Open access peer-reviewed chapter

Aptamers for Infectious Disease Diagnosis

Written By

Soma Banerjee and Marit Nilsen-Hamilton

Submitted: 11 November 2018 Reviewed: 20 May 2019 Published: 22 June 2019

DOI: 10.5772/intechopen.86945

From the Edited Volume

E. Coli Infections - Importance of Early Diagnosis and Efficient Treatment

Edited by Luis Rodrigo

Chapter metrics overview

1,478 Chapter Downloads

View Full Metrics

Abstract

Aptamers are in vitro-selected, nucleic acids with unique abilities to bind strongly and specifically to their selective targets (ligands) based on their three-dimensional structures. Target binding is generally associated with a change in aptamer structure, which provides a means of linking many output signals to the binding event. Being synthetic, aptamers are less expensive compared to antibodies. Aptamers are also more easily modified chemically or their sequence changed to optimize properties such target specificity, storability and stability. In this chapter we will discuss the potential benefits of applying aptamers to diagnostics with a focus on infectious disease and the unique challenges posed by aptamers for their successful incorporation into reliable aptasensors.

Keywords

  • aptamers
  • SELEX
  • aptasensors
  • portable diagnostic tools
  • electrochemical impedance spectroscopy

1. Introduction

Aptamers, first disclosed in 1990 by three groups [1, 2, 3], are ssDNA or RNA molecules capable of binding strongly and specifically to their target (ligand) molecules. Their target binding specificities and affinities are based on their sequence-specific 3D structures. Such properties of aptamers make them analogues of antibodies with unique advantages. For example, aptamers are relatively small (diam. ~2 nm) compared to antibodies (diam. ~15 nm), which allows them to bind targets that are inaccessible to the larger antibodies. Like antibodies, their properties are defined by the ionic conditions and pH in which they are placed. However, being shorter polymers, aptamers are generally more sensitive than antibodies to their physical and chemical environment.

In contrast to the time-consuming and expensive production and screening procedures for antibodies, aptamers can be produced faster and more cost effectively by a procedure known as Systemic Evolution of Ligands by Exponential Enrichment (SELEX). Once an aptamer sequence has been identified, its further production is by chemical synthesis, for which variation is negligible compared with the batch to batch variation of antibodies generated in animals or by cell culture. Their synthetic production makes aptamers accessible for selective chemical modifications to enhance their binding specificity or to increase their resistance to degradation. With such advantages over antibodies, aptamers have emerged as new generation molecular recognition elements [4]. In the current chapter, we focus on their impact in diagnosis of infectious disease agents. The reader is referred to other reviews of the application of aptamers to therapy, biosensing and molecular probing [5, 6, 7, 8, 9, 10].

Advertisement

2. Aptamers in diagnostics

Fast and accurate diagnosis is a key factor for the treatment of infectious disease. Molecular recognition by aptamers can be highly discriminatory such that they can distinguish between two closely related molecules, including conformational isomers [11] or highly related proteins [12, 13, 14]. One-to-one comparison between aptamers and antibodies as recognition elements for the same target molecule, demonstrated that aptamers can equal antibodies in their sensitivity and selectivity [15]. As an added advantage, aptamers showed tolerance to repeated regeneration and recycling after ligand binding.

The small sizes and homogeneous structures of aptamers allow them to be immobilized in a dense and well-oriented manner. The higher density of aptamer packing compared with antibodies increases the binding capacity of the sensors and extends their linear range of detection of analyte [9, 15]. With these aspects of aptamers considered, their application as recognition elements in analytical devices offers a multitude of advantages and brings a new dimension to diagnostics.

Aptamers are compatible as the recognition element with many sensor platforms, including quartz crystal microbalances (QCM), surface plasmon resonance (SPR), diamond field effect transistors (FET), electrochemical impedance spectroscopy (EIS), colorimetric and fluorescence-based optical detection. Of these, electrochemical impedance spectroscopy (EIS) has gained popularity as it offers rapid, low-cost, label-free detection with high signal to noise ratios and sensitive detection of target molecules when employing aptamers [16, 17]. EIS is more sensitive than other electrochemical techniques [18] and is one of the best techniques to analyze the properties of electrochemical systems [19]. EIS is a technique used to study the electrochemical response to the application of periodic small amplitude ac signal at different frequencies [20, 21]. It is useful to monitor the changes in the electrochemical properties of the system due to biorecognition events at the surface of modified electrodes. For example, the electrodes can be modified with aptamers to detect the presence of a target analyte. EIS produces high quality data by directly converting a biological event into an electrical signal. Moreover, EIS-based sensors are small and portable and can be employed outside of well-equipped laboratory. EIS is an attractive technique for biosensor development as it provides the advantages of real time monitoring and label-free detection and is compatible with flexible electronics, disposable sensors and wearable devices. Inkjet printing can be applied to produce aptamer-based EIS biosensors for their automated mass production with uniform aptamer deposition [22, 23, 24].

In some instances aptamers can also be used for therapeutics or for combined diagnostics and therapeutics (theranostics). Examples include (1) various aptamers to human immunodeficiency virus (HIV) proteins that either prevent virus entry or replication [7, 25, 26, 27, 28, 29], (2) the S-PS8.4 aptamer that recognizes Salmonella enterica and inhibits invasion of the bacteria into human monocytes [30], (3) two DNA aptamers A9 and B4 against the H9N2 avian influenza virus that inhibit viral infection [31], and (4) a DNA aptamer against MUC-1 that was applied in a nanocomposite for fluorescent imaging and demonstrated to inhibit the proliferation of colorectal (HT-29) and breast (MCF-7) cancer cells [32].

Although antibodies dominate the global market of diagnostics and therapeutics, several biotechnology companies have actively started exploring aptamers for diagnostics. The first to develop aptamer-based diagnostic arrays, SomaLogic employed SOMAmers (slow off rate modified aptamers) to detect many protein biomarkers for disease diagnosis [33, 34]. The combination of aptamers as the recognition elements, with long shelf lives at room temperature, inkjet printing methods for immobilizing them and EIS as the method of detection will enable the development of low cost, label free, and rapid response diagnostic devices that could be in disposable or wearable forms as well as in more conventional instrument formats.

2.1 Aptamers that detect biomarkers of microbial infections

Early detection of infectious disease is of primary importance for its management. The conventional methods of diagnosis, which include microbiological methods (isolation, growth and microscopy of pathogens), polymerase chain reaction (PCR) and immunological methods [35, 36] suffer from turn-around times of 24 h or longer. This is especially a problem when the patient(s) are located in remote areas from which samples must be sent to centralized laboratories for their analysis. It is also difficult to grow some microbes in culture, which limits their detection.

Viral diseases are generally detected by serology, immunology or PCR amplification of DNA/RNA fragments corresponding to the pathogen’s genome. Although it is sensitive and specific, the performance of PCR requires appropriate instruments, specialized reagents and experienced personnel. Immunological methods, which are widely used for diagnosis, employ antibodies specific to a protein or carbohydrate moiety that is unique to the target pathogen. Some popular immunological methods are the agglutination, ELISA and western blot assays. These well-established and time-tested assays are the work-horses of modern clinical laboratories. They are generally reliable and are likely to be the mainstay of the clinical technical repertoire well into the future. But, these assays limit the ability of communities to respond rapidly to microbial and viral outbreaks because they require laboratory equipment that is not readily portable and trained personnel to perform them. ELISA and western blots also depend on provided antibodies, which require cold storage to prevent their denaturation. Infectious disease outbreaks often start in regions that are distant from clinical laboratories. Therefore, the challenge for future diagnostics is to develop portable devices that require little expertise to perform. Current technology development is moving in this direction with portable PCR machines [37] and lateral flow immunology tests [38]. Whereas the former still requires trained personal to operate, the latter can often be readily used and interpreted by an untrained individual.

Diagnostics based on aptamers stand out as promising options for rapid, cost effective and specific detection of pathogens using devices that can be operated by minimally trained personnel. Many aptamers have been reported that recognize specific viruses and bacteria. Some were selected against recombinant proteins from the target microbe or virus. Others were selected against the intact microbe or virus. For example, RNA aptamer S-PS8.4, which specifically recognizes the type IVB pili of Salmonella enterica, was isolated using a recombinant pilin structural protein as the selection target [39] and incorporated into a potentiometric biosensor as a recognition element for S. enterica [40]. This aptasensor could detect a single CFU of target S. enterica and was specific for S. enterica, not recognizing either Escherichia coli or Lactobacillus casei. Using a cell-SELEX approach, two 62 nt DNA aptamers, SA17 and SA61, were selected against intact Staphylococcus aureus. As for S-PS8.4, these aptamers bound their S. aureus target with high affinities and specificities [41]. Many aptamers have been selected with specificities for particular microbe targets including or Campylobacter jejuni [42, 43], L. monocytogenes [44, 45, 46, 47, 48], Vibrio parahemolyticus [49], Shigella dysenteriae [50], Streptococcus pyogenes [51], Francisella tularensis [52], Pseudomonas aeruginosa [53] and the spores of anthrax Bacillus anthracis [54, 55]. Parasites are also good aptamer targets with aptamers identified that recognize Trypanosoma spp., Plasmodium spp., Leishmania spp., Entamoeba histolytica, and Cryptosporidium parvuum [56]. For the diagnosis of invasive fungal infections, DNA aptamers have been screened against (1 → 3)-β-D-glucans from cell wall of Candida albicans. Two selected DNA aptamers (AU1 and AD1) showed high binding affinities in the range of 100 nM and did not bind to the same domain of (1 → 3)-β-D-glucans. The application of these aptamers in a double-aptamer sandwich enzyme-linked oligonucleotide assay (ELONA) resulted in an assay sensitivity and specificity of the detection of ~92% [57]. For viruses, there are aptamers that recognize HIV intracellular proteins [7, 13, 58, 59, 60], the HIV envelope glycoprotein [28, 61], hepatitis C [62, 63, 64], influenza virus [65, 66, 67], herpes simplex virus 1 [68], dengue virus [69], zika virus [70] and ebola virus [71]. In another study, a device was reported for the multiplexed detection of the envelope proteins of Zika and chikungunya viruses. The detection takes place in a microfluidic channel containing microsized pillars with attached aptamers. These pillars increase the surface sensing area, thereby enabling the attachment of more aptamers and increasing the overall sensitivity of the sensor envelope proteins. The working principle of this device depends on the formation of a protein-mediated sandwich with an aptamer-functionalized gold nanoparticle (AuNP) and an unattached aptamer. The signal is obtained upon introduction of silver reagents into the channel, which is selectively deposited on the AuNP surface, providing a gray contrast in the testing zone. This colorimetric aptasensor is reported to detect clinically relevant concentrations of Zika and chikungunya envelope proteins in phosphate-buffered saline (1 pM) and calf blood (100 pM) with high specificity [72].

Many of the aptamers discussed have been employed as recognition elements in diagnostic tools of a wide variety of types with electrochemical sensors being a popular platform. Examples include a potentiometric carbon-nanotube system to detect trypanosomes in blood [73], a voltametric aptasensor for ultrasensitive detection of Mycobacterium tuberculosis (MTB) virulence factor antigen ESAT-6 [74] and an EIS aptasensor for influenza virus [67].

Along with fast read-out, another advantage of the electrochemical approach is high sensitivity. The aptamer-based detection threshold is sometimes lower than for RT-PCR as demonstrated for influenza virus [67]. In another example, aptamer conjugates with gold nanoparticles were sensitive enough to detect a single Staphylococcus aureus cell [41]. An aptasensor has been reported that detects attomolar concentrations of the variable surface glycoprotein from African trypanosomes as analyte in blood [73]. From these and other examples it was found that immobilization of aptamers on biosensor surfaces increases target binding affinity [75, 76, 77, 78]. The increased affinity is most likely due to two effects of immobilization: (1) immobilization creates a multivalent surface that decreases the rate at which the aptamer ligand can leave the surface (off-rate), which is the denominator in calculating the association constant (Ka = kon/koff), and (2) molecular crowding promotes aptamer folding to produce the appropriate ligand-binding structure [79, 80, 81].

Due to their low concentrations in the blood, infectious disease markers can be difficult to detect [82, 83]. Here, aptamers can play a different role of concentrating the target prior to their quantification. For this purpose, magnetic beads coated with aptamers specific for Trypanosoma cruzi were used to capture these parasites from the blood in which they are present at very low concentrations [84].

The stage is set for the development of commercial diagnostics for infectious disease agents. Some have already come to market such as OTA-Sense and Aflasense developed by Neoventures Biotechnology Inc. for detection of toxins in food samples, AptoCyto and AptoPrep developed by Aptsci Inc. for isolation of biomarker positive cells, SOMAscan from SomaLogic for diagnosis of several diseases, CibusDx a food pathogen Diagnostic platform developed by the USA-based start-up (Pronucleotein, Inc.), OLIGOBIND®Thrombin activity assay by Sekisui Diagnostics [85]. Aptasensors demonstrate remarkably short detection times compared with the conventional methods of ELISA and PCR. The advantage of a faster detection time is of utmost importance for identifying rapidly developing and epidemiologically dangerous diseases, such as influenza, Ebola and SARS (Severe Acute Respiratory Syndrome). The rapid detection capabilities of aptasensors and their ready portability will broaden their scope of acceptance in the field of diagnosis.

Advertisement

3. Aptasensors

Some of the most attractive features of aptamer-based sensors (aptasensors) are their stability to storage at ambient temperatures and their reusability. Moreover, their small size and versatility allow aptamers to be immobilized at high densities, which facilitate their multiplexing in miniaturized systems. Several signaling modes have been coupled to aptamer-based sensors [86, 87]. Some popular outputs for detection include fluorescence [88], chemiluminescence [55], electrochemical [89], field effects (FET) [90], surface plasmon resonance (SPR) [91], changes in resonating frequency of quartz crystal sensor (QCM) [92], surface acoustic waves (SAW) [93], mechanical (microcantilevers) [94]. In this chapter, we will focus mainly on aptamer-based biosensors with fluorescent or electrochemical outputs.

3.1 Fluorescent aptasensors

Aptasensors with fluorescence outputs are designed to take advantage of the flexibility of aptamers, which results in their frequently adopting alternate conformations in the presence or absence of their target molecules. For these sensors, aptamers are modified in key positions with fluorescent dyes that interact in Förster resonance energy transfer (FRET). Upon aptamer binding to its target, the associated structural change alters the distance between the fluorescent dyes and thus the efficiency of energy transfer. The signal change, manifested as an increase (signal-on mode) or decrease (signal-off mode) in fluorescence, is proportional to the extent of target binding. A representative “signal-on” fluorescent aptamer holds a fluorophore, usually at one end, which is quenched by a molecule that is attached to a proximate location in the unoccupied aptamer. Target binding separates fluorophore from quencher allowing recovery of the fluorescent signal, which provides a quantitative measure of the target concentration [95]. FRET can also be used in “signal-off” sensor designs in which the conformational change of the aptamer on target binding brings the donor and quencher into closer proximity with a resulting fluorescence quenching. Sensors based on the “signal-off” mode are usually less sensitive than those based on the “signal-on” mode, but they can help to improve target detection by low-affinity aptamers [88].

An alternative means of signaling an aptamer binding event using fluorescence is with an oligonucleotide (attenuator) that is complementary to a portion of the aptamer and remains bound to the aptamer in the absence of target molecule. The length of the attenuator and its placement on the aptamer must create a condition that prevents aptamer folding to the actively binding conformation, but the affinity of the aptamer for the attenuator should be less than for the target molecule. With these requirements fulfilled, the target molecule can successfully compete with the attenuator to bind the aptamer and release the attenuator. Such a design can be used for “signal-on” reporting if the target and aptamer are labeled with fluorescent dyes that are optimally placed to interact in FRET while the aptamer and attenuator are hybridized. This format can also be used for a “signal-off” system in which a single fluorophore is attached to the attenuator. When the aptamer binds the target molecule, the released attenuator adsorbs to surrounding gold nanoparticles (AuNPs), which quench the fluorescence [96]. Another aptasensor design used upconversion nanoparticles (UCNPs) as donors and AuNPs as acceptors for rapid, ultrasensitive and specific detection of bacteria (e.g., E. coli ATCC 8739) [97]. FRET-based aptasensors provide an efficient method for detecting pathogens and their released toxins in one step [98, 99, 100].

The concept of using a material to quench the fluorophore was applied to create a paper-based MoS2 nanosheet-mediated FRET aptasensor for rapid malaria diagnosis [96]. This format uses paper test strips impregnated with fluorescently-labeled aptamers and MoS2 nanosheets. The MoS2 quenches the fluorescence until the aptamers are released when they bind their targets. These aptasensors are facile, inexpensive and therefore attractive for point-of-care diagnosis, especially in low-resource areas. Similar “low-tech” FRET-based aptasensors have also been found to be ideal for spacecraft, such as for diagnosing microgravity-induced bone loss in outer space by monitoring urinary C-telopeptide [4, 101]. In such scenarios, where both space and lab resources are limited, handheld fluorometers such as the commercially available QuantiFluor™ (Promega Corp.) or other such portable fluorometers will provide much needed opportunities for point-of-care diagnostics. These applications benefit from the greater stability to ambient temperatures for storage of aptamers compared with antibodies.

3.2 Electrochemical aptasensors

Upon binding to their target molecules, aptamers fold their supple, single-stranded chains into distinct three-dimensional (3D) structures. This structural change can be employed for initiating electron-transfer when the aptamers are labeled with a redox-active moiety and immobilized on a conducting support. Several electrochemical aptasensors have been developed based on this strategy [87], which can also be classified into “signal-on” and “signal-off” aptasensors. For example, an electrochemical thrombin aptasensor was constructed by immobilizing a thrombin aptamer (TBA) labeled with redox-active methylene blue (MB) on an electrode [102]. After binding thrombin, the TBA adopts a G-quadruplex structure, which moves MB away from the electrode. This “signal-off” sensing format has the disadvantage of a decreasing signal with increasing target molecule. An example of the preferred “signal-on” format includes the TBA, which is immobilized on a gold electrode and tagged with a terminal electroactive ferrocene redox label [103]. In the absence of thrombin a low signal is produced because many of the conformations adopted by the aptamer do not bring the ferrocene close to the electrode. Upon binding thrombin, the TBA adopts a G-quadruplex conformation, bringing the ferrocene to the electrode to allow electron-transfer and a positive signal in the presence of target molecule.

Electrochemical signals can be amplified when catalytic events are part of the signaling mechanism. For example, an electrochemical aptasensor was developed to rapidly diagnose tuberculosis (TB) by detecting the Mycobacterium tuberculosis antigen, MPT64, in serum samples [104]. MPT64 exists in serum as a disulfide linked homo-multimer. With multiple target sites on the same multimeric particle, MPT64 can be detected by a sandwich assay with the same aptamer on each side of the sandwich. In this study, coil-like fullerene (C60)-doped polyaniline (C60-PAn) nanohybrids were used as redox nanoprobes and catalysts to initiate the oxidation of ascorbic acid. When linked to the MPT64 aptamer, these nanohybrids were brought close to a gold surface (also decorated with MPT64 aptamers) in a sandwich joined by MPT64 multimers. In this configuration the electrons released by the oxidation of ascorbic acid were transferred to the gold electrode. This simple yet elegant approach for TB diagnosis showed selectivity to target antigen over several other serum proteins, a wide linear range of detection from 0.02 to 1000 pg/mL and a detection limit of 20 fg/mL MPT64. The delayed diagnosis and misdiagnosis of patients with MTB infection is the leading cause behind the spread and high mortality rate of TB [105]. Therefore, the possibility of rapid and accurate detection of MTB by these aptasensors is of great significance for the early diagnosis and treatment of TB.

Electrochemical impedance spectroscopy (EIS), an electrochemical label-free detection method, can be an extremely sensitive method for target recognition at the electrode/electrolyte interface [106]. Here, aptamers are immobilized on a gold (Au) electrode and the remaining gold surface filled in by a self-assembled monolayer such as mercaptohexanol (6-MCH). This approach was used with a DNA aptamer as molecular recognition element for malaria detection, for which the response range of 1 pM–10 nM covered the diagnostically relevant concentration range of Plasmodium lactate dehydrogenase protein from the falciparum parasite species (PfLDH) [107, 108]. The aptasensor functioned well with a sample matrix of 10% human serum and could be regenerated for reuse by washing with 6 M urea.

Electrochemical aptasensors have been fabricated to be sufficiently small to insert into a vein for continuous, real-time measurement of specific molecular targets in situ in the living body. The limited surface area of these small devices leads to low faradaic currents and poor signal-to-noise ratios when deployed in the complex, fluctuating environments found in vivo. To circumvent this problem, an electrochemical roughening approach was developed to enhance the signal-to-noise ratios by increasing the microscopic surface area of gold electrodes, thereby allowing more redox reporter-modified aptamers to be packed onto the surface. These high surface area electrochemical aptasensors of less than 200 μm in diameter were used in a proof-of concept study to measure continuous drug pharmacokinetic profiles over a 3 h period in live rats [109].

Colorimetric detection is gaining popularity in the diagnostic field considering its low cost and the minimal training needed to identify and interpret the visible signal. A colorimetric approach was used to develop a diagnostic device for tuberculosis with aptamers that bind to antibodies against the MPT64 protein secreted by Mycobacterium tuberculosis. When adsorbed to Fe3O4 magnetic nanoparticles (MNPs) the aptamers decrease the ability of the particles to reduce oxygen to H2O2 [110]. Upon exposure of the MNP-aptamer suspension to anti-MPT64 antibodies, the aptamers preferentially bind to the antibodies, thereby increasing the available surface area of the MNPs with the resulting higher rates of H2O2 production. Inclusion of 2,2′-azino-bis(3-ethylbenzo-thiazoline-6-sulfonic acid) (ABTS), which is oxidized by H2O2 and converted to a colored product, signals the presence of anti-MPT64 antibodies [111]. Another format for a colorimetric aptasensor is a paper-based microfluidic chip. For this format, aptamers against bacteria considered as nosocomial and antibiotic-resistant were immobilized by ultraviolet crosslinking on a nitrocellulose membrane housed within the chip. Incubation with bacteria, washing and then the addition of biotinylated aptamers allowed the use of HRP-linked streptavidin to create a blue color based on the oxidation of tetramethyl benzidine (TMB) by the H2O2 product of HRP, which was trapped on the surface by way of its linkage to the streptavidin bound to the biotinylated aptamers attached to the target bacteria. This dual-aptamer microfluidic chip possesses many advantages such as rapid output (35 min), small size, higher specificity, and the capability to detect multiple pathogens simultaneously, which are ideal for point-of-care bacterial diagnostics [112]. A similar sandwich type aptasensor has been reported for the early diagnosis of periodontitis, in which chronic inflammation is caused by many factors including pathogenic bacteria. Periodontitis is one of the major causes of tooth loss in adults. Aptamers were targeted against a potential biomarker of this disease, odontogenic ameloblast-associated protein (ODAM). The lateral flow strip format used a cognate pair of aptamers that recognize different sites on ODAM. One aptamer (20 aptamer) was attached to gold nanoparticles that were mixed with the sample to capture the biomarker. The second aptamer (10 aptamer) was present in a line on the strip to capture nanoparticles with attached biomarker. A control line with DNA complementary to the 20 aptamer captured the particles that did not have attached biomarkers. The biosensor had a detection limit of 0.24 and 1.63 nM in buffer and saliva samples, respectively [113].

Advertisement

4. Considerations for further development of aptasensor applications in diagnostics

Of the many aptamers that have been selected, very few have been applied as recognition elements in sensors and fewer have reached the stage of commercial availability. In this section we will consider some of the reasons that aptasensors, with all their promise, have been slow to come to the market in diagnostic devices.

One reason for the limited breadth of application of aptamers to diagnostics is that many of the biosensor platforms are new to the concept of incorporating aptamers as sensors. Therefore, the research focus has been on developing sensor platforms that are compatible with aptamers. For proof-of-principle devices, aptamers that have been previously demonstrated to function well on a variety of sensor platforms have been chosen as recognition elements. Consequently, the TBA and ATP aptamers have been incorporated into many sensor platforms [114, 115]. However, these targets are not relevant biomarkers for disease. With many sensor platforms now validated using these “prototype” aptamers the field has the opportunity to move forward to incorporate and optimize some of the many available clinically relevant aptamers for diagnostic applications.

Nucleic acids are more flexible polymers than polypeptides. Whereas there are two rotatable bonds between each amino acid side-chain in a polypeptide, there are six rotatable bonds between each base in a nucleic acid. This additional flexibility gives aptamers the property of ready structural rearrangement with target binding and enables their incorporation into many sensor platforms that rely on these rearrangements for creating signals. Thus, the fundamental principles underpinning antibody and aptamer-based sensors are different. Whereas antibody-based sensors rely on the uniformity of antibody structure and their bivalency, aptamer-based sensors rely on the flexibility of the monovalent aptamer structure and the structural changes that occur on target binding. Consequently, a single sensor format can be readily adapted to many antibodies, but each aptamer-sensor combination must be optimized to benefit from the unique structural change of the relevant aptamer. Thus, although there may be some applications for sensors that use the TBA and ATP model aptamers [116], their repeated use in developing and testing aptamer-based sensors has delayed the development of aptasensors for relevant biomarkers.

With the importance of aptamer structure and conformational changes for sensor development, a second challenge for reliable aptasensor development lies in the dearth of known aptamer tertiary structures. This deficiency results from several circumstances. First, biophysical determination of nucleic acid structure by conventional methods such as NMR and X-ray crystallography is more challenging than for proteins. Second, when structures can be determined by biophysical means, these are often only the structures of the target-bound aptamers because the apo-aptamer structures are too flexible to be reduced to a single structure. Third, new aptamers are being reported at an increasingly rapid rate that is much faster than their structures can be determined. Finally, small modifications of existing aptamers that can have large effects on aptamer structure are often made for their application in sensors. Thus, aptasensors must be developed with little information about the tertiary structure of the aptamer employed and how it changes with target binding. The most likely route to obtaining molecular structure for most aptamers will be in silico modeling. Although not yet demonstrated to be adequate for accurately predicting the structures of short nucleic acids and how they change with target binding, molecular modeling techniques are improving and their successful integration into models for sensor mechanisms could eventually result in dependable strategies for engineering new aptamers and integrating them into sensors.

Many aptamer selection protocols require the availability of purified target molecules [117]. Protein target molecules are usually expressed as recombinant proteins by prokaryotic or eukaryotic cell cultures and then purified, frequently by affinity chromatography based on a capture tag. Like for obtaining antibodies, the protein targets must be pure. Difficulties can come if the recombinant protein is not post-translationally modified similarly to the native protein. For example, many biomarkers found in the blood are glycosylated in the native form, but the recombinant proteins produced by bacterial cells are not glycosylated. Due to their steric hindrance or by their altering the protein structure in the region of the aptamer epitope these modifications can make regions of a native eukaryotic protein inaccessible to aptamers generated against the recombinant protein equivalent expressed in prokaryotic cells [118]. An early screen for aptamers selected to bind non-glycosylated recombinant proteins should be to determine if they bind the native glycosylated protein. Approaches to selecting for aptamers that recognize glycan structure in the context of the protein will also be useful [119].

Cell-SELEX avoids the complication of the target lacking the native posttranslational modifications by selecting against the cell surface protein target in situ [120, 121]. However, the identities of targets obtained by Cell-SELEX are often not known. As well, Cell-SELEX is performed with cell lines that are different from normal or in situ cancer cells and that are cultured under conditions that differ from those in the body. In particular, cultured cells have adjusted to exist in serum, which is not present in vivo and to an environment of much higher oxygen content than cells in situ, both of which conditions might result in altered protein expression on the cell surface.

Another consideration for aptasensors is that they frequently must function with the target (analyte) in a complex sample matrix. Many sensors have been demonstrated to function well in simple buffers, while the most common biomedical samples (blood, serum, urine and saliva) are complex with many potentially interfering substances. Aptamer target binding affinity can be sensitive to the matrix in the range of dilutions commonly used to detect the analyte, such as 10 or 50% serum or urine [122, 123]. For biofluids, sampling methods must also be considered with a view to minimal invasiveness and small sample volumes.

As aptasensors are developed that avoid the pitfalls discussed here, these sensors will take their place beside the antibody-based assays and provide new capabilities such as continuous analyte monitoring and inexpensive devices that can be distributed to small clinics throughout the world and yet be connected by Bluetooth and other options to send their results to central clinics and distant physicians for improved monitoring of patients in rural and other isolated locations.

Advertisement

5. Prospects of aptasensors in next-generation diagnosis

Medical diagnostics is moving towards a future of individualized patient care. Individuals vary greatly in their response to specific drugs and in the rate at which these drugs are removed from the system [124, 125, 126, 127, 128]. Pharmacokinetic parameters have been reported to vary daily in the same patient [129] or with time of drug exposure [130] and between individuals depending on disease state [131, 132, 133], age [134], genetics [130, 135, 136, 137, 138, 139], concurrent medication [137], body fat composition [140], and even circadian rhythm [141]. Adherence to a drug intake regime is also a factor, particularly for care of the young and the aged [142, 143, 144]. For these many reasons, individualized diagnostics are considered a clinical necessity for improved patient treatment and for establishing effective therapeutic windows [145].

With the push towards individualized medicine as a desirable future approach for optimal patient care comes the need to move some diagnostics out of the clinic into the home. For this purpose, reliable inexpensive sensors that might be linked by wireless connections to clinical centers would be optimal. Some personalized diagnostics has long been available in the home. These include pregnancy tests and glucose monitors, which are based on antibody and enzymatic reactions. However, these are not linked to the larger medical care network. The course is now set for a huge expansion of personalized diagnostics that do not require trained operators on site, but that can transmit information to clinical specialists who can monitor a patient’s condition off-site and continuously. This is a niche for which aptasensors can provide a large diversity of options with their potential for long shelf-live under ambient conditions, simplicity of operation, ability to be designed for continuous use or repeated use, and compatibility with wearable sensor formats.

Aptasensors can also be applied to address the acute need for diagnostics during infectious disease epidemics by their placement in clinics located in isolated regions of the world and in individual physician’s offices that are distant from major well-equipped hospitals and clinical centers. With the additional capability of Bluetooth communication, centers of disease control can be quickly updated regarding the spread of infections, which will enable central authorities to rapidly initiate effective means of controlling a potential epidemic.

A future of diagnostics for individualized medicine and for the control of infectious disease outbreaks over vast regions will result from many cross-disciplinary collaborations that are already underway, which include experts in molecular biology, virology, medicine, engineering, diagnostics and other disciplines. With this effort, many of the aptasensors that are still now at the proof-of concept stage, are expected to become major contributors to a future of improved personalized health care for all people, including those living in remote regions, and will help to stem future outbreaks of infectious diseases.

References

  1. 1. Ellington AD, Szostak JW. In vitro selection of RNA molecules that bind specific ligands. Nature. 1990;346:818-822. DOI: 10.1038/346818a0
  2. 2. Robertson DL, Joyce GF. Selection in vitro of an RNA enzyme that specifically cleaves single-stranded DNA. Nature. 1990;344:467-468. DOI: 10.1038/344467a0
  3. 3. Tuerk C, Gold L. Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science. 1990;249:505-510
  4. 4. Dhiman A, Kalra P, Bansal V, Bruno JG, Sharma TK. Aptamer-based point-of-care diagnostic platforms. Sensors and Actuators B: Chemical. 2017;246:535-553. DOI: 10.1016/j.snb.2017.02.060
  5. 5. Banerjee J, Nilsen-Hamilton M. Aptamers: Multifunctional molecules for biomedical research. Journal of molecular medicine (Berlin, Germany). 2013;91:1333-1342. DOI: 10.1007/s00109-013-1085-2
  6. 6. Wandtke T, Woźniak J, Kopiński P. Aptamers in diagnostics and treatment of viral infections. Viruses. 2015;7:751. DOI: 10.3390/v7020751
  7. 7. Bala J, Chinnapaiyan S, Dutta RK, Unwalla H. Aptamers in HIV research diagnosis and therapy. RNA Biology. 2018;15:327-337. DOI: 10.1080/15476286.2017.1414131
  8. 8. Zhou J, Rossi J. Aptamers as targeted therapeutics: Current potential and challenges. Nature Reviews Drug Discovery. 2016;16:181. DOI: 10.1038/nrd.2016.199
  9. 9. Tombelli S, Minunni M, Mascini M. Aptamers-based assays for diagnostics, environmental and food analysis. Biomolecular Engineering. 2007;24:191-200. DOI: 10.1016/j.bioeng.2007.03.003
  10. 10. Tombelli S, Mascini M. Aptamers as molecular tools for bioanalytical methods. Current Opinion in Molecular Therapeutics. 2009;11:179-188
  11. 11. Geiger A, Burgstaller P, von der Eltz H, Roeder A, Famulok M. RNA aptamers that bind L-arginine with sub-micromolar dissociation constants and high enantioselectivity. Nucleic Acids Research. 1996;24:1029-1036. DOI: 10.1093/nar/24.6.1029
  12. 12. Jenison RD, Jennings SD, Walker DW, Bargatze RF, Parma D. Oligonucleotide inhibitors of P-selectin-dependent neutrophil-platelet adhesion. Antisense and Nucleic Acid Drug Development. 1998;8:265-279. DOI: 10.1089/oli.1.1998.8.265
  13. 13. Tombelli S, Minunni M, Luzi E, Mascini M. Aptamer-based biosensors for the detection of HIV-1 Tat protein. Bioelectrochemistry. 2005;67:135-141. DOI: 10.1016/j.bioelechem.2004.04.011
  14. 14. Ruslinda AR, 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 and Bioelectronics. 2013;40:277-282. DOI: 10.1016/j.bios.2012.07.048
  15. 15. Liss M, Petersen B, Wolf H, Prohaska E. An aptamer-based quartz crystal protein biosensor. Analytical Chemistry. 2002;74:4488-4495
  16. 16. Rohrbach F, Karadeniz H, Erdem A, Famulok M, Mayer G. Label-free impedimetric aptasensor for lysozyme detection based on carbon nanotube-modified screen-printed electrodes. Analytical Biochemistry. 2012;421:454-459. DOI: 10.1016/j.ab.2011.11.034
  17. 17. Liang G, Man Y, Jin X, Pan L, Liu X. Aptamer-based biosensor for label-free detection of ethanolamine by electrochemical impedance spectroscopy. Analytica Chimica Acta. 2016;936:222-228. DOI: 10.1016/j.aca.2016.06.056
  18. 18. Karimipour M, Heydari-Bafrooei E, Sanjari M, Johansson MB, Molaei M. A glassy carbon electrode modified with TiO2(200)-rGO hybrid nanosheets for aptamer based impedimetric determination of the prostate specific antigen. Microchimica Acta. 2018;186:33. DOI: 10.1007/s00604-018-3141-7
  19. 19. Ehsani A, Mahjani MG, Hosseini M, Safari R, Moshrefi R, Mohammad SH. Evaluation of Thymus vulgaris plant extract as an eco-friendly corrosion inhibitor for stainless steel 304 in acidic solution by means of electrochemical impedance spectroscopy, electrochemical noise analysis and density functional theory. Journal of Colloid and Interface Science. 2017;490:444-451. DOI: 10.1016/j.jcis.2016.11.048
  20. 20. Kongsuphol P, Ng HH, Pursey JP, Arya SK, Wong CC, Stulz E, et al. EIS-based biosensor for ultra-sensitive detection of TNF-α from non-diluted human serum. Biosensors and Bioelectronics. 2014;61:274-279. DOI: 10.1016/j.bios.2014.05.017
  21. 21. Grieshaber D, MacKenzie R, Vörös J, Reimhult E. Electrochemical biosensors—Sensor principles and architectures. Sensors (Basel, Switzerland). 2008;8:1400-1458. DOI: 10.3390/s80314000
  22. 22. da Costa TH, Song E, Tortorich RP, Choi JW. A paper-based electrochemical sensor using inkjet-printed carbon nanotube electrodes. ECS Journal of Solid State Science and Technology. 2015;4:S3044-S3047. DOI: 10.1149/2.0121510jss
  23. 23. Maddaus A, Curley P, Griswold MA, Costa BD, Hou S, Jeong KJ, et al. Design and fabrication of bio-hybrid materials using inkjet printing. Biointerphases. 2016;11:041002. DOI: 10.1116/1.4966164
  24. 24. Khan NI, Maddaus AG, Song E. A low-cost inkjet-printed aptamer-based electrochemical biosensor for the selective detection of lysozyme. Biosensors. 2018;8:7. DOI: 10.3390/bios8010007
  25. 25. Konopka K, Lee NS, Rossi J, Duzgunes N. Rev-binding aptamer and CMV promoter act as decoys to inhibit HIV replication. Gene. 2000;255:235-244
  26. 26. Bai J, Banda N, Lee NS, Rossi J, Akkina R. RNA-based anti-HIV-1 gene therapeutic constructs in SCID-hu mouse model. Molecular Therapy. 2002;6:770-782
  27. 27. Chaloin L, Lehmann MJ, Sczakiel G, Restle T. Endogenous expression of a high-affinity pseudoknot RNA aptamer suppresses replication of HIV-1. Nucleic Acids Research. 2002;30:4001-4008. DOI: 10.1093/nar/gkf522
  28. 28. Khati M, Schuman M, Ibrahim J, Sattentau Q , Gordon S, James W. Neutralization of infectivity of diverse R5 clinical isolates of human immunodeficiency virus type 1 by gp120-binding 2'F-RNA aptamers. Journal of Virology. 2003;77:12692-12698. DOI: 10.1128/jvi.77.23.12692-12698.2003
  29. 29. Dey AK, Khati M, Tang M, Wyatt R, Lea SM, James W. An aptamer that neutralizes R5 strains of human immunodeficiency virus type 1 blocks gp120-CCR5 interaction. Journal of Virology. 2005;79:13806-13810. DOI: 10.1128/JVI.79.21.13806-13810.2005
  30. 30. Davydova A, Vorobjeva M, Pyshnyi D, Altman S, Vlassov V, Venyaminova A. Aptamers against pathogenic microorganisms. Critical Reviews in Microbiology. 2016;42:847-865. DOI: 10.3109/1040841x.2015.1070115
  31. 31. 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:e0123060. DOI: 10.1371/journal.pone.0123060
  32. 32. Binaymotlagh R, Hajareh Haghighi F, Aboutalebi F, Mirahmadi-Zare SZ, Hadadzadeh H, Nasr-Esfahani MH. Selective chemotherapy and imaging of colorectal and breast cancer cells by a modified MUC-1 aptamer conjugated to a poly(ethylene glycol)-dimethacrylate coated Fe3O4–AuNCs nanocomposite. New Journal of Chemistry. 2019;43:238-248. DOI: 10.1039/c8nj04236e
  33. 33. Gold L, Ayers D, Bertino J, Bock C, Bock A, Brody EN, et al. Aptamer-based multiplexed proteomic technology for biomarker discovery. PLOS One. 2010;5:e15004. DOI: 10.1371/journal.pone.0015004
  34. 34. 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;7:3948-3961. DOI: 10.7150/thno.20725
  35. 35. Shinde SB, Fernandes CB, Patravale VB. Recent trends in in-vitro nanodiagnostics for detection of pathogens. Journal of Controlled Release. 2012;159:164-180. DOI: 10.1016/j.jconrel.2011.11.033
  36. 36. Kaittanis C, Santra S, Perez JM. Emerging nanotechnology-based strategies for the identification of microbial pathogenesis. Advanced Drug Delivery Reviews. 2010;62:408-423. DOI: 10.1016/j.addr.2009.11.013
  37. 37. Mendoza-Gallegos RA, Rios A, Garcia-Cordero JL. An affordable and portable thermocycler for real-time PCR made of 3D-printed parts and off-the-shelf electronics. Analytical Chemistry. 2018;90:5563-5568. DOI: 10.1021/acs.analchem.7b04843
  38. 38. van Grootveld R, van Dam GJ, de Dood C, de Vries JJC, Visser LG, Corstjens PLAM, et al. Improved diagnosis of active Schistosoma infection in travellers and migrants using the ultra-sensitive in-house lateral flow test for detection of circulating anodic antigen (CAA) in serum. European Journal of Clinical Microbiology and Infectious Diseases. 2018;37:1709-1716. DOI: 10.1007/s10096-018-3303-x
  39. 39. 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;49:4052-4060. DOI: 10.1128/aac.49.10.4052-4060.2005
  40. 40. Zelada-Guillén GA, Riu J, Düzgün A, Rius FX. Immediate detection of living bacteria at ultralow concentrations using a carbon nanotube based potentiometric aptasensor. Angewandte Chemie International Edition. 2009;48:7334-7337. DOI: 10.1002/anie.200902090
  41. 41. 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;3:1863-1863. DOI: 10.1038/srep01863
  42. 42. 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. 2008;19:427. DOI: 10.1007/s10895-008-0429-8
  43. 43. Dwivedi HP, Smiley RD, Jaykus LA. Selection and characterization of DNA aptamers with binding selectivity to Campylobacter jejuni using whole-cell SELEX. Applied Microbiology and Biotechnology. 2010;87:2323-2334. DOI: 10.1007/s00253-010-2728-7
  44. 44. Duan N, Ding X, He L, Wu S, Wei Y, Wang Z. Selection, identification and application of a DNA aptamer against Listeria monocytogenes. Food Control. 2013;33:239-243. DOI: 10.1016/j.foodcont.2013.03.011
  45. 45. 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;109:808-817. DOI: 10.1111/j.1365-2672.2010.04709.x
  46. 46. Suh SH, Choi SJ, Dwivedi HP, Moore MD, Escudero-Abarca BI, Jaykus L-A. Use of DNA aptamer for sandwich type detection of Listeria monocytogenes. Analytical Biochemistry. 2018;557:27-33. DOI: 10.1016/j.ab.2018.04.009
  47. 47. Suh SH, Dwivedi HP, Choi SJ, Jaykus LA. Selection and characterization of DNA aptamers specific for Listeria species. Analytical Biochemistry. 2014;459:39-45. DOI: 10.1016/j.ab.2014.05.006
  48. 48. Suh SH, Jaykus LA. Nucleic acid aptamers for capture and detection of Listeria spp. Journal of Biotechnology. 2013;167:454-461. DOI: 10.1016/j.jbiotec.2013.07.027
  49. 49. Duan N, Wu S, Chen X, Huang Y, Wang Z. Selection and identification of a DNA aptamer targeted to Vibrio parahemolyticus. Journal of Agricultural and Food Chemistry. 2012;60:4034-4038. DOI: 10.1021/jf300395z
  50. 50. Duan N, Ding X, Wu S, Xia Y, Ma X, Wang Z, et al. In vitro selection of a DNA aptamer targeted against Shigella dysenteriae. Journal of Microbiological Methods. 2013;94:170-174. DOI: 10.1016/j.mimet.2013.06.016
  51. 51. Hamula CLA, Le XC, Li XF. DNA aptamers binding to multiple prevalent M-types of Streptococcus pyogenes. Analytical Chemistry. 2011;83:3640-3647. DOI: 10.1021/ac200575e
  52. 52. Vivekananda J, Kiel JL. Anti-Francisella tularensis DNA aptamers detect tularemia antigen from different subspecies by aptamer-linked immobilized sorbent assay. Laboratory Investigation. 2006;86:610. DOI: 10.1038/labinvest.3700417
  53. 53. Wang KY, Zeng YL, Yang XY, Li WB, Lan XP. Utility of aptamer-fluorescence in situ hybridization for rapid detection of Pseudomonas aeruginosa. European Journal of Clinical Microbiology and Infectious Diseases. 2011;30:273-278. DOI: 10.1007/s10096-010-1074-0
  54. 54. Bruno JG, Carrillo MP. Development of aptamer beacons for rapid presumptive detection of bacillus spores. Journal of Fluorescence. 2012;22:915-924. DOI: 10.1007/s10895-011-1030-0
  55. 55. Bruno JG, Kiel JL. In vitro selection of DNA aptamers to anthrax spores with electrochemiluminescence detection. Biosensors and Bioelectronics. 1999;14:457-464
  56. 56. Ospina-Villa JD, Lopez-Camarillo C, Castanon-Sanchez CA, Soto-Sanchez J, Ramirez-Moreno E, Marchat LA. Advances on aptamers against protozoan parasites. Genes (Basel). 2018;9:584. DOI: 10.3390/genes9120584
  57. 57. Tang XL, Hua Y, Guan Q , Yuan CH. Improved detection of deeply invasive candidiasis with DNA aptamers specific binding to (1 → 3)--D-glucans from Candida albicans. European Journal of Clinical Microbiology and Infectious Diseases. 2016;35:587-595. DOI: 10.1007/s10096-015-2574-8
  58. 58. Kensch O, Connolly BA, Steinhoff HJ, McGregor A, Goody RS, Restle T. HIV-1 reverse transcriptase-pseudoknot RNA aptamer interaction has a binding affinity in the low picomolar range coupled with high specificity. The Journal of Biological Chemistry. 2000;275:18271-18278. DOI: 10.1074/jbc.M001309200
  59. 59. de Soultrait VR, Lozach PY, Altmeyer R, Tarrago-Litvak L, Litvak S, Andreola ML. DNA aptamers derived from HIV-1 RNase H inhibitors are strong anti-integrase agents. Journal of Molecular Biology. 2002;324:195-203
  60. 60. Kim SJ, Kim MY, Lee JH, You JC, Jeong S. Selection and stabilization of the RNA aptamers against the human immunodeficiency virus type-1 nucleocapsid protein. Biochemical and Biophysical Research Communications. 2002;291:925-931. DOI: 10.1006/bbrc.2002.6521
  61. 61. Dey AK, Griffiths C, Lea SM, James W. Structural characterization of an anti-gp120 RNA aptamer that neutralizes R5 strains of HIV-1. RNA. 2005;11:873-884. DOI: 10.1261/rna.7205405
  62. 62. Urvil PT, Kakiuchi N, Zhou DM, Shimotohno K, Kumar PK, Nishikawa S. Selection of RNA aptamers that bind specifically to the NS3 protease of hepatitis C virus. European Journal of Biochemistry. 1997;248:130-138
  63. 63. 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;267:3685-3694
  64. 64. 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:e8142. DOI: 10.1371/journal.pone.0008142
  65. 65. Gopinath SC, 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;87:479-487. DOI: 10.1099/vir.0.81508-0
  66. 66. Park SY, Kim S, Yoon H, Kim KB, Kalme SS, Oh S, et al. Selection of an antiviral RNA aptamer against hemagglutinin of the subtype H5 avian influenza virus. Nucleic Acid Therapeutics. 2011;21:395-402. DOI: 10.1089/nat.2011.0321
  67. 67. Bai C, Lu Z, Jiang H, Yang Z, Liu X, Ding H, et al. Aptamer selection and application in multivalent binding-based electrical impedance detection of inactivated H1N1 virus. Biosensors and Bioelectronics. 2018;110:162-167. DOI: 10.1016/j.bios.2018.03.047
  68. 68. Gopinath SCB, Hayashi K, Kumar PKR. Aptamer that binds to the gD protein of herpes simplex virus 1 and efficiently inhibits viral entry. Journal of Virology. 2012;86:6732-6744. DOI: 10.1128/JVI.00377-12
  69. 69. 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:e0131240. DOI: 10.1371/journal.pone.0131240
  70. 70. Kim DTH, Bao DT, Park H, Ngoc NM, Yeo SJ. Development of a novel peptide aptamer-based immunoassay to detect Zika virus in serum and urine. Theranostics. 2018;8:3629-3642. DOI: 10.7150/thno.25955
  71. 71. 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. Scientific Reports. 2018;8:12373. DOI: 10.1038/s41598-018-30590-8
  72. 72. Saraf N, Villegas M, Willenberg BJ, Seal S. Multiplex viral detection platform based on a aptamers-integrated microfluidic channel. ACS Omega. 2019;4:2234-2240. DOI: 10.1021/acsomega.8b03277
  73. 73. Zelada-Guillén GA, Tweed-Kent A, Niemann M, Göringer HU, Riu J, Rius FX. Ultrasensitive and real-time detection of proteins in blood using a potentiometric carbon-nanotube aptasensor. Biosensors and Bioelectronics. 2013;41:366-371. DOI: 10.1016/j.bios.2012.08.055
  74. 74. Li L, Yuan Y, Chen Y, Zhang P, Bai Y, Bai L. Aptamer based voltammetric biosensor for Mycobacterium tuberculosis antigen ESAT-6 using a nanohybrid material composed of reduced graphene oxide and a metal-organic framework. Microchimica Acta. 2018;185:379. DOI: 10.1007/s00604-018-2884-5
  75. 75. Daniel C, Roupioz Y, Gasparutto D, Livache T, Buhot A. Solution-phase vs surface-phase aptamer-protein affinity from a label-free kinetic biosensor. PLOS One. 2013;8:e75419. DOI: 10.1371/journal.pone.0075419
  76. 76. 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 and Bioelectronics. 2019;126:88-95. DOI: 10.1016/j.bios.2018.10.010
  77. 77. 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;84:8763-8770. DOI: 10.1021/ac3020643
  78. 78. Kang KH, Sachan A, Nilsen-Hamilton M, Shrotriya P. Aptamer functionalized microcantilever sensors for cocaine detection. Langmuir. 2011;27:14696-14702. DOI: 10.1021/la202067y
  79. 79. Dupuis NF, Holmstrom ED, Nesbitt DJ. Molecular-crowding effects on single-molecule RNA folding/unfolding thermodynamics and kinetics. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:8464-8469. DOI: 10.1073/pnas.1316039111 1316039111 [pii]
  80. 80. Paudel BP, Rueda D. Molecular crowding accelerates ribozyme docking and catalysis. Journal of the American Chemical Society. 2014;136:16700-16703. DOI: 10.1021/ja5073146
  81. 81. Nakano SI, Sugimoto N. The structural stability and catalytic activity of DNA and RNA oligonucleotides in the presence of organic solvents. Biophysical Reviews. 2016;8:11-23. DOI: 10.1007/s12551-015-0188-0
  82. 82. Sun K, Xia N, Zhao L, Liu K, Hou W, Liu L. Aptasensors for the selective detection of alpha-synuclein oligomer by colorimetry, surface plasmon resonance and electrochemical impedance spectroscopy. Sensors and Actuators B: Chemical. 2017;245:87-94. DOI: 10.1016/j.snb.2017.01.171
  83. 83. Hansen JA, Wang J, Kawde AN, Xiang Y, Gothelf KV, Collins G. Quantum-dot/aptamer-based ultrasensitive multi-analyte electrochemical biosensor. Journal of the American Chemical Society. 2006;128:2228-2229. DOI: 10.1021/ja060005h
  84. 84. Nagarkatti R, Bist V, Sun S, Fortes de Araujo F, Nakhasi HL, Debrabant A. Development of an aptamer-based concentration method for the detection of Trypanosoma cruzi in blood. PLOS One. 2012;7:e43533. DOI: 10.1371/journal.pone.0043533
  85. 85. Kaur H, Bruno JG, Kumar A, Sharma TK. Aptamers in the therapeutics and diagnostics pipelines. Theranostics. 2018;8:4016-4032. DOI: 10.7150/thno.25958
  86. 86. Song MS, Sekhon S, Shin WR, Kim H, Min J, Ahn JY, et al. Detecting and discriminating Shigella sonnei using an aptamer-based fluorescent biosensor platform. Molecules. 2017;22:825
  87. 87. Song S, Wang L, Li J, Fan C, Zhao J. Aptamer-based biosensors. TrAC Trends in Analytical Chemistry. 2008;27:108-117. DOI: 10.1016/j.trac.2007.12.004
  88. 88. Wang RE, Zhang Y, Cai J, Cai W, Gao T. Aptamer-based fluorescent biosensors. Current Medicinal Chemistry. 2011;18:4175-4184
  89. 89. Swensen JS, Xiao Y, Ferguson BS, Lubin AA, Lai RY, Heeger AJ, et al. Continuous, real-time monitoring of cocaine in undiluted blood serum via a microfluidic, electrochemical aptamer-based sensor. Journal of the American Chemical Society. 2009;131:4262-4266. DOI: 10.1021/ja806531z
  90. 90. Yoon H, Kim JH, Lee N, Kim BG, Jang J. A novel sensor platform based on aptamer-conjugated polypyrrole nanotubes for label-free electrochemical protein detection. Chembiochem. 2008;9:634-641. DOI: 10.1002/cbic.200700660
  91. 91. Polonschii C, David S, Tombelli S, Mascini M, Gheorghiu M. A novel low-cost and easy to develop functionalization platform. Case study: Aptamer-based detection of thrombin by surface plasmon resonance. Talanta. 2010;80:2157-2164. DOI: 10.1016/j.talanta.2009.11.023
  92. 92. Zhang X, Yadavalli VK. Surface immobilization of DNA aptamers for biosensing and protein interaction analysis. Biosensors and Bioelectronics. 2011;26:3142-3147. DOI: 10.1016/j.bios.2010.12.012
  93. 93. Treitz G, Gronewold TMA, Quandt E, Zabe-Kühn M. Combination of a SAW-biosensor with MALDI mass spectrometric analysis. Biosensors and Bioelectronics. 2008;23:1496-1502. DOI: 10.1016/j.bios.2008.01.013
  94. 94. Chen X, Pan Y, Liu H, Bai X, Wang N, Zhang B. Label-free detection of liver cancer cells by aptamer-based microcantilever biosensor. Biosensors and Bioelectronics. 2016;79:353-358. DOI: 10.1016/j.bios.2015.12.060
  95. 95. Kumar YVVA, Renuka RM, Achuth J, Mudili V, Poda S. Development of a FRET-based fluorescence aptasensor for the detection of aflatoxin B1 in contaminated food grain samples. RSC Advances. 2018;8:10465-10473. DOI: 10.1039/c8ra00317c
  96. 96. Emrani AS, Danesh NM, Lavaee P, Ramezani M, Abnous K, Taghdisi SM. Colorimetric and fluorescence quenching aptasensors for detection of streptomycin in blood serum and milk based on double-stranded DNA and gold nanoparticles. Food Chemistry. 2016;190:115-121. DOI: 10.1016/j.foodchem.2015.05.079
  97. 97. Jin B, Wang S, Lin M, Jin Y, Zhang S, Cui X, et al. Upconversion nanoparticles based FRET aptasensor for rapid and ultrasenstive bacteria detection. Biosensors and Bioelectronics. 2017;90:525-533. DOI: 10.1016/j.bios.2016.10.029
  98. 98. Nasirian V, Chabok A, Barati A, Rafienia M, Arabi MS, Shamsipur M. Ultrasensitive aflatoxin B1 assay based on FRET from aptamer labelled fluorescent polymer dots to silver nanoparticles labeled with complementary DNA. Microchimica Acta. 2017;184:4655-4662. DOI: 10.1007/s00604-017-2508-5
  99. 99. Lu X, Wang C, Qian J, Ren C, An K, Wang K. Target-driven switch-on fluorescence aptasensor for trace aflatoxin B1 determination based on highly fluorescent ternary CdZnTe quantum dots. Analytica Chimica Acta. 2019;1047:163-171. DOI: 10.1016/j.aca.2018.10.002
  100. 100. Khan IM, Zhao S, Niazi S, Mohsin A, Shoaib M, Duan N, et al. Silver nanoclusters based FRET aptasensor for sensitive and selective fluorescent detection of T-2 toxin. Sensors and Actuators B: Chemical. 2018;277:328-335. DOI: 10.1016/j.snb.2018.09.021
  101. 101. Bruno JG, Carrillo MP, Phillips T, Hanson D, Bohmann JA. DNA aptamer beacon assay for C-telopeptide and handheld fluorometer to monitor bone resorption. Journal of Fluorescence. 2011;44:2021. DOI: 10.1007/s10895-011-0903-6
  102. 102. Xiao Y, Lubin AA, Heeger AJ, Plaxco KW. Label-free electronic detection of thrombin in blood serum by using an aptamer-based sensor. Angewandte Chemie (International Ed. in English). 2005;44:5456-5459. DOI: 10.1002/anie.200500989
  103. 103. Radi AE, Acero Sánchez JL, Baldrich E, O’Sullivan CK. Reagentless, reusable, ultrasensitive electrochemical molecular beacon aptasensor. Journal of the American Chemical Society. 2006;128:117-124. DOI: 10.1021/ja053121d
  104. 104. Bai L, Chen Y, Bai Y, Chen Y, Zhou J, Huang A. Fullerene-doped polyaniline as new redox nanoprobe and catalyst in electrochemical aptasensor for ultrasensitive detection of Mycobacterium tuberculosis MPT64 antigen in human serum. Biomaterials. 2017;133:11-19. DOI: 10.1016/j.biomaterials.2017.04.010
  105. 105. Kumari P, Lavania S, Tyagi S, Dhiman A, Rath D, Anthwal D, et al. A novel aptamer-based test for the rapid and accurate diagnosis of pleural tuberculosis. Analytical Biochemistry. 2019;564-565:80-87. DOI: 10.1016/j.ab.2018.10.019
  106. 106. Chang BY, Park SM. Electrochemical impedance spectroscopy. Annual Review of Analytical Chemistry. 2010;3:207-229. DOI: 10.1146/annurev.anchem.012809.102211
  107. 107. Figueroa-Miranda G, Feng L, Shiu SCC, Dirkzwager RM, Cheung YW, Tanner JA, et al. Aptamer-based electrochemical biosensor for highly sensitive and selective malaria detection with adjustable dynamic response range and reusability. Sensors and Actuators B: Chemical. 2018;255:235-243. DOI: 10.1016/j.snb.2017.07.117
  108. 108. Cheung YW, Kwok J, Law AWL, Watt RM, Kotaka M, Tanner JA. Structural basis for discriminatory recognition of Plasmodium lactate dehydrogenase by a DNA aptamer. Proceedings of the National Academy of Sciences. 2013;110:15967. DOI: 10.1073/pnas.1309538110
  109. 109. Arroyo-Currás N, Scida K, Ploense KL, Kippin TE, Plaxco KW. High surface area electrodes generated via electrochemical roughening improve the signaling of electrochemical aptamer-based biosensors. Analytical Chemistry. 2017;89:12185-12191. DOI: 10.1021/acs.analchem.7b02830
  110. 110. Barros WRP, Wei Q , Zhang G, Sun S, Lanza MRV, Tavares AC. Oxygen reduction to hydrogen peroxide on Fe3O4 nanoparticles supported on printex carbon and graphene. Electrochimica Acta. 2015;162:263-270. DOI: 10.1016/j.electacta.2015.02.175
  111. 111. Cheon HJ, Lee SM, Kim SR, Shin HY, Seo YH, Cho YK, et al. Colorimetric detection of MPT64 antibody based on an aptamer adsorbed magnetic nanoparticles for diagnosis of tuberculosis. Journal of Nanoscience and Nanotechnology. 2019;19:622-626. DOI: 10.1166/jnn.2019.15905
  112. 112. Wang CH, Wu JJ, Lee GB. Screening of highly-specific aptamers and their applications in paper-based microfluidic chips for rapid diagnosis of multiple bacteria. Sensors and Actuators B: Chemical. 2019;284:395-402. DOI: 10.1016/j.snb.2018.12.112
  113. 113. Lee BH, Kim SH, Ko Y, Park JC, Ji S, Gu MB. The sensitive detection of ODAM by using sandwich-type biosensors with a cognate pair of aptamers for the early diagnosis of periodontal disease. Biosensors and Bioelectronics. 2019;126:122-128. DOI: 10.1016/j.bios.2018.10.040
  114. 114. Zhang Z, Oni O, Liu J. New insights into a classic aptamer: Binding sites, cooperativity and more sensitive adenosine detection. Nucleic Acids Research. 2017;45:7593-7601. DOI: 10.1093/nar/gkx517
  115. 115. Baird GS. Where are all the aptamers? American Journal of Clinical Pathology. 2010;134:529-531. DOI: 10.1309/AJCPFU4CG2WGJJKS
  116. 116. Jiang Y, Ma W, Ji W, Wei H, Mao L. Aptamer superstructure-based electrochemical biosensor for sensitive detection of ATP in rat brain with in vivo microdialysis. The Analyst. 2019. DOI: 10.1039/c8an02077a
  117. 117. Lakhin AV, Tarantul VZ, Gening LV. Aptamers: Problems. Solutions and Prospects. Acta Naturae. 2013;5:34-43
  118. 118. Liu J, Cao Z, Lu Y. Functional nucleic acid sensors. Chemical Reviews. 2009;109:1948-1998. DOI: 10.1021/cr030183i
  119. 119. Díaz-Fernández A, Miranda-Castro R, de-los-Santos-Álvarez N, Rodríguez EF, Lobo-Castañón MJ. Focusing aptamer selection on the glycan structure of prostate-specific antigen: Toward more specific detection of prostate cancer. Biosensors and Bioelectronics. 2019;128:83-90. DOI: 10.1016/j.bios.2018.12.040
  120. 120. Kaur H. Recent developments in cell-SELEX technology for aptamer selection. Biochimica et Biophysica Acta (BBA)—General Subjects. 2018;1862:2323-2329. DOI: 10.1016/j.bbagen.2018.07.029
  121. 121. 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. DOI: 10.3390/ijms17122079
  122. 122. 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;49:2238-2241. DOI: 10.1002/anie.200903998
  123. 123. Sachan A, Ilgu M, Kempema A, Kraus GA, Nilsen-Hamilton M. Specificity and ligand affinities of the cocaine aptamer: Impact of structural features and physiological NaCl. Analytical Chemistry. 2016;88:7715-7723. DOI: 10.1021/acs.analchem.6b01633
  124. 124. Babalik A, Ulus IH, Bakirci N, Kuyucu T, Arpag H, Dagyildiz L, et al. Pharmacokinetics and serum concentrations of antimycobacterial drugs in adult Turkish patients. The International Journal of Tuberculosis and Lung Disease. 2013;17:1442-1447. DOI: 10.5588/ijtld.12.0771
  125. 125. Casu GS, Hites M, Jacobs F, Cotton F, Wolff F, Beumier M, et al. Can changes in renal function predict variations in beta-lactam concentrations in septic patients? International Journal of Antimicrobial Agents. 2013;42:422-428. DOI: 10.1016/j.ijantimicag.2013.06.021
  126. 126. Varatharajan S, Panetta JC, Abraham A, Karathedath S, Mohanan E, Lakshmi KM, et al. Population pharmacokinetics of Daunorubicin in adult patients with acute myeloid leukemia. Cancer Chemotherapy and Pharmacology. 2016;78:1051-1058. DOI: 10.1007/s00280-016-3166-8
  127. 127. Andriguetti NB, Raymundo S, Antunes MV, Perassolo MS, Verza SG, Suyenaga ES, et al. Pharmacogenetic and pharmacokinetic dose individualization of the taxane chemotherapeutic drugs paclitaxel and docetaxel. Current Medicinal Chemistry. 2017;24:3559-3582. DOI: 10.2174/0929867324666170623093445
  128. 128. Motohashi S, Mino Y, Hori K, Naito T, Hosokawa S, Furuse H, et al. Interindividual variations in aprepitant plasma pharmacokinetics in cancer patients receiving cisplatin-based chemotherapy for the first time. Biological and Pharmaceutical Bulletin. 2013;36:676-681
  129. 129. Contin M, Alberghini L, Candela C, Benini G, Riva R. Intrapatient variation in antiepileptic drug plasma concentration after generic substitution vs stable brand-name drug regimens. Epilepsy Research. 2016;122:79-83. DOI: 10.1016/j.eplepsyres.2016.02.012
  130. 130. Pisupati J, Jain A, Burckart G, Hamad I, Zuckerman S, Fung J, et al. Intraindividual and interindividual variations in the pharmacokinetics of mycophenolic acid in liver transplant patients. Journal of Clinical Pharmacology. 2005;45:34-41. DOI: 10.1177/0091270004270145
  131. 131. Albrecht D, Turakhia MP, Ries D, Marbury T, Smith W, Dillon D, et al. Pharmacokinetics of tecarfarin and warfarin in patients with severe chronic kidney disease. Thrombosis and Haemostasis. 2017;117:2026-2033. DOI: 10.1160/TH16-10-0815
  132. 132. Deitchman AN, Derendorf H. Measuring drug distribution in the critically ill patient. Advanced Drug Delivery Reviews. 2014;77:22-26. DOI: 10.1016/j.addr.2014.08.014
  133. 133. Pena MA, Horga JF, Zapater P. Variations of pharmacokinetics of drugs in patients with cirrhosis. Expert Review of Clinical Pharmacology. 2016;9:441-458. DOI: 10.1586/17512433.2016.1135733
  134. 134. Salem F, Abduljalil K, Kamiyama Y, Rostami-Hodjegan A. Considering age variation when coining drugs as high versus low hepatic extraction ratio. Drug Metabolism and Disposition. 2016;44:1099-1102. DOI: 10.1124/dmd.115.067595
  135. 135. Nieuweboer AJ, Smid M, de Graan AM, Elbouazzaoui S, de Bruijn P, Eskens FA, et al. Role of genetic variation in docetaxel-induced neutropenia and pharmacokinetics. The Pharmacogenomics Journal. 2016;16:519-524. DOI: 10.1038/tpj.2015.66
  136. 136. Xie XC, Li J, Wang HY, Li HL, Liu J, Fu Q , et al. Associations of UDP-glucuronosyltransferases polymorphisms with mycophenolate mofetil pharmacokinetics in Chinese renal transplant patients. Acta Pharmacologica Sinica. 2015;36:644-650. DOI: 10.1038/aps.2015.7
  137. 137. Jiang F, Kim HD, Na HS, Lee SY, Seo DW, Choi JY, et al. The influences of CYP2D6 genotypes and drug interactions on the pharmacokinetics of venlafaxine: Exploring predictive biomarkers for treatment outcomes. Psychopharmacology (Berlin). 2015;232:1899-1909. DOI: 10.1007/s00213-014-3825-6
  138. 138. Bertholee D, Maring JG, van Kuilenburg ABP. Genotypes affecting the pharmacokinetics of anticancer drugs. Clinical Pharmacokinetics. 2017;56:317-337. DOI: 10.1007/s40262-016-0450-z
  139. 139. Elewa H, Wilby KJ. A review of pharmacogenetics of antimalarials and associated clinical implications. European Journal of Drug Metabolism and Pharmacokinetics. 2017;42:745-756. DOI: 10.1007/s13318-016-0399-1
  140. 140. Wong AL, Seng KY, Ong EM, Wang LZ, Oscar H, Cordero MT, et al. Body fat composition impacts the hematologic toxicities and pharmacokinetics of doxorubicin in Asian breast cancer patients. Breast Cancer Research and Treatment. 2014;144:143-152. DOI: 10.1007/s10549-014-2843-8
  141. 141. Janukonyte J, Parkner T, Lauritzen T, Christiansen JS, Laursen T. Circadian variation in the pharmacokinetics of steady state continuous subcutaneous infusion of growth hormone in adult growth hormone deficient patients. Growth Hormone and IGF Research. 2013;23:256-260. DOI: 10.1016/j.ghir.2013.09.002
  142. 142. Muzzarelli S, Brunner-La Rocca H, Pfister O, Foglia P, Moschovitis G, Mombelli G, et al. Adherence to the medical regime in patients with heart failure. European Journal of Heart Failure. 2010;12:389-396. DOI: 10.1093/eurjhf/hfq015
  143. 143. Suttorp M, Bornhauser M, Metzler M, Millot F, Schleyer E. Pharmacology and pharmacokinetics of imatinib in pediatric patients. Expert Review of Clinical Pharmacology. 2018;11:219-231. DOI: 10.1080/17512433.2018.1398644
  144. 144. Bar-Dayan Y, Shotashvily T, Boaz M, Wainstein J. Error in drugs consumption among older patients. American Journal of Therapeutics. 2017;24:e701-e705. DOI: 10.1097/MJT.0000000000000400
  145. 145. Dearling JLJ, Packard AB. Molecular imaging in nanomedicine—A developmental tool and a clinical necessity. Journal of Controlled Release. 2017;261:23-30. DOI: 10.1016/j.jconrel.2017.06.011

Written By

Soma Banerjee and Marit Nilsen-Hamilton

Submitted: 11 November 2018 Reviewed: 20 May 2019 Published: 22 June 2019