Open access peer-reviewed chapter

Aptamers for Diagnostics with Applications for Infectious Diseases

Written By

Muslum Ilgu, Rezzan Fazlioglu, Meric Ozturk, Yasemin Ozsurekci and Marit Nilsen-Hamilton

Submitted: 31 August 2018 Reviewed: 31 January 2019 Published: 10 April 2019

DOI: 10.5772/intechopen.84867

From the Edited Volume

Recent Advances in Analytical Chemistry

Edited by Muharrem Ince and Olcay Kaplan Ince

Chapter metrics overview

2,181 Chapter Downloads

View Full Metrics

Abstract

Aptamers are in vitro selected oligonucleotides (DNA, RNA, oligos with modified nucleotides) that can have high affinity and specificity for a broad range of potential targets with high affinity and specificity. Here we focus on their applications as biosensors in the diagnostic field, although they can also be used as therapeutic agents. A small number of peptide aptamers have also been identified. In analytical settings, aptamers have the potential to extend the limit of current techniques as they offer many advantages over antibodies and can be used for real-time biomarker detection, cancer clinical testing, and detection of infectious microorganisms and viruses. Once optimized and validated, aptasensor technologies are expected to be highly beneficial to clinicians by providing a larger range and more rapid output of diagnostic readings than current technologies and support personalized medicine and faster implementation of optimal treatments.

Keywords

  • aptamer
  • SELEX
  • biosensors
  • aptasensors
  • diagnostics
  • infectious diseases

1. Introduction

In 1868, a young Swiss physician Friedrich Miescher isolated a new biological compound from nuclei of white blood cells, which had not been described before. He named it “nuclein” [1]. Today, it is known as “deoxyribonucleic acid (DNA)” which is nucleic acid in nature and carries heritable information for biological organisms. However, with advancements in the molecular genetics field, scientists started to discover new functions of nucleic acids other than storing and transferring genetic information [2].

In the 1980s, research on human immunodeficiency virus (HIV) and adenoviruses revealed a new understanding of the importance of selective interactions between nucleic acids and proteins. These studies demonstrated that viruses express small RNAs, which bind to cellular or viral proteins with high specificity. In parallel with these discoveries, scientists focused on deciphering fundamental features of short RNAs that can fold into unique three-dimensional structures [3]. In 1990, three separate groups [4, 5, 6] invented the systematic evolution of ligands by exponential enrichment (SELEX) method by which they obtained nucleic acid molecules, similar to naturally occurring nucleic acids, which have high specificities and affinities toward their targets. They named these in vitro selected molecules “aptamers.” Since then, researchers have selected numerous aptamers against targets varying from small molecules to cells using SELEX [3].

1.1 Systematic evolution of ligands by exponential enrichment

SELEX is a technique to isolate an aptamer that is specific for the desired target from a randomized oligonucleotide (oligo) library by simulating evolution (systematic evolution of ligands). This in vitro technique includes a number of selection rounds (between 5 and 20) alternated with exponential amplification of the fittest oligonucleotides by PCR for DNA libraries or RT-PCR for RNA libraries (exponential enrichment).

In a typical SELEX experiment, the starting pool contains up to 1015 random oligonucleotide sequences. These sequences in the pool have a unique three-dimensional (3D) structures defined by the combination of interactions that include base pairing, stacking, sugar packing, and noncanonical intramolecular interactions. This structural complexity in the pool establishes a high probability of selecting an oligo that can interact avidly and specifically with the target of interest (aptamer). The intermolecular interaction between aptamer and target may include hydrogen bonds, salt bridges, van der Waals, and hydrophobic and electrostatic interactions.

Traditionally, SELEX is comprised of three main steps: incubation, separation, and amplification. The process involves incubating a “library” of oligos with randomized internal sequences of 20–60 nt with target molecules for a chosen period of time. Removal of unbound oligos from this mixture completes this initial step. Oligos that remain bound to the target are separated and amplified either by PCR or RT-PCR depending on the oligo type, DNA, or RNA. For DNA SELEX, biotin-labeled primers can be used in PCR, and the resulting double-stranded forms are separated using methods such as streptavidin bead capture. For RNA, T7 RNA polymerase promoter-containing primers are used in RT-PCR after which RNA libraries are amplified by in vitro transcription. The protocol follows the same amplification steps for several rounds (4–20) as for DNA SELEX (Figure 1). Negative selection can be included in either RNA or DNA SELEX protocols, which is achieved by passing the nucleic acid pool over a supporting matrix in the absence of the target. This step aims at eliminating the oligos that bind the matrix in a target-independent manner. Analogs of the target can also be included during selection rounds as competitors of binding if there is a means of separating the target from the analog competitors. This competition is expected to result in aptamers with higher specificity for the target over the analogs.

Figure 1.

Schematic representation of SELEX protocol.

Two alternative protocols are followed after these rounds to complete the SELEX. The selected oligos from the final round are cloned and sequenced for aptamer identification. Alternatively, high-throughput sequencing (or next-generation sequencing) can be employed to obtain sequence data from oligos present in the pool after different rounds of selection. Comparative sequence analysis allows pinpointing consensus sites that are potentially involved in target recognition [7, 8]. The most promising sequences are synthesized and characterized further, among which aptamers with nanomolar dissociation constants are frequently identified. In some instances, aptamers have also been isolated with picomolar dissociation constants.

In vitro selection can take months. To shorten the time for selecting high affinity aptamers, Golden et al. suggested a new method for SELEX: photochemical SELEX (PhotoSELEX) [9]. The technique involves the evolution of modified DNA aptamers, which are capable of forming a photoinduced covalent bond with their targets. Thus, these aptamers have greater specificity, and fewer selection rounds are required to select aptamers compared with the traditional SELEX methodologies.

1.2 Aptamer structure

Aptamer selection directly depends on environmental components during SELEX. This is because the ionic components and pH of the environment can dictate the predominant structures of oligos in the pool. Nucleic acids are negatively charged molecules that create an “ionic atmosphere” for ion-nucleic acid interactions, a freely joined sheath of ions surrounding the nucleic acids. These electrostatic interactions directly affect the structure of nucleic acids and thus the target binding by changing charge distribution. For specific ligand binding, it is crucial that the aptamer reliably forms the appropriate three-dimensional (3D) structure. However, structures of short oligonucleotides like aptamers are affected by the incubation temperature and the components of the operating buffer system such as the specific ions, ionic strength, and pH. Therefore, the affinities and future performance of aptamers depend on the buffers used during aptamer selection, and the choice of buffer present during selection requires attention when developing a SELEX protocol [10, 11, 12, 13].

Some factors present in the samples to be analyzed may have a negative impact on aptamer performance. For example, aptamers are susceptible to nucleases that are present in many biological samples. This is particularly true for RNA, having the 2'OH group, which can electrophilically attack the phosphate of the nucleic acid backbone. Nucleases promote this chemical property to catalyze hydrolysis of RNA, and this property also makes RNA more chemically labile to high pH and temperature compared with DNA. To counter this susceptibility to hydrolysis and to stabilize nucleic acids, many post-selection chemical modifications can be made. However, incorporating these modifications into the aptamer after its selection carries a large risk of altering the aptamer structure with a resulting loss of affinity for the target analyte. Alternatively, the less risky approach is to use chemically substituted nucleotide analogs during SELEX.

Secondary motifs in the tertiary structures of aptamers are diverse. Such motifs include the “stem-loop,” “hairpin structure,” “pseudoknot,” “internal bulge,” “kissing loop,” “three-way junction,” and “the G-quadruplex.” To understand these structures in detail, X-ray crystallography or nuclear magnetic resonance spectroscopy is utilized. But these techniques are laborious and expensive. To ease the process, computer algorithms have been developed to estimate the lowest free-energy structures using sequence-based modeling [14]. This makes it possible to quickly predict the secondary structure of oligonucleotides without needing many resources. However, computational approaches to obtaining 3D models of nucleic acid structures by modeling from primary sequence are still in the development phase, and the results are not as reliable as experimental methods [15, 16, 17].

Aptamers have a significant advantage over antibodies as components of sensing units. One advantage is that nucleic acid structures can be regenerated several times with little activity loss, whereas protein-based antibodies can only be used once or a few times before their functionality is lost. In contrast to antibodies or enzymes, nucleic acid aptamers are often highly stable and can be inexpensively synthesized with high reproducibility and purity. Like antibodies, they bind their targets with high affinity and specificity (Table 1). These properties are motivating the current flood of reports of aptamer-based biosensors employing a wide range of technologies.

Table 1.

Comparison of aptamers with antibodies.

1.3 Aptamers for diagnostic and therapeutic applications

Development of novel biosensors for various clinical diseases has become essential as new health issues emerged. To meet this goal, antibodies have been used extensively; however, more recently, aptamers have been recognized as promising alternatives for developing diagnostic devices.

A biosensor is a tool with the ability to provide a measurable signal as a result of biomolecular interactions. Biosensors generally consist of two components: a bioreceptor and a transducer. The bioreceptor binds specifically to the molecule of interest, and the transducer turns information from the binding event into a detectable signal. The components of the transducer include a detector and a reporter, which acts as a bridge between the bioreceptor and the detector.

The most critical part of the biosensor is the bioreceptor (or bio-recognition element). The success of the sensor directly depends on its high affinity and specificity. Because of their more flexible structures, aptamers can provide a substantial signal in combination with a larger number of detection methods than possible for antibodies. Aptamers are preferred over antibodies for biosensor applications because of their cost, stability, and reusability as well as the aforementioned advantages (see Table 1). Biosensors, called “aptasensors,” have been developed with many detectors including electrochemical, optical, microcantilever, and acoustic detectors [10].

Aptamers are also used in therapeutic applications, which will not be extensively discussed in this chapter. For discussions of therapeutic applications, the reader is directed to reviews that focus on this topic [18, 19, 20, 21]. The majority of therapeutic aptamers inhibit their target molecules, and some act as receptor agonists. Potential therapeutic aptamers against proteins including nucleolin, chemokine ligand 12, or thrombin have been described. Several RNA and DNA aptamers are undergoing clinical trials, yet only pegaptanib against vascular endothelial growth factor has so far been in the USA by the FDA for the treatment of vascular ocular disease [3].

To advance therapeutic applications, drug delivery systems with aptamers have been developed. For such applications, aptamers are needed that recognize cell surface proteins, which can be challenging to select because these proteins are difficult to purify in their natural conformations. The development of cell-based selection techniques has enabled aptamer selection against proteins in their native form while they are on the cell surface. These selections are performed against single-cell types [22] to obtain aptamers with the ability to bind cell surface proteins specifically expressed on the surface of the cell type used for selection. Delivery systems with these aptamers can carry a variety of cargos into cells by taking advantage of the surface protein internalization in response to aptamer-receptor binding.

Advertisement

2. Aptamer use in diagnostics

Diagnostics is one of the most dynamic fields in biosensor research. To support early diagnosis and individualized medicine, researchers seek to develop methods to detect identified biomarkers by more sensitive, time-saving, and cheaper methods. The perfect sensor for medical diagnostics should also be specific, reusable, easy to monitor, nonreactive, and stable with various biological samples. Aptamers can be developed to meet all these requirements. With their wide spectrum of possible targets, sensitive detection of virtually all toxins, drugs, peptides, proteins, metabolites, biomarkers, and cells is possible with aptamers.

Several types of aptasensors have been developed based on electrochemical, optical, mechanical, and acoustic approaches. An early use of aptamers as bio-recognition elements in aptasensors was reported in 1996 with an optical biosensor that utilized fluorescently labeled aptamers in a homogenous assay [23]. Later, aptamers were integrated onto solid supports, which provided an opportunity for real-time analyte detection. Most of these studies have been at the proof-of-principle level, and the majority of studies have been performed with the thrombin aptamer (TA), which has the advantage of a stable target and aptamer. Thrombin aptamer self-assembles into a highly stable G-quadruplex structure, and thrombin is a structurally stable globular protein found in blood. Future studies to optimize other aptasensors with less stable aptamers and analytes in complex matrices are likely to be challenging.

Aptamers fold their flexible, single-stranded chains into 3D structures, which may change upon binding to their cognate target molecules. This structure-switching characteristic has been capitalized on in many aptamer applications. An early example used electrochemical sensing with aptamers immobilized on an electrode surface, and target binding is observed by measuring electrochemical current variations. This system utilized an amperometric sandwich assay combining TA on a gold electrode which was used to capture the TA-labeled glucose dehydrogenase (GDH) [24].

Several types of aptasensors have been developed based on electrochemical, optical, mechanical, and acoustic approaches, which are discussed in the following sections.

2.1 Electrochemical aptasensors

Electrochemical aptasensors are constructed by attaching an aptamer that carries a redox-active moiety to an electrode surface. Such aptasensors can make use of voltammetric (amperometric), potentiometric, conductometric, or impedimetric assays for analyte detection. In some formats, the aptamer is labeled with an electroactive group and a structural change in the aptamer upon binding to the target analyte and changes the distance of the electroactive group from the electrode surface, resulting in the switching “on/off” of the electrochemical signal. Measurement of changes in electrochemical features after target binding has been used to determine target concentration [24]. So far, electrochemical aptasensors have been reported for a wide range of targets including PDGF, thrombin and immunoglobulin E (IgE), cocaine, theophylline, adenosine, aminoglycosides, and adenosine triphosphate (ATP) and inorganic ions such as potassium (K+) [10].

In addition to their innate structural changes on binding their target molecules, other structural constraints can be applied to aptamers that result in signals for detection by electrochemical approaches. The most commonly used additional constraint is an oligonucleotide (either connected with the aptamer or separate) that is complementary to part of the aptamer and, upon hybridizing with the aptamer, constrains its structure to an inactive form. For example, a biotin-tagged DNA aptamer for zeatin was hybridized with a complimentary “assist DNA” to form a Y-type DNA structure. Avidin-modified alkaline phosphatase (ALP) was attached to this structure with two biotins at the terminals of DNA aptamer. In the presence of zeatin, this complex was disrupted which leads to a decrease in the oxidation signal from p-nitrophenol (PNP) produced by the catalytic effect of ALP. From this, zeatin concentration in the range of 50 pM–50 nM was selectively measured with a detection limit of 16.6 pM [25].

Redox-active methylene blue (MB) has been used as an aptamer label and electrotransfer communication agent with the electrode. Methylene blue enables the detection of changes in aptamer conformation upon target binding. For example, an MB-labeled TA was used to construct an aptasensor by its immobilization on an electrode. The flexible conformation of the aptamer enabled the electrotransfer from MB to the electrode. The structural change upon analyte binding shielded MB in a “signal-off” mode. However, this mode is a disadvantage for diagnostics because the amperometric response decreases as a result of the association of the target thrombin with the aptamer.

Several approaches have been taken to develop aptasensors that operate in a “signal-on” mode. As an example, the TA was modified with an electroactive ferrocene group as the redox label at one end and a thiol group at the other end. The electrical contact of the electrode with the ferrocene label was affected by the long, flexible aptamer chain. Thrombin binding stabilizes the aptamer’s G-quadruplex conformation, which brings the ferrocene group closer to the electrode. This close proximity enables electron transfer between the electro-active ferrocene units and the electrode, thus producing a positive signal in the presence of thrombin. A similar approach to creating signal-on electrochemical aptasensors utilized the conformational change in the cocaine aptamer that occurs on binding its target. In the absence of the target, the aptamer on the solid surface stays in a partially folded form as a three-way junction. Cocaine binding decreases the distance for electron transfer and thus increases the signal [26].

Demirkol et al. generated an electrochemical aptasensor to detect E. coli O157:H7. The electrode surfaces were modified by cysteamine via self-assembled monolayer formation. The carboxyl-functionalized quantum dots and aptamers were conjugated to cysteamine-modified gold electrodes [27]. Ge et al. reported an affinity-mediated homogeneous electrochemical aptasensor using graphene-modified glassy carbon electrode (GCE) as the sensing platform. In this approach, the aptamer-target recognition is converted into an ultrasensitive electrochemical signal output with the aid of a novel T7 exonuclease (T7Exo)-assisted target-analog recycling amplification strategy, in which ingeniously designed methylene blue (MB)-labeled hairpin DNA reporters are digested in the presence of target and, then, converted to numerous MB-labeled long ssDNAs. The distinct difference in differential pulse voltammetry response between the designed hairpin reporters and the generated long ssDNAs on the graphene/GCE allows ultrasensitive detection of target biomolecules [28]. Lai et al. proposed a renewable electrochemical aptasensor for super sensitive Hg2+ determination [29]. The novel aptasensor, based on sulfur-nitrogen co-doped ordered mesoporous carbon (SN-OMC) and a thymine-Hg2+-thymine (T-Hg2+-T) mismatch structure, used ferrocene as signal molecules to achieve the conversion of signal to current. In the absence of Hg2+, the thiol-modified T-rich probe 1 spontaneously formed a hairpin structure by base pairing. After hybridizing with the ferrocene-labeled probe 2 in the presence of Hg2+, the hairpin structure of probe 1 was opened due to the preferential formation of the T-Hg2+-T mismatch structure, and the ferrocene signal molecules approached the modified electrode surface. Sulfur-nitrogen co-doped ordered mesoporous carbon with high specific surface area and ample active sites acted as a signal amplification element in electrochemical sensing. The sensitive determination of Hg2+ can be actualized by analyzing the relationship between the change of oxidation current caused by ferrocene and the Hg2+ concentrations [29]. Finally, Wang et al. combined the strengths of advanced aptamer technology, DNA-based nanostructure, and portable electrochemical devices to develop a nanotetrahedron (NTH)-assisted aptasensor for direct capture and detection of hepatocellular exosomes. The oriented immobilization of aptamers significantly improved their accessibility to suspended exosomes, and the NTH-assisted aptasensor could detect exosomes with 100-fold higher sensitivity when compared to the single-stranded aptamer-functionalized aptasensor [30].

Recently, nanoporous metal surfaces have been found as good sensor platforms for aptamers. Nanoporous gold [31, 32]-based sensors have been used with the TA and ATP aptamers and a redox probe to provide the electrons to the gold surface for sensitive detection of analyte by electrochemical impedance spectroscopy (EIS). The ATP aptamer was in a split format with the second half of the aptamer covalently linked with 3,4-diaminobenzoic acid (DABA), which created the EIS signal by its oxidation at the gold interface. For an analyte that can undergo redox reactions, such as bisphenol A (BPA), this property can be used to provide a signal [33]. Other nanoporous surfaces such as graphene oxide/Au composites and porous PtFe or PtTiAl ternary alloys have also been employed to measure breast cancer cells using the MUC-1 aptamer linked with the electroactive label thionine [34] or kanamycin with [Fe(CN)6]3−/4− [35]. The ability of aptamers to hybridize with other oligonucleotides was employed to create molecular gates over the pores in nanoporous gold surfaces. The gate, created with an aptamer highly specific for the avian influenza viruses (AIV) H5N1 hybridized to oligos linked to the nanoporous gold surface, was closed in the absence of AIV H5N1 but open when the virus bound and released the aptamer. The open pores allowed the entry of substrate and cofactor for lactate dehydrogenase, layered on a glassy electrode below the nanoporous gold. Cyclic voltammetry was used to detect the gold-catalyzed oxidation of the NADH produced as a result of LDH activity [34].

Nanoporous anodized aluminum oxide surfaces have more recently been used for providing a nanoporous surface through which electron movement can be controlled by aptamer-analyte binding. In these aptasensors, the aptamers are attached to a gold surface, which is provided by 2 nm gold nanoparticles [36] or by a surface coating created by sputtering [37]. The structural change in the aptamer due to binding of the analyte reduces access of [Fe(CN)6] 3−/4− to the gold surface [36], which can be measured by EIS. Even in the absence of a redox probe, a good EIS signal can be obtained due to a combination of steric hindrance and change in electrical conductance around the pores resulting from the structural changes that occur in the highly negatively charged aptamers upon binding their targets [37].

2.2 Optical aptasensors

Another type of biosensor that utilizes aptamers as bio-recognition elements is the optical sensor, for which fluorescent and colorimetric assays are the two widely used formats. In general, fluorescent detection is preferred due to its suitability for real-time detection and because there are many available labeling options as fluorophores and quenchers, which can easily be incorporated during aptamer synthesis.

To convert aptamers into fluorescent signaling probes, several strategies have been developed. A frequently used format places an aptamer sequence in a molecular beacon-like, hairpin structure in which ends are labeled either with two fluorophores or a fluorophore and a quencher. This system utilizes Förster resonance energy transfer (FRET), which relies on the energy transfer between donor and acceptor. Upon target binding, the structure is disrupted by separating the two ends, thus leading to a fluorescence signal. In this format, the use of organic fluorescent dyes or quantum dots (QDs) improved the assay performance and could also be used to detect drug delivery in cells. Another format places a fluorophore-labeled aptamer in a duplex structure with a complementary DNA sequence labeled with a quencher. The aptamer target successfully competes with the complementary DNA resulting in departure of the complementary strand from the aptamer and an accompanying increase in the fluorescence signal [26].

In optical analysis, simultaneous detection of several analytes is readily achieved by multiplexing. In one of the earliest examples, fluorescently labeled aptamers were immobilized on a glass surface. In this system, detection of thrombin and three cancer-biomarker proteins, inosine monophosphate dehydrogenase, vascular endothelial growth factor, and basic fibroblast growth factor, was achieved by fluorescence polarization even in the presence of human serum and E. coli cell lysates [10]. The use of aptamer-linked beads in a microarray setting brought a further sophistication of an “electronic tongue” that consists of a fluid delivery system and a fluorescence microscope attached to a digital camera for quantification [10].

Graphene oxide-based aptasensors can also be readily multiplexed [10]. As an example, a novel label-free fluorescent approach was constructed for H1N1 detection based on graphene-oxide and strand displacement reaction, using SYBR Green I (SGI) for signal amplification [38]. Another example is an assay for detection of the pathogenic bacterium, Pseudomonas aeruginosa. This assay was enabled by highly specific aptamers conjugated with photoluminescent carbon dots as the fluorescent probe and graphene oxide as the quencher, and it allowed detection of as low as 9 CFU mL−1P. aeruginosa [39]. Electrochemiluminescence is another output option for aptasensors [40].

2.3 Microcantilever aptasensors

Incorporating aptamers into microcantilever sensors offers the possibility of label-free target detection, low noise, high scalability, and small testing volumes [10]. Microcantilever-based sensing has been incorporated into several experimental chemical and biological sensing systems due to its small size, low cost, low sample volume, label-free detection, and ease of integration with microfluidic devices [41]. High-throughput analysis is achievable via microcantilever arrays for parallel processing, although they cannot easily be extensively multiplexed. Microcantilever aptasensors can be operated in either static or dynamic mode. In a liquid environment, the static mode can be more sensitive compared to the dynamic mode. In the static mode, one side of the microcantilever is functionalized with aptamers for analyte detection. Surface stress is generated when target analyte adsorbs onto the functionalized surface. The difference in surface stress between the top and bottom surface results in microcantilever bending, which can be upward (positive) or downward (negative) depending on the type of molecular interactions involved. Displacement of the beam can be detected by using readout techniques such as optical, piezoresistive, and capacitive. The optical technique is the most popular approach because it has high resolution and linear response and produces absolute displacement measurement. Detection of proteins and small molecule analytes, with an aptamer-decorated cantilever, can be achieved with many sensing means including interferometry [42, 43, 44] and piezoresistivity [45, 46]. Further development of microcantilever devices will require solutions to their sensitivity to vibration and the limits to which they can be multiplexed.

2.4 Acoustic aptasensors

Early work on acoustic aptasensors included the modification of gold-coated quartz crystals with aptamers. Target binding changes the frequency or phase shift which can be detected as a change in the input and output light. For example, the IgE DNA aptamer on a quartz crystal microbalance (QCM) format provided a detection limit of 3.3 ng.cm−2 of IgE [47].

Label-free and real-time quantification of proteins were also measured by the propagation of the acoustic wave in a surface acoustic wave (SAW) biosensor, which included an array of five sensor elements to detect human α-thrombin or HIV-1 Rev peptide. This system had a detection limit of 75 pg.cm−2 for both α-thrombin and HIV-1 Rev peptide as analytes [48]. The aptamers demonstrated a better linear response, stability, and reusability when compared with antibodies specific for IgE.

Surface plasmon resonance (SPR) sensors, similar to QCM and SAW aptasensors, rely on a change in refractive index due to target binding. Quartz crystal microbalance and surface plasmon resonance aptasensors for detection of HIV-1 Tat protein were found to have similar high specificities with the SPR sensor having a wider linear range [49]. The SPR aptasensor for retinol binding protein-4 was found more sensitive than an ELISA [50]. In developing SPR aptasensors, combined approaches resulted in a microfluidic device with interdigitated transducer creating high-frequency acoustic waves for target separation [51]. Biotinylated-thrombin aptamers were captured by streptavidin-functionalized polystyrene, which was pumped in the microchannel after incubating with sample. In the microfluidic device, SAW exposure leads to separation of thrombin captured in the polystyrene by its aptamer from the nontarget serum proteins.

Gold nanoparticles (AuNPs) have also been utilized in developing SPR aptasensors. A U-shaped fiber-optic SPR biosensor was developed for the rapid detection of BPA [52]. Incubation of bare AuNPs with BPA aptamer resulted in AuNPs/ssDNA complexes which are stable in high salt. Bisphenol A binding disrupted this complex, which resulted in the aggregation of AuNPs and enhanced refractive index of the solution in the fiber-optic SPR sensor. This system had a detection limit of 3.7 pg.mL−1, and linear range was 0.01–50 ng.mL−1.

Neves et al. generated two sensitive cocaine aptasensors that rely on an electromagnetic piezoelectric acoustic sensor (EMPAS) platform as the basis of ultra-high frequency with tuned signal-to-noise ratio [53]. The sensing interface consists of a S-(11-trichlorosilyl-undecanyl) benzenethiosulfonate (BTS) adlayer-coated quartz disc onto which a structure-switching cocaine aptamer was immobilized, completing the preparation of the MN4 cocaine aptamer with an apparent Kd of 45 ± 12 μM and limit of detection of 0.9 μM. The same group developed an MN6 cocaine aptasensor using an EMPAS platform that had apparent Kd of 27 ± 6 and a 0.3 μM detection limit [54].

Detection of cells using SPRs has some limitations that have been creatively overcome. The first limitation includes nonselective binding that causes the refractive index changes, which can be circumvented by reference flow cells to offset this effect [55]. Second is the sensing range, which is typically around 200 nm compared with cell dimensions that are in the micron range. Using long-range SPR, the depth was increased over 800 nm, which increased the sensitivity for cell detection [56]. Another drawback is its low-throughput, which has been resolved by SPR imaging technology.

Advertisement

3. Clinical perspective

As discussed throughout this chapter, aptamers can be evolved to have high affinity and specificity for a range of target molecules that includes small organics, peptides, protein, sugar, viruses, bacteria, parasites, live cell, and tissue (Tables 2 and 3). This characteristic paves the way for aptamers to have applications in various disciplines including sensing, medicine, pharmacology, and microbiology. Aptamer-based sensors have great promise as effective tools in the areas of diagnostics and therapeutics for clinical use [164, 165].

OrganismTargetAptamerBackboneBinding affinity (Kd)Reference
M. tb Beijing strainManLAMT9DNA668 ± 59 nM[57]
M. tb H37RvCE proteinCE24DNA0.375 μM[58]
CE proteinCE15DNA0.16 μM
M. tbCE proteinCSIR 2.11DNA[59]
MPT64MPT64-A1DNA[60, 61]
EsxG proteinG43RNA8.04 ± 1.90 nM[62]
EsxG proteinG78RNA78.85 ± 9.40 nM
Whole bacteriumMA1DNA12.02 nM[63]
Whole bacteriumAptamer 1DNA37 ± 4 nM[64]
S. typhimuriumOuter membrane proteinsAptamers 33 and 45DNA[65, 66]
OmpC proteinI-2RNA20 nM[67]
Whole bacteriumC4DNA[68]
Whole bacteriumST2PDNA6.33 ± 0.58 nM[69]
Whole bacteriumSAL 26DNA123 ± 23 nM[70]
Salmonella Paratyphi AWhole bacteriumApt22DNA47 ± 3 nM[71]
S. enteritidisMixtures of 10 strains of S. enteritidisS25RNA[72]
Whole bacteriumcrn1DNA0.971 μM[73]
Whole bacteriumcrn2DNA0.309 μM
S. aureusS. Endotoxin BAPTSEB1DNA[74]
S. Endotoxin C1C10DNA65.14 ± 11.64 nM.L−1[75]
Alpha toxinR12.06DNA93.7 ± 7.0 nM[76]
PeptidoglycanAntibac1DNA0.415 ± 0.047 μM[77]
PeptidoglycanAntibac2DNA1.261 ± 0.280 μM
Protein APA#2/8DNA172 ± 14 nM for the recombinant Protein A and 84 ± 5 nM for the native Protein A[78]
Whole bacteriumSA20DNA70.68 ± 39.22 nM[79, 80]
Whole bacteriumSA23DNA61.50 ± 22.43 nM
Whole bacteriumSA31DNA82.86 ± 33.20 nM
Whole bacteriumSA34DNA72.42 ± 35.23 nM
Whole bacteriumSA43DNA210.70 ± 135.91 nM
Whole bacteriumSA17DNA35 nM[81]
Whole bacteriumSA61DNA129 nM
L. monocytogenesInternalin AA8DNA[82]
Listeriolysin OLLO-3DNA[83]
Whole bacteriumLbi-17DNA[84]
Whole bacteriumLMCA2DNA2.01x10−12 M[85]
Whole bacteriumLMCA26DNA1.56x10−10 M
E. coli 0157LPSE-5, E-11, E-12, E-16 to E-19DNA[86]
E. coli 0158Whole bacteriumAM-6DNA107.6 ± 67.8 pmol[87]
E. coli K88Whole bacteriumApt B12DNA15 ± 4 nM[88]
E. coli KCTC 2571Whole bacteriumE1DNA12.4 nM[35, 89]
Whole bacteriumE2DNA25.2 nM
Whole bacteriumE10DNA14.2 nM
Whole bacteriumE12DNA16.8 nM
E. coli DH5αWhole cellEc3(31)RNA225 nM[90]
Whole cell8.28ADNA27.4 ± 18.7 nM[91]
M. tb H37RvWhole bacteriumNK2DNA[92]
ManLAMZXL1DNA436.3 ± 37.84 nM[93]
BCGManLAMBM2DNA8.59 ± 1 .23 nM[94]
M. tbAcetohydroxyacid synthaseMtb-Apt1DNA1.06 ± 0.10 μM[95]
Acetohydroxyacid synthaseMtb-Apt6DNA0.210 ± 0.05 μM
S. typhiType IVB piliS-PS8.4RNA8.56 nM[96, 97]
S. aureusSEAS3DNA36.93 ± 7.29 nM[98]
α-toxinAT-33DNA[99]
α-toxinAT-36DNA

Table 2.

Aptamers selected against bacteria.

OrganismTargetAptamerBackboneBinding affinity (Kd)Reference
HIV-1TatRNATatRNA120 ± 13 pM[49, 100, 101]
HCVE2 proteinZE2DNA1.05 ± 1 nM[102]
Core protein9–14RNA142 nM[103]
Core protein9–15RNA224 nM
Core proteinC4DNA[104, 105]
Influenza A virus (H3N2)HA(91–261)A22DNA[106]
HAClone BRNA200 pM[107]
Whole virusP30-10-16RNA188 pM[108]
Influenza A virus (H5N1)HAA10DNA[109]
Influenza A virus (H9N2)HA(101–257)C7DNA[110]
Influenza A virus (H5N1)HAHAS15–5RNA[111]
Influenza A virus (H1N1)HAD-26RNA67 fM[112]
Influenza A virus (H3N2)HAHA68DNA7.1 nM[113]
Influenza A virus (H5N1) and (H7N7)HAs from H5N1 and H7N78-3RNA170 fM[114]
Influenza A virusHA1 subunitApIDNA64.76 ± 18.24 nM[115]
Influenza A virusHA1 subunitApIIDNA69.06 ± 12.34 nM
Influenza A virusHA1 subunitApIIIDNA50.32 ± 14.07 nM
Influenza A virus (H9N2)HAA9DNA46.23 ± 5.46 nM[116]
Influenza A virus (H9N2)HAB4DNA7.38 ± 1.09 nM[117]
Influenza A virus (H5N1) and (H5N8)Whole virusIF22 and IF23DNA[118]
HBVSurface antigenHBs-A22RNA[119]
HPV 16E7 proteinG5α3N.4RNA1.9 μM[120]
HIV-1Gp120B40RNA21 ± 2 nM[121]
Gp120B40t77RNA31 ± 2 nM
Gp120A-1RNA52 nM[122]
Gp120BclON-mutDNA143 ± 79 nM[123]
Gp120F-thio-BclONDNA86 ± 17 nM
RT1.1RNARNA5 nM[124]
RTRT1t49DNA1 nM[125]
RT4.20DNA180 ± 70 pM[126]
RTR12–2DNA70 nM[127]
RT37NTDNA660 pM[128]
RTFA1FANA aptamerLow pM range[129]
5’-UTR of HIV-1 genomeRNApt16RNA280 ± 60 nM[130]
TAR RNA elementIV04DNA20 nM[131]
IntegraseT30695DNA0.5 ± 0.2 μM[132, 133]
Integrase93delDNA[56, 134, 135]
Nucleocapsid protein8–6RNA1.4x10−9 M[136]
Gag proteinDP6–22RNA100 ± 3.4 nM[137]
Rev proteinRBE(apt)RNA[138]
IntegraseS3R3RNA47 ± 3 nM[139]
HCVNS3 proteinG6–16RNA238 nM[140]
Truncated protease domain of NS3 proteinG9-IRNA10 nM[141]
Helicase domain of NS3G5RNA25 nM[142]
IRES domains III-IV3–07RNA9 nM[143]
IRESAP50RNA5 nM[144, 145]
IRES domain IIIf and IVHH-11RNA[146]
NS5B27vDNA132.3 ± 20 nM[147, 148]
NS5Br10/43RNA1.3 ± 0.3 nM[149]
NS5Br10/47RNA23.5 ± 6.7 nM
NS5BR-F t2RNA2.62 ± 0.90 nM[85, 150]
Influenza A virus (H5N2)Glycosylated HAHA12–16RNA[151]
Influenza A virus (H1N1, H5N1, H7N7 and H7N9)Residues in the N-terminal of the PAN of the influenza A virus polymerasePAN-2DNA247 ± 11 nM[152]
HBVTruncated P proteinS9RNA[153]
Core proteinApt.No.28DNA[154]
CapsidAO-01DNA180 ± 82 nM[155]
HPV 16E7 proteinA2RNA107 nM[156, 157]
E6 proteinF2RNA[158]
SARS-CoVHelicaseNG8DNA5 nM[159]
DENV-2Envelope protein domain IIIS15DNA200 nM[160]
RABVGlycoproteinGE54DNA307 nM[161]
EBOVEBOV sGP39SGP1ARNA27 nM[162]
ZikaNS1 proteinClone 2DNA24 pM[163]
NS1 proteinClone 10DNA134 nM

Table 3.

Aptamers selected against viruses.

3.1 Clinical diagnostics

In clinical practice, the quantification of single biomarkers is frequently not sufficient to support a confident diagnosis. Thus, multiple analyses are required for each diagnosis, many of which might rely on antibodies, the current workhorses of the diagnostic world. Although many antibodies are highly sensitive and specific for their antigen targets, they often suffer from batch-to-batch variation. ELISA assays, which are the main diagnostic platform for antibodies, have been improved with protocols that increase their efficiency and the specificity of their output. However, ELISA assays cannot be readily multiplexed to measure simultaneously the many biomarkers required for a confident diagnosis. A strength of aptamer biosensors is their ability to be multiplexed. Another strength is the range of technologies that can be applied to produce operating aptasensors [166, 167, 168]. With these strengths to drive it, aptasensor technology is one of the fastest growing biotechnology areas in diagnostics with an expectation of reaching about $250 million by 2020 [169, 170].

Aptamers can provide new opportunities for medical diagnostics beyond what is available with antibodies [18, 171]. For example, aptamers can be selected against non-immunogenic and toxic targets, to which antibodies cannot be elicited. These short, single-stranded oligonucleotides can be synthesized via simple chemical synthesis, making them easier and less costly to produce than antibodies [172]. When synthesized in cells containing DNAs encoding RNA aptamers, they can fold appropriately and recognize intracellular targets [173, 174, 175].

A limited number of aptasensors are in the pipeline to be used in the areas such as biomarker and microorganism detection and cancer clinical testing. The available aptasensors include (1) OTA-Sense for detection of Ochratoxin A (OTA), a toxin produced by fungi [165, 176]; (2) AflaSense for detection of aflatoxins [95, 177, 178, 179]; (3) AptoCyto for flow cytometry applications [165]; (4) AptoPrep as a kit including conjugated aptamers specific to CD-31, EGFR, HGFR, and ICAM-2 [165]; (5) SOMAscan as a platform with the ability to detect >1300 proteins from small volumes [180, 181, 182, 183, 184]; (6) an aptamer-based proteomics technology developed by Jung et al. for detecting non-small cell lung cancer [165]; and (7) OLIGOBIND for measuring the thrombin level in blood [185].

Highly sensitive tests are required to specifically detect cancer cells in body fluids over all others. Success in this effort requires the identification of biomarkers that are found only on tumor cells. Some aptamers have been identified that might be applied to detecting tumor cells in the blood. For example, cancer cells and normal cells have been distinguished by using electrochemical sensors, a SERS active bimetallic core-satellite nanostructure, porphyrin-based covalent organic framework based aptasensor [186], and a deterministic lateral displacement (DLD) pattern-based aptamer-tailed octopus chip [187]. Recently, some aptasensors have been tested in cancer studies. Prostate-specific antigen (PSA), mucin 1 (MUC1), PDGF-BB, and vascular endothelial growth factor (VEGF) are detected as cancer biomarkers in cancer cell lines [165, 188].

Many conventional diagnostic technologies for detecting virus and bacteria, including serologic-, nucleic acid-, and culture-based tests, are either time-consuming or expensive on account of the need for sophisticated equipment [188]. For example, the gold standard for laboratory diagnosis of acute viral infections is isolation and characterization of the virus or bacterium. Isolation and long replication times for some viruses and bacterial strains can delay confirmation of the initial diagnosis for more than a week. The most commonly used alternative method is the ELISA. However, cross-reactive antibodies against viruses, particularly when they are part of the same virus family, may confound the results of serologic tests and may lead to misinterpretation during the epidemiologic assessment in regions of the world where they are co-endemic [189]. For instance, among the flaviviruses, serological cross-reactivity between Zika virus and dengue virus confounds diagnosis of Zika virus infections in pregnant woman in regions where Dengue virus is also endemic [190]. Additionally, ELISA-detecting antibodies (IgG and IgM) that were produced against the virus do not identify the active infection or the virus particles. Because they can interact with different regions of the protein compared with antibodies, aptamers might be capable of distinguishing viruses that cannot be distinguished serologically [162]. Aptamers that have been reported to specifically bind flaviviruses and their protein products including Ebola [162], Zika [163], and dengue virus [160] should be tested for their abilities to distinguish these viruses.

The abilities to detect, identify, and quantify microbes and viruses and to identify virally infected cells are essential for their early diagnosis. An increasing number of aptamers have been isolated that bind specific microbes such as Escherichia coli, Bacillus thuringiensis, Campylobacter jejuni, and Campylobacter coli [191], various salmonella species, including S. enteritidis [72] and S. enterica [65], and staphylococcus species including S. aureus [79], S. typhimurium [67], and S. enteritidis [192] (Table 2). In addition to the aptamers recognizing flaviviruses discussed in the previous paragraph, there are also aptamers that bind HIV-1 and hepatitis C virus (HCV) [188], among others (Table 3).

Aptasensors that have been developed using a number of the aptamers just mentioned hold the promise of prompt management of infections with a decreasing incidence of morbidity. Another possible application of aptasensors is in vaccine development. The costs and extended time associated with the assessment of vaccine concentrations by ELISA might be overcome by the application of aptasensors [193].

3.2 Future directions in clinics

Application of aptamer technology in the clinic has the potential of solving some stubborn diagnostic problems. Here we discuss some examples.

Tuberculosis (TB) is caused by Mycobacterium tuberculosis and is one of the top ten causes of death worldwide [194]. The incidence of childhood TB is reported as a half a million cases with 74,000 deaths annually by the World Health Organization [195]. The most commonly used diagnostic tool for tuberculosis is the TB skin test. However, a false-positive test result in people vaccinated with the bacillus Calmette-Guerin (BCG) vaccine can be a confounding factor for diagnosis. Thus, the distinction of infection from disease, particularly in children, is still unclear. Additionally, microbiologic confirmation of bacteria in body fluids in childhood is difficult because of the poor bacillary count [195, 196, 197].

Cytomegalovirus (CMV) causes life-threatening infections in patients with solid organ transplantation, hematopoietic stem cell transplantation, and AIDS. Cytomegalovirus causes acute and latent infections and reactivates in immunosuppressed patients. Isolation and infection control procedures as well as proper management of VHFs commonly depend on an accurate diagnosis [198]. Current diagnostic tests targeting CMV DNA and CMV antigens are insufficient for discriminating acute and latent infections and detecting organ involvement. As a result, the results of these tests are frequently misinterpreted [199, 200, 201]. Viral hemorrhagic fevers (VHFs) are caused by a couple of viruses including Arenaviridae, Bunyaviridae, Filoviridae, and Flaviviridae. Fulminant and fatal disease processes are the common features of the VHFs and diagnosing and distinguishing of VHF from other tropical diseases may be problematic because of the indiscriminatory symptoms [202].

An accurate and reliable diagnosis of these and other infections will provide for appropriate management and will decrease morbidity and mortality. Thus, development of fast microorganism-focused tests will provide rapid accurate diagnosis of infection and organ involvements. It seems inevitable that, in the near future, many aptamer-based methods/tools will provide for early diagnosis that will enable rapid initiation of optimal treatment regimens for viral and bacterial diseases.

Advertisement

4. Future perspective

First identified in 1890 [203], antibodies and their means of interaction with their antigens were already being extensively studied in the early twentieth century, and their structures were reported in the 1960s by Gerald Edelman [204]. Continuing studies of antibodies and the means of eliciting them have resulted in a detailed understanding of how they interact with their antigens. This knowledge combined with an expansive array of available antibodies motivates scientists to incorporate them into new diagnostic tools. Thus, antibody use dominates the 45 billion dollars global diagnostics market.

The more recently discovered aptamers are generally compared with antibodies due to their functional similarity. Since the discovery of aptamers in the 1990s, over a thousand studies have been conducted on applications of aptamers for diagnostics. Aptamers that specifically target biomarkers and bacterial or viral virulence factors such as surface glycoproteins or secreted proteins have been generated. These studies have demonstrated the range of targets that can be recognized by aptamers and the number of sensor platforms into which aptamers can be incorporated, many of which have been discussed in this chapter. However, the more structurally flexible aptamers are not as readily plugged into standard diagnostic assays as antibodies. In this burgeoning and yet immature field of aptasensors, there is still much to be learned about how to control aptamer behavior.

More systematic studies are needed to optimize selection methods, and more aptamers need to be characterized structurally. The biological matrix in which the analyte will be measured in the final application platform should be considered at the beginning of the selection so as to use buffer and ionic compositions during SELEX that resemble the target matrix. Maturation of aptamers to increase specificity for their target analyte in the appropriate matrix and for effective performance in the chosen reporter platform is also extremely important.

Aptamers offer the allure of easier production, ease of chemical modification, smaller size, reusability, stability even at high temperatures, low cost, and a long shelf life. A variety of chemical modifications further enhance aptamer stability. A significant advantage over antibody-based assays is that aptamers can be reused for many cycles without losing potency with the analyte being removed between each cycle by heating or other means. These features hold promise for the continued incorporation of aptamers into various sensor platforms and for the further development and eventual commercial application of aptasensors for diagnostics.

Advertisement

Acknowledgments

Authors are thankful to the administrative staff members at IntechOpen for their invaluable help.

Advertisement

Conflict of interest

MNH is the owner of Aptalogic Inc. located in Ames, IA, USA.

References

  1. 1. Dahm R. Friedrich Miescher and the discovery of DNA. Developmental Biology. 2005;278(2):274-288. DOI: 10.1016/j.ydbio.2004.11.028
  2. 2. Lipfert J, Doniach S, Das R, Herschlag D. Understanding nucleic acid-ion interactions. Annual Review of Biochemistry. 2014;83:813-841. DOI: 10.1146/annurev-biochem-060409-092720
  3. 3. Song KM, Lee S, Ban C. Aptamers and their biological applications. Sensors. 2012;12(1):612-631. DOI: 10.3390/s120100612
  4. 4. Tuerk C, Gold L. Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science. 1990;249(4968):505-510. DOI: 10.1126/science.2200121
  5. 5. Ellington AD, Szostak JW. In vitro selection of RNA molecules that bind specific ligands. Nature. 1990;346:818-822. DOI: 10.1038/346818a0
  6. 6. Robertson DL, Joyce GF. Selection in vitro of an RNA enzyme that specifically cleaves single-stranded DNA. Nature. 1990;344(6265):467-468. DOI: 10.1038/344467a0
  7. 7. Hoinka J, Zotenko E, Friedman A, Sauna ZE, Przytycka TM. Identification of sequence-structure RNA binding motifs for SELEX-derived aptamers. Bioinformatics. 2012;28(12):i215-223. DOI: 10.1093/bioinformatics/bts210
  8. 8. Hoinka J, Berezhnoy A, Sauna ZE, Gilboa E, Przytycka TM. AptaCluster—A method to cluster HT-SELEX aptamer pools and lessons from its application. Lecture Notes in Computer Science. 2014;8394(LNBI):115-128. DOI: 10.1007/978-3-319-05269-4_9
  9. 9. Tombelli S, Minunni M, Mascini M. Analytical applications of aptamers. Biosensors & Bioelectronics. 2005;20(12):2424-2434. DOI: 10.1016/j.bios.2004.11.006
  10. 10. Ilgu M, Nilsen-Hamilton M. Aptamers in analytics. The Analyst. 2016;141(5):1551-1568. DOI: 10.1039/c5an01824b
  11. 11. Yu Q , Dai Z, Wu W, Zhu H, Ji L. Effects of different buffers on the construction of aptamer sensors. IOP Conference Series: Materials Science and Engineering. 2017;274(1):012117. DOI: 10.1088/1757-899X/274/1/012117
  12. 12. Bruno JG, Sivils JC. Studies of DNA aptamer oligreen and picogreen fluorescence interactions in buffer and serum. Journal of Fluorescence. 2016;26(4):1479-1487. DOI: 10.1007/s10895-016-1840-1
  13. 13. Baaske P, Wienken CJ, Reineck P, Duhr S, Braun D. Optical thermophoresis for quantifying the buffer dependence of aptamer binding. Angewandte Chemie International Edition. 2010;49(12):2238-2241. DOI: 10.1002/anie.200903998
  14. 14. Radom F, Jurek PM, Mazurek MP, Otlewski J, Jeleń F. Aptamers: Molecules of great potential. Biotechnology Advances. 2013;31(8):1260-1274. DOI: 10.1016/j.biotechadv.2013.04.007
  15. 15. Parisien M, Major F. The MC-Fold and MC-Sym pipeline infers RNA structure from sequence data. Nature. 2008;452(7183):51-55. DOI: 10.1038/nature06684
  16. 16. Lu XJ, Bussemaker HJ, Olson WK. DSSR: An integrated software tool for dissecting the spatial structure of RNA. Nucleic Acids Research. 2015;43(21):e142. DOI: 10.1093/nar/gkv716
  17. 17. Chojnowski G, Waleń T, Bujnicki JM. RNA bricks—A database of RNA 3D motifs and their interactions. Nucleic Acids Research. 2014;42(Database issue):D123-D131. DOI: 10.1093/nar/gkt1084
  18. 18. Banerjee J, Nilsen-Hamilton M. Aptamers: Multifunctional molecules for biomedical research. Journal of Molecular Medicine. 2013;91(12):1333-1342. DOI: 10.1007/s00109-013-1085-2
  19. 19. Zhou G, Latchoumanin O, Bagdesar M, Hebbard L, Duan W, Liddle C, et al. Aptamer-based therapeutic approaches to target cancer stem cells. Theranostics. 2017;7(16):3948-3961. DOI: 10.7150/thno.20725
  20. 20. Esposito CL, Catuogno S, Condorelli G, Ungaro P, De Franciscis V. Aptamer chimeras for therapeutic delivery: The challenging perspectives. Genes (Basel). 2018;9(11):529. DOI: 10.3390/genes9110529
  21. 21. Ismail SI, Alshaer W. Therapeutic aptamers in discovery, preclinical and clinical stages. Advanced Drug Delivery Reviews. 2018;134:51-64. DOI: 10.1016/j.addr.2018.08.006
  22. 22. Chen M, Yu Y, Jiang F, Zhou J, Li Y, Liang C, et al. Development of cell-SELEX technology and its application in cancer diagnosis and therapy. International Journal of Molecular Sciences. 2016;17(12):2079. DOI: 10.3390/ijms17122079
  23. 23. Drolet DW, Moon-McDermott L, Romig TS. An enzyme-linked oligonucleotide assay. Nature Biotechnology. 1996;14(8):1021-1025. DOI: 10.1038/nbt0896-1021
  24. 24. Ikebukuro K, Kiyohara C, Sode K. Electrochemical detection of protein using a double aptamer sandwich. Analytical Letters. 2004;37(14):2901-2909. DOI: 10.1081/AL-200035778
  25. 25. Zhou Y, Yin H, Wang Y, Sui C, Wang M, Ai S. Electrochemical aptasensors for zeatin detection based on MoS2 nanosheets and enzymatic signal amplification. The Analyst. 2018:143(21):5185-5190. DOI: 10.1039/C8AN01356J
  26. 26. Song S, Wang L, Li J, Fan C, Zhao J. Aptamer-based biosensors. Trends in Analytical Chemistry. 2008;27(2):108-117. DOI: 10.1016/j.trac.2007.12.004
  27. 27. Odacı-Demirkol D, Güleç K. Electrochemical sensing and fluorescence imaging of E. coli O157:H7 based on aptamer-conjugated semiconducting nanoparticles. Süleyman Demirel Üniversitesi Fen Bilimleri Enstitüsü Dergisi. 2018;22:421-426
  28. 28. Ge L, Wang W, Sun X, Hou T, Li F. Affinity-mediated homogeneous electrochemical aptasensor on a graphene platform for ultrasensitive biomolecule detection via exonuclease-assisted target-analog recycling amplification. Analytical Chemistry. 2016;88(4):2212-2219. DOI: 10.1021/acs.analchem.5b03844
  29. 29. Lai C, Liu S, Zhang C, Zeng G, Huang D, Qin L, et al. An electrochemical aptasensor based on sulfur-nitrogen Co-doped ordered mesoporous carbon and thymine-Hg2+-thymine mismatch structure for Hg2+ detection. ACS Sensors. 2018;3(12):2566-2573. DOI: 10.1021/acssensors.8b00926
  30. 30. Wang S, Zhang L, Wan S, Cansiz S, Cui C, Liu Y, et al. Aptasensor with expanded nucleotide using DNA nanotetrahedra for electrochemical detection of cancerous exosomes. ACS Nano. 2017;11(4):3943-3949. DOI: 10.1021/acsnano.7b00373
  31. 31. Qiu H, Sun Y, Huang X, Qu Y. A sensitive nanoporous gold-based electrochemical aptasensor for thrombin detection. Colloids Surfaces B Biointerfaces. 2010;79(1):304-308. DOI: 10.1016/j.colsurfb.2010.04.017
  32. 32. Kashefi-Kheyrabadi L, Mehrgardi MA. Aptamer-based electrochemical biosensor for detection of adenosine triphosphate using a nanoporous gold platform. Bioelectrochemistry. 2013;94:47-52. DOI: 10.1016/j.bioelechem.2013.05.005
  33. 33. Zhu B, Alsager OA, Kumar S, Hodgkiss JM, Travas-Sejdic J. Label-free electrochemical aptasensor for femtomolar detection of 17β-estradiol. Biosensors & Bioelectronics. 2015;70:398-403. DOI: 10.1016/j.bios.2015.03.050
  34. 34. Yan M, Sun G, Liu F, Lu J, Jinghua Yu XS. Investigations on the interface of nucleic acid aptamers and binding targets. Analytica Chimica Acta. 2013;798:33-39. DOI: 10.1039/C8AN01467A
  35. 35. Guo Y, Wang Y, Liu S, Yu J, Wang H, Wang Y, et al. Label-free and highly sensitive electrochemical detection of E. coli based on rolling circle amplifications coupled peroxidase-mimicking DNAzyme amplification. Biosensors & Bioelectronics. 2016;75:315-319. DOI: 10.1016/j.bios.2015.08.031
  36. 36. Peinetti AS, Ceretti H, Mizrahi M, González GA, Ramírez SA, Requejo FG, et al. Confined gold nanoparticles enhance the detection of small molecules in label-free impedance aptasensors. Nanoscale. 2015;7(17):7763-7769. DOI: 10.1039/c5nr01429h
  37. 37. Gosai A, Hau Yeah BS, Nilsen-Hamilton M, Shrotriya P. Label free thrombin detection in presence of high concentration of albumin using an aptamer-functionalized nanoporous membrane. Biosensors & Bioelectronics. 2019;126:88-95. DOI: 10.1016/j.bios.2018.10.010
  38. 38. Feng X, Liu K, Ning Y, Chen L, Deng L. A label-free aptasensor for rapid detection of H1N1 virus based on graphene oxide and polymerase-aided signal amplification. Nanomedicine and Nanotechnology. 2015;6(3):288. DOI: 10.4172/2157-7439.1000288
  39. 39. Sadeghi AS, Ansari N, Ramezani M, Abnous K, Mohsenzadeh M, Taghdisi SM, et al. Optical and electrochemical aptasensors for the detection of amphenicols. Biosensors & Bioelectronics. 2018;118:137-152. DOI: 10.1016/j.bios.2018.07.045
  40. 40. Du X, Jiang D, Hao N, Qian J, Dai L, Zhou L, et al. Building a three-dimensional nano-bio interface for aptasensing: An analytical methodology based on steric hindrance initiated signal amplification effect. Analytical Chemistry. 2016;88(19):9622-9629. DOI: 10.1021/acs.analchem.6b02368
  41. 41. Luka G, Ahmadi A, Najjaran H, Alocilja E, Derosa M, Wolthers K, et al. Microfluidics integrated biosensors: A leading technology towards lab-on-A-chip and sensing applications. Sensors (Basel). 2015;15(12):30011-30031. DOI: 10.3390/s151229783
  42. 42. Kang K, Sachan A, Nilsen-Hamilton M, Shrotriya P. Aptamer functionalized microcantilever sensors for cocaine detection. Langmuir. 2011;27(23):14696-14702. DOI: 10.1021/la202067y
  43. 43. Kang K, Nilsen-Hamilton M, Shrotriya P. Differential surface stress sensor for detection of chemical and biological species. Applied Physics Letters. 2008;93(14):14-17. DOI: 10.1063/1.2996411
  44. 44. Zhai L, Wang T, Kang K, Zhao Y, Shrotriya P, Nilsen-Hamilton M. An RNA aptamer-based microcantilever sensor to detect the inflammatory marker, mouse lipocalin-2. Analytical Chemistry. 2012;84(20):8763-8770. DOI: 10.1021/ac3020643
  45. 45. Lim YC, Kouzani AZ, Duan W, Dai XJ, Kaynak, A, Mair D. Design and evaluation of a microcantilever aptasensor. In: ISCAS 2014: Proceedings of the 2014 IEEE International Symposium on Circuits and Systems, IEEE, Piscataway, NJ. 2014. pp. 221-224. DOI: 10.1109/ISCAS.2014.6865105
  46. 46. Zhao Y, Schagerl M, Viechtbauer C, Loh KJ. Characterizing the conductivity and enhancing the Piezoresistivity of carbon nanotube-polymeric thin films. Materials (Basel). 2017;10(7):724. DOI: 10.3390/ma10070724
  47. 47. Liss M, Petersen B, Wolf H, Prohaska E. An aptamer-based quartz crystal protein biosensor. Analytical Chemistry. 2002;74(17):4488-4495. DOI: 10.1021/ac011294p
  48. 48. Schlensog MD, Gronewold TMA, Tewes M, Famulok M, Quandt E. A love-wave biosensor using nucleic acids as ligands. Sensors & Actuators,B:Chemical. 2004;101(3):308-315. DOI: 10.1016/j.snb.2004.03.015
  49. 49. Tombelli S, Minunni M, Luzi E, Mascini M. Aptamer-based biosensors for the detection of HIV-1 Tat protein. Bioelectrochemistry. 2005;67(2 Spec Issue):135-141. DOI: 10.1016/j.bioelechem.2004.04.011
  50. 50. Wilson R, Cossins A, Nicolau DV, Missailidis S. The selection of DNA aptamers for two different epitopes of thrombin was not due to different partitioning methods. Nucleic Acid Therapeutics. 2013;23(1):88-92. DOI: 10.1089/nat.2012.0386
  51. 51. Ahmad Raston NH, Nguyen VT, Gu MB. A new lateral flow strip assay (LFSA) using a pair of aptamers for the detection of vaspin. Biosensors & Bioelectronics. 2017;93:21-25. DOI: 10.1016/j.bios.2016.11.061
  52. 52. Luo Z, Zhang J, Wang Y, Chen J, Li Y, Duan Y. An aptamer based method for small molecules detection through monitoring salt-induced AuNPs aggregation and surface plasmon resonance (SPR) detection. Sensors & Actuators, B:Chemical. 2016;236:474-479. DOI: 10.1016/j.snb.2016.06.035
  53. 53. Neves MADD, Blaszykowski C, Bokhari S, Thompson M. Ultra-high frequency piezoelectric aptasensor for the label-free detection of cocaine. Biosensors & Bioelectronics. 2015;72:383-392. DOI: 10.1016/j.bios.2015.05.038
  54. 54. Neves MAD, Blaszykowski C, Thompson M. Utilizing a key aptamer structure-switching mechanism for the ultrahigh frequency detection of cocaine. Analytical Chemistry. 2016;88(6):3098-3106. DOI: 10.1021/acs.analchem.5b04010
  55. 55. Cai S, Yan J, Xiong H, Liu Y, Peng D, Liu Z. Investigations on the interface of nucleic acid aptamers and binding targets. The Analyst. 2018;143(22):5317-5338. DOI: 10.1039/c8an01467a
  56. 56. Faure-Perraud A, Métifiot M, Reigadas S, Recordon-Pinson P, Parissi V, Ventura M, et al. The guanine-quadruplex aptamer 93del inhibits HIV-1 replication ex vivo by interfering with viral entry, reverse transcription and integration. Antiviral Therapy. 2011;16(3):383-394. DOI: 10.3851/IMP1756
  57. 57. Tang XL, Wu SM, Xie Y, Song N, Guan Q , Yuan C, et al. Generation and application of ssDNA aptamers against glycolipid antigen ManLAM of Mycobacterium tuberculosis for TB diagnosis. The Journal of Infection. 2016;72(5):573-586. DOI: 10.1016/j.jinf.2016.01.014
  58. 58. Tang XL, Zhou YX, Wu SM, Pan Q , Xia B, Zhang XL. CFP10 and ESAT6 aptamers as effective mycobacterial antigen diagnostic reagents. The Journal of Infection. 2014;69(6):569-580. DOI: 10.1016/j.jinf.2014.05.015
  59. 59. Rotherham LS, Maserumule C, Dheda K, Theron J, Khati M. Selection and application of ssDNA aptamers to detect active TB from sputum samples. PLoS One. 2012;7(10):e46862. DOI: 10.1371/journal.pone.0046862
  60. 60. Qin L, Zheng R, Ma Z, Feng Y, Liu Z, Yang H, et al. The selection and application of ssDNA aptamers against MPT64 protein in Mycobacterium tuberculosis. Clinical Chemistry and Laboratory Medicine. 2009;47(4):405-411. DOI: 10.1515/CCLM.2009.097
  61. 61. Zhu C, Liu J, Ling Y, Yang H, Liu Z, Zheng R, et al. Evaluation of the clinical value of ELISA based on MPT64 antibody aptamer for serological diagnosis of pulmonary tuberculosis. BMC Infectious Diseases. 2012;12:96. DOI: 10.1186/1471-2334-12-96
  62. 62. Ngubane NAC, Gresh L, Pym A, Rubin EJ, Khati M. Selection of RNA aptamers against the M. tuberculosis EsxG protein using surface plasmon resonance-based SELEX. Biochemical and Biophysical Research Communications. 2014;449(1):114-119. DOI: 10.1016/j.bbrc.2014.04.163
  63. 63. Aimaiti R, Qin L, Cao T, Yang H, Wang J, Lu J, et al. Identification and application of ssDNA aptamers against H37Rv in the detection of Mycobacterium tuberculosis. Applied Microbiology and Biotechnology. 2015;99(21):9073-9083. DOI: 10.1007/s00253-015-6815-7
  64. 64. Zhang XQ , Feng Y, Yao QQ , He F. Selection of a new Mycobacterium tuberculosis H37Rv aptamer and its application in the construction of a SWCNT/aptamer/Au-IDE MSPQC H37Rv sensor. Biosensors & Bioelectronics. 2017;98:261-266. DOI: 10.1016/j.bios.2017.05.043
  65. 65. Joshi R, Janagama H, Dwivedi HP, Senthil Kumar TMA, Jaykus LA, Schefers J, et al. Selection, characterization, and application of DNA aptamers for the capture and detection of Salmonella enterica serovars. Molecular and Cellular Probes. 2009;23(1):20-28. DOI: 10.1016/j.mcp.2008.10.006
  66. 66. Ma X, Jiang Y, Jia F, Yu Y, Chen J, Wang Z. An aptamer-based electrochemical biosensor for the detection of Salmonella. Journal of Microbiological Methods. 2014;98(1):94-98. DOI: 10.1016/j.mimet.2014.01.003
  67. 67. Han SR, Lee SW. In vitro selection of RNA aptamer specific to Salmonella Typhimurium. Journal of Microbiology and Biotechnology. 2013;23(6):878-884. DOI: 10.4014/jmb.1212.12033
  68. 68. Moon J, Kim G, Lee S, Park S. Identification of Salmonella typhimurium-specific DNA aptamers developed using whole-cell SELEX and FACS analysis. Journal of Microbiological Methods. 2013;95(2):162-166. DOI: 10.1016/j.mimet.2013.08.005
  69. 69. Duan N, Wu S, Chen X, Huang Y, Xia Y, Ma X, et al. Selection and characterization of aptamers against Salmonella typhimurium using whole-bacterium systemic evolution of ligands by exponential enrichment (SELEX). Journal of Agricultural and Food Chemistry. 2013;61(13):3229-3234. DOI: 10.1021/jf400767d
  70. 70. Lavu PSR, Mondal B, Ramlal S, Murali HS, Batra HV. Selection and characterization of aptamers using a modified whole cell bacterium SELEX for the detection of Salmonella enterica Serovar Typhimurium. ACS Combinatorial Science. 2016;18(6):292-301. DOI: 10.1021/acscombsci.5b00123
  71. 71. Yang M, Peng Z, Ning Y, Chen Y, Zhou Q , Deng L. Highly specific and cost-efficient detection of Salmonella paratyphi A combining aptamers with single-walled carbon nanotubes. Sensors (Switzerland). 2013;13(5):6865-6881. DOI: 10.3390/s130506865
  72. 72. Hyeon JY, Chon JW, Choi IS, Park C, Kim DE, Seo KH. Development of RNA aptamers for detection of Salmonella enteritidis. Journal of Microbiological Methods. 2012;89(1):79-82. DOI: 10.1016/j.mimet.2012.01.014
  73. 73. Bayraç C, Eyidoğan F, Avni Öktem H. DNA aptamer-based colorimetric detection platform for Salmonella enteritidis. Biosensors & Bioelectronics. 2017;98:22-28. DOI: 10.1016/j.bios.2017.06.029
  74. 74. DeGrasse JA. A single-stranded DNA aptamer that selectively binds to Staphylococcus aureus enterotoxin B. PLoS One. 2012;7(3):e33410. DOI: 10.1371/journal.pone.0033410
  75. 75. Huang Y, Chen X, Duan N, Wu S, Wang Z, Wei X, et al. Selection and characterization of DNA aptamers against Staphylococcus aureus enterotoxin C1. Food Chemistry. 2015;166:623-629. DOI: 10.1016/j.foodchem.2014.06.039
  76. 76. Hong KL, Battistella L, Salva AD, Williams RM, Sooter LJ. In vitro selection of single-stranded DNA molecular recognition elements against S. Aureus alpha toxin and sensitive detection in human serum. International Journal of Molecular Sciences. 2015;16(2):2794-2809. DOI: 10.3390/ijms16022794
  77. 77. Ferreira IM, de Souza Lacerda CM, de Faria LS, Corrêa CR, de Andrade ASR. Selection of peptidoglycan-specific aptamers for bacterial cells identification. Applied Biochemistry and Biotechnology. 2014;174(7):2548-2556. DOI: 10.1007/s12010-014-1206-6
  78. 78. Stoltenburg R, Schubert T, Strehlitz B. In vitro selection and interaction studies of a DNA aptamer targeting Protein A. PLoS One. 2015;10(7):e0134403. DOI: 10.1371/journal.pone.0134403
  79. 79. Cao X, Li S, Chen L, Ding H, Xu H, Huang Y, et al. Combining use of a panel of ssDNA aptamers in the detection of Staphylococcus aureus. Nucleic Acids Research. 2009;37(14):4621-4628. DOI: 10.1093/nar/gkp489
  80. 80. Borsa BA, Tuna BG, Hernandez FJ, Hernandez LI, Bayramoglu G, Arica MY, et al. Staphylococcus aureus detection in blood samples by silica nanoparticle-oligonucleotides conjugates. Biosensors & Bioelectronics. 2016;86:27-32. DOI: 10.1016/j.bios.2016.06.023
  81. 81. Chang YC, Yang CY, Sun RL, Cheng YF, Kao WC, Yang PC. Rapid single cell detection of Staphylococcus aureus by aptamer-conjugated gold nanoparticles. Scientific Reports. 2013;3:1863. DOI: 10.1038/srep01863
  82. 82. Ohk SH, Koo OK, Sen T, Yamamoto CM, Bhunia AK. Antibody-aptamer functionalized fibre-optic biosensor for specific detection of Listeria monocytogenes from food. Journal of Applied Microbiology. 2010;109(3):808-817. DOI: 10.1111/j.1365-2672.2010.04709.x
  83. 83. Bruno JG, Phillips T, Montez T, Garcia A, Sivils JC, Mayo MW, et al. Development of a fluorescent enzyme-linked DNA aptamer-magnetic bead sandwich assay and portable fluorometer for sensitive and rapid listeria detection. Journal of Fluorescence. 2015;25(1):173-183. DOI: 10.1007/s10895-014-1495-8
  84. 84. Suh SH, Jaykus LA. Nucleic acid aptamers for capture and detection of Listeria spp. Journal of Biotechnology. 2013;167(4):454-461. DOI: 10.1016/j.jbiotec.2013.07.027
  85. 85. Lee SH, Ahn JY, Lee KA, Um HJ, Sekhon SS, Sun Park T, et al. Analytical bioconjugates, aptamers, enable specific quantitative detection of Listeria monocytogenes. Biosensors & Bioelectronics. 2015;68:272-280. DOI: 10.1016/j.bios.2015.01.009
  86. 86. Bruno JG, Phillips T, Carrillo MP, Crowell R. Plastic-adherent DNA aptamer-magnetic bead and quantum dot sandwich assay for campylobacter detection. Journal of Fluorescence. 2009;19(3):427-435. DOI: 10.1007/s10895-008-0429-8
  87. 87. Amraee M, Oloomi M, Yavari A, Bouzari S. DNA aptamer identification and characterization for E. coli O157 detection using cell based SELEX method. Analytical Biochemistry. 2017;536:36-44. DOI: 10.1016/j.ab.2017.08.005
  88. 88. Peng Z, Ling M, Ning Y, Deng L. Rapid fluorescent detection of Escherichia coli K88 based on DNA aptamer library as direct and specific reporter combined with immuno-magnetic separation. Journal of Fluorescence. 2014;24(4):1159-1168. DOI: 10.1007/s10895-014-1396-x
  89. 89. Kim YS, Song MY, Jurng J, Kim BC. Isolation and characterization of DNA aptamers against Escherichia coli using a bacterial cell-systematic evolution of ligands by exponential enrichment approach. Analytical Biochemistry. 2013;436(1):22-28. DOI: 10.1016/j.ab.2013.01.014
  90. 90. Dua P, Ren S, Lee SW, Kim J, Shin H, Jeong O, et al. Cell-SELEX based identification of an RNA aptamer for Escherichia coli and Its use in various detection formats. Molecules and Cells. 2016;39(11):807-813. DOI: 10.1111/j.1369-7625.2012.00810.x
  91. 91. Renders M, Miller E, Lam CH, Perrin DM. Whole cell-SELEX of aptamers with a tyrosine-like side chain against live bacteria. Organic & Biomolecular Chemistry. 2017;15(9):1980-1989. DOI: 10.1039/C6OB02451C
  92. 92. Chen F, Zhou J, Luo F, Mohammed AB, Zhang XL. Aptamer from whole-bacterium SELEX as new therapeutic reagent against virulent Mycobacterium tuberculosis. Biochemical and Biophysical Research Communications. 2007;357(3):743-748. DOI: 10.1016/j.bbrc.2007.04.007
  93. 93. Pan Q , Wang Q , Sun X, Xia X, Wu S, Luo F, et al. Aptamer against mannose-capped lipoarabinomannan inhibits virulent Mycobacterium tuberculosis infection in mice and rhesus monkeys. Molecular Therapy. 2014;22(5):940-951. DOI: 10.1038/mt.2014.31
  94. 94. Sun X, Pan Q , Yuan C, Wang Q , Tang X-L, Ding K, et al. A single ssDNA aptamer binding to ManLAM of BCG enhances immunoprotective effects against tuberculosis. Journal of the American Chemical Society. 2016;138(36):11680-11689. DOI: 10.1021/jacs.6b05357
  95. 95. Baig IA, Moon JY, Lee SC, Ryoo SW, Yoon MY. Development of ssDNA aptamers as potent inhibitors of Mycobacterium tuberculosis acetohydroxyacid synthase. Biochim Biophys Acta—Proteins Proteomics. 2015;1854(10):1338-1350. DOI: 10.1016/j.bbapap.2015.05.003
  96. 96. Pan Q , Zhang XL, Wu HY, He PW, Wang F, Zhang MS, et al. Aptamers that preferentially bind type IVB pili and inhibit human monocytic-cell invasion by Salmonella enterica Serovar typhi. Antimicrobial Agents and Chemotherapy. 2005;49(10):4052-4060. DOI: 10.1128/AAC.49.10.4052-4060.2005
  97. 97. Zelada-Guillen GA, Riu J, Düzgün A, Rius FX. Immediate detection of living bacteria at ultralow concentrations using a carbon nanotube based potentiometrie aptasensor. Angewandte Chemie International Edition. 2009;48(40):7334-7337. DOI: 10.1002/anie.200902090
  98. 98. Wang K, Wu D, Chen Z, Zhang X, Yang X, Yang CJ, et al. Inhibition of the superantigenic activities of staphylococcal enterotoxin A by an aptamer antagonist. Toxicon. 2016;119:21-27. DOI: 10.1016/j.toxicon.2016.05.006
  99. 99. Vivekananda J, Salgado C, Millenbaugh NJ. DNA aptamers as a novel approach to neutralize Staphylococcus aureus α-toxin. Biochemical and Biophysical Research Communications. 2014;444(3):433-438. DOI: 10.1016/j.bbrc.2014.01.076
  100. 100. Yamamoto R, Katahira M, Nishikawa S, Baba T, Taira K, Kumar PKR. A novel RNA motif that binds efficiently and specifically to the Tat protein of HIV and inhibits the trans-activation by Tat of transcription in vitro and in vivo. Genes to Cells. 2000;5(5):371-388. DOI: 10.1046/j.1365-2443.2000.00330.x
  101. 101. Minunni M, Tombelli S, Gullotto A, Luzi E, Mascini M. Development of biosensors with aptamers as bio-recognition element: The case of HIV-1 Tat protein. Biosensors & Bioelectronics. 2004;20(6):1149-1156. DOI: 10.1016/j.bios.2004.03.037
  102. 102. Chen F, Hu Y, Li D, Chen H, Zhang XL. CS-SELEX generates high-affinity ssDNA aptamers as molecular probes for hepatitis C virus envelope glycoprotein E2. PLoS One. 2009;4(12):e8142. DOI: 10.1371/journal.pone.0008142
  103. 103. Lee S, Kim YS, Jo M, Jin M, ki Lee D, Kim S. Chip-based detection of hepatitis C virus using RNA aptamers that specifically bind to HCV core antigen. Biochemical and Biophysical Research Communications. 2007;358(1):47-52. DOI: 10.1016/j.bbrc.2007.04.057
  104. 104. Shi S, Yu X, Gao Y, Xue B, Wu X, Wang X, et al. Inhibition of hepatitis C virus production by aptamers against the core protein. Journal of Virology. 2014;88(4):1990-1999. DOI: 10.1128/JVI.03312-13
  105. 105. Ghanbari K, Roushani M, Azadbakht A. Ultra-sensitive aptasensor based on a GQD nanocomposite for detection of hepatitis C virus core antigen. Analytical Biochemistry. 2017;534:64-69. DOI: 10.1016/j.ab.2017.07.016
  106. 106. Sung HJ, Kayhan B, Ben-Yedidia T, Arnon R. A DNA aptamer prevents influenza infection by blocking the receptor binding region of the viral hemagglutinin. The Journal of Biological Chemistry. 2004;279(46):48410-48419. DOI: 10.1074/jbc.M409059200
  107. 107. Misono TS, Kumar PKR. Selection of RNA aptamers against human influenza virus hemagglutinin using surface plasmon resonance. Analytical Biochemistry. 2005;342(2):312-317. DOI: 10.1016/j.ab.2005.04.013
  108. 108. Gopinath SCB, Misono TS, Kawasaki K, Mizuno T, Imai M, Odagiri T, et al. An RNA aptamer that distinguishes between closely related human influenza viruses and inhibits haemagglutinin-mediated membrane fusion. The Journal of General Virology. 2006;87(3):479-487. DOI: 10.1099/vir.0.81508-0
  109. 109. Cheng C, Dong J, Yao L, Chen A, Jia R, Huan L, et al. Potent inhibition of human influenza H5N1 virus by oligonucleotides derived by SELEX. Biochemical and Biophysical Research Communications. 2008;366(3):670-674. DOI: 10.1016/j.bbrc.2007.11.183
  110. 110. Choi SK, Lee C, Lee KS, Choe SY, Mo IP, Seong RH, et al. DNA aptamers against the receptor binding region of hemagglutinin prevent avian influenza viral infection. Molecules and Cells. 2011;32(6):527-533. DOI: 10.1007/s10059-011-0156-x
  111. 111. Park SY, Kim S, Yoon H, Kim K-B, Kalme SS, Oh S, et al. Selection of an antiviral RNA aptamer against hemagglutinin of the subtype H5 avian influenza virus. Nucleic Acid Therapy (Formerly Oligonucleotides). 2011;21(6):395-402. DOI: 10.1089/nat.2011.0321
  112. 112. Gopinath SCB, Kumar PKR. Aptamers that bind to the hemagglutinin of the recent pandemic influenza virus H1N1 and efficiently inhibit agglutination. Acta Biomaterialia. 2013;9(11):8932-8941. DOI: 10.1016/j.actbio.2013.06.016
  113. 113. Wongphatcharachai M, Wang P, Enomoto S, Webby RJ, Gramer MR, Amonsin A, et al. Neutralizing DNA aptamers against swine influenza H3N2 viruses. Journal of Clinical Microbiology. 2013;51(1):46-54. DOI: 10.1128/JCM.02118-12
  114. 114. Suenaga E, Kumar PKR. An aptamer that binds efficiently to the hemagglutinins of highly pathogenic avian influenza viruses (H5N1 and H7N7) and inhibits hemagglutinin-glycan interactions. Acta Biomaterialia. 2014;10(3):1314-1323. DOI: 10.1016/j.actbio.2013.12.034
  115. 115. Woo HM, Lee JM, Yim S, Jeong YJ. Isolation of single-stranded DNA aptamers that distinguish influenza virus hemagglutinin subtype H1 from H5. PLoS One. 2015;10(4):e0125060. DOI: 10.1371/journal.pone.0125060
  116. 116. Zhang Y, Yu Z, Jiang F, Fu P, Shen J, Wu W, et al. Two DNA aptamers against avian influenza H9N2 virus prevent viral infection in cells. PLoS One. 2015;10(3):e0123060. DOI: 10.1371/journal.pone.0123060
  117. 117. Lai HC, Wang CH, Liou TM, Bin LG. Influenza A virus-specific aptamers screened by using an integrated microfluidic system. Lab on a Chip. 2014;14(12):2002-2013. DOI: 10.1016/j.vetpar.2015.05.004
  118. 118. Nguyen VT, Bin SH, Kim BC, Kim SK, Song CS, Gu MB. Highly sensitive sandwich-type SPR based detection of whole H5Nx viruses using a pair of aptamers. Biosensors & Bioelectronics. 2016;86:293-300. DOI: 10.1016/j.bios.2016.06.064
  119. 119. Liu J, Yang Y, Hu B, Ma ZY, Huang HP, Yu Y, et al. Development of HBsAg-binding aptamers that bind HepG2.2.15 cells via HBV surface antigen. Virologica Sinica. 2010;25(1):27-35. DOI: 10.1007/s12250-010-3091-7
  120. 120. Toscano-Garibay JD, Benítez-Hess ML, Alvarez-Salas LM. Isolation and characterization of an RNA aptamer for the HPV-16 E7 oncoprotein. Archives of Medical Research. 2011;42(2):88-96. DOI: 10.1016/j.arcmed.2011.02.005
  121. 121. Rahim Ruslinda A, Tanabe K, Ibori S, Wang X, Kawarada H. Effects of diamond-FET-based RNA aptamer sensing for detection of real sample of HIV-1 Tat protein. Biosensors & Bioelectronics. 2013;40(1):277-282. DOI: 10.1016/j.bios.2012.07.048
  122. 122. Zhou J, Swiderski P, Li H, Zhang J, Neff CP, Akkina R, et al. Selection, characterization and application of new RNA HIV gp 120 aptamers for facile delivery of Dicer substrate siRNAs into HIV infected cells. Nucleic Acids Research. 2009;37(9):3094-3109. DOI: 10.1093/nar/gkp185
  123. 123. Prokofjeva M, Tsvetkov V, Basmanov D, Varizhuk A, Lagarkova M, Smirnov I, et al. Anti-HIV activities of intramolecular G4 and Non-G4 oligonucleotides. Nucleic Acid Therapeutics. 2017;27(1):56-66. DOI: 10.1089/nat.2016.0624
  124. 124. Tuerk C, MacDougal S, Gold L. RNA pseudoknots that inhibit human immunodeficiency virus type 1 reverse transcriptase. Proceedings of the National Academy of Sciences of the United States of America. 1992;89(15):6988-6992. DOI: 10.1073/pnas.89.15.6988
  125. 125. Schneider DJ, Feigon J, Hostomsky Z, Gold L. High-affinity ssDNA inhibitors of the reverse transcriptase of type 1 human immunodeficiency virus. Biochemistry. 1995;34(29):9599-9610. DOI: 10.1021/bi00029a037
  126. 126. Mosing RK, Mendonsa SD, Bowser MT. Capillary electrophoresis-SELEX selection of aptamers with affinity for HIV-1 reverse transcriptase. Analytical Chemistry. 2005;77(19):6107-6112. DOI: 10.1021/ac050836q
  127. 127. Somasunderam A, Ferguson MR, Rojo DR, Thiviyanathan V, Li X, O’Brien WA, et al. Combinatorial selection, inhibition, and antiviral activity of DNA thioaptamers targeting the RNase H domain of HIV-1 reverse transcriptase. Biochemistry. 2005;44(30):10388-10395. DOI: 10.1021/bi0507074
  128. 128. DeStefano JJ, Nair GR. Novel aptamer inhibitors of human immunodeficiency virus reverse transcriptase. Oligonucleotides. 2008;18(2):133-144. DOI: 10.1089/oli.2008.0103
  129. 129. Ferreira-Bravo IA, Cozens C, Holliger P, DeStefano JJ. Selection of 2′-deoxy-2′-fluoroarabinonucleotide (FANA) aptamers that bind HIV-1 reverse transcriptase with picomolar affinity. Nucleic Acids Research. 2015;43(20):9587-9599. DOI: 10.1093/nar/gkv1057
  130. 130. Sánchez-Luque FJ, Stich M, Manrubia S, Briones C, Berzal-Herranz A. Efficient HIV-1 inhibition by a 16 nt-long RNA aptamer designed by combining in vitro selection and in silico optimisation strategies. Scientific Reports. 2014;1(4):6242. DOI: 10.1038/srep06242
  131. 131. Boiziau C, Dausse E, Yurchenko L, Toulmé JJ. DNA aptamers selected against the HIV-1 trans-activation-responsive RNA element form RNA-DNA kissing complexes. The Journal of Biological Chemistry. 1999;274(18):12730-12737. DOI: 10.1074/jbc.274.18.12730
  132. 132. Jing N, Hogan ME. Structure-activity of tetrad-forming oligonucleotides as a potent anti- HIV therapeutic drug. The Journal of Biological Chemistry. 1998;273(52):34992-34999. DOI: 10.1074/jbc.273.52.34992
  133. 133. Esposito V, Pirone L, Mayol L, Pedone E, Virgilio A, Galeone A. Exploring the binding of d(GGGT)4 to the HIV-1 integrase: An approach to investigate G-quadruplex aptamer/target protein interactions. Biochimie. 2016;127:19-22. DOI: 10.1016/j.biochi.2016.04.013
  134. 134. De Soultrait VR, Lozach PY, Altmeyer R, Tarrago-Litvak L, Litvak S, Andréola ML. DNA aptamers derived from HIV-1 RNase H inhibitors are strong anti-integrase agents. Journal of Molecular Biology. 2002;324(2):195-203. DOI: 10.1016/S0022-2836(02)01064-1
  135. 135. Phan AT, Kuryavyi V, Ma J-B, Faure A, Andreola M-L, Patel DJ. From the cover: An interlocked dimeric parallel-stranded DNA quadruplex: A potent inhibitor of HIV-1 integrase. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(3):634-639. DOI: 10.1073/pnas.0406278102
  136. 136. Kim SJ, Kim MY, Jeong S, Kim SJ, Lee JH, You JC. Selection and stabilization of the RNA aptamers against the human immunodeficiency virus type-1 nucleocapsid protein. Biochemical and Biophysical Research Communications. 2002;291(4):925-931. DOI: 10.1006/bbrc.2002.6521
  137. 137. Ramalingam D, Duclair S, Datta SAK, Ellington A, Rein A, Prasad VR. RNA aptamers directed to human immunodeficiency virus type 1 gag polyprotein bind to the matrix and nucleocapsid domains and inhibit virus production. Journal of Virology. 2011;85(1):305-314. DOI: 10.1128/JVI.02626-09
  138. 138. Jensen KB, Green L, MacDougal-Waugh S, Tuerk C. Characterization of an in vitro-selected RNA ligand to the HIV-1 rev protein. Journal of Molecular Biology. 1994;235(1):237-247. DOI: 10.1016/S0022-2836(05)80030-0
  139. 139. Pang KM, Castanotto D, Li H, Scherer L, Rossi JJ. Incorporation of aptamers in the terminal loop of shRNAs yields an effective and novel combinatorial targeting strategy. Nucleic Acids Research. 2017;46(1):e6. DOI: 10.1093/nar/gkx980
  140. 140. Kumar PKR, Machida K, Urvil PT, Kakiuchi N, Vishnuvardhan D, Shimotohno K, et al. Isolation of RNA aptamers specific to the NS3 protein of hepatitis C virus from a pool of completely random RNA. Virology. 1997;237(2):270-282. DOI: 10.1006/viro.1997.8773
  141. 141. Fukuda K, Vishnuvardhan D, Sekiya S, Hwang J, Kakiuchi N, Taira K, et al. Isolation and characterization of RNA aptamers specific for the hepatitis C virus nonstructural protein 3 protease. European Journal of Biochemistry. 2000;267(12):3685-3694. DOI: 10.1046/j.1432-1327.2000.01400.x
  142. 142. Nishikawa F, Funaji K, Fukuda K, Nishikawa S. In vitro selection of RNA aptamers against the HCVNS3 helicase domain. Oligonucleotides. 2004;14:114. DOI: 10.1089/1545457041526335
  143. 143. Kikuchi K, Umehara T, Fukuda K, Kuno A, Hasegawa T, Nishikawa S. A hepatitis C virus (HCV) internal ribosome entry site (IRES) domain III-IV-targeted aptamer inhibits translation by binding to an apical loop of domain IIId. Nucleic Acids Research. 2005;33(2):683-692. DOI: 10.1093/nar/gki215
  144. 144. Romero-López C, Barroso-delJesus A, Puerta-Fernández E, Berzal-Herranz A. Interfering with hepatitis C virus IRES activity using RNA molecules identified by a novel in vitro selection method. Biological Chemistry. 2005;386(2):183-190. DOI: 10.1515/BC.2005.023
  145. 145. Romero-López C, Díaz-González R, Berzal-Herranz A. Inhibition of hepatitis C virus internal ribosome entry site-mediated translation by an RNA targeting the conserved IIIf domain. Cellular and Molecular Life Sciences. 2007;64(22):2994-3006. DOI: 10.1007/s00018-007-7345-y
  146. 146. Romero-López C, Lahlali T, Berzal-Herranz B, Berzal-Herranz A. Development of optimized inhibitor RNAs allowing multisite-targeting of the HCV genome. Molecules. 2017;22(5):E861. DOI: 10.3390/molecules22050861
  147. 147. Bellecave P, Andreola M-L, Ventura M, Tarrago-Litvak L, Litvak S, Astier-Gin T. Selection of DNA aptamers that bind the RNA-dependent RNA polymerase of hepatitis C virus and inhibit viral RNA synthesis in vitro. Oligonucleotides. 2003;13(6):455-463. DOI: 10.1089/154545703322860771
  148. 148. Bellecave P, Cazenave C, Rumi J, Staedel C, Cosnefroy O, Andreola ML, et al. Inhibition of hepatitis C virus (HCV) RNA polymerase by DNA aptamers: Mechanism of inhibition of in vitro RNA synthesis and effect on HCV-infected cells. Antimicrobial Agents and Chemotherapy. 2008;52(6):2097-2110. DOI: 10.1128/AAC.01227-07
  149. 149. Jones LA, Clancy LE, Rawlinson WD, White PA. High-affinity aptamers to subtype 3a hepatitis C virus polymerase display genotypic specificity. Antimicrobial Agents and Chemotherapy. 2006;50(9):3019-3027. DOI: 10.1128/AAC.01603-05
  150. 150. Lee CH, Lee YJ, Kim JH, Lim JH, Kim J-H, Han W, et al. Inhibition of hepatitis C Virus (HCV) replication by specific RNA aptamers against HCV NS5B RNA replicase. Journal of Virology. 2013;87(12):7064-7074. DOI: 10.1128/JVI.00405-13
  151. 151. Kwon HM, Lee KH, Han BW, Han MR, Kim DH, Kim DE. An RNA aptamer that specifically binds to the glycosylated hemagglutinin of avian influenza virus and suppresses viral infection in cells. PLoS One. 2014;9(5):e97574. DOI: 10.1371/journal.pone.0097574
  152. 152. Yuan S, Zhang N, Singh K, Shuai H, Chu H, Zhou J, et al. Cross-protection of influenza A virus infection by a DNA aptamer targeting the PA endonuclease domain. Antimicrobial Agents and Chemotherapy. 2015;59(7):4082-4093. DOI: 10.1128/AAC.00306-15
  153. 153. Feng H, Beck J, Nassal M, hong Hu K. A SELEX-screened aptamer of human hepatitis B virus RNA encapsidation signal suppresses viral replication. PLoS One. 2011;6(11):e27862. DOI: 10.1371/journal.pone.0027862
  154. 154. Zhang Z, Zhang J, Pei X, Zhang Q , Lu B, Zhang X, et al. An aptamer targets HBV core protein and suppresses HBV replication in HepG2.2.15 cells. International Journal of Molecular Medicine. 2014;34(5):1423-1429. DOI: 10.3892/ijmm.2014.1908
  155. 155. Orabi A, Bieringer M, Geerlof A, Bruss V. An aptamer against the matrix binding domain on the hepatitis B virus capsid impairs virion formation. Journal of Virology. 2015;89(18):9281-9287. DOI: 10.1128/JVI.00466-15
  156. 156. Nicol C, Bunka DHJ, Blair GE, Stonehouse NJ. Effects of single nucleotide changes on the binding and activity of RNA aptamers to human papillomavirus 16 E7 oncoprotein. Biochemical and Biophysical Research Communications. 2011;405(3):417-421. DOI: 10.1016/j.bbrc.2011.01.044
  157. 157. Nicol C, Cesur O, Forrest S, Belyaeva TA, Bunka DHJ, Blair GE, et al. An RNA aptamer provides a novel approach for the induction of apoptosis by targeting the HPV16 E7 oncoprotein. PLoS One. 2013;8:e64781. DOI: 10.1371/journal.pone.0064781
  158. 158. Belyaeva T, Nicol C, Cesur Ö, Travé G, Blair G, Stonehouse N. An RNA aptamer targets the PDZ-binding motif of the HPV16 E6 oncoprotein. Cancers (Basel). 2014;6(3):1553-1569. DOI: 10.3390/cancers6031553
  159. 159. Shum KT, Tanner JA. Differential inhibitory activities and stabilisation of DNA aptamers against the SARS coronavirus helicase. Chembiochem. 2008;9(18):3037-3045. DOI: 10.1002/cbic.200800491
  160. 160. Chen HL, Hsiao WH, Lee HC, Wu SC, Cheng JW. Selection and characterization of DNA aptamers targeting all four serotypes of dengue viruses. PLoS One. 2015;10(6):e0131240. DOI: 10.1371/journal.pone.0131240
  161. 161. Liang HR, Hu GQ , Li L, Gao YW, Yang ST, Xia XZ. Aptamers targeting rabies virus-infected cells inhibit street rabies virus in vivo. International Immunopharmacology. 2014;21(2):432-438. DOI: 10.1016/j.intimp.2014.03.020
  162. 162. Shubham S, Hoinka J, Banerjee S, Swanson E, Dillard JA, Lennemann NJ, et al. A 2′FY-RNA motif defines an aptamer for Ebolavirus secreted protein. Science Reports. 2018;8(1):12373. DOI: 10.1038/s41598-018-30590-8
  163. 163. Lee KH, Zeng H. Aptamer-based ELISA assay for highly specific and sensitive detection of Zika NS1 protein. Analytical Chemistry. 2017;89(23):12743-12748. DOI: 10.1021/acs.analchem.7b02862
  164. 164. Que-Gewirth NS, Sullenger BA. Gene therapy progress and prospects: RNA aptamers. Gene Therapy. 2007;14(4):283-291. DOI: 10.1038/sj.gt.3302900
  165. 165. Kaur H, Bruno JG, Kumar A, Sharma TK. Aptamers in the therapeutics and diagnostics pipelines. Theranostics. 2018;8(15):4016-4032. DOI: 10.7150/thno.25958
  166. 166. O’Sullivan CK. Aptasensors—The future of biosensing? Analytical and Bioanalytical Chemistry. 2002;372:44-48. DOI: 10.1007/s00216-001-1189-3
  167. 167. Shamah SM, Healy JM, Cload ST. Complex target SELEX. Accounts of Chemical Research. 2008;41:130-138. DOI: 10.1021/ar700142z
  168. 168. Lim YC, Kouzani AZ, Duan W. Aptasensors: A review. Journal of Biomedical Nanotechnology. 2010;6(2):93-105. DOI: 10.1166/jbn.2010.1103
  169. 169. Dhiman A, Kalra P, Bansal V, Bruno JG, Sharma TK. Aptamer-based point-of-care diagnostic platforms. Sensors Actuators, B Chemical. 2017;246:535-553. DOI: 10.1016/j.snb.2017.02.060
  170. 170. Sharma TK. The point behind translation of aptamers for point of care diagnostics. Aptamers (Synthetic Antibodies). 2016;3:36-42
  171. 171. Jayasena SD. Aptamers: An emerging class of molecules that rival antibodies in diagnostics. Clinical Chemistry. 1999;45:1628-1650
  172. 172. Zhang P, Zhao N, Zeng Z, Chang CC, Zu Y. Combination of an aptamer probe to CD4 and antibodies for multicolored cell phenotyping. American Journal of Clinical Pathology. 2010;134(4):586-593. DOI: 10.1309/AJCP55KQYWSGZRKC
  173. 173. Abdeevaa IA, Maloshenoka LG, Pogorelkoab GV, Mokrykovaa MV, Bruskin SA. RNA-aptamers—As targeted inhibitors of protein functions in plants. Journal of Plant Physiology. 2019;232:127-129. DOI: 10.1016/j.jplph.2018.10.026
  174. 174. Konopka K, Lee NS, Rossi J, Düzgüneş N. Rev-binding aptamer and CMV promoter act as decoys to inhibit HIV replication. Gene. 2000;255(2):235-244. DOI: 10.1016/S0378-1119(00)00334-6
  175. 175. Chaloin L. Endogenous expression of a high-affinity pseudoknot RNA aptamer suppresses replication of HIV-1. Nucleic Acids Research. 2002;30(18):4001-4008. DOI: 10.1093/nar/gkf522
  176. 176. Soh JH, Lin Y, Rana S, Ying JY, Stevens MM. Colorimetric detection of small molecules in complex matrixes via target-mediated growth of aptamer-functionalized gold nanoparticles. Analytical Chemistry. 2015;87(15):7644-7652. DOI: 10.1021/acs.analchem.5b00875
  177. 177. Ruscito A, DeRosa MC. Small-molecule binding aptamers: Selection strategies, characterization, and applications. Frontiers in Chemistry. 2016;4(14):1-14. DOI: 10.3389/fchem.2016.00014
  178. 178. Liu A, Zhang Y, Chen W, Wang X, Chen F. Gold nanoparticle-based colorimetric detection of staphylococcal enterotoxin B using ssDNA aptamers. European Food Research and Technology. 2013;237(3):323-329. DOI: 10.1007/s00217-013-1995-9
  179. 179. Shim WB, Kim MJ, Mun H, Kim MG. An aptamer-based dipstick assay for the rapid and simple detection of aflatoxin B1. Biosensors & Bioelectronics. 2014;62:288-294. DOI: 10.1016/j.bios.2014.06.059
  180. 180. Candia J, Cheung F, Kotliarov Y, Fantoni G, Sellers B, Griesman T, et al. Assessment of variability in the SOMAscan assay. Scientific Reports. 2017;7(1):14248. DOI: 10.1038/s41598-017-14755-5
  181. 181. De Groote MA, Nahid P, Jarlsberg L, Johnson JL, Weiner M, Muzanyi G, et al. Elucidating novel serum biomarkers associated with pulmonary tuberculosis treatment. PLoS One. 2013;8(4):e61002. DOI: 10.1371/journal.pone.0061002
  182. 182. Kraemer S, Vaught JD, Bock C, Gold L, Katilius E, Keeney TR, et al. From SOMAmer-based biomarker discovery to diagnostic and clinical applications: A SOMAmer-based, streamlined multiplex proteomic assay. PLoS One. 2011;6(10):e26332. DOI: 10.1371/journal.pone.0026332
  183. 183. Lollo B, Steele F, Gold L. Beyond antibodies: New affinity reagents to unlock the proteome. Proteomics. 2014;14(6):638-644. DOI: 10.1002/pmic.201300187
  184. 184. Mehan MR, Ayers D, Thirstrup D, Xiong W, Ostroff RM, Brody EN, et al. Protein signature of lung cancer tissues. PLoS One. 2012;7(4):e35157. DOI: 10.1371/journal.pone.0035157
  185. 185. Müller J, Friedrich M, Becher T, Braunstein J, Kupper T, Berdel P, et al. Monitoring of plasma levels of activated protein C using a clinically applicable oligonucleotide-based enzyme capture assay. Journal of Thrombosis and Haemostasis. 2012;10(3):390-398. DOI: 10.1111/j.1538-7836.2012.04623.x
  186. 186. Yan X, Song Y, Liu J, Zhou N, Zhang CL, He L, et al. Two-dimensional porphyrin-based covalent organic framework: A novel platform for sensitive epidermal growth factor receptor and living cancer cell detection. Biosensors & Bioelectronics. 2018;5663(18):30944-30948. DOI: 10.1016/j.bios.2018.11.047
  187. 187. Song Y, Shi Y, Huang M, Wang W, Wang Y, Cheng J, et al. Bioinspired engineering of multivalent aptamer-functionalized nanointerface to enhance capture and release of circulating tumor cells. Angewandte Chemie International Edition. 2018;131:1-6. DOI: 1002/anie.201809337
  188. 188. Hong P, Li W, Li J. Applications of aptasensors in clinical diagnostics. Sensors. 2012;12(2):1181-1193. DOI: 10.3390/s120201181
  189. 189. Bhat VG, Chavan P, Ojha S, Nair PK. Challenges in the laboratory diagnosis and management of dengue infections. The Open Microbiology Journal. 2015;9:33-37. DOI: 10.2174/1874285801509010033
  190. 190. Tsai WY, Youn HH, Brites C, Tsai JJ, Tyson J, Pedroso C, et al. Distinguishing secondary dengue virus infection from Zika virus infection with previous dengue by a combination of 3 simple serological tests. Clinical Infectious Diseases. 2017;65(11):1829-1836. DOI: 10.1093/cid/cix672
  191. 191. Kim YJ, Kim HS, Chon JW, Kim DH, Hyeon JY, Seo KH. New colorimetric aptasensor for rapid on-site detection of Campylobacter jejuni and Campylobacter coli in chicken carcass samples. Analytica Chimica Acta. 2018;1029:78-85. DOI: 10.1016/j.aca.2018.04.059
  192. 192. Labib M, Zamay AS, Kolovskaya OS, Reshetneva IT, Zamay GS, Kibbee RJ, et al. Aptamer-based impedimetric sensor for bacterial typing. Analytical Chemistry. 2012;84(19):8114-8117. DOI: 10.1021/ac302217u
  193. 193. Yoon SY, Gee G, Hong KJ, Seo SH. Application of aptamers for assessment of vaccine efficacy. Clinical and Experimental Vaccine Research. 2017;6:160-163. DOI: 10.7774/cevr.2017.6.2.160
  194. 194. WHO. Tuberculosis. 2018. Available from: https://www.who.int/news-room/fact-sheets/detail/tuberculosis
  195. 195. WHO. Global Tuberculosis Report 2015. DOI: 978-92-4-156450-2
  196. 196. Elhassan MM, Elmekki MA, Osman AL, Hamid ME. Challenges in diagnosing tuberculosis in children: A comparative study from Sudan. International Journal of Infectious Diseases. 2016;43:25-29. DOI: 10.1016/j.ijid.2015.12.006
  197. 197. Hilman BC. Pediatric tuberculosis: Problems in diagnosis and issues in management. The Journal of the Louisiana State Medical Society: Official organ of the Louisiana State Medical Society. 1998;150(12):601-610
  198. 198. Racsa LD, Kraft CS, Olinger GG, Hensley LE. Viral hemorrhagic fever diagnostics. Clinical Infectious Diseases. 2016;62:214-219. DOI: 10.1093/cid/civ792
  199. 199. Miyazaki D, Shimizu D, Shimizu Y, Inoue Y, Inoue T, Higaki S, et al. Diagnostic efficacy of real-time PCR for ocular cytomegalovirus infections. Graefe's Archive for Clinical and Experimental Ophthalmology. 2018;256(12):2413-2420. DOI: 10.1007/s00417-018-4111-9
  200. 200. Nolan N, Halai UA, Regunath H, Smith LP, Rojas-Moreno C, Salzer W. Primary cytomegalovirus infection in immunocompetent adults in the United States—A case series. IDCases. 2017;10:123-126. DOI: 10.1016/j.idcr.2017.10.008
  201. 201. O’Hara KM, Pontrelli G, Kunstel KL. An introduction to gastrointestinal tract CMV disease. Journal of the Academy of Physician Assistants. 2017;30:48-52. DOI: 10.1097/01.JAA.0000524712.40590.76
  202. 202. Marty AM, Jahrling PB, Geisbert TW. Viral hemorrhagic fevers. Clinics in Laboratory Medicine. 2006;26(2):345-386. DOI: 10.1016/j.cll.2006.05.001
  203. 203. Behring K. Ueber das Zustandekommen der Diphtherie-Immunitüt und der Tetanus-Immunitüt bei Thieren. Dtsch Medizinische Wochenschrift. 1890;16:1113-1114. DOI: 10.1055/s-0029-1207589
  204. 204. Edelman G. Antibody structure and molecular immunology. Annals of the New York Academy of Sciences. 1971;190:5-25

Written By

Muslum Ilgu, Rezzan Fazlioglu, Meric Ozturk, Yasemin Ozsurekci and Marit Nilsen-Hamilton

Submitted: 31 August 2018 Reviewed: 31 January 2019 Published: 10 April 2019