch contained haplotypes 5, 6, and 7. Notably, three haplotypes (5.17%) were represented by a single sequence. The nucleotide diversity of PvDBPII in Arbaminch (π = 0.00363) surpassed that of the other sites (Table 3).

The Neighbour-Joining method was employed to infer the evolutionary history, with evolutionary distances calculated using the Tajima-Nei method, measured in the number of base substitutions per site. The phylogenetic tree, based on PvDBPII SNPs, indicated a shared profile among parasites from different study sites, with a few exceptions. Notably, the majority of isolates from Gambella (9 out of 12, or 75%) exhibited a close relationship with the Sal-1 reference strain, while isolates from Arbaminch showed the least relationship (3 out of 11, or 27%) (Fig. 3).

Fig. 3
figure 3

Neighbour-Joining phylogenetic tree based on PvDBP-II-gene SNPs of Arbaminch (n = 11), Gambella (n = 12), Metehara (n = 14), Dubti (n = 14) and Shewarobit (n = 7) isolates

Genetic differentiation and haplotypes network among global isolates

A total of 39 haplotypes were identified from 223 PvDBPII gene sequences, consisting of 58 sequences from Ethiopia and 165 obtained from the NCBI archive (global isolates). The average number of nucleotide differences (k) was 2.411, with 40 segregating sites, a nucleotide diversity of 0.00373, and a haplotype diversity of 0.845 ± 0.015. The genetic differentiation between Ethiopian isolates and those from Iran was minimal, as indicated by an FST value of 0.00. However, there was moderate differentiation between Ethiopian isolates and those from Sudan, with an FST value of 0.0697. Ethiopian isolates were most differentiated from PNG isolates (0.1846). Notably, isolates from Papua New Guinea exhibited significant genetic differentiation from those in Sudan and Mexico, with FST values of 0.3051 and 0.3336, respectively (Table 4).

Table 4 Nucleotide and haplotype diversity of PvDBPII gene and Fst of global populations

In the network analysis, the haplotype prevalence ranged from 0.45% to 27.3%. Both haplotypes H1 (Sal-1 like) and H4 were present across isolates from all geographical areas of the globe. Haplotypes H5 and H9 were shared among all geographic locations except those from Iran and Brazil, respectively. The most prevalent haplotype H1 (27.3%) was shared with multiple populations followed by H4 (23.8%). Twenty-six haplotypes (66.67%) were represented by a single sequence. Three haplotypes (H10, H11, and H12) were exclusive to Ethiopian isolates, not shared with those from other geographical areas. The network comprised a total of 48 connections among haplotypes, with 36 (75%) differing by one nucleotide, and 12 (25%) varying by two nucleotides (Fig. 4).

Fig. 4
figure 4

Networking of PvDBPII obtaining from Ethiopian P. vivax isolates and isolates from different malaria endemic countries; Brazil, Iran, Mexico, Papua New Guinea, and Sudan

Discussion

The analysis of PvDBPII gene sequences from diverse geographical regions sheds light on several noteworthy patterns of parasite evolution and implications, such as evaluation of malaria vaccine candidates [34]. The range of observed haplotype prevalence, spanning from 0.45% to 27.3%, serves as an indicator of the inherent genetic diversity within P. vivax populations. This observation is particularly relevant considering that the ongoing vaccine development is solely grounded in the DBPII haplotype of the laboratory-adapted P. vivax strain Sal-1 [11]. This study was the first to report polymorphism on the isolates from five different malaria endemic sites in the country. PvDBPII polymorphisms were investigated in 58 P. vivax isolates from endemic study sites to expand the knowledge on PvDBPII gene population diversity in Ethiopia and compare the data with those from other endemic countries in the world. In general, this study identified 7 haplotypes from Ethiopian isolates and 39 haplotypes from the global isolates.

Parasite genetic diversity has been considered a potential indicator of transmission intensity [35, 36], particularly in areas nearing malaria elimination where routine surveillance and case reporting may be less reliable [35]. In examining the molecular polymorphism within the 58 PvDBPII sequences, revealed a relatively lower nucleotide diversity (π = 0.00267 ± 0.00023) compared to documented studies in various regions. Notably, the nucleotide diversity observed in this study was lower than findings in Mexico (π = 0.00304) [32], Sudan (π = 0.00497) [13], China (π = 0.0059) [37], Ethiopia (π = 0.00608) [38], Ecuador (0.00736) [39], Brazil (π = 0.0081) [30], and Thailand (π = 0.012) [35]. These variations underscore the dynamics of malaria transmission intensity and emphasize the importance of region-specific assessments in understanding the genetic landscape of P. vivax [35]. The overall haplotype diversity observed in Ethiopia was 0.731, surpassing the reported diversity in Mexico (0.556) [32], and aligning closely with isolates from the Amazon region, Brazil (0.715) [40]. However, it was relatively lower than the haplotype diversity documented in Thailand (0.849) [35], Eastern Myanmar (0.850) [41], Ethiopia (0.874) [38], Sudan (0.900), China (0.91) [37], Sri Lanka (0.922) [34], Brazil (0.934) [30], Ecuador (0.940) [39], and Brazil (0.98) [42]. These comparative diversity measures underscore the unique genetic landscape of P. vivax populations and highlight the importance of regional context in understanding haplotype diversity [37].

In this study, seven haplotypes were identified from a pool of 58 PvDBPII sequences. This represents a relatively low number of haplotypes compared to a prior study in Ethiopia where nine distinct haplotypes were discerned from 23 isolates [38]. Although the phylogenetic tree based on PvDBPII SNPs indicated shared profiles among parasites from different study sites, with few exceptions, the haplotype diversity varied across sites, ranging from 0.8545 in Arbaminch to 0.4091 in Gambella. This may be due to the difference in transmission intensity of P. vivax at different part of the country [35]. Among the 58 Ethiopian sequences, the highest number which was 24 (41.4%) sequences were identical to Sal-1 reference sequence and shared with multiple populations. The identification of haplotypes identical to the Sal-1 reference sequence, and those shared across populations, may hold significance for vaccine development [37]. The DBPII Sal-1 variant is currently utilized in the development of a PvDBP-based vaccine, making these haplotypes potentially attractive candidates for further vaccine research and development [40].

For development of PvDBP-based vaccine it is important to assess the levels of DBPII genetic diversity among P. vivax field isolates globally [11]. Comparing the genetic differentiation among isolates from different regions provides insights into the evolutionary dynamics of P. vivax. The negligible genetic differentiation between Ethiopian isolates and those from Iran suggests a degree of gene flow, shared ancestry or stronger functional constraints on PvDBPII [35]. In contrast, moderate differentiation with isolates from Sudan indicates regional distinctions in parasite populations. Notably, isolates from Papua New Guinea display substantial genetic differentiation from those in Sudan and Mexico, suggesting distinct evolutionary trajectories in these regions and or the geographic barrier between the countries inhibiting gene flow [38].

The haplotype network configuration of PvDBPII provided the information about intragenic recombination in these gene fragments. Network analysis further elucidates the relationships among haplotypes, revealing both single and multiple nucleotide variations. A haplotype network was constructed to establish the relationships among the PvDBPII haplotypes from the 223 isolates of six malaria endemic countries from different continents in the world. A total of 39 haplotypes were identified with haplotype prevalence ranging from 0.45% to 27.3%. Such a finding advanced the knowledge on the parasite population dynamics in this country for the rational design of effective interventions to block disease transmission [37].There were three haplotypes only with the Ethiopian isolates, which does not shared with isolates of different geographical areas. Existing of such unique geographical sequences, may be due to biogeographic limitations, or may be due to the possible recent introduction of the parasite to that geographic area [34].

Conclusion

This study highlights the intricate genetic landscape of PvDBPII in P. vivax populations in Ethiopia and globally. The prevalence and distribution of haplotypes, regional variations, and genetic differentiation underscore the complex interplay of factors shaping the evolutionary dynamics of this parasite. Understanding these dynamics is crucial for tailoring effective malaria control strategies and advancing the knowledge on P. vivax biology. Such insights enable investigators to select dominant haplotypes crucial for designing an effective blood-stage vaccine thereby playing a pivotal role in the control and elimination of P. vivax.

Availability of data and materials

All relevant data are within the manuscript. The data that support the findings of this study are available from the corresponding author on reasonable request.

Abbreviations

BPs:

Base Pairs

DARC:

Duffy Antigen Receptor for Chemokine

DBS:

Dried blood spots

DNA:

Deoxyribonucleic acid

F ST :

Fixation Index

Hd:

Haplotype diversity

MEGA:

Diversity Molecular evolutionary Genetic Analysis

NCBI:

National Centre for Biotechnology Information

Pv/Pf :

Plasmodium vivax and Plasmodium falciparum

PvDBP:

P. vivax Duffy Binding Protein

RBCs:

Red blood cells

RDT:

Rapid diagnostic tests

RT PCR:

Real Time Polymerase Chain Reaction

SNP:

Single nucleotide polymorphism

SPSS:

Statistical package for social sciences

References

  1. WHO. World malaria report 2023. Geneva: World Health Organization; 2023.

    Google Scholar 

  2. Poespoprodjo JR, Fobia W, Kenangalem E, Lampah DA, Hasanuddin A, Warikar N, et al. Vivax malaria: a major cause of morbidity in early infancy. Clin Infect Dis. 2009;48:1704–12.

    Article  PubMed  Google Scholar 

  3. Baird KJ. Evidence and implications of mortality associated with acute Plasmodium vivax malaria. Clin Microbiol Rev. 2013;26:36–57.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Moreira CM, Abo-Shehada M, Price RN, Drakeley CJ. A systematic review of sub-microscopic Plasmodium vivax infection. Malar J. 2015;14:360.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Ford A, Kepple D, Williams J, Kolesar G, Ford CT, Abebe A, et al. Gene polymorphisms among Plasmodium vivax geographical isolates and the potential as new biomarkers for gametocyte detection. Front Cell Infect Microbiol. 2022;11: 789417.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Liu W, Li Y, Shaw KS, Learn GH, Plenderleith LJ, Malenke JA, et al. African origin of the malaria parasite Plasmodium vivax. Nat Commun. 2014;5:3346.

    Article  ADS  PubMed  Google Scholar 

  7. Ketema T, Bacha K, Getahun K, Del Portillo HA, Bassat Q. Plasmodium vivax epidemiology in Ethiopia 2000–2020: a systematic review and meta-analysis. PLoS Negl Trop Dis. 2021;15: e0009781.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Chaudhuri A, Polyakova J, Zbrzenat V, Williams K, Gulati S, Pogo AO. Cloning of glycoprotein D cDNA, which encodes the major subunit of the Duffy blood group system and the receptor for the Plasmodium vivax malaria parasite. Proc Natl Acad Sci USA. 1993;90:10793–7.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Tournamille C, Colin Y, Cartron JP, Le Van Kim C. Disruption of a GATA motif in the Duffy gene promoter abolishes erythroid gene expression in Duffy-negative individuals. Nat Genet. 1995;10:224–8.

    Article  CAS  PubMed  Google Scholar 

  10. Howes RE, Patil AP, Piel FB, Nyangiri OA, Kabaria CW, Gething PW, et al. The global distribution of the Duffy blood group. Nat Commun. 2011;2:266.

    Article  ADS  PubMed  Google Scholar 

  11. de Sousa TN, Kano FS, de Brito CFA, Carvalho LH. The duffy binding protein as a key target for a Plasmodium vivax vaccine: Lessons from the Brazilian Amazon. Mem Inst Oswaldo Cruz. 2014;109:608–17.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Grimberg BT, Udomsangpetch R, Xainli J, McHenry A, Panichakul T, Sattabongkot J, et al. Plasmodium vivax invasion of human erythrocytes inhibited by antibodies directed against the Duffy binding protein. PLoS Med. 2007;4: e337.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Hoque MR, Elfaki MMA, Ahmed MA, Lee SK, Muh F, Ali Albsheer MM, et al. Diversity pattern of Duffy binding protein sequence among Duffy-negatives and Duffy-positives in Sudan. Malar J. 2018;17:297.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Singh AP, Ozwara H, Kocken CH, Puri SK, Thomas AW, Chitnis C. Targeted deletion of Plasmodium knowlesi Duffy binding protein confirms its role in junction formation during invasion. Mol Microbiol. 2005;55:1925–34.

    Article  CAS  PubMed  Google Scholar 

  15. Ranjan A, Chitnis C. Mapping regions containing binding residues within functional domains of Plasmodium vivax and Plasmodium knowlesi erythrocyte-binding proteins. Proc Natl Acad Sci USA. 1999;96:14067–72.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Siddiqui AA, Xainli J, Schloegel J, Carias L, Ntumngia F, Shoham M, et al. Fine specificity of Plasmodium vivax Duffy Binding Protein binding engagement of the Duffy antigen on human erythrocytes. Infect Immun. 2012;80:2920–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Chen E, Salinas ND, Ntumngia FB, Adams JH, Tolia NH. Structural analysis of the synthetic Duffy binding protein (DBP) antigen DEKnull relevant for Plasmodium vivax malaria vaccine design. PLoS Negl Trop Dis. 2015;9: e0003644.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Nateghpour M, Haghi AM, Naderi B, Sepehrizadeh Z. Genetic polymorphism of Plasmodium vivax Duffy binding protein in malarious areas in southeastern of Iran. J Parasit Dis. 2017;41:1132–8.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Cole-Tobian J, King C. Diversity and natural selection in Plasmodium vivax Duffy binding protein gene. Mol Biochem Parasitol. 2003;127:121–32.

    Article  CAS  PubMed  Google Scholar 

  20. Xainli J, Adams J, King C. The erythrocyte binding motif of Plasmodium vivax Duffy binding protein is highly polymorphic and functionally conserved in isolates from Papua New Guinea. Mol Biochem Parasitol. 2000;111:253–60.

    Article  CAS  PubMed  Google Scholar 

  21. Stanisic DI, Javati S, Kiniboro B, Lin E, Jiang J, Singh B, et al. Naturally acquired immune responses to P. vivax Merozoite Surface Protein 3α and Merozoite surface protein 9 are associated with reduced risk of P. vivax malaria in Young Papua New Guinean children. PLoS Negl Trop Dis. 2013;7: e2498.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Abebe A, Bouyssou I, Mabilotte S, Dugassa S, Assefa A, Juliano JJ, et al. Potential hidden Plasmodium vivax malaria reservoirs from low parasitemia Duffy-negative Ethiopians: molecular evidence. PLoS Negl Trop Dis. 2023;17: e0011326.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Gunalan K, Lo E, Hostetler JB, Yewhalaw D, Mu J. Role of Plasmodium vivax Duffy-binding protein 1 in invasion of Duffy-null Africans. Proc Natl Acad Sci USA. 2016;113:6271–6.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, et al. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28:1647–9.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Waterhouse AM, Procter JB, Martin DMA, Clamp M, Barton GJ. Jalview Version 2 - a multiple sequence alignment editor and analysis workbench. Bioinformatics. 2009;25:1189–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35:1547–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Rozas J, Ferrer-Mata A, Sanchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, et al. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol. 2017;34:3299–302.

    Article  CAS  PubMed  Google Scholar 

  28. Leigh JW, Bryant D. POPART: Full-feature software for haplotype network construction. Methods Ecol Evol. 2015;6:1110–6.

    Article  Google Scholar 

  29. Balloux F, Lugon-Moulin N. The estimation of population differentiation with microsatellite markers. Mol Ecol. 2002;11:155–65.

    Article  PubMed  Google Scholar 

  30. Sousa TN, Tarazona-Santos EM, Wilson DJ, Madureira AP, Falco PR, Fontes CJ, et al. Genetic variability and natural selection at the ligand domain of the Duffy binding protein in brazilian Plasmodium vivax populations. Malar J. 2010;9:344.

    Article  Google Scholar 

  31. Babaeekho L, Zakeri S, Djadid N. Genetic mapping of the Duffy binding protein (DBP) ligand domain of Plasmodium vivax from unstable malaria region in the Middle East. Am J Trop Med Hyg. 2009;80:112–8.

    Article  CAS  Google Scholar 

  32. González-Cerón L, Cerritos R, Corzo-Mancilla J, Santillán F. Diversity and evolutionary genetics of the three major Plasmodium vivax merozoite genes participating in reticulocyte invasion in southern Mexico. Parasit Vectors. 2015;8:651.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Cole-Tobian JL, Cortés A, Baisor M, Kastens W, Xainli J, Bockarie M, et al. Age-acquired immunity to a Plasmodium vivax invasion ligand, the duffy binding protein. J Infect Dis. 2002;186:531–9.

    Article  CAS  PubMed  Google Scholar 

  34. Premaratne PH, Aravinda BR, Escalante AA, Udagama PV. Genetic diversity of Plasmodium vivax Duffy Binding Protein II (PvDBPII) under unstable transmission and low intensity malaria in Sri Lanka. Infect Genet Evol. 2011;11:1327–39.

    Article  CAS  PubMed  Google Scholar 

  35. Tapaopong P, da Silva G, Chainarin S, Suansomjit C, Manopwisedjaroen K, Cui L, et al. Genetic diversity and molecular evolution of Plasmodium vivax Duffy binding protein and merozoite surface protein-1 in northwestern Thailand. Infect Genet Evol. 2023;113: 105467.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Gwarinda HB, Tessema SK, Raman J, Greenhouse B, Birkholtz LM. Parasite genetic diversity reflects continued residual malaria transmission in Vhembe District, a hotspot in the Limpopo Province of South Africa. Malar J. 2021;20:96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Shi TQ, Shen HM, Chen SB, Kassegne K, Cui YB, Xu B, et al. Genetic diversity and natural selection of Plasmodium vivax Duffy binding protein-II from China-Myanmar border of Yunnan Province, China. Front Microbiol. 2021;12: 758061.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Golassa L, Messele A, Oriero EC, Amambua-Ngwa A. Sequence analysis of Plasmodium vivax Duffy binding proteins reveals the presence of unique haplotypes and diversifying selection in Ethiopian isolates. Malar J. 2021;20:312.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Núñez A, Ntumngia FB, Guerra Y, Adams JH, Sáenz FE. Genetic diversity and natural selection of Plasmodium vivax reticulocyte invasion genes in Ecuador. Malar J. 2023;22:225.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Kano FS, Sanchez BAM, Sousa TN, Tang ML, Saliba J, Oliveira FM, et al. Plasmodium vivax Duffy binding protein: baseline antibody responses and parasite polymorphisms in a well-consolidated settlement of the Amazon Region. Trop Med Int Health. 2012;17:989–1000.

    Article  PubMed  Google Scholar 

  41. Hu Y, Wang L, Mbenda HGN, Soe MT, Yu C, Feng H, et al. Genetic diversity, natural selection and haplotype grouping of Plasmodium vivax Duffy-binding protein genes from eastern and western Myanmar borders. Parasit Vectors. 2019;12:542.

    Article  Google Scholar 

  42. Almeida-De-Oliveira NK, Lima-Cury L, De Abreu-Fernandes R, De Rosa LA, De Pina-Costa A, De Souza P-D-S, et al. Extensive genetic diversity of Plasmodium vivax dbp-II in Rio de Janeiro Atlantic Forest and Brazilian Amazon Basin: evidence of positive selection. Malar J. 2020;19:81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to acknowledge; (I) Drexel University, Philadelphia, USA, (II) Ethiopian Public Health Institute, and (III) Human, Heredity and Health in Africa (H3Africa) [H3A-18-002] for the materials support. We would also like to thank all the laboratory staff at the study sites, the staff at Aklilu Lema Institute of Pathobiology, staff at Ethiopian Public Health Institute, and the staff at Drexel University. We need to express our sincere gratitude to all the regional health offices and the study sites for their cooperation. Last but not least, our sincere thanks go to all the participants who took part in this study.

Funding

The authors received no specific funding for this work. But materials used for data collection was supported by Aklilu Lema Institute of Pathobiology and Ethiopian Public Health Institute, and laboratory analysis was supported by Drexel University, Philadelphia, USA.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization and design: AA, LG. Drafting proposal: AA. Revision: AA, LG, EL, CD, DM, SD, DA, AA2, TS. Data collection: AA, LG, SD. Laboratory analysis: AA, LG, EL, CD, DM. Validation: AA, LG, EL, CD, DM, SD, DA, AA2, TS. Writing manuscript—draft: AA. Writing manuscript—review and editing: AA, LG, EL, CD, DM, SD, DA, AA2, TS.

Corresponding authors

Correspondence to Abnet Abebe or Lemu Golassa.

Ethics declarations

Ethics approval and consent to participate

Ethical approval was obtained from Aklilu Lemma Institute of Pathobiology Institutional Review Board (Ref No: ALIPB IRB/34/2013/20), Addis Ababa University, and from the national research ethical review committee (Ref No: 7/2-514/m259/35), Addis Ababa, Ethiopia. An official letter of support was sent to the study sites and the study was initiated after obtaining permission from the respective health facilities. Written informed consent and/or assent was obtained from all study participants, using the local language where appropriate.

Consent for publication

All authors have read and agreed to publish this article.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

Figure S1. A Conting view of consensus sequence coverage of Ethiopian isolates against the referance sequence. B Consensus sequence coverage of Global isolates (223 sequences and referance).

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Abebe, A., Dieng, C.C., Dugassa, S. et al. Genetic differentiation of Plasmodium vivax duffy binding protein in Ethiopia and comparison with other geographical isolates. Malar J 23, 55 (2024). https://doi.org/10.1186/s12936-024-04887-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12936-024-04887-1

Keywords