Open access peer-reviewed chapter

Human and Simian Malaria in the Greater Mekong Subregion and Challenges for Elimination

Written By

Jeffrey Hii, Indra Vythilingam and Arantxa Roca-Feltrer

Submitted: 20 June 2017 Reviewed: 30 March 2018 Published: 18 July 2018

DOI: 10.5772/intechopen.76901

From the Edited Volume

Towards Malaria Elimination - A Leap Forward

Edited by Sylvie Manguin and Vas Dev

Chapter metrics overview

1,487 Chapter Downloads

View Full Metrics

Abstract

In recent years malaria initiatives have increasingly shifted from malaria control to a focus on achieving malaria elimination in the Southeast Asia region. However, this region experiences unique challenges in this transition due to its distinctive malaria ecosystem (mainly related to forests) and high volume of population movement (both within and between countries). These bioenvironmental factors increase the exposure of populations at higher risk due to their close association with forest, and contributes to outdoor and residual transmission. Given that this region has also historically been the source of resistance to anti-malarial drugs, the potential spread of artemisinin resistance via global transportation routes would pose a major threat to malaria control and elimination efforts worldwide. In addition, other factors also hinder the malaria elimination goal such as importation of parasite infection, uncontrolled monkey malaria (Plasmodium knowlesi), or the fact that many countries in this region experience mixed infections where P. vivax becomes a more predominant species as overall malaria transmission decreases. This chapter addresses these challenges in detail and provide recommendations and key priorities to overcome these obstacles to accelerate efforts for achieving malaria elimination.

Keywords

  • malaria
  • elimination
  • Greater Mekong Subregion
  • drug-resistance
  • Plasmodium knowlesi
  • vivax malaria
  • residual transmission

1. Introduction

In the Greater Mekong Subregion (GMS)1, malaria is still a substantial public health problem, especially along international borders and forested areas, adversely putting populations such as migrants, refugees, and forest workers most at risk. In 2013, there were 447,800 malaria cases and 342 deaths in the GMS, with close to 700 million people living in risk areas [1]. Between 2012 and 2016, the reported number of malaria cases in the GMS fell by 74% (Figure 1) and malaria deaths by 91% in the same period (Figure 2).

Figure 1.

Declining trends of malaria transmission in the Greater Mekong Subregion (GMS) since 2012 (source: [2]).

Figure 2.

Declining trend of malaria deaths in the Greater Mekong Subregion (GMS) since 2012 (source: [2]).

Mid-year estimates for 2017 point to a further decline in cases [2]. Contributing to these impressive results, all six countries of the Subregion are making significant headway towards a common target: eliminating malaria by the year 2030 at the latest.

These goals will not be easy to achieve. Despite these reductions and the subsequent move towards elimination, malaria remains an important cause of morbidity for an estimated 32 million inhabitants, especially in remote areas with low population densities and limited healthcare services and infrastructure, located in and near forested areas, which often lie close to international borders [3, 4]. In many places, the population groups most affected are ethnic minorities and forest-goers who are rapidly becoming the most important source of transmission in areas where main vectors are present. Within these groups, cultural and linguistic barriers often constrain malaria control efforts due to their high mobility and low access to interventions to prevent, diagnose and treat malaria.

In some areas the malaria situation has deteriorated by armed conflict affecting access to malaria control services. Population movements are a key feature in the GMS and are largely occupationally/economically driven; occur within borders and across borders; involve multiple factors and complex dynamics of movement; and affect different subsets of moving populations [5], thus further complicating the epidemiology and control of the disease [6]. The rapid increase in the number of large infrastructure and agricultural development projects in the region is also having a significant impact on the epidemiology of communicable diseases in general, and malaria in particular [7]. This chapter addresses several key challenges faced by elimination programmes to contain the unacceptably high disease burden against the background of rapidly declining incidence.

Advertisement

2. Resistance to artemisinin and ACT: current and future approaches

Antimalarial drug resistance is not a new biological phenomenon. In the 1970s and 1980s, Plasmodium falciparum—the parasite species responsible for the most common and deadliest form of malaria—developed widespread resistance to previous antimalarial medicines, such as chloroquine and sulfadoxine-pyrimethamine (SP) [8]. Artemisinin based combination therapies (ACTs), introduced in the 1990s, are currently the most effective antimalarial drugs [9] and represent the first line-treatment for uncomplicated falciparum malaria in all endemic countries.

Although artemisinin usually kills all malaria parasites, the use of a combination of drugs—as opposed to monotherapy—helps ensure that any remaining parasites will be killed by the partner drug before the resistant parasites can spread. According to the World Health Organisation (WHO), clinical artemisinin (and its derivatives) resistance is defined as delayed parasite clearance and represents a partial/relative resistance that has thus far only affected ring-stage malaria parasites [10]. In Southeast Asia, however, some malaria parasites have already developed resistance to artemisinin-based drugs; a recent report of a single multidrug resistant malaria parasite lineage (PfPailin) with associated piperaquine resistance in Vietnam and its implications of subsequent transnational spread is of international concern [11]. Artemisinin resistance was first reported along the Thailand-Cambodia border in 2008 [12, 13] and has continued to spread in all Greater Mekong Subregion countries [14, 15, 16, 17, 18]. In addition, artemisinin resistance has been involved in selecting for resistance to ACT partner drugs, resulting in high late treatment failure rates with dihydroartemisinin-piperaquine in Cambodia [14, 19, 20, 21, 22, 23, 24, 25] and with artesunate-mefloquine on the Thai-Myanmar border [26].

There are many factors that are thought to have contributed to the emergence and spread of artemisinin resistance in the GMS. One important factor is thought to be the use of oral artemisinin monotherapy (AMT) in place of WHO-recommended ACTs (as unregulated artemisinin or artesunate monotherapy has been available since mid-1970s in the region). In Myanmar, private healthcare facilities and healthcare providers who prioritize consumers’ demand instead of recommended practices were more likely to stock oral AMT [26, 27]. Malaria elimination strategies should include targeted interventions to effectively reach these outlets. Fortunately, a major achievement during the resistance containment (and more recently elimination) activities has ceased the use of artemisinin monotherapies. ACT watch methods are monitoring displacement of oral AMTs, a major objective of the resistance containment strategy [28], and data will feed into regional score cards such as the Asia Pacific Leaders Malaria Alliance Access to Quality Medicines Task Force and the World Health Organisation (WHO) Emergency Response to Artemisinin Resistance (ERAR), which are vested in supporting national programs in tracking progress towards halting the availability and use of oral AMTs [28]. In Southeast Asia, where malaria transmission is generally low and emergence of resistance has been documented in multiple independent locations [29]; containment programmes have been converted into elimination of P. falciparum strategies to ensure halting the spread of resistance entirely.

Other contributing factors are the use of substandard and counterfeit anti-malarial drugs and the difficulty of controlling malaria within migrant and hard-to-reach populations [30]. Given the transnational nature of this problem, the establishment of effective mechanisms for cross-country surveillance, information exchange and coordinated action is also necessary. This includes reinforcing existing institutional frameworks for regional health cooperation, particularly the Association of Southeast Asian Nations, and their potential to support enhanced capacities and cooperation to address this challenge [31]. Lastly, selection pressure—genetic mutations of wild-type genes in the parasite render them insusceptible to antimalarial drug treatment—is also thought to be important. The use of antimalarial drugs in patients with parasites containing mutations can eliminate susceptible parasites but leave resistant mutants to survive and reproduce [32].

More recently another potential contributing factor has been hypothesized. Given that there are parasite isolates that do not infect some Anopheles species, it is thought that artemisinin-resistant parasites are spreading so fast in Southeast Asia because they infect most or all native Anopheles species (e.g., Anopheles dirus and An. minimus), including African vector counterparts such as An. coluzzii (formerly Anopheles gambiae M form) [33]. The ability of artemisinin-resistant parasite clones to infect three highly genetically diverse vectors suggests that these resistant parasites have enhanced their transmission in the region and could effectively spread in sub-Saharan Africa, where most of the world’s malaria mortality, morbidity, and transmission occurs [33, 34].

Since there are no equally effective alternative drugs to treat malaria, the spread of artemisinin resistance through India (Asia) to Africa and beyond could be a catastrophic setback to global efforts to control and eliminate the disease. Infection and mortality rates could dramatically increase in both regions, reversing the progress made towards malaria control and elimination efforts. The spread of artemisinin resistance would in turn expose the partner drugs in ACTs to greater selection pressure for the development of resistance and increased failure rates for the treatment of uncomplicated malaria. For severe malaria, the recent change in recommended treatment from quinine to artesunate [35] increased survival by 25%, and many endemic countries have adopted (or are adopting) this policy [36, 37]. Reverting back to quinine because of artemisinin resistance would also jeopardise all these gains achieved in the management of severe malaria.

The spread of ACT resistance requires constant and comprehensive monitoring across regions. Continuous monitoring of drug resistance in malaria-endemic countries along with contributing factors is a key and will enable health authorities and practitioners to prevent drug resistance from spreading. WHO issues regular reports about the status of artemisinin resistance in malaria endemic countries [38], provides updates on the status of resistance to artemisinins and ACT, and maintains a network of sentinel sites performing therapeutic efficacy studies of first and second-line antimalarial drugs [38, 39].

Advertisement

3. Targeting interventions in hard-to-reach population groups

Although most of malaria endemic countries in Southeast Asia have incorporated malaria elimination goals in their national strategic plans, yet this region experiences high volume of population movement (both within and between countries) causing a great hindrance in achieving their elimination targets given the increased risk of importation of infection, spread of drug resistance, and challenges in providing healthcare services to mobile populations at higher risk of malaria [40, 41].

It is the movement of populations that results in importation of new infections leading to a source of local transmission [42, 43]. Cross-border movement of populations has contributed to establishment of “hot-spots” of high transmission along international borders [44, 45], and spread of drug resistance [6], because mobile populations often experience delays in receiving diagnosis and treatment, have improper health-seeking behaviour or self-medicate [88], and are subject to lower levels of surveillance [41, 46, 47]. Population mobility in the GMS is strongly associated with shifting land use, including large rural infrastructure projects and agricultural industries that attract migrant labor and influence human-vector contact. With the recent Association of Southeast Asian Nations (ASEAN) Economic Community agreement, allowing free movement of goods, services and labor between ASEAN countries [48]; population movement is expected to rise even more in the coming years [6].

In addition, the epidemiology of malaria in many parts of Southeast Asia is shifting toward migratory labor force that gets exposed to vectors in the forest, construction sites, and has variable access to healthcare services [46, 47, 49, 50, 51, 52, 53]. Since forested regions are concentrated along borders and much of the cross-border movement is from the migrant labor population, malaria prevalence in these pockets was hypothesized to represent foci of hot-spots. Following this rationale, the increased malaria risk in these groups was recently documented in a cross-border malaria project conducted in the Thai-Cambodian, Lao-Cambodia and Vietnam-Cambodian borders. In this study [45], it was observed that the odds of infection in security/armed forces and forest-goers was 8 and 13 times higher compared to low-risk occupations (e.g., teachers, traders, salesmen, etc.). Mechanisms and risk reduction strategies should be in place to appropriately cover these special occupational high-risk groups.

Therefore, although population mobility is a key factor to take into account when addressing drug resistance, it suffers from a range of challenges that limit countries’ capacity to effectively engage and deliver interventions to migrant and mobile populations (MMPs). In addition, outdoor biting mosquitoes represent a major challenge for vector control for MMPs working during the night or sleeping outdoors, as well as forest-fringe communities.

Another challenge is the large proportion of asymptomatic infections within geographical clusters of high malaria transmission (hot-spots), where infections with low and sub-microscopic parasite densities are highly prevalent in MMP and other risk groups [54]. Asymptomatic carriers can repeatedly fuel transmission to surrounding areas as the vector population expands during the wet season [55, 56, 57]. Whilst groups of homesteads consisting of asymptomatic carriers can act as stable clusters over several years [7], it is likely that the flight range of 800 m for An. dirus may account for increased probability of repeated mosquito feeding in the same house and clustering of cases over the dry season in Southeastern Thailand [58]. Recent clusters of malaria infection among the parasite reservoir responsible for preserving malaria over the dry season in Ratanakiri Province (northeastern Cambodia) may also explain recurrent transmission at the onset of the rainy season when the vector populations expand [59]. This reservoir is often not (completely) covered by control strategies [60] and parasite specific approaches are non-existent [61]. Programmatic interventions to interrupt transmission in “hidden” asymptomatic reservoir must focus on individuals with malaria infection at early stage, as asexual parasitemia left untreated will eventually produce gametocytes, and diagnostics for the sexual stage are limited [62].

This represents an important hindrance to malaria elimination as these infections are unlikely to be detected by passive surveillance and conventional diagnostic tools, and therefore require additional approaches to effectively reach all infections [63]. A combination of methods, or new diagnostics, may be required to detect infections in these asymptomatic parasite reservoirs. Also, a cross-sectoral response, involving non-health government agencies and the private sector addressing the links between malaria transmission, mobility and labor, will play an important role in responding to drug resistance and achieving elimination in the Southeast Asia region. Preliminary studies of the use of peer outreach workers to conduct screening of suspected cases, providing health education, and distributing nets in hot-spot areas in or near the forest, suggest that it is feasible to target high-risk populations in a culturally appropriate and evidence-based manner to reach the goal of elimination in Pursat Province, Cambodia [64]. Mobile Malaria Workers or peer outreach activities often face logistic challenges including muddy roads, river crossings, and transportation difficulties that make it hard to quickly respond to all infections. The recent President Malaria Initiative (PMI) studies show this is a potential resource that can be piloted or replicated across GMS countries (John Hustedt, personal communication).

Lastly, persisting low health-care coverage and access in remote locations remains an important challenge for mobile populations and migrant workers in some Southeast Asian countries, limiting the ability of malaria programmes to effectively capture these groups through the routine surveillance system, but most importantly to adequately provide the necessary preventive measures and care needed [65]. It is encouraging, however, to see that malaria infection rates in people who had sought treatment, or blood-smear examined in a previous malaria episode, and/or who knew how to prevent malaria (e.g., sleeping under a mosquito net), tend to be lower than those that did not seek treatment or had inadequate malaria knowledge [45]. This highlights the importance of scaling up and expanding the reach of point of access care and dissemination of information, such as through border posts or at large development or construction areas that are likely to host high-risk malaria occupational groups. These posts can potentially be used as effective channels to target and deliver specific interventions such as Behavioral Change Communication (BCC) materials, insecticide-treated uniforms or hammock nets.

Therefore, there is an urgent need to develop appropriate and sustainable malaria services for MMPs in different settings, in the context of the spread of artemisinin resistance and malaria elimination in the GMS. Different types of mobility require different malaria control interventions and therefore elimination strategies that should be based on an in-depth understanding of malaria risk in each group [66]. A population movement framework can assist in improved targeting of malaria (and other public health interventions) by going beyond a simple labeling of risk groups to develop a better understanding of risk behaviour and vulnerabilities. The implementation of the framework should be carefully evaluated to identify the changes in coverage, access, and effectiveness of the programme efforts to serve MMPs [67].

Advertisement

4. Residual and outdoor transmission: how much and where?

In 2012, global malaria transmission was reported as mainly attributable to 51 Anopheles species, with an average of about 3 major species per country [68]. Biological factors that determine whether a species becomes a major local threat are its competence for transmitting human malaria parasites, its anthropophilic versus zoophilic preference, and its abundance in relation to its ability to multiply, survive, and compete for resources with other Anopheles species. The third of these factors is regulated by the ecosystem’s carrying capacity for potent vectors depending on their ecological niches [69]. Species of several Anopheles complexes are either major or secondary malaria vectors depending on their geographical range of distribution [70]. The peculiarity of these sibling species within a complex is that they cannot be distinguished using morphological criteria. However, several Asian malaria vectors within the Dirus, Leucosphyrus, Minimus, Maculatus, Culicifacies, Sundaicus, Subpictus complexes or groups show similar morphological characteristics, different ecological traits and vector competencies and overlapping geographical distribution with other vectors and non-vectors [70, 85, 92]. As some of these sibling species occur sympatrically and differ in their ability to transmit malaria and in their behaviour, the use of molecular tools to differentiate the vectors from the non-vectors is essential to target the correct species in vector control programs.

Malaria vector control relies largely on Long-Lasting Insecticidal Nets (LLINs) and Indoor Residual Spraying (IRS), along with Larval Source Management (LSM) as a supplementary measure appropriate in certain settings. These core interventions are highly efficacious for control of susceptible malaria vectors when implemented at universal coverage; LLINs and IRS contributed to a 48% reduction in malaria infection prevalence and 47% reduction in mortality worldwide between 2000 and 2013 [71]. However, malaria transmission can persist even when LLINs and/or IRS are effectively implemented and malaria vectors are susceptible to the insecticides used. This may be due to a combination of vector and human behaviour and bionomical characteristics, which compromise inadequate control measures against early and/or outdoor biting mosquitoes, and human activity away from protected houses or places at peak biting times. With current efforts focusing on malaria elimination [72], there is considerable interest in vector behaviour that is not influenced by application of core interventions (i.e., conventional IRS and Insecticide Treated Nets (ITNs)), such as feeding earlier and resting outdoors when humans are not protected. For example, an unprecedented malaria outbreak, related to illegal rosewood logging, occurred in 2014 with a seven-fold increase of cases in 1 year in Ubon Ratchathani Province, Northeastern Thailand [73]. Insecticide-susceptible and exophilic An. dirus s.l. were collected from a forested area in Ban Pakla and Chong Ta Ou Thai border control station, including An. maculatus s.l. collected remote villages with potentially low insecticide pressure [73]. These susceptible vector species are less amenable to control interventions due to their behaviour and their interactions with humans contribute to persistent residual transmission and represent barrier to success [74, 75].

From a geographical perspective, residual malaria parasite transmission has been reported across numerous transmission settings, even with good access and usage of LLINs or well-implemented IRS [76, 77, 78, 79, 80]. From the programmatic perspective, residual malaria transmission (RMT) is defined as the persistent malaria transmission that occurs once universal coverage of LLINs and maximal coverage of IRS have been achieved. Identification and elucidation of RMT requires the following pre-conditions: (a) comprehensive and up-to-date LLIN and IRS coverage data, where coverage is defined as 100% access and usage of ITN/LLIN or IRS [81]; (b) outdoor human activity or behaviour to allow identification of outdoor sites and “gaps” in protection, not only before sleeping time, but also for people that remain outdoors during the night. In many countries of the GMS, LLIN and IRS distribution data are sparse or not readily available. Where these data are available, it is often compiled at a relatively high administrative level, such as district or province. Malaria transmission at the community level can vary considerably within a small area and thus to investigate RMT at this level, LLIN and IRS coverage data by village are necessary. Furthermore, LLIN coverage figures quoted at the province or district level often do not match the actual situation at the community level, perhaps due to inequities in distribution, inaccurate population estimates, and calculation of procurement need, limited replacement of outdated and damaged LLIN; the outcome of which could lead to an underestimation of the magnitude of RMT.

As malaria is becoming more and more restricted to hard-to-reach population groups, alternative or adapted control strategies are required who are somehow marginalised, poor, on the fringes of the public health system, living in dwellings that are either very close to the forest or harbour people who are exposed to the forest through their occupation (e.g., development sites and seasonal labour areas) or mobility behaviour. As shown in Table 1, the risk of RMT in the malaria foci is spread over the entire night, from dusk-to-dawn, requiring a combination of complementary vector control measures, such as long-lasting insecticide hammock nets (LLIHN) that can be used during different periods of the night. However, the use of LLIHN, single LLIN/ITN or topical repellents in the field may not be acceptable due to cultural and linguistic barriers of ethnic minorities and MMPs for which specific acceptability studies should be conducted to guide the feasibility of these vector control tools.

District, province, country Ecotype % Access to LLIN % Use of LLIN Proportion of Anopheles bites or infective bites in relation to sleeping time Reference
Eastern region: Borkeo & O’Chum districts, Rattanakiri Province; Western region: Pailin & Pursat Provinces; Cambodia Forest plots & villages 68.4% (Ratanakiri) 70.7% (forest workers) After 22:00 h 71% [53, 102]
Forest plots 69.2% (Pailin); 81.8% (Pursat) 66.3% (forest workers) Before 22:00 h 29%
Ma Noi and Phuoc Binh communes, Ninh Thuan Province, south-central Vietnam Village NA 85% Before 22:00 h 45% (bites only) [83]
Way to the forest NA NA Before 19:00 h 13% (bites of An maculatus)
Forest plots NA 53% Before 21:00 h 64% (bites only)
Tha Song Yang, Tak Province, Thailand Villagea 78% 80% Before 21:00 h and 05:00 h
20.0% Suan Oi
33.7% Pha Man
37.6% farm huts
[82]
Hamletsb 100% 75–95%
Farm huts NA NA
Son Thai commune, Khan Hoa Province, central Vietnam Village 78% 95% Before 20:00 and 05:00 h
26% farm huts
37% forest
Farm huts NA 62.7%c
Forest NA 25%c

Table 1.

Overview of residual malaria parasite transmission (RMT) in various ecological settings in Greater Mekong Subregion.

Suan Oi village.


Pha Man & Komonae hamlets, Thailand [82].


Regular use of LLIN.


NA: not available.

Another driver of RMT is mega and micro-development projects impacting the forest or creating new conditions suitable for vector species, and often attracting a substantial workforce from various horizons across borders and cultural boundaries [6]. Their sleeping or residential places can have additional vulnerabilities if they are remote, comprising mainly ethnic minorities, or in conflict areas, all of which can hinder access to the public health system. Another key concern is to restrict or mitigate the widespread dispersal of parasites by these elusive population groups.

Just as forest workers often stay in the forests for several days and sleep exposed to vectors [84], rubber tappers also work in plantations at night with higher likelihoods of being bitten by Anopheles mosquitoes, in particular vector species of the Dirus Complex [3, 85]; they all have poor access to healthcare services [86]. Plantation work is seasonal, and manpower is often composed of highly mobile seasonal migrants, but little is known about their patterns of movement. More malaria infections were observed in people with temporary labour positions and plantation workers at the Thailand-Myanmar border [87], but this was not confirmed due to a very high proportion of the study participants opting to perform forest or field activities, and a very low number opting to work in rubber plantation [59]. Many migrants that arrive for rubber tapping settle beyond the harvest season [87] and go on to work on other cash crops (e.g., rice, cassava, fruit orchards). On return to their usual settlements, they contribute to the spread of malaria within and across international borders [41, 43]. By creating hot-spots of malaria and disproportionately affecting people with certain high-risk occupations [86, 89], residual transmission under these circumstances has so far hindered progress towards elimination.

Advertisement

5. Correct identification of malaria vectors and Plasmodium detection

High levels of malaria transmission occurring in forest-fringe areas of Southeast Asia is explained by movements of people in search of forest products and exposure to many highly efficient vector species that have adapted to forest ecotypes [66, 85, 102, 103]. The wide diversity of both the deep-forest (e.g., Leucosphyrus Group of mosquitoes), forest-fringe and deforested area main vectors (e.g., An. minimus, An. maculatus s.l., An. culicifacies s.l., An. fluviatilis s.l., An. letifer, An. donaldi), as well as their great potential to adapt to habitat changes, means that the consequences of deforestation on malaria transmission in Southeast Asia are difficult to predict and unlikely to be unidirectional [104]. Whilst An. dirus and An. baimaii, main vectors of the Dirus complex, can find tree-crop plantations suitable for breeding, a close association between malaria and rubber plantations has been demonstrated [4, 105, 106, 107, 108], contributing to high larval and pupal density during the rainy season [90, 91] and low numbers during the cool-dry season [92, 109], or provide conditions that are similar to this vector’s natural habitat [110]. This ecological adaptation in human settlements and shaded plantations contributes to outdoor transmission among rubber tappers.

The identification of secondary or incidental vector species poses new challenges as shown by mixed results of sporozoite-positivity using nested Polymerase chain reaction (PCR) and routine circumsporozoite enzyme-linked immunosorbent assay (CSP-ELISA) (Table 2). Confirmation of all positive CSP-ELISA results by a second CSP-ELISA test on the heated ELISA lysate, especially in zoophilic species showed a relatively high proportion of false positives (40%) [93]. On the other hand, PCR analysis of Deoxyribonucleic acid (DNA) extracted from the head and thorax alone, along with sequence data, revealed five Anopheles species (An. hyrcanus, An. barbirostris s.s., An. barbirostris clade III, An. nivipes, and An. peditaeniatus) infected with Plasmodium falciparum, which are not considered major vectors in the GMS [94]. Similarly, out of 11 P. falciparum CSP positive samples from Bangladesh, seven turned out to be positive by PCR suggesting that An. maculatus, An. jeyporiensis and An. nivipes play important roles in malaria transmission in Kuhalong District [95]. In Vietnam, the role of a secondary vector, An. pampanai infected with P. vivax, was also reported in the Binh Phuoc Province [96]. Morphological mis-identification of the closely related sympatric species, such as An. aconitus, An. pampanai and An. varuna are common [99, 100]. Morphological identification of Anopheles specimens prior to PCR assays allows them to be sort out at the group or complex level but does not permit species identification [85]. PCR assays must be applied for a reliable identification to the species level, which ensures that data received by malaria vector control programmes are suitable for targeting the correct vector species [101]. Given the low infection rates among many of these species especially in elimination phase, it is important for field entomologists to assess various species’ role in malaria transmission in the eco-epidemiological context. When changing objectives from control to elimination of malaria in Southeast Asia, the need to focus not only in the so-called main vector species, but also on secondary vectors is increasingly important.

Morphological Anopheles species* Nested PCR, Cambodia [93] Circumsporozoite ELISA, Thailand [97] Prior heating of eluate and circumsporozoite ELISA, Bangladesh [98] PCR confirmation of ELISA-positives Bangladesh [95]
Total collection (%) Positive/total Total collection (%) Positive /total Total collection (%) Positive /total Total collection (%) Positive /total
An. maculatus s.l 21.43 4/640 4.3 2/97
An. annularis s.l. 14.43 3/431 0.78 1/19
An. kochi 0.93 1/44
An. barbirostris s.s. 6.6 3/55 3.52 1/105 2.9 1/140 7.4 1/186
An. peditaeniatus 5.08 3/139
An. hyrcanus 0.09 2/2
An. nigerrimus 0.87 1/21 4.1 1/104
An. philippinensis 3/219 24.7 25/1169
An. vagus 41.9 25/1978
An. nivipes 10.8 1/264
An. jeyporiensis 3.1 1/142 18.9 2/479
An. karwari 5.16 11/244 1.7

Table 2.

Sporozoite infectivity rates of less known (secondary) vectors along the Bangladesh-Thailand-Cambodia corridor.

Molecular identification was specifically conducted on Anopheles barbirostris s.s. and An. barbirostris clade III; An. hyrcanus and An. hyrcanus s.s; An. peditaeniatus and An. nivipes, and morphological identification for the other Anopheles species.


Deforestation may deplete the populations of deep-forest vectors and so initially reduce malaria transmission; in some localities this depletion may be followed by the invasion of other efficient vector species resulting in increased transmission. With the exception of two longitudinal studies examining the effects of progressive land use changes from pre-development forest to oil palm cultivation on the distribution of disease vectors and malaria incidence [111], there is a striking lack of primary research directly measuring the impact of deforestation on malaria in Southeast Asia [104]. Recent studies showed that An. dirus s.l. was abundant in rubber plantations in Myanmar [109] and An. baimaii (molecularly identified) adults were caught from human landing collections in Wae Kha Mi, Mon State, the site of an acceptability study of permethrin-treated clothing [110]. In Lao PDR, a total of 46 An. dirus s.l. were collected, of which 31 were from immature rubber plantations, nine from mature rubber plantations, five from secondary forests and one from the rural village [105] (Tangena Julie-Ann, personal communication).

Advertisement

6. Plasmodium knowlesi: an additional challenge to malaria elimination

Plasmodium knowlesi, a simian malaria parasite, is now considered the 5th parasite affecting humans [112]. All countries in Southeast Asia have reported cases of P. knowlesi with the exception of Lao PDR and Timor Leste [113]. Since most countries are now working towards malaria elimination, it is pertinent to pay serious attention to malaria cases especially in areas where malaria has been reduced to very low levels. A good example is Sabah, Malaysian Borneo where large numbers of P. knowlesi were diagnosed in areas where P. falciparum and P. vivax were occurring in very low numbers [114]. Malaysia is working towards malaria elimination by 2020 and currently more than 60% of the malaria cases are due to P. knowlesi (MOH personal communication).

Recently, an increasing number of cases of P. knowlesi were reported from Kalimantan and Ache in Indonesia [115, 116] where malaria was in process of being eliminated. In Northern Sumatra, Indonesia where they are working towards malaria elimination, they recorded only 614 (16.5%) positive malaria cases by microscopy out of 3731 people examined [117]. However, PCR detected malaria parasites in 1169 (31.3%) individuals. Of these, 74.9% were mono-infection and 25.1% were multiple infection. P. falciparum constituted 24.8%, P. vivax 33.9%, P. malariae 9.3%, and P. knowlesi 32% [114] of the cases. It was also found that the primers developed from the SICAvar gene were more sensitive than the SSU rRNA gene [117]. It is obvious that parasite species are being mis-identified and many people who are asymptomatic are also missed by conventional microscopy [117, 118]. Thus, it is important to develop Rapid Diagnostic Tests (RDTs) that can be used by field workers to detect accurately malaria parasite species, especially P. knowlesi, and also additional laboratories should be established to conduct molecular assays for malaria diagnosis in the context of malaria elimination.

Deforestation and changes in the environment are the key factors leading to a surge of P. knowlesi malaria [119]. This parasite occurs in Macaca fascicularis (long-tailed) and Macaca nemestrina (pig-tailed) monkeys and its distribution is limited by some species of the Leucosphyrus Group of Anopheles mosquitoes [120]. These species are found biting in greater abundance in forest and farms compared to villages [121, 122]. However, in Sabah, Malaysian Borneo, it was found that An. balabacensis was abundant in villages as well [123], and sporozoite-positive specimens were reported in addition to farms and forest [123], while infective mosquitoes were found only in the forested sites and farms in Sarawak (Borneo) and Pahang (Peninsular), Malaysia [121, 122]. In addition, vector studies have also been conducted in Vietnam [124, 125] where the species An. dirus has been incriminated as the simian malaria vector in Khanh Phu—South Central Vietnam. Studies were conducted in the forest and forest-fringe areas near Nga Hai village where both human malaria parasites, P. falciparum and P. vivax, were found along with P. knowlesi in order to determine the potential role of An. dirus as bridge vectors of Plasmodium parasites from monkeys to humans [126]. Based on these studies, it was possible for An. dirus to pick up infection from humans and macaques during the mosquito’s lifespan. However, since there have been no reports of epidemics of P. knowlesi, it is believed that humans are infected by mosquitoes which acquired infection from the macaques. Perhaps even likely given that confirmed vectors of human plasmodia in Southeast Asia also become naturally infected by the monkey malaria species [127]. A recent case control study conducted in Sabah revealed that the age group >15, predominantly males, working in farms, plantations, forested areas, and with travel history, were independently associated with the risk of acquiring knowlesi malaria [128]. It also highlighted that IRS was associated with decrease of risk [128].

There are only few investigations on record in understanding bionomics of vectors transmitting P. knowlesi malaria. In order to implement vector control activities, the bionomics of the vectors must be understood. Based on few studies, it has been shown that the vectors are biting in the early part of the night from 18:00 h to 21:00 h and mostly outdoors [121, 122, 123, 129]. In these rural areas, people go to bed by 22:00 h and they are up by 05:00 h. The results showed that only 39.79% of An. balabacensis [123], 43.8% of An. latens [121] and 12.8% of An. cracens [122] were found biting during this sleeping time. Thus, current vector control measures like IRS and ITNs are not appropriate for the exophagic and exophilic vectors. The forests in Southeast Asia is providing a favorable environment with high percentage of macaques being positive for P. knowlesi [130, 131, 132], and with the presence of the vectors, it is going to be a daunting task to eliminate malaria. On a global scale, malaria has been reduced to low levels due to the scaling up of ITNs, IRS, ACTs, and intermittent preventive treatment to infants and pregnant women [133]. Thus, it is obvious that new tools are urgently required for successful malaria elimination.

It is known that the two human malaria species (P. falciparum and P. vivax), which infects millions of people actually were of zoonotic origin (from the African apes), which evolved thousands of years ago [134, 135]. Thus, there is always a possibility that in the future P. knowlesi and other simian malarias may become established in humans, especially when human malaria is eliminated. However, currently human-to-human transmission of knowlesi malaria by mosquitoes has not been established. This is crucial in the light of malaria elimination and more focused research is needed on this topic if we are to succeed with malaria elimination.

Changing landscape affects Anopheles distribution, mosquito density and diversity in Malaysia, and more globally Southeast Asia [105, 111, 136, 137, 138]. It has been shown that with loss of forest cover, cases of P. knowlesi have increased in Sabah [119]. Land use change has also led to increase of malaria cases due to various factors such as increase of macaques in small forest patches along with the colonization of the main vectors [119, 136]. It is interesting to note that An. balabacensis, the predominant vector of human and simian malaria, was found in great abundance in logged forest, followed by thinly logged virgin jungle reserve and was lowest in primary forest [136]. This vector was also found to be biting humans more at ground level compared to canopy level [136]. It is therefore important to include both the public health and agro-forestry sectors in controlling malaria vectors in the country. Studies from Thailand also indicate that if landscape management should be used for malaria control in northern Thailand, large-scale reduction and fragmentation of forest cover would be needed [139, 140]. Such drastic actions, however, do not align with current global objectives concerning forest and biodiversity conservation.

The vectors of simian malaria described to date were An. hackeri (Leucosphyrus Group) [141] recorded biting mainly the macaques and large numbers were collected resting on Nipah palm trees in Selangor in 1960s; An. cracens (Dirus Complex) [122] biting both macaques and humans and found mainly in the forest and farms; An. latens (Leucosphyrus Complex) [121] was the predominant mosquito in the forest compared to farm and village, and was biting macaques at ground level and at six meters in the canopy compared to three meters. The biting ratio of monkey versus human for An. latens was 1:1.3 [121]. An. introlatus (Leucosphyrus Complex) [142]was biting in the early part of the night from 19:00 h to 21:00 h and was the predominant mosquito in Hulu Selangor where cases of P. knowlesi were reported. Most recently, An. balabacensis (Leucosphyrus Complex) has been incriminated as vector of P. knowlesi in Sabah [123], as well as human malaria and Bancroftian lymphatic filariasis due to Wuchereria bancrofti [143, 144, 145].

Although an increased number of countries are successfully eliminating human malaria in recent years, no country has yet eliminated non-human malaria, which adds another layer of complexity to be addressed. The complex situation of malaria is Southeast Asia is very unique from the rest of the tropical countries. More effort is needed to study the host switching mechanisms between the parasites in humans, macaques and vectors. A series of review papers have been published over the years and all these have indicated the importance of addressing the problem caused by P. knowlesi, if malaria elimination is to be successful in the region [113, 146, 147, 148, 149, 150, 151].

Advertisement

7. Targeting vivax malaria: a bottleneck to malaria elimination

As opposed to P. falciparum infection, which does not have latency (dormant), P. vivax causes two distinct infection syndromes, one that actively proliferates and the other latent due to hypnozoites. Each of these P. vivax forms requires distinct therapeutic treatments and the latent form cannot be diagnosed [152]. Most acute attacks of P. vivax in endemic areas originate from hypnozoites, and unless that reservoir is aggressively attacked, elimination of transmission may be an unrealistic goal.

Treatment of latent vivax represents an important challenge as the only known therapies are 8-aminoquinoline drugs, which results in acute hemolytic anemia in patients deficient in glucose-6-phosphate dehydrogenase (G6PD)—a highly polymorphic inherited disorder affecting 1–30% of residents of malaria-endemic nations [153]. The single low dose of primaquine against gametocytes of P. falciparum does not threaten the G6PD deficient subjects [154]. Another challenge is that the parasitemia of vivax malaria patients is typically an order of magnitude lower than falciparum malaria, causing larger proportions of parasitemia to fall below diagnostics detection thresholds [152]. In addition, vivax malaria patients may exhibit very low parasitemia, and yet become severely ill. These fundamental distinctions between the two dominant human malarias explain why P. vivax is relatively unaffected by interventions tailored to control P. falciparum calling for new strategies needed for combatting vivax malaria [155].

In addition, P. vivax has the ability to develop at lower temperature than P. falciparum and has a shorter sporogonic cycle in the vector, which results in P. vivax extending beyond tropical climates into temperate regions. This ability, combined with its early-biting, outdoor-feeding and outdoor-resting behavior of vector mosquito species, also makes them less susceptible to vector control measures such as IRS, which have proven effective against transmission of P. falciparum [156]. Also having dormant forms in the liver (hypnozoites) mean that one successful infection will generate a number of parasitological and clinical episodes without reinfection. Therefore, recurrent cases cannot be prevented via vector control, though, paradoxically, successful transmission control of vivax malaria could reduce the disease burden more than that of P. falciparum, because avoiding one infection will result in preventing a number of clinical episodes over several years [155].

Vivax malaria is diagnosed late, because infected people get ill with low parasite densities, which cannot be detected with current diagnostics, such as RDTs and microscopy. Delayed diagnosis means not only delayed treatment (hence prolonged morbidity, especially anemia) but also ability to transmit over an extended period. This is further amplified by the fact that mature gametocytes appear simultaneously with asexual forms—hence transmission occurs before diagnosis and treatment [157, 158].

As recently described [156], an effective P. vivax control and elimination toolbox should include:

  1. Practical point-of-care G6PD deficiency diagnostics allowing wider access to safe primaquine therapy or with tafenoquine—a related single dose hypnozoitocide recently developed by GSK and Medicines for Malaria Venture (MMV); the latter has been submitted to the United States Food and Drug Administration (FDA) seeking approval of single-dose tafenoquine for the radical cure (prevention of relapse) of vivax malaria in patients 16 years of age and older [159];

  2. More sensitive point-of-care diagnostics for detecting intrinsically lower parasitemia, including sub-patent and asymptomatic infections;

  3. Validated strategies for relapse prevention in special population groups, i.e., pregnant women, young infants, G6PD deficient and G6PD unknowns in which 8-aminoquinoline is contraindicated;

  4. Clinical care algorithms acknowledging risk of severe and threatening syndromes despite seemingly non-threatening levels of parasitemia; and

  5. Interventions of proven efficacy to minimize human contact with often zoophilic, exophagic and exophilic Anopheles species of great diversity.

In conclusion, the malaria community needs to address these challenges and create a viable strategy to achieve vivax elimination goals, providing novel solutions for overcoming critical bottlenecks. This process needs to begin now to enhance treatment practice for 8-aminoquinoline drugs based radical cure. Highlighting the benefits of radical cure for the patient and community will improve prescription practice and patient adherence [160]. Coupling this with improved access to adequate G6PD testing will pave the way for the introduction of tafenoquine, with huge potential to accelerate the elimination of P. vivax.

Advertisement

8. Socio-ecological and adaptive management of malaria ecosystem in areas approaching malaria elimination

WHO has recently proposed sustainable prevention and control of diseases emerging within complex, dynamic, adaptive systems, such as malaria, based on interdisciplinary and approaches addressing environmental and social health determinants holistically [161]. More insights into transmission dynamics and the possibility of intersectoral ecosystem management programs for malaria elimination and control are urgently needed. An ecosystem approach to successful reduction of vector-borne disease burden [162, 163] can lead to considerable health gains [Available at: http://www.maweb.org/documents/document.317.aspx.pdf].

Once local entomological inoculation rates (EIRs) have been reduced to a level of unstable transmission the infectious reservoir can be eliminated via several approaches without a threat of malaria re-emergence from reintroduction of parasites. At this point, use of time-limited mass drug administrations (MDA) campaign at high coverage should be sufficient to effectively clear the majority of remaining P. falciparum cases, and may be considered for epidemic control as part of the initial response, along with the urgent introduction of other interventions [164]. This can be supplemented by screening and treatment programmes based on WHO Global Malaria Programme’s T3: Test, Treat, Track initiative supporting malaria-endemic countries in their efforts to achieve universal coverage with diagnostic testing and antimalarial treatment, as well as in strengthening their malaria surveillance system [WHO T3: Test, Treat, Track. Scaling up diagnostic testing, treatment and surveillance for malaria. World Health Organisation; 2012. http://www.who.int/malaria/publications/atoz/t3_brochure/en/]. Healthcare workers or locally trained and supervised community volunteer networks can apply this method to effectively limit reintroduction of parasites from other areas to a minimum, and apply additional active case management, e.g., the systematic detection and treatment of parasitemia using highly sensitive RDTs can reduce the risk attributed to any unscreened or asymptomatic cases.

Depending on the local situations, supplementary measures, in addition to LLINs or IRS, such as repellents or treated clothing for high-risk individuals, offer special precautionary preventive protection [1, 110, 165, 166]. Passive case management should suffice for treating any symptomatic infections as they may occur. This, however, assumes at least a periodic provision of health services at all locations, including remote ones. A transdisciplinary approach integrates different scientific perspectives [167, 168] and provides a formal platform for stakeholder participation in the research and development of new information, ideas and strategies, their testing and eventual application.

Participatory approaches that engage local communities in a complex social-ecological mapping process are a vital starting point for identifying community-applicable solutions and leveraging community capacity for local interventions [169, 170] and promoting integrative and equitable collaboration within partnership of researchers and communities [170, 171]. Ownership of continuous surveillance, monitoring, treatment and preventive efforts should be transferred to members of local communities, assuming collective responsibility for their continuous well-being.

Advertisement

9. Conclusions

This review attempts to consolidate the challenges of operational research for innovations in designing interventions [172], according to the current situation and progress made, for achieving malaria elimination in Southeast Asia. As the entry of artemisinin resistant parasites to India could be the first step in their spread to Africa, the current priority must be to address this problem in Southeast Asia before it can become a threat in Africa. Continuous monitoring of drug resistance in conjunction with analysis and proper interpretation is critical to guide the appropriate action for effective treatment. While P. falciparum elimination in the GMS is realistic, feasible and particularly urgent in the context of drug resistance, the main challenges are to ensure community participation and plan for the preservation of ACT potency so that the dosing regimens and surveillance for resistance are rigorously pursued to sustain their efficacy for as long as possible [172].

We support a priority focus on MMP and other high-risk groups to contain the spread of artemisinin resistance and new hot-spots, however, implementation challenges should be considered when planning future interventions. More efforts are needed in documenting the malaria risk among different types of MMPs, innovative tools and interventions, as well as designing implementation in a way that can be evaluated, lessons learned, and programmes adapted in an on-going process [172]. New ways of evaluating MMP interventions (including highly sensitive RDTs) are needed, as routine health information systems have limitations and might not allow capturing the information and data needed, and existing type of surveys might not be sufficient for monitoring interventions for MMP.

Malaria programmes need to heed the recent revised WHO recommendations for achieving universal coverage with LLINs or IRS for populations at risk [173]. The coverage of key interventions is critically low in some countries and sub-optimal in most others, threatening progress across the region as a whole [174]. Malaria programmes are encouraged to evaluate the magnitude (and drivers) of the residual transmission in their country, regarding both mosquito and human behavior. This information will provide a boost for industry and academic partners to develop new vector control methods and paradigms for outdoor and residual transmission.

The current precarious funding situation could undermine elimination efforts and result in a resurgence of disease. The threat posed to regional and global malaria control and elimination efforts by artemisinin resistant P. falciparum parasites is imminent and potentially severe. In many Asian countries, operational feasibility of P. vivax elimination is lower than that for P. falciparum [27]. Therefore, creating a viable strategy to achieve vivax elimination goals should include improvements in access to safe treatment to 8-aminoquinoline drugs based radical cure together with improved access to adequate G6PD testing in P. vivax endemic countries.

Whilst human P. knowlesi is still largely a zoonosis, all indications suggest that human-to human transmission can take place, and probably is taking place in some situations [175]. More research is required to substantiate the body of evidence for human-to-human transmission, laboratory diagnosis and clinical management, and mapping vectors of P. knowlesi and environmental risk factors.

The challenge for elimination programmes is dealing with dynamic, social-ecological systems for which an entirely different kind of thinking and scientific framework is required. The retooling for this next phase is more challenging this time since it requires malaria experts and managers to understand complex systems, thinking and practices. This thinking and actions are more or less contrary to conventional understandings of disease control, which tend to be top down and not guided by concepts like resilience and adaptive management developed as part of so-called ecosystem approach/management.

Advertisement

Acknowledgments

The authors would like to specifically thank Sergio Lopes and Allan Schapira for their comments and inputs to the manuscript. We thank Irwin Chavez for supporting EndNote citations and references. The authors thank WHO for approval to reproduce Figures 1 and 2 taken from [2]. Data for Thailand and Vietnam in Table 1 was supported by the Special Programme for Research and Training in Tropical Diseases (TDR), in partnership with the Global Malaria Programme of the World Health Organisation (WHO) to develop multi-country, multi-institutional research on Investigation of the magnitude and causes of residual malaria parasite transmission in selected settings.

Conflict of interest

All authors declare no conflict of interest related to the writing of this chapter.

List of acronyms

ACTArtemisinin-based combination therapy
AMTArtemisinin monotherapy
ASEANAssociation of Southeast Asian nations
BCCBehavioral change communication
DNADeoxyribonucleic acid
CSPCircumsporozoite
EIREntomological inoculation rate
ELISAEnzyme-linked immunosorbent assay
ERAREmergency response to artemisinin resistance
FDAFood and drug administration
G6PDGlucose-6-phosphate dehydrogenase
GMSGreater Mekong Subregion
IRSIndoor residual spraying
ITNInsecticide treated nets
LLINLong lasting insecticidal nets
LLIHNLong lasting insecticidal hammock nets
LSMLarval source management
MDAMass drug administration
MMPMobile and migrant populations
MMVMedicines for malaria venture
MOHMinistry of Health
PCRPolymerase chain reaction
PMIPresident Malaria Initiative
RDTRapid diagnostic test
RMTResidual malaria transmission
SESSocio-ecological system
SPSulfadoxine-pyrimethamine
WHOWorld Health Organisation

References

  1. 1. WHO. Strategy for Malaria Elimination in the Greater Mekong Subregion: 2015-2030. Manila: WHO Regional Office for the Western Pacific; 2015. 80 p
  2. 2. WHO. Countries of the Greater Mekong making inroads in the race towards elimination. 2017. Available from: http://apps.who.int/iris/bitstream/10665/259602/1/WHO-HTM-GMP-2017.23-eng.pdf [Accessed: 2017-10-05]
  3. 3. Suwonkerd W, Ritthison W, Ngo CT, Tainchum K, Bangs MJ, Chareonviriyaphap T. Vector biology and malaria transmission in Southeast Asia. In: Manguin S, editor. Anopheles Mosquitoes-New Insights into Malaria Vectors. Rijeka, Croatia: InTech Open Access; 2013. DOI: 10.5772/56347
  4. 4. Bhumiratana A, Sorosjinda-Nunthawarasilp P, Kaewwaen W, Maneekan P, Pimnon S. Malaria-associated rubber plantations in Thailand. Travel Medicine and Infectious Disease. 2013;11:37-50. DOI: 10.1016/j.tmaid.2012.11.002
  5. 5. WHO. Informal Consultation on Malaria and Mobile and Migrant Populations: Addressing Priority Gaps in the Context of Malaria Elimination in the GMS. Bangkok, Thailand: World Health Organization; 2016. 54 p
  6. 6. Jitthai N. Migration and malaria. The Southeast Asian Journal of Tropical Medicine and Public Health. 2013;44:166-200 discussion 306-167
  7. 7. O’Shannassy T. Towards malaria elimination in the greater Mekong subregion. The Southeast Asian Journal of Tropical Medicine and Public Health. 2013;44(Supp):45
  8. 8. White NJ. Antimalarial drug resistance. The Journal of Clinical Investigation. 2004;113:1084. DOI: 10.1172/JCI200421682
  9. 9. Nosten F, White NJ. Artemisinin-based combination treatment of falciparum malaria. American Journal of Tropical Medicine and Hygiene. 2007;77(6_Suppl):181-192. DOI: 0.4269/ajtmh.2007.77.181
  10. 10. Status report on artemisinin and ACT resistance. 2017. Available from: http://www.who.int/malaria/publications/atoz/artemisinin-resistance-april2017/en/ [Accessed: 2017-11-7]
  11. 11. Imwong M, Hien TT, Thuy-Nhien NT, Dondorp AM, White NJ. Spread of a single multidrug resistant malaria parasite lineage (PfPailin) to Vietnam. The Lancet Infectious Diseases. 2017;17:1022-1023. DOI: 10.1016/S1473-3099(17)30524-8
  12. 12. Noedl H, Se Y, Schaecher K, Smith BL, Socheat D, Fukuda MM. Evidence of artemisinin-resistant malaria in western Cambodia. New England Journal of Medicine. 2008;359:2619-2620. DOI: 10.1056/NEJMc0805011
  13. 13. Dondorp AM, Nosten F, Yi P, Das D, Phyo AP, Tarning J, Lwin KM, Ariey F, Hanpithakpong W, Lee SJ. Artemisinin resistance in Plasmodium falciparum malaria. New England Journal of Medicine. 2009;361:455-467. DOI: 10.1056/NEJMoa0808859
  14. 14. Amaratunga C, Sreng S, Suon S, Phelps ES, Stepniewska K, Lim P, Zhou C, Mao S, Anderson JM, Lindegardh N. Artemisinin-resistant Plasmodium falciparum in Pursat province, western Cambodia: A parasite clearance rate study. The Lancet Infectious Diseases. 2012;12:851-858. DOI: 10.1016/S1473-3099(12)70181-0
  15. 15. Hien TT, Thuy-Nhien NT, Phu NH, Boni MF, Thanh NV, Nha-Ca NT, Thai CQ, Van Toi P, Thuan PD, Merson L. In vivo susceptibility of Plasmodium falciparum to artesunate in Binh Phuoc Province, Vietnam. Malaria Journal. 2012;11:355. DOI: 10.1186/1475-2875-11-355
  16. 16. Phyo AP, Nkhoma S, Stepniewska K, Ashley EA, Nair S, McGready R, ler Moo C, Al-Saai S, Dondorp AM, Lwin KM. Emergence of artemisinin-resistant malaria on the western border of Thailand: A longitudinal study. Lancet. 2012;379:1960-1966. DOI: 10.1016/S0140-6736(12)60484-X
  17. 17. Ashley EA, Dhorda M, Fairhurst RM, Amaratunga C, Lim P, Suon S, Sreng S, Anderson JM, Mao S, Sam B. Spread of artemisinin resistance in Plasmodium falciparum malaria. New England Journal of Medicine. 2014;371:411-423. DOI: 10.1056/NEJMoa1314981
  18. 18. Huang F, Takala-Harrison S, Jacob CG, Liu H, Sun X, Yang H, Nyunt MM, Adams M, Zhou S, Xia Z. A single mutation in K13 predominates in southern China and is associated with delayed clearance of Plasmodium falciparum following artemisinin treatment. The Journal of Infectious Diseases. 2015;212:1629-1635. DOI: 10.1093/infdis/jiv249
  19. 19. Menard D, Dondorp A. Antimalarial drug resistance: A threat to malaria elimination. Cold Spring Harbor Perspectives in Medicine. 2017;7:a025619. DOI: 10.1101/cshperspect.a025619
  20. 20. Leang R, Barrette A, Bouth DM, Menard D, Abdur R, Duong S, Ringwald P. Efficacy of dihydroartemisinin-piperaquine for treatment of uncomplicated Plasmodium falciparum and Plasmodium vivax in Cambodia, 2008 to 2010. Antimicrobial Agents and Chemotherapy. 2013;57:818-826. DOI: 10.1128/AAC.00686-12
  21. 21. Leang R, Taylor WR, Bouth DM, Song L, Tarning J, Char MC, Kim S, Witkowski B, Duru V, Domergue A. Evidence of Plasmodium falciparum malaria multidrug resistance to artemisinin and piperaquine in western Cambodia: Dihydroartemisinin-piperaquine open-label multicenter clinical assessment. Antimicrobial Agents and Chemotherapy. 2015;59:4719-4726. DOI: 10.1128/AAC.00835-15
  22. 22. Lon C, Manning JE, Vanachayangkul P, So M, Sea D, Se Y, Gosi P, Lanteri C, Chaorattanakawee S, Sriwichai S. Efficacy of two versus three-day regimens of dihydroartemisinin-piperaquine for uncomplicated malaria in military personnel in northern Cambodia: An open-label randomized trial. PLoS One. 2014;9:e93138. DOI: 10.1371/journal.pone.0093138
  23. 23. Saunders DL, Vanachayangkul P, Lon C. Dihydroartemisinin–piperaquine failure in Cambodia. New England Journal of Medicine. 2014;371:484-485. DOI: 10.1056/NEJMc1403007
  24. 24. Duru V, Khim N, Leang R, Kim S, Domergue A, Kloeung N, Ke S, Chy S, Eam R, Khean C. Plasmodium falciparum dihydroartemisinin-piperaquine failures in Cambodia are associated with mutant K13 parasites presenting high survival rates in novel piperaquine in vitro assays: Retrospective and prospective investigations. BMC Medicine. 2015;13:305. DOI: 10.1186/s12916-015-0539-5
  25. 25. Spring MD, Lin JT, Manning JE, Vanachayangkul P, Somethy S, Bun R, Se Y, Chann S, Ittiverakul M, Sia-ngam P. Dihydroartemisinin-piperaquine failure associated with a triple mutant including kelch13 C580Y in Cambodia: An observational cohort study. The Lancet Infectious Diseases. 2015;15:683-691. DOI: 10.1016/S1473-3099(15)70049-6
  26. 26. Carrara VI, Lwin KM, Phyo AP, Ashley E, Wiladphaingern J, Sriprawat K, Rijken M, Boel M, McGready R, Proux S. Malaria burden and artemisinin resistance in the mobile and migrant population on the Thai–Myanmar border, 1999-2011: An observational study. PLoS Medicine. 2013;10:e1001398. DOI: 10.1371/journal.pmed.1001398
  27. 27. Thein ST, Sudhinaraset M, Khin HSS, McFarland W, Aung T. Who continues to stock oral artemisinin monotherapy? Results of a provider survey in Myanmar. Malaria Journal. 2016;15:334. DOI: 10.1186/s12936-016-1392-5
  28. 28. ACTwatch G, Novotny J, Singh A, Dysoley L, Sovannaroth S, Rekol H. Evidence of successful malaria case management policy implementation in Cambodia: Results from national ACTwatch outlet surveys. Malaria Journal. 2016;15:194. DOI: 10.1186/s12936-016-1200-2
  29. 29. Ménard D, Khim N, Beghain J, Adegnika AA, Shafiul-Alam M, Amodu O, Rahim-Awab G, Barnadas C, Berry A, Boum Y. A worldwide map of Plasmodium falciparum K13-propeller polymorphisms. New England Journal of Medicine. 2016;374:2453-2464. DOI: 10.1056/NEJMoa1513137
  30. 30. Newton PN, Dondorp A, Green M, Mayxay M, White NJ. Counterfeit artesunate antimalarials in Southeast Asia. Lancet. 2003;362:169. DOI: 10.1016/S0140-6736(03)13872-X
  31. 31. Lamy M, Liverani M. Tackling substandard and falsified medicines in the Mekong: National responses and regional prospects. Asia Pacific Policy Studies. 2015;2:245-254. DOI: 10.1002/app5.87
  32. 32. Krishna S, Pulcini S, Fatih F, Staines H. Artemisinins and the biological basis for the PfATP6/SERCA hypothesis. Trends in Parasitology. 2010;26:517-523. DOI: 10.1016/j.pt.2010.06.014
  33. 33. Laurent BS, Miller B, Burton TA, Amaratunga C, Men S, Sovannaroth S, Fay MP, Miotto O, Gwadz RW, Anderson JM. Artemisinin-resistant Plasmodium falciparum clinical isolates can infect diverse mosquito vectors of Southeast Asia and Africa. Nature Communications. 2015;6:8614. DOI: 10.1038/ncomms9614
  34. 34. Fairhurst R, Dondorp A. Artemisinin-resistant Plasmodium falciparum malaria. Microbiology Spectrum. 2016;4(3). DOI: 10.1128/microbiolspec.EI10-0013-2016
  35. 35. WHO: Guidelines for the treatment of malaria—2nd edition. In: WHO Guidelines for the Treatment of Malaria. 2nd ed. Geneva, Switzerland: World Health Organisation; 2010.
  36. 36. Dondorp A, Nosten F, Stepniewska K, Day N, White N. Artesunate versus quinine for treatment of severe falciparum malaria: A randomised trial. Lancet. 2005;366:717-725. DOI: 10.1016/S0140-6736(05)67176-0
  37. 37. Dondorp AM, Fanello CI, Hendriksen IC, Gomes E, Seni A, Chhaganlal KD, Bojang K, Olaosebikan R, Anunobi N, Maitland K. Artesunate versus quinine in the treatment of severe falciparum malaria in African children (AQUAMAT): An open-label, randomised trial. Lancet. 2010;376:1647-1657. DOI: 10.1016/S0140-6736(10)61924-1
  38. 38. WHO. Updates on Drug Resistance. 2017. Available from: http://www.who.int/malaria/areas/drug_resistance/updates/en/ [Accessed: 2017-11-7]
  39. 39. WHO. Efficacy Database. 2017. Available from: http://www.who.int/malaria/areas/drug_resistance/drug_efficacy_database/en/ [Accessed: 2017-11-7]
  40. 40. Prothero RM. Population movements and problems of malaria eradication in Africa. Bulletin of the World Health Organization. 1961;24:405
  41. 41. Martens P, Hall L. Malaria on the move: Human population movement and malaria transmission. Emerging Infectious Diseases. 2000;6:103. DOI: 10.3201/eid0602.000202
  42. 42. Prothero RM. Disease and mobility: A neglected factor in epidemiology. International Journal of Epidemiology. 1977;6:259-267. DOI: 10.1093/ije/6.3.259
  43. 43. Pindolia DK, Garcia AJ, Wesolowski A, Smith DL, Buckee CO, Noor AM, Snow RW, Tatem AJ. Human movement data for malaria control and elimination strategic planning. Malaria Journal. 2012;11:205. DOI: 10.1186/1475-2875-11-205
  44. 44. Delacollette C, D'Souza C, Christophel E, Thimasarn K, Abdur R, Bell D, Dai TC, Gopinath D, Lu S, Mendoza R. Malaria trends and challenges in the Greater Mekong Subregion. The Southeast Asian Journal of Tropical Medicine and Public Health. 2009;40:674
  45. 45. Edwards HM, Canavati SE, Rang C, Ly P, Sovannaroth S, Canier L, Khim N, Menard D, Ashton RA, Meek SR. Novel cross-border approaches to optimise identification of asymptomatic and artemisinin-resistant Plasmodium infection in mobile populations crossing Cambodian borders. PLoS One. 2015;10:e0124300. DOI: 10.1371/journal.pone.0124300
  46. 46. Bhumiratana A, Intarapuk A, Sorosjinda-Nunthawarasilp P, Maneekan P, Koyadun S. Border malaria associated with multidrug resistance on Thailand-Myanmar and Thailand-Cambodia borders: Transmission dynamic, vulnerability, and surveillance. BioMed Research International. 2013;2013:363417. DOI: 10.1155/2013/363417
  47. 47. Lynch C, Roper C. The transit phase of migration: Circulation of malaria and its multidrug-resistant forms in Africa. PLoS Medicine. 2011;8:e1001040. DOI: 10.1371/journal.pmed.1001040
  48. 48. ASEAN. Blueprint. 2008. Available from: https://archive.org/…/ASEANeconomiccommunityblueprint/ASEAN%20E [Accessed: 2017-11-9]
  49. 49. WHO. Informal consultation on improving access to malaria control services for migrant/mobile populations in the context of the emergency response to artemisinin resistance in the GMS. 2014. Available from: http://www.searo.who.int/entity/malaria/documents/access_malaria_services…GMS/en/ [Accessed: 2017-11-12]
  50. 50. Ministry of Health. Cambodia malaria survey. 2004. Available from: http://malariasurveys.org/documents/CMBS2004report15jul.pdf 2005 [Accessed: 2017-11-12]
  51. 51. Ministry of Health. Cambodia malaria survey. 2007. Available from: http://malariasveys.org/documents/CMS2007%20report_22Sep10_FINAL.pdf. 2008 [Accessed: 2017-11-12]
  52. 52. Ministry of Health. Cambodia malaria survey. 2010. Available from: http://malariasurveys.org/documents/CMS%202010%20GF%20Report%20(FINAL).pdf. 2011 [Accessed: 2017-11-12]
  53. 53. Ministry of Health. Cambodia malaria survey. 2013 Available at: http://www.malariaconsortium.org/resources/publications/624/cambodia-malaria-survey-2013. 2014. [Accessed 2017-11-12]
  54. 54. Okell LC, Ghani AC, Lyons E, Drakeley CJ. Submicroscopic infection in Plasmodium falciparum-endemic populations: A systematic review and meta-analysis. The Journal of Infectious Diseases. 2009;200:1509-1517. DOI: 10.1086/644781
  55. 55. Bousema T, Griffin JT, Sauerwein RW, Smith DL, Churcher TS, Takken W, Ghani A, Drakeley C, Gosling R. Hitting hotspots: Spatial targeting of malaria for control and elimination. PLoS Medicine. 2012;9:e1001165. DOI: doi.org/10.1371/journal.pmed.1001165
  56. 56. Mosha JF, Sturrock HJ, Greenwood B, Sutherland CJ, Gadalla NB, Atwal S, Hemelaar S, Brown JM, Drakeley C, Kibiki G. Hot spot or not: A comparison of spatial statistical methods to predict prospective malaria infections. Malaria Journal. 2014;13:53. DOI: 10.1186/1475-2875-13-53
  57. 57. Bousema T, Stresman G, Baidjoe AY, Bradley J, Knight P, Stone W, Osoti V, Makori E, Owaga C, Odongo W. The impact of hotspot-targeted interventions on malaria transmission in Rachuonyo South District in the Western Kenyan Highlands: A cluster-randomized controlled trial. PLoS Medicine. 2016;13:e1001993. DOI: /10.1371/journal.pmed.1001993
  58. 58. Rosenberg R, Andre RG, Somchit L. Highly efficient dry season transmission of malaria in Thailand. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1990;84:22-28
  59. 59. Durnez L, Pareyn M, Mean V, Kim S, Khim N, Menard D, Coosemans M, Sochantha T, Sluydts V. Identification and characterization of areas of high and low risk for asymptomatic malaria infections at sub-village level in Ratanakiri, Cambodia. Malaria Journal. 2018;17:27. DOI: 10.1186/s12936-017-2169-1
  60. 60. Sturrock HJ, Hsiang MS, Cohen JM, Smith DL, Greenhouse B, Bousema T, Gosling RD. Targeting asymptomatic malaria infections: Active surveillance in control and elimination. PLoS Medicine. 2013;10:e1001467. DOI: 10.1371/journal.pmed.1001467
  61. 61. Moonen B, Cohen JM, Snow RW, Slutsker L, Drakeley C, Smith DL, Abeyasinghe RR, Rodriguez MH, Maharaj R, Tanner M. Operational strategies to achieve and maintain malaria elimination. The Lancet. 2010;376:1592-1603. DOI: 10.1016/S0140-6736(10)61269-X
  62. 62. Lindblade KA, Steinhardt L, Samuels A, Kachur SP, Slutsker L. The silent threat: Asymptomatic parasitemia and malaria transmission. Expert Review of Anti-Infective Therapy. 2013;11:623-639. DOI: 10.1586/eri.13.45
  63. 63. Harris I, Sharrock WW, Bain LM, Gray K-A, Bobogare A, Boaz L, Lilley K, Krause D, Vallely A, Johnson M-L. A large proportion of asymptomatic Plasmodium infections with low and sub-microscopic parasite densities in the low transmission setting of Temotu Province, Solomon Islands: Challenges for malaria diagnostics in an elimination setting. Malaria Journal. 2010;9:254. DOI: 10.1186/1475-2875-9-254
  64. 64. CMEP: Response to Abnormal Increase of Malaria Cases in Pursat within Cambodia Malaria Elimination Project (CMEP). Cambodia: President’s Malaria Initiative. University Research Co., LLC; 2017
  65. 65. WPRO WHO. ERAR Dissemination Meeting to Share Findings of the Assessment of Malaria Surveillance Systems in Greater Mekong Subregion Countries. 2015. Available from: http://www.who.int/malaria/areas/greater_mekong/technical-reports/en/ [Accessed: 2017-11-12]
  66. 66. Grietens KP, Gryseels C, Dierickx S, Bannister-Tyrrell M, Trienekens S, Uk S, Phoeuk P, Suon S, Set S, Gerrets R. Characterizing types of human mobility to inform differential and targeted malaria elimination strategies in Northeast Cambodia. Scientific Reports. 2015;5:16836. DOI: 10.1038/srep16837
  67. 67. Guyant P, Canavati SE, Chea N, Ly P, Whittaker MA, Roca-Feltrer A, Yeung S. Malaria and the mobile and migrant population in Cambodia: A population movement framework to inform strategies for malaria control and elimination. Malaria Journal. 2015;14:252. DOI: 10.1186/s12936-015-0773-5
  68. 68. WHO. World malaria report 2012. Global Malaria Program. 2013. Available from: http://www.who.int/malaria/publications/world_malaria_report_2013/en [Accessed: 2107-11-12]
  69. 69. Yang G-J, Brook BW, Whelan PI, Cleland S, Bradshaw CJ. Endogenous and exogenous factors controlling temporal abundance patterns of tropical mosquitoes. Ecological Applications. 2008;18:2028-2040. DOI: 10.1890/07-1209.1
  70. 70. Sinka ME, Bangs MJ, Manguin S, Chareonviriyaphap T, Patil AP, Temperley WH, Gething PW, Elyazar IR, Kabaria CW, Harbach RE. The dominant Anopheles vectors of human malaria in the Asia-Pacific region: Occurrence data, distribution maps and bionomic précis. Parasites & Vectors. 2011;4:89. DOI: 10.1186/1756-3305-4-89
  71. 71. WHO. World Malaria Report 2013: World Health Organization; Geneva, Switzerland; 2014
  72. 72. APMEN. Report of the fourth annual business and technical meeting of the Asia-Pacific Malaria Elimination Network (APMEN). 2012. Available from: http://apmen.org/events/annual-meeting/4/apmen-iv.html/ [Accessed: 2017-12-5]
  73. 73. Sumarnrote A, Overgaard HJ, Marasri N, Fustec B, Thanispong K, Chareonviriyaphap T,Corbel V. Status of insecticide resistance in Anopheles mosquitoes in Ubon Ratchathani province, Northeastern Thailand. Malaria Journal. 2017;16:299. DOI: 10.1186/s12936-017-1948
  74. 74. Lyttleton C. Deviance and resistance: Malaria elimination in the greater Mekong subregion. Social Science & Medicine. 2016;150:144-152. DOI: 1016
  75. 75. Ferguson HM, Dornhaus A, Beeche A, Borgemeister C, Gottlieb M, Mulla MS, Gimnig JE, Fish D, Killeen GF. Ecology: A prerequisite for malaria elimination and eradication. PLoS Medicine. 2010;7:e1000303. DOI: 10.1371/journal.pmed.1000303
  76. 76. Achee NL, Bangs MJ, Farlow R, Killeen GF, Lindsay S, Logan JG, Moore SJ, Rowland M, Sweeney K, Torr SJ. Spatial repellents: From discovery and development to evidence-based validation. Malaria Journal. 2012;11:164. DOI: 10.1186/1475-2875-11-164
  77. 77. Molineaux L, Gramiccia G, Organization WH. The Garki project: research on the epidemiology and control of malaria in the Sudan savanna of West Africa. In: Molineaux L, Gramiccia G. The Garki Project: Research on the Epidemiology and Control of Malaria in the Sudan Savanna of West Africa. 1980
  78. 78. Govella NJ, Chaki PP, Killeen GF. Entomological surveillance of behavioural resilience and resistance in residual malaria vector populations. Malaria Journal. 2013;12:124. DOI: 10.1186/1475-2875-12-124
  79. 79. Durnez L, Coosemans M. Residual transmission of malaria: An old issue for new approaches. In: Manguin S, editor. Anopheles Mosquitoes-New Insights into Malaria Vectors. Rijeka, Croatia: InTech Open Access; 2013. DOI: 10.5772/55925
  80. 80. Killeen GF. Characterizing, controlling and eliminating residual malaria transmission. Malaria Journal. 2014;13:330. DOI: 10.1186/1475-2875-13-330
  81. 81. WHO Malaria Policy Advisory Committee (MPAC) meeting. 2017. Available from: http://www.apps.who.int/iris/bitstream/10665/259396/1/WHO-HTM-GMP-2017.21-eng.pdf [Accessed 2017-12-5]
  82. 82. Hii J, Edwards HM. Studies to examine magnitude and identify causes (study sites in Thailand and Viet Nam). In: Workshop on Residual Malaria Transmission, Dar-Es-Salam, Tanzania, TDR/WHO, IHI and GMP/WHO. 2017
  83. 83. Van Bortel W, Trung HD, Hoi-le-X VHN, Van-Chut N, Luu ND, Roelants P, Coosemans M. Malaria transmission and vector behaviour in a forested malaria focus in central Vietnam and the implications for vector control. Malaria Journal. 2010;9:373. DOI: 10.1186/1475-2875-9-373
  84. 84. Tangena JAA, Thammavong P, Lindsay SW, Brey PT. Risk of exposure to potential vector mosquitoes for rural workers in Northern Lao PDR. PLOS Neglected Tropical Diseases. 2017;11:e0005802. DOI: 10.1371/journal.pntd.0005802
  85. 85. Manguin S, Garros C, Dusfour I, Harbach R, Coosemans M. Bionomics, taxonomy, and distribution of the major malaria vector taxa of Anopheles subgenus Cellia in Southeast Asia: An updated review. Infection, Genetics and Evolution. 2008;8:489-503. DOI: 10.1016/j.meegid.2007.11.004
  86. 86. Hlaing T, Wai KT, Oo T, Sint N, Min T, Myar S, Lon KN, Naing MM, Tun TT, Maung NLY. Mobility dynamics of migrant workers and their socio-behavioral parameters related to malaria in Tier II, Artemisinin Resistance Containment Zone, Myanmar. BMC Public Health. 2015;15:886. DOI: 10.1186/s12889-015-2241-0
  87. 87. Parker DM, Matthews SA, Yan G, Zhou G, Lee M-C, Sirichaisinthop J, Kiattibutr K, Fan Q, Li P, Sattabongkot J. Microgeography and molecular epidemiology of malaria at the Thailand-Myanmar border in the malaria pre-elimination phase. Malaria Journal. 2015;14:198. DOI: 10.1186/s12936-015-0712-5
  88. 88. Wangroongsarb P, Satimai W, Khamsiriwatchara A, Thwing J, Eliades JM, Kaewkungwal J, Delacollette C. Respondent-driven sampling on the Thailand-Cambodia border. II. Knowledge, perception, practice and treatment-seeking behaviour of migrants in malaria endemic zones. Malaria Journal. 2011;10:117. DOI: 10.1186/1475-2875-10-117
  89. 89. Tipmontree R, Fungladda W, Kaewkungwal J, Tempongko MSB, Schelp FP. Migrants and malaria risk factors: A study of the Thai-Myanmar border. The Southeast Asian Journal of Tropical Medicine and Public Health. 2009;40:1148
  90. 90. Rattanarithikul R, Mongkolpanya K, Noigamol C, Chanaimongkol S, Mahapibul P, Nakngen S. Dry season distribution of mosquito larvae in the bed of the Mekong River, northeastern Thailand. Journal of the American Mosquito Control Association. 1994;10:197-201
  91. 91. Rattanarithikul R, Green C, Panyim S, Noigamol C, Chanaimongkol S, Mahapibul P. Larval habitats of malaria vectors and other Anopheles mosquitoes around a transmission focus in northwestern Thailand. Journal of the American Mosquito Control Association. 1995;11:428-433
  92. 92. Hii J, Rueda L. Malaria vectors in the Greater Mekong Subregion: Overview of malaria vectors and remaining challenges. The Southeast Asian Journal of Tropical Medicine and Public Health. 2013;44:73-165 discussion 306-167
  93. 93. Durnez L, Van Bortel W, Denis L, Roelants P, Veracx A, Trung HD, Sochantha T, Coosemans M. False positive circumsporozoite protein ELISA: A challenge for the estimation of the entomological inoculation rate of malaria and for vector incrimination. Malaria Journal. 2011;0:195. DOI: 10.1186/1475-2875-10-195
  94. 94. Laurent BS, Oy K, Miller B, Gasteiger EB, Lee E, Sovannaroth S, Gwadz RW, Anderson JM, Fairhurst RM. Cow-baited tents are highly effective in sampling diverse Anopheles malaria vectors in Cambodia. Malaria Journal. 2016;15:440. DOI: 10.1186/s12936-016-1488
  95. 95. Alam MS, Chakma S, Khan WA, Glass GE, Mohon AN, Elahi R, Norris LC, Podder MP, Ahmed S, Haque R. Diversity of anopheline species and their Plasmodium infection status in rural Bandarban, Bangladesh. Parasites & Vectors. 2012;5:150. DOI: 10.1186/1756-3305-5-150
  96. 96. Ngo CT, Dubois G, Sinou V, Parzy D, Le HQ, Harbach RE, Manguin S. Diversity of Anopheles mosquitoes in Binh Phuoc and Dak Nong Provinces of Vietnam and their relation to disease. Parasites & Vectors. 2014;7:316. DOI: 10.1186/1756-3305-7-316
  97. 97. Sriwichai P, Samung Y, Sumruayphol S, Kiattibutr K, Kumpitak C, Payakkapol A, Kaewkungwal J, Yan G, Cui L, Sattabongkot J. Natural human Plasmodium infections in major Anopheles mosquitoes in western Thailand. Parasites & Vectors. 2016;9:17. DOI: 10.1186/s13071-016-1295-x
  98. 98. Bashara K, Tunob N, Ahmed TU, Howlader AJ. False positivity of circumsporozoite protein (CSP)–ELISA in zoophilic anophelines in Bangladesh. Acta Tropica. 2013;125:220-225
  99. 99. Harrison BA. The Myzomyia Series of Anopheles (Cellia) in Thailand, with emphasis on intra-interspecific variations (Diptera: Culicidae). Contributions of the American Entomological Institute. 1980;17:1-195
  100. 100. Van Bortel W, Harbach RE, Trung HD, Roelants P, Backeljau T, Coosemans M. Confirmation of Anopheles varuna in Vietnam, previously misidentified and mistargeted as the malaria vector Anopheles minimus. American Journal of Tropical Medicine and Hygiene. 2001;65:729-732
  101. 101. Erlank E, Koekemoer LL, Coetzee M. The importance of morphological identification of African anopheline mosquitoes (Diptera: Culicidae) for malaria control programmes. Malaria Journal. 2018;17:43. DOI: 10.1186/s12936-018-2189-5
  102. 102. Durnez L, Mao S, Denis L, Roelants P, Sochantha T, Coosemans M. Outdoor malaria transmission in forested villages of Cambodia. Malaria Journal. 2013;12:329. DOI: 10.1186/1475-2875-12-329
  103. 103. Gryseels C, Durnez L, Gerrets R, Uk S, Suon S, Set S, Phoeuk P, Sluydts V, Heng S, Sochantha T. Re-imagining malaria: Heterogeneity of human and mosquito behaviour in relation to residual malaria transmission in Cambodia. Malaria Journal. 2015;14:165. DOI: 10.1186/s12936-015-0689-0
  104. 104. Guerra C, Snow R, Hay S. A global assessment of closed forests, deforestation and malaria risk. Annals of Tropical Medicine and Parasitology. 2006;100:189. DOI: 10.1179/136485906X91512
  105. 105. Tangena JAA, Thammavong P, Malaithong N, Inthavong T, Ouanesamon P, Brey PT, Lindsay SW. Diversity of mosquitoes (Diptera: Culicidae) attracted to human subjects in rubber plantations, secondary forests, and villages in Luang Prabang Province, Northern Lao PDR. Journal of Medical Entomology. 2017;54:1589-1604. DOI: 10.1093/jme/tjx071
  106. 106. Tangena JAA, Thammavong P, Wilson AL, Brey PT, Lindsay SW. Risk and control of mosquito-borne diseases in Southeast Asian rubber plantations. Trends in Parasitology. 2016;32:402-415. DOI: 10.1016/j.pt.2016.01.009
  107. 107. Kaewwaen W, Bhumiratana A. Landscape ecology and epidemiology of malaria associated with rubber plantations in Thailand: Integrated approaches to malaria ecotoping. Interdisciplinary perspectives. Infectious Diseases. 2015;2015:909106. DOI: 10.1155/2015
  108. 108. Satitvipawee P, Wongkhang W, Pattanasin S, Hoithong P, Bhumiratana A. Predictors of malaria-association with rubber plantations in Thailand. BMC Public Health. 2012;12:1115. DOI: 10.1186/1471-2458-12-1115
  109. 109. Oo T. Biology and vector competence of the anopheline mosquitoes of Myanmar with special consideration of Anopheles dirus [thesis]. University of Heidelber; 2003
  110. 110. Crawshaw AF, Maung TM, Shafique M, Sint N, Nicholas S, Li MS, Roca-Feltrer A, Hii J. Acceptability of insecticide-treated clothing for malaria prevention among migrant rubber tappers in Myanmar: A cluster-randomized non-inferiority crossover trial. Malaria Journal. 2017;16:92. DOI: 10.1186/s12936-017-1737-8
  111. 111. Chang MS, Hii J, Buttner P, Mansoor F. Changes in abundance and behaviour of vector mosquitoes induced by land use during the development of an oil palm plantation in Sarawak. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1997;91:382-386. DOI: 10.1016/S0035-9203(97)90248-0
  112. 112. White N. Plasmodium knowlesi: The fifth human malaria parasite. Clinical Infectious Diseases. 2008;46:172. DOI: 10.1086/524889
  113. 113. Lee KS, Vythilingam I. Plasmodium knowlesi: Emergent human malaria in Southeast Asia. In: Lim YAL, Vythilingam I, editors. Parasites and their Vectors A Special Focus on Southeast Asia. Vienna: Springer; 2013. pp. 5-31
  114. 114. William T, Rahman HA, Jelip J, Ibrahim MY, Menon J, Grigg MJ, Yeo TW, Anstey NM, Barber BE. Increasing incidence of Plasmodium knowlesi malaria following control of P. falciparum and P. vivax malaria in Sabah, Malaysia. PLOS Neglected Tropical Diseases. 2013;7:e2026. DOI: 10.1371/journal.pntd.0002026
  115. 115. Setiadi W, Sudoyo H, Trimarsanto H, Sihite BA, Saragih RJ, Juliawaty R, Wangsamuda S, Asih PBS, Syafruddin D. A zoonotic human infection with simian malaria, Plasmodium knowlesi, in Central Kalimantan, Indonesia. Malaria Journal. 2016;15:218. DOI: 10.1186/s12936-016-1272-z
  116. 116. Herdiana H, Cotter C, Coutrier FN, Zarlinda I, Zelman BW, Tirta YK, Greenhouse B, Gosling RD, Baker P, Whittaker M. Malaria risk factor assessment using active and passive surveillance data from Aceh Besar, Indonesia, a low endemic, malaria elimination setting with Plasmodium knowlesi, Plasmodium vivax, and Plasmodium falciparum. Malaria Journal. 2016;15:468. DOI: 10.1186/s12936-016-1523-z
  117. 117. Lubis IN, Wijaya H, Lubis M, Lubis CP, Divis P, Beshir KB, Sutherland CJ. Contribution of Plasmodium knowlesi to multispecies human malaria infections in North Sumatera, Indonesia. The Journal of Infectious Diseases. 2017;215:1148-1155. DOI: 10.1093/infdis/jix091
  118. 118. Fornace KM, Nuin NA, Betson M, Grigg MJ, William T, Anstey NM, Yeo TW, Cox J, Ying LT, Drakeley CJ. Asymptomatic and submicroscopic carriage of Plasmodium knowlesi malaria in household and community members of clinical cases in Sabah, Malaysia. The Journal of Infectious Diseases. 2015;213:784-787. DOI: 10.1093/infdis/jiv475
  119. 119. Fornace KM, Abidin TR, Alexander N, Brock P, Grigg MJ, William T, Menon J, Drakeley C, Cox J. Association between landscape factors and spatial patterns of Plasmodium knowlesi infections in Sabah, Malaysia. Emerging Infectious Diseases. 2016;22. DOI: 10.3201/eid2202.150656
  120. 120. Moyes CL, Henry AJ, Golding N, Huang Z, Singh B, Baird JK, Newton PN, Huffman M, Duda KA, Drakeley CJ. Defining the geographical range of the Plasmodium knowlesi reservoir. PLOS Neglected Tropical Diseases. 2014;8:e2780. DOI: 10.1371/journal.pntd.0002780
  121. 121. Tan CH, Vythilingam I, Matusop A, Chan ST, Singh B. Bionomics of Anopheles latens in Kapit, Sarawak, Malaysian Borneo in relation to the transmission of zoonotic simian malaria parasite Plasmodium knowlesi. Malaria Journal. 2008;7:52. DOI: 10.1186/1475-2875-7-52
  122. 122. Jiram AI, Vythilingam I, NoorAzian YM, Yusof YM, Azahari AH, Fong MY. Entomologic investigation of Plasmodium knowlesi vectors in Kuala Lipis, Pahang, Malaysia. Malaria Journal. 2012;11:213. DOI: 10.1186/1475-2875-11-213
  123. 123. Wong ML, Chua TH, Leong CS, Khaw LT, Fornace K, Wan-Sulaiman W-Y, William T, Drakeley C, Ferguson HM, Vythilingam I. Seasonal and spatial dynamics of the primary vector of Plasmodium knowlesi within a major transmission focus in Sabah, Malaysia. PLOS Neglected Tropical Diseases. 2015;9:e0004135. DOI: 10.1371/journal.pntd.0004135
  124. 124. Marchand RP, Culleton R, Maeno Y, Quang NT, Nakazawa S. Co-infections of Plasmodium knowlesi, P. falciparum, and P. vivax among humans and Anopheles dirus mosquitoes, southern Vietnam. Emerging Infectious Diseases. 2011;17:1232-1239. DOI: 10.3201/eid1707.101551
  125. 125. Nakazawa S, Marchand RP, Quang NT, Culleton R, Manh ND, Maeno Y. Anopheles dirus co-infection with human and monkey malaria parasites in Vietnam. International Journal for Parasitology. 2009;39:1533-1537. DOI: 10.1016/j.ijpara.2009.08.005
  126. 126. Maeno Y, Quang NT, Culleton R, Kawai S, Masuda G, Nakazawa S, Marchand RP. Humans frequently exposed to a range of non-human primate malaria parasite species through the bites of Anopheles dirus mosquitoes in South-central Vietnam. Parasites & Vectors. 2015;8:376. DOI: 10.1186/s13071-015-0995-y
  127. 127. Baird JK. Asia-Pacific malaria is singular, pervasive, diverse and invisible. International Journal for Parasitology. 2017;47:371-377. DOI: 10.1016/j.ijpara.2016.06.006
  128. 128. Grigg MJ, Cox J, William T, Jelip J, Fornace KM, Brock PM, von Seidlein L, Barber BE, Anstey NM, Yeo TW. Individual-level factors associated with the risk of acquiring human Plasmodium knowlesi malaria in Malaysia: A case-control study. Lancet Planet Health. 2017;1:e97-e104. DOI: 10.1016/S2542-5196(17)30031-1
  129. 129. Manin BO, Ferguson HM, Vythilingam I, Fornace K, William T, Torr SJ, Drakeley C, Chua TH. Investigating the contribution of peri-domestic transmission to risk of zoonotic malaria infection in humans. PLoS Neglected Tropical Diseases. 2016;10:e0005064. DOI: 10.1371/journal.pntd.0005064
  130. 130. Akter R, Vythilingam I, Khaw LT, Qvist R, Lim YA-L, Sitam FT, Venugopalan B, Sekaran SD. Simian malaria in wild macaques: First report from Hulu Selangor district, Selangor, Malaysia. Malaria Journal. 2015;14:1-9. DOI: 10.1186/s12936-015-0856-3
  131. 131. Li M. Identification and molecular characterization of simian malaria parasites in wild monkeys of Singapore. [thesis]. National University of Singapore; 2011
  132. 132. Lee KS, Divis PCS, Zakaria SK, Matusop A, Julin RA, Conway DJ, Cox-Singh J, Singh B. Plasmodium knowlesi: Reservoir hosts and tracking the emergence in humans and macaques. PLoS Pathogens. 2011;7:e1002015. DOI: 10.1371/journal.ppat.1002015
  133. 133. Takken W, Knols BGJ. Malaria vector control: Current and future strategies. Trends in Parasitology. 2009;25:101-104. DOI: 10.1016/j.pt.2008.12.002
  134. 134. Liu W, Li Y, Learn GH, Rudicell RS, Robertson JD, Keele BF, Ndjango J-BN, Sanz CM, Morgan DB, Locatelli S. Origin of the human malaria parasite Plasmodium falciparum in gorillas. Nature. 2010;467:420-425. DOI: 10.1038/nature09442
  135. 135. Liu W, Li Y, Shaw KS, Learn GH, Plenderleith LJ, Malenke JA, Sundararaman SA, Ramirez MA, Crystal PA, Smith AG. African origin of the malaria parasite Plasmodium vivax. Nature Communications. 2014;5:3346. DOI: 10.1038/ncomms4346
  136. 136. Brant HL, Ewers RM, Vythilingam I, Drakeley C, Benedick S, Mumford JD. Vertical stratification of adult mosquitoes (Diptera: Culicidae) within a tropical rainforest in Sabah, Malaysia. Malaria Journal. 2016;15:370. DOI: 10.1186/s12936-016-1416-1
  137. 137. Yasuoka J, Levins R. Impact of deforestation and agricultural development on anopheline ecology and malaria epidemiology. American Journal of Tropical Medicine and Hygiene. 2007;76:450-460. DOI: 10.4269/ajtmh.2007.76.450
  138. 138. Singhasivanon P, Thimasarn K, Yimsamran S, Linthicum K, Nualchawee K, Dawreang D, Kongrod S, Premmanisakul N, Maneeboonyang W, Salazar N. Malaria in tree crop plantations in south-eastern and western provinces of Thailand. The Southeast Asian Journal of Tropical Medicine and Public Health. 1999;30:399-404
  139. 139. Overgaard HJ, Ekbom B, Suwonkerd W, Takagi M. Effect of landscape structure on anopheline mosquito density and diversity in northern Thailand: Implications for malaria transmission and control. Landscape Ecology. 2003;18:605-619. DOI: 10.1023/A-102
  140. 140. Overgaard HJ, Suwonkerd W, Hii J. The malaria landscape: Mosquitoes, transmission, landscape, insecticide resistance, and integrated control in Thailand. In: Serge-Morand DJP, Lefait-Robin R, Apiwathnasorn C, editors. Socio-Ecological Dimensions of Infectious Diseases in Southeast Asia. UK: Springer; 2015. p. 123-154
  141. 141. Wharton RH, Eyles DE. Anopheles hackeri, a vector of Plasmodium knowlesi in Malaya. Science. 1961;134:279. DOI: 10.1126/science.134.3474.279
  142. 142. Vythilingam I, Lim YA, Venugopalan B, Ngui R, Leong CS, Wong ML, Khaw L, Goh X, Yap N, Sulaiman WYW. Plasmodium knowlesi malaria an emerging public health problem in Hulu Selangor, Selangor, Malaysia (2009-2013): Epidemiologic and entomologic analysis. Parasites & Vectors. 2014;7:436. DOI: 10.1186/1756-3305-7-436
  143. 143. Hii JKL, Kan S, Pereria M, Parmar SS, Campos RL, Chan MKC. Transmission dynamics and estimates of malaria vectorial capacity for An. balabacensis and An. flavirostris on Bangii island, Sabah, Malaysia. Annals of Tropical Medicine and Parasitology. 1988;82:91-101. DOI: 10.1080/00034983.1988.11812214
  144. 144. Hii JLK, Kan S, Pereira M, Parmar SS, Campos RL, Chan MKC. Bancroftian filariasis and malaria in island and hinterland populations in Sabah, Malaysia. Tropical and Geographical Medicine. 1985;37:93-101
  145. 145. Manguin S, Bangs M, Pothikasikorn J, Chareonviriyaphap T. Review on global co-transmission of human Plasmodium species and Wuchereria bancrofti by Anopheles mosquitoes. Infection, Genetics and Evolution. 2010;10:159-177
  146. 146. Cox-Singh J. Zoonotic malaria: Plasmodium knowlesi, an emerging pathogen. Current Opinion in Infectious Diseases. 2012;25:530-536. DOI: 10.1097/QCO.0b013e3283558780
  147. 147. Cox-Singh J, Singh B. Knowlesi malaria: Newly emergent and of public health importance? Trends in Parasitology. 2008;24:406-410. DOI: 10.1016/j.pt.2008.06.001
  148. 148. Singh B, Daneshvar C. Human infections and detection of Plasmodium knowlesi. Clinical Microbiology Reviews. 2013;26:165-184. DOI: 10.1128/CMR.00079-12
  149. 149. Vythilingam I, Hii J. Simian malaria parasites: Special emphasis on Plasmodium knowlesi and their Anopheles vectors in Southeast Asia. In: Manguin S, editor. Anopheles Mosquitoes—New Insights into Malaria Vectors. Rijeka, Croatia: InTech Open Access; 2013. pp. 487-510. DOI: 10.5772/54491
  150. 150. Vythilingam I, Wong M, Wan-Yussof W. Current status of Plasmodium knowlesi vectors: A public health concern? Parasitology. 2018;145:32-40. DOI: 10.1017/S0031182016000901
  151. 151. Barber BE, Rajahram GS, Grigg MJ, William T, Anstey NM. World malaria report: Time to acknowledge Plasmodium knowlesi malaria. Malaria Journal. 2017;16:135. DOI: 10.1186/s12936-017-1787-y
  152. 152. Baird JK, Valecha N, Duparc S, White NJ, Price RN. Diagnosis and treatment of Plasmodium vivax malaria. American Journal of Tropical Medicine and Hygiene. 2016;95:35-51. DOI: 10.4269/ajtmh.16-0171
  153. 153. Howes RE, Piel FB, Patil AP, Nyangiri OA, Gething PW, Dewi M, Hogg MM, Battle KE, Padilla CD, Baird JK. G6PD deficiency prevalence and estimates of affected populations in malaria endemic countries: A geostatistical model-based map. PLoS Medicine. 2012;9:e1001339. DOI: 10.1371/journal.pmed.1001339
  154. 154. WHO. Single dose primaquine as a gametocytocide in Plasmodium falciparum malaria. 2012. Available from: http://www.who.int/malaria/publications/atoz/who_pq_policy_recommendation/en/ [Accessed: 2017-12-5]
  155. 155. Olliaro PL, Barnwell JW, Barry A, Mendis K, Mueller I, Reeder JC, Shanks GD, Snounou G, Wongsrichanalai C. Implications of Plasmodium vivax biology for control, elimination, and research. American Journal of Tropical Medicine and Hygiene. 2016;95:4-14. DOI: 10.4269/ajtmh.16-0160
  156. 156. WHO. Control and elimination of Plasmodium vivax malaria: A technical brief. World Health Organization; 2015. Available from: http://www.who.int/malaria/publications/atoz/9789241509244/en/ [Accessed: 2018-01-27]
  157. 157. Mendis K, Sina BJ, Marchesini P, Carter R. The neglected burden of Plasmodium vivax malaria. American Journal of Tropical Medicine and Hygiene. 2001;64:97-106. DOI: 10.4269/ajtmh.2001.64.97
  158. 158. Bousema T, Drakeley C. Epidemiology and infectivity of Plasmodium falciparum and Plasmodium vivax gametocytes in relation to malaria control and elimination. Clinical Microbiology Reviews. 2011;24:377-410. DOI: 10.1128/CMR.00051-10
  159. 159. Alonso PL, Tanner M. Public health challenges and prospects for malaria control and elimination. Nature Medicine. 2013;19:150-155. DOI: 10.1038/nm.3077
  160. 160. Vivax Working Group. Targeting vivax malaria in the Asia Pacific: The Asia Pacific malaria elimination network Vivax working group. Malaria Journal. 2015;14:484. DOI: 10.1186/s12936-015-0958-y
  161. 161. WHO. Research priorities for the environment, agriculture and infectious diseases of poverty. 2013. Available from: http://apps.who.int/iris/bitstream/10665/78129/1/WHO_TRS_976_eng.pdf [Accessed: 2017-12-5]
  162. 162. Campbell-Lendrum D, Molyneux D. Ecosystems and vector-borne disease control. In: Epstein P, Githeko A, Rabinovich J, Weinstein P, editors. Ecosystems and Human Well-Being: Policy Responses. Vol. 3. Washington, USA: Island Press; 2005. pp. 353-372
  163. 163. Stewart-Ibarra M. A social-ecological systems approach to dengue-chikungunya-zika in urban coastal Ecuador. In: A social-ecological Systems Approach to Dengue-chikungunya-zika in Urban Coastal Ecuador. Vol. 2017. 2016
  164. 164. WHO. Mass drug administration for malaria. A practical field manual. Malaria Policy Advisory Committee. 2017. Available from: http://www.who.int/malaria/mpac/mpac-mar2017-MDA-draft-manual-session7-presentation.pdf [Accessed: 2017-12-6]
  165. 165. Wilson AL, Chen-Hussey V, Logan JG, Lindsay SW. Are topical insect repellents effective against malaria in endemic populations? A systematic review and meta-analysis. Malaria Journal. 2014;13:446. DOI: 10.1186/1475-2875-13-446
  166. 166. Banks S, Murray N, WilderSmith A, Logan J. Insecticide-treated clothes for the control of vector-borne diseases: A review on effectiveness and safety. Medical and Veterinary Entomology. 2014;28:14-25. DOI: 10.1111/mve.12068
  167. 167. Parkes MW, Bienen L, Breilh J, Hsu L-N, McDonald M, Patz JA, Rosenthal JP, Sahani M, Sleigh A, Waltner-Toews D. All hands on deck: Transdisciplinary approaches to emerging infectious disease. EcoHealth. 2005;2:258-272. DOI: 10.1007/s12275-013-2393-5
  168. 168. Wilcox B, Kueffer C. Transdisciplinarity in EcoHealth: Status and future prospects. EcoHealth. 2008;5:1-3. DOI: 10.1007/s10393-008-0161-5
  169. 169. Okello-Onen J, Mboera LE, Mugisha S. Malaria research and management need rethinking: Uganda and Tanzania case studies. In: Charron DF, editor. Ecohealth Research in Practice. New York: Springer; 2012. pp. 139-151. DOI: 10.1007/978-1-4614-0517-7
  170. 170. Richter CH, Xu J, Wilcox BA. Opportunities and challenges of the ecosystem approach. Futures. 2015;67:40-51. DOI: 10.1016/j.futures.2014.12.002
  171. 171. Minkler M, Wallerstein N. Introduction to community-based participatory research. In: Wallerstein MMN, editor. Community-Based Participatory Research for Health: From Process to Outcomes. San Francisco: John Wiley & Sons; 2008. pp. 5-23
  172. 172. Guyant P, Corbel V, Guérin PJ, Lautissier A, Nosten F, Boyer S, Coosemans M, Dondorp AM, Sinou V, Yeung S. Past and new challenges for malaria control and elimination: The role of operational research for innovation in designing interventions. Malaria Journal. 2015;14:279. DOI: 10.1186/s12936-015-0802-4
  173. 173. WHO. Revised recommendations for achieving universal coverage with long-lasting insecticidal nets in malaria control. 2017. Available from: http://www.who.int/entity/malaria/mpac/mpac-oct2017-erg-universal-coverage-session9-presentation.pdf [Accessed: 2018-01-28]
  174. 174. Hewitt S, Delacollette C, Poirot E. Malaria control in the Greater Mekong Subregion: An overview of the current response and its limitations. The Southeast Asian Journal of Tropical Medicine and Public Health. 2013;44:249-305 discussion 306-247
  175. 175. WHO. Expert Consultation on Plasmodium knowlesi malaria to guide malaria elimination strategies. 2017. Available from: ww.wpro.who.int/mvp/meetings/exp_con_pknowlesi_2017/en/ [Accessed: 2018-01-27]

Notes

  • Cambodia, The People’s Republic of China (specifically Yunnan Province), the Lao People’s Democratic Republic (Lao PDR), Myanmar, Thailand, and Vietnam.

Written By

Jeffrey Hii, Indra Vythilingam and Arantxa Roca-Feltrer

Submitted: 20 June 2017 Reviewed: 30 March 2018 Published: 18 July 2018