MicroRNA biosensors

From Wikipedia, the free encyclopedia
Figure illustrating the workflow of miRNA detection. miRNAs can be detected with complex biosensors such as electrochemical biosensors and optical biosensors. miRNA Biosensors utilize nanomaterials, recognition elements, and amplification elements for sensitive and specific detection of miRNAs. Created with BioRender.com

MicroRNA (miRNA) biosensors are analytical devices that involve interactions between the target miRNA strands and recognition element on a detection platform to produce signals that can be measured to indicate levels or the presence of the target miRNA. Research into miRNA biosensors shows shorter readout times, increased sensitivity and specificity of miRNA detection and lower fabrication costs than conventional miRNA detection methods.[1]

miRNAs are a category of small, non-coding RNAs in the range of 18-25 base pairs in length.[1] miRNAs regulate cellular processes such as gene regulation post-transcriptionally, and are abundant in body fluids such as saliva, urine and circulatory fluids such as blood. Also, miRNAs are found in animals and plants and have regulatory functions that affect cellular mechanisms. miRNAs are highly associated with diseases such as cancers and cardiovascular diseases. In cancer, miRNAs have oncogenic or tumor suppressor roles and are promising biomarkers for disease diagnosis and prognosis.[1] Many techniques exist in clinical and research settings for analyzing miRNA biomarkers. However, inherent limitations with current methods, such as high cost, time and personnel training requirements, and low detection sensitivity and specificity, create the need for improved miRNA detection methods.[1]

Background[edit]

miRNAs are associated with physiological and pathological processes; hence, measuring them in fields like human health, agriculture, and environmental testing is in demand. Here are some key aspects of the necessity of detection of miRNAs:

  • Potential biomarkers: miRNAs have specific expression in diseases such as cancer, cardiovascular diseases, and autoimmune diseases, which can be beneficial for early detection, prognosis and monitoring for response to treatments.[2] Furthermore, because miRNAs are in body fluids like urine, saliva, and blood, detecting miRNAs is less invasive than methods such as biopsies. This is more comfortable for patients and can facilitate more frequent monitoring of their disease.[3]
  • Molecular mechanisms: As miRNAs have regulatory roles in gene expression and signaling pathways, studying them can give the etiology of diseases and targeting them can provide therapeutic options.[4][5]
  • Personalized medicine: Because the specific expression of miRNAs offers a promising avenue for enhancing personalized medicine, they provide a deeper understanding of individual disease risk, treatment response, and prognosis, which help clinicians make better informed clinical decisions.[6]

History of miRNA detection technology[edit]

Early and current detection methods[edit]

The first miRNA (lin-4) was detected by Victor Ambros in Caenorhabditis elegans in 1993.[7] The first detection method was Northern blotting (1977), which had low sensitivity. Following that was Reverse Transcription Polymerase Chain Reaction (RT-PCR) (1990), which had high detection sensitivity.[8][9]

  • Northern Blotting: Northern blotting involves hybridizing miRNA probes (short nucleic acid sequences) with miRNAs, followed by their separation on a gel and transfer to a membrane. The probes are labeled with radioactive isotopes (raising safety and environmental concerns), enzymes, or fluorescent markers. The quantity of RNA present is inferred from the probe signal’s intensity. While Northern Blotting is highly specific and helpful for validating high-throughput methods like RNA-seq, it requires a large sample volume, is time-consuming, and lacks precision in quantification analysis.[10][11]
  • Real-Time Reverse Transcription–Polymerase Chain Reaction (Real-time RT-PCR): This method starts with converting miRNA into cDNA using reverse transcriptase enzymes. The cDNA is then amplified using sequence-specific primers, a process monitored by fluorescent dyes or probes. Real-time RT-PCR is noted for its sensitivity and specificity. However, it faces challenges such as the need for standardization, technical complexities (e.g., primer design, sample preparation), time-intensive processes, and high costs.[12][13]

High-throughput Methods:

  • Microarrays (1990): Microarrays enable the detection of thousands of miRNAs in a single experiment. They consist of a solid surface to which complementary miRNA sequences are attached. Introducing miRNAs allows them to bind to these probes, with the amount of miRNA measured by the fluorescence intensity. Microarrays are cost-effective compared to real-time RT-PCR and NGS but have limitations in detecting low quantities of miRNAs and distinguishing between miRNAs with similar sequences.[14][15]
  • Next Generation Sequencing (NGS) (2005): NGS begins with RNA extraction and reverse transcription into cDNA, followed by adaptor ligation and amplification. The cDNA is then sequenced on an NGS platform, producing millions of short reads. Expert bioinformaticians and sophisticated tools must align and analyze the data and map reads to reference miRNA sequences for miRNA discovery and identification. NGS offers high sensitivity and specificity for detecting low-quantity miRNAs and identifying miRNAs differing by a single nucleotide.[16][17]

Principles of microRNA biosensors[edit]

Three essential elements make up miRNA biosensors:

  • Biological recognition element: they can detect specific target molecules and have different types, including antibodies, antigens, DNA/RNA, aptamers, enzymes, and MIPs (molecularly imprinted polymers).[18][19]
  • Transducer: following recognition, the transducer is an element required to convert changes in the recognition element to a measurable signal. Based on the type of signal they produce, they are categorized into electrochemical, optical, and mechanical transducers[20]
  • Signal processor: computational elements that amplify and process the signals produced from transducers and can be demonstrated by numerical values and digital readouts.[21][22]

Specificity in miRNA detection[edit]

The term “specificity” in the context of miRNA biosensors refers to the ability of the biosensor to identify a particular miRNA within a sample that contains various components and miRNAs with similar sequences. The challenge in achieving this specificity derives from the small size of miRNAs, which may differ from each other by only one nucleotide. Consequently, designing biosensors capable of precisely recognizing the target miRNA is essential.[23]

Sensitivity in miRNA detection[edit]

Sensitivity in miRNA biosensors refers to their ability to detect target miRNAs in low concentrations within samples. Since miRNAs are typically found in small amounts, biosensors are engineered to identify concentrations as low as femtomolar (10^-15) or attomolar (10^-18) levels. Achieving such high sensitivity involves enhancements to recognition elements, amplification, and signal processing techniques. The LoD (limit of detection) is used to determine the concrete value of sensitivity in biosensors, which indicates the lowest concentration of miRNA that can be separated from the background (zero) signal with a specified level of confidence.[24]

The dynamic range in miRNA biosensors refers to the concentrations over which the biosensor can accurately detect the target miRNAs, extending from the lowest detectable LoD to the maximum concentration that can be measured without necessitating sample dilution.[25][26]

Types of microRNA biosensors[edit]

Electrochemical biosensors[edit]

General mechanism of a label-based electrochemical miRNA biosensor. Created with BioRender.com

Electrochemical biosensors present significant advantages to miRNA detection over conventional miRNA analysis methods. Using simple electronics reduces production costs and increases ease of use in portable system configurations. This allows for a broader scope of use, including environmental, clinical and food analysis applications.[27]

miRNA electrochemical biosensor detection relies on measuring the changes in the electrode-property or electroactive compound redox signal in the transduction of electrochemically active reporter species and hybridization between the target miRNA and complementary probe. Various materials can be made into the transduction element, including silver, gold, graphite or nanoparticle variations of such materials. Detection of electrochemical property changes allows for real-time analysis and kinetics data, an advantage biosensor methods such as optical biosensors lack. Light pollution is not a limitation of electrochemical miRNA biosensors. However, amplification techniques such as rolling circle amplification (RCA) may be required when miRNA concentrations are insufficient to produce an electrical signal.[28]

1. Voltammetric and amperometric electrochemical biosensors[edit]

Electrochemical miRNA biosensors can be designed to infer voltammetric or amperometric measurements. Upon hybridization of the miRNA target with its complementary probe sequence, voltammetric miRNA biosensors detect the change in current based on a controlled increase or decrease in electric potential on the detection platform. Amperometric-based biosensors detect the change in electric current at a fixed positive electric potential.[1] Recent developments in voltammetric and amperometric miRNA biosensors can be classified as label-based or label-free biosensors, indicating whether or not electroactive labels on the miRNA target are used as the naming suggests.[1]

  • Voltammetric and amperometric label-free (direct detection) miRNA biosensors
First published in 2009, label-free (direct detection) electrochemical miRNA biosensors function without labelling the target miRNA with electrocatalytic nanoparticle tags or hybridization indicators.[29] Label-free miRNA biosensors were initially based on DNA detection through guanine electrooxidation measurements, with the lower detection limit being 5 nM of miRNA. Since then, electrode materials have been developed to increase the sensitivity of detection down to less than 1 pM, such as with graphene and ionic-liquid modified electrodes.[30][31] For example, Wu et al. (2013) increased the conductivity of the electrode surface of an amperometric biosensor with a multilayer consisting of Nafion, thionine and palladium nanoparticles, which immobilized the target miRNA on the electrode surface for a lower limit of detection of 1.87 pM.[32] Label-free miRNA biosensors detect signals before and after the hybridization of electroactive nucleic acid bases.[33] For instance, doxorubicin-loaded gold nanoparticles (AuNps) have been integrated with a double-loop hairpin probe that hybridizes with the target miRNA to form heteroduplexes, in which duplex specific nucleases hydrolyze DNA in the heteroduplex structures to released target miRNA strands for amplification in a signal amplification system. The limit of detection in such a system is 0.17pM.[34]
  • Voltammetric and amperometric label-based (indirect detection) miRNA biosensors
Label-based (indirect detection) electrochemical miRNA biosensors require electrocatalytic or redox active molecule or nanoparticle labelling of the miRNA target or complementary capture probes for detection. Generally, label-based approaches offer significantly greater sensitivity of miRNA detection than label-free methods, with sensitivity reaching the fM-aM range.[35][36]
An example is AuNp-superlattice-based miRNA biosensors utilizing the small molecule cationic dye toluidine blue to detect miRNA-21. Toluidine blue acts as a miRNA intercalative label through electrostatic interaction with the negatively charged backbone phosphate groups. On the biosensor, toluidine blue is a redox indicator to measure the oxidation peak current of toluidine blue and indicated hybridization of miRNA. The LoD levels reached 78 aM.[37]

2. Amplification (enzyme)-based electrochemical miRNA biosensors[edit]

Electrochemical detection or amplification strategies for miRNA biosensors have been developed using enzyme-based methods. Amplification of miRNA is often a necessary component of biosensor detection as miRNA concentrations are found in low abundance, and amplification of target miRNA strands will increase the sensitivity of detection.[38] Additionally, inherent properties of miRNA include short strand length and high sequence homology, which present a challenge with detection sensitivity and specificity.[1]

Various methods, such as duplex-specific nuclease enzymes and polymerase extension, can amplify miRNA targets to reach LoD in the fM range.[1] Isothermal amplification techniques are widely used enzyme-based miRNA amplification techniques, given the advantages of cost and time-reduction associated with ease of use compared to polymerase chain reaction (PCR) methods. Isothermal methods amplify nucleic acids at a constant temperature, which removes the thermal cycling requirement as used in PCR and does not require specific enzymes for spatial recognition sites in the target miRNA.[1] A commonly used isothermal technique for miRNA detection is rolling circle amplification (RCA). In the RCA of miRNA targets, the miRNA binds to a complementary circular DNA template, which is continuously and exponentially amplified through the synthesis of long single-stranded DNA.[1] Research with gold electrode electrochemical biosensors has shown that RCA initiated on the electrode has provided LoD levels of 50 aM.[39] RCA's isothermal nature and ease of use allow it to be used in clinical diagnostic and resource-lacking laboratory settings and in point-of-care biosensor devices.[1]

Optical miRNA biosensors[edit]

Upon hybridization of the target miRNA tagged with a nucleic acid probe and an optically active reporter, label-based optical biosensors transduce the absorbance or fluorescence optical signal into quantifiable data. The reporters can be either quantum dots or dye labels.[1] On the other hand, label-free optical miRNA biosensors detect changes in the refractive index (RI) at the recognition element, which are caused by the binding of the target miRNA to its bioreceptor. The electromagnetic field probes the RI changes, characterized as an evanescent wave. The electromagnetic fields are generated by guided or resonant optical modes that travel in the transducer element.[40] Additionally, label-free optical miRNA biosensors are insensitive to unbound or background RNA or DNA molecules, as optical detection is confined to the sensing recognition surface. This is beneficial for miRNA detection in small volumes and is an advantage over other label-based miRNA biosensors, as signal detection is based on measuring the total number of miRNA in the sample.[40]

  • Surface Plasmon resonance-based optical miRNA biosensors
Surface plasmon resonance (SPR) based miRNA biosensors are a label-free method that detects RI changes after target miRNA binds to its probes and forms a complex. Detection involves propagating a surface plasmon wave (SPW) across the metal-dielectric interface surface layer of the biosensor in a Kretschmann configuration.[40] The SPW decays exponentially, where the changes in the SPW propagation constant are measured as the constant is sensitive to change in the RI.[41] A practical example of a label-based SPR-based miRNA biosensor is miR-21 detection with a LoD of 1 fM. The biosensor utilized graphene oxide–gold nanoparticles integrated with the sandwiching of the target miRNA between two DNA probes to amplify the SPR signal and have secondary hybridization through miR-21 report probes.[42]

Electromechanical biosensors[edit]

Electromechanical biosensors represent an integration of electrical and mechanical engineering disciplines, employing a detection strategy that hinges on the hybridization of miRNAs to specific probes anchored on the sensor’s surface. Subsequent alterations in parameters such as stress or mass are then transduced into electrical signals. A notable implementation involves Atomic Force Microscopy (AFM), which has successfully identified has-mir-194 and has-mir-205 in samples related to colon and bladder cancer.[43] The underlying mechanism of this approach is AFM’s ability to delineate the variations in stiffness across the gold surface of the biosensor, facilitating the detection of miRNA hybridization events. Another pivotal component in electromechanical biosensors is the gold-coated piezoelectric cantilever sensor, which is adept at recognizing hybridized miRNA.[44] Although electromechanical biosensors are highly sensitive to miRNAs, it is difficult to measure them in samples with high amounts of different molecules.[1]

Nanomaterials used in miRNA biosensors[edit]

Nanomaterials are used for their unique characteristics to facilitate the detection of miRNAs. Here, we discuss some features of nanomaterials used in miRNA biosensors.[35]

  • Gold nanoparticles (AuNps): AuNps enhance miRNA detection signals and facilitate the stable conjugation of recognition elements into miRNAs.[45] AuNps have excellent catalytic properties, conductivity, high surface area and interface energy and can be modified with molecules such as oligonucleotide aggregates for high affinity binding with specific substrates.[46]
In electrochemical miRNA biosensors, AuNps allow for ease of functionalization for electrochemical reactions that involve changes in potential, current, conductivity, or impedance in detecting target miRNA binding on the detection surface.[46] In optical biosensors, AuNps exhibit unique and tunable optical properties beneficial for SPR miRNA biosensors.[46] When AuNps are exposed to light, propagating surface plasmons needed for detecting receptor-bonded miRNAs are created from a resonant interaction between the electromagnetic field of light and the electron-charged oscillations on the metal surface. This is due to AuNps exhibiting a high density of conduction band electrons and its nanoparticle size allowing multiple angular shifts for more reflectance angles.[46][47]
  • Graphene: Graphene is a member of the carbon nanomaterials family and stands out for its biocompatibility, electrical conductivity, light molecular weight, stability, and affordability, making it an exceptional choice for miRNA biosensor applications. It demonstrates excellent responsiveness to chemical, optical, and mechanical stimuli. Graphene is predominantly utilized in electrical and optical miRNA biosensors.[48] A notable recent application involves using laser-induced self-N-doped porous graphene in miRNA biosensors, capable of detecting miRNA hsa-miR-486-5p at concentrations as low as 10 fM. This approach combines cost-effectiveness with high reproducibility, offering significant advantages for conditions like preeclampsia.[49]
  • Terahertz (THz) Metamaterial with Gold Nanoparticles: THz metamaterial is artificially synthesized and designed to interact with THz frequency waves. When combined with AuNps and after binding with target miRNA, they produce higher changes in THz spectral regions. For instance, a miRNA biosensor based on these materials could detect the miRNA-21 from clinical samples with a LoD of 14.54 aM.[45]

Technologies and principles of multiplex miRNA biosensors[edit]

Multiplex miRNA biosensors are designed to detect multiple types of miRNAs simultaneously with high specificity and sensitivity. This capability is essential for several reasons: First, it allows for detecting various miRNAs within a single sample that may contribute to disease, enabling comprehensive monitoring during treatment while facilitating high-throughput screening. Second, it can significantly reduce cost and time by allowing the simultaneous analysis of data from multiple miRNAs.[50][51] Here are some recent technologies in multiplex miRNA biosensors:

  • DNA-PAINT based using a DNA origami-based sensor platform - this miRNA biosensor has a unique geometric barcoding system and can detect up to 4 miRNAs at the same time. The 52nm distance intervals between strands enable the platform to distinguish between single mismatches to the LoD of 11 fM to 388 fM.[52]
  • CRISPR-Multiplex Biosensor- this platform utilizes various technologies, including electrochemical microfluidics and Cas13a, to enable the amplification-free detection of eight miRNAs. It features a design with four divided channels for electrochemical analysis.[53]

Applications[edit]

Diagnostic and prognostic applications[edit]

Since the initial discovery of miRNAs, large databases of miRNAs have been identified in humans, plants and animals. As many miRNAs are associated with disease onset and development, miRNAs are a suitable biomarker for biosensor detection in clinical settings.[1] Considerations must be taken into account of the biological sample source for miRNA targets. Clinical miRNA sample analysis commonly comes in blood, plasma, serum, seminal fluid, saliva, urine, and tissue-derived miRNAs.[54] In the context of cancer, biosensor detection of miRNAs is most conveniently performed in the form of liquid biopsies, as circulatory miRNAs are found in the highest abundance in liquid samples.[55]

  • Point-of-Care (POC) testing
Research into POC diagnostic tests has resulted in the development of microfluidic biosensors capable of early diagnostic clinical analysis of cancer-associated miRNAs, which produce cost- and time-efficient results with increased sensitivity and specificity over traditional methods.[56] Liquid biopsy droplet-based microfluidic biosensors can be fabricated into POC devices for ease of use by integrating with pre-existing devices and interfaces and can extend utilization beyond traditional laboratory settings and those without sophisticated instruments.[57] An example of developments in POC testing for prostate cancer is where miR-21 in low concentrations of urine samples was detected with a limit of detection of 2 nM on screen-printed, label-based electrochemical biosensor chips. Detection was rapid, with results produced in less than two hours.[58]

Agriculture management[edit]

Besides clinical usage, miRNA biosensors have been adapted for managing agriculture plant stress and growth and disease analysis, as plant miRNAs are associated with growth regulatory mechanisms. An example is electrochemical biosensors fabricated for detecting miR-319a, a miRNA associated with phytohormone response that regulates rice seedling growth regulation. Isothermal alkaline phosphatase catalytic signal amplification of the target miRNA strands was integrated with a three-electrode system to detect miR319a to LoD levels of 1.7 fM.[59] AuNp label-based optical biosensors were tested for detecting miRNA-1886, an indicator of drought stress in tomato plants. They found that decreasing irrigation levels increased the concentration of miRNA-1886 at a range of 100 to 6800 fM.[60]

Research applications[edit]

1. Molecular and cellular biology[edit]

As miRNAs are one of the main regulators of genes, detection and measuring them in cells and molecular levels can be helpful to decipher miRNA interactions with other molecules. For instance, a study by Bandi et al. found that miR-15a and miR-16 function in tumorigenesis of non-small cell lung cancer (NSCLC) cell lines.[61] miRNA biosensors also have a significant role in the elucidation of disease mechanisms. For example, a study on cardiovascular diseases found that miRNA biosensors based on DNA tetrahedron nanostructure can recognize miR-133a in aM levels, which is helpful for further studies on myocardial infarction.[62]

2. Drug discovery and development[edit]

Because of their high-throughput potential, miRNA biosensors can significantly accelerate drug discovery by evaluating various drugs on miRNA expression levels to observe which drug can target unregulated miRNAs in diseases. Furthermore, miRNA biosensors can monitor the expression of miRNA expression in real-time to observe which changes happen in different concentrations of drugs, and this is especially crucial in early-phase clinical trials for drug dosage optimization. In addition, by testing various miRNA expressions, researchers can discover relations between diseases and miRNAs’ expression[63][64][65]

Limitations to miRNA biosensors[edit]

While miRNA biosensors hold considerable promise for miRNA detection, several critical challenges must be addressed:

  • Sensitivity and Specificity: The low abundance of miRNAs in complex biological samples, such as blood, necessitates enhancing biosensor sensitivity to detect miRNAs at levels beyond femtomolar concentrations. Additionally, due to the high sequence similarity among miRNAs, improving the specificity of these biosensors is essential to differentiate between miRNAs based on single nucleotide differences.[66]
  • Sample Preparation: Extracting miRNAs from samples presents significant difficulties. The process is complex and requires optimization to ensure the purity and integrity of the miRNAs for accurate detection.[67]
  • Stability of miRNA Biosensor: The stability of miRNA biosensors is compromised by environmental conditions, particularly for components like aptamers and antibodies. This issue is especially pertinent for point-of-care (POC) devices, which require robustness and longevity to be effectively used in various settings.[68]
  • Standardization: A significant limitation in the field is the absence of standardized guidelines and universal reference miRNAs for comparing results across blood and plasma samples. Establishing reliable normalizers, characterized by consistent expression and stability across all samples, is crucial for accurately interpreting miRNA levels.[69][70][71]

Addressing these challenges is essential for advancing and adopting miRNA biosensor technologies.

Future directions[edit]

The significance of miRNA in diagnostics and the recent advancements in miRNA detection from various sample sources, particularly in clinical settings, underscore the need for enhancing miRNA biosensor technologies. The future of miRNA biosensor optimization encompasses several key areas:

  • Furthering nanomaterial integration research: The materials, including graphene, gold nanoparticles, and quantum dots, can significantly improve the biosensors’ specificity and sensitivity, making them more effective in detecting miRNAs.[72]
  • Multiplex detection: Efforts are underway to refine miRNA biosensors for the simultaneous detection of multiple miRNA types, especially those within the same family, from small-volume samples; in this regard, artificial intelligence can aid in distinguishing between miRNA types and correlating them with clinical outcomes. Such advancements would be particularly beneficial for point-of-care (POC) devices, simplifying sample preparation, enhancing user-friendliness, and enabling physicians to monitor miRNA levels in real-time remotely.[67]
  • Encapsulation technologies: Encapsulation technologies aim to safeguard the biosensors’ sensitive components from environmental threats, ensuring their durability and reliability.[73]
  • Standardization of miRNA research and development: The development of standardized guidelines and the identification of universal genes for miRNA expression comparison will facilitate the accurate evaluation of miRNA biosensors across different clinical scenarios.[1]
  • Clinical Sample Analysis: The study of prospective and retrospective analyses of clinical samples and comparing miRNA biosensor results with those obtained via real-time qPCR and sequencing technologies can assess biosensor performance under varied clinical conditions.[74][75]

These advancements suggest a focused trajectory for miRNA biosensor development, aiming at technological enhancements that promise improved diagnostic capabilities and clinical applications.

References[edit]

  1. ^ a b c d e f g h i j k l m n o Kilic, Tugba; Erdem, Arzum; Ozsoz, Mehmet; Carrara, Sandro (2018-01-15). "microRNA biosensors: Opportunities and challenges among conventional and commercially available techniques". Biosensors and Bioelectronics. 99: 525–546. doi:10.1016/j.bios.2017.08.007. ISSN 0956-5663. PMID 28823978.
  2. ^ Calin, George Adrian; Dumitru, Calin Dan; Shimizu, Masayoshi; Bichi, Roberta; Zupo, Simona; Noch, Evan; Aldler, Hansjuerg; Rattan, Sashi; Keating, Michael; Rai, Kanti; Rassenti, Laura; Kipps, Thomas; Negrini, Massimo; Bullrich, Florencia; Croce, Carlo M. (2002-11-26). "Frequent deletions and down-regulation of micro- RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia". Proceedings of the National Academy of Sciences. 99 (24): 15524–15529. Bibcode:2002PNAS...9915524C. doi:10.1073/pnas.242606799. ISSN 0027-8424. PMC 137750. PMID 12434020.
  3. ^ Ps, Mitchell (2008). "Circulating microRNAs as stable blood-based markers for cancer detection". Proc Natl Acad Sci U S A. 105 (30): 10513–10518. Bibcode:2008PNAS..10510513M. doi:10.1073/pnas.0804549105. PMC 2492472. PMID 18663219.
  4. ^ Krützfeldt, Jan; Rajewsky, Nikolaus; Braich, Ravi; Rajeev, Kallanthottathil G.; Tuschl, Thomas; Manoharan, Muthiah; Stoffel, Markus (December 2005). "Silencing of microRNAs in vivo with 'antagomirs'". Nature. 438 (7068): 685–689. Bibcode:2005Natur.438..685K. doi:10.1038/nature04303. ISSN 1476-4687. PMID 16258535. S2CID 4414240.
  5. ^ Bartel, David P (January 2004). "MicroRNAs". Cell. 116 (2): 281–297. doi:10.1016/s0092-8674(04)00045-5. ISSN 0092-8674. PMID 14744438.
  6. ^ Garzon, Ramiro; Calin, George A.; Croce, Carlo M. (2009-02-01). "MicroRNAs in Cancer". Annual Review of Medicine. 60 (1): 167–179. doi:10.1146/annurev.med.59.053006.104707. ISSN 0066-4219. PMID 19630570.
  7. ^ Lindow, Morten; Kauppinen, Sakari (2012-10-29). "Discovering the first microRNA-targeted drug". Journal of Cell Biology. 199 (3): 407–412. doi:10.1083/jcb.201208082. ISSN 1540-8140. PMC 3483128. PMID 23109665.
  8. ^ Alwine, J C; Kemp, D J; Stark, G R (December 1977). "Method for detection of specific RNAs in agarose gels by transfer to diazobenzyloxymethyl-paper and hybridization with DNA probes". Proceedings of the National Academy of Sciences. 74 (12): 5350–5354. Bibcode:1977PNAS...74.5350A. doi:10.1073/pnas.74.12.5350. ISSN 0027-8424. PMC 431715. PMID 414220.
  9. ^ Mocharla, H.; Mocharla, R.; Hodes, M. E. (1990-09-14). "Coupled reverse transcription-polymerase chain reaction (RT-PCR) as a sensitive and rapid method for isozyme genotyping". Gene. 93 (2): 271–275. doi:10.1016/0378-1119(90)90235-j. ISSN 0378-1119. PMID 1699848.
  10. ^ Trayhurn, Paul (March 1996). "Northern blotting". Proceedings of the Nutrition Society. 55 (1B): 583–589. doi:10.1079/PNS19960051. ISSN 1475-2719. PMID 8832821.
  11. ^ Green, Michael R.; Sambrook, Joseph (2022-02-01). "Analysis of RNA by Northern Blotting". Cold Spring Harbor Protocols. 2022 (2): pdb.top101741. doi:10.1101/pdb.top101741. ISSN 1940-3402. PMID 35105788. S2CID 246474418.
  12. ^ Chen, Caifu; Tan, Ruoying; Wong, Linda; Fekete, Richard; Halsey, Jason (2011), Park, Daniel J. (ed.), "Quantitation of MicroRNAs by Real-Time RT-QPCR", PCR Protocols, Methods in Molecular Biology, vol. 687, Totowa, NJ: Humana Press, pp. 113–134, doi:10.1007/978-1-60761-944-4_8, ISBN 978-1-60761-944-4, PMID 20967604, retrieved 2024-03-01
  13. ^ Hardikar, Anandwardhan A.; Farr, Ryan J.; Joglekar, Mugdha V. (2014-01-27). "Circulating microRNAs: Understanding the Limits for Quantitative Measurement by Real-Time PCR". Journal of the American Heart Association. 3 (1): e000792. doi:10.1161/jaha.113.000792. ISSN 2047-9980. PMC 3959687. PMID 24572259.
  14. ^ Li, Wei; Ruan, Kangcheng (June 2009). "MicroRNA detection by microarray". Analytical and Bioanalytical Chemistry. 394 (4): 1117–1124. doi:10.1007/s00216-008-2570-2. ISSN 1618-2650. PMID 19132354. S2CID 42414331.
  15. ^ Nelson, Peter T.; Baldwin, Don A.; Scearce, L. Marie; Oberholtzer, J. Carl; Tobias, John W.; Mourelatos, Zissimos (November 2004). "Microarray-based, high-throughput gene expression profiling of microRNAs". Nature Methods. 1 (2): 155–161. doi:10.1038/nmeth717. ISSN 1548-7105. PMID 15782179. S2CID 17917521.
  16. ^ Lopez, Juan Pablo; Diallo, Alpha; Cruceanu, Cristiana; Fiori, Laura M.; Laboissiere, Sylvie; Guillet, Isabelle; Fontaine, Joelle; Ragoussis, Jiannis; Benes, Vladimir; Turecki, Gustavo; Ernst, Carl (2015-07-01). "Biomarker discovery: quantification of microRNAs and other small non-coding RNAs using next generation sequencing". BMC Medical Genomics. 8 (1): 35. doi:10.1186/s12920-015-0109-x. ISSN 1755-8794. PMC 4487992. PMID 26130076.
  17. ^ Song, Chunjiao; Chen, Huan; Wang, Tingzhang; Zhang, Weiguang; Ru, Guomei; Lang, Juan (April 2015). "Expression profile analysis of microRNAs in prostate cancer by next-generation sequencing". The Prostate. 75 (5): 500–516. doi:10.1002/pros.22936. ISSN 0270-4137. PMID 25597612. S2CID 13078880.
  18. ^ Malik, Sumit; Singh, Joginder; Goyat, Rohit; Saharan, Yajvinder; Chaudhry, Vivek; Umar, Ahmad; Ibrahim, Ahmed A.; Akbar, Sheikh; Ameen, Sadia; Baskoutas, Sotirios (2023-09-01). "Nanomaterials-based biosensor and their applications: A review". Heliyon. 9 (9): e19929. Bibcode:2023Heliy...919929M. doi:10.1016/j.heliyon.2023.e19929. ISSN 2405-8440. PMC 10559358. PMID 37809900.
  19. ^ Chambers, James P.; Arulanandam, Bernard P.; Matta, Leann L.; Weis, Alex; Valdes, James J. (January 2008). "Biosensor Recognition Elements". Current Issues in Molecular Biology. 10 (1–2): 1–12. doi:10.21775/cimb.010.001. ISSN 1467-3045. PMID 18525101.
  20. ^ Nirschl, Martin; Reuter, Florian; Vörös, Janos (September 2011). "Review of Transducer Principles for Label-Free Biomolecular Interaction Analysis". Biosensors. 1 (3): 70–92. doi:10.3390/bios1030070. ISSN 2079-6374. PMC 4264362. PMID 25586921.
  21. ^ Mehrotra, Parikha (2016-05-01). "Biosensors and their applications – A review". Journal of Oral Biology and Craniofacial Research. 6 (2): 153–159. doi:10.1016/j.jobcr.2015.12.002. ISSN 2212-4268. PMC 4862100. PMID 27195214.
  22. ^ "Chemical Sensors and Biosensors | Wiley". Wiley.com. Retrieved 2024-02-24.
  23. ^ Ouyang, Tinglan; Liu, Zhiyu; Han, Zhiyi; Ge, Qinyu (2019-03-05). "MicroRNA Detection Specificity: Recent Advances and Future Perspective". Analytical Chemistry. 91 (5): 3179–3186. doi:10.1021/acs.analchem.8b05909. ISSN 0003-2700. PMID 30702270.
  24. ^ Du, Hui; Strohsahl, Christopher M.; Camera, James; Miller, Benjamin L.; Krauss, Todd D. (2005-06-01). "Sensitivity and Specificity of Metal Surface-Immobilized "Molecular Beacon" Biosensors". Journal of the American Chemical Society. 127 (21): 7932–7940. doi:10.1021/ja042482a. ISSN 0002-7863. PMID 15913384.
  25. ^ Opperwall, Stacey R.; Divakaran, Anand; Porter, Elizabeth G.; Christians, Jeffrey A.; DenHartigh, Andrew J.; Benson, David E. (2012-09-25). "Wide Dynamic Range Sensing with Single Quantum Dot Biosensors". ACS Nano. 6 (9): 8078–8086. doi:10.1021/nn303347k. ISSN 1936-0851. PMID 22924857.
  26. ^ Lavín, Álvaro; Vicente, Jesús De; Holgado, Miguel; Laguna, María F.; Casquel, Rafael; Santamaría, Beatriz; Maigler, María Victoria; Hernández, Ana L.; Ramírez, Yolanda (July 2018). "On the Determination of Uncertainty and Limit of Detection in Label-Free Biosensors". Sensors. 18 (7): 2038. Bibcode:2018Senso..18.2038L. doi:10.3390/s18072038. ISSN 1424-8220. PMC 6068557. PMID 29949904.
  27. ^ Erdem, A. (2008). "Electrochemical Sensor Technology Based on Nanomaterials for Biomolecular Recognitions". In Vaseashta, A.; Mihailescu, I. N. (eds.). Functionalized Nanoscale Materials, Devices and Systems. NATO Science for Peace and Security Series B: Physics and Biophysics. Dordrecht: Springer Netherlands. pp. 273–278. doi:10.1007/978-1-4020-8903-9_17. ISBN 978-1-4020-8903-9.
  28. ^ Drummond, T. Gregory; Hill, Michael G.; Barton, Jacqueline K. (October 2003). "Electrochemical DNA sensors". Nature Biotechnology. 21 (10): 1192–1199. doi:10.1038/nbt873. ISSN 1546-1696. PMID 14520405. S2CID 7299424.
  29. ^ Lusi, E. A.; Passamano, M.; Guarascio, P.; Scarpa, A.; Schiavo, L. (2009-04-01). "Innovative Electrochemical Approach for an Early Detection of microRNAs". Analytical Chemistry. 81 (7): 2819–2822. doi:10.1021/ac8026788. ISSN 0003-2700. PMID 19331434.
  30. ^ Kilic, Tugba; Kaplan, Merve; Demiroglu, Sibel; Erdem, Arzum; Ozsoz, Mehmet (2016). "Label-Free Electrochemical Detection of MicroRNA-122 in Real Samples by Graphene Modified Disposable Electrodes". Journal of the Electrochemical Society. 163 (6): B227–B233. doi:10.1149/2.0481606jes. ISSN 0013-4651.
  31. ^ Yaralı, Ece; Kanat, Erkin; Erac, Yasemin; Erdem, Arzum (February 2020). "Ionic Liquid Modified Single-use Electrode Developed for Voltammetric Detection of miRNA-34a and its Application to Real Samples". Electroanalysis. 32 (2): 384–393. doi:10.1002/elan.201900353. ISSN 1040-0397. S2CID 203945157.
  32. ^ Wu, Xiaoyan; Chai, Yaqin; Yuan, Ruo; Su, Huilan; Han, Jing (2013-01-22). "A novel label-free electrochemical microRNA biosensor using Pd nanoparticles as enhancer and linker". Analyst. 138 (4): 1060–1066. Bibcode:2013Ana...138.1060W. doi:10.1039/C2AN36506E. ISSN 1364-5528. PMID 23291596.
  33. ^ Erdem, Arzum; Congur, Gulsah (2014-01-15). "Label-free voltammetric detection of MicroRNAs at multi-channel screen printed array of electrodes comparison to graphite sensors". Talanta. 118: 7–13. doi:10.1016/j.talanta.2013.09.041. ISSN 0039-9140. PMID 24274264.
  34. ^ Tao, Yiyi; Yin, Dan; Jin, Mingchao; Fang, Jie; Dai, Tao; Li, Yi; Li, Yuxia; Pu, Qinli; Xie, Guoming (2017-10-15). "Double-loop hairpin probe and doxorubicin-loaded gold nanoparticles for the ultrasensitive electrochemical sensing of microRNA". Biosensors and Bioelectronics. 96: 99–105. doi:10.1016/j.bios.2017.04.040. ISSN 0956-5663. PMID 28475957.
  35. ^ a b Tran, Hoang Vinh; Piro, Benoit (2021-03-19). "Recent trends in application of nanomaterials for the development of electrochemical microRNA biosensors". Microchimica Acta. 188 (4): 128. doi:10.1007/s00604-021-04784-3. ISSN 1436-5073. PMID 33740140. S2CID 232273687.
  36. ^ Liu, Sihan; Yang, Zhehan; Chang, Yuanyuan; Chai, Yaqin; Yuan, Ruo (2018-11-15). "An enzyme-free electrochemical biosensor combining target recycling with Fe3O4/CeO2@Au nanocatalysts for microRNA-21 detection". Biosensors and Bioelectronics. 119: 170–175. doi:10.1016/j.bios.2018.08.006. ISSN 0956-5663. PMID 30125878. S2CID 52052784.
  37. ^ Tian, Liang; Qian, Kun; Qi, Jinxu; Liu, Qinyao; Yao, Chen; Song, Wei; Wang, Yihong (2018-01-15). "Gold nanoparticles superlattices assembly for electrochemical biosensor detection of microRNA-21". Biosensors and Bioelectronics. 99: 564–570. doi:10.1016/j.bios.2017.08.035. ISSN 0956-5663. PMID 28826000.
  38. ^ He, Lei; Huang, Rongrong; Xiao, Pengfeng; Liu, Yuan; Jin, Lian; Liu, Hongna; Li, Song; Deng, Yan; Chen, Zhu; Li, Zhiyang; He, Nongyue (2021-05-01). "Current signal amplification strategies in aptamer-based electrochemical biosensor: A review". Chinese Chemical Letters. 32 (5): 1593–1602. doi:10.1016/j.cclet.2020.12.054. ISSN 1001-8417. S2CID 234215285.
  39. ^ Miao, Peng; Wang, Bidou; Meng, Fanyu; Yin, Jian; Tang, Yuguo (2015-03-18). "Ultrasensitive Detection of MicroRNA through Rolling Circle Amplification on a DNA Tetrahedron Decorated Electrode". Bioconjugate Chemistry. 26 (3): 602–607. doi:10.1021/acs.bioconjchem.5b00064. ISSN 1043-1802. PMID 25692917.
  40. ^ a b c Carrascosa, Laura G.; Huertas, César S.; Lechuga, Laura M. (2016-06-01). "Prospects of optical biosensors for emerging label-free RNA analysis". TrAC Trends in Analytical Chemistry. 80: 177–189. doi:10.1016/j.trac.2016.02.018. ISSN 0165-9936. S2CID 101649813.
  41. ^ Lai, Meimei; Slaughter, Gymama (November 2019). "Label-Free MicroRNA Optical Biosensors". Nanomaterials. 9 (11): 1573. doi:10.3390/nano9111573. ISSN 2079-4991. PMC 6915498. PMID 31698769.
  42. ^ Wang, Qing; Liu, Rongjuan; Yang, Xiaohai; Wang, Kemin; Zhu, Jinqing; He, Leiliang; Li, Qing (2016-02-01). "Surface plasmon resonance biosensor for enzyme-free amplified microRNA detection based on gold nanoparticles and DNA supersandwich". Sensors and Actuators B: Chemical. 223: 613–620. doi:10.1016/j.snb.2015.09.152. ISSN 0925-4005.
  43. ^ Husale, Sudhir; Persson, Henrik H. J.; Sahin, Ozgur (December 2009). "DNA nanomechanics allows direct digital detection of complementary DNA and microRNA targets". Nature. 462 (7276): 1075–1078. Bibcode:2009Natur.462.1075H. doi:10.1038/nature08626. ISSN 1476-4687. PMC 2966338. PMID 20010806.
  44. ^ Johnson, Blake N.; Mutharasan, Raj (2012-12-04). "Sample Preparation-Free, Real-Time Detection of microRNA in Human Serum Using Piezoelectric Cantilever Biosensors at Attomole Level". Analytical Chemistry. 84 (23): 10426–10436. doi:10.1021/ac303055c. ISSN 0003-2700. PMID 23101954.
  45. ^ a b Yang, Ke; Li, Jining; Lamy de la Chapelle, Marc; Huang, Guorong; Wang, Yunxia; Zhang, Jinbao; Xu, Degang; Yao, Jianquan; Yang, Xiang; Fu, Weiling (2021-03-01). "A terahertz metamaterial biosensor for sensitive detection of microRNAs based on gold-nanoparticles and strand displacement amplification". Biosensors and Bioelectronics. 175: 112874. doi:10.1016/j.bios.2020.112874. ISSN 0956-5663. PMID 33293192. S2CID 228079846.
  46. ^ a b c d Coutinho, Catarina; Somoza, Álvaro (2019-03-01). "MicroRNA sensors based on gold nanoparticles". Analytical and Bioanalytical Chemistry. 411 (9): 1807–1824. doi:10.1007/s00216-018-1450-7. ISSN 1618-2650. PMID 30390112. S2CID 53256743.
  47. ^ Lee, Jin-Ho; Cho, Hyeon-Yeol; Choi, Hye Kyu; Lee, Ji-Young; Choi, Jeong-Woo (July 2018). "Application of Gold Nanoparticle to Plasmonic Biosensors". International Journal of Molecular Sciences. 19 (7): 2021. doi:10.3390/ijms19072021. ISSN 1422-0067. PMC 6073481. PMID 29997363.
  48. ^ Zhang, Congcong; Miao, Pei; Sun, Mingyuan; Yan, Mei; Liu, Hong (2019-08-05). "Progress in miRNA Detection Using Graphene Material–Based Biosensors". Small. 15 (38): e1901867. doi:10.1002/smll.201901867. ISSN 1613-6810. PMID 31379135.
  49. ^ Wan, Zhengfen; Umer, Muhammad; Lobino, Mirko; Thiel, David; Nguyen, Nam-Trung; Trinchi, Adrian; Shiddiky, Muhammad J.A.; Gao, Yongsheng; Li, Qin (August 2020). "Laser induced self-N-doped porous graphene as an electrochemical biosensor for femtomolar miRNA detection". Carbon. 163: 385–394. Bibcode:2020Carbo.163..385W. doi:10.1016/j.carbon.2020.03.043. hdl:10072/392615. ISSN 0008-6223.
  50. ^ Pimalai, Dechnarong; Putnin, Thitirat; Waiwinya, Wassa; Chotsuwan, Chuleekorn; Aroonyadet, Noppadol; Japrung, Deanpen (2021-09-08). "Development of electrochemical biosensors for simultaneous multiplex detection of microRNA for breast cancer screening". Microchimica Acta. 188 (10): 329. doi:10.1007/s00604-021-04995-8. ISSN 1436-5073. PMID 34495394. S2CID 237437043.
  51. ^ Yeung, Wing Kiu; Chen, Huai-Yi; Sun, Juan-Jie; Hsieh, Tung-Han; Mousavi, Mansoureh Z.; Chen, Hsi-Hsien; Lee, Kuang-Li; Lin, Heng; Wei, Pei-Kuen; Cheng, Ji-Yen (2018-09-24). "Multiplex detection of urinary miRNA biomarkers by transmission surface plasmon resonance". Analyst. 143 (19): 4715–4722. Bibcode:2018Ana...143.4715Y. doi:10.1039/C8AN01127C. ISSN 1364-5528. PMID 30188550.
  52. ^ Kocabey, Samet; Chiarelli, Germán; Acuna, Guillermo P.; Ruegg, Curzio (2023-03-15). "Ultrasensitive and multiplexed miRNA detection system with DNA-PAINT". Biosensors and Bioelectronics. 224: 115053. doi:10.1016/j.bios.2022.115053. ISSN 0956-5663. PMID 36608362.
  53. ^ Bruch, Richard; Johnston, Midori; Kling, André; Mattmüller, Thorsten; Baaske, Julia; Partel, Stefan; Madlener, Sibylle; Weber, Wilfried; Urban, Gerald A.; Dincer, Can (2021-04-01). "CRISPR-powered electrochemical microfluidic multiplexed biosensor for target amplification-free miRNA diagnostics". Biosensors and Bioelectronics. 177: 112887. doi:10.1016/j.bios.2020.112887. hdl:20.500.11850/466547. ISSN 0956-5663. PMID 33493854.
  54. ^ Weber, Jessica A; Baxter, David H; Zhang, Shile; Huang, David Y; How Huang, Kuo; Jen Lee, Ming; Galas, David J; Wang, Kai (2010-11-01). "The MicroRNA Spectrum in 12 Body Fluids". Clinical Chemistry. 56 (11): 1733–1741. doi:10.1373/clinchem.2010.147405. ISSN 0009-9147. PMC 4846276. PMID 20847327.
  55. ^ El Aamri, Maliana; Yammouri, Ghita; Mohammadi, Hasna; Amine, Aziz; Korri-Youssoufi, Hafsa (November 2020). "Electrochemical Biosensors for Detection of MicroRNA as a Cancer Biomarker: Pros and Cons". Biosensors. 10 (11): 186. doi:10.3390/bios10110186. ISSN 2079-6374. PMC 7699780. PMID 33233700.
  56. ^ D’Agata, Roberta; Spoto, Giuseppe (2019-07-01). "Advanced methods for microRNA biosensing: a problem-solving perspective". Analytical and Bioanalytical Chemistry. 411 (19): 4425–4444. doi:10.1007/s00216-019-01621-8. ISSN 1618-2650. PMID 30710205. S2CID 73441549.
  57. ^ Chin, Curtis D.; Chin, Sau Yin; Laksanasopin, Tassaneewan; Sia, Samuel K. (2013), Issadore, David; Westervelt, Robert M. (eds.), "Low-Cost Microdevices for Point-of-Care Testing", Point-of-Care Diagnostics on a Chip, Berlin, Heidelberg: Springer Berlin Heidelberg, pp. 3–21, doi:10.1007/978-3-642-29268-2_1, ISBN 978-3-642-29267-5, S2CID 30403891, retrieved 2024-02-24
  58. ^ Cimmino, Wanda; Migliorelli, Davide; Singh, Sima; Miglione, Antonella; Generelli, Silvia; Cinti, Stefano (2023-07-01). "Design of a printed electrochemical strip towards miRNA-21 detection in urine samples: optimization of the experimental procedures for real sample application". Analytical and Bioanalytical Chemistry. 415 (18): 4511–4520. doi:10.1007/s00216-023-04659-x. ISSN 1618-2650. PMC 10328899. PMID 37000212.
  59. ^ Zhou, Yunlei; Yin, Huanshun; Li, Jie; Li, Bingchen; Li, Xue; Ai, Shiyun; Zhang, Xiansheng (2016-05-15). "Electrochemical biosensor for microRNA detection based on poly(U) polymerase mediated isothermal signal amplification". Biosensors and Bioelectronics. 79: 79–85. doi:10.1016/j.bios.2015.12.009. ISSN 0956-5663. PMID 26700579.
  60. ^ Asefpour Vakilian, Keyvan (2019-12-01). "Gold nanoparticles-based biosensor can detect drought stress in tomato by ultrasensitive and specific determination of miRNAs". Plant Physiology and Biochemistry. 145: 195–204. doi:10.1016/j.plaphy.2019.10.042. ISSN 0981-9428. PMID 31706222. S2CID 209577118.
  61. ^ Bandi, Nora; Zbinden, Samuel; Gugger, Mathias; Arnold, Marlene; Kocher, Verena; Hasan, Lara; Kappeler, Andreas; Brunner, Thomas; Vassella, Erik (1 July 2009). "miR-15a and miR-16 Are Implicated in Cell Cycle Regulation in a Rb-Dependent Manner and Are Frequently Deleted or Down-regulated in Non–Small Cell Lung Cancer". Cancer Research. 69 (13): 5553–5559. doi:10.1158/0008-5472.CAN-08-4277. PMID 19549910.
  62. ^ Zhu, Liping; Ye, Jing; Wang, Shuang; Yan, Mengxia; Zhu, Qiuju; Huang, Jianshe; Yang, Xiurong (2019). "Dual amplification ratiometric biosensor based on a DNA tetrahedron nanostructure and hybridization chain reaction for the ultrasensitive detection of microRNA-133a". Chemical Communications. 55 (77): 11551–11554. doi:10.1039/C9CC05592D. PMID 31490470. S2CID 201844715.
  63. ^ Moro, Giulia; Fratte, Chiara Dalle; Normanno, Nicola; Polo, Federico; Cinti, Stefano (2023-12-18). "Point-of-Care Testing for the Detection of MicroRNAs: Towards Liquid Biopsy on a Chip". Angewandte Chemie International Edition. 62 (51): e202309135. doi:10.1002/anie.202309135. hdl:10278/5036502. ISSN 1433-7851. PMID 37672490.
  64. ^ Farshchi, Fatemeh; Saadati, Arezoo; Fathi, Nazanin; Hasanzadeh, Mohammad; Samiei, Mohammad (2021-03-18). "Flexible paper-based label-free electrochemical biosensor for the monitoring of miRNA-21 using core–shell Ag@Au/GQD nano-ink: a new platform for the accurate and rapid analysis by low cost lab-on-paper technology". Analytical Methods. 13 (10): 1286–1294. doi:10.1039/D1AY00142F. ISSN 1759-9679. PMID 33624680. S2CID 232036763.
  65. ^ Gao, Zhiqiang; Deng, Huimin; Shen, Wei; Ren, Yuqian (2013-02-05). "A Label-Free Biosensor for Electrochemical Detection of Femtomolar MicroRNAs". Analytical Chemistry. 85 (3): 1624–1630. doi:10.1021/ac302883c. ISSN 0003-2700. PMID 23323518.
  66. ^ Godoy, Paula M.; Barczak, Andrea J.; DeHoff, Peter; Srinivasan, Srimeenakshi; Etheridge, Alton; Galas, David; Das, Saumya; Erle, David J.; Laurent, Louise C. (2019-12-17). "Comparison of Reproducibility, Accuracy, Sensitivity, and Specificity of miRNA Quantification Platforms". Cell Reports. 29 (12): 4212–4222.e5. doi:10.1016/j.celrep.2019.11.078. ISSN 2211-1247. PMC 7499898. PMID 31851944.
  67. ^ a b Johnson, Blake N.; Mutharasan, Raj (2014-03-03). "Biosensor-based microRNA detection: techniques, design, performance, and challenges". Analyst. 139 (7): 1576–1588. Bibcode:2014Ana...139.1576J. doi:10.1039/C3AN01677C. ISSN 1364-5528. PMID 24501736.
  68. ^ Bail, Sophie; Swerdel, Mavis; Liu, Hudan; Jiao, Xinfu; Goff, Loyal A.; Hart, Ronald P.; Kiledjian, Megerditch (2010-05-01). "Differential regulation of microRNA stability". RNA. 16 (5): 1032–1039. doi:10.1261/rna.1851510. ISSN 1355-8382. PMC 2856875. PMID 20348442.
  69. ^ Meyer, Swanhild U.; Pfaffl, Michael W.; Ulbrich, Susanne E. (2010-12-01). "Normalization strategies for microRNA profiling experiments: a 'normal' way to a hidden layer of complexity?". Biotechnology Letters. 32 (12): 1777–1788. doi:10.1007/s10529-010-0380-z. ISSN 1573-6776. PMID 20703800. S2CID 5791279.
  70. ^ Chugh, Pauline; Dittmer, Dirk P. (September 2012). "Potential pitfalls in microRNA profiling". WIREs RNA. 3 (5): 601–616. doi:10.1002/wrna.1120. ISSN 1757-7004. PMC 3597218. PMID 22566380.
  71. ^ D’haene, Barbara; Mestdagh, Pieter; Hellemans, Jan; Vandesompele, Jo (2012), Fan, Jian-Bing (ed.), "MiRNA Expression Profiling: From Reference Genes to Global Mean Normalization", Next-Generation MicroRNA Expression Profiling Technology: Methods and Protocols, Methods in Molecular Biology, vol. 822, Totowa, NJ: Humana Press, pp. 261–272, doi:10.1007/978-1-61779-427-8_18, ISBN 978-1-61779-427-8, PMID 22144205, retrieved 2024-03-01
  72. ^ Kangkamano, Tawatchai; Numnuam, Apon; Limbut, Warakorn; Kanatharana, Proespichaya; Vilaivan, Tirayut; Thavarungkul, Panote (April 2018). "Pyrrolidinyl PNA polypyrrole/silver nanofoam electrode as a novel label-free electrochemical miRNA-21 biosensor". Biosensors and Bioelectronics. 102: 217–225. doi:10.1016/j.bios.2017.11.024. ISSN 0956-5663. PMID 29149687.
  73. ^ Wang, Mo; Yin, Huanshun; Fu, Zhengliang; Guo, Yunlong; Wang, Xinxu; Zhou, Yunlei; Ai, Shiyun (2014-10-01). "A label-free electrochemical biosensor for microRNA detection based on apoferritin-encapsulated Cu nanoparticles". Journal of Solid State Electrochemistry. 18 (10): 2829–2835. doi:10.1007/s10008-014-2531-y. ISSN 1433-0768. S2CID 95204739.
  74. ^ Wang, Zhong-Yu; Sun, Ming-Hui; Zhang, Qun; Li, Pei-Feng; Wang, Kun; Li, Xin-Min (July 2023). "Advances in Point-of-Care Testing of microRNAs Based on Portable Instruments and Visual Detection". Biosensors. 13 (7): 747. doi:10.3390/bios13070747. ISSN 2079-6374. PMC 10377738. PMID 37504145.
  75. ^ Iacomino, Giuseppe (February 2023). "miRNAs: The Road from Bench to Bedside". Genes. 14 (2): 314. doi:10.3390/genes14020314. ISSN 2073-4425. PMC 9957002. PMID 36833241.