Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to main content
Advertisement
  • Loading metrics

Mathematical models of Plasmodium vivax transmission: A scoping review

  • Md Nurul Anwar ,

    Roles Conceptualization, Data curation, Formal analysis, Methodology, Visualization, Writing – original draft, Writing – review & editing

    nurul.anwar@unimelb.edu.au

    Affiliations School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia, Department of Mathematics, Bangabandhu Sheikh Mujibur Rahman Science and Technology University, Gopalganj, Bangladesh

  • Lauren Smith,

    Roles Data curation, Formal analysis, Validation, Writing – review & editing

    Affiliations The Walter and Eliza Hall Institute of Medical Research, Melbourne, Victoria, Australia, Department of Medical Biology, University of Melbourne, Melbourne, Victoria, Australia

  • Angela Devine,

    Roles Conceptualization, Formal analysis, Writing – review & editing

    Affiliations Division of Global and Tropical Health, Menzies School of Health Research, Charles Darwin University, Darwin, Australia, Health Economics Unit, Centre for Health Policy, Melbourne School of Population and Global Health, The University of Melbourne, Parkville, Australia

  • Somya Mehra,

    Roles Formal analysis, Writing – review & editing

    Affiliation School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia

  • Camelia R. Walker,

    Roles Formal analysis, Writing – review & editing

    Affiliation School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia

  • Elizabeth Ivory,

    Roles Formal analysis, Writing – review & editing

    Affiliation School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia

  • Eamon Conway,

    Roles Formal analysis, Writing – review & editing

    Affiliations The Walter and Eliza Hall Institute of Medical Research, Melbourne, Victoria, Australia, Department of Medical Biology, University of Melbourne, Melbourne, Victoria, Australia

  • Ivo Mueller,

    Roles Formal analysis, Supervision, Writing – review & editing

    Affiliations The Walter and Eliza Hall Institute of Medical Research, Melbourne, Victoria, Australia, Department of Medical Biology, University of Melbourne, Melbourne, Victoria, Australia

  • James M. McCaw,

    Roles Formal analysis, Methodology, Supervision, Writing – review & editing

    Affiliations School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia, Centre for Epidemiology and Biostatistics, Melbourne School of Population and Global Health, The University of Melbourne, Parkville, Australia

  • Jennifer A. Flegg,

    Roles Formal analysis, Methodology, Supervision, Writing – review & editing

    Affiliation School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia

  • Roslyn I. Hickson

    Roles Conceptualization, Formal analysis, Methodology, Supervision, Writing – review & editing

    Affiliations School of Mathematics and Statistics, The University of Melbourne, Parkville, Australia, Australian Institute of Tropical Health and Medicine, James Cook University, Townsville, Australia, Commonwealth Scientific and Industrial Research Organisation, Townsville, Australia

  • Article
  • Authors
  • Metrics
  • Comments
  • Media Coverage

Abstract

Plasmodium vivax is one of the most geographically widespread malaria parasites in the world, primarily found across South-East Asia, Latin America, and parts of Africa. One of the significant characteristics of the P. vivax parasite is its ability to remain dormant in the human liver as hypnozoites and subsequently reactivate after the initial infection (i.e. relapse infections). Mathematical modelling approaches have been widely applied to understand P. vivax dynamics and predict the impact of intervention outcomes. Models that capture P. vivax dynamics differ from those that capture P. falciparum dynamics, as they must account for relapses caused by the activation of hypnozoites. In this article, we provide a scoping review of mathematical models that capture P. vivax transmission dynamics published between January 1988 and May 2023. The primary objective of this work is to provide a comprehensive summary of the mathematical models and techniques used to model P. vivax dynamics. In doing so, we aim to assist researchers working on mathematical epidemiology, disease transmission, and other aspects of P. vivax malaria by highlighting best practices in currently published models and highlighting where further model development is required. We categorise P. vivax models according to whether a deterministic or agent-based approach was used. We provide an overview of the different strategies used to incorporate the parasite’s biology, use of multiple scales (within-host and population-level), superinfection, immunity, and treatment interventions. In most of the published literature, the rationale for different modelling approaches was driven by the research question at hand. Some models focus on the parasites’ complicated biology, while others incorporate simplified assumptions to avoid model complexity. Overall, the existing literature on mathematical models for P. vivax encompasses various aspects of the parasite’s dynamics. We recommend that future research should focus on refining how key aspects of P. vivax dynamics are modelled, including spatial heterogeneity in exposure risk and heterogeneity in susceptibility to infection, the accumulation of hypnozoite variation, the interaction between P. falciparum and P. vivax, acquisition of immunity, and recovery under superinfection.

Author summary

Malaria is a mosquito-borne disease and causes significant morbidity and mortality. P. vivax is one of the five species that cause malaria and the parasites can stay hidden within the human liver after an infection from an infected mosquito bite. These hidden parasites are known as hypnozoites and can activate later, causing further infections. Mathematical modelling techniques have been used since the 19th century to understand how P. vivax malaria spreads through a population. In this study, we provide a review of all the mathematical models that have been developed until May 2023 to capture P. vivax transmission. We discuss some key aspects that are crucial when developing P. vivax transmission models and highlight the models that capture these aspects. We aim to assist researchers in the field of P. vivax malaria by providing this summary and identifying the areas where further focus is needed.

1 Introduction

Malaria remains a significant public health problem, with an estimated 247 million cases and 619,000 deaths reported worldwide in 2021 alone [1]. Malaria is most prevalent in the World Health Organisation (WHO) African Region, while the South-East Asia Region has the second-highest estimated malaria burden globally. Plasmodium vivax is currently the most geographically widespread of the malaria parasites, resulting in significant associated global morbidity and mortality [25]. P. vivax has been responsible for approximately 45% of malaria cases in the WHO South-East Asia Region since 2000 and is widely prevalent in countries across Asia, Latin America, and the Pacific Islands [1, 3, 5]. P. vivax has often been overlooked and mistakenly considered as “benign” in the past [5, 6]. More recent research has produced evidence that, in addition to causing severe illness, P. vivax infection can cause long-term health consequences such as anaemia, impaired cognitive development, and chronic kidney disease [710]. The economic impact of P. vivax malaria is also significant, as the disease can lead to decreased productivity, increased healthcare costs, and reduced economic growth in endemic areas [11].

Mathematical modelling is an important tool that allows us to understand dynamic systems in various fields ranging from physics and engineering to social sciences and biology [12]. Mathematical modelling can provide valuable insight into infectious disease dynamics and plays an important role in informing public health policy and decision-making [13, 14]. Infectious disease modelling has been widely used to understand the transmission of malaria, particularly Plasmodium falciparum, and the impact of interventions to control and eliminate malaria [15, 16]. Modelling of P. vivax transmission differs from P. falciparum modelling, due to the need to account for recurrent infections caused by the activation of hypnozoites, a dormant liver stage of the parasite.

P. vivax parasites are introduced into the human body through infectious Anopheles mosquito bites. P. vivax parasites then travel to the liver, where they undergo a series of developmental and replication stages [17, 18] before the liver-stage parasites are released into the blood, causing blood-stage infections. Individuals experiencing a blood-stage infection may become symptomatic, with symptoms such as fever and fatigue, or be asymptomatic. One of the significant characteristics of P. vivax infection is that, as part of the parasites’ life-cycle, they can remain dormant in the liver for weeks or months [19] as hypnozoites that can cause further blood-stage infections (called relapses) upon reactivation. Importantly, between 79 and 96% of P. vivax cases are due to relapses (in the absence of anti-relapse treatment) [2023]. It can be challenging to distinguish a relapse from other types of recurrent malaria, such as a reinfection (i.e. malaria due to a new infectious bite) or a recrudescence (i.e. recurrence of malaria due to incomplete elimination of blood-stage infections, often associated with treatment failure) [24]. Relapse dynamics typically follow temperate or tropical phenotypes, relating to the period between primary infection and hypnozoite activation [25]. In tropical regions, relapses occur frequently within a few weeks to a few months, whereas in temperate regions, relapses typically occur between six to 12 months after initial infection. This variation in relapse frequency relates to vector dynamics and the transmission potential of P. vivax. In temperate regions, slower-relapsing hypnozoites may allow the parasites to survive colder months when mosquitoes are less prevalent, whereas, in tropical regions, a faster relapsing frequency may allow the parasite to maximise its transmission potential [26, 27]. As relapses contribute to the majority of blood-stage infections, it is important to capture these relapse dynamics when modelling P. vivax disease transmission.

The methods of incorporating hypnozoites and their associated relapse dynamics vary across the P. vivax modelling literature. Modellers have often adopted the approach of assuming a binary state (presence or absence) for hypnozoites harboured within an individual [20, 2831]. The P. vivax hypnozoite reservoir (i.e. the number of hypnozoites) is known to be non-binary [32, 33]. Due to this, more recent P. vivax models have attempted to incorporate the complex hypnozoite dynamics and capture the impact of the hypnozoite reservoir on transmission dynamics [4, 32, 3436].

The methods used to capture P. vivax immunity also vary across the modelling literature. When individuals are first infected with malaria, they naturally develop some level of immunity. This immunity can be defined as the body’s state of resistance to the infection, and, with each subsequent infection, this acquired immunity is enhanced [37]. Modellers may consider different types of immunity when modelling P. vivax transmission. This includes immunity against new infections, protection against severe malaria, anti-parasite immunity (i.e. the ability to control parasite density upon infection), clinical immunity (i.e. protection against clinical disease), and transmission-blocking immunity (i.e. immunity that reduces the probability of parasite transmission to mosquitoes) [3840].

One of the primary reasons for modelling infectious disease transmission is to understand the potential impact of treatment strategies on incidence. In terms of P. vivax, a combination therapy, known as radical cure, is needed to target both the acute infection and the dormant hypnozoite reservoir [4143]. The two drugs include: (i) a drug that clears parasites from the blood (such as chloroquine or artemisinin-based combination therapy); and (ii) an 8-aminoquinoline drug that clears hypnozoites from the liver (such as primaquine or tafenoquine). Targeting the hypnozoite reservoir is crucial in controlling or eliminating P. vivax, as transmission can be re-established from the reactivation of hypnozoites [32]. Incorporating Glucose-6-phosphate dehydrogenase deficiency (G6PD) testing is recommended before administering primaquine or tafenoquine as these drugs can cause life-threatening haemolysis in individuals with G6PD deficiency, an enzymopathy affecting up to 30% of individuals in malaria-endemic regions [44].

Other interventions that have been modelled include vector control, mass drug administration (MDA), mass screening and testing (MSaT), and P. vivax serological testing and treatment (PvSeroTAT). Vector control measures are recommended by the WHO in order to achieve elimination [45]. MDA is an effective intervention for controlling malaria and was advocated by the WHO in the 1950s to control malaria transmission [46]. MDA involves treating the entire population, or a well-defined sub-population, in a geographic location regardless of their infection status [46, 47], such that both individuals who are infected and non-infected are treated. In a radical cure MDA intervention, individuals are given artemisinin-based combination therapy to clear blood-stage parasites and primaquine (or tafenoquine) to clear hypnozoites. Due to the risks associated with radical cure treatment in G6PD–deficient individuals, mass administration of radical cure is not recommended by the WHO without first screening for G6PD deficiency [4850]. Another strategy for reducing and eliminating malaria is MSaT. This involves identifying and treating infected individuals within a specific geographical location by mass testing of all individuals regardless of their symptom status [51]. MSaT is effective in reducing malaria transmission in areas with low to moderate malaria prevalence. However, its success depends on the availability of accurate diagnostic tools, effective antimalarial drugs, and strong community participation [52, 53]. PvSeroTAT is a method for identifying individuals with recent blood-stage infections who are potential hypnozoite carriers by measuring antibodies and providing treatment with radical cure [54]. This method can identify individuals likely harbouring a hypnozoite reservoir, therefore allowing targeted treatment. Mathematical modelling has been used to understand how these different intervention strategies may impact P. vivax transmission [29, 55, 56].

In this article, we synthesise the findings of a scoping review of existing mathematical models for population-level P. vivax transmission to provide a comprehensive overview of the modelling frameworks and methods used to characterise P. vivax dynamics. In Section 2, we provide the search and inclusion criteria. We discuss the search results (Fig 1) in Section 3 as per the categorical structure in Fig 2 before concluding remarks and open problems are presented in Section 4.

2 Methods

We conducted a literature search on the 21st of May 2023, using the databases PubMed and Google Scholar to capture all relevant studies using the search terms “hypnozoite”, “malaria”, “vivax”, and “mathematical model” with Boolean operators. We screened the titles, abstracts and full text of articles for the following inclusion criteria:

  • the paper either applied or described a mathematical model of population-level P. vivax transmission dynamics, and;
  • the mathematical model of P. vivax incorporated hypnozoite dynamics, as this is a distinguishing feature of P. vivax parasites compared to other Plasmodium spp.

We excluded papers that:

  • were only concerned with the within-host dynamics of P. vivax. Although within-host models of P. vivax dynamics are important for understanding P. vivax transmission, they were not directly relevant to the aim of our study (i.e. to identify and compare mathematical models of population-level P. vivax transmission). Papers that modelled dynamics at both the within-host and population level (i.e. multi-scale models) were included.
  • only used or described mathematical models of Plasmodium species other than P. vivax (e.g. a mathematical model of P. falciparum infectious disease dynamics). Models that accounted for both P. vivax and another Plasmodium species were included.
  • were currently only available as a preprint.

Search terms were conducted in English only, and only literature published in English were considered. No limitations regarding study location, publication status (e.g. accepted, but no preprints), publication type, or publication year were included. To enhance the probability of finding all relevant literature, we screened all references within the articles that met our inclusion and exclusion criteria. Articles were then downloaded to identify key components, which are discussed in Section 3.

We categorised models depending on whether they used a stochastic or deterministic approach, and whether they were compartmental or agent-based. Deterministic models have no random variation and typically utilise a compartmental structure within a population to form differential equations to track the rate of flows between compartments. Stochastic models incorporate random variation and are useful for questions and scenarios where small population numbers or extinction are involved. In terms of P. vivax infectious disease modelling, agent-based models explicitly model P. vivax transmission dynamics at an individual-level, for example, modelling the interaction between humans and vectors and associating respective state variables and parameters to each individual and vector. In our review, we found that almost all stochastic models were also agent-based, so even though these features are not mutually exclusive, we categorised models as (i) deterministic compartmental models or (ii) stochastic agent-based models.

3 Search results

The initial search yielded 2289 articles, which was reduced to 1005 unique articles after removing duplicates between the two databases. After screening at the title level, a further 901 studies were excluded as they did not fulfil the selection criteria in Section 2. After screening the abstracts, a further 63 studies were excluded due to either (i) no underlying mathematical model being described or (ii) the model was for P. falciparum parasites only. Five additional studies were included from the selected studies’ references that were not initially identified. A total of 48 studies were finally selected for review (see Fig 1 for a summary of the selection process).

thumbnail
Fig 1. Summary of the article selection process, illustrating papers included and excluded at each stage of the review process.

https://doi.org/10.1371/journal.pcbi.1011931.g001

3.1 Model frameworks

In infectious disease dynamics, modelling frameworks typically involve a combination of mathematical models, statistical analyses, and computer simulations that aim to capture the complex dynamics of disease transmission. The Ross-Macdonald model [57], a compartmental model initially developed to describe malaria transmission dynamics, has been widely used as a modelling framework for P. vivax transmission. This modelling approach has been adapted to investigate a range of vector-borne infectious diseases, and has helped inform public health policies and intervention strategies. The first mathematical model describing P. vivax transmission was introduced by—to the best of our knowledge—Zoysa et al. (1988) [38] in a Ross-Macdonald style modelling approach. Following this, many models have now been developed.

Out of the 48 studies identified that incorporated a P. vivax transmission model, 38 (79%) utilised a deterministic and differential equation (compartmental) framework [4, 20, 24, 2830, 3235, 38, 39, 56, 5882] and nine (19%) used a stochastic and agent-based framework [21, 54, 55, 8388] (Fig 2). Only one study (2%) used both deterministic and stochastic frameworks [20] to model P. vivax transmission. Robinson et al. (2015) [20] developed the model in a deterministic framework but implemented a stochastic version of the model as a continuous-time Markov chain. For simplicity, we categorise this model as deterministic in Fig 2. Deterministic models are often the first choice amongst modellers due to their simplicity in comparison to stochastic models. Deterministic models are useful for understanding disease dynamics in large populations. Stochastic models provide more realistic and accurate representations of complex systems when dealing with small population sizes or low disease prevalence, as they can account for the randomness and variability observed in real life [89, 90].

thumbnail
Fig 2. A summary of the 48 P. vivax transmission models currently available in the literature (published as of May 21, 2023) [4, 20, 21, 24, 2830, 3235, 38, 39, 5456, 5888].

Related models (either modified or motivated by) are connected with a dashed line. Similar/same models are connected with a solid line. The coloured boxes represent key features incorporated in the models (see legend). The hexagonal boxes with the same name represent that the model was also implemented in other frameworks. The timescale (non-linear) is shown on the left.

https://doi.org/10.1371/journal.pcbi.1011931.g002

In contrast to the compartmental differential equation framework, agent-based models represent a system as a collection of individual agents that interact with each other based on a set of rules or behaviours [91, 92]. The main difference between compartmental and agent-based modelling frameworks is that a compartmental model uses aggregate variables or compartments to represent the system, while agent-based models use individuals (agents) [91]. Out of the eight studies that used an agent-based model to capture the dynamics of P. vivax transmission, only two studies modelled both the human and mosquito populations as agents [83, 84]. The other agent-based models modelled the mosquito populations as a deterministic compartmental process, such that they combined ordinary differential equations for mosquitoes with an agent-based model for humans [21, 54, 55, 8588]. Modelling mosquito dynamics as a deterministic process is an approximate strategy if the size of the mosquito population is very large and P. vivax is not near elimination. In this case, the average behaviour of the stochastic dynamics agrees with those of a deterministic process [9294]. The actual behaviour of the system depends on the interactions between individuals and mosquitoes, instead of averages. Treating the mosquito compartment as a deterministic process means that modelling elimination is not straightforward, as there will always be some non-zero number of infectious mosquitoes remaining that can trigger an infection in humans again [87].

Environmental features, ecology, and mosquito habitat locations were explicitly included when modelling malaria spread in the agent-based models that modelled both the human and mosquito population as agents [83, 84]. The most recent agent-based models modelling P. vivax dynamics [21, 54, 85, 86, 88] have evolved from a model introduced by White et al. (2018) [55]. The White et al. (2018) [55] model has been adapted to capture disease epidemiology in particular geographical settings [85], and to study the impact of different interventions (drugs or vaccination) [21, 86, 88].

While agent-based models have many advantages, their use poses several challenges. One of the main challenges of agent-based models is the difficulty in parameterising and calibrating the model, given the large number of agents and their interactions [9597]. For example, despite being an agent-based model, parameterisation is done using an ordinary differential equation system that describes the process in several models [21, 54, 55, 85]. Despite these challenges in parameterisation, agent-based models also often offer a more intuitive representation of epidemiological processes. The computational demands of agent-based models can be challenging [98], although with improving computer technology, this has become less of a concern [99].

3.2 Population-level multiscale models

Multiscale disease modelling incorporates at least two interacting scales and provides insights into disease dynamics across these scales that cannot be obtained from a single scale alone [100]. Here we only focus on within-host population models as ‘multiscale models’. For P. vivax, multiscale modelling approaches can incorporate the complex hypnozoite dynamics and their relapse effects on onward disease transmission. Most models in the existing literature only capture the population-level impact of P. vivax (boxes with a light lime green border in Fig 2). Few models capture both within-host and population-level impacts (boxes with a strong blue border in Fig 2) [21, 32, 35, 54, 55, 67, 85, 86]. The very first multiscale model for P. vivax transmission was developed by White et al. (2018) [55], and modelled the within-host hypnozoite dynamics using an agent-based model that considered heterogeneity in exposure to mosquito bites. This built on White et al. (2014) [32], which was the first to develop a within-host model that captured the dynamics of P. vivax hypnozoites. This multiscale model considered the variability in the size of hypnozoite inoculum across each mosquito bite and was subsequently used to parameterise a separate transmission model that captured the entire structure of the hypnozoite reservoir [55]. The White et al. (2014) [32] within-host model for temperate settings assumed collective dormancy. This means that the hypnozoites established by each mosquito bite progress through the dormancy states as a group or batch. This assumption may be biologically unrealistic due to the independence of individual hypnozoite activation and clearance dynamics within liver cells [101]. The other within-host models that were adapted from White et al. (2018) [55] applied the same assumption regarding batch hypnozoite behaviour [21, 54, 85, 86].

Recent work by Mehra et al. (2020) [101] relaxed the collective dormancy assumption. This enabled them to characterise the long-latency period of hypnozoite dynamics (a period of latency prior to hypnozoite activation) modelled (light purple bordered box in Fig 2) in White et al. (2014) [32] in analytical form. Later work by Mehra and colleagues embedded the activation-clearance model governing a single hypnozoite in an epidemiological framework [34]. This framework accounts for successive mosquito bites, where each bite can simultaneously establish multiple hypnozoites [34, 102], and explores the epidemiological consequence of radical cure treatment on a single individual. Anwar et al. (2022) [35] have since developed a multiscale model motivated by White et al. (2014) [32] by embedding the framework of Mehra et al. (2022) [34] for short-latency hypnozoites (deriving the relapse rate by averaging the distribution of hypnozoite burden, which is dependent on the force of reinfection) into a simple population-level model that provides key insights into both within-host level and population level dynamics. The within-host and population models were coupled at each time step (thus producing a multiscale model) to incorporate key parameters that describe the hypnozoite dynamics. This multiscale model can provide the hypnozoite distributions within the population and, more importantly, reduces the infinite compartmental structure of White et al. (2014) [32] into three compartments and relaxes the artificial truncation needed in White et al. (2014) [32] for numerical simulation. Mehra et al. (2022) [36] proposed an alternative approach, constructing a Markov population process to couple host and vector dynamics whilst accounting for (short-latency) hypnozoite accrual and superinfection as per the within-host framework proposed in Mehra et al. (2022) [34]. In the infinite population size limit, Mehra et al. (2022) [36] recovered a functional law of large numbers for this Markov population process, comprising an infinite compartment deterministic model. This infinite compartment model was then reduced into a system of integrodifferential equations based on the expected prevalence of blood-stage infection derived at the within-host scale [34]. This construction yielded population-level distributions of superinfection and hypnozoite burden, and has been generalised to allow for additional complexity, such as long-latency hypnozoites and immunity [36].

3.3 Hypnozoite dynamics and variation

The eradication of P. vivax is challenging due to the presence of the hypnozoite reservoir, which is undetectable and causes new infections long after the initial infection. In developing the first mathematical model for P. vivax, Zoysa et al. (1991) were also the first to model the effect of hypnozoite relapse on P. vivax transmission [39]. Since most P. vivax blood-stage infections are due to the reactivation of hypnozoites rather than new primary infections (in the absence of anti-relapse treatment), it is crucial that mathematical models incorporate the size of the hypnozoite reservoir [103107]. Zoysa et al. (1991) [39] assumed that the transmission dynamics could be accounted for by modelling a hypnozoite reservoir of size two (to account for up to two relapses). This assumption was later followed by Fujita et al. (2006) [30]. In reality, the average size of the hypnozoite reservoir is likely to be more than two in endemic settings, particularly those with high transmission intensity [33]. Despite having the relapse characteristic that makes P. vivax parasites unique, Nah et al. (2010) [63] and Aldila et al. (2021) [77] did not incorporate relapses in their P. vivax transmission model. In their model, individuals did not harbour hypnozoites when infected with P. vivax and hence did not experience relapse after recovery from blood-stage infection.

Modelling hypnozoite dynamics and relapses can be addressed with varying levels of complexity. The common modelling approaches to embedding hypnozoite dynamics into transmission models are the (i) binary hypnozoite model, (ii) batch hypnozoite model, and (iii) density hypnozoite model. The binary hypnozoite model is the simplest one, assuming that an individual can be either infected with or without hypnozoites [33]. Individuals experience relapse and clear hypnozoites at a constant rate. The batch hypnozoite model assumes that each infectious mosquito bite contributes to a batch of hypnozoites where each batch of hypnozoite is independent of the other. Individuals can experience multiple relapses from each batch, and the overall relapse rate depends on the total batches of hypnozoites present. Hypnozoites are subject to clearance at a constant rate. This reduces one batch of hypnozoites to a single set of dynamics, which is still truncated at a maximum number of batches [55]. The density hypnozoite model is the most complex but most biologically accurate among the approaches. In this approach, hypnozoite numbers within individuals are explicitly modelled, where each infectious mosquito bite can contribute a number of hypnozoites. Hypnozoites can either die or activate at a constant rate, and the dynamics of each hypnozoite are independent and identically distributed [34, 55, 101].

Most P. vivax transmission models consider the hypnozoite reservoir as a single compartment (i.e. binary approach), rather than explicitly accounting for a variable number of hypnozoites in the reservoir [20, 24, 28, 29, 33, 56, 5866, 6876, 7881, 83, 84]. Only a handful of models account for the variability in hypnozoite inoculation across mosquito bites (i.e. hypnozoite density model, boxes with a bright pink border in Fig 2) [4, 32, 34, 35]. If the size of the hypnozoite reservoir is modelled explicitly, the number of compartments in the model increases substantially. The very first model that accounted for the variation in hypnozoites across mosquito bites was introduced by—to the best of our knowledge—White et al. (2014) [32] for a short-latency strain (where hypnozoites can activate immediately after establishment). To account for the variation of hypnozoites across bites, White et al. (2014) modelled a system with an infinite number of compartments to represent individuals with different numbers of hypnozoites. In practice, this is truncated at 2(Lmax + 1) ordinary differential equations (for human population only), where Lmax is the maximum number of hypnozoites considered. In their model, the hypnozoite reservoir within individuals increases with new infectious bites and decreases with both activation and death of hypnozoites. This infinite compartmental system makes the model very complex, particularly when other important structures must also be incorporated, such as individual heterogeneity in bite exposure. An agent-based model later developed by White et al. (2018) [55], and other models that utilise this agent-based model, consider variation in hypnozoites within individuals, but do not account for the variability in hypnozoites across mosquito bites [21, 54, 85, 86]. Furthermore, instead of explicitly modelling hypnozoites independently, they impose the batch hypnozoite model.

The multiscale model developed by Anwar et al. (2022) [35] accounted for the variation of hypnozoites dynamics across bites. Unlike the White et al. (2014) [32] model, Anwar et al. (2022) only utilised three compartments at the population level by embedding the within-host model (short-latency) developed by Mehra et al. (2022) [34] as a system of integrodifferential equations. This relaxes the artificial truncation for the maximum number of hypnozoites used within the White et al. (2014) [32] model. Under a constant force of reinfection, Anwar et al. (2022) [35] analytically proved that the multiscale model [35] exhibits an identical steady-state hypnozoite distribution as the infinite ordinary differential equation model structure in White et al. (2014) [32]. The advantage of the multiscale model by Anwar et al. (2022) [35] is that the population-level component is considerably simpler than the 2(Lmax + 1) ordinary differential equations of White et al. (2014) [32]. The transmission models proposed by Mehra et al. (2022) [36] likewise account for variation in hypnozoite batch sizes, with Mehra et al. (2022) [36] additionally accommodating long-latency hypnozoite dynamics. The models of Mehra et al. [36] are formulated as systems of integrodifferential equations, informed by the within-host framework of Mehra et al. (2022) [34]. The analyses of Anwar and Mehra et al. [35, 36] provided insights into hypnozoite dynamics (e.g. the average size of a hypnozoite reservoir within the population and the average relapse rate), in addition to disease dynamics.

Furthermore, hypnozoites have two different relapse patterns: (i) fast relapsing in tropical regions and (ii) slow relapsing in temperate regions [25]. Depending on the setting the model is being developed for, these relapse patterns may also need to be taken into account. In most of the models, the tropical relapse pattern is achieved with a shorter latent period [4, 20, 24, 2830, 35, 38, 39, 56, 5862, 6466, 6872, 7476, 7882]. Only a few models have considered both temperate and tropical relapse patterns [3234]. White et al. (2014) [32] and Mehra et al. (2022) [34] explicitly accounted for both temperate and tropical relapse patterns by modelling the hypnozoite dynamics. In their temperate model, after hypnozoites were established they undergo a dormancy phase, which is enforced through additional model compartments. In the dormancy phase, the hypnozoites do not activate and can only die. To accommodate the tropical pattern, this dormancy phase is ignored [32, 34]. White et al. (2016) [33] later captured the temperate relapse pattern in a separate transmission model (with a binary hypnozoite model) by considering additional compartments where hypnozoite-positive individuals stay for some duration of time before they experience relapse. Kim et al. (2020) [73] modelled only the temperate relapse pattern in Korea through a survival function to obtain the time to relapse.

3.4 Superinfection

Superinfection with malaria is a common phenomenon, especially in high transmission settings, and can be defined as when an individual has more than one blood-stage infection with the same malaria-causing parasite species at a given time [108]. For P. falciparum malaria, when an infected individual (primary infection) receives a second infectious mosquito bite, they can become infected with two different parasite broods. In reference to P. vivax malaria, individuals can harbour hypnozoites in the liver even after they recover from a primary infection. Therefore, relapsing hypnozoites can provide another pathway to superinfection for individuals infected with P. vivax [108, 109].

When modelling P. vivax dynamics, it is important to consider the impact of superinfection on recovery and transmission, especially in high transmission settings, as the abundance of mosquitoes and the contribution of hypnozoite activation can frequently trigger superinfection. Superinfection can potentially delay recovery from infection [110, 111]. Most of the literature that incorporates superinfection in P. vivax transmission models (boxes with a brown border in Fig 2) [21, 2830, 32, 55, 72, 80, 85, 86] does so via the recovery rate [2830, 32]. The superinfection phenomenon was first introduced into malaria models by Macdonald (1950) [112], who assumed “The existence of infection is no barrier to superinfection, so that two or more broods or organisms may flourish side by side”. In the malaria modelling literature, it has been assumed that each brood could be cleared independently at a constant rate. Following this assumption, Dietz et al. (1974) [111] proposed a recovery rate under superinfection for P. falciparum malaria, derived at equilibrium under a constant force of reinfection. This form of the recovery rate was adopted in most studies that included superinfection via the recovery rate. This approach is straightforward when hypnozoites are integrated into the model as a binary state (i.e. an individual either has or does not have hypnozoites) [2830]. Since White et al. (2014) [32] accounts for the variation of hypnozoites, they modified the recovery rate proposed by Dietz et al. (1974) [111] to account for the additional burden of hypnozoites; however, Mehra et al. (2022) [36] argued that this modified recovery rate does not hold in the presence of hypnozoite accrual.

Generally, there are two approaches when incorporating superinfection: (i) using a corrected recovery rate that explicitly accounts for the history of past infections in the population and hypnozoite accrual dynamics [4, 36, 113] and (ii) coupling the prevalence of blood-stage infection (derived under a within-host model that accounts for superinfection) directly to the proportion of infected mosquitoes [36]. The within-host model of Mehra et al. (2022) [34] included superinfection, with each blood-stage infection (whether primary or relapse) being cleared independently, while the population-level model developed by Anwar et al. (2022) [35], which built on work of Mehra et al. (2022) [34], did not incorporate superinfection. A correction to account for superinfection, based on the recovery rate formulated by Nåsell et al. (2013) [113], was proposed in Mehra et al. (2023) [36] and incorporated in later work by Anwar et al. (2023) [4].

Superinfection was incorporated in later work, where it was assumed that different batches of hypnozoites originated from different mosquito bites [21, 54, 55, 86]. Silal et al. (2019) [72] assumed that superinfection increased the severity of the disease. That is, individuals will transition from lower to higher severity classes with a certain probability due to multiple infections. The only other study incorporating a superinfection-like phenomenon was Aldila et al. (2021), who modelled P. vivax and P. falciparum co-infection and assumed that P. vivax dominates P. falciparum [77], which does not closely resemble the definition of superinfection. This study assumed that if an individual was currently infected with P. falciparum, they would become infected with P. vivax if they received an infectious bite from a mosquito that was infected with P. vivax. The assumption that P. vivax parasites dominate P. falciparum results in the individual being infected with only P. vivax, which is not supported by the empirical biological evidence that shows that the parasitaemic load is much higher for P. falciparum [114]. Accordingly, it may not be reasonable to consider this to be a valid model of superinfections.

3.5 P. vivax and P. falciparum co-infection

Within the Asia-Pacific region, the horn of Africa, and South America, both P. vivax and P. falciparum parasites are common [72, 87]. For example, in 2019 in Cambodia, co-infection with both P. vivax and P. falciparum accounted for about 17% of malaria cases [115]. In co-endemic regions, P. falciparum infections are often followed by P. vivax infection, giving rise to the hypothesis that P. falciparum infections trigger P. vivax hypnozoite activation [72, 116118]. The high risk of P. vivax parasitaemia after P. falciparum infection is possibly related to reactivation of hypnozoites [119121]. Hypnozoites may be activated when P. falciparum parasites have been introduced into the body [122] or when the individual is exposed to Anopheles specific proteins [27]. This increased risk of P. vivax blood-stage infection following a P. falciparum infection could alternatively be explained by spatial or demographic heterogeneity in exposure and thus infection risk. Individuals either living in areas where both P. vivax and P. falciparum are highly prevalent or those that engage in an activity bringing them in frequent contact with infected mosquitoes (e.g. forest work) are more likely to be exposed to both parasites than the average person. Having a P. falciparum episode indicates the person has recently been exposed to infectious mosquito bites and is thus likely to have hypnozoites from previous exposure events (that may be triggered or activated spontaneously) or acquire a new primary P. vivax infections following recovery from P. falciparum infection [123125]. The lack of diagnostics to differentiate primary infections and relapses further complicates determining when an individual is infected with P. vivax hypnozoites. This makes it challenging to disentangle whether P. falciparum infections cause relapses through the reactivation of hypnozoites.

It is also not yet clearly understood how P. vivax and P. falciparum interact, if they compete within the host or if one species causes some, if any, protection against the other [126, 127]. A systematic review and meta-analysis showed that mixed infections (P. falciparum and P. vivax) can often cause a high rate of severe infection regardless of infection order [128]. This evidence was in contrast to a previous study which suggested that severe mixed infections were more likely to happen when P. vivax infection occurred on top of an existing P. falciparum infection (i.e. superinfection), whereas the reverse scenario, P. falciparum infection on top of an existing P. vivax infection, were more likely to result in a lower risk of severe malaria [129]. Furthermore, there is likely ascertainment bias associated with mixed infections in areas with co-circulating parasite strains, as efforts might be biased towards P. falciparum detection [130]. This is likely to be particularly common during episodes of clinical malaria when parasitaemia of one species greatly exceeds the other, and the innate host immune response may suppress both infections. Gaining a better understanding of these cross-species interactions and adjusting accounting for this co-existing phenomenon in the co-endemic region will require multi-species transmission models. Only a handful of mathematical models included both these Plasmodium species [56, 61, 62, 72, 77, 83]. While both P. vivax and P. falciparum species are included in a single model by Aldila et al. (2021) [77], this model did not account for P. vivax relapses. Five studies included both species but used two independent models for each species, which did not allow for interactions between species [56, 61, 62, 77, 83].

Whether it is important to model species interaction depends on the particular geographical setting. If both parasites are co-endemic in a setting, and the research question being considered relates to both species, then it may be important to use a model that can capture the interactions between the parasite species [72, 87, 116]. To the best of our knowledge, the first model that accounts for the interaction between both species was developed by Silal et al. (2019) [72]. In this study, a separate model (deterministic, meta-population) for both species was proposed, and these two models were entangled at each time step to incorporate interactions between the species, including treatment, triggering, and masking (non-P. falciparum infections are misdiagnosed as P. falciparum). Following this work, the first agent-based model transmission model accounting for both P. vivax and P. falciparum infections and treatment was developed by Walker et al. (2023) [87]. This model had reduced complexity compared with Silal et al.’s (2019) co-infection model, but used many of the same parameter values [72] (co-infection models shown with a vivid orange bordered box in Fig 2).

3.6 Immunity

Immunity against disease acquired through infection is usually referred to as adaptive immunity, and the primary function of adaptive immunity is to destroy foreign pathogens [131, 132]. Naturally acquired immunity to malaria is characterised by relatively rapid acquisition of immunity against severe disease and a more gradual establishment of immunity against uncomplicated malaria, while sterile immunity against infections is never achieved [133136]. In co-endemic areas, clinical immunity to P. vivax is more rapidly acquired than that due to P. falciparum [135].

How immunity is accounted for in mathematical models of malaria varies since different models consider different types of immunity. For example, immunity against new infections, immunity against severe malaria, anti-parasite immunity (i.e. the ability to control parasite density upon infection), clinical immunity (i.e. protection against clinical disease), and transmission-blocking immunity (i.e. reducing the probability of parasite transmission to mosquitoes). Immunity against new infections and severe malaria is assumed to be acquired through infection. This reduces the probability of reinfection from an infectious mosquito bite and has been modelled using up to two immunity levels [38, 39]. This type of immunity is assumed to be boosted by infection [133]. Acquiring some partial immunity (i.e. some degree of protection against malaria) following infection that wanes over time, is most common among published models [56, 5862, 64, 69, 70, 79, 84]. Some assumptions regarding immunity include that, if treated, individuals acquire some level of immunity that reduces the probability of reinfection (i.e. gain immunity against new infection) and that this wanes over time [32, 65]. The assumption regarding permanent immunity against malaria is not considered valid, as immunity often wanes rapidly when immune adults leave malaria-endemic regions [137]. Despite this, two models assumed that recovered individuals become permanently immune to P. vivax [68, 77]. Another study assumed that only a fixed proportion of individuals are immune against P. vivax rather than explicitly incorporating immunity into the model [83]. Strategies for incorporating immunity into P. vivax transmission models thus widely vary, where some assumptions are more realistic and appropriate than others.

Individuals who have not previously experienced malaria infection almost invariably become infected when first exposed to infectious mosquito bites, as immunity against malaria has not yet developed [137]. Repeated exposure to infectious bites will still likely result in infection, though these individuals may be protected against severe malaria or death [137]. Silal et al. (2019) [72] applied the opposite assumption and assumed that repeated exposure to infectious bites would likely result in severe infection however, with less probability per bite. With increasing exposure, naturally acquired immunity will also give some level of protection against symptomatic malaria. Adults living in endemic areas are more likely to have developed protective immunity compared to children due to repeated exposure over their lifetime. Adults living in endemic areas are likely to have experienced substantially more infectious mosquito bites compared to children due to age (and therefore lengthened opportunity to acquire infectious mosquito bites), greater skin surface area, and more time spent outside in environments with a higher prevalence of mosquitoes [55, 138, 139].

Immunity should be considered when using mathematical models to capture underlying disease dynamics. The assumption regarding immunity varies among models (boxes with a purple border in Fig 2). The only model that explicitly accounts for the acquisition of immunity that increases with new bites was developed by White et al. (2018) [55]. The assumption in regards to both anti-parasite immunity (ability to reduce parasite density upon infection) and clinical immunity (protection against clinical disease) depends on age and exposure to mosquito bites which is modelled using partial differential equations [55]. They also assumed that children acquired immunity through their birth parent’s immunity, which then decayed exponentially from birth. Models that were adapted from Whiteet al. (2018) [55] also allow for the acquisition of immunity [21, 54, 85, 86]. However, the immunity acquired from a primary infection may protect more strongly against relapses (which are genetically related to the primary infection) than against a new, genetically distinct primary infection. That is, hypnozoites established from an infectious bite, when reactivated, may be less likely to cause clinical infection. This is because the parasites could be genetically identical or related, which could elicit a more protective immune response due to familiarity with the primary infection [117, 140]. Thus, relapses from the same batch of hypnozoites may only cause asymptomatic infections. Despite this, no models to date have fully accounted for the relationship between relapse and immunity. Model assumptions regarding the acquisition of immunity may be too simple to capture the true underlying biology and dynamics.

Furthermore, as repeated exposure to infectious bites or relapse from the same batch of hypnozoites might cause only asymptomatic infections, this reservoir of asymptomatic infections in addition to the hypnozoite reservoir might greatly affect the overall transmission dynamics. This is recognised as an obstacle in malaria elimination [141, 142]. Asymptomatic malaria infections have received less research attention than symptomatic infections in major studies [143, 144] and even when developing mathematical models. Only a few models have explicitly considered asymptomatic infections [55, 72, 74, 87].

3.7 Effect of interventions for malaria control

In most of the models included in this review, it was assumed that treatment would be targeted towards infected individuals [21, 30, 32, 6466, 70, 79, 80, 83, 85, 145], but a range of other interventions can contribute to malaria control. The impact of a pre-erythrocytic vaccine on reducing P. vivax transmission has been studied and hypothesised that even with low efficacy, it can substantially reduce transmission [67]. The impact of radical cure treatment on the infected population has also been studied in models [21, 30, 32, 64, 85] where only a handful of models incorporate G6PD testing [21, 64, 85]. Furthermore, 11 studies (23%) evaluated the effect of MDA on disease transmission, despite few national control programs considering MDA for P. vivax control [4, 20, 28, 29, 34, 5456, 72, 87, 88] (boxes with a dark green border in Fig 2). Since MDA is recommended as an important tool to reduce asymptomatic P. falciparum infection, it is also likely to be of great importance for P. vivax elimination [20, 146, 147]. One study examined the effect of multiple MDAs and MSaTs (up to two rounds) with different drug combinations (blood-stage drug only, blood-stage drug and primaquine, or blood-stage drug and tafenoquine), finding that MDA with tafenoquine following G6PD screening could significantly reduce transmission compared to MSaT, given that no tools were available at the time to identify individuals with hypnozoites [20]. The effect of long-lasting insecticide nets along with MDA was studied using an agent-based model in Papua New Guinea, where the model predicted that MDA could reduce P. vivax transmission by between 58% and 86% [55]. The same agent-based model was later used to investigate the effect of multiple treatment strategies, including MDA, MSaT with light microscopy detection of blood-stage parasitemia, and P. vivax serological test and treatment (PvSeroTAT) [54, 88], as well as the effect of chloroquine and primaquine with vector control [85], and the potential effect of three different types of vaccines that target different stages of the P. vivax life cycle [86] in different geographical settings. The impact of different intervention scenarios, including five annual rounds of MDA on both P. falciparum and P. vivax across the Asia-Pacific, has been studied using a metapopulation model [72]. The only mixed-species agent-based model [87] was used to investigate different treatment scenarios, including current practice, accelerated radical cure, and unified radical cure provided with and without MDA (radical cure was with 14 days of primaquine and a G6PD test while the MDA was with blood-stage treatments only).

The only within-host model that accounted for the effect of MDA on each of the hypnozoites and infections was proposed by Mehra et al. (2022) [34]. This model provided base analytical expressions for the effect of multiple rounds of MDA on hypnozoite dynamics and provided the epidemiological impact of one round of MDA on a single individual. Anwar et al. (2023) recently embedded Mehra et al.’s work [34] and extended the model to study the effect of multiple MDA rounds (up to N rounds) on both within-host and population-level [4]. To the best of our knowledge, no other multiscale model has been developed that explicitly accounts for the effect of multiple rounds of MDA. The interval between MDA rounds when multiple rounds are considered, varies across different models, mostly with either one week [28, 29] or six months [20, 54] with different level of coverage. The only model that provides optimal intervals if multiple MDA rounds were under consideration is by Anwar et al. (2023) [4].

4 Open questions and conclusion

Mathematical modelling is a powerful tool for understanding, analyzing, and predicting complex real-world phenomena, as well as simulating different scenarios, testing hypotheses, and making informed decisions based on the results. Mathematical models have proven useful to characterise P. vivax transmission in different parts of the world and provide insights into the effect of different strategies to achieve elimination, including treatment, vaccination, and vector control. In this work, we provided a review of the existing mathematical models that capture P. vivax disease progression and transmission. P. vivax transmission dynamics are particularly challenging to model given the difficulties discerning relapses from reinfections and recrudescences. The choice of transmission model framework comes down to the research question at hand.

While mathematical models can provide key insights without the expense of large trials or epidemiological studies, it is important to recognize that mathematical models are not perfect representations of reality and are always subject to limitations, uncertainties, and assumptions. Therefore, using mathematical models in conjunction with empirical data, expert knowledge, and critical thinking is essential to obtain meaningful and reliable results.

Across the different approaches of mathematical modelling of P. vivax, there were varying assumptions regarding parasite dynamics and acquisition of immunity. Some models were motivated to capture realistic biological aspects of the parasite [32, 35, 38], or epidemiological and public health aspects [4, 20, 21, 2830, 33, 34, 39, 5456, 64, 65, 67, 72, 74, 79, 8387], whereas some models were motivated to construct a novel or extended mathematical model of P. vivax dynamics, i.e., focusing on the mathematical aspects of P. vivax dynamics [24, 5863, 66, 6871, 73, 75, 77, 78, 80]. As the dynamics of these type of models are well established, we argue that more importance should be placed on using these models to address the current hurdles and setbacks in achieving P. vivax elimination. For example, the effect of new drugs, emerging drug resistance, and the potential effect of vaccination (when it becomes available). Modelling different scenarios with the available tools under the current recommendations is crucial to inform decision-making regarding malaria elimination. Furthermore, given that some of the biological aspects of P. vivax are well understood, we argue that researchers should shift their focus to modelling these important aspects.

The spatial distribution of P. vivax transmission is heterogeneous, and the number of hypnozoites that an individual harbours might vary significantly; this contributes directly to the risk of hypnozoite reactivation and P. vivax relapse [148, 149]. This heterogeneity can be partially captured by modelling individuals’ movement using metapopulations and including parasite movement between different sub-populations [55, 85]. In addition to the heterogeneity in bite exposure, individuals might be heterogeneous with regard to their likelihood of becoming infected. This heterogeneity is referred to as individuals’ susceptibility to infection and is affected by a number of factors including infection history, age, gender, race, and genetics and highly impacts the disease dynamics [150152]. However, none of the current models except Corder et al. [74] explicitly consider this spatial heterogeneity in bite exposure or susceptibility to infection. Given the high degree of heterogeneity of P. vivax risk in almost all populations as well as heterogeneity in susceptibility to infection, future model development should focus on this. As more than 80% of P. vivax infections may be due to relapse (in the absence of anti-relapse treatment), and multiple hypnozoites can be established from each infectious bite, modelling the dynamics of hypnozoite variation and activation is crucial [20, 34, 35]. Another important aspect that requires more detailed attention is the interaction between multiple species of Plasmodium, particularly in areas where P. falciparum and P. vivax are co-endemic (Asia, the Horn of Africa, and the Americas). Studies show that there is a high risk of P. vivax parasitaemia after P. falciparum infection that is possibly related to reactivation of hypnozoites [119121]. This is in line with the hypothesis that P. falciparum infection might trigger underlying P. vivax infection [72, 116118]. Hence, we argue that this hypothetical triggering phenomenon should be investigated when modelling P. falciparum and P. vivax interactions.

Future P. vivax modelling efforts should also account for superinfections. Where mosquito abundance is high, transmission intensity is also likely to be high if malaria parasites are present [109, 153155]. In these scenarios, infected individuals are likely to experience multiple episodes of infection at once (i.e. superinfection). Superinfection can significantly delay recovery time, leaving ample opportunity for onward transmission from the infected individual to susceptible mosquitoes. P. vivax models should hence account for the transmission dynamics associated with superinfection. Immunity against P. vivax strongly correlates to past exposure; therefore, focus should also be placed on modelling the acquisition (and waning) of immunity related to superinfection, as multiple concurrent exposures may boost immunity more than singular exposures [36, 55]. Furthermore, as parasites from relapse are either genetically identical or related to a previous primary infection, they are more efficiently targeted by naturally acquired immune responses previously developed from the primary infection than further, genetically unrelated primary infections. As a consequence, relapses are less likely to be associated with (severe) clinical symptoms [117, 140]. This interplay between immunity and relapse has not been fully addressed in any models developed to date. Furthermore, the reservoir of asymptomatic infections as a consequence of clinical immunity can further contribute to onward transmission and should also be accounted for along with immunity. Given these important biological aspects, we suggest that future modelling should focus on developing the above-mentioned key areas: (i) spatial heterogeneity in exposure risk and heterogeneity in susceptibility to infection, (ii) accumulation of hypnozoites variation, (iii) P. falciparum and P. vivax interactions, (iv) acquisition of immunity, and (v) recovery under superinfection. Different modelling communities have recently started focusing on these areas recently, for example, modelling hypnozoite dynamics [34, 101], multispecies interactions (P. falciparum and P. vivax) [72, 87], bite exposure immunity [55] and superinfection [4, 34, 36].

Future modelling efforts should also consider focusing on the role of splenic infections in P. vivax transmission dynamics, prompted by recent findings indicating a large P. vivax parasite biomass within the spleen [156]. Additionally, improving the accuracy of models for case management, accounting for dosing, adherence, and drug pharmacokinetics/pharmacodynamics should also be the future focus as these play a crucial role in malaria elimination [157159].

No model currently includes all of the above factors that play a role in P. vivax transmission due to the complexity the resulting model would have, and not all of the factors may need to be modelled to answer the research questions at hand. The principle of parsimony is important in modelling, whereby a complex model is not necessarily better, as complex models are associated with more parameters and hence additional uncertainty and complexity in calibrating to data [160, 161]. Additionally, a model developed for a particular setting might not be suitable for other settings. Therefore, when developing models to explore P. vivax disease progression with a focus on answering specific research questions, mathematical epidemiologists and modellers should consider relevant aspects within the context of existing recommendations in a specific geographical setting.

To address the outstanding research questions identified here, a suitably skilled interdisciplinary team is required. We hope that this review can contribute to developing the common language needed for communication between different scientists by highlighting the progress of P. vivax transmission models to date.

References

  1. 1. WHO. World malaria report 2022. World Health Organization; 2022.
  2. 2. Antinori S, Galimberti L, Milazzo L, Corbellino M. Biology of human malaria plasmodia including Plasmodium knowlesi. Mediterranean Journal of Hematology and Infectious Diseases. 2012;4(1). pmid:22550559
  3. 3. Battle KE, Lucas TC, Nguyen M, Howes RE, Nandi AK, Twohig KA, et al. Mapping the global endemicity and clinical burden of Plasmodium vivax, 2000–17: a spatial and temporal modelling study. The Lancet. 2019;394(10195):332–343.
  4. 4. Anwar MN, Hickson RI, Mehra S, Price DJ, McCaw JM, Flegg MB, et al. Optimal interruption of P. vivax malaria transmission using mass drug administration. Bulletin of Mathematical Biology. 2023;85(6):43. pmid:37076740
  5. 5. Price RN, Commons RJ, Battle KE, Thriemer K, Mendis K. Plasmodium vivax in the Era of the Shrinking P. falciparum Map. Trends in Parasitology. 2020;36(6):560–570. pmid:32407682
  6. 6. Price RN, Tjitra E, Guerra CA, Yeung S, White NJ, Anstey NM. Vivax malaria: neglected and not benign. The American Journal of Tropical Medicine and Hygiene. 2007;77(6_Suppl):79–87. pmid:18165478
  7. 7. Breman JG, Alilio MS, White NJ. Defining and defeating the intolerable burden of malaria III. Progress and perspectives. The American Journal of Tropical Medicine and Hygiene. 2007;77(6_Suppl):vi–xi.
  8. 8. Tjitra E, Anstey NM, Sugiarto P, Warikar N, Kenangalem E, Karyana M, et al. Multidrug-resistant Plasmodium vivax associated with severe and fatal malaria: a prospective study in Papua, Indonesia. PLoS Medicine. 2008;5(6). pmid:18563962
  9. 9. Kochar DK, Das A, Kochar SK, Saxena V, Sirohi P, Garg S, et al. Severe Plasmodium vivax malaria: a report on serial cases from Bikaner in northwestern India. The American Journal of Tropical Medicine and Hygiene. 2009;80(2):194–198. pmid:19190212
  10. 10. Baird JK. Evidence and implications of mortality associated with acute Plasmodium vivax malaria. Clinical Microbiology Reviews. 2013;26(1):36–57. pmid:23297258
  11. 11. Devine A, Battle KE, Meagher N, Howes RE, Dini S, Gething PW, et al. Global economic costs due to vivax malaria and the potential impact of its radical cure: A modelling study. PLoS Medicine. 2021;18(6):e1003614. pmid:34061843
  12. 12. Keeling MJ, Rohani P. Modeling infectious diseases in humans and animals. Princeton University Press; 2011.
  13. 13. Barnabas RV, Laukkanen P, Koskela P, Kontula O, Lehtinen M, Garnett GP. Epidemiology of HPV 16 and cervical cancer in Finland and the potential impact of vaccination: mathematical modelling analyses. PLoS Med. 2006;3(5):e138. pmid:16573364
  14. 14. Huppert A, Katriel G. Mathematical modelling and prediction in infectious disease epidemiology. Clinical Microbiology and Infection. 2013;19(11):999–1005. pmid:24266045
  15. 15. Mandal S, Sarkar RR, Sinha S. Mathematical models of malaria-a review. Malaria Journal. 2011;10(1):1–19. pmid:21777413
  16. 16. Smith NR, Trauer JM, Gambhir M, Richards JS, Maude RJ, Keith JM, et al. Agent-based models of malaria transmission: a systematic review. Malaria Journal. 2018;17(1):1–16. pmid:30119664
  17. 17. Jones MK, Good MF. Malaria parasites up close. Nature Medicine. 2006;12(2):170–171. pmid:16462797
  18. 18. Kebaier C, Voza T, Vanderberg J. Kinetics of mosquito-injected Plasmodium sporozoites in mice: fewer sporozoites are injected into sporozoite-immunized mice. PLoS Pathogens. 2009;5(4). pmid:19390607
  19. 19. Imwong M, Snounou G, Pukrittayakamee S, Tanomsing N, Kim JR, Nandy A, et al. Relapses of Plasmodium vivax infection usually result from activation of heterologous hypnozoites. The Journal of Infectious Diseases. 2007;195(7):927–933. pmid:17330781
  20. 20. Robinson LJ, Wampfler R, Betuela I, Karl S, White MT, Suen CSLW, et al. Strategies for understanding and reducing the Plasmodium vivax and Plasmodium ovale hypnozoite reservoir in Papua New Guinean children: a randomised placebo-controlled trial and mathematical model. PLoS Med. 2015;12(10):e1001891. pmid:26505753
  21. 21. Huber JH, Koepfli C, España G, Nekkab N, White MT, Alex Perkins T. How radical is radical cure? Site-specific biases in clinical trials underestimate the effect of radical cure on Plasmodium vivax hypnozoites. Malaria Journal. 2021;20(1):1–15. pmid:34930278
  22. 22. Adekunle AI, Pinkevych M, McGready R, Luxemburger C, White LJ, Nosten F, et al. Modeling the dynamics of Plasmodium vivax infection and hypnozoite reactivation in vivo. PLoS Neglected Tropical Diseases. 2015;9(3). pmid:25780913
  23. 23. Commons RJ, Simpson JA, Watson J, White NJ, Price RN. Estimating the proportion of Plasmodium vivax recurrences caused by relapse: a systematic review and meta-analysis. The American Journal of Tropical Medicine and Hygiene. 2020;103(3):1094. pmid:32524950
  24. 24. Ghosh M, Olaniyi S, Obabiyi OS. Mathematical analysis of reinfection and relapse in malaria dynamics. Applied Mathematics and Computation. 2020;373:125044.
  25. 25. Lover AA, Coker RJ. Quantifying effect of geographic location on epidemiology of Plasmodium vivax malaria. Emerging Infectious Diseases. 2013;19(7):1058. pmid:23763820
  26. 26. Battle KE, Karhunen MS, Bhatt S, Gething PW, Howes RE, Golding N, et al. Geographical variation in Plasmodium vivax relapse. Malaria Journal. 2014;13(1):144. pmid:24731298
  27. 27. Hulden L, Hulden L. Activation of the hypnozoite: a part of Plasmodium vivax life cycle and survival. Malaria Journal. 2011;10(1):90. pmid:21496287
  28. 28. Ishikawa H, Ishii A, Kaneko A. The prevalence of Plasmodium vivax in Vanuatu Islands: Computer simulation of malaria control trails. Journal of the Faculty of Environmental Science and Technology. 2000;5(1):1–6.
  29. 29. Ishikawa H, Ishii A, Nagai N, Ohmae H, Harada M, Suguri S, et al. A mathematical model for the transmission of Plasmodium vivax malaria. Parasitology International. 2003;52(1):81–93. pmid:12543150
  30. 30. Fujita K, Tian Tian C, Nishina T, Ishikawa H. Modeling of re-emerging Plasmodium vivax in the Northern Area of the Republic of Korea Based on a Mathematical Model. Journal of the Faculty of Environmental Science and Technology. 2006;11(1):1–7.
  31. 31. Champagne C, Gerhards M, Lana J, Espinosa BG, Bradley C, González O, et al. Using observed incidence to calibrate the transmission level of a mathematical model for Plasmodium vivax dynamics including case management and importation. Mathematical Biosciences. 2022;343:108750. pmid:34883106
  32. 32. White MT, Karl S, Battle KE, Hay SI, Mueller I, Ghani AC. Modelling the contribution of the hypnozoite reservoir to Plasmodium vivax transmission. Elife. 2014;3:e04692. pmid:25406065
  33. 33. White MT, Shirreff G, Karl S, Ghani AC, Mueller I. Variation in relapse frequency and the transmission potential of Plasmodium vivax malaria. Proceedings of the Royal Society B: Biological Sciences. 2016;283(1827):20160048. pmid:27030414
  34. 34. Mehra S, Stadler E, Khoury D, McCaw JM, Flegg JA. Hypnozoite dynamics for Plasmodium vivax malaria: the epidemiological effects of radical cure. Journal of Theoretical Biology. 2022; p. 111014. pmid:35016895
  35. 35. Anwar MN, Hickson RI, Mehra S, McCaw JM, Flegg JA. A Multiscale Mathematical Model of Plasmodium Vivax Transmission. Bulletin of Mathematical Biology. 2022;84(8):1–24. pmid:35778540
  36. 36. Mehra S. Epidemic models for malaria: superinfection [MSc thesis]. The University of Melbourne; 2022.
  37. 37. Bruce-Chwatt LJ, et al. Essential Malariology. William Heinemann Medical Books Ltd; 1985.
  38. 38. de Zoysa AP, Herath PR, Abhayawardana T, Padmalal U, Mendis KN. Modulation of human malaria transmission by anti-gamete transmission blocking immunity. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1988;82(4):548–553. pmid:3076711
  39. 39. de Zoysa A, Mendis C, Gamage-Mendis A, Weerasinghe S, Herath P, Mendis KN. A mathematical model for Plasmodium vivax malaria transmission: estimation of the impact of transmission-blocking immunity in an endemic area. Bulletin of the World Health Organization. 1991;69(6):725. pmid:1786621
  40. 40. Doolan DL, Dobaño C, Baird JK. Acquired immunity to malaria. Clinical Microbiology Reviews. 2009;22(1):13–36. pmid:19136431
  41. 41. Wells TN, Burrows JN, Baird JK. Targeting the hypnozoite reservoir of Plasmodium vivax: the hidden obstacle to malaria elimination. Trends in Parasitology. 2010;26(3):145–151. pmid:20133198
  42. 42. Taylor WR, Thriemer K, von Seidlein L, Yuentrakul P, Assawariyathipat T, Assefa A, et al. Short-course primaquine for the radical cure of Plasmodium vivax malaria: a multicentre, randomised, placebo-controlled non-inferiority trial. The Lancet. 2019;394(10202):929–938. pmid:31327563
  43. 43. Poespoprodjo JR, Burdam FH, Candrawati F, Ley B, Meagher N, Kenangalem E, et al. Supervised versus unsupervised primaquine radical cure for the treatment of falciparum and vivax malaria in Papua, Indonesia: a cluster-randomised, controlled, open-label superiority trial. The Lancet Infectious Diseases. 2022;22(3):367–376. pmid:34710363
  44. 44. Recht J, Ashley EA, White NJ. Use of primaquine and glucose-6-phosphate dehydrogenase deficiency testing: divergent policies and practices in malaria endemic countries. PLoS Neglected Tropical Diseases. 2018;12(4):e0006230. pmid:29672516
  45. 45. Zuber JA, Takala-Harrison S. Multidrug-resistant malaria and the impact of mass drug administration. Infection and Drug Resistance. 2018;11:299. pmid:29535546
  46. 46. Hsiang MS, Hwang J, Tao AR, Liu Y, Bennett A, Shanks GD, et al. Mass drug administration for the control and elimination of Plasmodium vivax malaria: an ecological study from Jiangsu province, China. Malaria Journal. 2013;12(1):1–14. pmid:24175930
  47. 47. Newby G, Hwang J, Koita K, Chen I, Greenwood B, Von Seidlein L, et al. Review of mass drug administration for malaria and its operational challenges. The American Journal of Tropical Medicine and Hygiene. 2015;93(1):125. pmid:26013371
  48. 48. WHO. Second focused review meeting of the Malaria Elimination Oversight Committee (MEOC): report of a virtual meeting, 28 June–1 July 2021. 2021;.
  49. 49. Howes RE, Piel FB, Patil AP, Nyangiri OA, Gething PW, Dewi M, et al. G6PD deficiency prevalence and estimates of affected populations in malaria endemic countries: a geostatistical model-based map. PLoS Medicine. 2012;9(11). pmid:23152723
  50. 50. Watson J, Taylor WR, Bancone G, Chu CS, Jittamala P, White NJ. Implications of current therapeutic restrictions for primaquine and tafenoquine in the radical cure of vivax malaria. PLoS Neglected Tropical Diseases. 2018;12(4):e0006440. pmid:29677199
  51. 51. Singh A, Rajvanshi H, Singh MP, Bhandari S, Nisar S, Poriya R, et al. Mass screening and treatment (MSaT) for identifying and treating asymptomatic cases of malaria-malaria elimination demonstration project (MEDP), Mandla, Madhya Pradesh. Malaria Journal. 2022;21(1):1–9.
  52. 52. WHO. World malaria report 2020: 20 years of global progress and challenges. 2020;.
  53. 53. Kim S, Luande VN, Rocklöv J, Carlton JM, Tozan Y. A systematic review of the evidence on the effectiveness and cost-effectiveness of mass screen-and-treat interventions for malaria control. The American Journal of Tropical Medicine and Hygiene. 2021;105(6):1722. pmid:34491213
  54. 54. Obadia T, Nekkab N, Robinson LJ, Drakeley C, Mueller I, White MT. Developing sero-diagnostic tests to facilitate Plasmodium vivax Serological Test-and-Treat approaches: modeling the balance between public health impact and overtreatment. BMC Medicine. 2022;20(1):98. pmid:35300700
  55. 55. White MT, Walker P, Karl S, Hetzel MW, Freeman T, Waltmann A, et al. Mathematical modelling of the impact of expanding levels of malaria control interventions on Plasmodium vivax. Nature Communications. 2018;9(1):1–10. pmid:30120250
  56. 56. Águas R, Ferreira MU, Gomes MGM. Modeling the effects of relapse in the transmission dynamics of malaria parasites. Journal of Parasitology Research. 2012;2012. pmid:21966590
  57. 57. Macdonald G, et al. The epidemiology and control of malaria. The Epidemiology and Control of Malaria. 1957;.
  58. 58. Kammanee A, Kanyamee N, Tang I. Basic reproduction number for the transmission of Plasmodium vivax malaria. The Southeast Asian Journal of Tropical Medicine and Public Health. 2001;32(4):702–706. pmid:12041542
  59. 59. Pongsumpun P, Tang IM. Transmission model for Plasmodium vivax malaria. In: Proceedings of the 3rd WSEAS/IASME International Conference on Dynamical Systems and Control; 2007. p. 276–281.
  60. 60. Pongsumpun P, Tang I. Plasmodium Vivax Malaria Transmission in a Network of Villages. Eng and Tech. 2008; p. 333–337.
  61. 61. Pongsumpun P, Tang IM. Mathematical model for the transmission of P. falciparum and P. vivax malaria along the Thai-Myanmar border. International Journal of Biological and Medical Sciences. 2008;3(3).
  62. 62. Pongsumpun P, Tang IM. Impact of cross-border migration on disease epidemics: case of the P. falciparum and P. vivax malaria epidemic along the Thai-Myanmar border. Journal of Biological Systems. 2010;18(01):55–73.
  63. 63. Nah K, Kim Y, Lee JM. The dilution effect of the domestic animal population on the transmission of P. vivax malaria. Journal of Theoretical Biology. 2010;266(2):299–306. pmid:20619273
  64. 64. Chamchod F, Beier JC. Modeling Plasmodium vivax: relapses, treatment, seasonality, and G6PD deficiency. Journal of Theoretical Biology. 2013;316:25–34. pmid:22959914
  65. 65. Roy M, Bouma MJ, Ionides EL, Dhiman RC, Pascual M. The potential elimination of Plasmodium vivax malaria by relapse treatment: insights from a transmission model and surveillance data from NW India. PLoS Neglected Tropical Diseases. 2013;7(1). pmid:23326611
  66. 66. Huo HF, Qiu GM. Stability of a mathematical model of malaria transmission with relapse. In: Abstract and Applied Analysis. vol. 2014. Hindawi; 2014.
  67. 67. White M, Amino R, Mueller I. Theoretical implications of a pre-erythrocytic Plasmodium vivax vaccine for preventing relapses. Trends in parasitology. 2017;33(4):260–263. pmid:28077251
  68. 68. Prihantini , Pratama YB, Muhammad A, et al. Stability analysis model of Plasmodium vivax from Anopheles in human infection using SIDR population compartment with treatment RTS-S/AS01 vaccine in Yogyakarta. EDUCATUM Journal of Science, Mathematics and Technology. 2017;4(2):43–50.
  69. 69. Kammanee A, Tansuiy O. A mathematical model of transmission of Plasmodium vivax malaria with a constant time delay from infection to infectious. Communications of the Korean Mathematical Society. 2019;34(2):685–699.
  70. 70. Gebremichael SM, Mekonnen TT. Relapse Effect on the Dynamics of Malaria in Humans and Mosquitoes: A Mathematical Model Analysis. IOSR Journal of Mathematics. 2019;15(5):46–59.
  71. 71. Kim JE, Choi Y, Lee CH. Effects of climate change on Plasmodium vivax malaria transmission dynamics: A mathematical modeling approach. Applied Mathematics and Computation. 2019;347:616–630.
  72. 72. Silal SP, Shretta R, Celhay OJ, Gran Mercado CE, Saralamba S, Maude RJ, et al. Malaria elimination transmission and costing in the Asia-Pacific: a multi-species dynamic transmission model. Wellcome Open Research. 2019;4:62.
  73. 73. Kim S, Byun JH, Park A, Jung IH. A mathematical model for assessing the effectiveness of controlling relapse in Plasmodium vivax malaria endemic in the Republic of Korea. PLoS One. 2020;15(1):e0227919. pmid:31978085
  74. 74. Corder RM, Ferreira MU, Gomes MGM. Modelling the epidemiology of residual Plasmodium vivax malaria in a heterogeneous host population: a case study in the Amazon Basin. PLoS Computational Biology. 2020;16(3):e1007377. pmid:32168349
  75. 75. Nyaberi HO, Wakwabubi VW. Mathematical Modeling of the Dynamics of Infectious Disease With Relapse. Asian Journal of Mathematics and Computer Research. 2020; p. 28–37.
  76. 76. Baihaqi MA, Adi-Kusumo F. Modelling malaria transmission in a population with SEIRSp method. In: AIP Conference Proceedings. vol. 2264. AIP Publishing; 2020.
  77. 77. Aldila D. A superinfection model on malaria transmission: analysis on the invasion basic reproduction number. Commun Math Biol Neurosci. 2021;2021:Article–ID.
  78. 78. Abimbade SF, Olaniyi S, Ajala OA. Recurrent malaria dynamics: insight from mathematical modelling. The European Physical Journal Plus. 2022;137(3):292.
  79. 79. Tasman H, Aldila D, Dumbela PA, Ndii MZ, Herdicho FF, Chukwu CW. Assessing the Impact of Relapse, Reinfection and Recrudescence on Malaria Eradication Policy: A Bifurcation and Optimal Control Analysis. Tropical Medicine and Infectious Disease. 2022;7(10):263. pmid:36288004
  80. 80. Olaniyi S, Ajala OA, Abimbade SF. Optimal Control Analysis of a Mathematical Model for Recurrent Malaria Dynamics. In: Operations Research Forum. vol. 4. Springer; 2023. p. 14.
  81. 81. Ahkrizal A, Jaharuddin J, Nugrahani EH. Dynamics System in the SEIR-SI Model of the Spread of Malaria with Recurrence. Jambura Journal of Biomathematics (JJBM). 2023;4(1):31–36.
  82. 82. Olaniyi S, Abimbade SF, Ajala OA, Chuma FM. Efficiency and economic analysis of intervention strategies for recurrent malaria transmission. Quality & Quantity. 2023; p. 1–19.
  83. 83. Pizzitutti F, Pan W, Barbieri A, Miranda JJ, Feingold B, Guedes GR, et al. A validated agent-based model to study the spatial and temporal heterogeneities of malaria incidence in the rainforest environment. Malaria Journal. 2015;14:1–19.
  84. 84. Gharakhanlou NM, Mesgari MS, Hooshangi N. Developing an agent-based model for simulating the dynamic spread of Plasmodium vivax malaria: A case study of Sarbaz, Iran. Ecological Informatics. 2019;54:101006.
  85. 85. Nekkab N, Lana R, Lacerda M, Obadia T, Siqueira A, Monteiro W, et al. Estimated impact of tafenoquine for Plasmodium vivax control and elimination in Brazil: A modelling study. PLoS Medicine. 2021;18(4):e1003535. pmid:33891582
  86. 86. White M, Chitnis CE. Potential role of vaccines in elimination of Plasmodium vivax. Parasitology International. 2022;90:102592. pmid:35489701
  87. 87. Walker CR, Hickson RI, Chang E, Ngor P, Sovannaroth S, Simpson JA, et al. A model for malaria treatment evaluation in the presence of multiple species. Epidemics. 2023;44:100687. pmid:37348379
  88. 88. Nekkab N, Obadia T, Monteiro WM, Lacerda MV, White M, Mueller I. Accelerating towards P. vivax elimination with a novel serological test-and-treat strategy: a modelling case study in Brazil. The Lancet Regional Health–Americas. 2023;22. pmid:37250687
  89. 89. Allen LJ, Burgin AM. Comparison of deterministic and stochastic SIS and SIR models in discrete time. Mathematical Biosciences. 2000;163(1):1–33. pmid:10652843
  90. 90. Beran J. Statistics for long-memory processes. vol. 61. CRC press; 1994.
  91. 91. Volker G, Railsback SF. Individual-based modeling and ecology; 2005.
  92. 92. Bonabeau E. Agent-based modeling: Methods and techniques for simulating human systems. Proceedings of The National Academy of Sciences. 2002;99(suppl_3):7280–7287. pmid:12011407
  93. 93. Mohd MH. Revisiting discrepancies between stochastic agent-based and deterministic models. Community Ecology. 2022;23(3):453–468.
  94. 94. Yao K, Gao J. Law of large numbers for uncertain random variables. IEEE Transactions on Fuzzy Systems. 2015;24(3):615–621.
  95. 95. Crooks A, Castle C, Batty M. Key challenges in agent-based modelling for geo-spatial simulation. Computers, Environment and Urban Systems. 2008;32(6):417–430.
  96. 96. Laubenbacher R, Jarrah AS, Mortveit HS, Ravi S. Agent-based modeling, Mathematical formalism for. Complex Social and Behavioral Systems: Game Theory and Agent-Based Models. 2020; p. 683–703.
  97. 97. Colman AM. The complexity of cooperation: Agent-based models of competition and collaboration; 1998.
  98. 98. Samuelson DA, Macal CM. Agent-Based Simulation Comes of Age: Software opens up many new areas of application. OR/MS Today. 2006;33(4):34–39.
  99. 99. Darrin D, Darrin D. Pros & Cons of Agent-based modeling; 2021. Available from: https://educationalresearchtechniques.com/2020/08/07/pros-cons-of-agent-based-modeling/.
  100. 100. Garabed RB, Jolles A, Garira W, Lanzas C, Gutierrez J, Rempala G. Multi-scale dynamics of infectious diseases; 2020.
  101. 101. Mehra S, McCaw JM, Flegg MB, Taylor PG, Flegg JA. An Activation-Clearance Model for Plasmodium vivax Malaria. Bulletin of Mathematical Biology. 2020;82(2):32. pmid:32052192
  102. 102. Mehra S, McCaw JM, Flegg MB, Taylor PG, Flegg JA. Antibody Dynamics for Plasmodium vivax Malaria: A Mathematical Model. Bulletin of Mathematical Biology. 2021;83(1):1–27. pmid:33387082
  103. 103. Baird JK. Real-world therapies and the problem of vivax malaria; 2008.
  104. 104. Betuela I, Rosanas-Urgell A, Kiniboro B, Stanisic DI, Samol L, de Lazzari E, et al. Relapses contribute significantly to the risk of Plasmodium vivax infection and disease in Papua New Guinean children 1–5 years of age. The Journal of Infectious Diseases. 2012;206(11):1771–1780. pmid:22966124
  105. 105. Commons RJ, Simpson JA, Thriemer K, Hossain MS, Douglas NM, Humphreys GS, et al. Risk of Plasmodium vivax parasitaemia after Plasmodium falciparum infection: a systematic review and meta-analysis. The Lancet Infectious Diseases. 2019;19(1):91–101. pmid:30587297
  106. 106. Commons RJ, Simpson JA, Thriemer K, Humphreys GS, Abreha T, Alemu SG, et al. The effect of chloroquine dose and primaquine on Plasmodium vivax recurrence: a WorldWide Antimalarial Resistance Network systematic review and individual patient pooled meta-analysis. The Lancet Infectious Diseases. 2018;18(9):1025–1034. pmid:30033231
  107. 107. Luxemburger C, van Vugt M, Jonathan S, McGready R, Looareesuwan S, White NJ, et al. Treatment of vivax malaria on the western border of Thailand. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1999;93(4):433–438. pmid:10674098
  108. 108. Smith DL, Drakeley CJ, Chiyaka C, Hay SI. A quantitative analysis of transmission efficiency versus intensity for malaria. Nature Communications. 2010;1(1):1–9. pmid:21045826
  109. 109. Portugal S, Drakesmith H, Mota MM. Superinfection in malaria: Plasmodium shows its iron will. EMBO Reports. 2011;12(12):1233–1242. pmid:22081142
  110. 110. Smith DL, Battle KE, Hay SI, Barker CM, Scott TW, McKenzie FE. Ross, Macdonald, and a theory for the dynamics and control of mosquito-transmitted pathogens. PLoS Pathogens. 2012;8(4):e1002588. pmid:22496640
  111. 111. Dietz K, Molineaux L, Thomas A. A malaria model tested in the African savannah. Bulletin of the World Health Organization. 1974;50(3-4):347. pmid:4613512
  112. 112. Macdonald G. Tropical Diseases Bulletin. 1950;47 (10):907–915. pmid:14798656
  113. 113. Nasell I. Hybrid models of tropical infections. vol. 59. Springer Science & Business Media; 2013.
  114. 114. Battle KE, Baird JK. The global burden of Plasmodium vivax malaria is obscure and insidious. PLoS Medicine. 2021;18(10):e1003799. pmid:34618814
  115. 115. Chhim S, Piola P, Housen T, Herbreteau V, Tol B. Malaria in Cambodia: a retrospective analysis of a changing epidemiology 2006–2019. International Journal of Environmental Research and Public Health. 2021;18(4):1960. pmid:33670471
  116. 116. Snounou G, White NJ. The co-existence of Plasmodium: sidelights from falciparum and vivax malaria in Thailand. Trends in Parasitology. 2004;20(7):333–339. pmid:15193565
  117. 117. White NJ. Determinants of relapse periodicity in Plasmodium vivax malaria. Malaria Journal. 2011;10(1):297. pmid:21989376
  118. 118. Lin JT, Bethell D, Tyner SD, Lon C, Shah NK, Saunders DL, et al. Plasmodium falciparum gametocyte carriage is associated with subsequent Plasmodium vivax relapse after treatment. PLoS One. 2011;6(4):e18716. pmid:21533092
  119. 119. Taylor AR, Watson JA, Chu CS, Puaprasert K, Duanguppama J, Day NPJ, et al. Resolving the cause of recurrent Plasmodium vivax malaria probabilistically;10(1):5595.
  120. 120. Hossain MS, Commons RJ, Douglas NM, Thriemer K, Alemayehu BH, Amaratunga C, et al. The risk of Plasmodium vivax parasitaemia after P. falciparum malaria: An individual patient data meta-analysis from the WorldWide Antimalarial Resistance Network;17(11):e1003393.
  121. 121. Commons RJ, Simpson JA, Thriemer K, Hossain MS, Douglas NM, Humphreys GS, et al. Risk of Plasmodium vivax parasitaemia after Plasmodium falciparum infection: a systematic review and meta-analysis;19(1):91–101.
  122. 122. Shanks GD, White NJ. The activation of vivax malaria hypnozoites by infectious diseases. The Lancet Infectious Diseases. 2013;13(10):900–906. pmid:23809889
  123. 123. Amratia P, Psychas P, Abuaku B, Ahorlu C, Millar J, Oppong S, et al. Characterizing local-scale heterogeneity of malaria risk: a case study in Bunkpurugu-Yunyoo district in northern Ghana. Malaria Journal. 2019;18:1–14.
  124. 124. Haque U, Magalhães RJS, Reid HL, Clements AC, Ahmed SM, Islam A, et al. Spatial prediction of malaria prevalence in an endemic area of Bangladesh. Malaria Journal. 2010;9(1):1–10.
  125. 125. Hofmann NE, Karl S, Wampfler R, Kiniboro B, Teliki A, Iga J, et al. The complex relationship of exposure to new Plasmodium infections and incidence of clinical malaria in Papua New Guinea. Elife. 2017;6:e23708. pmid:28862132
  126. 126. Muh F, Kim N, Nyunt MH, Firdaus ER, Han JH, Hoque MR, et al. Cross-species reactivity of antibodies against Plasmodium vivax blood-stage antigens to Plasmodium knowlesi. PLoS Neglected Tropical Diseases. 2020;14(6):e0008323. pmid:32559186
  127. 127. Cox-Singh J, Singh B. Knowlesi malaria: newly emergent and of public health importance? Trends in Parasitology. 2008;24(9):406–410. pmid:18678527
  128. 128. Kotepui M, Kotepui KU, De Jesus Milanez G, Masangkay FR. Plasmodium spp. mixed infection leading to severe malaria: A systematic review and meta-analysis. Scientific Reports. 2020;10(1):11068. pmid:32632180
  129. 129. Mohapatra M, Dash L, Bariha P, Karua P. Profile of mixed species (Plasmodium vivax and falciparum) malaria in adults. J Assoc Physicians India. 2012;60:20–24. pmid:23777020
  130. 130. Stresman G, Sepúlveda N, Fornace K, Grignard L, Mwesigwa J, Achan J, et al. Association between the proportion of Plasmodium falciparum and Plasmodium vivax infections detected by passive surveillance and the magnitude of the asymptomatic reservoir in the community: a pooled analysis of paired health facility and community data. The Lancet Infectious Diseases. 2020;20(8):953–963. pmid:32277908
  131. 131. Mayer G. Microbiology and Immunology On-Line Textbook. USC School of Medicine Available [online]< http://pathmicro.med.sc.edu/ghaffar/innate.htm>, retrieved on 10th May. 2010;.
  132. 132. Chaffey N. Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K. and Walter, P. Molecular biology of the cell. 4th edn.; 2003.
  133. 133. Barua P, Beeson JG, Maleta K, Ashorn P, Rogerson SJ. The impact of early life exposure to Plasmodium falciparum on the development of naturally acquired immunity to malaria in young Malawian children. Malaria Journal. 2019;18:1–12. pmid:30658632
  134. 134. Grüner AC, Mauduit M, Tewari R, Romero JF, Depinay N, Kayibanda M, et al. Sterile protection against malaria is independent of immune responses to the circumsporozoite protein. PloS One. 2007;2(12):e1371. pmid:18159254
  135. 135. Mueller I, Galinski MR, Tsuboi T, Arevalo-Herrera M, Collins WE, King CL. Natural acquisition of immunity to Plasmodium vivax: epidemiological observations and potential targets. Advances in Parasitology. 2013;81:77–131. pmid:23384622
  136. 136. López C, Yepes-Pérez Y, Hincapié-Escobar N, Díaz-Arévalo D, Patarroyo MA. What is known about the immune response induced by Plasmodium vivax malaria vaccine candidates? Frontiers in Immunology. 2017;8:126. pmid:28243235
  137. 137. Langhorne J, Ndungu FM, Sponaas AM, Marsh K. Immunity to malaria: more questions than answers. Nature Immunology. 2008;9(7):725–732. pmid:18563083
  138. 138. Port G, Boreham P, Bryan JH. The relationship of host size to feeding by mosquitoes of the Anopheles gambiae Giles complex (Diptera: Culicidae). Bulletin of Entomological Research. 1980;70(1):133–144.
  139. 139. Carnevale P, Frezil J, Bosseno M, Le Pont F, Lancien J. The aggressiveness of Anopheles gambiae A in relation to the age and sex of the human subjects. Bulletin of the World Health Organization. 1978;56(1):147–154. pmid:307444
  140. 140. Joyner CJ, Brito CF, Saney CL, Joice Cordy R, Smith ML, Lapp SA, et al. Humoral immunity prevents clinical malaria during Plasmodium relapses without eliminating gametocytes. PLoS Pathogens. 2019;15(9):e1007974. pmid:31536608
  141. 141. Laishram DD, Sutton PL, Nanda N, Sharma VL, Sobti RC, Carlton JM, et al. The complexities of malaria disease manifestations with a focus on asymptomatic malaria. Malaria Journal. 2012;11(1):1–15. pmid:22289302
  142. 142. Galagan SR, Prue CS, Khyang J, Khan WA, Ahmed S, Ram M, et al. The practice of jhum cultivation and its relationship to Plasmodium falciparum infection in the Chittagong Hill Districts of Bangladesh. The American Journal of Tropical Medicine and Hygiene. 2014;91(2):374. pmid:24821843
  143. 143. Drakeley C, Corran P, Coleman P, Tongren J, McDonald S, Carneiro I, et al. Estimating medium-and long-term trends in malaria transmission by using serological markers of malaria exposure. Proceedings of the National Academy of Sciences. 2005;102(14):5108–5113. pmid:15792998
  144. 144. Ibrahim AO, Bello IS, Ajetunmobi AO, Ayodapo A, Afolabi BA, Adeniyi MA. Prevalence of asymptomatic malaria infection by microscopy and its determinants among residents of Ido-Ekiti, Southwestern Nigeria. Plos One. 2023;18(2):e0280981. pmid:36787321
  145. 145. Kim JR, Nandy A, Maji AK, Addy M, Dondorp AM, Day NP, et al. Genotyping of Plasmodium vivax reveals both short and long latency relapse patterns in Kolkata. PloS One. 2012;7(7):e39645. pmid:22808048
  146. 146. Grueninger H, Hamed K. Transitioning from malaria control to elimination: the vital role of ACTs. Trends in Parasitology. 2013;29(2):60–64. pmid:23228225
  147. 147. Okell LC, Griffin JT, Kleinschmidt I, Hollingsworth TD, Churcher TS, White MJ, et al. The potential contribution of mass treatment to the control of Plasmodium falciparum malaria. PLOS One. 2011;6(5):e20179. pmid:21629651
  148. 148. Popovici J, Ménard D. Challenges in antimalarial drug treatment for vivax malaria control. Trends in Molecular Medicine. 2015;21(12):776–788. pmid:26611336
  149. 149. Stadler E, Cromer D, Mehra S, Adekunle AI, Flegg JA, Anstey NM, et al. Population heterogeneity in Plasmodium vivax relapse risk. PLOS Neglected Tropical Diseases. 2022;16(12):e0010990. pmid:36534705
  150. 150. Smith D, Dushoff J, Snow R, Hay S. The entomological inoculation rate and Plasmodium falciparum infection in African children. Nature. 2005;438(7067):492–495. pmid:16306991
  151. 151. Filion GJ, Paul RE, Robert V. Transmission and immunity: the importance of heterogeneity in the fight against malaria. Trends in Parasitology. 2006;22(8):345–348. pmid:16797234
  152. 152. Valim C, Mezzetti M, Maguire J, Urdaneta M, Wypij D. Estimation of vaccine efficacy in a repeated measures study under heterogeneity of exposure or susceptibility to infection. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 2008;366(1874):2347–2360. pmid:18407892
  153. 153. Bashar K, Tuno N. Seasonal abundance of Anopheles mosquitoes and their association with meteorological factors and malaria incidence in Bangladesh. Parasites & Vectors. 2014;7(1):1–10. pmid:25233890
  154. 154. Reimer LJ, Thomsen EK, Koimbu G, Keven JB, Mueller I, Siba PM, et al. Malaria transmission dynamics surrounding the first nationwide long-lasting insecticidal net distribution in Papua New Guinea. Malaria Journal. 2016;15:1–11. pmid:26753618
  155. 155. Keven JB, Katusele M, Vinit R, Rodríguez-Rodríguez D, Hetzel MW, Robinson LJ, et al. Vector composition, abundance, biting patterns and malaria transmission intensity in Madang, Papua New Guinea: assessment after 7 years of an LLIN-based malaria control programme. Malaria Journal. 2022;21(1):1–15. pmid:34983530
  156. 156. Kho S, Qotrunnada L, Leonardo L, Andries B, Wardani PA, Fricot A, et al. Hidden biomass of intact malaria parasites in the human spleen. New England Journal of Medicine. 2021;384(21):2067–2069. pmid:34042394
  157. 157. Grobusch MP, Schlagenhauf P. Primaquine and the power of adherence in radical cure. The Lancet Infectious Diseases. 2022;22(3):304–305. pmid:34710364
  158. 158. Rathmes G, Rumisha SF, Lucas TC, Twohig KA, Python A, Nguyen M, et al. Global estimation of anti-malarial drug effectiveness for the treatment of uncomplicated Plasmodium falciparum malaria 1991–2019. Malaria Journal. 2020;19:1–15. pmid:33081784
  159. 159. Mallayasamy S, Chaturvedula A, Blaschke T, Fossler MJ. A systematic evaluation of effect of adherence patterns on the sample size and power of a clinical study. CPT: Pharmacometrics & Systems Pharmacology. 2018;7(12):818–828. pmid:30291680
  160. 160. Killeen GF, McKenzie FE, Foy BD, Schieffelin C, Billingsley PF, Beier JC. A simplified model for predicting malaria entomologic inoculation rates based on entomologic and parasitologic parameters relevant to control. The American Journal of Tropical Medicine and Hygiene. 2000;62(5):535. pmid:11289661
  161. 161. Amadi M, Shcherbacheva A, Haario H. Agent-based modelling of complex factors impacting malaria prevalence. Malaria Journal. 2021;20(1):1–15. pmid:33858432