Mahsa Bagi,
Sina Jamalzadegan,
Anastasiia Steksova and
Qingshan Wei*
Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, NC 27605, USA. E-mail: qwei3@ncsu.edu
First published on 21st July 2025
The detection of RNA biomarkers is crucial for diagnosing many urgent diseases such as infections and cancer. Conventional RNA detection techniques such as RT-PCR, LAMP, and microarrays are effective, but often face limitations in terms of speed, sensitivity, and equipment demands. In recent years, CRISPR/Cas systems have emerged as versatile platforms for RNA detection, which offer high specificity, programmability, and adaptability across a wide range of diagnostic applications. This review first categorizes different CRISPR-based RNA detection systems according to the CRISPR effectors employed, including Cas13, Cas12, Cas14, Cas9, and newly characterized enzymes such as Cas7–11 and Cas10, detailing their mechanisms of target recognition, cleavage activity, and signal generation. The CRISPR detection platforms are coupled with or without pre-amplification steps to meet the different sensitivity needs. Preamplification-based systems integrate CRISPR with methods like RT-PCR and isothermal amplification to enhance sensitivity. In parallel, preamplification-free strategies, such as split-crRNA or split-activator systems, are gaining attention for their balanced assay performance and simplicity, which are especially attractive for point-of-care (POC) settings. Then, the diagnostic applications of these technologies are explored across two major domains: infectious disease detection and cancer biomarker identification via miRNAs, demonstrating the clinical potential of CRISPR-based RNA detection platforms. In addition, we explore ongoing challenges such as improving sensitivity in amplification-free formats, and developing field-deployable, cost-effective systems. The review concludes by outlining emerging trends and future directions in CRISPR-based RNA diagnostics, emphasizing their transformative potential in clinical settings.
To date, a wide range of RNA detection techniques have been developed to target various RNA types, including microRNAs (miRNAs) and viral RNAs, across both in vitro and in vivo environments. Among the most commonly used methods are reverse transcription polymerase chain reaction (RT-PCR),6 quantitative reverse transcription PCR (RT-qPCR),7 northern blotting,8 RNA sequencing (RNA-Seq),9 in situ hybridization (ISH),10 and microarray11 technologies. Each of these approaches offers distinct advantages, yet they also present notable limitations. Specifically, many of these methods require expensive instrumentation, non-portable setups, labor-intensive protocols, and the expertise of trained laboratory personnel.12 To overcome these challenges, there is a growing need to develop RNA detection strategies that are compatible with portable, cost-effective, fast-response, and user-friendly devices. These innovations are particularly valuable for critical biomedical and biological applications, including the early diagnosis of viral infections, transcriptomic profiling in cancer and neurodegenerative diseases, and the monitoring of gene expression dynamics in cellular therapies and drug delivery systems. Consequently, advancing rapid and cost-effective RNA detection technologies remains a crucial objective to enhance the accessibility, timeliness, and impact of RNA-based diagnostics and therapeutic monitoring.
Clustered regularly interspaced short palindromic repeats (CRISPR) and CRISPR-associated (Cas) enzymes, first identified in bacteria in 1987,13 represent a natural defense mechanism against invading genetic elements such as plasmids and phages.14 The transformative potential of the CRISPR–Cas system for genome editing was first demonstrated in 2012, when researchers successfully repurposed it for precise and programmable DNA cleavage.14 Later, around 2016–2017, specific Cas enzymes, particularly Cas13,15 were shown to possess high efficacy for nucleic acid detection, enabling the direct targeting of RNA molecules and the development of novel RNA-based diagnostic technologies. Compared to conventional RNA detection methods, CRISPR/Cas-based platforms offer rapid, highly sensitive, and highly specific detection that can be integrated into portable biosensor and point-of-care (POC) assay formats.16–18 Several types of Cas enzymes have been employed for RNA detection depending on whether the strategy relies on direct or indirect detection, pre-amplification-based or amplification-free approaches, and the choice of readout techniques, including fluorescent and non-fluorescent modalities.19
Several previous reviews have discussed the applicability of CRISPR-based diagnostic assays for detecting RNA nucleic acids.20–25 However, many existing works consider RNA detection within the broader context of nucleic acid diagnostics, which, while informative, may not fully elaborate on the unique challenges and strategies pertinent to the diverse array of RNA species, ranging from messenger RNAs (mRNAs), miRNAs, long non-coding RNAs (lncRNAs), to viral RNAs.21,26 From an experimental perspective, it is also critical to comprehensively discuss both preamplification-free and preamplification-based CRISPR mechanisms for RNA detection. This distinction is crucial for better understanding their respective advantages, limitations, and compatibility with different RNA targets, yet previous reviews often treat these strategies in isolation.25 In addition, it is important to highlight that CRISPR platforms for RNA detection are relevant not only for pathogenic RNAs, such as viral genomes, but also for non-pathogenic host RNAs that play key roles in complex diseases like cancer and neurodegeneration. Previous reviews, tend to concentrate on pathogen-oriented detection and may overlook other applications.21,26 Even the more targeted reviews27 that emphasize Cas13 and modified Cas9 systems for RNA imaging and detection, may not comprehensively address the full range of emerging RNA-targeting Cas effectors or offer a systematic comparison of their molecular mechanisms and application scopes across all RNA classes. As such, there is a need for a timely review that not only introduces CRISPR-based RNA detection as a transformative approach but also systematically categorizes strategies by Cas enzyme and RNA target, delves into their core biological mechanisms and comparative performance, and finally broadly surveys their applications and challenges, thereby bridging fundamental science with real-world utility.
In this review, we first provide an overview of RNA detection and its significance in modern diagnostics, outlining traditional RNA detection methods along with their inherent limitations (Fig. 1). We then introduce CRISPR-based RNA detection as a promising and transformative approach that addresses many of these challenges. Following this, we categorize the various CRISPR-based RNA detection strategies, including platforms built upon various Cas enzymes. For each platform, we comprehensively explore the fundamental biological mechanisms involved, distinguishing between preamplification-based and amplification-free detection strategies, and we summarize their unique advantages and disadvantages. Next, we highlight recent applications of CRISPR-based RNA detection across a wide range of fields, from infectious disease diagnostics to cancer biomarker detection, and from neurodegenerative disease monitoring to agricultural and environmental surveillance. Finally, we prospect the current challenges that still limit the widespread adoption of these platforms in real-world applications, considering critical dimensions such as sensitivity, specificity, scalability, and regulatory pathways. We also propose our future perspectives to advance CRISPR/Cas-based RNA detection technologies and bring them closer to their full clinical and environmental potential. To summarize, Sections 1–5 detail various CRISPR–Cas based detection platforms and their characteristics, while Sections 6–8 explore their diverse diagnostic and monitoring applications across different domains.
![]() | ||
Fig. 1 RNA detection with various CRISPR–Cas systems and their applications in detecting viral RNA and cancer-related RNA biomarkers. |
![]() | ||
Fig. 2 Detection mechanisms of different Cas effectors. (a) Schematic illustration of Cas 9-based cis-cleavage for dsDNA target. (b) Schematic of Cas12-induced cis-cleavage and trans-cleavage (non-target cleavage) mechanism for dsDNA target detection. (c) Schematic of Cas13-assisted cis-cleavage and trans-cleavage (non-target cleavage) for ssRNA target detection.17 |
In contrast, the Cas13 family exclusively targets RNA. Upon recognition of its specific single-stranded RNA (ssRNA) target, Cas13's two HEPN (higher eukaryotes and prokaryotes nucleotide-binding) domains become catalytically active, triggering promiscuous degradation of surrounding non-target ssRNA (Fig. 2c). This trans-cleavage property is utilized in diagnostic platforms by introducing engineered reporter RNA molecules, the cleavage of which produces a detectable signal, such as fluorescence or a colorimetric change.28 Cas13-based diagnostics first emerged with the SHERLOCK system developed by Zhang's group in 2017 and have since evolved significantly, with numerous amplification strategies developed to improve sensitivity and adaptability for different applications.30
Type of enzyme | Preamplification | Target | LOD | Read out | Rxn time | Ref. |
---|---|---|---|---|---|---|
LbuCas13a | RT-PCR | SARS-CoV-2 | 10 nM | Lateral flow and flour-plate reader | ∼90 min for fluorescence 35 min for lateral flow | 41 |
Cas13a | RT-PCR | SARS-CoV-2 RNA, HBV RNA, LMP1 gene of EBV, EBNA gene of EBV | 0.1 aM | Fluorescence (naked eye under blue light) | 40 min | 31 |
LbuCas13a | RT-PCR | SARS-CoV-2 | 2 aM | Fluorescence, colorimetric, or electrochemical methods | 90 min | 42 |
Cas13a | RT-PCR, RT-RAA | Hepatitis delta virus (HDV) | 10 copies per μL | Fluorescence and lateral flow strip | 60 min for fluorescence; 33 min for lateral flow | 43 |
LwaCas13a | LAMP | miRNA-7 (ciRS-7) | 1 fM | Fluorescence | 30 min | 44 |
LwCas13a | RT-LAMP | SARS-CoV-2 | 42 copies per reaction | Fluorescence, lateral flow | 70 min | 45 |
Cas13 | RT-RPA | Mitochondrial DNA or RNA, reverse transcription of environmental RNA (eRNA) | 22.6 ng μL−1 | Lateral flow/flour-plate reader | 60 min | 46 |
LwCas13a | RPA | Y. ruckeri DNA and transcripts sRNAs | 2 fM | Fluorescence or lateral flow | 70 min | 47 |
LwaCas13a | RPA | V. alginolyticus | 10 copies per μL | Fluorescence or lateral flow | 50 min | 48 |
LwaCas13a | RT-RPA | Fusarium graminearum and Fusarium verticillioides | Few copies of DNA/RNA targets | Fluorescence, lateral flow | 26 min | 49 |
LwaCas13a | CHA | SARS-CoV-2 | 84 aM | Fluorescence | 45 min | 50 |
Cas13a | CHA, HCR | SARS-CoV-2 | 285 fM | Fluorescence | 70 min | 51 |
Cas13a | HCR | Biomarker such as brain natriuretic peptide (BNP) | 3.2 fg mL−1 | Electrochemiluminescence (ECL) | 80 min | 52 |
LbuCas13a | bHCR | miRNA-106a | 8.55 aM | Polyacrylamide gel electrophoresis (PAGE), fluorescence and SERS | 60 min | 36 |
LwaCas13a | CHDC | miRNA-17, miRNA-155, TTF-1 mRNA, miRNA-19b, miRNA-210 and EGFR mRNA | 50 aM | Electrochemical biosensor | 36 min | 33 |
Cas13a | Endonuclease-mediated cyclic fluorescent amplification | SARS-CoV-2 | 74.13 aM | Fluorescence | 60 min | 53 |
LwaCas13a | Endonuclease cycle amplification | sja-miR-2c5p | 83.2 fM | Magnetic upconversion nanoparticles (MUCNPs) as a biosensor | 60 min | 54 |
Cas13a | DNAzyme-mediated signal amplification | miRNA-21 | 27 fM | Colorimetry | 30 min | 55 |
Cas13a | Entropy-driven cyclic amplification strategy | SARS-CoV-2 | 7.39 aM | Electrochemiluminescence (ECL) | 80 min | 56 |
Cas13a | T7 | circRNA, miRNA, piRNA, and 16S rRNA | 1.65 aM | QD-based FRET nano sensor | 60 min | 34 |
Cas13a | Rolling circle transcription (RCT) | piR-hsa-14 | 3.32 fM | Flour-plate reader | 70 min | 57 |
LwaCas13a | Transcription mediated amplification (TMA) | P. jirovecii-mitochondrial large subunit ribosomal RNA | 2 copies per μL | Fluorescence | 60 min | 58 |
Cas13a | Mn/NiCo2O4 nanozyme as a signal amplifier | miRNA-143 | Tens of aM | Colorimetric and fluorometric | 60 min | 59 |
Cas13a | Exo-III activity on the Ag+-aptamer | miRNA-155 | 5.12 fM | Colorimetry | 75 min | 37 |
Cas13a | No | miRNA-21 | 10 aM | Fluorescence | 60 min | 40 |
Cas13a | No | E. coli, bacterial RNA | 0.65 fM | Electrochemiluminescence, bipolar electrode-ECL lateral flow chip | 20 min | 60 |
LwaCas13a | No | SARS-CoV-2 | 100 aM | Capillary sensor (RNA-cross-linking DNA hydrogel film) | 30 min | 61 |
Cas13a | No | SARS-CoV-2 | 10 fM | Fluorescence | 60 min | 62 |
LwaCas13a | No | Synthetic RNA, pseudovirus | 100 fM | On-chip total internal reflection fluorescence microscopy | ∼45 min | 63 |
Cas13a | No | miRNA-21 | 9 fM | Photoelectrochemical (PEC) biosensing platform | 60 min | 64 |
Cas13a | No | Enterovirus B, Lassa, dengue, influenza A | 1010 copies per μL | Fluorescence | — | 65 |
Cas13a | No | Human circular RNA | 0.089 fM | Electrochemical biosensor | ∼10 min | 66 |
LbCas13a | No | SARS-CoV-2 | 2 aM | Fluorescence | 15 min | 39 |
LwCas13a | No | Y. pestis, F. tularensis, Chlamydia psittaci, B. mallei, B. pseudomallei, and Brucella melitensis | 1 pM | FAM-RNA-MB electrochemical signal probe | 25 min | 67 |
LwCas13a | No | SARS-CoV-2 | 26.2 and 53.5 copies per μL | Electrochemical biosensor | 200 min | 68 |
LwaCas13a | No | Ebola RNA | 291 aM | Fluorescence | 40 min | 69 |
LbuCas13a | No | SARS-CoV-2 | 10 aM | Fluorescence, digital droplets | ∼10–20 min | 70 |
LbuCas13a | No | miRNA-21 | 75 aM | Fluorescent | 30 min | 71 |
LwCas13a | No | SARS-CoV-2 | ∼200 copies per sample | Lateral-flow assay (LFA) | 2 min lateral flow plus SHERLOCK | 72 |
Cas13a | No | lncRNA H19 | — | Fluorescence | — | 73 |
6× His-twinstrep-SUMO-huLwCas13a | Light-triggered exponential amplification, RCA | miRNA-21 | 1 fM | ssDNA reporter with photocleavable linker | 80 min | 38 |
Cas13a–Cas12 | Cas13a–12a amplification | miRNA-155 | 0.35 fM | Fluorescence | 75 min | 35 |
Hybrid Cas12a and Cas13a with SpyTag–SpyCatcher | RT-RPA | SARS-CoV-2 | 10 copies per reaction tube | Fluorescence | 74 |
Building on the same principle of integrating isothermal amplification with Cas13 detection, Xiao Wang et al. in 202432 reported a system for detecting synthetic monkeypox virus (MPXV) using recombinase-aided amplification (RAA) followed by T7 transcription and Cas13a–crRNA targeting (Fig. 3d). The output relies on a fluorogenic RNA aptamer (mango III), which emits fluorescence upon binding to the TO3 fluorophore. When Cas13a remains inactive, the aptamer stays intact, and fluorescence is observed. However, if the target is present, Cas13a becomes activated and cleaves the RNA aptamer, disrupting its structure and suppressing fluorescence. This fluorescent change, detectable under UV light, offers a simple and equipment-free readout. Impressively, the system achieved detection of as few as 4 copies of the target RNA within 40 minutes.
![]() | ||
Fig. 3 (a) Schematic illustrating detection of RNA target using PddCas13a.40 (b) Schematic of Cas13-triggered PC-ssDNA release from the MB@PC-NAC, padlock generation and under UV light and RCA amplification.38 (c) Cas-CHDC-powered electrochemical RNA-sensing technology (COMET) chip: off-chip signal amplification (i), on-chip RNA measurements (ii), and square wave voltammetry (SWV) readout (iii).33 (d) Schematic illustrating detection with RAA-Cas13a-Apt system.32 (e) Schematic illustrating RNA detection using CRISPR/Cas13a-triggered exonuclease-iii-assisted colorimetric assay.37 (f) Schematic illustration of Cas13a–12a amplification fluorescence biosensing platform.35 Pdd: polydisperse droplet digital; PC: photocleavable linker; MB: methylene blue; PC-NAC: photocleavable linker-nucleic acid probe; RCA: rolling circle amplification; CHDC: catalytic hairpin DNA circuit; RAA: recombinase-aided amplification; Apt: aptamer. |
Taking a different approach to amplify, Yan Sheng and colleagues (2021)33 developed a reusable electrochemical biosensor that combines Cas13a with a catalytic hairpin DNA circuit (CHDC) (Fig. 3c). After Cas13a is activated by a target RNA, it initiates the CHDC, an enzyme-free signal amplification mechanism made up of two DNA hairpins. The first hairpin undergoes structural changes upon hybridizing with a nucleic acid trigger, while the second hairpin displaces the initial trigger in a strand-exchange reaction. Multiple cycles of hybridization process amplify the detection signal, which is then observed by an integrated screen-printed electrode chip. The detection, performed using square-wave voltammetry, achieves a LOD of 50 aM with a total process time of 36 minutes, highlighting the potential for low-cost, reusable, and highly sensitive RNA diagnostics.
Researchers have also identified additional preamplification strategies that offer unique advantages to enhance signal amplification further and broaden diagnostic applicability. One such method further refines sensitivity through a transcription-driven amplification mechanism. In a 2024 study, Wen-jing Liu and colleagues introduced a nanosensor system that couples Cas13a with quantum dot (QD)-based fluorescence readout (Table 1). Upon target RNA recognition, Cas13a cleaves a substrate probe that subsequently serves as a promoter for T7 RNA polymerase. The resulting RNA transcripts are amplified and hybridized with both biotinylated capture probes and Cy5-labeled reporter probes anchored to a single QD.34 This proximity facilitates fluorescence resonance energy transfer (FRET) between the QD and Cy5, generating a detectable signal with remarkable sensitivity—achieving a LOD of 1.65 aM in 60 minutes. The strategy elegantly combines enzymatic signal amplification with nanotechnology to offer precise detection in complex clinical samples, such as breast cancer tissue.
In a complementary direction, Dan Zhao et al. (2023)35 introduced a dual-enzyme platform that integrates Cas13a with Cas12a in a sequential activation cascade (Fig. 3f). In this system, the target RNA activates Cas13a, which then cleaves a bulge structure within a blocker strand immobilized on magnetic beads. This cleavage event releases a primer strand that subsequently activates Cas12a. Once triggered, Cas12a initiates its well-known trans-cleavage activity, cutting a fluorescent ssDNA reporter and producing a measurable signal. This elegant two-stage mechanism enhances the detection sensitivity by combing two CRISPR activities, enabling signal amplification while maintaining sequence specificity. It exemplifies how multiple Cas systems can be strategically combined to maximize diagnostic performance.
Another creative use of amplification comes from the work of Jingjing Zhang et al. (2023),36 who merged Cas13a activity with a branched hybridization chain reaction (bHCR) (Table 1). In this system, Cas13a's trans-cleavage is used to initiate the bHCR by cleaving an RNA substrate that releases an initiator. This initiator then hybridizes to a DNA hairpin, triggering a chain reaction where subsequent hairpins unfold and extend the signal cascade. The resulting branched DNA structures are then detected using a surface-enhanced Raman scattering (SERS) sensor composed of silver nanorods. This method achieved a detection limit of 8.55 aM within 60 minutes, offering not only high sensitivity but also compatibility with SERS-based biosensing for label-free, multiplex-capable detection.
Building on these innovative signal amplification designs, Yunxiao Li et al. (2024)37 introduced an exonuclease-III (Exo-III)-assisted Cas13a detection platform, integrating a clever aptamer-based reporting mechanism (Fig. 3e). Here, target activation triggers Cas13a to cleave a structured probe, releasing an intermediate strand that unlocks a silver ion (Ag+)-binding aptamer. This structural change forms a protruding 3′-terminus, which is specifically recognized and cleaved by Exo-III, releasing Ag+ ions. The liberated silver ions are then chelated again by excess aptamer, generating a visible color change. This multi-step cascade achieved a detection limit of 5.12 fM in around 110 minutes, demonstrating the effectiveness of integrating enzymatic cleavage and metal-ion signaling into CRISPR diagnostics.
Taking signal amplification a step further, Tao Hu et al. (2023)38 presented a light-triggered system combining Cas13a and rolling circle amplification (RCA) (Fig. 3b). Upon target binding, Cas13a cleaves RNA substrates embedded in a photocleavable complex, enabling the light-controlled release of short DNA oligos with 5′ phosphate ends. These oligos serve as triggers for RCA, generating long ssDNA products that emit strong fluorescence signals. This system achieves an excellent detection limit of 1 fM for RNA targets in 80 minutes.
One such example is described by Dou Wang (2023),39 who developed a digital detection platform using magnetic beads to enrich and capture RNA targets (Table 1). Upon recognition, the Cas13a–crRNA complex initiates cleavage of a reporter molecule. The cleavage products are then compartmentalized into a femtoliter-scale microwell array, allowing individual fluorescent events to be resolved and quantified digitally. This digital assay approach minimizes background noise and supports highly sensitive detection.
A different but equally elegant strategy was proposed by Ke Wang et al. (2025)40 through the development of polydisperse droplet digital (Pdd) Cas13a, a system utilizing polydisperse water-in-oil droplets as independent reaction chambers (Fig. 3a). In this setup, RNA molecules—including noncoding RNAs relevant to cancer—are randomly partitioned into thousands of droplets. Following incubation, only those droplets that contain target RNA activate Cas13a and emit a fluorescent signal. After 60 minutes, the droplets can be visualized under a fluorescence microscope, achieving a remarkable LOD as low as 10 aM. By combining microfluidics with Cas13-based collateral cleavage, this droplet system provides a scalable and precise method for digital RNA quantification without amplification.
Type of enzyme | Preamplification | Target RNA | LOD | Read out | Rxn time | Ref. |
---|---|---|---|---|---|---|
Cas12a | PCR | Infected cell culture | aM | Fluorescence | ∼1 h | 77 |
Cas12a | RT-PCR | SARS-CoV-2 | 0.5 copies per mL | Fluorescence | 110 minutes | 78 |
Cas12a | RT-LAMP | SARS-CoV-2 | 5 copies | Fluorescent detection by naked eye under blue light | 81 | |
LbCas12a | LAMP | RNAseP POP7 mRNA | 16 copies per μL | — | 110 minutes | 109 |
Cas12a | RPA/RT-RPA | Ebola virus | 11 aM | μPAD readout | 1–4 h | 110 |
Cas12a | RPA | HRSV | 100 copies per mL | Fluorescence | 60 minutes | 111 |
Cas12a | RPA | HIV | 200 copies per test | Glucose meter | ∼70 minutes | 112 |
AsCas12a, LbCas12a | RPA/RT-RPA | Rice black-streaked dwarf virus (RBSDV) | 1 aM | Lateral flow and fluorescence | 30–60 minutes | 49 |
Cas12a | RT-RPA | Potato virus X (PVX), and potato virus Y (PVY) | fM levels | Fluorescence | ∼20 minutes | 113 |
Cas12a | RT-RPA | Conserved fragments within the VP2 gene of the norovirus GII.2 subtype | 10 copies per μL | Lateral flow | 25–35 minutes | 114 |
LbCas12a | RPA/RT-RPA | Apple necrotic mosaic virus (ApNMV), apple stem pitting virus (ASPV), apple stem grooving virus (ASGV), apple chlorotic leaf spot virus (ACLSV) and apple scar skin viroid (ASSVd) | 25 viral copies per reaction | Oligonucleotide-conjugated gold nanoparticles | ∼50 minutes | 115 |
LbCas12a | RPA/RT-RPA | microRNAs | ∼aM | G-quadruplex (G4) containing a hairpin structure as the reporter | 1 hour preamplification + 40 minutes CRISPR rxn | 116 |
Cas12a | RPA | miRNA-21 | 3.43 aM | Fluorescence | 40 min preamplification + 30 min cleavage reaction | 117 |
Cas12a | RPA/RT-RPA | SARS-CoV-2 | 0.38 copies per μL | Fluorescence detection using a smartphone-based device | 15 minutes | 79 |
SuCas12a2 | RPA | Transcripts of the EhPrx and p1 genes | 102 copies per reaction | Fluorescence | 40 min preamplification + 45 min cleavage reaction | 118 |
MeCas12a (manganese-enhanced Cas12a) | RT-RAA | MERS-CoV | 5 copies | Fluorescence | ∼45 min | 82 |
Cas12a | RT-RAA | SARS-CoV-2 | 1 copy per μL | Fluorescence | 15 min preamplification + 10 min cleavage reaction | 119 |
Cas12a | RCA | SARS-CoV-2 | 604 fM | Portable glucose meter (PGM) | 3 h | 120 |
Cas12a | RCA | miRNAs | 0.52 aM | Electrochemical | ∼4 h | 121 |
Cas12a | RCA | miRNA-21 | 16 aM | Chemiluminescence (CL) | ∼4 h | 122 |
Cas12a | Hyperbranched rolling circle amplification (HRCA) | miRNA-21 | 10.02 fM | Fluorescence | 223 min preamplification + 50 min cleavage reaction | 123 |
LbCas12a | HRCA | miRNA-143 | 1 fM | Gold nanoparticle (AuNP)-based visual assay | 45–50 minutes | 124 |
Cas12a | Branched rolling circle amplification (BRCA) | Primers incorporating ncRNA sequences of circulating CRC-associated RNAs (piRNA) | 12 pM | Fluorescence | One hour | 125 |
FnCas12a | Multiply-primed rolling circle amplification (MRCA) | SARS-CoV-2 | 1.625 copies per reaction | Lateral flow and fluorescence | 20 to 60 minutes | 83 |
Cas12a | 3D DNA walker cascade amplification | miRNA-214 | 20.42 fM | Fluorescence | 126 | |
LbCas12a | 3D DNA walker cascade amplification | miRNA-141 | 0.331 fM | Electrochemiluminescence (ECL) | 1 day and a few hours | 127 |
Cas12a | Reverse transcription-free exponential amplification reaction (RTF-EXPAR) | SARS-CoV-2 | 3.77 aM | Fluorescence | 40 minutes | 128 |
Cas12a | EXPAR | miRNA-21 | 3.2 fM | Electrochemical | 80 minutes preamplification + 40 minutes cleavage reaction | 85 |
Cas12a | EXPAR | miRNA-155 | 85 aM | Dual-channel fluorescence and colorimetric signal output | 80 minutes preamplification + 20 minutes cleavage reaction | 84 |
LbCas12a | CHA | miRNAs | ∼100 fM | ECL | 2 hours | 129 |
LbCas12a | CHA | miRNA-122 | 2.04 fM | Photoelectrochemical (PEC) biosensor | 3 hours preamplification + 1 hour cleavage reaction | 130 |
LbCas12a | CHA | miRNA-21 | 0.48 fM | ECL | — | 131 |
LbCas12a | CHA | miR-128-3p | 2.5 to 8.98 fM | Fluorescence | 75 minutes preamplification + 60 minutes cleavage reaction | 87 |
Cas12a | HCR | miRNA-141 | 3.3 fM | ECL | 3 hours preamplification + 1 hour cleavage reaction | 132 |
LbCas12a | HCR | miRNA-21 | 75.4 aM | Fluorescence | 2 hours preamplification + 8 minutes cleavage reaction | 86 |
Cas12a | T7 RNA polymerase | miRNAs | 1 pM | Fluorescence | 160 minutes preamplification + 150 minutes cleavage reaction | 89 |
FnCas12a | RT-HDA | Deformed wing virus (DWV) RNA | 500 fM | Fluorescence | 90 minutes preamplification + 45 minutes cleavage reaction | 133 |
AsCas12a | HDA | SARS-CoV-2, SCGB2A2 (mammaglobin A) RNA | 0.6 copies per μL | Lateral flow and fluorescence | 90 minutes | 92 |
Cas12a | Target-induced transcription process via split-T7 promoter | miRNA-21 isothermal amplification | 43.9 fM | Fluorescence | ∼1 h | 134 |
Cas12a | RCA + CHE | Rabies viral RNA | 2.8 pM | ECL | ∼11 h | 135 |
Cas12a | RCA + CHA | microRNA-320d | 0.342 fM | ECL | 3 hours RCA + 3 hours CHA + 2 hours Cas12a rxn | 136 |
Cas12a | Dual-signal amplification (Exo III amplification + RCA) | miRNA-21 | 6.01 fM | Fluorescence | 5.5 hours | 137 |
LbCas12a | Strand displacement reaction (SDR) | miRNA-21 | 2.42 fM | Fluorescence | ∼4 h | 88 |
Cas12a | Toehold-mediated strand displacement reaction (TSDR) | SARS-CoV-2 | 40 aM | Electrochemical | 90 minutes preamplification + 40 minutes cleavage reaction | 138 |
Cas12a | Entropy-driven catalysis (EDC) cycle amplification | miRNA-21 | 1.5 fM | PEC | ∼4 h | 90 |
Cas12a | Strand displacement amplification (SDA) | miRNA-let-7a | 6.28 pM | Distance-based biosensor | 1 hours preamplification + 30 minutes cleavage reaction | 139 |
Cas12a | SDA | miRNA-21 | 0.5 fM | Single-particle inductively coupled plasma-mass spectrometry | 2 hours and 15 minutes | 91 |
Cas12a | Self-primer-initiated (SPI)-CRISPR–Cas12a-assisted amplification | miRNAs | 254 aM | Functionalized gold nanoparticle (AuNP)-based color generation | 60 minutes preamplification + 30 minutes cleavage reaction | 140 |
Cas12a | T7 exonuclease-assisted target recycling | bacterial 16S rRNA (rRNA) | 3.6 pM | Split G-quadruplex (G4) catalytic signal output | — | 141 |
Cas12a | — | DENV | 100 fM | Electrochemical | ∼30 min | 142 |
Cas12a | — | HCV RNA, miRNA-155 | 767 pM | Fluorescence | 1 h | 98 |
Cas12a | — | SARS-CoV-2 | 50 copies per μL | Fluorescence | ∼71 min | 93 |
Cas12a | — | SARS-CoV-2 | 50 copies per μL | PGM | — | 94 |
Cas12a | — | SARS-CoV-2 | 10 fM | Fluorescence | 30 minutes | 101 |
Cas12a | — | miRNA-21 | 10 fM | Fluorescence | 5 minutes | 96 |
Cas12a | — | SARS-CoV-2 | 2.2 pM | Fluorescence | 45 minutes | 95 |
Cas12a | — | miRNA-21 | 301 fM | Lateral flow assay | 60 minutes | 143 |
Cas12a | — | miRNA-19a | 856 aM | Fluorescence | 20 minutes | 99 |
Cas12a | — | miRNA-10b | 5.53 fM | Reverse fluorescence-enhanced lateral flow test strip (rFLTS) | 40 minutes | 144 |
Cas12a | — | miRNA-21 | 10 pM | Lateral flow and fluorescence | 30 minutes | 102 |
Cas12b | RT-LAMP | SARS-CoV-2, human adenovirus, herpes simplex virus | 1 copy per μl | Fluorescence | — | 105 |
eBrCas12b | RT-LAMP | HCV | — | Fluorescence | ∼1 h | 103 |
Cas12b | LAMP | mRNAs | 10−8 nM | Fluorescence | — | 145 |
AapCas12b | LAMP | SARS-CoV-2 | 10 copies per reaction | Lateral flow and fluorescence | — | 146 |
AapCas12b | RT-RPA | SARS-CoV-2 | 8 copies per μL | Fluorescence | 30 to 60 minutes | 106 |
BhCas12b | No | CJ8421_04975 mRNA from Campylobacter jejuni | 1 μM | Fluorescence | 45 min | 147 |
Cas12j | EXPAR | miRNA-92a, miRNA-122, and miRNA-155 | 3.2 fM | Fluorescence | 30 minutes | 107 |
Cas12g | PCR | — | — | Gel electrophoresis | 30 minutes | 76 |
Cas12g | — | — | — | Running the reaction samples on 20% PAGE TBE-urea denaturing gels stained with GelRed nucleic acid gel stain, and the results were visualized using alpha innotech (fluouchem TM) | — | 108 |
Cas12g1 | — | — | 100–125 pM | Gel electrophoresis | — | 148 |
Building on this, Ning et al. (2022) applied a Cas12a-based platform to detect RNA variants, again using reverse transcription by either RT-PCR or RT-RPA to generate DNA amplicons as targets for Cas12a detection. Notably, their system could discriminate between different viral variants by targeting mutations within the PAM or seed region and showed potential for smartphone-based POC diagnostics.78
![]() | ||
Fig. 4 (a) CRISPR–FDS assay optimization using spiked saliva samples. Steps 1–3 tested different lysis buffer dilutions to enhance RNA release and fluorescence detection.79 (b) Schematic showing CRISPR/LbCas12a trans-cleavage activity on gold nanoparticles (AuNPs) for direct RNA detection.96 (c) Schematic illustrating the core principle and overall workflow of the RT-LAMP and Cas12a-based RNA detection assay.80 (d) Schematic of iSCAN-V2: a one-tube assay combining reverse transcription, amplification, and Cas12b-mediated collateral cleavage for fluorescent RNA detection from swab samples.106 (e) The process of viral RNA detection using reverse transcription recombinase-aided amplification (RT-RAA) combined with MeCas12a-based detection.82 (f) Workflow of the EXP-J reaction for lung cancer diagnosis.107 (g) Demonstration of the enzyme-catalyzed rolling circle amplification-assisted CRISPR/FnCas12a assay through stepwise reactions, with a conceptual diagram illustrating the padlock probe and PAM-free design.83 RT-LAMP: reverse transcription loop-mediated isothermal amplification; iSCAN: in vitro specific CRISPR-based assay for nucleic acids; RT-RAA: reverse transcription recombinase-aided amplification; EXP-J: exponential junction (EXP-junction) amplification. |
Building on the momentum of RPA, researchers also explored loop-mediated isothermal amplification (RT-LAMP) as a complementary strategy. Unlike multi-step processes, LAMP allowed amplification and Cas12a-mediated detection to occur in a single vessel, reducing contamination risks (Fig. 4c).80 Systems like opvCRISPR exemplified this streamlined design, achieving near single-molecule detection sensitivity and offering simple visual readouts, making sophisticated molecular diagnostics accessible beyond laboratory environments.81 As efforts continued to refine speed and field applicability, recombinase-aided amplification (RT-RAA) emerged as another promising technique (Fig. 4e).82 Its low-temperature operation and rapid amplification kinetics made it ideal for time-critical diagnostics. When integrated into a centrifugal microfluidic platform, RT-RAA combined with Cas12a achieved detection of viral RNA at single-copy sensitivity within 30 minutes, with diagnostic accuracy matching that of RT-PCR. This demonstrated that high performance could be maintained even in settings where complex equipment is impractical.
While many platforms focused on cDNA intermediates, others sought to simplify the workflow further by bypassing reverse transcription altogether. Rolling circle amplification (RCA) provided such an opportunity. RCA-based systems like OPERATOR83 used RNA-templated DNA ligation to initiate multiply-primed amplification, producing products that directly activated Cas12a without prior cDNA synthesis (Fig. 4g). Ingenious variations like CRISPR–PGM even converted nucleic acid detection into glucose readouts measurable by handheld meters, illustrating how CRISPR and RCA could be combined for affordable, decentralized diagnostics.83
In pursuit of even faster and more sensitive assays, researchers turned to an exponential amplification reaction (EXPAR). EXPAR has emerged as a powerful partner to Cas12a in RNA diagnostics, offering rapid, highly sensitive, and versatile detection. By leveraging EXPAR's strong amplification capability alongside Cas12a's collateral cleavage activity, several studies have demonstrated detection of miRNAs and viral RNAs at femtomolar to attomolar concentrations.84 One key advantage of this approach is the high specificity achieved through Cas12a's sequence-specific recognition, which significantly reduces false positives often associated with isothermal amplification alone. Additionally, multiplex detection has been demonstrated using EXPAR-generated dual ssDNA triggers to activate Cas12a in an “AND” logic circuit, enabling the simultaneous detection of multiple miRNA targets with a single crRNA complex. EXPAR–Cas12a platforms have also introduced dual-mode signal outputs (e.g., fluorescence and colorimetric readouts), increasing result reliability and flexibility. Smartphone integration for POC testing (POCT) has further enhanced the portability and accessibility of these systems.85
Alongside these enzyme-driven amplifications, catalytic hairpin assembly (CHA) offered a simpler, enzyme-free alternative. Integrating CHA with Cas12a enabled one-step, highly sensitive RNA detection, sidestepping multiple handling steps and reducing contamination risk. Creative adaptations pushed detection sensitivity into the femtomolar range, while combination strategies involving photoelectrochemical and electrochemiluminescence methods enhanced signal clarity and minimized background noise. Notably, platforms like MCM-CRISPR/Cas12a, combining CHA with hybridization chain reaction (HCR),86 demonstrated detection of low-abundance non-coding RNAs with remarkable precision.87
Beyond the major amplification strategies discussed, other methods such as hybridization chain reaction (HCR),86 strand displacement reaction (SDR),88 binding-induced primer-triggered cascade (BPTC),89 enzyme-driven cascade amplification (EDC),90 and strand displacement amplification (SDA)91 have also been employed in Cas12a-based RNA detection. Though less commonly used, these approaches provide valuable enhancements in sensitivity, specificity, or workflow flexibility. Among these, one particularly notable system combined SDA with Cas12a and utilized single-particle inductively coupled plasma mass spectrometry (spICP-MS) for detection. This platform integrated SDA with Cas12a trans-cleavage and employed gold nanoparticle labels to achieve a quantification limit of 0.5 aM. With a 45-minute reaction time, this system offers an optimal balance of sensitivity and detection kinetics for clinical applications.91 Another innovative approach is the flap endonuclease, Taq ligase and CRISPR–Cas for diagnostics (X) (FELICX) platform, which uniquely uses flap endonuclease (FEN) for target recognition, enabling PAM-independent detection. This is followed by adaptor ligation and Cas12a activation, and when combined with RTx-HDA, FELICX achieved RNA detection at 1 aM sensitivity within 90 minutes. Its independence from guide RNA optimization and compatibility with both DNA and RNA targets make FELICX an exceptionally versatile and practical platform for molecular diagnostics.92
In some of these systems, although the detection phase is preamplification-free, the RNA target is first reverse transcribed into cDNA before Cas12a engagement which is due to the intrinsic nature of Cas12a that will be activated by DNA targets. In one such example, viral RNA was reverse transcribed using a gene-specific primer and a commercial reverse transcription kit. The resulting cDNA acted as a trigger for Cas12a, allowing the detection step to proceed without preamplification. This system maintained high sensitivity, reaching detection limits in the femtomolar range, despite skipping the preamplification phase.93 Another approach under the same category departs from direct activation of Cas12a by instead employing the RNA as a competitive inhibitor. In this method, Cas12a is initially activated by a target DNA sequence (tcDNA), which triggers its nonspecific ssDNA cleavage activity. When the target RNA is present, it competes with the DNA activator for binding to the crRNA–Cas12a complex, effectively suppressing Cas12a's trans-cleavage function. This change in enzymatic behaviour is then monitored electrochemically. Unlike the previous example, this system does not require reverse transcription, relying instead on RNA's ability to modulate Cas12a activity through competitive inhibition. This approach maintains an amplification-free and label-free detection process and reveals a different dimension of Cas12a's programmability, where signal generation can occur through inhibition rather than activation.94
Progressing further within the same category, recent studies have shown that Cas12a can, under certain optimized conditions, be directly activated by RNA targets, challenging the previous assumption that Cas12a only responds to DNA triggers. In one such example, functionalized magnetic beads were employed to capture the target RNA alongside signal probes containing activation sequences for Cas12a. Once bound, magnetic separation isolated the RNA-probe complex, and the addition of Cas12a and crRNA allowed the probe itself to activate Cas12a, leading to collateral cleavage of fluorescent reporters. Notably, the use of manganese ions in place of magnesium was found to enhance Cas12a's cleavage efficiency, enabling RNA detection with high precision in clinical settings, all without preamplification or reverse transcription. This technique delivered results within 45 minutes with a detection limit of 2.2 pM.95
In a parallel strategy, another group introduced a nanomaterial-enhanced system, where gold nanoparticle-based nanobeacons (Au-nanobeacons) were used as reporters. In this case, Cas12a's collateral activity was directly triggered by RNA, specifically miRNA, without any DNA intermediates or pre-treatment. The Au-nanobeacons offered both faster kinetics and enhanced sensitivity compared to traditional fluorescent reporters. With a detection limit reaching as low as 10 fM in just five minutes of reaction, and validated performance in serum, this method pushes the boundaries of what Cas12a can achieve in its unamplified, native state (Fig. 4b).96 In addition, a recent crRNA spacer regulation-based SCas12aV2 assay demonstrated robust and programmable Cas12a activation by leveraging stem–loop-structured crRNAs, enabling sensitive and modular detection of structured RNA targets, including SARS-CoV-2, miRNAs, and circRNAs, with high specificity and one-pot isothermal readout capability.97
A separate yet related method employed a competitive displacement mechanism for signal enhancement. This technique utilized both a full-length crRNA and a split crRNA that compete for Cas12a binding. Initially, the full crRNA activates Cas12a upon binding to the RNA, triggering trans-cleavage. Subsequently, the split crRNA replaces the full one, initiating a secondary round of activation—thus creating a cascade signal amplification. While still based on complete crRNA recognition of the RNA target, this approach enhances sensitivity by exploiting the dynamic binding interactions between full and split crRNAs.99 More recently, a flexible “splice-at-will” crRNA engineering strategy was introduced, enabling precise and sensitive Cas12a-mediated detection of ultrashort RNAs (as short as 6–8 nt) by reconstituting functional crRNAs through modular hybridization at nearly any site within the spacer region using a DNA splint and target RNA.100
Together, these studies demonstrate how split activators, whether composed of synthetic DNA, endogenous RNA, or a combination can be paired with complete crRNA to create preamplification-free Cas12a systems. These systems elegantly bypass the need for reverse transcription by selectively engaging different regions of the crRNA with engineered or native sequences, broadening the toolkit for rapid, sensitive, and versatile RNA diagnostics.
A notable example of this concept is the PARC–Cas12a platform (proximity-activated RNA-guided Cas12a), where the crRNA is split within its 5′ handle domain. In this system, activation of Cas12a is contingent on the reconstitution of the guide RNA, which occurs only when the split fragments are spatially brought together. This is achieved through the hybridization of the fragments to a shared trigger molecule, such as an RNA sequence, aptamer, or small molecule. This proximity-dependent assembly transforms Cas12a into a programmable biosensor, enabling logic-gated detection schemes and multiplexing. The system has also been successfully integrated into platforms like arrayPARC–Cas12a and ICP-MS-based assays, enhancing its diagnostic versatility.101
Another important contribution in this space is the development of SCas12a, a system that fully separates the scaffold RNA from the spacer RNA. Here, the target RNA, such as a mature miRNA, directly serves as the spacer component. In the presence of a complementary ssDNA activator and the scaffold RNA, the miRNA completes the crRNA assembly and triggers Cas12a's trans-cleavage activity. This method achieves direct RNA detection without preamplification or reverse transcription and does so with high sensitivity in the femtomolar range. Moreover, SCas12a exhibits strong discrimination power, capable of distinguishing closely related sequences, as well as single-nucleotide variants.102
A major advancement in this area came with the engineering of a thermostable Cas12b variant, BrCas12b, derived from Brevibacillus sp. and further optimized through hydrophobic core mutations to enhance its trans-cleavage activity at elevated temperatures. This engineered enzyme, eBrCas12b, retains robust performance up to 67 °C, allowing seamless integration with RT-LAMP reactions.103 Utilizing this capability, the SPLENDID platform was developed, a single-pot LAMP-mediated detection assay clinically validated for RNA detection in human serum and saliva. The system demonstrated high specificity and accuracy, completing detection within one hour and outperforming standard RT-LAMP by significantly reducing false positives.103 Similarly, the STOPCovid assay used AapCas12b from Alicyclobacillus acidiphilus in a one-pot RT-LAMP setup, optimized through guide RNA engineering and the addition of taurine to accelerate reaction kinetics.104
Moving beyond conventional detection, the WS-RADICA platform introduced a digital format for Cas12b-based diagnostics. Combining warm-start RT-LAMP with Cas12b in a droplet-based digital system, this method achieved a detection limit as low as 1 copy per μL, offering both qualitative and quantitative output. Notably, WS-RADICA performed well under challenging conditions, showing tolerance to common inhibitors like EDTA and SDS, and providing quantification accuracy comparable to RT-qPCR and RT-dPCR, while offering faster results and broader application to both RNA and DNA viruses.105
In addition to RT-LAMP, RT-RPA has also been explored as a lower-temperature alternative. While less widely reported, some systems have demonstrated effective coupling of Cas12b with RT-RPA for rapid RNA detection, particularly when rapid deployment or reduced thermal requirements are needed. However, further engineering is typically required to adapt Cas12b's activity to RPA's temperature window. The iSCAN-V2 platform is another example of using RT-RPA, offering a one-pot, preamplification-based assay where Cas12b outperformed Cas12a in both signal strength and speed (Fig. 4d). The assay achieved a LOD of 8 copies per μL, and in clinical validation, it showed 93.75% sensitivity and 100% specificity, reliably detecting RNA in samples with cycle threshold (Ct) ≤ 30 in under an hour.106
Overall, Cas12b-based systems have rapidly evolved into highly sensitive, amplification-compatible diagnostic tools. Their thermal stability, compatibility with one-pot formats, and adaptability to digital platforms make them particularly promising for clinical, field-based, and POC RNA diagnostics, especially in settings requiring speed, specificity, and minimal equipment.
Structural and biochemical analyses have confirmed Cas12g's capacity for RNA detection. It employs a dual-guide RNA system, a crRNA and a tracrRNA, or a fused single-guide RNA (sgRNA), to recognize RNA sequences with no requirement for a PAM or protospacer flanking sequence (PFS). This feature removes a major limitation in CRISPR targeting and simplifies assay design for RNA diagnostics. Importantly, studies have shown that Cas12g demonstrates comparable RNA detection sensitivity to top-performing Cas13 effectors, while maintaining thermostability, broadening its potential for both laboratory and POC applications.108
While no fully realized Cas12g-based diagnostic assay has yet been reported, recent cryo-EM and crystal structure analyses have provided critical insight into its molecular mechanism. These studies revealed Cas12g's bilobed architecture and highlighted key structural components, including the REC and NUC lobes, zinc finger motifs, and the lid motif within the RuvC domain, offering a blueprint for engineering future RNA detection platforms.76,108
Type of enzyme | Preamplification | Target | LOD | Read out | Rxn time | Ref. |
---|---|---|---|---|---|---|
Cas10 subunit from the type III-A CRISPR complex | RT-LAMP | SARS-CoV-2 | 108 copies per reaction | Fluorometric + colorimetric detection | ∼1 h | 165 |
Cas10 subunit of the type III-B Cmr complex | RT-LAMP | SARS-CoV-2 | 800 aM | Fluorescence | 180 minutes | 166 |
Cas10 subunit within the type III-A CRISPR complex | — | SARS-CoV-2 | 106 copies per μL | Fluorescence, gel electrophoresis, thin-layer chromatography (TLC) | 10 minutes | 175 |
Cas10 subunit within the type III-B CRISPR complex | — | SARS-CoV-2 | 8 fM | Fluorescence | 30–50 minutes | 167 |
Cas10 subunit within the type III-A CRISPR complex | Isothermal amplification | SARS-CoV-2 | aM | Fluorescence | — | 176 |
Cas10 | No | T. congolense 7SL-sRNA | 10–100 fM | Lateral flow and fluorescence | ∼2 h | 168 |
Cas10 subunit of the Staphylococcus epidermidis Cas10–Csm complex | No | Target RNA | — | Denaturing PAGE and autoradiography | 30–60 minutes | 177 |
Cas10 subunit within various Csm complexes | No | Target RNA | — | Denaturing gel electrophoresis and autoradiography | Up to 120 minutes | 170 |
Cas10 subunit of the Lactobacillus delbrueckii subsp. | No | miRNA-155 | 500 pM–2 nM | Fluorescence | 30–60 minutes | 169 |
Cas10 subunit of the Sulfolobus tokodaii Csm (StCsm) complex | No | SARS-CoV-2 | — | Agarose gel electrophoresis with EB staining | 60 minutes | 178 |
Cas7–11 | Rt-RPA | SARS-CoV-2 | 2 fM | Fluorescence | ∼2 h | 171 |
Type III-E Cas7–11 | — | Various target RNAs | — | Fluorescence | — | 172 |
Cas14a | EDC | miRNA-10b | 2.1 pM | Fluorescence spectrophotometry | 2 hours | 159 |
Cas14a1 | Asymmetric PCR | Bacterial RNAs (Streptococcus pyogenes and Eberthella typhi) | 105 CFU per mL | Fluorescence | 2 hours | 162 |
Cas14a | RCA | miRNA156a from banana | 1.839 pM | Fluorescence | 2 hours | 160 |
Cas14a | SDA | miRNA-21 | 680 fM | Fluorescence | 1 hours | 164 |
Cas14 | RT-LAMP | RNA2 gene of the red-spotted grouper nervous necrosis virus (RGNNV) | 63.4 aM | Fluorescence | 2 hours | 161 |
Cas14a | Toehold-containing three-way junction (TWJ) | RNA viruses | — | Magnetic separation enhanced colorimetry | — | 179 |
Cas14 (AsCas12f1) | No | CJ8421_04975 mRNA from Campylobacter jejuni | — | Fluorescence | — | 147 |
Cas14a1 | Transcription by T7 RNA polymerase | 16S rRNA of bacteria (specifically E. typhi) | 0.6 aM | Fluorescence | 1 hours | 163 |
Pyrococcus furiosus (Pfu) Cas3 | No | ssRNA target | 0.1–1 nM | Fluorescence | 15 minutes | 180 |
E. coli Cas3 | RT-LAMP | SARS-CoV-2 | <102 copies | Lateral flow and fluorescence | 32–42 minutes | 173 |
fastCas9n (derived from Streptococcus pyogenes Cas9 (SpCas9)) | — | Salmonella typhimurium 16S rRNA, Escherichia coli O157:H7 16S rRNA, synthetic SARS-CoV-2 genes (Orf1ab-a, Orf1ab-b, S gene, E gene, N gene), and HIV virus RNA | ∼10 copies per rxn (20 μL volume) | Lateral flow and fluorescence | 50 minutes preamplification + 5–10 minutes cleavage reaction | 149 |
dCas9 | RT-RPA | Bunyavirus RNA, the causative agent of severe fever with thrombocytopenia syndrome (SFTS) | 0.63 aM | Single microring resonator (SMR) biosensor | 30 minutes | 150 |
dCas9 | LAMP | HIV-1 | 0.96 copies per mL | Bright field microscopy | 70 minutes | 152 |
dCas9 | No | Epstein-Barr virus encoded RNA (EBER) | — | Fluorescence in situ hybridization (FISH) | 20 minutes or less | 157 |
dCas9 | RCA | miRNAs (miRNA-195) | fM level | Colorimetric | ∼4 hours | 153 |
Cas9 | No | SARS-CoV-2 | 3 × 108 copies of RNA | Gel electrophoresis, fluorescence | — | 156 |
Cas9 | No | Target RNA | Picomolar level | Denaturing polyacrylamide gel electrophoresis (PAGE) | — | 158 |
Cas9 | RCA | miRNA-21, miRNA-221 | 90 fM | Fluorescence | 2 hours preamplification + 30–60 minutes cleavage reaction | 154 |
Cas9 | PCR | RFP transgene | — | Targeted sequencing | — | 181 |
Cas9 | Exponential amplification reaction (EXPAR) | L. monocytogenes hemolysin (hly) mRNA | 0.82 aM | Fluorescence | — | 151 |
Cas9 | NASBA | Zika virus | — | Colorimetric | — | 182 |
SpyCas9 | RAA | Respiratory syncytial virus A | 98 copies per μL | Fluorescence | 2 hours 25 minutes | 183 |
dCas9 | RT-RPA | SARS-CoV-2 | 2.5 copies per μL | Lateral flow assay (LFA) | 55 minutes | 155 |
Recognizing the need for even faster and more sensitive detection, other strategies combined Cas9 cleavage with exponential amplification reactions (CAS-EXPAR). After site-specific cleavage of DNA or cDNA by Cas9/sgRNA, short fragments are generated that trigger an EXPAR cascade, exponentially amplifying the signal for highly sensitive, real-time RNA detection.151 The evolution toward digital detection led to the development of dCRISTOR, which uses dCas9-engineered magnetic micromotors. Following LAMP152 amplification of RNA, the micromotors bind specific amplicons and change their motion when exposed to a magnetic field, allowing deep learning algorithms to digitally classify positive or negative detection events. Further innovations explored rolling circle amplification (RCA) paired with Cas9 systems. In the RACE method, a padlock probe is ligated to the target miRNA, generating a circular template that undergoes RCA to produce long ssDNA, which is then recognized and cleaved by Cas9.153 Cleavage of a TaqMan probe releases fluorescence, directly indicating the presence of the target. Similarly, the RCH system harnesses RCA to bring together split horseradish peroxidase (HRP) fragments mediated by dCas9 binding, resulting in a visible color change upon substrate reaction (Fig. 5c).153,154
![]() | ||
Fig. 5 (a) Scheme of the EDC–Cas14a system for miRNA detection, where the target generates multiple activators that trigger Cas14a/sgRNA collateral cleavage for signal amplification.159 (b) Workflow of the CONAN RNA detection assay, including RNA extraction, RT-LAMP at 62 °C for 20–30 minutes, the CONAN reaction at 37 °C for 10 minutes, and lateral flow detection at room temperature for 2 minutes.173 (c) Schematic of the RCH workflow for miRNA detection using RCA to amplify target signals into large DNA structures. CRISPR–dCas9 with split-HRP fusion enables secondary amplification by binding to RCA products.153 (d) Outline of the type III CRISPR–Cas (Cmr)/NucC assay for RNA detection, where target RNA binding activates Cas10 to produce cA3 molecules. cA3 activates NucC, leading to degradation of a dsDNA reporter labeled with a fluorophore–quencher pair.168 EDC: entropy-driven circuit; CONAN: CRISPR–Cas3-operated nucleic acid detection; RCH: rolling circle hybridization; RCA: rolling circle amplification; dCas9: catalytically dead CRISPR-associated protein 9; HRP: horseradish peroxidase. |
Finally, aiming for the simplest and most portable format, the Vigilant platform was introduced. This system fuses dCas9 with the VirD2 relaxase to enable a lateral flow assay, where amplified biotinylated targets are captured by VirD2–dCas9 complexes tagged with FAM-labeled oligonucleotides. A standard lateral flow strip detects these complexes visually through streptavidin and anti-FAM antibodies, allowing rapid and accessible RNA diagnostics.155
Beyond preamplification-assisted methods, research has revealed that Cas9 can directly enable amplification-free RNA detection, offering simpler and faster diagnostic possibilities. A major breakthrough was LEOPARD, where reprogrammed tracrRNA allowed Cas9 to bind RNA targets and cleave fluorescent DNA reporters, enabling multiplexed RNA detection without amplification.156 Expanding into imaging, Chen et al. introduced RCasFISH, using dCas9 combined with sgRNAs bearing MS2 aptamers. Upon binding to RNA inside fixed cells, fluorescent MS2 proteins visualize RNA targets without amplification, preserving spatial information with high specificity.157
Even more striking, Strutt et al. discovered that certain Cas9 homologs, such as Staphylococcus aureus Cas9 (SauCas9) and Campylobacter jejuni Cas9 (CjeCas9), can inherently recognize and cleave RNA without the need for a PAM or PAMmer, instead solely guided by sgRNA. This native RNA-targeting ability expands Cas9's potential for amplification-free RNA diagnostics and direct RNA manipulation in cells.158
Several preamplification strategies have been employed to harness Cas14's potential in RNA detection. Among these, a dual amplification system developed by Shu et al. that integrates an entropy-driven circuit (EDC) with Cas14a showed significant performance enhancement. This one-pot setup continuously produces ssDNA activators that stimulate Cas14a's trans-cleavage, improving sensitivity by 100-fold over Cas14a alone (Fig. 5a).159 In another approach, a rolling circle amplification (RCA)–Cas14 system enabled the detection of plant miRNAs without requiring reverse transcription. The ligation-triggered amplification and Cas14a activation provided single-nucleotide resolution and a detection limit as low as 1 pM, demonstrating applicability in plant molecular diagnostics.160 In viral detection, Cas14a was paired with a RT-LAMP assay, using primers engineered with PAM sequences to enable Cas14a activation. This system achieved an impressive LOD of aM level, demonstrating exceptional sensitivity and compatibility with simplified magnetic bead-based RNA extraction.161
Additionally, Cas14a1 has been shown to function in asymmetric PCR-based assays for DNA diagnostics, with reported LODs down to the attomolar range. While primarily applied to DNA targets such as the SMN1 exon 7 deletion, these results support Cas14a1's potential for high-sensitivity nucleic acid detection. Most notably, Cas14a1 has also been shown to support amplification-free RNA detection.162 A platform known as ATCas-RNA demonstrated that RNA can directly trigger Cas14a1's trans-cleavage activity without undergoing degradation. This system exhibited excellent specificity, including the ability to distinguish single-base mismatches, and achieved a remarkable LOD of 1 aM. By removing the need for reverse transcription or amplification, this method represents a major step forward in streamlining RNA diagnostics.163
In summary, Cas14a and Cas14a1 have proven to be flexible, powerful tools for RNA detection, operating effectively across diverse preamplification strategies, including EDC,159 RCA,160 SDA,164 and RT-LAMP,161 and excelling even in amplification-free147 formats. Their compact size, high sensitivity, and programmability make them strong contenders for portable, rapid, and precise RNA diagnostics.
A notable example is Cas10, the signature effector of type III CRISPR–Cas systems, which has been successfully adapted for RNA detection using RT-LAMP-based preamplification.165 In one study, the type III-A TtCsm complex from Thermus thermophilus was repurposed for RNA detection. Upon RNA target binding, the Cas10 subunit exhibited polymerase activity, producing cyclic oligoadenylates (cOAs), pyrophosphates, and protons. These by-products were detected using a combination of colorimetric dyes (e.g., phenol red, malachite green), fluorometric reporters (calcein), and TtCsm6-activated fluorescent cleavage. Sensitivity was significantly enhanced by combining RT-LAMP with T7 transcription, reducing the detection limit from ∼108 copies per reaction to 106 copies per mL in clinical samples, with results available in under an hour.165
A second study introduced SCOPE, a diagnostic platform based on the type III-B Cmr complex, which similarly used RNA-activated Cas10 to generate cOAs. These activated a downstream CARF-domain RNase (TTHB144), which cleaved quenched RNA reporters to produce a fluorescent signal. In a two-step RT-LAMP–CRISPR workflow, SCOPE achieved a detection limit of 40 aM (∼25 copies per μL) and demonstrated robust sensitivity even in a one-pot format (LOD: 800 aM). Reaction times varied between 35 minutes (two-step) and up to 180 minutes (one-pot), depending on the configuration.166
Beyond preamplification-based detection, recent studies have also explored amplification-free RNA diagnostics using Cas10. One approach utilized the VmeCmr complex from Vibrio metoecus, where RNA binding triggers Cas10 to generate cOAs, activating the nuclease NucC, which cleaves a DNA reporter (Fig. 5d).167,168 Other amplification-free approaches have leveraged the direct RNA cleavage activity of Cas10-containing complexes. The Staphylococcus epidermidis Cas10–Csm complex demonstrated crRNA-guided ssRNA cleavage without amplification, while the LdCsm system from Lactobacillus delbrueckii achieved miRNA detection via Cas10-mediated collateral DNase activity, with a detection limit of 500 pM to 1 nM in buffer and ∼2 nM in serum.169 Additionally, the StCsm complex from Sulfolobus tokodaii was used in RNA editing applications, further supporting Cas10's intrinsic RNA-targeting capabilities.170
While sensitivity remains slightly lower than platforms like DETECTR or SHERLOCK, the stability of protein probes and programmability of the protease system offer promising avenues for future clinical applications and assay expansion. In a separate approach, researchers developed a programmable RNA sensor named CASP, utilizing Cas7–11 (DiCas7–11) from Desulfonema ishimotonii as the RNA-binding component. Upon target RNA binding, Cas7–11 activates an associated protease (Csx29), triggering downstream gene expression via release of transcriptional effectors. A mutant version of Cas7–11 with deactivated ribonuclease activity showed 20-fold improved sensitivity. While highly sensitive to mid-to-high expression RNAs, the system faces challenges with low-abundance targets, suggesting further enhancement through signal amplification strategies.172
To better understand the characteristics and advantages of different CRISPR–Cas platforms for RNA diagnostics, we systematically compared the detection performance of major Cas enzyme families, such as Cas13, Cas12, Cas9, Cas14, Cas10, Cas7–11 (craspase), and Cas3, based on their molecular targets, collateral activities, preamplification requirements, sensitivity, specificity, assay time, thermal stability, and POC compatibility (Table 4). Each system exhibits distinct biochemical properties that influence its suitability for diagnostic applications. For instance, Cas13 systems are favored for direct RNA detection with high specificity and no need for reverse transcription,68,72 while Cas12 enzymes, especially Cas12g, offer high thermal stability, better accessibility, and compatibility with visual or smartphone-based detection.108 Emerging systems like Cas10 and Cas7–11 expand detection mechanisms to include signal relay and protease-based readouts, respectively.168,171 Table 4 summarizes the key features of these systems, highlighting the trade-offs between sensitivity, speed, and ease of deployment in resource-limited settings.
Feature | Cas13 | Cas12 | Cas9 | Cas14 | Cas10 | Cas7–11 (craspase) | Cas3 |
---|---|---|---|---|---|---|---|
Target | RNA | Primarily DNA (via RT); Cas12g directly targets RNA | Primarily DNA; engineered for RNA | Primarily ssDNA; Cas14a1 adapted for RNA | RNA | RNA | Primarily DNA; adapted to RNA via RT |
Collateral activity | Trans-cleaves RNA reporters | trans-Cleaves ssDNA reporters | Recently discovered: trans-cleaves both RNA and ssDNA reporters | trans-Cleaves ssDNA | Indirect trans-cleavage by triggering cyclic oligoadenylates (cOAs) synthesis and activation of downstream RNases | trans-Cleaves RNA reporters | trans-Cleaves ssDNA reporters |
Preamplification requirement | Both preamplification-based and preamplification-free detection | Typically required; Cas12g is an exception | Often requires preamplification, but preamplification-free possible | Both preamplification and preamplification-free | Both preamplification and preamplification-free | Both preamplification and preamplification-free | Mostly preamplification |
Sensitivity (w/o preamplification) | As low as 10 aM | As low as 10 fM | Picomolar range | As low as 1 aM | As low as 8 fM | 25 pM | 0.1–1 nM |
Sensitivity (w/preamplification) | ∼0.1 aM | ∼0.5 aM or single-copy detection | As low as 0.63 aM | As low as 0.6 aM | As low as 800 aM | ∼2 fM | Single-copy sensitivity |
Turnover rate | 1–740 s−1 | 0.01–20 s−1 | — | <0.02 s−1 | 0.01–0.5 s−1 for Csm6 | — | 0.01–10 s−1 |
Assay speed | 15–40 min | 5–30 min | 5–70 min | 1–2 hours | 10–180 min | ∼2 hours | 15–42 min |
Specificity (e.g., SNP detection) | High | Very high | High | Very high | High | — | — |
Thermal stability | Variable; generally, requires cold chain | High for engineered variants (e.g., eBrCas12b at 67 °C) | — | — | Thermostable (TtCsm, TthB144) | — | Thermostable (TsiCas3) |
POC compatibility | Excellent: portable, adaptable to fluorescent, lateral flow, electrochemical | Excellent: smartphone-based, paper-based, visual readouts | Smartphone, lateral flow, digital detection (e.g., dCRISTOR) | Portable, magnetic bead-based, visual detection | Colorimetric/fluorescent, rugged for low-resource use | Fluorescent protein cleavage, protease-based sensing | Lateral flow, robust for field use |
Limitations | Cold chain, reagent cost | Needs reverse transcription (except Cas12g), format complexity | Needs reverse transcription step, complex multiplexing, primarily DNA-targeting | Originally DNA-based, limited RNA usage without engineering | Lower sensitivity vs. Cas12/13, limited SNP selectivity | Lower sensitivity | Weak ssRNA cleavage, lower sensitivity |
In the realm of Cas12 enzymes, researchers leveraged both Cas12a and Cas12b to build diverse diagnostic platforms for SARS-CoV-2 (Fig. 6b).184 One amplification-free method used Cas12a with magnetic bead-assisted separation and manganese-enhanced fluorescence signaling, enabling detection within 45 minutes with a LOD of 2.2 pM.95 Another Cas12a-based system integrated toehold-mediated strand displacement reactions (TSDR), converting RNA into DNA activators that triggered Cas12a's transcleavage activity, achieving ultrasensitive electrochemical detection down to 40 aM.138 In a parallel effort, Ning et al. demonstrated an RT-RPA-Cas12a-based assay capable of ultrasensitive and quantitative detection of SARS-CoV-2 directly from saliva samples, bypassing the need for conventional nasal swabs. Remarkably, this work was also among the first to integrate CRISPR diagnostics with smartphone readout, enabling portable and accessible testing. The combination of high sensitivity, non-invasive sampling, and smartphone compatibility positioned this platform as a promising tool for point-of-care COVID-19 diagnostics.79 Separately, the iSCAN-V2 platform integrated RT-RPA with Cas12b, showing enhanced performance over Cas12a and achieving reliable detection at 40 copies per μL using a low-cost fluorescence readout device.106
![]() | ||
Fig. 6 (a) The control flow of HIV-1 detection using the dCRISTOR (dCas9) assay with CNN-MOT, combining inactivation, extraction-free RT-LAMP, micromotor binding to LAMP amplicons, and motion-based detection.152 (b) Working mechanism of a CRISPR–Cas12a-based biosensor that allows quantitative RNA detection through a portable personal glucose meter readout.184 (c) Workflow of HCR combined with Cas12a for miR-21 detection.86 (d) Overview of the digital dual CRISPR–Cas-powered single EV evaluation (ddSEE) system design.190 (e) miRNA detection by miRoll-Cas for prostate cancer diagnosis.191 dCRISPR: digital clustered regularly interspaced short palindromic repeats; CNN-MOT: convolutional neural network classification-based multiobject tracking algorithm; HCR: hybridization chain reaction; miRoll-Cas: microRNA rolling circle amplification-Cas system. |
Cas13a added another dimension with the SATCAS platform, which integrated reverse transcription, amplification, and detection in a one-pot setup. It achieved single-copy detection (∼aM range) within 40 minutes, proving effective in clinical validation.31 Meanwhile, type III CRISPR enzymes introduced alternative mechanisms, Cas10 in the SCOPE platform produced cyclic oligoadenylates (cOAs) upon RNA recognition, activating CARF-domain nucleases for signal generation.166 A type III-B complex (VmeCmr) utilized cA3 signaling to trigger NucC nuclease activity, allowing direct SARS-CoV-2 RNA detection at 2 fM without preamplification.171
Additionally, Cas3 has been adapted into the CONAN assay, which utilizes EcoCas3's collateral ssDNA cleavage activity for SARS-CoV-2 RNA detection, offering a rapid and low-cost diagnostic alternative.173 In addition to Cas3-based systems, Cas9 has also been adapted for diagnostic use through the Vigilant platform. The vigilant platform repurposes dCas9 fused with VirD2 to detect SARS-CoV-2 RNA after RT-RPA amplification. By binding the target sequence and linking to a FAM-labelled probe, it enables visual detection on a lateral flow strip. Vigilant offers a low detection limit (2.5 copies per μL), high sensitivity (96.4%), and perfect specificity (100%), making it a simple and cost-effective tool for point-of-care diagnostics.155
Cas13a has also been explored for HCV diagnostics, especially for amplification-free strategies. Screening 13 distinct crRNAs revealed that pooling them significantly enhanced Cas13a's trans-cleavage efficiency, improving assay sensitivity for RNA detection. The same approach was extended to hepatitis D virus (HDV), where CRISPR–Cas13a was combined with RT-PCR or RT-RAA for detection. These methods offered strong performance and were visualized using lateral flow strips, demonstrating the system's potential in both laboratory and POC settings.185
Another approach combines RT-RAA amplification with Cas13a-mediated detection for HIV-1 RNA. This system uses a degenerate base-binding Cas13a–crRNA complex. Results are visualized using a lateral flow strip, offering a portable, rapid, and user-friendly format. The assay demonstrated a limit of detection of 1 copy per μL, with 91.81% sensitivity and 100% specificity in clinical evaluations, detecting viral loads as low as 112 copies per mL.187
Additionally, the dCRISTOR (Fig. 6a) assay uses dCas9-functionalized magnetic micromotors to bind amplified HIV-1 RNA and convert target presence into a binary digital signal. The system combines extraction-free LAMP, magnetic motion tracking, and deep learning for simple, label-free detection. It achieved 100% sensitivity and specificity in plasma samples, with a detection limit of 0.96 copies per μL, making it a low-cost and effective tool for point-of-care HIV diagnostics.152
A key innovation is the Cas-Roller assay, which employs Cas13a from Leptotrichia wadei (LwaCas13a) for direct RNA detection without amplification. Upon recognition of Ebola RNA, the Cas13a–crRNA complex is activated and cleaves a specially designed RNA-modified DNA hairpin probe attached to gold nanoparticles. The cleavage releases a single-stranded DNA “leg” that initiates a DNA nanomachine via catalytic hairpin assembly (CHA), leading to an amplified fluorescent signal. This system achieved a limit of detection as low as 291 aM (∼175 copies per μL) and completed the detection process in about 40 minutes at 37 °C.69
In a separate approach, Cas12a was incorporated into a microfluidic paper-based analytical device (mPAD), where it was combined with reverse transcription-recombinase polymerase amplification (RT-RPA) to detect synthetic Ebola genomic RNA. This method achieved a limit of detection of 11 aM, underscoring the ultra-sensitive capabilities of CRISPR-based detection when paired with isothermal amplification techniques.110
A notable development is the LOC-CRISPR microfluidic system, which employs Cas12a to detect multiple respiratory viruses, including SARS-CoV-2 variants (BA.1, BA.2, BA.5), H1N1, H3N2, influenza B virus (IVB), and human respiratory syncytial virus (HRSV). This chip-based platform integrates nucleic acid extraction, isothermal amplification (RPA), and Cas12a-mediated cleavage in a sealed, contamination-free format. Clinical validation showed 97.8% sensitivity and 100% specificity, with detection possible for viral RNA concentrations as low as 100 copies per mL within 60 minutes.111 Another Cas12a-based method, designed for rabies virus RNA, combines target binding-induced isothermal amplification with Cas12a's trans-cleavage activity. This electrochemiluminescence biosensor achieved a detection limit of 2.8 pM, offering high sensitivity and specificity without relying on complex instrumentation.135
For DENV detection, an electrochemical biosensor was developed using the CRISPR–Cas12a system. This method exploits the trans-cleavage activity of Cas12a, activated upon recognition of a DNA analog of DENV-4 RNA, to cleave a methylene blue (MB)-linked ssDNA probe immobilized on gold nanoparticles. Cleavage of the probe results in a measurable decrease in electrochemical signal, enabling sensitive detection. This system achieved a detection limit of 100 fM for DENV-4 RNA without any RNA amplification and demonstrated high specificity by discriminating against other DENV serotypes and unrelated viral RNA sequences, including ZIKV.142
In contrast, the ZIKV study employed a CRISPR–Cas13b system not for detection but for targeted inhibition of viral replication in mammalian cells. crRNAs designed against conserved regions of the ZIKV genome successfully guided Cas13b to cleave viral RNA, reducing infection levels. A fluorescent reporter system (mCherry fused to the ZIKV capsid) was used to quantify viral load, offering a potential foundation for diagnostic development.188 To improve strain-level detection of Zika virus, researchers developed NASBACC, a low-cost CRISPR/Cas9-based diagnostic module integrated with NASBA amplification. By exploiting Cas9's sequence-specific cleavage, NASBACC can distinguish between closely related Zika strains based on single-nucleotide differences. This adds a crucial layer of specificity to paper-based diagnostics, enabling accurate strain identification while remaining portable and suitable for low-resource settings.182
One notable strategy uses Cas13a in an amplification-free electrochemiluminescence (ECL) biosensor for detecting Escherichia coli O157:H7 RNA. This system exploits Cas13a's transcleavage activity to cleave self-enhanced ECL probes upon target recognition. The result is signal amplification without the need for nucleic acid amplification or co-reactants. This method demonstrated rapid detection (∼20 minutes), a wide linear detection range, and strong specificity, including successful application to clinical urine and blood samples.60
Another approach incorporates Cas12a in a method called T7/G4-CRISPR, which enables sensitive detection of bacterial RNA such as 16S rRNA. This assay converts a single RNA target into many DNA activators via a toehold-mediated strand displacement and T7 exonuclease-assisted recycling circuit. The DNA activators trigger Cas12a's trans-cleavage, preventing assembly of a split G-quadruplex (G4) reporter, leading to a measurable fluorescence decrease. The platform showed improved sensitivity over direct detection, was able to distinguish single-nucleotide variants, and was validated against clinical urine samples, avoiding complex thermal cycling.141
More recently, Cas14a1 has been shown to directly respond to RNA, leading to the development of the ATCas-RNA platform. This system uses RNA to activate Cas14a1's trans ssDNA cleavage activity, achieving high specificity and ultralow detection limits (attomolar range). It was successfully applied to detect bacterial RNA in contaminated milk samples, extending Cas14a1's utility beyond its original DNA-targeting scope.162
CRISPR-based systems, particularly those involving Cas12a and Cas13a, have been extensively adapted to detect cancer-related miRNAs with high sensitivity and specificity. By integrating these enzymes with amplification strategies or signal transduction mechanisms, researchers have developed robust platforms capable of detecting miRNAs in clinical samples such as blood, serum, urine, and tissue lysates.44,145
In the following subsections, we explore how different CRISPR–Cas-based systems have been applied to detect miRNA biomarkers associated with specific types of cancer, including lung, breast, colorectal, prostate, and glioblastoma. Each subsection highlights the key miRNAs linked to these cancers and summarizes CRISPR strategies developed for their detection.
Among these, miR-21 stands out as the most commonly targeted marker, consistently overexpressed in lung cancer samples. It has been detected using a range of CRISPR–Cas systems from serum, plasma, and cell lysates with attomolar sensitivity.86 Similarly, miR-155, another key biomarker, has been identified using Cas12a-driven electrochemical sensors84 and the LdCsm system,169 showing excellent specificity suitable for point-of-care applications.
Broader panels including miR-17, miR-92a, and EGFR mRNA33 have also been investigated using Cas13a-based electrochemical biosensors. These platforms can measure multiple targets on a single chip, distinguishing early-stage NSCLC from healthy samples. Cas12j, a more recently explored enzyme, was used to sensitively detect miR-21 and miR-92a in plasma, delivering results comparable to RT-qPCR standards.107
Further innovations extended to diagnostic validation. For example, miR-195 was used in an RCA-CRISPR-split-HRP (RCH) assay to confirm the specificity of NSCLC-targeted miRNA detection. The platform correctly showed no differential miR-195 expression between NSCLC patients and healthy individuals, reinforcing its accuracy.33
In another study, the Cas9-based RACE system was used to detect miR-221, miR-21, and miR-222 simultaneously. By coupling RCA with Cas9-mediated cleavage, the system generated fluorescent signals correlating to miRNA abundance in extracellular vesicles from lung cancer samples. Its results matched RT-qPCR, supporting its potential for clinical use.154
The detection of miR-21, one of the most studied biomarkers, has led to the development of several CRISPR/Cas12a-based approaches. These included systems combining split T7 polymerase transcription with Cas12a for fluorescence-based detection,134 electrochemical sensing platforms that simultaneously detected miR-21 and miR-155,85 and Au-nanobeacon biosensors achieving direct, attomolar-level sensitivity without the need for RNA amplification.96 Additional strategies, such as entropy-driven catalysis cycles, hybridization chain reaction (HCR) controllers, and hyperbranched rolling circle amplification (HRCA) coupled with Cas12a, further boosted detection sensitivity, reaching femtomolar limits (Fig. 6c).86 In another approach, platforms like SCas12a, employing split crRNA designs, enabled amplification-free and multiplexed detection of mature miRNAs, maintaining strong agreement with traditional RT-qPCR results.102
On the other hand, for miR-10b, another microRNA associated with breast cancer, an entropy-driven circuit (EDC) integrated with Cas14a enabled dual amplification, dramatically enhancing sensitivity and allowing detection even in complex samples like serum and cell lysates. This method capitalized on continuous ssDNA activator production to robustly trigger Cas14a's trans-cleavage activity.159
Among these technologies, the COMET system stands out, using Cas13a in combination with a catalytic hairpin DNA circuit (CHDC) to achieve two-stage signal amplification. This system successfully detected a panel of RNAs associated with non-small-cell lung carcinoma (NSCLC), including miR-17, miR-155, miR-19b, and miR-210, demonstrating its capacity to distinguish cancer patients from healthy individuals through analysis of serum samples. The high sensitivity of the COMET platform, capable of attomolar detection, highlights its adaptability for other cancers like colorectal cancer.33
In another approach, researchers utilized an asymmetric CRISPR assay based on Cas12a, leveraging competitive crRNA binding to enhance cascade signal amplification without the need for reverse transcription. This system demonstrated successful quantitative detection of miR-19a in plasma samples from bladder cancer patients, achieving excellent correlation with RT-qPCR results.99
Additionally, the EXP-J assay, employing Cas12j, further advanced miRNA detection by coupling exponential amplification reaction (EXPAR) with Cas12j's trans-cleavage activity. Applied to lung cancer biomarkers, the method effectively detected miR-21 and miR-92a in plasma samples, yielding results that closely matched traditional RT-qPCR data. The success of EXP-J in clinical samples suggests strong potential for its application in detecting colorectal cancer-associated miRNAs.107
The detection strategies center around the trans-cleavage activity of Cas12a. Upon recognizing a target-specific crRNA and a corresponding activator DNA sequence, Cas12a is activated and begins cleaving surrounding single-stranded DNA (ssDNA) reporters like fluorescence-based systems, and electrochemiluminescence (ECL)-based biosensors, generating strong and measurable signals. Signal amplification played a crucial role in boosting sensitivity.132 Strategies like the 3D DNA walker helped convert a single miRNA-141 molecule into multiple DNA activators, while hybridization chain reaction (HCR) circuits amplified the pre-crRNA needed for Cas12a activation. These smart designs allowed researchers to achieve extremely low detection limits, reaching the femtomolar and even attomolar range.127
While the primary applications were broader than just prostate cancer, the strong focus on miR-141 and miR-21 detection clearly signals the high potential of CRISPR-based diagnostics for prostate cancer screening. With their rapid detection times, high sensitivity, and ability to adapt to portable formats, these systems pave the way for future POC applications and early cancer diagnostics.
In a complementary study, researchers designed a novel sensor that integrated upconverted nanoparticles (UCNPs) with the CRISPR/Cas12a system, leveraging a dual enzymatic amplification strategy using exonuclease III (Exo III) and phi29 DNA polymerase. Upon near-infrared light activation, a hairpin probe on the UCNPs bound to miRNA-21, triggering Exo III-mediated recycling and subsequent rolling circle amplification (RCA) via phi29 polymerase. The resulting amplified RNA sequences activated Cas12a, which cleaved a fluorescent reporter to produce a strong, detectable signal. This dual amplification platform achieved a remarkable limit of detection of 6.01 fM for miRNA-21 and demonstrated excellent performance in real biological samples, including serum and cell lysates from cancer cell lines like MCF-7 and HeLa.137 Moreover, both studies confirmed that these CRISPR-based approaches offer excellent specificity, showing minimal cross-reactivity with non-target miRNAs, a critical feature for clinical application.
In essence, the combination of innovative crRNA designs, advanced signal amplification strategies, and the precise activity of Cas12a is pushing CRISPR technology toward becoming a transformative tool for cancer biomarker detection, offering the promise of earlier and more accurate cancer diagnostics.
Recent studies have demonstrated the utility of CRISPR-based platforms for detecting various plant-related RNA targets with high sensitivity and future prospects for crop field compatibility. For instance, tomato spotted wilt virus (TSWV) in tomato and thrips was detected using an indirect Cas13a-based assay following RPA amplification, achieving a detection limit of 2.26 × 102 copies per μL with a sensor response in 20 minutes.198 The glyphosate resistance gene transcript in soybean was detected indirectly using LwaCas13a via the SHERLOCK platform, providing a LOD of 2 aM.199 Banana ripeness profiling was demonstrated through direct detection of 1.839 pM.160 Tomato brown rugose fruit virus (ToBRFV), a major threat to solanaceous crops, was detected indirectly via LbCas12a in conjunction with RT-LAMP, offering detection sensitivity comparable to RT-qPCR and enabling visual detection in field settings.200
Mixed infections involving tobacco mosaic virus (TMV), tobacco etch virus (TEV), and potato virus X (PVX) were diagnosed using both Cas12a (indirect, via RT) and Cas13a/d (direct), with Cas13 also enabling viral load quantification and compatibility with lateral flow strips.201 Rice black-streaked dwarf virus (RBSDV) was detected indirectly using RfxCas13d, integrated with isothermal amplification for field deployment, achieving detection of only a few RNA copies.49 Maize chlorotic mottle virus (MCMV), a serious quarantine pathogen in maize production, was detected using Cas12a paired with RT-RAA, allowing visual fluorescence-based detection at dilutions as low as 10−5 from 2000 ng of total RNA.202 These recent findings highlight the significant potential of CRISPR diagnostic platforms for detecting RNA targets in agricultural applications.196,197
A recent study demonstrated the applicability of the Cas13a platform to detect RNA targets from the 16S rRNA gene of Cyprinus carpio and Oryzias latipes. C. carpio is an invasive species that disrupts aquatic ecosystems, while O. latipes is a native model species widely used in ecological research. In addition, the study showed that Cas13a could detect environmental RNA (eRNA) from filtered water samples, highlighting its potential for monitoring invasive species to mitigate biodiversity threats, tracking endangered species in aquatic systems, and supporting habitat conservation. This platform enables on-site detection of live organisms in ecosystems and facilitates real-time biodiversity surveillance in dynamic environmental conditions. The Cas13a assay achieved a LOD as low as 1 copy per μL for C. carpio RNA and 1000 copy per μL for O. latipes RNA.46
In addition, future perspectives have proposed the integration of CRISPR platforms for RNA detection into marine biomonitoring systems, aiming to enable rapid, sensitive, and on-site detection of environmental RNA targets such as those from harmful algal blooms (HABs), marine pathogens, and invasive species, with the support of AI-driven CRISPR RNA design for enhanced precision and field deployment.204
From the biological aspect, factors such as variability in efficacy among different Cas enzymes, off-target (non-target) detection issues, indirect or direct RNA detection mechanism, the frequent need for preamplification steps to achieve ultra-high sensitivity, the use of low-yield or non-specific guide RNAs for various diseases, and the biological stability of reagents all critically influence the efficiency of CRISPR/Cas platforms.16,19,205 Minimizing off-target effects and designing unique, high-yield, and efficient guide RNAs (gRNAs) through the use of machine learning and bioinformatics platforms could significantly reduce non-specific detection and enhance robust performance across diverse sequence contexts—both of which remain key areas of ongoing research.206,207 Also, synergistic innovation in both crRNA and DNA activator design is recommended to enhance the biological aspects of CRISPR/Cas platforms for RNA detection, considering the following key aspects. For crRNA design, future advancements are expected to focus on structurally dynamic architectures such as split,102 caged,208 or switchable209 crRNAs, which allow conditional activation of the CRISPR–Cas system in response to specific RNA targets. Additionally, for in vivo diagnostic platforms, incorporating chemically modified nucleotides is expected to improve crRNA stability against nuclease degradation.210 On the DNA activator side, to reduce background signals, utilizing refined toehold-mediated strand displacement systems is recommended.211,212 Also, integrating aptamers213 into activators can expand detection capabilities beyond nucleic acids. Finally, to facilitate the construction of more intelligent diagnostic platforms capable of multi-input processing and decision-making, the development of nucleic acid-based logic circuits (e.g., AND, OR gates) is recommended.214 Together, these innovative strategies provide promising opportunities to significantly increase the sensitivity, specificity, and flexibility of next-generation CRISPR diagnostics.
From the result readout aspect, the lack of digitized and portable platforms restricts the broader deployment of CRISPR diagnostics. Many existing systems still rely on conventional laboratory instruments such as plate readers or fluorescence microscopes, limiting their use in POC or field settings.215 The development of integrated, portable, and user-friendly readout systems, such as smartphone-based readers, is critical to unlocking the full potential of CRISPR technologies for decentralized testing.216–218 From the sample preparation aspect, the complexity of upstream sample handling remains a major bottleneck. Many biological samples require nucleic acid extraction and purification to remove inhibitors that can interfere with amplification or Cas activity.12,219 Therefore, innovations such as rapid lysis methods compatible with CRISPR reactions, microfluidic sample prep systems,220 and integrated “sample-to-answer”221 platforms are paramount to truly enable field-deployable CRISPR diagnostics.
From the scalability and cost-effectiveness aspect, several factors inhibit the mass adoption of CRISPR-based diagnostics. These include the relatively high costs and availability of reagents (Cas enzymes, guide RNAs, reporters), the complexity of device fabrication for POC use, and the need for regulatory approval processes such as those required by the FDA.12 Also the need for cold chain storage of Cas proteins and RNA components significantly restricts the deployment of CRISPR-based diagnostics in remote or resource-limited settings. Addressing this issue requires the development of thermostable Cas variants222 that can remain functional at ambient temperatures.
Finally, from the real clinical application aspect,223 challenges such as multiplexed detection where multiple targets need to be detected simultaneously remain.224 While the programmability of CRISPR offers exciting potential for multiplexing, practical issues such as guide RNA cross-reactivity, optimization of multiple Cas–gRNA reactions in a single assay, and clear signal differentiation still need to be systematically addressed. Additionally, while delivery of CRISPR-based detection systems into cells is less critical for ex vivo diagnostics, it becomes a major challenge for any potential in vivo applications, demanding further advancements in safe and efficient delivery methods.225
Particularly, Cas13's innate RNA-targeting capability has driven major advances in amplification-free detection, while Cas12a and Cas12b have demonstrated powerful one-pot isothermal amplification strategies suited for POC deployment. Cas9, traditionally a DNA-targeting enzyme, has been ingeniously re-engineered for RNA diagnostics, expanding its utility across imaging, amplification, and digital quantification platforms. Meanwhile, class 1 systems like Cas3 and Cas10 contribute unique signal amplification routes and ruggedness for low-resource applications. Emerging platforms, including Cas7–11-based protease sensors and Cas14-driven ultrasensitive assays, further push the boundaries of detection sensitivity and modular design.
Despite tremendous progress, ongoing challenges such as improving sensitivity in amplification-free systems, multiplexing capacity, and field-deployable integration remain focal points for future development. Nonetheless, the convergence of CRISPR technology with advanced biosensing, microfluidics, and AI-driven signal analysis signals a promising future for decentralized, rapid, and precise RNA diagnostics. As these tools continue to mature, their integration into clinical workflows has the potential to revolutionize early disease detection, therapeutic monitoring, and pandemic preparedness on a global scale.
This journal is © The Royal Society of Chemistry 2025 |