Aptamers isolated against mosquito-borne pathogens

Abstract

Mosquito-borne diseases are a major threat to public health. The shortcomings of diagnostic tools, especially those that are antibody-based, have been blamed in part for the rising annual morbidity and mortality caused by these diseases. Antibodies harbor a number of disadvantages that can be clearly addressed by aptamers as the more promising molecular recognition elements. Aptamers are defined as single-stranded DNA or RNA oligonucleotides generated by SELEX that exhibit high binding affinity and specificity against a wide variety of target molecules based on their unique structural conformations. A number of aptamers were developed against mosquito-borne pathogens such as Dengue virus, Zika virus, Chikungunya virus, Plasmodium parasite, Francisella tularensis, Japanese encephalitis virus, Venezuelan equine encephalitis virus, Rift Valley fever virus and Yellow fever virus. Intrigued by these achievements, we carry out a comprehensive overview of the aptamers developed against these mosquito-borne infectious agents. Characteristics of the aptamers and their roles in diagnostic, therapeutic as well as other applications are emphasized.

Graphic abstract

Introduction

Mosquito-borne infections have been identified as the leading cause of high mortality and morbidity worldwide (Hoenen et al. 2006; Price and Thio 2011; Siegel et al. 2016). Pathogens that cause mosquito-borne infections have a broad geographic range and can cause human illnesses at various levels of severity. Many of them also infect wild animals, including rodents, bats, birds and monkeys. In certain cases, animals act as a hidden viral reservoir in a sylvatic cycle (Kuno et al. 1998). The mosquitoes imbibe the pathogen when blood-feeding on an infected host to support the replication of the pathogen in their organism. As such, they deliver a sufficiently large inoculum of the pathogen into the host. In order to infect a recipient host, a specific level of viraemia is usually required. Thus, only vertebrate species that contain a distinct degree of viraemia are considered as the amplifying hosts (Bárdos and Rosický 1979). When infected mosquitoes spread the virus to humans, it creates a condition in which humans are unable to transmit the virus further and become dead-end hosts (Fig. 1) (Bowen and Calisher 1976; Walton et al. 1973). Transmission of mosquito-borne pathogens from male to female occurs during copulation [horizontal or venereal transmission (VT)] or even from female to the offspring [vertical or transovarial transmission (TOT)]. Under TOT conditions, the mosquito vector plays a role as a long-term reservoir of the virus. Some pathogens only have a few hosts and vectors, while others replicate in a large number of hosts and vectors. Aside from that, some have a widespread distribution, whereas others are restricted to a limited geographic area (Morales et al. 2017). The transmission rate is strongly influenced by factors such as the population density of mosquito vectors, vertebrate hosts, floods, droughts and habitats such as the wetlands, shallow water reservoirs or sewage systems (Hubálek 2008). The complex interplay of these multiple factors necessitates an early diagnosis of these diseases to cure and prevent further viral progression, thus saving the lives of the infected. Available serological based diagnostic methods are largely beleaguered by the disadvantages of the antibodies such as reversible denaturation property, large batch-to-batch variation and temperature insensitive. An alternative molecular recognition element (MRE) known as aptamer has many advantages that have the potential to alleviate these drawbacks.

Fig. 1
figure1

The schematic diagram of pathogen transmission by the mosquito

The term “aptamer” comes from the combination of Latin “aptus” which means “to fit” and the Greek “meros” which means “particle” (Ellington and Szostak 1990). Aptamers are single-stranded oligonucleotides that are capable of binding to their cognate target molecules with high affinity and specificity based on their unique structural conformations (Ellington and Szostak 1990; Tuerk and Gold 1990). The aptamer-target interactions are governed by hydrogen bonding, electrostatic interactions, van der Waals forces and shape complementarity (Gelinas et al. 2016; Kong and Byun 2013; Nomura et al. 2010; Sun et al. 2014; Zhou et al. 2016). Since RNA aptamers contain a 2′-OH group, they can form more diverse three-dimensional (3D) structures compared to DNA aptamers, making them more flexible in binding to a broad spectrum of targets (Butcher and Pyle 2011; Hendrix et al. 2005; Sun et al. 2014). On the other hand, the absence of a hydroxyl group at the 2′ position of the deoxyribose sugar makes DNA aptamers more stable than RNA, eventually increasing the shelf-life of DNA aptamers (Lakhin et al. 2013; Zhu et al. 2015). Aptamers are generated in vitro from a random oligonucleotide library through a repetitive amplification and selection process known as systematic evolution of ligands by exponential enrichment (SELEX) (Fig. 2) (Ellington and Szostak 1990; Tuerk and Gold 1990). In several studies, different SELEX methods have been explored and modified to isolate high-affinity aptamers. Affinity chromatography and magnetic bead-based SELEX, nitrocellulose membrane filtration-based SELEX and capillary electrophoresis-based SELEX are the few examples (Gopinath 2007; Mencin et al. 2014; Song et al. 2012).

Fig. 2
figure2

The schematic diagram of SELEX: a library preparation, b incubation, c binding, d partitioning, e elution, f amplification/ssDNA generation/in vitro transcription, and g cloning and sequencing

To date, many aptamers have been selected against a wide range of targets such as metal ions, hormones, antibiotics, vitamins, toxins, proteins, cells, tissues, bacteria and viruses (Chiu et al. 2018; Kwon et al. 2020a, b; Li et al. 2019; Raducanu et al. 2020; Song et al. 2017; Stoltenburg et al. 2016; Su et al. 2020; Wang et al. 2019; Ye et al. 2019; Zhang et al. 2020). Aptamers have a low molecular weight, which allows for rapid tissue penetration, as well as other benefits such as greater structural flexibility, lower immunogenicity, prolonged shelf life, reusability due to reversible denaturation property, no batch-to-batch variation, lower production cost and ease of synthesis (Breaker 1997; Kruspe et al. 2014; Mairal et al. 2008; Wang et al. 2015; Zhang et al. 2019; Zhou and Rossi 2017). In addition, aptamers can selectively distinguish between very similarly structured biomolecules such as caffeine and theophylline, which differ only by a methyl group (Hermann and Patel 2000; Jenison et al. 1994; Wrist et al. 2020). Owing to their superior properties, aptamers have received a great deal of attention in numerous applications ranging from diagnostics to therapeutics, targeted drug delivery and bio-sensing (Alshamaileh et al. 2017; Citartan et al. 2019; Khan et al. 2018; Shigdar et al. 2013). Due to the aforementioned factors, aptamers are considered as an alternative to antibodies (Fig. 3).

Fig. 3
figure3

The illustration of the advantages of aptamers over antibodies

The advantages of aptamers have propelled extensive research into developing these molecules against a broad range of mosquito-borne infectious agents. An overview of these aptamers against mosquito-borne infectious agents was provided in this current study. This is the first endeavor to compile and underscore all the existing aptamers for mosquito-borne infectious agents including Dengue virus (DENV), Zika virus (ZIKV), Chikungunya virus (CHIKV), Plasmodium parasite, Francisella tularensis, Japanese encephalitis virus (JEV), Venezuelan equine encephalitis virus (VEEV), Rift Valley fever virus (RVFV) and Yellow fever virus (YFV). Moreover, the epidemiology of mosquito-borne diseases is briefly highlighted in this review (Table 1). The majority of aptamers were developed for diagnostic purposes as potential substituents for antibody-based pathogen detection that are still beset with setbacks. This review also includes several aptamers that have undergone therapeutic functionality testing and other applications.

Table 1 Mosquito-borne infectious agents and their respective majorly involved transmitting mosquitoes

Aptamers against Dengue Virus (DENV)

Globally, there are more than two billion people in tropical and subtropical areas that are at risk for dengue fever, with 50–100 million people infected and 24,000 deaths annually (Diamond and Pierson 2015). Dengue infection is caused by DENV, which is a member of the Flaviviridae family and the genus Flavivirus. Four closely related dengue serotypes (DENV types 1–4) cause human diseases including non-specific viral syndrome, fatal dengue hemorrhagic fever (DHF), dengue shock syndrome (DSS) and even death if not treated appropriately. This virus is transmitted to humans through the bite of infected female Aedes mosquitoes (WHO 2009). The rising global spread of DENV and the lack of an approved vaccine or anti-viral therapeutic has prompted extensive research towards diagnostics as well as therapeutics (Guzman et al. 2010). DENV has an 11-kb positive single-stranded RNA (ssRNA) genome. Three structural proteins (capsid (C), membrane (M), and envelope (E)) and seven non-structural proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5) are encoded by the genome; with type I cap structure at the 5′-end. The protein translation is achieved with the aid of the host translation machinery (Mukhopadhyay et al. 2005; Potisopon et al. 2014; Whitehead et al. 2007; Zhou et al. 2007).

Current diagnostics of DENV are enzyme-linked immunosorbent assay (ELISA) and reverse transcription polymerase chain reaction (RT-PCR) (Poloni et al. 2010). ELISA requires antibodies, which are time-consuming to produce, often unstable and have batch-to-batch variation. On the other hand, RT-PCR is a complicated procedure that takes a long time and necessitates the use of a costly thermocycler. As such, many efforts have paved the way towards aptamer-based diagnostics of DENV. Fletcher et al. (2010) achieved a significant milestone by developing a biosensor utilizing an oligonucleotide linker module based on restriction endonuclease EcoRI in a complex with an aptamer. The linker has a stem region that is complementary to the target sequence, in this case, a specific region of the viral genome. In the presence of the virus, the linker binds to the specific region of the genome. The linker’s complementary sequence to the aptamer sequence allows the aptamer to be released from the EcoRI-bound complex. As such, the released EcoRI enzyme will rapidly cleave multiple signaling molecules, generating detectable signals proportional to the number of viral copies. This is the first sensor that relies on the ‘amplifying effect’ of the restriction enzyme for the detection of the virus. Besides that, Basso et al. (2019) used a hybrid nanomaterial formed by magnetic nanoparticles γ-Fe2O3 conjugated to gold nanoparticles modified with DNA aptamers on the surface to develop a simple and fast colorimetric immunosensor for the detection of DENV. Fourier-transform infrared (FTIR) spectroscopy, transmission electron microscopy (TEM), quartz crystal microbalance (QCM), UV–Vis and localized surface plasmon resonance (SPR) were among the techniques used to characterize the biosensor. Colorimetric changes upon the aptamer-target binding allow rapid visual detection of the virus without the requirement of any special equipment. As the DENV’s structure resembles that of another virus from the same genus, the binding specificity of the aptamer was crosschecked against ZIKV and YFV. The aptamer did not cross-react with ZIKV or YFV according to the UV–Vis results. One possible challenge that the researchers might have to deal with is the ambiguity of the colorimetric result, which could stem from the equivocality of the naked-eye observatory-based interpretation. In an innovative aptamer design, Kwon et al. (2020a, b) created a star-shaped DNA architecture carrying five molecular beacon-like motifs constructed to display ten DENV E protein domain III (ED3)-targeting aptamers that are specific and precisely fit the spatial arrangement of ED3 clusters on the DENV viral surface. The researchers endeavored to elevate the sensing sensitivity by incorporating an element of avidity into the sensor by conflating multiple aptamers into one single unit. The construct showed a high binding avidity for the ED3 protein and is promising for the development of a DENV detection sensor that outperforms the existing gold standard methods such as ELISA and quantitative reverse transcription polymerase chain reaction (RT-qPCR). Recently, several aptasensors targeting NS1 protein were developed. In one study, aptamer against NS1 protein of DENV was used for the development of fluorescence aptasensor. The fluorescence emission of the NS1 aptamer labelled with 6-carboxyfluorescein (FAM) at the 5′-end was quenched due to the structural conformation change upon binding to the target NS1. A limit of detection (LOD) of 2.51 nM and 8.13 nM in buffer and serum, respectively were achieved (Mok et al. 2021). Junior et al. (2021) have also isolated aptamers against the NS1 protein of DENV and used them in an electrochemical sensor. The aptamers were immobilized on the surface of the gold electrode with the aid of 6-mercapto-1-hexanol (MCH) to form a self-assembled monolayer while blocking was achieved with BSA. The LOD achieved was down to 0.025 ng/mL, which is within the clinical range of NS1. The specificity of the sensor was also tested against the E protein of DENV, which showed insignificant interaction. These aptamers against NS1 proteins should also be extended to point-of-care diagnostic systems such as lateral flow assay.

Several DENV proteins were subjected to aptamer isolation. These aptamers unveil therapeutic properties by antagonizing the interaction of these viral proteins, which would otherwise facilitate pathogenesis. Chen et al. (2015) used SELEX to obtain a DNA aptamer, S15, which binds to DENV-2 E protein domain III (ED3) with a dissociation constant of 200 nM. The Quadfinder prediction and circular dichroism experimentation results revealed that S15 is capable of forming a parallel G-quadruplex. Both the quadruplex structure and the sequence at the 5′-end of the aptamer were predicted to be an important region for the binding against the targeted protein domain. Furthermore, NMR titration results indicated that the highly conserved loop between the βA and βB strands of ED3 serves as the binding region for the aptamer. Interestingly, the S15 aptamer was unable to bind to the denatured ED3 in the western blot assay, suggesting that the aptamer recognizes a conformational epitope instead of a linear epitope. Although S15 aptamer was isolated against DENV-2, the authors also discovered that it could neutralize all four DENV serotypes. Using a similar protein target, ED3, Gandham et al. (2014) identified DENTA-1, an ssDNA thioaptamer with a dissociation constant of 154 nM that binds to the antibody-binding site within ED3. In addition, Jung et al. (2017) have generated an RNase-resistant 2′-fluoro-modified RNA aptamer that directly binds to DENV-2 methyltransferase (MTase) and truncated it down to 45-mer RNA sequence without losing binding specificity. The authors have unleashed the therapeutic value of the aptamer, as it was found to disrupt the function of the DENV MTase, which is responsible for the sequential guanine N-7 methylation and the ribose 2′-O methylation of the type-I cap of the DENV RNA. The 45-mer aptamer not only showed binding against DENV2 MTase with the Kd of 28 ± 2.1 nM but also binds to DENV3 MTase with the Kd of 15.6 ± 1.03 nM and was able to disrupt its methylation activity.

Most of the aptamers used so for were directed against the DENV proteins. Targeting intracellular host protein was also proven to be useful in a DENV therapeutic intervention. Balinsky et al. (2013) proved that the addition of a nucleolin (NCL)-binding aptamer (AS1411) would disrupt the association between DENV C protein and multifunctional host protein NCL. NCL is a multifunctional cellular protein that plays a significant role in an array of different cellular processes, including ribosome biogenesis, protein transport, chromatin remodeling, translational regulation, RNA processing and stability. The AS1411 aptamer reduces DENV C protein colocalization, causes a significant reduction of viral titers following DENV infection. Furthermore, the authors demonstrated that the treatment with AS1411 affected the viral DENV C's migration characteristics. Apart from therapeutic and diagnostic uses, aptamer were also used in the identification of the viral RNA binding proteins involved in DENV replication. Dong et al. (2015) have devised a novel affinity purification strategy focused on the fusion of the streptavidin-binding RNA aptamer S1 sequence to the 3′ end of DENV 5′–3′ UTR RNA to isolate DENV-2 RNA-binding proteins (RBPs) from living mammalian cells. The streptavidin magnetic beads were coated with the S1 aptamer. This stratagem allowed the DENV 5′–3′ UTR RNA to isolate endogenous viral ribonucleoprotein (RNP) from the mammalian cell extract. This method led to the identification of several novel hosts DENV RBPs via liquid chromatography with tandem mass spectrometry (LC–MS). The outcome revealed RPS8, one of the RBPs discovered, which is linked to DENV replication. The protein identified can be targeted for therapeutic application in suppressing the pathogenesis of DENV. In summary, in the realm of DENV, aptamers have found various roles, ranging from diagnostics, therapeutics to viral protein studies.

Aptamers against Zika Virus (ZIKV)

In recent years, ZIKV, which is associated with Guillain–Barre’ syndrome in adults and microcephaly in newborns, has caused sporadic outbreaks (Calvet et al. 2016). The World Health Organization (WHO) declared it as a Public Health Emergency of International Concern in February 2016. ZIKV is an arthropod-borne positive ssRNA virus that belongs to Flavivirus genus of the Flaviviridae family. It is primarily transmitted via mosquito bites of Aedes aegypti (Ae. aegypti) (Sher et al. 2019). However, there have been reports of mother-to-child, blood and sexual transmission (Calvet et al. 2016; Saiz et al. 2016). The ZIKV reproductive cycle begins when the viral particle attaches to the host’s cytoplasmic membrane via an E protein that promotes endocytosis. The viral membrane then fuses with the endosomal membrane, releasing positive ssRNA into the cytoplasm of the host cell. After that, translation starts and the resulting polyprotein is cleaved into three structural (C, pre-membrane (prM) and E) and seven NS proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B and NS5) (Zmurko et al. 2015).

Even in the case of ZIKV infection, aptamers have proven to be an expedient tool. Numerous aptamers were generated against a number of ZIKV proteins. In a study conducted by Lee and Zeng (2017), two aptamers (Aptamer 2 and 10) were selected after 7 rounds of SELEX cycles against ZIKV NS1 protein. Aptamer 2 showed the best binding affinity of 24 pM, whereas aptamer 10 showed the best binding specificity when compared to dengue NS1 serotypes. The LOD for ZIKV NS1 in the sandwich ELISA using aptamer-2-antibody pair were 0.1~1, 1~10 and > 10 ng/mL in buffer, 10% human serum and 100% human serum, respectively. In addition to Lee and Zeng’s research, ssDNA aptamers were developed using SELEX against ZIKV NS1 and NS5 proteins for sensing applications (Abalo et al. 2019; Alves et al. 2018; Morais et al. 2018, 2019). Aside from DNA and RNA, a peptide aptamer was also designed against ZIKV. Kim et al. (2018) used three bioinformatics methods, including BCPreds, ABCpred, and Bepipred, to devise a peptide aptamer against ZIKV E protein called Z_10.8 peptide (KRAVVSCAEA). The peptide aptamer selection was done based on several parameters such as binding affinity, root-mean-square deviation, the position of amine residue of lysine at the N-terminus and the position of the interactive sites in a docking analysis. The equilibrium dissociation constant estimated for the Z_10.8 peptide was 706.0 ± 177.9 nM. The diagnostic functionality of the aptamer was validated by fluorescence-linked sandwich immunosorbent assay (FLISA) and peptide-linked sandwich FLISA using ZIKV-spiked human serum and urine. The LOD for the sandwich FLISA was estimated at 1 × 104, for the 50% cell culture infectious dose (TCID) 50 mL. Researchers have also shown that even the matrix effect of serum or urine did not affect the performance of Z_10.8-linked sandwich FLISA. Interestingly, human serum interfered lesser with the peptides as opposed when using antibodies. Moreover, the smaller size of the peptides makes them less liable to non-specific electrostatic attraction from the components of serum.

As label-free assays are equally as sensitive as label-based approaches, many studies are gradually moving towards these strategies. Dolai and Tabib-Azar (2019) have created a label-free sensor that operates using an aptamer specific against ZIKV C protein. The aptamer was thiolated and immobilized on the surface of the 433 MHz Lithium Niobate (LiNbO3) microbalance sensors. As compared to the standard 5 MHz quartz microbalance, the whole virus can be detected with a sensitivity of 370 Hz/ng, which is approximately 400 times better. Dolai and Tabib-Azar (2020) have shown the applicability of quartz crystal microbalance (QCM) as a label-free approach of virus sensing. Similar researchers have reported the development of a paper-based potentiometric sensor to detect the whole ZIKV based on standard printer papers functionalized with aptamers. The paper sensor is made up of two to three 10 mm paper segments with conducting silver paint contact patches on both ends. When ZIKV was added, the two silver paint contacts reproducibly became more negative, yielding a minimum sensitivity of 0.26 nV/ZIKV and a minimum detectable signal of 2.4 × 107 ZIKV. Moreover, the authors also developed a proof-of-concept device using a liquid crystalline display (LCD) powered by the sensor to read the sensor output. In a colorimetric-based diagnostic application, aptamers specific for the ZIKV were conjugated to the surface of the gold nanoparticles. The aptamer-gold nanoparticles conjugate aggregate in the presence of the target ZIKV, causing the colour to change from red to blue. The LOD for live ZIKV was achieved at 1.0 × 105 PFU (Bosak et al. 2019). Though many sensors of various mode of sensing were developed thus far, more efforts should be directed towards the development of point-of-care-based strategy, such as lateral flow assay.

Apart from using aptamers generated against ZIKV, attempts to use other aptamers specific against certain dyes are also useful for the diagnostic detection of ZIKV infection. For instance, Kikuchi et al. (2019) designed a split DNA aptamer (SDA) hybridization probe employing two DNA strands. The innovation in it is that the SDA shows low binding affinity to the dapoxyl dye in normal conditions. However, in the presence of a specific analyte, it changes its structural conformation to form a dye-binding site (dapoxyl), amplifying the fluorescence signal by up to 120-fold. The developed SDA could selectively detect a conserved region of the ZIKV after isothermal nucleic acid sequence-based amplification (NASBA) reaction, which is a nucleic acid amplification strategy that operates only at a single temperature and has the potential to replace real-time PCR (qPCR). The effort by the researchers is laudable, as this assay is operatable at room temperature, in light of the isothermal nucleic acid amplification assay. This can be incorporated into a mobile system such as lateral flow assay, which as a whole fulfills the criterion of a point-of-care system.

Aptamers against Chikungunya Virus (CHIKV)

Chikungunya is an infectious disease caused by CHIKV, which was discovered in febrile human serum in Tanzania, Africa in 1953 (Robinson 1955). The common symptoms for CHIKV infections include fever, severe joint pain, muscle pain, joint swelling, fatigue and rash. Although it is not life-threatening, complex clinicopathological manifestations and increased number of mortalities were reported during the Indian Ocean outbreak in 2004 to 2005 (Enserink 2007). Subsequent to this outbreak, CHIKV has spread to almost all parts of the world; Africa, Asia, Europe and America (Kumar et al. 2008; Pialoux et al. 2007; Rezza et al. 2007). Mainly transmitted by Aedes species mosquitoes, CHIKV transmission cycles can be either man-mosquito-man (urban cycle) or animal-mosquito-man (sylvatic cycle). The sylvatic cycle involves forest-dwelling Aedes mosquitoes (Ae. furcifer, Ae. africanus, Ae. fulgens, A.luteocephalus, Ae. camptorhynchites) and non-human primates, which serve as reservoir hosts and are more common in Africa (Diallo et al. 1999). To date, these two mosquitoes (Ae. aegypti and Ae. albopictus) are the principal vectors involved in the urban transmission cycle and because of their wide distribution, the number of Chikungunya infection rise globally. With two open reading frames with the size of 11.8 kb that encodes for four NS proteins (NS1, NS2, NS3 and NS4) and five structural proteins (C, E3, E2, 6K and E1), CHIKV is a positive-sense ssRNA virus (Solignat et al. 2009). Since Chikungunya fever’s clinical symptoms are very similar to those of other tropical infections like DENV or leptospirosis, laboratory confirmation is in dire need. Virus isolation, molecular detection and antigen detection are the most appropriate test for samples collected between days 1 to 5 days from the illness onset. Meanwhile, samples collected after more than 5 days of illness onset preferably detect antibody (IgM and IgG) in the patient’s sample. This can be achieved by using ELISA, immunofluorescent assay (IFA), plaque reduction neutralization test (PRNT) and haemagglutination inhibition assay (HIA) (Pialoux et al. 2007). The current methods mentioned require high-end laboratory equipment such as a biosafety cabinet and fluorescence microscope, which are not usually available in laboratories with resource-limited setting. The urgency of CHIKV detection calls for a diagnostic system that is able to meet the ASSURED criteria suggested by WHO: affordable, sensitive, specific, user-friendly, rapid and robust, equipment-free and deliverable to those who need it (Urdea et al. 2006). Several endeavors were made to fulfill this aspiration using aptamers as the MREs. Bruno et al. (2012) developed DNA aptamers against CHIKV by using CHIKV E1 peptide as the target in SELEX. ELASA was used to rank the binding affinity of the aptamers based on the average absorbance at 405 nm. They successfully reported that both the capture and the conjugate aptamers could be used to detect CHIKV E peptide in a lateral flow chromatographic assay, despite not reporting on the binding affinity of each aptamer. On the other hand, Saraf et al. (2019) developed a multiplex framework for ZIKV and CHIKV E proteins using an aptamer-conjugated microfluidic channel approach. This aptamer-based device is one of the diagnostic platforms that can be used without the need for any fluorescent reporter because it is modular, compact and simple to use. In this format, the capture aptamer is immobilized on the surface of the microfluidic channel and the protein captured is detectable by a secondary aptamer conjugated with gold nanoparticles (AuNPs). As an electron-transferring agent, the presence of AuNPs leads to the deposition of silver reagents over the surface of AuNPs. A silver staining technique was used in this study to further amplify the colour change. The intensity of the colour depends on the concentration of analyte-bound aptamer-AuNPs. Therefore, silver deposition increases as the number of AuNPs increases, leading to intense colour change. This method achieves a detection limit of 1 pM of viral target protein in phosphine-buffered saline and 10 pM in diluted calf blood. The signal amplification strategy that is based on the usage of AuNPs and the deposition of silver should be used as an exemplary assay in other virus detection as well.

Aptamers against Plasmodium parasite

Plasmodium parasite is the causative agent of malaria. Transmitted by mosquito Anopheles, there are five Plasmodium species known to infect humans such as P. falciparum, P. vivax, P. ovale, P. malariae as well as P. knowlesi (Ashley et al. 2018). Despite global efforts to eradicate malaria, there were still an estimated 229 million cases and 409,000 deaths in 87 malaria-endemic countries in 2019 (WHO 2020), where the majority of the cases were contributed by P. falciparum and P. vivax. The gold standard for malarial diagnosis is the examination of stained blood films on light microscopy but has setbacks as it is a laborious, tedious, time-intensive technique and requires trained personnel (Ashley et al. 2018). As an effective malarial treatment requires early diagnosis, malaria diagnosis has been pivoted towards the development of rapid diagnostic kits (RDTs). Nonetheless, the current antibody-centric RDTs suffer from several caveats such as cost and thermal instability at high storage temperatures especially in tropical and subtropical countries (Jorgensen et al. 2006; Rafael et al. 2006). Thus, it is imperative to search for an alternative MRE with a much lower cost of synthesis and higher thermal stability.

In pursuant to this, several aptamers were isolated against biomarkers such as var2CSA (Birch et al. 2015), thirteen-amino acid long peptide (Frith et al. 2018), glutamate dehydrogenase (GDH) (Singh et al. 2018), High Mobility Group Box 1 (HMGB1) (Joseph et al. 2019), lactate dehydrogenase (LDH) (Cheung et al. 2013, 2020; Jain et al. 2016a; Lee et al. 2012) and the whole P. falciparum-infected red blood cells (Birch et al. 2015; Oteng et al. 2020). Among all, Plasmodium lactate dehydrogenase stands out as the most important target, as evidenced by the highest number of aptamers generated against this biomarker. Several aptamers developed against Plasmodium lactate dehydrogenase named pL1, 2008s and P38 were employed on multiple platforms such as gold nanoparticles (Cheung et al. 2013, 2020; Jain et al. 2016a, 2016b; Lee et al. 2012), graphene oxide (Jain et al. 2016a), silver nanoclusters (Wang et al. 2016), molybdenum disulfide (MoS2) nanosheet, magnetic microparticles (Kim and Searson 2017), DNA tweezer (Shiu et al. 2017) and DNA origami (Godonoga et al. 2016; Tang et al. 2018). Apart from the direct identification of LDH protein, signal production of the malarial diagnosis could also be carried out based on the enzymatic property of LDH that reduces pyruvate to l-lactate. Predicating on this enzymatic activity, an assay namely Aptamer-Tethered Enzyme Capture assay (APTEC) was developed, achieving a LOD of 4.9 ng/mL or 4.33 ± 1.66 pg/µL, as reported by two groups of researchers (Cheung et al. 2020; Dirkzwager et al. 2015). Most of the aptasensing platforms developed thus far were sensitive enough for the detection of the clinically relevant amount of LDH protein, which is around 3–15 pg/μL in plasma (Martin et al. 2009). In addition to the application of aptamers in the diagnosis of malaria, aptamers can also be used as therapeutic agents (Cui et al. 2015). For instance, Niles et al. (2009) demonstrated a successful inhibition of the growth of parasites and the formation of hemozoin by heme-binding aptamers. This inhibitory action of the aptamer is laudable as it is able to inhibit the accumulation of hemozoin, which is detrimental to red blood cells. Severe malaria is most likely caused by the sequestration of the infected erythrocytes by P. falciparum erythrocyte M protein 1 (PfEMP1), which serves as a key molecule in modulating the interaction of parasite and host (Flick and Chen 2004). Sequestration of parasitized erythrocytes from peripheral blood can be achieved by a cytoadherence strategy known as rosette formation, a phenomenon where uninfected erythrocytes agglutinate around parasitized erythrocytes (David et al. 1988). This cytoadherence was achievable by RNA aptamers generated by Barfod et al. (2009), whereby the aptamers generated showed adherence to the non-infected erythrocytes with a rosette disrupting capacity at 33 nM and with 100% disruption at 387 nM in a live cell assay. Nik Kamarudin et al. (2020) have isolated RNA aptamers that could potentially interrupt the interaction between PfEMP1 and CD36 receptor, a host endothelial surface protein. The researchers have proven the inhibitory action of the aptamer by promoting anti-cytoadherence, which could be useful for malaria adjunt therapy. Efforts should be intensified into bringing these aptamers into clinical trials to assess their efficacy for malaria adjunct therapy.

Aptamers against Francisella tularensis

Francisella tularensis, which is a Gram-negative non-motile Coccobacillus, is the causative agent responsible for Tularemia infection (Mandell et al. 2005). Tularemia is transmitted to humans by insect bites, ingestion of contaminated food or water and contact with infected animals. Apart from that, the transmission of tularemia was also proven to take place by the intermediate Ae. egypti (Bäckman et al. 2015). Additionally, an enzootic cycle is associated with the pathogenesis involving wild animals such as rodents and blood-sucking insects (Ellis et al. 2002). Due to high infectivity and the ability to cause lethal disease by aerosol, it is categorized as a class ‘A’ agent by the Centers for Disease Control and Prevention (CDC). In fact, F. tularensis could have a massive impact on public health if it was exploited by terrorists as a possible biological weapon (Day et al. 2009; Rotz et al. 2002; Willke et al. 2009). While various PCR platforms are used to detect F. tularensis, it is important to note that the gold standard for confirming F. tularensis detection is in vitro cultivation, which requires growth on cysteine or thioglycolate enriched medium and incubation times of 2–4 days at 37 °C (Dennis et al. 2001; Ellis et al. 2002). The viability of PCR to inhibitors and the time-consuming nature of the culture process calls for the possible applications of aptamers.

Only a few studies have been focused on the development of aptamers against F. tularensis. Vivekananda and Kiel (2006) selected DNA aptamers specific to the commercially available Francisella tularensis subspecies japonica bacterial antigen. The sandwich aptamer-linked immobilized sorbent assay (ALISA) and dot blot assay results exhibited the specificity of the aptamer cocktail to tularemia bacterial antigens from three different subspecies japonica, holarctica and tularensis, but not to unrelated Bartonella henselae, which was used as a negative control. Tularemia bacterial antigen was detected at concentrations as low as 25 ng. Apart from that, Lamont et al. (2014) used aptamers to perform a two-step enrichment process for improved cultivation and detection of F. tularensis in lettuce and soil. The first process utilizes logarithmic-phase F. tularensis spent culture filtrate to supplement standard culture medium to enhance F. tularensis growth in the presence of residual bacteria from food and environmental matrices. Within the spent culture supernatant, ultra-performance liquid chromatography (UPLC)/MS analysis found several unique chemicals including carnosine, which had a matching m/z ratio. Carnosine is a chemical that can cause enhanced growth and biofilm formation in E. coli. At low inoculums, adding 0.625 mg/mL of carnosine to conventional F. tularensis medium increased the growth of F. tularensis by ten fold. The second procedure used a DNA aptamer cocktail capture assay to concentrate and isolate F. tularensis amidst other bacteria present in food and environmental matrices in order to further enrich F. tularensis cells. The researchers have shown that using the aptamer as the capturing agent prior to enrichment resulted in a detection range of 1–106 CFU/mL, which is better than that of without any aptamer-based enrichment. The aptamer-based enrichment can also be used in a nucleic acid amplification assay such as PCR or RT-PCR, which can tremendously improve the LOD of F. tularensis.

Aptamers against Japanese encephalitis virus (JEV)

Japanese encephalitis (JE) is an acute viral infection of the central nervous system that is caused by JEV, which belongs to the genus Flavivirus in the family Flaviviridae. Only 0.1 to 1% of JEV infections result in encephalitis, while the rest are asymptomatic or characterized by a mild febrile illness (Grossman et al. 1973). There are approximately 67,900 cases of JEV-induced encephalitis globally each year with fatality rate ranges from 20 to 30% (Campbell et al. 2011). The majority of JE epidemic cases occur throughout Asian and Oceanian countries, with over three billion people are at risk of JEV transmission (Campbell et al. 2011; Wang and Liang 2015). JEV is an enveloped virus of about 50 nm virion size and consists of a positive-sense ssRNA genome of approximately 11 kb in length, which encodes for three structural proteins (C, prM and E) and seven NS proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5) (Desingu et al. 2017; Lu et al. 2017; Sumiyoshi et al. 1992). Initially, JEV is transmitted from pigs to mosquitoes (Barzon and Palù 2018; Imoto et al. 2010; Sasaki et al. 1982). When JEV-infected mosquitoes feed on human blood rather than their primary amplifying hosts, JEV replicates in the human skin and lymphoid organs. Next, JEV is transported by the bloodstream to the peripheral organs before crossing the blood–brain barrier to infect neurons and trigger infiltration of inflammatory cells (Myint et al. 2007; Pearce et al. 2018). Therefore, the WHO integrated vaccination programs into routine health control schedules in JEV-endemic countries to prevent JEV infection (Bharucha et al. 2020). These vaccines are both safe and reliable, offering long-lasting protection. However, it does not inhibit JEV circulation and may not be 100% effective, as demonstrated by Tandale et al. (2018).

The gold standard for JE diagnosis, which measures anti-JEV IgM and IgG antibodies in the blood serum or the cerebrospinal fluid is by PRNT (Hills et al. 2009). PRNT is laborious, time-consuming and necessitates a high level of biosafety. Another method is via ELISA, which suffers from the sensitivity of only 50–70% and cross-reactivity with other closely related Flaviviruses (Dubot-Pérès et al. 2015; Johnson et al. 2016; Maeda and Maeda 2013; Robinson et al. 2010). RT-qPCR assay is punctuated by the hurdle of the absence of JEV total RNA in blood and cerebrospinal fluid. Aptamer can be an elegant MRE that would be able to alleviate the disadvantages associated with the aforementioned assays (Bharucha et al. 2018; Nan et al. 2018; Pantawane et al. 2018; Saron et al. 2018).

So far, the only target used for aptamer selection was JEV MTase. JEV MTase catalyzes the methylation of the N-7 position of the guanine and the 2′-OH position of the first ribonucleotide in the viral genome (Ray et al. 2006). Han and Lee (2016) reported the first isolation of a 24-mer truncated RNA aptamer modified with 2'-O-methyl pyrimidines against the JEV MTase by using SELEX. The aptamer isolated bind specifically to JEV MTase with a high affinity (Kd ~ 16 nM) and is also able to inhibit both N-7 and 2′-OH-MTase activity of JEV MTase. JEV production and replication in cells were suppressed up to 80% compared to control. Moreover, the cellular positive and negative JEV genome RNA levels were decreased by 65% and 76%, respectively. In the therapeutic realm of JEV, the biggest challenge for the aptamer is the need to cross the blood–brain barrier, which consists of tight and adherent junction between adjacent endothelial cells (Chen and Liu 2012). This can be achieved by conjugating the aptamer with the aptamer against the transferrin receptor, which is a protein abundant on the endothelial cells of the blood–brain barrier (Li et al. 2020).

Aptamers against Venezuelan equine encephalitis virus (VEEV)

The Venezuelan equine encephalitis virus (VEEV) is an Alphavirus belonging to the Togaviridae family that causes epidemics and equine epizootics. It is made up of an approximately 11.4 kb positive-sense RNA genome that encodes for four NS proteins (NS1 to NS4) and a structural polyprotein that is proteolytically cleaved into the C and the E glycoproteins E2 and E1. The NS proteins play a major role in viral replication (Murphy, 1995; Rice and Strauss 1981; Strauss and Strauss 1986). Rodent hosts especially the cotton spiney rat and mosquito vectors such as Aedes and Culex mosquitoes play a role in the enzoonotic cycle of VEEV (Carrara et al. 2005, 2007; Deardorff and Weaver 2010; Deardorff et al. 2011; Ortiz et al. 2008; Smith et al. 2007). The diagnostic detection of VEEV involves the use of the PCR method and plaque assay (Linssen et al. 2000; Sahu et al. 1994). However, the usage of both the detection methods is limited during the viraemic phase of infection.

So far, only one aptamer was isolated for the diagnostic detection of VEEV. The target protein chosen for the isolation of aptamer was C protein. C protein is a viral protein that interacts with viral genomic RNA to form a nucleocapsid (NC) in the cytoplasm during viral assembly. Kang et al. (2007) reported a phosphorothioate RNA aptamer for the detection of the C protein of VEEV. The aptamer was modified with phosphorothioate at the 5′ end to improve the binding affinity of the aptamer. Due to its high binding affinity of 7.1 ± 2.4 nM, one particular aptamer, 16_1 aptamer, was identified as the best thioaptamer targeting the C protein of VEEV by chemiluminescence electrophoretic mobility shift assay (CL EMSA). Human cellular extracts were also used to evaluate the thioaptamer’s specificity, whereby in the gel shift assay, some band shifts were observed, which reflect on the possible non-specific interaction of the aptamer with the components of the human cellular extract. However, the researchers have not identified the proteins involved, thus this aptamer must be used cautiously in diagnostic applications that involve human cellular extracts. Besides C protein, another protein that can be targeted for aptamer selection in the future is NS1 protein, which plays an important role in pathogenesis, analogous to NS1 protein in DENV.

Aptamers against Rift Valley fever virus (RVFV)

Rift Valley fever virus (RVFV), a Bunyavirus from the genus Phlebovirus, is the etiological agent of Rift Valley fever (RVF). RVF is widespread across a large portion of Africa (Bird et al. 2009; Daubney et al. 1931). Mosquito vectors that have been linked to RVFV transmission are believed to be of Culex and Aedes genera (Chevalier et al. 2010). The RVFV consists of a tripartite negative-sense ssRNA genome, which is referred to as large (L), medium (M) and small (S). The L portion contains the coding regions for the RNA-dependent RNA polymerase (Müller et al. 1994). Meanwhile, the S portion expresses the nucleoprotein (N). The M segment encodes for a glycosylated 78-kDa protein, a non-glycosylated 14-kDa protein, two E glycoproteins Gn and Gc in a single open reading frame (Suzich et al. 1990). Virus isolation, histopathology, antigen detection, antibody detection and nucleic acid-based molecular assays such as nested RT-PCR methods, qPCR, multiplex PCR-based macroarray assay, real-time reverse-transcription loop-mediated isothermal amplification (RT-LAMP) and recombinant polymerase amplification (RPA) are among the latest RVFV diagnostic strategies (Bird et al. 2007; Drosten et al. 2002; Euler et al. 2012; Fukushi et al. 2012; Garcia et al. 2001; Jr et al. 1989; Kortekaas et al. 2013; Meegan et al. 1989; Mwaengo et al. 2012; Odendaal et al. 2014; Roux et al. 2009; Sall et al. 2002; Swanepoel et al. 1986; Venter et al. 2014; Wal et al. 2012; Weidmann et al. 2008). Besides that, numerous serological diagnosis methods are also available such as virus neutralization assay, ELISA, Indirect Immunofluorescence, agar gel immunodiffusion (AGID), radioimmunoassays and complement fixation and HIA (Pepin et al. 2010). However, these methods are time-consuming, lack specificity and require skilled personnel.

Direct detection of RVFV with aptamers by targeting the specific biomarkers can be a feasible diagnostic method. Due to its multifunction and viral-specific nature, the N protein is considered as a novel biomarker for diagnostic and therapeutic purposes among the other proteins encoded by the genome. The N protein protects the viral genome from degradation and prevents the formation of double-stranded RNA during replication, which could ignite antiviral response from host (Ruigrok et al. 2011). As such, aptamer selection has largely centered on N protein. Ellenbecker et al. (2012) successfully isolated a specific RNA aptamer targeting RVFV N protein using SELEX. Interestingly, the BLAST analysis data revealed the presence of GAUU and/or pyrimidine/guanine motifs within the coding region of the genome, implying that N protein might interact with non-terminal viral RNA sequences during replication. The aptamer was truncated and it retained an affinity of 2.6 µM for the N protein. In another EllenBecker’s experiment, he and his colleagues have conducted an in silico approach to isolate RNA aptamers and compared them with the experimentally acquired aptamer, MBE87. To compare the predicted secondary structures of the known aptamers against RVFV-N protein, they have designed an algorithm that uses a distance matrix and multidimensional scaling. The overall in silico-generated RNA aptamers were further validated in vitro with filter binding assay. The in silico-generated aptamers, AS-1 and AS4, did not bind with the similar binding affinity as the MBE87 sequence, the aptamer that survived 16 rounds of selection in the previous experiment. However, AS-1 and AS4 bind to the N protein and inhibit the viral function. The presence of the GAUU motif in both the experimentally derived and in silico-derived aptamers was predicted to be responsible for the high binding affinity (Ellenbecker et al. 2015). So far, the only protein targeted for aptamer selection is N protein. The isolation of aptamers should also be expanded beyond the target N protein, for example by using NS protein or even E glycoprotein as the SELEX target.

Aptamers against Yellow fever virus (YFV)

Yellow fever virus (YFV) is a Flavivirus with a positive-strand RNA genome that codes for three structural proteins and seven NS proteins, analogous to DENV, ZIKV and JEV. Despite the application of effective vaccines since 1930s, the disease still lingers. As the early symptoms caused by YFV are not significant, the diagnosis of YFV remains a hurdle. The non-specificity of certain serological methods also remains a challenge. Early detection of the disease is important to take on-time medical action. The qPCR, RT-LAMP and helicase-dependent amplification assays (HDA) were established as molecular methods for the diagnosis. However, they require expensive lab instruments and also skilled personnel (Escadafal et al. 2014).

The methylated 5′ cap is essential for mRNA stability and translation efficiency (Furuichi and Shatkin 2000). The viral protease/helicase activities of (NS2B/NS3) and the MTase of NS5 are prominent YFV biomarkers. So far, only one aptamer has been isolated against YFV, which was by Falk and Weisblum (2014). The aptamer against YFV MTase was functionalized at its 5′ end with fluorescein (FL-dT10). The FL-dT10 showed binding to YFV MTase with a binding affinity of 231 nM. As a whole, the isolation of aptamer against proteins associated with YFV is very limited thus far, which can be due to the low frequency of occurrence of yellow fever globally.

Concluding remarks

The current diagnostic tools for viral infections especially the ones that are incumbent upon antibodies as the MREs have their limitations. These drawbacks are addressable with the usage of aptamers based on their advantages over antibodies. In the future, we believe that early detection of mosquito-borne diseases will greatly improve disease control. In this review, we have overviewed many studies demonstrating the potential usage of aptamers in mosquito-borne infectious diseases agents, as outlined in Table 2. However, there are lack of aptamers available for many other mosquito-borne pathogens such as West Nile virus, Filariasis, Saint Louis encephalitis, Western equine encephalitis, Eastern equine encephalitis, Ross River fever, Barmah Forest fever, La Crosse encephalitis and newly detected Keystone virus. Future studies should also center on generating aptamers against these rare mosquito-borne pathogens. In the aspect of therapeutics, many aptamers have yet to enter human clinical trials. Therefore, clinical trials are needed to assess the actual efficacy of aptamers in therapeutics prior to the administration of aptamer-based treatments in humans.

Table 2 Summary data on aptamers selected against mosquito-borne pathogens

Future perspectives should focus on improving the stability of aptamers by incorporating chemical modifications that enhance the stability (Odeh et al. 2020). In silico approaches such as molecular docking and molecular dynamics could also be very useful to predict and characterize specific interactions between the aptamer and the target, in order to choose the best aptamer candidates (Navien et al. 2021). In addition, the incorporation of aptamers into portable biosensors could fulfill the requirements of point-of-care diagnostics (Guo et al. 2020). Apart from that, it is important to gain more insights into the basic knowledge of mosquito-borne infections that allows the visibility of novel biomarkers in order to select the best target candidates for the aptamer-based diagnostics. Finally, it is also essential to attract the attention of researchers, the public and industries for the transition of research to the stage of the development of marketable diagnostic prototypes (Ospina-villa et al. 2020). Taken together, aptamers are the dawning MREs, paving the trajectory towards a much more amenable diagnostic strategy of mosquito-borne pathogens.

Data availability

Material submitted is original; all authors are in agreement to have the article published.

References

  1. Abalo AA, Argondizzo APC, Morais LM et al (2019) Fluorescence spectroscopy study of single-stranded nucleic acid aptamer species against NS5 Zika virus. AIP Conf Proc 2186(1):1–5. https://doi.org/10.1063/1.5138052

    CAS  Article  Google Scholar 

  2. Alshamaileh H, Wang T, Xiang D et al (2017) Aptamer-mediated survivin RNAi enables 5-fluorouracil to eliminate colorectal cancer stem cells. Sci Rep 7(1):1–9. https://doi.org/10.1038/s41598-017-05859-z

    CAS  Article  Google Scholar 

  3. Alves LN, Abalo AA, Argondizzo APC et al (2018) Selection and evaluation of the binding of aptamers against NS5 Zika virus using fluorescence spectroscopy. AIP Conf Proc 2040(1):1–5. https://doi.org/10.1063/1.5079160

    CAS  Article  Google Scholar 

  4. Anderson GW Jr, Saluzzo JF, Ksiazek TG et al (1989) Comparison of in vitro and in vivo systems for propagation of Rift Valley fever virus from clinical specimens. Res Virol 140(2):129–138. https://doi.org/10.1016/s0923-2516(89)80090-1

    Article  PubMed  Google Scholar 

  5. Ashley AA, Phyo AP, Woodrow CJ (2018) Malaria. Lancet 391(10130):1608–1621. https://doi.org/10.1016/S0140-6736(18)30324-6

    Article  PubMed  Google Scholar 

  6. Bäckman S, Näslund J, Forsman M et al (2015) Transmission of tularemia from a water source by transstadial maintenance in a mosquito vector. Sci Rep 5(1):7793. https://doi.org/10.1038/srep07793

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  7. Balinsky CA, Schmeisser H, Ganesan S et al (2013) Nucleolin interacts with the Dengue virus capsid protein and plays a role in formation of infectious virus particles. J Virol 87(24):13094–13106. https://doi.org/10.1128/JVI.00704-13

    Article  PubMed  PubMed Central  Google Scholar 

  8. Bárdos V, Rosický B (1979) A proposal for the evaluation of vertebrates as to their role in the circulation of arboviruses. Folia Parasitol 26(1):89–91

    Google Scholar 

  9. Barfod A, Persson T, Lindh J (2009) In vitro selection of RNA aptamers against a conserved region of the Plasmodium falciparum erythrocyte membrane protein 1. Parasitol Res 105(6):1557–1566. https://doi.org/10.1007/s00436-009-1583-x

    Article  PubMed  PubMed Central  Google Scholar 

  10. Barzon L, Palù G (2018) Recent developments in vaccines and biological therapies against Japanese encephalitis virus. Expert Opin Biol Ther 18(8):851–864. https://doi.org/10.1080/14712598.2018.1499721

    CAS  Article  PubMed  Google Scholar 

  11. Basso CR, Crulhas BP, Magro M et al (2019) A new immunoassay of hybrid nanomater conjugated to aptamers for the detection of dengue virus. Talanta 197:482–490. https://doi.org/10.1016/j.talanta.2019.01.058

    CAS  Article  PubMed  Google Scholar 

  12. Bharucha T, Sengvilaipaseuth O, Vongsouvath M et al (2018) Development of an improved RT-qPCR Assay for detection of Japanese encephalitis virus (JEV) RNA including a systematic review and comprehensive comparison with published methods. PLoS ONE 13(3):1–18. https://doi.org/10.1371/journal.pone.0194412

    CAS  Article  Google Scholar 

  13. Bharucha T, Shearer FM, Vongsouvath M et al (2020) A need to raise the bar: a systematic review of temporal trends in diagnostics for Japanese encephalitis virus infection, and perspectives for future research. Int J Infect Dis 95:444–456. https://doi.org/10.1016/j.ijid.2020.03.039

    Article  PubMed  PubMed Central  Google Scholar 

  14. Birch CM, Hou HW, Han J et al (2015) Identification of malaria parasite-infected red blood cell surface aptamers by inertial microfluidic SELEX (I-SELEX). Sci Rep 5(1):1–16. https://doi.org/10.1038/srep11347

    Article  Google Scholar 

  15. Bird BH, Bawiec DA, Ksiazek TG et al (2007) Highly sensitive and broadly reactive quantitative reverse transcription-PCR assay for high-throughput detection of Rift Valley fever virus. J Clin Microbiol 45(11):3506–3513. https://doi.org/10.1128/JCM.00936-07

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  16. Bird BH, Ksiazek TG, Nichol ST et al (2009) Rift Valley fever virus. J Am Vet Med Assoc 234(7):883–893. https://doi.org/10.2460/javma.234.7.883

    Article  PubMed  Google Scholar 

  17. Bosak A, Saraf N, Willenberg A et al (2019) Aptamer–gold nanoparticle conjugates for the colorimetric detection of arboviruses and vector mosquito species. RSC Adv 9(41):23752–23763. https://doi.org/10.1039/c9ra02089f

    CAS  Article  Google Scholar 

  18. Bowen GS, Calisher CH (1976) Virological and serological studies of Venezuelan equine encephalomyelitis in humans. J Clin Microbiol 4(1):22–27. https://doi.org/10.1128/jcm.4.1.22-27.1976

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  19. Breaker RR (1997) DNA aptamers and DNA enzymes. Curr Opin Chem Biol 1(1):26–31. https://doi.org/10.1016/s1367-5931(97)80105-6

    CAS  Article  PubMed  Google Scholar 

  20. Bruno JG, Carrillo MP, Richarte AM et al (2012) Development, screening, and analysis of DNA aptamer libraries potentially useful for diagnosis and passive immunity of arboviruses. BMC Res Notes 5(633):1–12. https://doi.org/10.1186/1756-0500-5-633

    CAS  Article  Google Scholar 

  21. Butcher SE, Pyle AM (2011) The molecular interactions that stabilize RNA tertiary structure: RNA motifs, patterns, and networks. Acc Chem Res 44(12):1302–1311. https://doi.org/10.1021/ar200098t

    CAS  Article  PubMed  Google Scholar 

  22. Calvet G, Aguiar RS, Melo ASO et al (2016) Detection and sequencing of Zika virus from amniotic fluid of fetuses with microcephaly in Brazil: a case study. Lancet 16(6):653–660. https://doi.org/10.1016/S1473-3099(16)00095-5

    Article  Google Scholar 

  23. Campbell GL, Hills SL, Fischer M et al (2011) Estimated global incidence of Japanese encephalitis: a systematic review. Bull World Health Organ 89(10):766–774. https://doi.org/10.2471/BLT.10.085233

    Article  PubMed  PubMed Central  Google Scholar 

  24. Carrara A, Gonzales M, Ferro C et al (2005) Venezuelan equine encephalitis virus infection of spiny rats. Emerg Infect Dis 11(5):663–669. https://doi.org/10.3201/eid1105.041251

    Article  PubMed  PubMed Central  Google Scholar 

  25. Carrara A, Coffey LL, Aguilar PV et al (2007) Venezuelan equine encephalitis virus infection of cotton rats. Emerg Infect Dis 13(8):1158–1165. https://doi.org/10.3201/eid1308.061157

    Article  PubMed  PubMed Central  Google Scholar 

  26. Chen H, Hsiao W, Lee H et al (2015) Selection and characterization of DNA aptamers targeting all four serotypes of Dengue viruses. PLoS ONE 10(6):1–13. https://doi.org/10.1371/journal.pone.0131240

    CAS  Article  Google Scholar 

  27. Chen Y, Liu L (2012) Modern methods for delivery of drugs across the blood-brain barrier. Adv Drug Deliv Rev 64(7):640–665. https://doi.org/10.1016/j.addr.2011.11.010

    CAS  Article  PubMed  Google Scholar 

  28. Cheung Y, Kwok J, Law AWL et al (2013) Structural basis for discriminatory recognition of Plasmodium lactate dehydrogenase by a DNA aptamer. Proc Natl Acad Sci USA 110(40):15967–15972. https://doi.org/10.1073/pnas.1309538110

    Article  PubMed  PubMed Central  Google Scholar 

  29. Cheung Y, Röthlisberger P, Mechaly AE et al (2020) Evolution of abiotic cubane chemistries in a nucleic acid aptamer allows selective recognition of a malaria biomarker. Proc Natl Acad Sci U S A 117(29):16790–16798. https://doi.org/10.1073/pnas.2003267117

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  30. Chevalier V, Pépin M, Plée L et al (2010) Rift Valley fever: a threat for Europe? Euro Surveill 15(10):19506

    CAS  Article  Google Scholar 

  31. Chiang C, Beljanski V, Yin K et al (2015) Sequence-specific modifications enhance the broad spectrum antiviral response activated by RIG-I agonists. J Virol 89(15):8011–8025. https://doi.org/10.1128/JVI.00845-15

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  32. Chiu AS, Sankarapani V, Drabek R et al (2018) Inhibition of vitamin C oxidation by DNA aptamers. Aptamers 2:1–20

    Google Scholar 

  33. Citartan M, Kaur H, Presela R et al (2019) Aptamers as the chaperones (Aptachaperones) of drugs-from siRNAs to DNA nanorobots. Int J Pharm 567:1–16. https://doi.org/10.1016/j.ijpharm.2019.118483

    CAS  Article  Google Scholar 

  34. Cnossen EJ, Silva AG, Marangoni K et al (2017) Characterization of oligonucleotide aptamers targeting the 5´-UTR from Dengue virus. Future Med Chem 9(6):541–552. https://doi.org/10.4155/fmc-2016-0233

    CAS  Article  PubMed  Google Scholar 

  35. Cui L, Mharakurwa S, Ndiaye D et al (2015) Antimalarial drug resistance: literature review and activities and findings of the ICEMR network. Am J Trop Med Hyg 93(3):57–68. https://doi.org/10.4269/ajtmh.15-0007

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  36. Daubney R, Hudson JR, Garnham PC (1931) Enzootic hepatitis or Rift Valley fever. An undescribed virus disease of sheep cattle and man from East Africa. J Pathol Bacteriol 34(4):545–579. https://doi.org/10.1002/path.1700340418

    Article  Google Scholar 

  37. David PH, Handunnetti SM, Leech JH et al (1988) Rosetting: a new cytoadherence property of malaria-infected erythrocytes. Am J Trop Med Hyg 38(2):289–297. https://doi.org/10.4269/ajtmh.1988.38.289

    CAS  Article  PubMed  Google Scholar 

  38. Day JB, Nguyen H, Sharma SK et al (2009) Effect of dehydrated storage on the survival of Francisella tularensis in infant formula. Food Microbiol 26(8):932–935. https://doi.org/10.1016/j.fm.2009.06.005

    CAS  Article  PubMed  Google Scholar 

  39. Deardorff ER, Estrada-Franco JG, Freier JE et al (2011) Candidate vectors and rodent hosts of Venezuelan equine encephalitis virus, Chiapas, 2006–2007. Am J Trop Med Hyg 85(6):1146–1153. https://doi.org/10.4269/ajtmh.2011.11-0094

    Article  PubMed  PubMed Central  Google Scholar 

  40. Deardorff ER, Weaver SC (2010) Vector competence of Culex (Melanoconion) taeniopus for equine-virulent subtype IE strains of Venezuelan equine encephalitis virus. Am J Trop Med Hyg 82(6):1047–1052. https://doi.org/10.4269/ajtmh.2010.09-0556

    Article  PubMed  PubMed Central  Google Scholar 

  41. Dennis DT, Inglesby TV, Henderson DA et al (2001) Tularemia as a biological weapon: medical and public health management. JAMA 285(21):2763–2773. https://doi.org/10.1001/jama.285.21.2763

    CAS  Article  PubMed  Google Scholar 

  42. Desingu PA, Ray PK, John JK et al (2017) First complete genome sequence of genotype III Japanese encephalitis virus isolated from a stillborn piglet in India. Genome Announc 5(3):1503–1516. https://doi.org/10.1128/genomeA.01503-16

    Article  Google Scholar 

  43. Diallo M, Thonnon J, Traore-Lamizana M et al (1999) Vectors of Chikungunya virus in senegal: current data and transmission cycles. Am J Trop Med Hyg 60(2):281–286. https://doi.org/10.4269/ajtmh.1999.60.281

    CAS  Article  PubMed  Google Scholar 

  44. Diamond MS, Pierson TC (2015) Molecular insight into Dengue virus pathogenesis and its implications for disease control. Cell 162(3):488–492. https://doi.org/10.1016/j.cell.2015.07.005

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  45. Dirkzwager RM, Kinghorn AB, Richards JS et al (2015) APTEC: aptamer-tethered enzyme capture as a novel rapid diagnostic test for malaria. Chem Commun (camb) 51(22):4697–4700. https://doi.org/10.1039/C5CC00438A

    CAS  Article  Google Scholar 

  46. Dolai S, Tabib-Azar M (2019) 433 MHz Lithium Niobate microbalance aptamer-coated whole Zika virus sensor with 370 Hz/ng sensitivity. IEEE Sens J 20(8):1–6. https://doi.org/10.1109/JSEN.2019.2961611

    Article  Google Scholar 

  47. Dolai S, Tabib-Azar M (2020) Whole virus detection using aptamers and paper-based sensor potentiometry. Med Devices Sens. https://doi.org/10.1002/mds3.10112

    Article  PubMed  PubMed Central  Google Scholar 

  48. Dong Y, Yang J, Ye W et al (2015) Isolation of endogenously sssembled RNA-protein complexes using affinity purification based on streptavidin aptamer S1. Int J Mol Sci 16(9):22456–22472. https://doi.org/10.3390/ijms160922456

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  49. Drosten C, Göttig S, Schilling S et al (2002) Rapid detection and quantification of RNA of Ebola and Marburg viruses, Lassa virus, Crimean-Congo hemorrhagic fever virus, Rift Valley fever virus, Dengue virus, and yellow fever virus by real-time reverse transcription-PCR. J Clin Microbiol 40(7):2323–2330. https://doi.org/10.1128/JCM.40.7.2323

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  50. Dubot-Pérès A, Sengvilaipaseuth O, Chanthongthip A et al (2015) How many patients with anti-JEV IgM in cerebrospinal fluid really have Japanese encephalitis? Lancet Infect Dis 15(12):1376–1377. https://doi.org/10.1016/S1473-3099(15)00405-3

    Article  PubMed  Google Scholar 

  51. Ellenbecker M, Sears L, Li P et al (2012) Characterization of RNA aptamers directed against the nucleocapsid protein of Rift Valley fever virus. Antiviral Res 93(3):330–339. https://doi.org/10.1016/j.antiviral.2012.01.002

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  52. Ellenbecker M, Goddard JS, Sundet A et al (2015) Computational prediction and biochemical characterization of novel RNA aptamers to Rift Valley fever virus nucleocapsid protein. Comput Biol Chem 58:120–125. https://doi.org/10.1016/j.compbiolchem.2015.06.005

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  53. Ellington AD, Szostak JW (1990) In vitro selection of RNA molecules that bind specific ligands. Nature 346(6287):818–822. https://doi.org/10.1038/346818a0

    CAS  Article  PubMed  Google Scholar 

  54. Ellis J, Oyston PCF, Green M et al (2002) Tularemia. Clin Microbiol Rev 15(4):631–646. https://doi.org/10.1128/CMR.15.4.631

    Article  PubMed  PubMed Central  Google Scholar 

  55. Enserink M (2007) Infectious diseases. Chikungunya: no longer a third world disease. Science 318(5858):1860–1861. https://doi.org/10.1126/science.318.5858.1860

    CAS  Article  PubMed  Google Scholar 

  56. Escadafal C, Faye O, Sall AA et al (2014) Rapid molecular assays for the detection of yellow fever virus in low-resource settings. PLoS Negl Trop Dis 8(3):1–8. https://doi.org/10.1371/journal.pntd.0002730

    Article  Google Scholar 

  57. Euler M, Wang Y, Nentwich O et al (2012) Recombinase polymerase amplification assay for rapid detection of Rift Valley fever virus. J Clin Virol 54(4):308–312. https://doi.org/10.1016/j.jcv.2012.05.006

    CAS  Article  PubMed  Google Scholar 

  58. Falk SP, Weisblum B (2014) Aptamer displacement screen for flaviviral RNA methyltransferase inhibitors. J Biomol Screen 19(8):1147–1153. https://doi.org/10.1177/1087057114533147

    CAS  Article  PubMed  Google Scholar 

  59. Ferlin J, Farhat R, Belouzard S et al (2018) Investigation of the role of GBF1 in the replication of positive-sense single-stranded RNA viruses. J Gen Virol 99(8):1086–1096. https://doi.org/10.1099/jgv.0.001099

    CAS  Article  PubMed  Google Scholar 

  60. Fletcher SJ, Phillips LW, Milligan AS et al (2010) Toward specific detection of Dengue virus serotypes using a novel modular biosensor. Biosens Bioelectron 26(4):1696–1700. https://doi.org/10.1016/j.bios.2010.07.046

    CAS  Article  PubMed  Google Scholar 

  61. Flick K, Chen Q (2004) var genes, PfEMP1 and the human host. Mol Biochem Parasitol 134(1):3–9. https://doi.org/10.1016/j.molbiopara.2003.09.010

    CAS  Article  PubMed  Google Scholar 

  62. Fontenille D, Diallo M, Mondo M et al (1997) First evidence of natural vertical transmission of yellow fever virus in Aedes aegypti, its epidemic vector. Trans R Soc Trop Med Hyg 91(5):533–535. https://doi.org/10.1016/s0035-9203(97)90013-4

    CAS  Article  PubMed  Google Scholar 

  63. Frith K, Fogel R, Goldring JPD et al (2018) Towards development of aptamers that specifically bind to lactate dehydrogenase of Plasmodium falciparum through epitopic targeting. Malar J 17(1):191. https://doi.org/10.1186/s12936-018-2336-z

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  64. Fukushi S, Nakauchi M, Mizutani T et al (2012) Antigen-capture ELISA for the detection of Rift Valley fever virus nucleoprotein using new monoclonal antibodies. J Virol Methods 180(1–2):68–74. https://doi.org/10.1016/j.jviromet.2011.12.013

    CAS  Article  PubMed  Google Scholar 

  65. Furuichi Y, Shatkin AJ (2000) Viral and cellular mRNA capping: past and prospects. Adv Virus Res 55:135–184. https://doi.org/10.1016/s0065-3527(00)55003-9

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  66. Gandham SHA, Volk DE, Rao LGL et al (2014) Thioaptamers targeting Dengue virus type-2 envelope protein domain III. Biochem Biophys Res Commun 453(3):309–315. https://doi.org/10.1016/j.bbrc.2014.09.053

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  67. Garcia S, Crance JM, Billecocq A et al (2001) Quantitative real-time PCR detection of Rift Valley fever virus and its application to evaluation of antiviral compounds. J Clin Microbiol 39(12):4456–4461. https://doi.org/10.1128/JCM.39.12.4456

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  68. Gelinas AD, Davies DR, Janjic N (2016) Embracing proteins: Structural themes in aptamer–protein complexes. Curr Opin Struct Biol 36:122–132. https://doi.org/10.1016/j.sbi.2016.01.009

    CAS  Article  PubMed  Google Scholar 

  69. Glover K, Coombs KM (2020) ZIKV infection induces DNA damage response and alters the proteome of gastrointestinal cells. Viruses 12(7):771. https://doi.org/10.3390/v12070771

    CAS  Article  PubMed Central  Google Scholar 

  70. Glover KKM, Gao A, Zahedi-Amiri A et al (2019) Vero cell proteomic changes induced by Zika virus infection. Proteomics 19(4):1–6. https://doi.org/10.1002/pmic.201800309

    CAS  Article  Google Scholar 

  71. Godonoga M, Ting-Yu L, Oshima A et al (2016) A DNA aptamer recognising a malaria protein biomarker can function as part of a DNA origami assembly. Sci Rep 6:1–12. https://doi.org/10.1038/srep21266

    CAS  Article  Google Scholar 

  72. Gopinath SCB (2007) Methods developed for SELEX. Anal Bioanal Chem 387(1):171–182. https://doi.org/10.1007/s00216-006-0826-2

    CAS  Article  PubMed  Google Scholar 

  73. Grossman RA, Edelman R, Willhight M et al (1973) Study of Japanese encephalitis virus in Chiang Mai Valley, Thailand III. Human seroepidemiology and inapparent infections. Am J Epidemiol 98(2):133–149. https://doi.org/10.1093/oxfordjournals.aje.a121538

    CAS  Article  PubMed  Google Scholar 

  74. Guo X, Wen F, Zheng N et al (2020) Aptamer-based biosensor for detection of mycotoxins. Front Chem 8(195):1–19. https://doi.org/10.3389/fchem.2020.00195

    CAS  Article  Google Scholar 

  75. Guzman MG, Halstead SB, Artsob H et al (2010) Dengue: a continuing global threat. Nat Rev Microbiol 8(12 Suppl):7–16. https://doi.org/10.1038/nrmicro2460

    CAS  Article  Google Scholar 

  76. Han SR, Lee S (2016) Inhibition of Japanese encephalitis virus (JEV) replication by specific RNA aptamer against JEV methyltransferase. Biochem Biophys Res Commun 483(1):687–693. https://doi.org/10.1016/j.bbrc.2016.12.081

    CAS  Article  PubMed  Google Scholar 

  77. Hendrix DK, Brenner SE, Holbrook SR (2005) RNA structural motifs: building blocks of a modular biomolecule. Q Rev Biophys 38(3):221–243. https://doi.org/10.1017/S0033583506004215

    CAS  Article  PubMed  Google Scholar 

  78. Hermann T, Patel DJ (2000) Adaptive recognition by nucleic acid aptamers. Science 287(5454):820–825. https://doi.org/10.1126/science.287.5454.820

    CAS  Article  PubMed  Google Scholar 

  79. Hills S, Dabbagh A, Jacobson J et al (2009) Evidence and rationale for the World Health Organization recommended standards for Japanese encephalitis surveillance. BMC Infect Dis 9(214):1–9. https://doi.org/10.1186/1471-2334-9-214

    Article  Google Scholar 

  80. Hoenen T, Groseth A, Falzarano D et al (2006) Ebola virus: unravelling pathogenesis to combat a deadly disease. Trends Mol Med 12(5):206–215. https://doi.org/10.1016/j.molmed.2006.03.006

    CAS  Article  PubMed  Google Scholar 

  81. Hubálek Z (2008) Mosquito-borne viruses in Europe. Parasitol Res 103(1):29–43. https://doi.org/10.1007/s00436-008-1064-7

    Article  Google Scholar 

  82. Imoto J, Ishikawa T, Yamanaka A et al (2010) Needle-free jet injection of small doses of Japanese encephalitis DNA and inactivated vaccine mixture induces neutralizing antibodies in miniature pigs and protects against fetal death and mummification in pregnant sows. Vaccine 28(46):7373–7380. https://doi.org/10.1016/j.vaccine.2010.09.008

    CAS  Article  PubMed  Google Scholar 

  83. Jain P, Das S, Chakma B et al (2016a) Aptamer-graphene oxide for highly sensitive dual electrochemical detection of Plasmodium lactate dehydrogenase. Anal Biochem 514:32–37. https://doi.org/10.1016/j.ab.2016.09.013

    CAS  Article  PubMed  Google Scholar 

  84. Jain P, Chakma B, Singh NK et al (2016b) Aromatic surfactant as aggregating agent for aptamer-gold nanoparticle-based detection of Plasmodium lactate dehydrogenase. Mol Biotechnol 58(7):497–508. https://doi.org/10.1007/s12033-016-9946-x

    CAS  Article  PubMed  Google Scholar 

  85. Jenison RD, Gill SC, Pardi A et al (1994) High-resolution molecular discrimination by RNA. Science 263(5152):1425–1429. https://doi.org/10.1126/science.7510417

    CAS  Article  PubMed  Google Scholar 

  86. Johnson BW, Goodman CH, Jee Y et al (2016) Differential diagnosis of Japanese encephalitis virus infections with the Inbios JE Detect TM and DEN Detect TM MAC-ELISA kits. Am J Trop Med Hyg 94(4):820–828. https://doi.org/10.4269/ajtmh.15-0631

    Article  PubMed  PubMed Central  Google Scholar 

  87. Jorgensen P, Chanthap L, Rebueno A et al (2006) Malaria rapid diagnostic tests in tropical climates: the need for a cool chain. Am J Trop Med Hyg 74(5):750–754

    Article  Google Scholar 

  88. Joseph DF, Nakamoto JA, Ruiz OAG et al (2019) DNA aptamers for the recognition of HMGB1 from Plasmodium falciparum. PLoS ONE 14(4):1–20. https://doi.org/10.1371/journal.pone.0211756

    CAS  Article  Google Scholar 

  89. Jung JI, Han SR, Lee S (2017) Development of RNA aptamer that inhibits methyltransferase activity of Dengue virus. Biotechnol Lett 40(2):315–324. https://doi.org/10.1007/s10529-017-2462-7

    CAS  Article  PubMed  Google Scholar 

  90. Junior BB, Batistuti MR, Pereira AS et al (2021) Electrochemical aptasensor for NS1 detection: towards a fast dengue biosensor. Talanta 233:122527. https://doi.org/10.1016/j.talanta.2021.122527

    CAS  Article  Google Scholar 

  91. Kamarudin NAAN, Sat JNA, Zaidi NFM et al (2020) Evolution of specific RNA aptamers via SELEX targeting recombinant human CD36 protein: a candidate therapeutic target in severe malaria. Asian Pac J Trop Biomed 10(1):23–32. https://doi.org/10.4103/2221-1691.273091

    CAS  Article  Google Scholar 

  92. Kang J, Lee MS, Watowich SJ et al (2007) Combinatorial selection of a RNA thioaptamer that binds to Venezuelan equine encephalitis virus capsid protein. FEBS Lett 581(13):2497–2502. https://doi.org/10.1016/j.febslet.2007.04.072

    CAS  Article  PubMed  Google Scholar 

  93. Kenney JL, Adams AP, Gorchakov R et al (2012) Genetic and anatomic determinants of enzootic Venezuelan equine encephalitis virus infection of Culex (Melanoconion) taeniopus. PLoS Negl Trop Dis 6(4):1–13. https://doi.org/10.1371/journal.pntd.0001606

    Article  Google Scholar 

  94. Khan NI, Maddaus AG, Song E (2018) A low-cost inkjet-printed aptamer-based electrochemical biosensor for the selective detection of lysozyme. Biosensors 8(1):1–18. https://doi.org/10.3390/bios8010007

    CAS  Article  Google Scholar 

  95. Kikuchi N, Reed A, Gerasimova YV et al (2019) Split dapoxyl aptamer for sequence-selective analysis of nucleic acid sequence based amplification amplicons. Anal Chem 91(4):2667–2671. https://doi.org/10.1021/acs.analchem.8b03964

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  96. Kim C, Searson PC (2017) Detection of Plasmodium lactate dehydrogenase ( pLDH ) antigen in buffer using aptamer-modified magnetic microparticles for capture, oligonucleotide-modified quantum dots for detection, and oligonucleotide-modified gold nanoparticles for signal amplification. Bioconjug Chem 28(9):2230–2234. https://doi.org/10.1021/acs.bioconjchem.7b00328

    CAS  Article  PubMed  Google Scholar 

  97. Kim DTH, Bao DT, Park H et al (2018) Development of a novel peptide aptamer-based immunoassay to detect Zika virus in serum and urine. Theranostics 8(13):3629–3642. https://doi.org/10.7150/thno.25955

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  98. Kong HY, Byun J (2013) Nucleic acid aptamers: New methods for selection, stabilization, and application in biomedical science. Biomol Ther 21(6):423–434. https://doi.org/10.4062/biomolther.2013.085

    Article  Google Scholar 

  99. Kortekaas J, Kant J, Vloet R et al (2013) European ring trial to evaluate ELISAs for the diagnosis of infection with Rift Valley fever virus. J Virol Methods 187(1):177–181. https://doi.org/10.1016/j.jviromet.2012.09.016

    CAS  Article  PubMed  Google Scholar 

  100. Kruspe S, Mittelberger F, Szameit K et al (2014) Aptamers as drug delivery vehicles. ChemMedChem 9(9):1998–2011. https://doi.org/10.1002/cmdc.201402163

    CAS  Article  PubMed  Google Scholar 

  101. Kumar NP, Joseph R, Kamaraj T et al (2008) A226V mutation in virus during the 2007 Chikungunya outbreak in Kerala, India. J Gen Virol 89:1945–1948. https://doi.org/10.1099/vir.0.83628-0

    CAS  Article  PubMed  Google Scholar 

  102. Kuno G, Chang GJ, Tsuchiya KR et al (1998) Phylogeny of the genus Flavivirus. J Virol 72(1):73–83. https://doi.org/10.1128/JVI.72.1.73-83.1998

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  103. Kwon J, Lee Y, Lee T et al (2020a) Aptamer-based field-effect transistor for detection of avian influenza virus in chicken serum. Anal Chem 92(7):5524–5531. https://doi.org/10.1021/acs.analchem.0c00348

    CAS  Article  PubMed  Google Scholar 

  104. Kwon PS, Ren S, Kwon S et al (2020b) Designer DNA architecture offers precise and multivalent spatial pattern-recognition for viral sensing and inhibition. Nat Chem 12(1):26–35. https://doi.org/10.1038/s41557-019-0369-8

    CAS  Article  PubMed  Google Scholar 

  105. Lakhin AV, Tarantul VZ, Gening LV (2013) Aptamers: Problems, solutions and prospects. Acta Nat 5(4):34–43

    CAS  Article  Google Scholar 

  106. Lamont EA, Wang P, Enomoto S et al (2014) A combined enrichment and aptamer pulldown assay for Francisella tularensis detection in food and environmental matrices. PLoS ONE 9(12):1–19. https://doi.org/10.1371/journal.pone.0114622

    CAS  Article  Google Scholar 

  107. Lee KH, Zeng H (2017) Aptamer-based ELISA assay for highly specific and sensitive detection of Zika NS1 protein. Anal Chem 89(23):12743–12748. https://doi.org/10.1021/acs.analchem.7b02862

    CAS  Article  PubMed  Google Scholar 

  108. Lee S, Song K, Jeon W et al (2012) A highly sensitive aptasensor towards Plasmodium lactate dehydrogenase for the diagnosis of malaria. Biosens Bioelectron 35(1):291–296. https://doi.org/10.1016/j.bios.2012.03.003

    CAS  Article  PubMed  Google Scholar 

  109. Li W, Wang S, Zhou L et al (2019) An ssDNA aptamer selected by Cell-SELEX for the targeted imaging of poorly differentiated gastric cancer tissue. Talanta 199:634–642. https://doi.org/10.1016/j.talanta.2019.03.016

    CAS  Article  PubMed  Google Scholar 

  110. Li X, Yang Y, Zhao H et al (2020) Enhanced in vivo blood−brain barrier penetration by circular tau−transferrin receptor bifunctional aptamer for tauopathy therapy. J Am Chem Soc 142(8):3862–3872. https://doi.org/10.1021/jacs.9b11490

    CAS  Article  PubMed  Google Scholar 

  111. Linssen B, Kinney RM, Aguilar P et al (2000) Development of reverse transcription-PCR assays specific for detection of equine encephalitis viruses. J Clin Microbiol 38(4):1527–1535. https://doi.org/10.1128/JCM.38.4.1527-1535.2000

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  112. Lu B, Qin Y, Li B et al (2017) Full-length genome sequence of Japanese encephalitis virus strain FC792, isolated from Guangxi. China Genome Announc 5(48):1–2. https://doi.org/10.1128/genomeA.01054-17

    Article  Google Scholar 

  113. Maeda A, Maeda J (2013) Review of diagnostic plaque reduction neutralization tests for flavivirus infection. Vet J 195(1):33–40. https://doi.org/10.1016/j.tvjl.2012.08.019

    Article  PubMed  Google Scholar 

  114. Mairal T, Ozalp VC, Sánchez PL et al (2008) Aptamers: molecular tools for analytical applications. Anal Bioanal Chem 390(4):989–1007. https://doi.org/10.1007/s00216-007-1346-4

    CAS  Article  PubMed  Google Scholar 

  115. Mandell G, Bennett JE, Dolin R (2005) Mandell, Douglas, and Bennett’s principles and practice of infectious diseases. Clin Infect Dis 41:277

    Article  Google Scholar 

  116. Martin SK, Rajasekariah G, Awinda G et al (2009) Unified parasite lactate dehydrogenase and histidine-rich protein ELISA for quantification of Plasmodium falciparum. Am J Trop Med Hyg 80(4):516–522. https://doi.org/10.4269/ajtmh.2009.80.516

    CAS  Article  PubMed  Google Scholar 

  117. Meegan J, Guenno BL, Ksiazek T et al (1989) Rapid diagnosis of Rift Valley fever: a comparison of methods for the direct detection of viral antigen in human sera. Res Virol 140(1):59–65. https://doi.org/10.1016/s0923-2516(89)80085-8

    CAS  Article  PubMed  Google Scholar 

  118. Mencin N, Šmuc T, Vraničar M et al (2014) Optimization of SELEX: Comparison of different methods for monitoring the progress of in vitro selection of aptamers. J Pharm Biomed Anal 91:151–159. https://doi.org/10.1016/j.jpba.2013.12.031

    CAS  Article  PubMed  Google Scholar 

  119. Mok J, Jeon J, Jo J et al (2021) Novel one-shot fluorescent aptasensor for dengue fever diagnosis using NS1-induced structural change of G-quadruplex aptamer. Sens Actuators B 343:130077. https://doi.org/10.1016/j.snb.2021.130077

    CAS  Article  Google Scholar 

  120. Morais LM, Alves LN, Argondizzo APC et al (2018) DNA aptamer as molecular tool for ZIKV NS1 protein detection. AIP Conf Proc 2040(1):1–5. https://doi.org/10.1063/1.5079162

    CAS  Article  Google Scholar 

  121. Morais LM, Argondizzo APC, Silva D et al (2019) Aptamers against the Zika virus NS1 protein, for a serological diagnostic assay development. AIP Conf Proc 2186:1–5. https://doi.org/10.1063/1.5138061

    CAS  Article  Google Scholar 

  122. Morales MA, Fabbri CM, Zunino GE et al (2017) Detection of the mosquito-borne flaviviruses, West Nile, Dengue, Saint Louis encephalitis, Ilheus, Bussuquara, and yellow fever in free-ranging black howlers (Alouatta caraya) of Northeastern Argentina. PLoS Negl Trop Dis 11(2):1–13. https://doi.org/10.1371/journal.pntd.0005351

    Article  Google Scholar 

  123. Mukhopadhyay S, Kuhn RJ, Rossmann MG (2005) A structural perspective of the flavivirus life cycle. Nat Rev Microbiol 3(1):13–22. https://doi.org/10.1038/nrmicro1067

    CAS  Article  PubMed  Google Scholar 

  124. Müller R, Poch O, Delarue M et al (1994) Rift Valley fever virus L segment: correction of the sequence and possible functional role of newly identified regions conserved in RNA-dependent polymerases. J Gen Virol 75(6):1345–1352. https://doi.org/10.1099/0022-1317-75-6-1345

    Article  PubMed  Google Scholar 

  125. Murphy FA et al (1995) Virus taxonomy, 6th report of the International Committee on Taxonomy of Viruses. Arch Virol Suppl 10:1–586

    Google Scholar 

  126. Mwaengo D, Lorenzo G, Iglesias J et al (2012) Detection and identification of Rift Valley fever virus in mosquito vectors by quantitative real-time PCR. Virus Res 169(1):137–143. https://doi.org/10.1016/j.virusres.2012.07.019

    CAS  Article  PubMed  Google Scholar 

  127. Myint KSA, Gibbons RV, Perng GC et al (2007) Unravelling the neuropathogenesis of Japanese encephalitis. Trans R Soc Trop Med Hyg 101(10):955–956. https://doi.org/10.1016/j.trstmh.2007.04.004

    CAS  Article  PubMed  Google Scholar 

  128. Nan S, Li F, Nie K et al (2018) TaqMan real-time RT-PCR assay for detecting and differentiating Japanese encephalitis virus. Biomed Environ Sci 31(3):208–214. https://doi.org/10.3967/bes2018.026

    Article  Google Scholar 

  129. Navien TN, Thevendran R, Hamdani HY et al (2021) In silico molecular docking in DNA aptamer development. Biochimie 180:54–67. https://doi.org/10.1016/j.biochi.2020.10.005

    CAS  Article  PubMed  Google Scholar 

  130. Niles JC, Derisi JL, Marletta MA (2009) Inhibiting Plasmodium falciparum growth and heme detoxification pathway using heme-binding DNA aptamers. Proc Natl Acad Sci USA 106(32):13266–13271. https://doi.org/10.1073/pnas.0906370106

    Article  PubMed  PubMed Central  Google Scholar 

  131. Nomura Y, Sugiyama S, Sakamoto T et al (2010) Conformational plasticity of RNA for target recognition as revealed by the 2.15. A structure of a human IgG–aptamer complex. Nucleic Acids Res 38(21):7822–7829. https://doi.org/10.1093/nar/gkq615

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  132. Odeh F, Nsairat H, Aishaer W et al (2020) Aptamers chemistry: chemical modifications and conjugation strategies. Molecules 25(1):1–51. https://doi.org/10.3390/molecules25010003

    CAS  Article  Google Scholar 

  133. Odendaal L, Fosgate GT, Romito M et al (2014) Sensitivity and specificity of real-time reverse transcription polymerase chain reaction, histopathology, and immunohistochemical labeling for the detection of Rift Valley fever virus in naturally infected cattle and sheep. J Vet Diagn Invest 26(1):49–60. https://doi.org/10.1177/1040638713516759

    CAS  Article  PubMed  Google Scholar 

  134. Ortiz DI, Kang W, Weaver SC (2008) Susceptibility of Ae. aegypti (Diptera: Culicidae) to infection with epidemic (subtype IC) and enzootic (subtypes ID, IIIC, IIID) Venezuelan equine encephalitis complex alphaviruses. J Med Entomol 45(6):1117–1125. https://doi.org/10.1603/0022-2585(2008)45[1117:soaadc]2.0.co;2

    Article  PubMed  Google Scholar 

  135. Ospina-villa JD, Cisneros-sarabia A, Sánchez-Jiménez MM et al (2020) Current advances in the development of diagnostic tests based on aptamers in parasitology: a systematic review. Pharmaceutics 12(11):1–21. https://doi.org/10.3390/pharmaceutics12111046

    CAS  Article  Google Scholar 

  136. Oteng EK, Gu W, McKeague M (2020) High-efficiency enrichment enables identification of aptamers to circulating Plasmodium falciparum-infected erythrocytes. Sci Rep 10(1):1–11. https://doi.org/10.1038/s41598-020-66537-1

    CAS  Article  Google Scholar 

  137. Pantawane PB, Dhanze H, Kumar GVPPSR et al (2018) TaqMan real-time RT-PCR assay for detecting Japanese encephalitis virus in swine blood samples and mosquitoes. Anim Biotechnol 30(3):267–272. https://doi.org/10.1080/10495398.2018.1481417

    CAS  Article  PubMed  Google Scholar 

  138. Pearce JC, Learoyd TP, Langendorf BJ et al (2018) Japanese encephalitis: the vectors, ecology and potential for expansion. J Travel Med 25(1):16–26. https://doi.org/10.1093/jtm/tay009

    Article  Google Scholar 

  139. Pepin M, Bouloy M, Bird BH et al (2010) Rift Valley fever virus (Bunyaviridae: Phlebovirus): an update on pathogenesis, molecular epidemiology, vectors, diagnostics and prevention. Vet Res 41(61):1–40. https://doi.org/10.1051/vetres/2010033

    CAS  Article  Google Scholar 

  140. Pialoux G, Gaüzère B, Jauréguiberry S et al (2007) Chikungunya, an epidemic arbovirosis. Lancet Infect Dis 7(5):319–327. https://doi.org/10.1016/S1473-3099(07)70107-X

    Article  PubMed  Google Scholar 

  141. Poloni TR, Oliveira AS, Alfonso HL et al (2010) Detection of dengue virus in saliva and urine by real time RT-PCR. Virol J 7(22):1–4. https://doi.org/10.1186/1743-422X-7-22

    CAS  Article  Google Scholar 

  142. Potisopon S, Priet S, Collet A et al (2014) The methyltransferase domain of dengue virus protein NS5 ensures efficient RNA synthesis initiation and elongation by the polymerase domain. Nucleic Acids Res 42(18):11642–11656. https://doi.org/10.1093/nar/gku666

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  143. Price JC, Thio CL (2011) Liver disease in the HIV-infected individual. Clin Gastroenterol Hepatol 8(12):1002–1012. https://doi.org/10.1016/j.cgh.2010.08.024.Liver

    Article  Google Scholar 

  144. Raducanu V, Rashid F, Zaher MS et al (2020) A direct fluorescent signal transducer embedded in a DNA aptamer paves the way for versatile metal-ion detection. Sens Actuators B 304:1–11. https://doi.org/10.1016/j.snb.2019.127376

    CAS  Article  Google Scholar 

  145. Rafael ME, Taylor T, Magill A et al (2006) Reducing the burden of childhood malaria in Africa: the role of improved diagnostics. Nature 444(1):39–48. https://doi.org/10.1038/nature05445

    Article  PubMed  Google Scholar 

  146. Rashid M, Zahedi-Amiri A, Glover KKM et al (2020) Zika virus dysregulates human Sertoli cell proteins involved in spermatogenesis with little effect on tight junctions. PLoS Negl Trop Dis 14(6):1–23. https://doi.org/10.1371/journal.pntd.0008335

    CAS  Article  Google Scholar 

  147. Ray D, Shah A, Tilgner M et al (2006) West Nile virus 5’-cap structure is formed by sequential guanine N-7 and ribose 2’-O methylations by nonstructural protein 5. J Virol 80(17):8362–8370. https://doi.org/10.1128/JVI.00814-06

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  148. Rezza G, Nicoletti L, Angelini R et al (2007) Infection with Chikungunya virus in Italy: an outbreak in a temperate region. Lancet 370(9602):1840–1846

    CAS  Article  Google Scholar 

  149. Rice CM, Strauss JH (1981) Nucleotide sequence of the 26S mRNA of Sindbis virus and deduced sequence of the encoded virus structural proteins. Proc Natl Acad Sci USA 78(4):2062–2066. https://doi.org/10.1073/pnas.78.4.2062

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  150. Robinson JS, Featherstone D, Vasanthapuram R et al (2010) Evaluation of three commercially available Japanese encephalitis virus IgM enzyme-linked immunosorbent assays. Am J Trop Med Hyg 83(5):1146–1155. https://doi.org/10.4269/ajtmh.2010.10-0212

    Article  PubMed  PubMed Central  Google Scholar 

  151. Robinson MC (1955) An epidemic of virus disease in Southern province, Tanganyika territory, in 1952–1953. Trans R Soc Trop Med Hyg 49(1):28–32. https://doi.org/10.1016/0035-9203(55)90080-8

    CAS  Article  PubMed  Google Scholar 

  152. Rotz LD, Khan AS, Lillibridge SR (2002) Public health assessment of potential biological terrorism agents. Emerg Infect Dis 8(2):225–230. https://doi.org/10.3201/eid0802.010164

    Article  PubMed  PubMed Central  Google Scholar 

  153. Roux CAL, Kubo T, Grobbelaar AA et al (2009) Development and evaluation of a real-time reverse transcription-loop-mediated isothermal amplification assay for rapid detection of Rift Valley fever virus in clinical specimens. J Clin Microbiol 47(3):645–651. https://doi.org/10.1128/JCM.01412-08

    CAS  Article  PubMed  Google Scholar 

  154. Ruigrok RWH, Crépin T, Kolakofsky D (2011) Nucleoproteins and nucleocapsids of negative-strand RNA viruses. Curr Opin Microbiol 14(4):504–510. https://doi.org/10.1016/j.mib.2011.07.011

    CAS  Article  PubMed  Google Scholar 

  155. Sahu SP, Alstad AD, Pedersen DD et al (1994) Diagnosis of eastern equine encephalomyelitis virus infection in horses by immunoglobulin M and G capture enzyme-linked immunosorbent assay. J Vet Diagn Invest 6(1):34–38. https://doi.org/10.1177/104063879400600107

    CAS  Article  PubMed  Google Scholar 

  156. Saiz J, Vázquez-Calvo Á, Blázquez AB et al (2016) Zika virus the latest newcomer. Front Microbiol 7(496):1–19. https://doi.org/10.3389/fmicb.2016.00496

    Article  Google Scholar 

  157. Sall AA, Macondo EA, Sène OK et al (2002) Use of reverse transcriptase PCR in early diagnosis of Rift Valley fever. Clin Diagn Lab Immunol 9(3):713–715. https://doi.org/10.1128/CDLI.9.3.713

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  158. Saraf N, Villegas M, Willenberg BJ et al (2019) Multiplex viral detection platform based on a aptamers-integrated microfluidic channel. ACS Omega 4(1):2234–2240. https://doi.org/10.1021/acsomega.8b03277

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  159. Saron WAA, Rathore APS, Ting L et al (2018) Flavivirus serocomplex cross-reactive immunity is protective by activating heterologous memory CD4 T cells. Sci Adv 4(7):1–14. https://doi.org/10.1126/sciadv.aar4297

    CAS  Article  Google Scholar 

  160. Sasaki O, Karoji Y, Kuroda A et al (1982) Protection of pigs against mosquito-borne Japanese encephalitis virus by immunization with a live attenuated vaccine. Antiviral Res 2(6):355–360. https://doi.org/10.1016/0166-3542(82)90005-5

    CAS  Article  PubMed  Google Scholar 

  161. Sharma A, Knollmann-Ritschel B (2019) Current understanding of the molecular basis of Venezuelan equine encephalitis virus pathogenesis and vaccine development. Viruses 11(2):1–32. https://doi.org/10.3390/v11020164

    CAS  Article  Google Scholar 

  162. Sher AA, Glover KKM, Coombs KM (2019) Zika Virus infection disrupts astrocytic proteins involved in synapse control and axon guidance. Front Microbiol 10(596):1–20. https://doi.org/10.3389/fmicb.2019.00596

    Article  Google Scholar 

  163. Shigdar S, Macdonald J, O’Connor M et al (2013) Aptamers as theranostic agents: modifications, serum stability and functionalisation. Sensors 13(10):13624–13637. https://doi.org/10.3390/s131013624

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  164. Shiu SC, Cheung Y, Dirkzwager RM et al (2017) Aptamer-mediated protein molecular recognition driving a DNA tweezer nanomachine. Adv Biosys 1(2):1–6. https://doi.org/10.1002/adbi.201600006

    CAS  Article  Google Scholar 

  165. Siegel RL, Miller KD, Jemal A (2016) Cancer statistics. CA Cancer J Clin 66(1):7–30. https://doi.org/10.3322/caac.21332

    Article  Google Scholar 

  166. Singh NK, Chakma B, Jain P et al (2018) Protein-induced fluorescence enhancement based detection of Plasmodium falciparum glutamate dehydrogenase using carbon dot coupled specific aptamer. ACS Comb Sci 20(6):350–357. https://doi.org/10.1021/acscombsci.8b00021

    CAS  Article  PubMed  Google Scholar 

  167. Smith DR, Adams AP, Kenney JL et al (2008) Venezuelan equine encephalitis virus in the mosquito vector Aedes taeniorhynchus: Infection initiated by a small number of susceptible epithelial cells and a population bottleneck. Virology 372(1):176–186. https://doi.org/10.1016/j.virol.2007.10.011

    CAS  Article  PubMed  Google Scholar 

  168. Smith DR, Arrigo NC, Leal G et al (2007) Infection and dissemination of Venezuelan equine encephalitis virus in the epidemic mosquito vector, Aedes taeniorhynchus. Am J Trop Med Hyg 77(1):176–187

    Article  Google Scholar 

  169. Solignat M, Gay B, Higgs S et al (2009) Replication cycle of chikungunya: a re-emerging arbovirus. Virology 393(2):183–197. https://doi.org/10.1016/j.virol.2009.07.024

    CAS  Article  PubMed  Google Scholar 

  170. Song C, Chen C, Che X et al (2017) Detection of plant hormone abscisic acid (ABA) using an optical aptamer-based sensor with a microfluidics capillary interface. In Proceedings of the 2017 IEEE 30th International Conference on Micro Electro Mechanical Systems (MEMS), Las Vegas, NV, USA, pp 370–373. https://doi.org/10.1109/MEMSYS.2017.7863418

  171. Song K, Lee S, Ban C (2012) Aptamers and their biological applications. Sensors 12(1):612–631. https://doi.org/10.3390/s120100612

    Article  PubMed  PubMed Central  Google Scholar 

  172. Stoltenburg R, Krafčiková P, Víglaský V et al (2016) G-quadruplex aptamer targeting Protein A and its capability to detect Staphylococcus aureus demonstrated by ELONA. Sci Rep 6:1–12. https://doi.org/10.1038/srep33812

    CAS  Article  Google Scholar 

  173. Strauss EG, Strauss JH (1986) Structure and replication of the alphavirus genome. In: Schlesinger S, Schlesinger MJ (eds) The togaviridae and flaviviridae. The viruses. Springer, Boston, MA, pp 35–90. https://doi.org/10.1007/978-1-4757-0785-4_3

  174. Su C, Tsai M, Lin C et al (2020) Dual aptamer assay for detection of Acinetobacter baumannii on an electromagnetically-driven microfluidic platform. Biosens Bioelectron 159(112148):1–7. https://doi.org/10.1016/j.bios.2020.112148

    CAS  Article  Google Scholar 

  175. Sumiyoshi H, Hoke CH, Trent DW (1992) Infectious Japanese encephalitis virus RNA can be synthesized from in vitro-ligated cDNA templates. J Virol 66(9):5425–5431. https://doi.org/10.1128/JVI.66.9.5425-5431.1992

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  176. Sun H, Zhu X, Lu PY et al (2014) Oligonucleotide aptamers: New tools for targeted cancer therapy. Mol Ther Nucleic Acids 3(8):1–14. https://doi.org/10.1038/mtna.2014.32

    CAS  Article  Google Scholar 

  177. Suzich JA, Kakach LT, Collett MS (1990) Expression strategy of a Phlebovirus: Biogenesis of proteins from the Rift Valley fever virus M segment. J Virol 64(4):1549–1555. https://doi.org/10.1128/JVI.64.4.1549-1555.1990

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  178. Swanepoel R, Struthers JK, Erasmus MJ et al (1986) Comparison of techniques for demonstrating antibodies to Rift Valley fever virus. J Hyg Camb 97(2):317–329. https://doi.org/10.1017/s0022172400065414

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  179. Tandale BV, Khan SA, Kushwaha KP et al (2018) Effectiveness of Japanese encephalitis SA 14–14-2 live attenuated vaccine among Indian children: retrospective 1:4 matched case-control study. J Infect Public Health 11(5):711–719. https://doi.org/10.1016/j.jiph.2018.04.011

    Article  Google Scholar 

  180. Tang MSL, Shiu SC, Godonoga M et al (2018) An aptamer-enabled DNA nanobox for protein sensing. Nanomedicine 14(4):1161–1168. https://doi.org/10.1016/j.nano.2018.01.018

    CAS  Article  PubMed  Google Scholar 

  181. Tuerk C, Gold L (1990) Systematic evolution of ligands by exponential enrichment : RNA ligands to bacteriophage T4 DNA polymerase. Science 249(4968):505–510. https://doi.org/10.1126/science.2200121

    CAS  Article  PubMed  Google Scholar 

  182. Urdea M, Penny LA, Olmsted SS et al (2006) Requirements for high impact diagnostics in the developing world. Nature 444(1):73–79. https://doi.org/10.1038/nature05448

    Article  PubMed  Google Scholar 

  183. Venter M, Zaayman D, Niekerk SV et al (2014) Macroarray assay for differential diagnosis of meningoencephalitis in southern Africa. J Clin Virol 60(1):50–56. https://doi.org/10.1016/j.jcv.2014.02.001

    CAS  Article  PubMed  Google Scholar 

  184. Vivekananda J, Kiel JL (2006) Anti-Francisella tularensis DNA aptamers detect tularemia antigen from different subspecies by aptamer-linked immobilized sorbent assay. Lab Invest 86(6):610–618. https://doi.org/10.1038/labinvest.3700417

    CAS  Article  PubMed  Google Scholar 

  185. Wal FJVD, Achterberg RP, Boer SMD et al (2012) Bead-based suspension array for simultaneous detection of antibodies against the Rift Valley fever virus nucleocapsid and Gn glycoprotein. J Virol Methods 183(2):99–105. https://doi.org/10.1016/j.jviromet.2012.03.008

    CAS  Article  PubMed  Google Scholar 

  186. Walton TE Jr, AlwarezBuckwalter ORM et al (1973) Experimental infection of horses with enzootic and epizootic strains of Venezuelan equine encephalomyelitis virus. J Infect Dis 128(3):271–282. https://doi.org/10.1093/infdis/128.3.271

    CAS  Article  PubMed  Google Scholar 

  187. Wang H, Liang G (2015) Epidemiology of Japanese encephalitis: past, present, and future prospects. Ther Clin Risk Manag 11:435–448. https://doi.org/10.2147/TCRM.S51168

    Article  PubMed  PubMed Central  Google Scholar 

  188. Wang J, Gao T, Luo Y et al (2019) In vitro selection of a DNA Aptamer by Cell-SELEX as a molecular probe for cervical cancer recognition and imaging. J Mol Evol 87(2–3):72–82. https://doi.org/10.1007/s00239-019-9886-8

    CAS  Article  PubMed  Google Scholar 

  189. Wang T, Gantier MP, Xiang D et al (2015) EpCAM Aptamer-mediated survivin silencing sensitized cancer stem cells to doxorubicin in a breast cancer model. Theranostics 5(12):1456–1472. https://doi.org/10.7150/thno.11692

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  190. Wang W, Cheung Y, Dirkzwager RM et al (2016) Specific and sensitive detection of Plasmodium falciparum lactate dehydrogenase by DNA-scaffolded silver nanoclusters combined with an aptamer. Analyst. https://doi.org/10.1039/C6AN02417C

    Article  PubMed  PubMed Central  Google Scholar 

  191. Weidmann M, Sanchez-Seco MP, Sall AA et al (2008) Rapid detection of important human pathogenic Phleboviruses. J Clin Virol 41(2):138–142. https://doi.org/10.1016/j.jcv.2007.10.001

    CAS  Article  PubMed  Google Scholar 

  192. Whitehead SS, Blaney JE, Durbin AP et al (2007) Prospects for a dengue virus vaccine. Nat Rev Microbiol 5(7):518–528. https://doi.org/10.1038/nrmicro1690

    CAS  Article  PubMed  Google Scholar 

  193. Willke A, Meric M, Grunow R et al (2009) An outbreak of oropharyngeal tularaemia linked to natural spring water. J Med Microbiol 58(1):112–116. https://doi.org/10.1099/jmm.0.002279-0

    CAS  Article  PubMed  Google Scholar 

  194. WHO (2009) Dengue guidelines for diagnosis, treatment, prevention and control: New edition. World Health Organization. https://apps.who.int/iris/handle/10665/44188

  195. WHO (2020) World malaria report 2020: 20 years of global progress and challenges, Geneva, World Health Organization. https://www.who.int/publications/i/item/9789240015791

  196. Wrist A, Sun W, Summers RM (2020) The theophylline aptamer: 25 years as an important tool in cellular engineering research. ACS Synth Biol 9(4):682–697. https://doi.org/10.1021/acssynbio.9b00475

    CAS  Article  PubMed  Google Scholar 

  197. Ye W, Liu T, Zhang W et al (2019) Marine toxins detection by biosensors based on aptamers. Toxins 12(1):1–22. https://doi.org/10.3390/toxins12010001

    CAS  Article  PubMed Central  Google Scholar 

  198. Zhang W, Yang F, Ou D et al (2019) Prediction, docking study and molecular simulation of 3D DNA aptamers to their targets of endocrine disrupting chemicals. J Biomol Struct Dyn 37(16):4274–4282. https://doi.org/10.1080/07391102.2018.1547222

    CAS  Article  PubMed  Google Scholar 

  199. Zhang Z, Tian Y, Huang P et al (2020) Using target-specific aptamers to enhance the peroxidase-like activity of gold nanoclusters for colorimetric detection of tetracycline antibiotics. Talanta 208:1–8. https://doi.org/10.1016/j.talanta.2019.120342

    CAS  Article  Google Scholar 

  200. Zhou G, Wilson G, Hebbard L et al (2016) Aptamers: a promising chemical antibody for cancer therapy. Oncotarget 7(12):13446–13463

    Article  Google Scholar 

  201. Zhou J, Rossi J (2017) Aptamers as targeted therapeutics: current potential and challenges. Nat Rev Drug Discov 16(3):181–202. https://doi.org/10.1038/nrd.2016.199

    CAS  Article  PubMed  Google Scholar 

  202. Zhou Y, Ray D, Zhao Y et al (2007) Structure and function of Flavivirus NS5 methyltransferase. J Virol 81(8):3891–3903. https://doi.org/10.1128/JVI.02704-06

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  203. Zhu Q, Liu G, Kai M (2015) DNA aptamers in the diagnosis and treatment of human diseases. Molecules 20(12):20979–20997. https://doi.org/10.3390/molecules201219739

    CAS  Article  PubMed  PubMed Central  Google Scholar 

  204. Zmurko J, Neyts J, Dallmeier K (2015) Flaviviral NS4b, chameleon and jack-in-the-box roles in viral replication and pathogenesis, and a molecular target for antiviral intervention. Rev Med Virol 25(4):205–223. https://doi.org/10.1002/rmv.1835

    CAS  Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Citartan M and Tang TH were supported by USM Research University Grant (1001.CIPPT.8011095). Navien TN was supported by USM Fellowship (IPS/Fellowship2019/IPG).

Funding

Citartan M and Tang TH were supported by USM Research University Grant (1001.CIPPT.8011095). Navien TN was supported by USM Fellowship (IPS/Fellowship2019/IPG).

Author information

Affiliations

Authors

Contributions

TNN: Conceptualization, Original Draft Writing, Review and Editing. TSY: Writing. AA: Writing. THT: Review, Editing and Supervision. MC: Conceptualization, Review, Editing and Supervision.

Corresponding authors

Correspondence to Thean-Hock Tang or Marimuthu Citartan.

Ethics declarations

Conflicts of interest

Author’s declared to not have any conflict of interest.

Consent to participate

The authors hereby consent to participate.

Ethical approval

Material submitted is original; all authors are in agreement to have the article published.

Additional information

Publisher's Note

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

Rights and permissions

Reprints and Permissions

About this article

Verify currency and authenticity via CrossMark

Cite this article

Navien, T.N., Yeoh, T.S., Anna, A. et al. Aptamers isolated against mosquito-borne pathogens. World J Microbiol Biotechnol 37, 131 (2021). https://doi.org/10.1007/s11274-021-03097-0

Download citation

Keywords

  • Aptamer
  • Mosquito
  • Pathogen
  • Disease
  • SELEX