DOI:
10.1039/D5AN00428D
(Minireview)
Analyst, 2025, Advance Article
Bacterial detection with electrochemical, SERS, and electrochemical SERS sensors
Received
15th April 2025
, Accepted 30th July 2025
First published on 31st July 2025
Abstract
Bacteria are responsible for a plethora of infectious diseases, with potentially serious complications, including sepsis, especially in young, elderly, and immunocompromised populations. Quick and accurate bacterial detection is becoming critically important, in many areas, e.g. food safety, medical diagnostics, and public health. However, the currently available bacterial detection methods, such as plate culture, flow cytometry, enzyme-linked immunosorbent assay, and polymerase chain reaction, face limitations, being either too time consuming, too costly, or not diagnostically accurate. Alternative approaches seek to provide rapid and accurate diagnostics; electrochemical sensors and optical assays based on surface enhanced Raman scattering (SERS) aim to minimize cost and processing time while improving diagnostic accuracy. Here, we provide a review of recent reports utilizing these techniques for bacterial detection in various settings, as well as their combination, namely, electrochemical surface enhanced Raman spectroscopy (EC-SERS).
1. Introduction
Pathogenic bacteria are a major threat to human health. Between 2025 and 2050, it is estimated that 92 million deaths will be attributed to bacteria;1 by 2050, bacterial infections will account for 10 million annual deaths, surpassing the current cancer-related death toll.2 Moreover, the rise of antibiotic-resistant bacteria, such as methicillin-resistant Staphylococcus aureus (MRSA), causes a concerning increase in mortality, with an estimated 700
000 individuals killed worldwide by antibiotic resistant strains every year.2 Furthermore, bacteria are implicated in food safety, as even small numbers of bacteria can compromise food quality and lead to spoilage or contamination.3 Bacterial biofilms in infrastructure further compound the threat, as they can become a persistent cause of nosocomial infections, especially when caused by antibiotic-resistant strains; biofilms also contribute to the corrosion of equipment in other settings, for example pipes used in water distribution and oil and gas production.4
Given the impact of bacteria on human health, as well as in materials and the food industry, there is a growing need for the development of straightforward, quick, and sensitive biosensors that allow bacterial identification and monitoring. The “gold standard” in bacterial detection is bacterial culture, which requires a laboratory setting and several hours. Other well-established methods include the performance of polymerase chain reaction (PCR), flow cytometry, or mass spectrometry techniques, all of which are resource intensive, costly, and can have long turnaround times, making them unsuitable for routine bacterial detection and monitoring.5 New diagnostic techniques based on electrical and optical sensors, are emerging, aiming to complement existing assays, and provide specific bacterial detection quickly and at low costs.
Electrochemical (EC) sensors have become an established technology, with many applications over the last 50 years;6 they offer respectable sensitivity and selectivity, ease of fabrication, quick detection times, and potential for miniaturization.7 A successful paradigm, the commonly-used glucometer, utilizes screen-printed electrodes, and is commonly used at home by patients with diabetes.8 Electrochemical sensors monitor changes in electrical signals at the electrode surface to infer the presence of the analyte of interest. Combined with biorecognition elements, such as antibodies, aptamers, enzymes, and phages, EC sensors can become a powerful tool for the rapid and specific detection of pathogenic bacteria.9
Surface enhanced Raman spectroscopy (SERS) is another emerging alternative for bacterial detection.10 It is an optical spectroscopic technique based on metallic nanostructures, that produces spectral “fingerprints” of molecules under laser illumination, identifying many analytes, including ones related to bacteria, with unparalleled sensitivity.11 SERS is extremely versatile, but has, thus far, only been reported in the laboratory setting. SERS can be combined with electrochemistry, in a hybrid electro-optical method, EC-SERS, merging the strengths of both techniques.12
Furthermore, the increased demand for automated point-of-care detection and identification of bacteria has led to the development of lab-on-chip devices that incorporate these sensing technologies.12–15 Microfluidic integration is increasingly applied to pathogen detection, due to multiple advantages offered by the miniaturisation of the analytical processes. High sensitivity and selectivity of detection can be achieved on-chip, while simultaneously decreasing the sample volume needed for detection and accelerating the workflow. This results in decreased operating costs, as well as less generated waste.
In this mini review we present recent reports of EC, SERS, and EC-SERS sensors with applications related to bacterial detection, in healthcare and other settings. The three methods are illustrated in Fig. 1.
 |
| Fig. 1 (a) Electrochemistry-based approach utilizing aptamers for bacterial detection. An electrochemical reporter molecule can be used to provide indirect signal. (b) SERS nanoparticles provide spectral signals for label-free bacterial detection. (c) EC-SERS uses electrodes and metal nanoparticles for label-free SERS-based bacterial detection. Different Raman peaks are preferentially enhanced by the applied voltage, facilitating detection. | |
As one of the major focal points in this review is the detection of bacteria, it is important to distinguish between the two major bacterial types, Gram positive (+ve) and Gram negative (−ve) bacteria. This categorization originates from crystal violet staining, commonly used in microbiology, which penetrates and stains the thicker outer layer of Gram +ve bacteria.16 Gram −ve bacteria, such as E. coli, Salmonella spp. (species within the Salmonella genus) and Klebsiella spp. are often associated with food spoilage and the pathogenesis of sepsis and more severe infections.17,18 Gram +ve bacteria, such as Staphylococcus spp. and Listeria spp. are commonly associated with respiratory and skin infections and gastrointestinal issues.19 Moreover, the antibiotic resistance of S. aureus is becoming increasingly more threatening, highlighting the need for rapid bacterial identification.19
When discussing sensors, the term Limit of Detection (LOD) is used extensively. LOD is defined as “the lowest analyte concentration likely to be reliably distinguished from the blank and at which detection is feasible”.20 The suitability of LOD varies between applications and sample matrices. Whereas chemical analytes are reported in units of molar, for bacteria, LOD is given in Colony Forming Units (CFU) per millilitre or gram, i.e. how many bacterial colonies develop after applying the nominal quantity of sample to a culture plate. For example, according to Public Health Laboratory Services guidelines in Europe, the limit of Gram −ve bacteria in food is <100 CFU g−1.17 In contrast, clinical samples have varying diagnostic bacterial cutoff limits – for the diagnosis of urinary-tract infections, a bacterial count higher than 104 CFU mL−1 is usually indicative of infection;21 whereas, during sepsis, the concentration of bacteria found in blood is much lower (between 1 and 100 CFU mL−1 in adult patients).22
2. Electrochemical bacterial sensors
An electrochemical biosensor functions by generating an electrical signal that correlates with the concentration of an analyte of interest, often a biomarker. It consists of three primary electrodes: the sensing electrode, which provides the signal for biomarker detection; a reference electrode, which provides a stable potential; and a counter electrode, which facilitates current flow to complete the circuit. These components are immersed in the liquid sample, typically an aqueous solution with electrolytes to provide electrical conductivity.23 EC measurements can be performed in various modes, such as cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS) and differential pulse voltammetry (DPV). Reviews covering electrochemical sensors extensively are available.23–26
EC sensors can be categorized as either labelled or label-free. Label-free sensors rely on the analyte of interest to induce the change in electron flow, whereas labelled sensors use a reporter molecule to facilitate current flow. In both cases, a receptor molecule needs to be tethered to the sensing electrode, to provide affinity for the target. Label-free detection relies on the specific binding between the receptor and the target and is often simpler and more cost-effective, reduces contamination from additional labelling steps, has a shorter readout time, and is typically preferred for point-of-care systems.27 However, compared to labelled approaches, this technique is limited by lower sensitivity, non-specific binding, and the generation of false-positive responses, hence care must be taken to ensure diagnostic reliability.27,28 On the other hand, labelled detection employs additional components, other than the receptor, such as redox reporters (ferrocene, methylene blue), enzymes, nanoparticles, and other species. These moieties are either attached to the receptor or the target (via a “sandwich assay”) and provide a strong signal upon target binding. A major benefit of labelled sensors is increased sensitivity and specificity, especially in complex matrices.29 For example, the widely used blood glucose meter detects glucose through enzymatic oxidation.30 However, labelled sensors are often held back by increased costs, time-consuming reactions, and the risk of contamination through the introduction of additional steps.28,29 The sections below present recent label-free and labelled electrochemical approaches for sensing bacteria.
2.1 Label free EC sensing
There is an increasing demand for the development of label-free sensors, as their low fabrication costs and quick turnaround times make them attractive for fieldwork and point-of-care testing.27 Notably, the global market for label-free detection has grown by about 9% over the past five years and is expected to reach $3 billion in value by the end of 2025, while the point-of-care industry is expected to reach a value of $72 billion by 2027.31,32
Label-free EC sensors feature a working electrode functionalized with receptor molecules that directly bind the target analyte, such as aptamers, antibodies, or other moieties. Target binding is detected via changes in the electrical characteristics of the circuit without the need of an additional reporter.
Fig. 2 shows examples of three different label-free EC sensors for the detection of bacteria using aptamers (a), antibodies (b) and imprinted polymers (c).
 |
| Fig. 2 Bacteria captured by receptors provide direct EC signal. (a) Aptamers,33 (b) antibodies,34 and (c) imprinted polymers35 have been reported for label free EC bacterial detection. | |
2.1.1. Aptamer-based sensors. Aptasensors are sensors utilizing aptamers for target selectivity. Aptamers make a popular receptor for label-free sensors, due to their low production cost, high sensitivity, and ease of integration. Readers are referred to existing reviews covering the mechanisms of aptamer detection and conventional aptamer applications in sensing, including bacterial detection.36–39Recently, aptamers have been included in more advanced applications, typically as parts of nanocomposites. For example, a sensor utilizing a graphene oxide/poly deep eutectic solvent/nickel oxide was combined with aptamers specific to S. aureus, reporting a LOD of 10 CFU mL−1.40 Nanoparticles are often coupled with aptamers and used in EC sensors as a means of increasing surface area and sensitivity. For example, carbon nanofibers, gold nanoparticles and an aptamer specific to Acinetobacter baumannii, a pathogen associated with meningitis and sepsis, were used for detection of the pathogen in water and urine with a LOD of 0.6 CFU mL−1.41 The rapid detection of E. coli was investigated, using aptamer-conjugated silver nanoparticles on screen printed carbon electrodes, reporting a LOD of 150 CFU mL−1.33 In a different study, a glassy carbon electrode was modified by silicon-methylene blue nanoparticles and gold nanoparticles, followed by aptamer attachment via Au–N bond. The sensor reported sensitive detection of Listeria monocytogenes in fresh produce samples.42
2.1.2. Immunosensors. Single-antibody immunosensors (without the need of a secondary antibody) are commonly used as receptors in label-free sensors. For example, anti-S. aureus antibodies were immobilized on lipoic acid structured on screen printed electrodes made of gold nanoparticles coated with iron oxide, reporting the sensitive (0.10 CFU mL−1) detection in milk and meat samples.34 Similarly, antibodies have been used in conjunction with a nanoporous platinum layer, which increased the surface area, reporting rapid detection of S. typhimurium down to 10 CFU mL−1 concentrations.43 A fluidic integrated electrochemical-cell-chip has also been constructed, where antibodies against Aeromonas salmonicida were immobilized on microfabricated gold electrodes, enabling bacterial detection in seawater.44Similarly to aptasensors, nanoparticles and nanocomposites are often used in combination with antibodies to provide superior detection. One study described the use of zinc oxide nanorods on a sensor board, combined with anti-S. aureus antibodies, reporting sensitive and rapid detection on human hand skin.45 Another study reports the use of ferric oxide ionic liquid nanocomposite on gold electrodes, followed by polyglutamic acid deposition, on which anti-Salmonella antibodies were deposited, resulting in bacterial detection in milk.46 A similar study on the same pathogen and sample matrix reported the use of ferric oxide on graphene as a nanocomposite combined with antibodies, reporting similar levels of detection.47
Finally, label-free immunosensors can be used for the simultaneous detection of multiple bacterial species. A recent example describes the use of a graphene oxide-copper (GO-Cu) metal–organic framework (MOF) nanocomposite on dual screen-printed electrodes, followed by functionalization by pyrene and a different antibody immobilized on each electrode.48 The sensor reported simultaneous detection of Mycoplasma pneumoniae and Legionella pneumophila in tap water samples.
2.1.3. Polymers. Polymer coatings on sensors have been explored as a versatile tool. Polymer films can provide affinity/specificity for the target molecule, facilitate the attachment of detection moieties such as antibodies, or provide a secondary signal, such as fluorescence.For example, an EC immunosensor based on a polydopamine film on the surface of a copper–iron (Cu–Fe) MOF nanocomposite was developed. The anti-Salmonella antibodies were immobilized onto the polydopamine film, enabling Salmonella detection in water and milk samples.49 Similarly, another sensor based on an indium tin oxide (ITO) electrode used electropolymerized poly melamine and poly glutamic acid film, on which anti-Shigella antibodies were immobilized, reporting sensitive detection in food samples.50 Another example reports the use of fluorescent nanofibers obtained by the electrospinning of quantum dot mixture in polyacylonitrile solution, before being deposited on the surface of gold electrodes, followed by the attachment of anti-Streptococcus agalactiae antibodies.51 The system reported real-time sensitive detection in diluted blood plasma.
Imprinted polymers offer selectivity for various analytes without the use of other targeting elements and are becoming increasingly popular in label-free sensors. One study reports the detection of L. monocytogenes in aqueous samples through the use of an imprinted polymer sensor. To produce the imprinted polymer, bacteria and pyrolle were mixed, and following electropolymerization, the bacteria were removed, leaving behind imprinted cavities in their shape. The study reported a LOD of 70 CFU mL−1.35 Imprinted polymers can also be combined with aptamers, creating hybrid imprinted polymers. The hybrid approach has gained popularity, as it allows for more sensitive detection of molecules, as the incorporation of recognition elements in molecularly-imprinted polymers (MIPs) is hypothesized to result in more homogenous binding sites that are readily accessible by the target.52 One recent example describes the development of a hybrid imprinted sensor, combining aptamers, magnetic nanoparticles and polymers. Using a magnetic nanoparticle-modified glassy carbon electrode as base, aptamer-modified gold nanoparticles were deposited on the sensor's surface. Following attachment of S. aureus, o-phenylenediamine was electro-polymerized, and the bacteria were subsequently washed away, revealing aptamer-lined cavities. The sensor was reported to sensitively detect S. aureus in milk, conduit water, and apple juice samples.53
Additionally, polymers can be combined with aptamers or antibodies in non-imprinted sensors. One such example describes the use of self-polymerized polydopamine on glassy-carbon electrodes, combined with amino-functionalized vertically-ordered mesoporous silica film, on which aptamers were bound, reporting V. parahaemolyticus detection in shrimp samples.54 Moreover, polydopamine has been used as a cross-linker on an MXene platform to immobilize antibodies against Listeria monocytogenes.55
Label-free EC sensing of bacteria has been integrated with microfluidic devices of various configurations. Microfluidic devices featuring electrodes functionalised with antibodies56,57 or aptamers58 have been used for selective bacteria detection, demonstrating fast detection times, and depending on the device configuration, achieving low LODs. For example, a microfluidic device was developed which used fluidic focusing to direct bacteria to interdigitated electrode arrays (IDEAs) functionalized with antibodies for impedance sensing. This device demonstrated a LOD as low as 39 CFU mL−1 for E. coli 0157:H7,56 while a similar device was reported to have achieved a LOD of 7 CFU mL−1 for Salmonella sp.57
Beyond simple detection, however, an advantage that microfluidic devices offer is the ability to trap bacteria on-chip in order to monitor cell behaviour and cell growth over time. This can be further combined with electrochemical sensing for cell growth analysis and antibiotic susceptibility testing.59,60
2.2 Labelled EC sensing
The majority of published electrochemical work for bacterial detection describes labelled approaches. It is important to clarify that the term “labelled”, when used in the context of electrochemical sensors, refers to the use of secondary reporters (e.g., aptamers/antibodies, electrochemical labels, molecular markers) in addition to a primary detection moiety. These secondary reporters serve to provide a measurable electrochemical signal in the presence of the target. Fig. 3 shows examples of labelled EC sensing utilizing (a) aptamers, (b) antibodies, (c) peptides, and (d) imprinted polymers, all in conjunction with secondary labelled molecules.
 |
| Fig. 3 Examples of labelled electrochemical sensor for bacterial detection. (a) Using S. aureus aptamers in conjunction with DNA cycling amplification based on CRISPR-Cas12a.61 (b) MXene-based sensor utilizing antibodies specific for S. aureus.62 (c) MXene-based sensor utilizing antimicrobial peptides for E. coli detection.63 (d) Glassy-carbon based sensor utilizing an imprinted polymer film for the detection of S. aureus.64 | |
2.2.1. DNA amplification. Recently, there has been an increase in the number of scientific works utilizing amplification techniques to improve the EC signal, and enzymatic DNA amplification methods are a popular choice. It is important to note that in this context whole bacteria are the detection target – DNA serves solely as a convenient means of signal amplification, due to the well-established nature of DNA amplification methods.One recent example, illustrated in Fig. 3a, used enzymatic tyramide signal amplification technology to detect S. aureus. Specifically, an aptamer was used to immobilize the bacteria on the sensors, followed by a secondary aptamer labelled with Horseradish Peroxidase (HRP), which catalyzed the amplification technique, achieving a low LOD of 3 CFU mL−1.61 Another amplification approach is described in a study which utilized saltatory rolling circle amplification combined with the enzyme Cas12a to detect S. aureus. The binding of the bacteria onto thiolated, methylene blue-labelled DNA strands activated the cleavage activity of the enzyme, releasing methylene blue into the solution and resulting in a drop of the electrochemical signal, allowing detection with LOD of 3 CFU mL−1.65 Another amplification approach is described in a study, where the detection of S. aureus by aptamers caused the release of DNA by the aptamer, triggering a cascade which resulted in the binding of ferrocene-labelled DNA, achieving increased detection sensitivity, with a LOD of 0.3 CFU mL−1 in both sterile water and seawater samples.66 CRISPR/Cas12a amplification has also been used to detect E. coli in conjunction with aptamers in both water and milk, with LOD of 5 CFU mL−1 in both samples, as well as for the detection of Vibrio parahaemolyticus, in both aqueous and mussel samples, with a LOD of 1.5 CFU mL−1.67,68 Hybrid approaches have also been developed, through the use of aptamers for the capture of Salmonella enteritidis combined with Cas12a and MOF-based electrochemiluminescence emitters, with a LOD of 37 CFU mL−1, followed by successful detection of the bacteria in waste water, milk, and whole blood.69
There are also examples of sensors utilizing an enzymatic, non-CRISPR amplification cascade. One such example describes the use of lubricant-infused surface DNAzymes, which bind onto species-specific proteins present on the cell wall of bacteria, resulting in successful detection in both water and milk samples.71 Another similar sensor used photoactive DNAzymes modified with semiconductor structures between titanium dioxide (TiO2) nanomaterials, enabling the sensitive detection of E. coli in water, even in the presence of other bacteria.72 Finally, a dual-mode electrochemical-electrochemilumine-scence sensor has been described, where hybridized aptamer-silver nanocomposites bound and killed S. aureus, resulting in the release of hybridized DNA into the solution, initiating an endonuclease cascade which resulted in the hybridization of ruthenium-labelled sequences, thus enabling sensitive detection (1 CFU mL−1).73
Non-enzymatic amplification. Non-enzymatic signal amplification sensors utilize nanostructures as amplification labels, complementary DNA and RNA sequences and recycled hybridization reactions, instead of enzymes. Such examples have been reported, one of which integrated gold nanoparticles with a MOF and aptamers specific to S. aureus. This study used a dual DNA amplification approach which resulted in a low LOD (1 CFU mL−1) in aqueous samples, as well as in orange juice and commercial milk.74 Similarly, another dual amplification approach has been described, where aptamers specific to S. aureus were used in combination with polymethylene blue nano-dumbbells and a hybridization chain reaction, resulting in LOD of 1 CFU mL−1.75 An interesting, non-enzymatic approach, which combines MIPs with MOF-assisted redox recycling amplification, has been reported, which utilized methylene blue as a label and recorded the redox recycling of iron facilitated by methylene blue, resulting in LOD of 1 CFU mL−1.76 A novel approach utilizing quantum dots and mesoporous silica nanospheres as amplification labels was able to detect S. aureus in both aqueous and milk samples with LOD of 2 CFU mL−1.77 A dual amplification approach using aptamers specific to Pseudomonas aeruginosa, MIL-101(Cr)/multi-walled carbon nanotubes, and a carbon nitride complex with silver nanoparticles as a label, achieved LOD of 1 CFU mL−1 in both aqueous solutions and diluted blood serum.78
2.2.2. MXenes. MXenes are 2D materials composed of carbides and nitrides of transition metals commonly used for electrochemical bacterial detection. Thanks to their conductivity, extensive surface area, hydrophilic nature and facile functionalization, they have attracted the interest of researchers.79Antibodies have been combined with MXenes to produce novel and sensitive sensors. One example describes the use of MXene@Au on a glassy carbon electrode combined with antibodies specific for Methicillin-resistant S. aureus (MRSA), employing DPV and ferroceneboronic acid as a label, and detecting bacteria with an LOD of 3.8 CFU mL−1 in cerebrospinal fluid samples, as shown in Fig. 3b.62 Antibodies were also used for the development of a sensor capable of detecting Vibrio parahaemolyticus, using antibodies to immobilize the bacteria followed by a MXene/gold nanobipyramid/antimicrobial peptide signal probe and a ferrocene-labelled DNAzyme, resulting in an LOD of 6 CFU mL−1 in aqueous samples.80
Aptamers have similarly been used with MXenes. For example, an aptasensor specific for S. aureus, was reported based on recycling of DNAzyme activation on gold nanoparticles-functionalized MXene nanomaterials, using methylene blue-labelled DNA probes, achieving LOD of 5 CFU mL−1 in aqueous samples and retaining its activity in food samples.81 Aptamers can also be paired with imprinted polymers, as was the case in a study which developed a hybrid polymer specific for S. aureus, paired with in situ growth of a MOF on MXene tagged with methylene blue and vancomycin as a label, resulting in LOD of 10 CFU mL−1 in aqueous solutions and spiked food samples.82 Moreover, aptamers can be combined with signal tags, such as two-dimensional Zn-MOF on MXene sheets, allowing detection of E. coli, S. aureus and S. typhimurium by using the suitable aptamer.83 An interesting example of an MXene aptasensor describes the use of 2D titanium carbide (Ti3C2Tx) MXene, on which aptamers were bound. Following that, a ferrocene-based MOF containing gold nanoparticles and 4-mercaptophenylboronic acid, which binds to endotoxin, was used as a signal label, which also functioned as a Faraday cage.84
MXene nanocomposites can also be combined with enzymatic amplification strategies, such as with the CRISPR/Cas12a enzymatic amplification cascade. One such sensor utilized ss-DNA sequences to capture S. Typhimurium, enabling detection in aqueous and chicken meat samples, with a LOD of 160 CFU mL−1.85 Non-CRISPR enzymatic amplification is also possible; aptamers specific to S. aureus were attached to gold nanoparticles-functionalized MXene nanomaterials, and subsequently hybridized with DNA sequences containing the blocked DNAzyme.81 Upon bacterial binding, the DNAzyme was released, initiating a DNA recycling cascade, resulting in bacterial quantification down to 5 CFU mL−1.
Additionally, MXene sensors can be developed without any biorecognition molecules. One such example is an impedimetric imprinted sensor for the detection of Salmonella, where mixture of pyrrole, MXene and Salmonella was electrochemically polymerized onto a glassy carbon electrode, resulting on a MIP with LOD of 23 CFU mL−1 in aqueous and food samples.86
Other labelled sensors. Antimicrobial peptides are also used in sensors, as illustrated in Fig. 3c. One example describes the development of an antimicrobial peptide/MXene platform specific to E. coli, followed by a phage-apoferritin@CuO2 label, which resulted in a LOD of 30 CFU mL−1 for live bacteria and 6 CFU mL−1 for all bacteria tested, in aqueous and food samples.63 Furthermore, aptamers have also been used in conjunction with antibiotics. One such example describes the use of vancomycin as a receptor, and the use of an aptamer as a labelling molecule, reporting successful detection of S. aureus and B. cereus in water, milk and bovine serum.87There have been reports of other sensors utilizing aptamers. For example, an aptasensor specific for E. coli has been developed, where methylene blue-labelled aptamers were used on gold screen-printed electrodes.88
There is interest in the development of dual-mode sensors combining electrochemistry and a secondary detection method. One such sensor combined electrochemistry and electroluminescence and used dual-bipolar electrodes, capture-DNA and luminol filled anodes and ferrocene-labelled aptamers, resulting in the sensitive detection of E. coli (LOD 1 CFU mL−1).89 Aptamers have also been combined with glucose oxidase labelled probe-DNA, releasing the enzyme upon recognition of S. aureus, allowing it to react with glucose, causing a decrease in signal and the sensitive quantification of bacteria.90 Another example describes the use of functionalized gold nanoparticles utilizing tetrahedral DNA nanostructures to bind Vibrio parahaemolyticus, combined with CuO2 nanodot-mediated MOF nanozymes, releasing copper ions upon bacterial recognition, which can be subsequently detected by both colorimetry and square wave voltammetry.91
Aptamers can also be combined with imprinted polymers. One such example, shown in Fig. 3d, describes the combination of aptamers with bacterial imprinted polymer film, plus 6-(ferrocenyl)-hexanethiol functionalized gold nanoparticles, capable of detecting E. coli at the single cell level, while also demonstrating satisfactory detection in milk.64
In addition, sensitive aptasensors can be developed through the inclusion of nanomaterials. For example, a S. aureus specific aptamer was immobilized on a 3D reduced graphene oxide and chitosan film, followed by hybridization with a DNA-zeolitic imidazolate framework-8-ferrocene labelled sequence, the displacement of which allowed for bacterial quantification.92
Numerous sensors utilizing antibodies have been reported in literature, either in combination with a secondary antibody or nanoparticles. In particular, “sandwich” approaches are common with antibodies, providing superior detection and selectivity compared to single antibodies. One example used laser-scribed, graphene oxide electrodes, combined with anti-E. coli antibodies plus HRP-labelled secondary antibodies, allowing bacterial detection in water and artificial urine samples with a LOD of 283 CFU mL−1.93 Similarly, a 3D printed sensor containing electroplated bismuth film on the sensor surface, which utilizes anti-S. typhimurium antibodies followed by streptavidin labelled with quantum dots, achieved sensitive detection in chicken broth (1 CFU per 25 mL).94 Electrodeposition has also been used to deposit chitosan on glassy carbon electrodes, on which anti-S. Typhimurium antibodies were bound, in combination with HRP-labelled antibodies.95
3. SERS for bacterial detection
Surface enhanced Raman spectroscopy (SERS) was first observed in 1974 when an electrochemical apparatus with roughened silver electrodes was illuminated using a Raman laser; Fleischmann et al. noted exceptionally high quality Raman spectra for pyridine adsorbed on the electrodes.96 This effect was identified as a distinct phenomenon shortly thereafter and coined as SERS. Its spectral signal offers excellent molecular specificity, typically referred to as a molecular “fingerprint”, while the enhancement afforded by nano-engineered substrates allows even single-molecule sensitivity.97 SERS has found many applications in molecular imaging, medical diagnostics, chemical analysis, food safety and environmental monitoring, including in clinical and food samples.10,11,98,99
By tuning the plasmonic resonances of metallic nanostructures with the excitation laser, the unique “fingerprint” signal of the target can be enhanced by 1014–1015, which makes SERS a powerful analytical technique, including for bacterial detection.100 The first reported use for bacterial classification using silver nanocolloids was in 2004 by Roger M. Jarvis;101 more recently, novel nanomaterials have emerged for this application, including nanowires, nanoislands and nanopancakes.101–104 Additionally, SERS can be paired with microfluidic chips, which can improve sample processing and assay accuracy, as well as lateral-flow assays capable of detecting multiple pathogens.105,106
It is important to distinguish between direct SERS and SERS tags (labelled or indirect SERS). Direct SERS refers to the simplest layout, where intrinsic molecular vibrational modes of the target analyte are enhanced by plasmonic nanostructures and provide its unique spectrum identifying the molecule. In contrast, indirect SERS involves using tags with Raman reporters with affinity to the target analyte, that rely on the reporter's spectrum – not the analyte's – for detection. Additionally, molecular chemosensors can serve for indirect SERS sensing, as they are subjected to changes in their spectral profile and/or intensity due to recognition events involving the target analyte.107
3.1 Direct SERS sensing
Label-free SERS approaches are preferred for bacterial detection, as evidenced in the literature. This is due to the method's accessibility and low cost. The employed nanomaterials can be divided in two categories: colloidal suspensions and nanostructures on a substrate. For direct SERS sensing, extensive spectral processing and analysis is required for target identification, particularly in complex media. Traditionally this is performed with multivariate analysis methods such as principal component analysis (PCA), least-squares regression (CLS, PLS, etc.), often followed by discriminant analysis (DA); more recently methods based on machine learning have been reported. It is worth noting that most studies report detection of the presence and identification of the bacterial species, while quantitative detection (along with LOD) is typically not provided.
3.1.1 Nanoparticles in solution.
Silver nanoparticles. Silver-based nanomaterials are commonly used in SERS, as silver produces high enhancement to the Raman scattering signal, low toxicity to non-pathogenic cells, has notable thermal stability, near ideal dielectric properties, and lower cost, compared to gold.102 An important consideration for the detection of bacteria is their peptide structure, which exhibits an affinity for silver, facilitating the attachment of silver-based structures for SERS.102A simple approach is the deposition of silver nanoparticles (AgNPs) mixed with the sample on silicon wafers or a glass slide, to form the “coffee-ring” stain. Once dried, the deposited nanoparticles are interrogated with a Raman laser to provide the spectral “fingerprint” of the sample.108 This method has been used along with a multiscale deep-learning method, reporting the discrimination of four different strains of Shigella.109 A similar approach reported the use of a machine learning model for differentiation between microorganisms (Lactobacillus spp., Candida spp.) and detection of bacterial vaginosis.110 Alternatively, Raman interrogation can be performed in liquid droplets. Droplet optical microcavities on aluminized hydrophobic glass slides were used for the detection of lactic acid bacteria (Lacticaseibacillus paracasei subsp. tolerans) reporting a high detection rate (>95%).111
Silver coated gold nanoparticles have been successfully employed for spectral analysis of E. coli, S. aureus and S. typhimurium. Spectral analysis was performed to characterize bacteria in mixed-species samples.112 Silver nanoparticles have been used for the identification and classification of multiple pathogens in beef.113 The colloid was mixed with the bacterial solution to optimize the SERS signal, and, following signal acquisition, the bacteria were differentiated. The sensor was capable of classifying S. typhimurium, E. coli, S. aureus, L. monocytogenes, L. innocua and L. welshimeri with reported average accuracy of 95.65%.
The rise of antibiotic resistance causes widespread concern, as the emergence of drug-resistant bacteria is becoming one of the leading causes of illness and mortality, and rapid detection of antibiotic-resistant bacteria and proper antibiotic guidance can be lifesaving.70 SERS has been used for the discrimination between antibiotic resistant and susceptible bacteria. Silver nanoparticles have been used as SERS substrates to successfully differentiate between carbapenem antibiotic-resistant and antibiotic-sensitive Klebsiella pneumoniae bacteria through spectral analysis using a convolutional neural network (CNN) model.114
Similarly, silver and gold nanoparticles were mixed with both antibiotic resistant and susceptible K. pneumoniae strains, and following spectral acquisition and analysis, antibiotic resistance was reportedly predicted with 99.46% accuracy.115 A recent study developed a residual neural network model, capable of not only identifying carbapenem antibiotic resistant bacteria but also discriminating them by species and enzyme type, as well as successfully predicting corresponding sensitive antibiotics.116 Similar methods have been employed for the detection and characterization of methicillin-resistant and susceptible Staphylococcus aureus, through spectral analysis using deep-learning models.117
Bacteria infected with bacteriophages can also be characterized using SERS. Silver nanoparticles were used to monitor the changes in S. aureus after interaction with bacteriophages. Specifically, after infection, SERS was performed to monitor temporal changes in pellets of infected bacteria.118 A similar study differentiated between phage-resistant and phage-susceptible methycillin resistant S. aureus, reporting 100% accuracy and specificity, in bacterial pellets and supernatant.119
AgNPs have also been used in microfluidic devices, where the sensitive SERS detection can be combined with the precise cell handling that microfluidics facilitates, to further enhance bacterial detection. As a result, on-chip enrichment of bacterial samples using a nanoporous membrane has been applied to improve detection sensitivity120 and optical tweezers have even been used to achieve single cell immobilisation for SERS interrogation, significantly enhancing the stability of the SERS signal recorded.121
In a separate instance, an optofluidic device was used to achieve parallel quantification and identification of pathogens on-chip at concentrations as low as 4 CFU mL−1.122 Quantitative detection was achieved by automated counting of the spectral events recorded as the bacterial cells flowed through the photonic fibre. Microfluidic devices also enable the in situ synthesis of the nanoparticles. AgNPs were synthesized in situ in a microfluidic device for the detection of E. coli in food samples. The AgNP synthesis was guided by aptamers that had been previously bound to the bacteria cells and which enabled uniform distribution of the AgNPs on the cell surface, enhancing the specificity and sensitivity of SERS detection.123
Gold and other materials. Gold nanoparticles have also been used for the successful monitoring of the vaginal microbiome.124 AuNPs were mixed into aqueous bacterial solutions and synthetic vaginal fluid solution, before being deposited on a glass slide covered with aluminum foil and being allowed to dry. The study reported the successful detection of Lactobacillus species, L. crispatus and L. iners in complex solutions. E. coli, S. aureus, and P. aeruginosa have been detected through the use of aluminum foil integrated pegylated gold nanoparticles as a SERS substrate.125 Notably, the bacteria in this study were used directly from the culture without any pretreatment, and the gold nanopartciles were drop-casted on the aluminum slide, reporting sensitive detection with the least bacterial concentration necessary for clinical diagnosis of urinary tract infections. Non-pathogenic bacteria have also been investigated with SERS. Bacteria capable of removing sulfur from fossil fuels though the metabolism of dibenzothiophene can also be detected using SERS combined with gold nanoparticles. One such example reports the use of gold nanoparticles for the identification and characterization of sulfur-metabolizing bacteria (Chryseobacterium sp.IS, Gordonia sp.4N, Mycolicibacterium sp.J2, and Rhodococcus sp.J16) leading to bacterial differentiation with high accuracy in soil and sludge samples.126Over the past ten years, molybdenum disulphide (MoS2) has attracted a lot of interest in the field of materials science and in SERS because of its layered structure, which resembles graphite, and because it exhibits unique anisotropic electrochemical, electrical, and optical properties.127 It can often be combined with tin sulfide (SnS2), forming a nanocomposite that is more economical than gold or silver. A study used one such nanocomposite, functionalized with L-cysteine, reporting accurate and specific detection of E. coli.127
Silicon-based SERS substrates have also been reported in published literature. Aptamer-coated magnetic beads were used to bind to and isolate target bacteria (E. coli, S. Typhimurium, S. aureus).128 Following that, the bacteria were isolated and pipetted onto a silicon-based SERS substrate, which, combined with machine learning and spectral analysis, reported identification and classification of bacteria.
3.1.2 Non-colloidal nanostructures. Metal nanostructures can be grown on a surface to provide the plasmonic enhancement necessary for SERS. Although this approach is in general more complex and costlier than the use of colloidal aggregates, it can be considerably more controllable and reproducible. A more versatile approach is to tether pre-synthesized nanoparticles on a larger fixed structure such as a polymer, producing a more stable nanoparticle arrangement without the need of extensive nanofabrication.
Silver-based sensors. A simple nanostructured SERS substrate was reported with silver nanoparticles combined with the elastomer polydimethylsiloxane (PDMS), forming a sponge-like structure through the “mold transfer-surface embedding” method.100 This substrate also provided the benefit of reduced risk of shedding and oxidation of silver nanoparticles, prolonging their usage. The sponge structure successfully and rapidly entrapped E. coli in close proximity to the nanoparticles and reported sensitive detection (LOD 1 CFU mL−1) in water and milk samples and differentiation between six bacterial strains using an ML model. Silver nanocubes embedded in PDMS have been also employed for the detection of E. coli and S. typhimurium on eggshells. The nanocubes were prepared by a sulfide-mediated polyol method, followed by arrangement on PDMS membranes. Following spectra acquisition, analysis was performed to differentiate between bacterial spectra with high accuracy reported.129Nanospheres, nanowires, and nanoseeds have been successfully combined on an amine-functionalized PDMS surface.102 The amine groups displayed high affinity for silver, allowing the orderly arrangement of nanospheres and nanowires on the PDMS surface, creating ideal hotspots for bacterial detection with high density. The sensor was capable of detecting Pectobacterium carotovorum subsp. Carotovorum, a plant pathogen, in kimchi cabbage juice, at single cell level.
Paper has also been used as a support for plasmonic nanostructures. Two-arm zigzag nanorods have been fabricated on a porous and fibrous cellulose substrate using glancing angle deposition method, which reported a tenfold increase in signal compared to nanorods deposited on a glass substrate, owing to the wicking ability of the paper substrate.130 Detection of P. aeruginosa, E. coli, and S. aureus was reported, as well as distinction between Gram +ve and Gram −ve bacteria.
Nanowires are also widely used as SERS substrates. One example described the use of silver coated nanowire SERS chips, which formed a 3D matrix containing tightly placed metal nanoparticles, providing ideal hotspots for bacterial attachment.131 Combined with a neural network model, the sensor was reported to differentiate between twelve bacterial strains and two strains of antibiotic-resistant E. coli in real time. Silver nanowires have been used in a three-layer arrangement, reporting the sensitive detection of Enterobacter, Staphylococcus and S. albus. The composite was constructed with nanowires at the center, ZIF-8 as a shell, and AgNPs on the exterior, and was capable of entrapping and inactivating bacteria, as well as enhancing localized surface plasmon resonance via ZIF-8 control. The sensor was capable of semiquantitative analysis as well as characterization and differentiation of the bacteria in water samples.132
A microfluidic device utilised a membrane filter to trap and enrich E. coli samples and then detect the SERS produced via a silver film substrate incorporated in the same device. The enrichment resulted in a 4-order of magnitude improvement in detection sensitivity over what could be achieved off-chip.133 An example of a SERS-active substrate used for microfluidic interrogation after the microfluidic device fabrication is found in a different study, where a microfluidic microwell device was used to trap bacteria. This was combined with SERS detection to monitor the growth of trapped bacteria and perform antibiotic susceptibility tests.134 A silver island-film coated glass substrate was placed on the microwell device to provide the SERS enhancement.134,135
Gold-based sensors. One example of gold nanoparticle application for bacterial detection describes the use of an inverted pyramid microcavity array, a microchannel cover plate and a multilayer gold nanoparticle substrate at the bottom of the microcavities, using only a few picoliters of sample to identify live bacteria.136 Each microcavity only contained one cell or no bacteria at all, and the substrate provided sensitive, single-cell bacterial detection.E. coli and S. aureus were detected using gold nanocubes on two-dimensional (2D) delaminated nano mica platelets, forming a hybrid substrate.137 Through self-assembly, the nanocubes and the mica platelets were able to form hotspots, enhancing the Raman signal. The combined hydrophilicity and hydrophobicity of the substrate attracted both S. aureus and E. coli, reporting sensitive detection for both bacteria.
An example microfluidic device incorporating a SERS active substrate to detect E. coli contained a uniform layer of upconversion nanoparticles with silica coating and gold nanoparticles formed through a self-assembly and chemical plating process.138
MXene has gained popularity in SERS sensors, in addition to electrochemical ones, as they have demonstrated antibacterial properties, which can potentially make them an attractive target for an antibacterial strategy.139 In a recent study, an MXene-gold nanoparticle nanocomposite was synthesized through a self-assembly method before being deposited on a cellulose substrate by freeze-drying. The paper sensor reported sensitive detection of E. coli and MRSA in porcine skin and blood serum, in addition to bactericidal effects.139 A similar study, utilizing MXene and gold nanoparticle nanostructures on electrostatic self-assembly, reported rapid and label-free detection of E. coli without tedious pre-processing.140 Additionally, laser irradiation increased the substrate temperature to 50 °C within 200 seconds, resulting in the release of reactive oxygen species, efficiently killing the bacteria.
Gold nanostars were used as a solid-state substrate on etched stainless steel by chemical replacement reaction, reporting detection of E. coli, S. epidermidis, S. enteritidis and S. aureus in beverage samples.141 Following Raman acquisition, spectral analysis allowed the identification and classification of the pathogens.
Vertically-aligned gold nanowires have been successfully used for the detection of S. aureus, E. coli, P. aeruginosa and S. epidermidis in tear samples.142 Coupled with a microfluidic chamber producing turbulent flow, which promoted contact of the bacteria with the nanowires, the sensor allowed the sensitive detection and discrimination of bacteria.
3.2 SERS tags
Unlike direct SERS sensing, which seeks the spectral signature of the analyte itself, indirect sensing uses SERS nanoparticles with a pre-encoded spectrum as reporters to infer the presence of the analyte. These nanoparticles are typically bright, chemically stable, and rely on targeting molecules (e.g. antibodies) to bind to their target. They can serve for molecular imaging in living organisms or incorporated in point-of-care diagnostic assays.143 In contrast to label-free sensors, the majority of published work on labelled SERS sensors utilizes gold substrates, due to the ease of functionalization with thiolated receptors and its chemical inertness.
Silver coated, gold core–shell nanorods have been combined with aptamers for the sensitive detection of S. Typhimurium.144 In particular, 4-mercaptobenzoic acid (4-MBA) was used to provide the Raman spectrum and also to bind aptamers onto Au@Ag nanorods. Upon sample addition, the aptamers preferentially bound to the bacteria, reducing 4-MBA concentration and resulting in a quantifiable decrease in SERS intensity.
Gold nanoparticles are often paired with aptamers. One such example describes the use of gold colloid nanoparticles, coupled with a blocked aptamer with high affinity for E. coli.145 In the presence of E. coli, the blocked aptamer released the blocker DNA, hybridizing with the capture DNA on the SERS probe and kickstarting a hybridization chain reaction, resulting in double-stranded DNA formation and reduction of the SERS signal intensity of 4-mercaptobenzonitrile. The sensor reported high specificity and selectivity for E. coli in tap water and milk, using a portable Raman spectrometer.
SERS tags can be combined with magnetic elements to help separate them from within complex media. Gold-core, silver outer shell magnetic round nanoparticles, functionalized with concanavalin A, which specifically binds terminal α-D-mannosyl and α-D-glucosyl on bacteria, formed a sandwich SERS sensor which was used for the sensitive (LOD 1 CFU mL−1) detection of E. coli, S. aureus and P. aeruginosa, following exposure to a magnetic field, in water and milk samples.146 Raspberry-like, wheat germ agglutinin-modified gold based magnetic nanoparticles have been reported as a bacterial entrapment method, followed by aptamer-functionalized gold nanoparticles.147 This sensor was also combined with co-freezing aptamer/5,5′-dithiobis-(2-nitrobenzoic acid) gold nanoparticles as signal tags. The sensor reported sensitive and rapid (>20 minutes) detection of E. coli in different sample types, including human urine, serum, and skimmed milk, as well as DNA from Mycobacterium tuberculosis. Superparamagnetic iron oxide nanocomposites, combined with a layer of self-assembled gold nanoparticles have been applied for the detection of K. pneumoniae and A. baumannii.148 In particular, the gold nanoparticles were functionalized with vancomycin, which functioned as a universal capture probe for both bacteria strains. Following bacterial capture, gold nanoparticles functionalized with SERS tags (5,5-dithiobis-(2-nitrobenzoic acid), 4-mercaptobenzoic acid) and aptamers were used as signal target probes. The dual-recognition strategy resulted in sensitive (10 CFU mL−1), simultaneous detection of bacteria in milk, orange juice, chicken, and sputum samples. Gold nanoparticles were also combined with aptamer functionalized-macroporous magnetic silica photonic microsphere.149 This 3D substrate reported sensitive detection of E. coli, owing to the aptamer's specific capture and the formation of uniform hot-spots caused by the 3D substrate. Gold nanoparticles functionalized with 5,5′-dithiobis (2-nitrobenzoic acid) (DTNB) and a secondary aptamer were used as signal tags. Importantly, the sensor integrated sample pretreatment and sample separation, further shortening the turnaround time. Successful detection was reported for water, milk and minced pork samples.149
Antibodies and nanogapped, self-assembled SERS tags (Si@Au/Au) have been reported in literature for the simultaneous detection of S. pneumoniae and FluA H1N1 virus in clinical samples.150 In particular, following the formation of a silicon dioxide nanoparticle layer, continuous assembly of two layers of dense gold nanoparticles resulted in an ultranarrow intraparticle gap. The nanoparticles were functionalized via DTNB and 4-MBA probes, providing sites for anti S. pneumoniae and anti-H1N1 antibodies. The study reported bacterial detection (LOD 16 CFU mL−1) in aqueous and saliva samples.
An interesting study reported the use of silver-etched silver-coated gold ‘nanopancakes’ which were made as a dual-function SERS substrate and as a reservoir for silver ions.104 The nanopancakes were decorated with a Prussian blue analogue, capable of binding with bacteria through a cyano group. Bacteria were first tagged with 4-mercaptophenyl boronic acid (4-MPBA). Upon mixing, the nanopancakes provided intrinsic SERS signal from the bacteria as well as from the 4-MPBA (which served as an internal standard). The study reported detection and antibacterial activity against E. coli, S. aureus, and P. aeruginosa, with a low LOD in aqueous samples, blood, and biofilms.
Magnetic sample enrichment in microfluidic devices has also been reported. Antibody-modified MNPs were used to selectively capture E. coli from complex milk samples in a capillary-driven microfluidic device. SERS was subsequently detected using 4-aminothiophenol (4-ATP)-labelled gold nanorods as SERS probes.151 A microfluidic platform was reported for the simultaneous detection of S. aureus and E. coli O157:H7 which utilized MNPs functionalized with antimicrobial peptides to selectively capture the bacteria, followed by SERS interrogation using bacteria-specific aptamer-functionalized SERS nanoparticles with peaks in the spectral silent region (around 2100 cm−1). In this case, performing both the magnetic enrichment and the SERS detection in a microfluidic chip provided flexibility in rapidly switching between magnetic enrichment and elution of the samples.152
Urchin-shaped, platinum layered nanoparticles based on a core of gold and silver nanoparticles have been reportedly used for the multimodal detection of E. coli, S. aureus, and P. aeruginosa in aqueous and human blood samples, using a lateral flow setup.106 4-MBPA was used as an antibody analogue, reporting high affinity for the bacteria used in the study. Interestingly, this study used a triple colorimetric-SERS-photothermal-catalytic method, reporting simultaneous bacterial detection.
A variety of SERS nanoparticles have been used in microfluidic devices to enable the sensitive and selective detection of bacteria. SERS-tagged gold nanostars have been functionalized with a monoclonal antibody to rapidly detect Listeria monocytogenes bacteria on-chip with high specificity.153 Gold nanoparticles with different Raman reporter molecules were used to selectively detect E. coli O157:H7 and Salmonella enterica. The nanostars were functionalised with either fluorescein or rhodamine 6G (R6G) and streptavidin (SA), followed by tagging with species-specific antibodies for targeting, resulting in two distinct and specific SERS-nanotags.154 The system was successfully demonstrated for E. coli detection in food samples.155 Finally, SERS detection can be further combined on-chip with tools such as CRISPR/Cas12a genome editing, broadening the scope of detection that is currently achievable. Gold nanostar@4-mercaptobenzoic acid@goldnanoshell structures (AuNS@4-MBA@Au@DNA) conjugated with thiolated ssDNA were used in one such microfluidic platform for the SERS detection of Salmonella typhimurium in food samples.156
4. Electrochemical SERS
Recently, there has been increasing interest in the combination of electrochemistry with SERS. EC sensors are often held back by their lack of specificity and SERS often faces challenges in the acquisition of reproducible signals. EC-SERS aims to enhance the Raman signal by applying controlled electrode polarization on the SERS active electrode.157 This approach can allow the detection of the target molecule at a desired applied voltage as well as the formation of a biologically relevant electric field. This can provide insights on the rotational configuration of molecules under certain voltages, as well as the electrochemical stability of analytes.158
4.1 Introduction to EC-SERS
Electrochemistry and SERS have always been intertwined, as the first SERS spectra were obtained from an electrochemical cell in the 1970s.96 The signal enhancement of SERS is attributed to two mechanisms: electromagnetic (plasmonic) and chemical enhancement; both can be modulated by the application of electric potential.159,160 The aim of EC-SERS is to measure Raman (SERS) spectra of adsorbed species on the surface of the electrode, while also measuring any changes caused by varying the applied voltage. Changes in the voltage can lead to species adsorption/desorption, reorientation and composition of the electrochemical double layer.12,161 The combination of these effects leads to improved analytical performance, which has found applications in molecular identification and characterization as well as detection of low-affinity species, including drugs, metabolites and other chemicals.162 Published literature has reported that the combination of electrochemistry and SERS leads to an average 10-fold improvement in signal enhancement compared to SERS alone; the varying voltage can promote analyte adsorption to the SERS substrate boosting the distance-dependent electromagnetic enhancement, but also facilitate charge transfer mechanisms related to the chemical enhancement.163
An EC-SERS setup consists of a Raman spectrometer, a potentiostat, and an electrochemical cell with the working electrode used as a SERS substrate.157 Working electrode materials are similar to those used in electrochemical applications and most commonly include gold, silver, platinum and glassy carbon, as their plasmonic properties allow generation of SERS signal.157 Screen-printed electrodes have gained popularity as EC-SERS substrates given their low cost, reliability, efficiency and simplicity, and are utilized by the majority of published work on EC-SERS.164 Furthermore, advances in nanomaterials have led to the development of novel and improved SERS substrates, such as bimetal complexes and nanowires, nanoislands, and nanopancakes.87,93–95,148
In the past three years, there has been an increasing number of published work describing the use of EC-SERS for the detection of plasticizers (phthalate esters and bisphenol A),165,166 insecticides and pesticides,6,164,167–172 drugs,173–183 pigments,184 nucleic acids,185 viral lysates,186 endotoxin,187 protein markers,188 and even whole bacteria.189
Fig. 4 shows examples of EC-SERS applications for bacterial detection, categorized as direct bacterial detection (a), detection of bacterial metabolites (b), and detection of bacterial-derived DNA (c).
 |
| Fig. 4 Examples of EC-SERS applications for bacterial detection. (a) EC-SERS label-free detection of E. coli and P. syringiae.189,190 (b) EC-SERS detection of bacterial metabolites.163,191 (c) EC-SERS detection of bacteria-derived cDNA, from E. coli and E. faecium.192 | |
4.2 EC-SERS for bacterial detection
A relatively recent EC-SERS application is in the more sensitive detection of bacteria and their metabolites. Since 2018, when the first EC-SERS application for a bacterial detection was described by Lynk et al., numerous other published works have described EC-SERS applications for various bacteria species and metabolites, using various EC techniques. Commonly used EC techniques applied with SERS include voltage stepping, cyclic voltammetry (CV) and differential pulse voltammetry (DPV), all of which can be used in electrochemical sensors by themselves.163,190 Table 1 lists examples of detection of bacteria and their metabolites using EC-SERS.
Table 1 Bacteria and bacterial metabolites detection using EC-SERS
Bacterial species/analyte |
Label |
Substrate |
Wavelength |
EC |
Sample type |
Ref. |
Whole bacteria |
Label free |
|
|
|
|
|
E. coli, B. megaterium, Bacterial metabolites |
|
AgNPs and SPEs |
780 nm |
Applied voltage |
Aqueous solutions |
163 |
B. megaterium, E. coli |
|
AgNPs and SPEs |
780 nm |
Applied voltage |
Aqueous solutions (pbs) |
190 |
Pseudomonas syringae pv. actinidiae |
|
AgNPs with calcium and iodide ions |
633 nm |
Enhancement at −0.8 V, DPV, CV |
Aqueous solutions |
189 |
Whole bacteria |
Labeled |
|
|
|
|
|
V. vulnificus |
Detection antibodies |
Mxene, rhodamine 6, Au nanorods |
780 nm |
CV |
Aqueous solutions |
203 |
Mycobacterium tuberculosis |
Antibodies |
AgNPs, carbon SPE |
785 nm |
Voltage stepping |
Aqueous, urine, sputum |
201 |
Bacterial metabolites |
Label free |
|
|
|
|
|
Uric acid |
|
Au/Ag substrate layers |
532 nm |
Applied voltage |
Aqueous solutions |
198 |
Pyocyanin from Pseudomonas |
|
AgNPs and SPEs |
532 nm |
CV |
Aqueous solutions |
191 |
Uric acid |
|
Ag SPEs |
785 nm |
CV |
Synthetic urine |
199 |
Uric acid |
|
Polycarbonate, Au nanospheres |
632 nm |
Applied voltage |
Aqueous solutions, simulated urine |
200 |
Pseudomonas metabolites |
|
AuNPS block copolymer (BCP) membrane, nanopore electrode arrays (NEAs) |
785 nm |
CV, SWV, DPV |
Aqueous solutions |
195 |
DPA from Bacillus spores |
|
AgNPs/cuttlebone-derived organic matrix |
785 nm |
Applied voltage |
Water, milk |
194 |
Bacterial metabolites & DNA |
Labeled |
|
|
|
|
|
Guanine, adenine as bacterial metabolites |
Aptamer |
SPEs, AgNPs |
785 nm |
CV |
Aqueous solutions |
158 |
DNA from E. faecium and S. aureus |
Capture DNA with cyanine 5 |
Ag nanopillars |
785 nm |
CV EC deposition of Au on metallic nanopillars |
H2O, 100× diluted blood |
192 |
Endotoxin |
Aptamer |
Au@Ag nanocubes |
785 nm |
CV, EIS |
Aqueous solutions |
187 |
Lynk et al., in their 2018 publication, reported the use of screen-printed electrodes, on which three layers of silver nanoparticle paste were deposited by drop-coating.163 Following a rinse with 0.5 M KCl, the electrode was incubated with a bacterial suspension for 16 hours, before being rinsed and dried prior to use. Firstly, spectra for bacterial metabolites and nucleotide breakdown products (adenine, hypoxanthine, xanthine, guanine, uric acid, AMP, and guanosine) were obtained, by applying voltage in −0.1 V steps, until a final potential of −1.0 V was reached. Next, E. coli and B. megaterium were assessed, using the same method and voltage application. In both cases (metabolites and bacteria), the application of voltage resulted in clearer spectra with more well-defined and more intense peaks compared to the control spectra, highlighting the enhancement provided by electrochemistry.
The same group, in a subsequent study published in 2019, further examined the effect of EC-SERS on bacterial detection.190 Using the same protocol (screen printed electrodes (SPEs) and silver nanoparticles), the study further investigated the initial findings and reported that the electrochemical enhancement was caused by the release of bacterial metabolites from both species (E. coli and B. megaterium), such as adenine, xanthine, hypoxanthine, guanine, guanosine, adenosine monophosphate and uric acid, all of which dominate the SERS features of bacteria. Such metabolites are released by bacteria upon exposure to environmental stressors, and due to their affinity for the SERS substrate, result in spectral enhancement.
The same year, EC-SERS was employed for the detection of pyocyanin, a blue-coloured characteristic metabolite produced by P. aeruginosa, which is heavily involved in pathogenesis.191 As pyocyanin is known for being both redox-active and Raman-active, it was anticipated that EC-SERS would provide valuable information. Overall, it was reported that following application of external potential, reduced pyocyanin generated weaker spectra compared to its oxidized form. In addition, the sensor reported the early detection of pyocyanin in low concentrations, which could enable rapid Pseudomonas detection.
The combination of electric field and SERS as an EC-SERS system was explored using gold nano-islands (20 nm) which were deposited on a 3D laser-scribed graphene (LSG) substrate to increase the sensitivity and reproducibility of SERS detection.193 The authors amplified the SERS signal (633 nm laser; 10 seconds integration time) by applying an electric-field stimulus to the Au-LSG SERS substrate and measured the electric-field-induced chemical enhancement (CE) SERS signals of uremic toxins (urea and creatine) for the determination of chronic kidney diseases. The detected molecules displayed a 4-fold increase in Raman signal at low voltage potentials (−0.2 V), demonstrating the capability of enhancing the SERS signal and its sensitivity.
A recent paper describes the use of applied voltage, DPV and CV for the label-free detection of Pseudomonas syringae pv. actinidiae (Psa), a kiwifruit pathogen responsible for economic losses worldwide.189 In this work, silver nanoparticles with iodide and calcium ions on a nanostructured silver electrode were used as a SERS substrate. Following preparation of the bacterial culture and resuspension in ultrapure water, the modified electrode was immersed into the culture for 16 hours, followed by air drying and spectral acquisition. Following a screening process, it was reported that application of −0.8 V led to a significant enhancement in SERS spectra, possible caused by potential-induced reorientation of Psa molecules on the electrode surface.
Additionally, EC-SERS can also be used for the detection of bacterial spores. The use of silver nanoparticles in a cuttlebone-derived organic matrix was reported. Following electrochemical enhancement through the application of 2 V, the sensor reported sensitive detection of Bacillus cereus spores in water and milk samples, with LOD of 0.3 nM.194
Novel SERS composites can be exploited for the detection of bacterial metabolites, with one such example describing the use of a pH-responsive block copolymer membrane with nanopore electrode arrays filled with gold nanoparticles, capable of isolating and detecting phenazine-1-carboxylic acid, an important Pseudonomas metabolite, in bacterial cultures.195 The pH-responsive membrane ensured the transfer of the negatively-charged analyte in acidic culture medium, across its membrane and into the gold-nanoparticle filled nanopores. Square-wave voltammetry performed from −0.8 to +0.2 V on different Pseudomonas phenazines revealed that each species generated distinct spectra, aiding their characterization.
A common bacterial metabolite is uric acid, often produced by aerobic bacteria as part of the classical metabolic pathway of purines.196 Some bacteria can break down uric acid, such as Bacillus spp., P. aeruginosa, K. pneumoniae and Lactobacillus spp.197 There have been reports of label-free EC-SERS sensors in literature designed for the detection of uric acid. One such sensor describes the use of a multilayered gold and silver substrate, constructed by depositing monodisperse gold and silver nanoparticles on the carbon working electrode surface of screen printed electrodes.198 By applying voltage in a stepwise approach, it was realized that application of −0.9 V resulted in the best Raman spectral enhancement in aqueous samples. Another similar study described the use of silver, screen-printed electrodes, combined with cyclic voltammetry between −0.40 V and +0.40 V, reporting uric acid detection even in complex samples, such as synthetic urine.199 Finally, a more recent study reported the use of a nanostructured, polycarbonate nanocone array, decorated with a tightly packed array of gold nanoparticles.200 Voltage application at −1.0 V was reported to assist uric acid detection, in both aqueous and synthetic urine samples.
4.2.1. Labelled EC-SERS applications. Labelled EC-SERS is considered an indirect detection approach. Commonly used EC-SERS labels can include antibodies, aptamers and peptides.A point of care sensor utilizing disposable carbon screen printed electrodes combined with silver nanoparticles with a polyethylene coating has been developed for the detection of Mycobacterium tuberculosis. M. tuberculosis is the causative pathogen of tuberculosis, infects around one third of the global population and is a leading bacterium-based infectious disease-causing pathogen worldwide.201 Disposable carbon screen printed electrodes were combined with silver nanoparticles with a polyethylene coating, which attracted bacteria, anti-M. tuberculosis antibodies and a customized voltage-stepping sequence-based protocol. The application of both anodic and cathodic voltage in sequence was able to specifically detect and identify mycobacteria, including three clinically important MTB strains. Additionally, the sensor differentiated between mycobacteria and Gram +ve and −ve bacteria (E. coli and S. aureus). The study reported M. tuberculosis detection in aqueous samples, as well as in urine and sputum samples, with little to no sample interference.
Bacterial DNA can be sensitively detected using a combination of electrochemistry and SERS. Electrochemical deposition was performed to deposit gold onto nanopillars, containing a plethora of hot-spots, on which DNA strands complementary to the target DNA were attached.192 The complementary DNA strands were labelled with Cyanine 5, a Raman active dye. During the EC-SERS measurement, a potential of +0.3 V was applied to the working electrode for five minutes, and the spectra were obtained with an acquisition time of 30 seconds. CV was also performed to assess complementarity and nonspecific binding of DNAs. The sensor reported sensitive detection of DNA for both species with LOD of 0.05 nM in aqueous solutions, high specificity and was even able to detect DNA with the same sensitivity in 100× diluted human whole blood. Aptamers can be combined with EC-SERS for the labelled detection of bacterial DNA hybridization and detection of individual nucleotides. One example reports the use of DNA probes immobilized on silver nanoparticles arranged on a screen-printed electrode.158 The hybridization process was monitored by measurement of adenine, guanine, cytosine, and thymine. Voltage application was overall found to enhance the SERS signal for each nucleotide assessed.
4.3 Other uses of electric fields for SERS
4.3.1. Electrochemiluminescence. ECL is closely related to electrochemistry; the applied voltage drives a redox reaction between a luminophore and a sacrificial co-reactant resulting in the emission of light, which in turn serves as the sensor signal.202 ECL has been used along with SERS for bacterial detection. Specifically, MXene and antibodies were combined, reporting detection of V. vulnificus, an especially high-risk Vibrio strain, responsible for gastroenteritis, sepsis, necrotizing wound infection and death.203 The sensor achieved sensitive bacterial detection in physiological and seawater samples. Although the ECL and SERS measurements were performed separately, this work demonstrates that the two techniques are closely intertwined, allowing cross-verification of results between the two methods. A recent, novel labelled approach reported the use of dual-mode ECL and SERS sensing of bacterial endotoxin, one of main components of Gram −ve bacterial cell walls, which functions as a major mediator for sepsis in the human body.187 In particular, the study reported a complex, adjustable cascade where the binding of endotoxin to the sensor initiated rolling-cycle amplification, causing the release of large amounts of poly-C DNA, which gradually etched the silver shell of silver-coated, gold core nanocubes, through removal of Ag+ ions. Subsequently, the poly-C DNA was reduced to silver nanoclusters. This reaction resulted in the loss of dye-signal and the disappearance of hot-spots, resulting in a noticeable, quantifiable reduction in SERS signal. The study reported sensitive detection of endotoxin in aqueous samples, as well as dilute milk and serum samples.
4.3.2. Field-effect transistors. As electric potentials can modulate SERS on EC electrodes, a similar approach can be pursued by using field-effect transistors (FETs). The combination of an electric field and SERS provides a unique approach to modulate the electronic state of the SERS substrate and enhance the Raman effect by controlling the electric field around the hot-spot nanostructures and any adjoining molecules. Tuning the applied potential of the metallic electrode contacts in the electric field-stimulated Raman device increases plasmonic enhancement and facilitates electron transfer to the molecule. The combination of electric field-stimulation and SERS has not been fully explored concurrently, with most studies employing FET structures as sensing platforms for SERS, serving as a supplemental detection technique for analytes.204–206A recent study reported an electrically tunable back-gate FET on silica/silicon substrates with pre-patterned Au/Ti electrodes comprising a tungsten/molybdenum oxide hybrid as the SERS active layer deposited via spin-coating (Fig. 5a).207 The study reported that the applied gate bias (from 0 V to +15 V) tunes the SERS enhancement factor and Raman intensity (Fig. 5b). The SERS performance was evaluated using rhodamine 6G as the Raman reporter, excited by 532 nm and 785 nm lasers, exhibiting a LOD as low as 10−8 M. According to the authors, when a positive voltage is applied to the gate electrode, the band bends downwards, and electrons from the source/drain electrodes accumulate near the interface between the hybrid semiconductor and the gate dielectric. This enhances the electron density near the Fermi level, ensuring that more electrons can be transferred to the molecular orbital. The reported study provides a substantial contribution to the understanding of the fundamental properties of FET-SERS mechanism using metal oxide semiconductors.
 |
| Fig. 5 (a) Optical image, 3D schematic structure, and cross-sectional SEM image of a WOx/MoOx-based FET-SERS substrate. (b) SERS spectra of 10−4 M R6G adsorbed on the WOx/MoOx-based SERS substrate with applied voltages from 0 to +15 V.207 | |
The limited number of studies highlights the research gap in electric field-stimulated SERS and its potential for analyte detection. Moreover, the FET-SERS platform shows potential for bacterial detection, presenting a flexible method for quickly and accurately identifying harmful microorganisms.
4.3.3. Electrokinetics. Following a different approach, electric fields shaped by microelectrodes can be used alongside SERS. For example, dielectrophoresis (DEP) has been reported to concentrate or isolate bacteria, facilitating detection. Specifically, negative DEP has been combined with SERS, reporting rapid concentration, isolation and label-free identification of E. coli.208 The SERS substrate was composed of poly(ethylene terephthalate) polymer foil, a layer of indium tin oxide subjected to dielectric barrier discharge and a layer of silver. DEP assisted the separation, concentration and controlled deposition of bacteria on the SERS-active area of the substrate. Longer DEP times (180 seconds) resulted in improved E. coli deposition and clearer SERS spectra. The study reported bacterial detection in aqueous solution, apple juice and urine samples.
5. Discussion
It is evident that bacterial detection remains a hot area of research. Miniaturized and rapid sensors are necessary to supplant the currently used lab-based protocols. Electrochemistry, SERS, and their combination, all exhibit potential towards this goal. These methods have been employed for bacterial detection, identification, and quantitation, in a variety of different settings.
To produce useful and relevant sensors, however, a few key parameters must be kept in mind: timeliness, ease of use, sensitivity and quantitation.
5.1. Time is of the essence when dealing with bacterial infections in a healthcare setting
A major limitation of bacterial detection is the reliance on bacterial cultures. Most detection methods require bacterial cultures either to amplify the numbers of bacteria above the detection threshold, or to ensure the bacteria are in a medium suitable for the sensor. Cultures take time to grow (typically 24 hours or more) taking a toll on the benefits these new technologies provide. For the deployment of a useful sensor, reliance on cultures should be eliminated. Methods such as SERS may be able to detect low abundance of bacteria, obviating the need for time consuming cultures.
5.2. Any commercial application of sensors requires ease of use
Electrochemistry is an established technique that may be conducive to miniaturization and automation. The equipment needed is cost-effective and lightweight. However, most EC sensors found in the literature require several steps of sample handling and reagents, making their field deployment cumbersome. SERS may be simpler to perform compared to EC, especially when using non-colloidal substrates, but data analysis is more complex, and the equipment needed a little more costly. One key technology that can greatly expedite and support the integration of EC-SERS and its related methodologies into commercial applications is microfluidics. Microfluidics can miniaturize and parallelize the analysis, performing several assays with a small volume sample. Tedious steps such as sample washes, serial reagent additions, and nanoparticle mixing can be automated. Additionally, microfluidic devices can integrate other mechanisms to facilitate detection, such as FETs, or electrokinetic traps for sample enrichment. Magnetic nanoparticles can be used, offering yet another modality for detection, complementing the electrical and optical signals. For example, a platform based on droplet microfluidics and magnetoresistive sensing has been employed for the selective detection of K. pneumoniae in urine.209
5.3. Sensitivity and quantitation are crucial for most applications
New sensor materials have allowed EC to push down its LOD, especially when secondary labels are used. EC sensors for bacteria have been reported in different contexts, demonstrating that EC detection is a robust strategy. Also, EC sensors are – by design – specific to only a single target, meaning that while they may be appropriate for dedicated applications, they will not detect other pathogens that may be present in the sample.
On the other hand, SERS boasts extremely low limits of detection, revealing low abundance targets. In the case of direct SERS, it offers the possibility of exploratory analysis, where multiple bacterial species may be detected with the same platform, making it more versatile and cost effective. However, making quantitative measurements using SERS is challenging; most studies provide qualitative results, such as yes/no detection or classification of bacterial species. This may be sufficient for applications such as sepsis or detection of Salmonella, where the presence of a single pathogen is indicative of infection, but it is limiting in applications such as UTI diagnosis or detection of Campylobacter spp in raw poultry, where quantitation is necessary.
The combination of the two methods as EC-SERS aims to leverage their strengths synergistically. The two techniques supplement each other, SERS providing the low limits of detection while EC ensures specificity.
One limitation in the field lies in the inconsistency of reporting of LODs. Most studies, regardless of the sensing methodology, report a theoretical LOD, based on extrapolation of measurements in a linear sensing region. Few studies report the last measurable concentration different than the blank.
Additionally, low target concentrations become increasingly harder to detect as the sample volume is decreased via miniaturization. I.e., low abundance targets will not be sufficiently represented in small sample volumes. There may be a compromise to be reached between miniaturization, sensitivity, and sample enrichment (through culture).
6. Conclusion
The combination of EC, SERS, and other related methods has the potential to produce reliable sensors, with application-relevant sensitivities, democratizing bacterial detection and saving lives and operational costs. To achieve this, it is necessary for researchers to have usability in mind, as the final product will need to be deployed very widely to make a meaningful impact.
While EC sensors are more mature, they are limited by the use of specific recognition moieties, limiting their generalizability. SERS, on the other hand, can be used to identify many bacterial species, but struggles to provide quantitative information. EC-SERS may be able to harness the strength of both, providing sufficient sensitivity, quantitation, and identification of bacterial species.
To date, the number of reports on EC-SERS in the literature remains limited. This may be due to the distinct skillsets required to employ the two techniques. The few examples demonstrated, however, seem to be more effective than ones using the parent techniques alone. Specifically, compared to SERS alone, EC-SERS is more likely to provide quantitative information; whereas the spectra acquired at different voltages contain a wealth of information that can be used to identify bacterial species with higher specificity than EC alone.
Moving forward, we recommend that the research community working in the bacterial diagnostics field be aware of these limitations. While the existing literature showcases the capabilities of the detection methodologies, there is limited discussion of application specific requirements. For example, regulatory requirements are different for different fields (e.g., health care and food safety) but also within the same fields (e.g., diagnostics of sepsis vs. UTIs). Conforming to such requirements is essential for any product's approval by the relevant regulatory authorities. This may be an important reason for the limited commercial success these methods have shown to date. Additionally, few studies report applications such as antimicrobial resistance, indicating a major gap in knowledge and potential area of research. One thing, however, remains constant: the need for rapid and cost-effective bacterial detection will only increase in the future.
Author contributions
Conceptualization: CA, KK; funding acquisition: CA; writing – original draft: KK, TL, MC; writing –review & editing: CA.
Conflicts of interest
There are no conflicts to declare.
Data availability
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Acknowledgements
This work received funding from the European Union's Horizon 2020 research and innovation program under the Marie Skłodowska-Curie grant agreement no. 101024362, and by the Republic of Cyprus “Restart 2016-2020 Programmes” through the Research and Innovation Foundation (CODEVELOP-AG-SH-HE/0823/0079).
References
- M. Naghavi, S. E. Vollset, K. S. Ikuta, L. R. Swetschinski, A. P. Gray, E. E. Wool, G. Robles Aguilar, T. Mestrovic, G. Smith, C. Han, R. L. Hsu, J. Chalek, D. T. Araki, E. Chung, C. Raggi, A. Gershberg Hayoon, N. Davis Weaver, P. A. Lindstedt, A. E. Smith, U. Altay, N. V. Bhattacharjee, K. Giannakis, F. Fell, B. McManigal, N. Ekapirat, J. A. Mendes, T. Runghien, O. Srimokla, A. Abdelkader, S. Abd-Elsalam, R. G. Aboagye, H. Abolhassani, H. Abualruz, U. Abubakar, H. J. Abukhadijah, S. Aburuz, A. Abu-Zaid, S. Achalapong, I. Y. Addo, V. Adekanmbi, T. E. Adeyeoluwa, Q. E. S. Adnani, L. A. Adzigbli, M. S. Afzal, S. Afzal, A. Agodi, A. J. Ahlstrom, A. Ahmad, S. Ahmad, T. Ahmad, A. Ahmadi, A. Ahmed, H. Ahmed, I. Ahmed, M. Ahmed, S. Ahmed, S. A. Ahmed, M. A. Akkaif, S. Al Awaidy, Y. Al Thaher, S. O. Alalalmeh, M. T. AlBataineh, W. A. Aldhaleei, A. A. S. Al-Gheethi, N. B. Alhaji, A. Ali, L. Ali, S. S. Ali, W. Ali, K. Allel, S. Al-Marwani, A. Alrawashdeh, A. Altaf, A. B. Al-Tammemi, J. A. Al-Tawfiq, K. H. Alzoubi, W. A. Al-Zyoud, B. Amos, J. H. Amuasi, R. Ancuceanu, J. R. Andrews, A. Anil, I. A. Anuoluwa, S. Anvari, A. E. Anyasodor, G. L. C. Apostol, J. Arabloo, M. Arafat, A. Y. Aravkin, D. Areda, A. Aremu, A. A. Artamonov, E. A. Ashley, M. O. Asika, S. S. Athari, M. M. D. W. Atout, T. Awoke, S. Azadnajafabad, J. M. Azam, S. Aziz, A. Y. Azzam, M. Babaei, F.-X. Babin, M. Badar, A. A. Baig, M. Bajcetic, S. Baker, M. Bardhan, H. J. Barqawi, Z. Basharat, A. Basiru, M. Bastard, S. Basu, N. S. Bayleyegn, M. A. Belete, O. O. Bello, A. Beloukas, J. A. Berkley, A. S. Bhagavathula, S. Bhaskar, S. S. Bhuyan, J. A. Bielicki, N. I. Briko, C. S. Brown, A. J. Browne, D. Buonsenso, Y. Bustanji, C. G. Carvalheiro, C. A. Castañeda-Orjuela, M. Cenderadewi, J. Chadwick, S. Chakraborty, R. M. Chandika, S. Chandy, V. Chansamouth, V. K. Chattu, A. A. Chaudhary, P. R. Ching, H. Chopra, F. R. Chowdhury, D.-T. Chu, M. Chutiyami, N. Cruz-Martins, A. G. da Silva, O. Dadras, X. Dai, S. D. Darcho, S. Das, F. P. De la Hoz, D. M. Dekker, K. Dhama, D. Diaz, B. F. R. Dickson, S. G. Djorie, M. Dodangeh, S. Dohare, K. G. Dokova, O. P. Doshi, R. K. Dowou, H. L. Dsouza, S. J. Dunachie, A. M. Dziedzic, T. Eckmanns, A. Ed-Dra, A. Eftekharimehrabad, T. C. Ekundayo, I. El Sayed, M. Elhadi, W. El-Huneidi, C. Elias, S. J. Ellis, R. Elsheikh, I. Elsohaby, C. Eltaha, B. Eshrati, M. Eslami, D. W. Eyre, A. O. Fadaka, A. F. Fagbamigbe, A. Fahim, A. Fakhri-Demeshghieh, F. O. Fasina, M. M. Fasina, A. Fatehizadeh, N. A. Feasey, A. Feizkhah, G. Fekadu, F. Fischer, I. Fitriana, K. M. Forrest, C. Fortuna Rodrigues, J. E. Fuller, M. A. Gadanya, M. Gajdács, A. P. Gandhi, E. E. Garcia-Gallo, D. O. Garrett, R. K. Gautam, M. W. Gebregergis, M. Gebrehiwot, T. G. Gebremeskel, C. Geffers, L. Georgalis, R. M. Ghazy, M. Golechha, D. Golinelli, M. Gordon, S. Gulati, R. D. Gupta, S. Gupta, V. K. Gupta, A. D. Habteyohannes, S. Haller, H. Harapan, M. L. Harrison, A. I. Hasaballah, I. Hasan, R. S. Hasan, H. Hasani, A. H. Haselbeck, M. S. Hasnain, I. I. Hassan, S. Hassan, M. S. Hassan Zadeh Tabatabaei, K. Hayat, J. He, O. E. Hegazi, M. Heidari, K. Hezam, R. Holla, M. Holm, H. Hopkins, M. M. Hossain, M. Hosseinzadeh, S. Hostiuc, N. R. Hussein, L. D. Huy, E. D. Ibáñez-Prada, A. Ikiroma, I. M. Ilic, S. M. S. Islam, F. Ismail, N. E. Ismail, C. D. Iwu, C. J. Iwu-Jaja, A. Jafarzadeh, F. Jaiteh, R. Jalilzadeh Yengejeh, R. D. G. Jamora, J. Javidnia, T. Jawaid, A. W. J. Jenney, H. J. Jeon, M. Jokar, N. Jomehzadeh, T. Joo, N. Joseph, Z. Kamal, K. K. Kanmodi, R. S. Kantar, J. A. Kapisi, I. M. Karaye, Y. S. Khader, H. Khajuria, N. Khalid, F. Khamesipour, A. Khan, M. J. Khan, M. T. Khan, V. Khanal, F. F. Khidri, J. Khubchandani, S. Khusuwan, M. S. Kim, A. Kisa, V. A. Korshunov, F. Krapp, R. Krumkamp, M. Kuddus, M. Kulimbet, D. Kumar, E. A. P. Kumaran, A. Kuttikkattu, H. H. Kyu, I. Landires, B. K. Lawal, T. T. T. Le, I. M. Lederer, M. Lee, S. W. Lee, A. Lepape, T. L. Lerango, V. S. Ligade, C. Lim, S. S. Lim, L. W. Limenh, C. Liu, X. Liu, X. Liu, M. J. Loftus, H. I. M Amin, K. L. Maass, S. B. Maharaj, M. A. Mahmoud, P. Maikanti-Charalampous, O. M. Makram, K. Malhotra, A. A. Malik, G. D. Mandilara, F. Marks, B. A. Martinez-Guerra, M. Martorell, H. Masoumi-Asl, A. G. Mathioudakis, J. May, T. A. McHugh, J. Meiring, H. N. Meles, A. Melese, E. B. Melese, G. Minervini, N. S. Mohamed, S. Mohammed, S. Mohan, A. H. Mokdad, L. Monasta, A. Moodi Ghalibaf, C. E. Moore, Y. Moradi, E. Mossialos, V. Mougin, G. D. Mukoro, F. Mulita, B. Muller-Pebody, E. Murillo-Zamora, S. Musa, P. Musicha, L. A. Musila, S. Muthupandian, A. J. Nagarajan, P. Naghavi, F. Nainu, T. S. Nair, H. H. R. Najmuldeen, Z. S. Natto, J. Nauman, B. P. Nayak, G. T. Nchanji, P. Ndishimye, I. Negoi, R. I. Negoi, S. A. Nejadghaderi, Q. P. Nguyen, E. A. Noman, D. C. Nwakanma, S. O'Brien, T. J. Ochoa, I. A. Odetokun, O. A. Ogundijo, T. R. Ojo-Akosile, S. R. Okeke, O. C. Okonji, A. T. Olagunju, A. Olivas-Martinez, A. A. Olorukooba, P. Olwoch, K. I. Onyedibe, E. Ortiz-Brizuela, O. Osuolale, P. Ounchanum, O. T. Oyeyemi, P. A. Mahesh Padukudru, J. L. Paredes, R. R. Parikh, J. Patel, S. Patil, S. Pawar, A. Y. Peleg, P. Peprah, J. Perdigão, C. Perrone, I.-R. Petcu, K. Phommasone, Z. Z. Piracha, D. Poddighe, A. J. Pollard, R. Poluru, A. Ponce-De-Leon, J. Puvvula, F. N. Qamar, N. H. Qasim, C. D. Rafai, P. Raghav, L. Rahbarnia, F. Rahim, V. Rahimi-Movaghar, M. Rahman, M. A. Rahman, H. Ramadan, S. K. Ramasamy, P. S. Ramesh, P. W. Ramteke, R. K. Rana, U. Rani, M.-M. Rashidi, D. Rathish, S. Rattanavong, S. Rawaf, E. M. M. Redwan, L. F. Reyes, T. Roberts, J. V. Robotham, V. D. Rosenthal, A. G. Ross, N. Roy, K. E. Rudd, C. J. Sabet, B. A. Saddik, M. R. Saeb, U. Saeed, S. Saeedi Moghaddam, W. Saengchan, M. Safaei, A. Saghazadeh, N. Saheb Sharif-Askari, A. Sahebkar, S. S. Sahoo, M. Sahu, M. Saki, N. Salam, Z. Saleem, M. A. Saleh, Y. L. Samodra, A. M. Samy, A. Saravanan, M. Satpathy, A. E. Schumacher, M. Sedighi, S. Seekaew, M. Shafie, P. A. Shah, S. Shahid, M. J. Shahwan, S. Shakoor, N. Shalev, M. A. Shamim, M. A. Shamshirgaran, A. Shamsi, A. Sharifan, R. P. Shastry, M. Shetty, A. Shittu, S. Shrestha, E. E. Siddig, T. Sideroglou, J. Sifuentes-Osornio, L. M. L. R. Silva, E. A. F. Simões, A. J. H. Simpson, A. Singh, S. Singh, R. Sinto, S. S. M. Soliman, S. Soraneh, N. Stoesser, T. Z. Stoeva, C. K. Swain, L. Szarpak, T. Y. Sree Sudha, S. Tabatabai, C. Tabche, Z. M.-A. Taha, K.-K. Tan, N. Tasak, N. Y. Tat, A. Thaiprakong, P. Thangaraju, C. C. Tigoi, K. Tiwari, M. R. Tovani-Palone, T. H. Tran, M. Tumurkhuu, P. Turner, A. J. Udoakang, A. Udoh, N. Ullah, S. Ullah, A. G. Vaithinathan, M. Valenti, T. Vos, H. T. L. Vu, Y. Waheed, A. S. Walker, J. L. Walson, T. Wangrangsimakul, K. G. Weerakoon, H. F. L. Wertheim, P. C. M. Williams, A. A. Wolde, T. M. Wozniak, F. Wu, Z. Wu, M. K. K. Yadav, S. Yaghoubi, Z. S. Yahaya, A. Yarahmadi, S. Yezli, Y. E. Yismaw, D. K. Yon, C.-W. Yuan, H. Yusuf, F. Zakham, G. Zamagni, H. Zhang, Z.-J. Zhang, M. Zielińska, A. Zumla, S. E. H. H. Zyoud, S. H. Zyoud, S. I. Hay, A. Stergachis, B. Sartorius, B. S. Cooper, C. Dolecek and C. J. L. Murray, Global burden of bacterial antimicrobial resistance 1990–2021: a systematic analysis with forecasts to 2050, Lancet, 2024, 404(10459), 1199–1226 CrossRef PubMed.
- C. Willyard, The drug-resistant bacteria that pose the greatest health threats, Nature, 2017, 543(7643), 15 CrossRef CAS PubMed.
- J. M. Lorenzo, P. E. Munekata, R. Dominguez, M. Pateiro, J. A. Saraiva and D. Franco, Chapter 3 - Main Groups of Microorganisms of Relevance for Food Safety and Stability: General Aspects and Overall Description, in Innovative Technologies for Food Preservation, ed. F. J. Barba, A. S. Sant'Ana, V. Orlien and M. Koubaa, Academic Press, 2018, pp. 53–107 Search PubMed.
- J. Chen, S. M. Andler, J. M. Goddard, S. R. Nugen and V. M. Rotello, Integrating recognition elements with nanomaterials for bacteria sensing, Chem. Soc. Rev., 2017, 46(5), 1272–1283 RSC.
- X. Zhou, Z. Hu, D. Yang, S. Xie, Z. Jiang, R. Niessner, C. Haisch, H. Zhou and P. Sun, Bacteria Detection: From Powerful SERS to Its Advanced Compatible Techniques, Adv. Sci., 2020, 7(23), 2001739 CrossRef CAS PubMed.
- R. Moldovan, K. Milenko, E. Vereshchagina, B.-C. Iacob, K. Schneider, C. Farcău and E. Bodoki, EC-SERS, detection of thiabendazole in apple juice using activated screen-printed electrodes, Food Chem., 2023, 405, 134713 CrossRef CAS PubMed.
- S. Ding, X. Chen, B. Yu and Z. Liu, Electrochemical biosensors for clinical detection of bacterial pathogens: advances, applications, and challenges, Chem. Commun., 2024, 60(71), 9513–9525 RSC.
- L. M. Castle, D. A. Schuh, E. E. Reynolds and A. L. Furst, Electrochemical Sensors to Detect Bacterial Foodborne Pathogens, ACS Sens., 2021, 6(5), 1717–1730 CrossRef CAS PubMed.
- M. A. Khaleque, S. I. Hossain, M. R. Ali, M. Aly Saad Aly, H. S. Abuelmakarem, M. S. Al Mamun and M. Z. Hossain Khan, Bioreceptor modified electrochemical biosensors for the detection of life threating pathogenic bacteria: a review, RSC Adv., 2024, 14(39), 28487–28515 RSC.
- D. Cialla-May, A. Bonifacio, T. Bocklitz, A. Markin, N. Markina, S. Fornasaro, A. Dwivedi, T. Dib, E. Farnesi, C. Liu, A. Ghosh and J. Popp, Biomedical SERS - the current state and future trends, Chem. Soc. Rev., 2024, 53(18), 8957–8979 RSC.
- M. Constantinou, K. Hadjigeorgiou, S. Abalde-Cela and C. Andreou, Label-Free Sensing with Metal Nanostructure-Based Surface-Enhanced Raman Spectroscopy for Cancer Diagnosis, ACS Appl. Nano Mater., 2022, 5(9), 12276–12299 CrossRef CAS PubMed.
- R. Moldovan, E. Vereshchagina, K. Milenko, B. C. Iacob, A. E. Bodoki, A. Falamas, N. Tosa, C. M. Muntean, C. Farcau and E. Bodoki, Review on combining surface-enhanced Raman spectroscopy and electrochemistry for analytical applications, Anal. Chim. Acta, 2022, 1209, 339250 CrossRef CAS PubMed.
- A. Fernández-la-Villa, D. F. Pozo-Ayuso and M. Castaño-Álvarez, Microfluidics and electrochemistry: an emerging tandem for next-generation analytical microsystems, Curr. Opin. Electrochem., 2019, 15, 175–185 CrossRef.
- C. Wang, G. Weng, J. Li, J. Zhu and J. Zhao, A review of SERS coupled microfluidic platforms: From configurations to applications, Anal. Chim. Acta, 2024, 1296, 342291 CrossRef CAS PubMed.
- Z. Altintas, M. Akgun, G. Kokturk and Y. Uludag, A fully automated microfluidic-based electrochemical sensor for real-time bacteria detection, Biosens. Bioelectron., 2018, 100, 541–548 CrossRef CAS PubMed.
- T. J. Beveridge, Use of the gram stain in microbiology, Biotech. Histochem., 2001, 76(3), 111–118 CrossRef CAS PubMed.
- M. N. Kokkinakis, G. A. Fragkiadakis, N. E. Lapidakis and A. N. Kokkinaki, Assessing microbiological quality of ready-to-eat prepacked sandwiches, in Crete, Greece, J. Food Sci. Technol., 2020, 57(11), 4220–4227 CrossRef CAS PubMed.
- I. Alexandraki and C. Palacio, Gram-negative versus Gram-positive bacteremia: what is more alarmin(g)?, Crit. Care, 2010, 14(3), 161 CrossRef PubMed.
- N. Woodford and D. M. Livermore, Infections caused by Gram-positive bacteria: a review of the global challenge, J Infect, 2009, 59(Suppl 1), S4–S16 CrossRef PubMed.
- D. A. Armbruster and T. Pry, Limit of blank, limit of detection and limit of quantitation, Clin. Biochem. Rev., 2008, 29(Suppl 1), S49–S52 Search PubMed.
- G. Schmiemann, E. Kniehl, K. Gebhardt, M. M. Matejczyk and E. Hummers-Pradier, The diagnosis of urinary tract infection: a systematic review, Dtsch. Arztebl. Int., 2010, 107(21), 361–367 Search PubMed.
- S. Puttaswamy, B. D. Lee and S. Sengupta, Novel electrical method for early detection of viable bacteria in blood cultures, J. Clin. Microbiol., 2011, 49(6), 2286–2289 CrossRef PubMed.
- J. L. Hammond, N. Formisano, P. Estrela, S. Carrara and J. Tkac, Electrochemical biosensors and nanobiosensors, Essays Biochem., 2016, 60(1), 69–80 CrossRef PubMed.
- D. Grieshaber, R. MacKenzie, J. Vörös and E. Reimhult, Electrochemical Biosensors - Sensor Principles and Architectures, Sensors, 2008, 8(3), 1400–1458 CrossRef CAS PubMed.
- N. J. Ronkainen, H. B. Halsall and W. R. Heineman, Electrochemical biosensors, Chem. Soc. Rev., 2010, 39(5), 1747–1763 RSC.
- A. Singh, A. Sharma, A. Ahmed, A. K. Sundramoorthy, H. Furukawa, S. Arya and A. Khosla, Recent Advances in Electrochemical Biosensors: Applications, Challenges, and Future Scope, Biosensors, 2021, 11(9), 336 CrossRef CAS PubMed.
- R. Cancelliere, E. Paialunga, A. Grattagliano and L. Micheli, Label-free electrochemical immunosensors: A practical guide, TrAC, Trends Anal. Chem., 2024, 180, 117949 CrossRef CAS.
- T. F. Hanifa Lestari, I. Irkham, U. Pratomo, S. Gaffar, S. N. Zakiyyah, I. Rahmawati, S. N. Topkaya and Y. W. Hartati, Label-free and label-based electrochemical detection of disease biomarker proteins, ADMET DMPK, 2024, 12(3), 463–486 Search PubMed.
- E. P. Randviir and C. E. Banks, A review of electrochemical impedance spectroscopy for bioanalytical sensors, Anal. Methods, 2022, 14(45), 4602–4624 RSC.
- M. Frigoli, M. P. Krupa, G. Hooyberghs, J. W. Lowdon, T. J. Cleij, H. Dilien, K. Eersels and B. van Grinsven, Electrochemical Sensors for Antibiotic Detection: A Focused Review with a Brief Overview of Commercial Technologies, Sensors, 2024, 24(17), 5576 CrossRef CAS PubMed.
- P. Zhang and R. Wang, Label-Free Biosensor, Biosensors, 2023, 556 CrossRef CAS PubMed.
- A. R. Khan, W. L. Hussain, H. C. Shum and S. U. Hassan, Point-of-care testing: a critical analysis of the market and future trends, Front. Lab Chip Technol., 2024, 3, 1394752 CrossRef.
- A. H. Dabhade, R. P. Verma, B. Paramasivan, A. Kumawat and B. Saha, Development of silver nanoparticles and aptamer conjugated biosensor for rapid detection of E. coli in a water sample, 3 Biotech, 2023, 13(7), 244 CrossRef PubMed.
- H. S. Magar, H. Abdelghany, M. N. Abbas, U. Bilitewski and R. Y. A. Hassan, Fast analysis of Staphylococcus aureus in food products using disposable label-free nano-electrochemical immunosensor chips, Microchem. J., 2023, 193, 109097 CrossRef CAS.
- V. Liustrovaite, M. Pogorielov, R. Boguzaite, V. Ratautaite, A. Ramanaviciene, G. Pilvenyte, V. Holubnycha, V. Korniienko, K. Diedkova, R. Viter and A. Ramanavicius, Towards Electrochemical Sensor Based on Molecularly Imprinted Polypyrrole for the Detection of Bacteria—Listeria monocytogenes, Polymers, 2023, 15(7), 1597 CrossRef CAS PubMed.
- B. Sequeira-Antunes and H. A. Ferreira, Nucleic Acid Aptamer-Based Biosensors: A Review, Biomedicines, 2023, 11(12), 3201 CrossRef CAS PubMed.
- S. Weaver, M. H. Mohammadi and N. Nakatsuka, Aptamer-functionalized capacitive biosensors, Biosens. Bioelectron., 2023, 224, 115014 CrossRef CAS PubMed.
- J. Pan, W. Xu, W. Li, S. Chen, Y. Dai, S. Yu, Q. Zhou and F. Xia, Electrochemical Aptamer-Based Sensors with Tunable Detection Range, Anal. Chem., 2023, 95(1), 420–432 CrossRef CAS PubMed.
- M. Jarczewska, Ł. Górski and E. Malinowska, Electrochemical aptamer-based biosensors as potential tools for clinical diagnostics, Anal. Methods, 2016, 8(19), 3861–3877 RSC.
- M. R. Rabiei, A. H. Rezayan, M. Hosseini and A. B. Pebdeni, Electrochemical Aptasensor for Selective and Sensitive Detection of Staphylococcus aureus Using Graphene oxide/poly Deep Eutectic Solvent/Nickel Oxide Nanocomposite Modified Electrode, ChemistrySelect, 2024, 9(3), e202302099 CrossRef CAS.
- M. Geravand, Y. Erfani, N. Nematpour, M. Khosravani, R. Rahimnia and M. Adabi, Nanoarchitectonics of aptamer-based electrochemical biosensor utilizing electerospun carbon nanofibers and gold nanoparticles for Acinetobacter baumannii detection, Microchem. J., 2024, 200, 110437 CrossRef CAS.
- M. Zhang, Y. Chen, S. G. Liu and X. Shi, Highly sensitive detection of L.monocytogenes using an electrochemical biosensor based on Si@MB/AuNPs modified glassy carbon electrode, Microchem. J., 2023, 194, 109357 CrossRef CAS.
- M. Zamzami, A. Ahmad, S. Alamoudi, H. Choudhry, S. Hosawi, G. Rabbani, E.-S. Shalaan and B. Arkook, A highly sensitive and specific Gold Electrode-Based electrochemical immunosensor for rapid On-Site detection of Salmonella enterica, Microchem. J., 2024, 199, 110190 CrossRef CAS.
- N. B. Messaoud, M. B. dos Santos, A. Vieira, A. Garrido-Maestu, B. Espiña and R. B. Queirós, A novel portable label-free electrochemical immunosensor for ultrasensitive detection of Aeromonas salmonicida in aquaculture seawater, Anal. Bioanal. Chem., 2022, 414(22), 6591–6600 CrossRef CAS PubMed.
- H.
A. de Godoy, A. M. Faria, N. A. V. Roza, L. Bach-Toledo, F. M. Simabuco, N. K. Scharlack, R. B. de Oliveira, A. E. C. Antunes, R. Arthur and T. Mazon, Point-of-Care Electrochemical Immunosensor Applied against Nosocomial Infection: Staphylococcus aureus Detection in Human Hand Skin, ACS Infect. Dis., 2024, 10(6), 1949–1957 CrossRef CAS PubMed.
- S. Li, Y. Zhang, M. Guo, Z. Yi, M. Hu, C. Xiong, G. Huang and J. Zhang, Rapid detection of Salmonella in milk by labeling-free electrochemical immunosensor based on an Fe3O4−ionic liquid-modified electrode, Talanta, 2024, 270, 125576 CrossRef CAS PubMed.
- K. Feng, T. Li, C. Ye, X. Gao, X. Yue, S. Ding, Q. Dong, M. Yang, G. Huang and J. Zhang, A novel electrochemical immunosensor based on Fe3O4@graphene nanocomposite modified glassy carbon electrode for rapid detection of Salmonella in milk, J. Dairy Sci., 2022, 105(3), 2108–2118 CrossRef CAS PubMed.
- K. Pandiyaraj, R. A. Elkaffas, M. I. H. Mohideen and S. Eissa, Graphene oxide/Cu–MOF-based electrochemical immunosensor for the simultaneous detection of Mycoplasma pneumoniae and Legionella pneumophila antigens in water, Sci. Rep., 2024, 14(1), 17172 CrossRef CAS PubMed.
- Z. Yi, Y. Zhang, X. Gao, S. Li, K. Li, C. Xiong, G. Huang and J. Zhang, Sensitive electrochemical immunosensor for rapid detection of Salmonella in milk using polydopamine/CoFe-MOFs@Nafion modified gold electrode, Int. J. Food Microbiol., 2024, 425, 110870 CrossRef CAS PubMed.
- M. R. Ali, M. S. Bacchu, S. Das, S. Akter, M. M. Rahman, M. Saad Aly and M. Z. H. Khan, Label free flexible electrochemical DNA biosensor for selective detection of Shigella flexneri in real food samples, Talanta, 2023, 253, 123909 CrossRef CAS PubMed.
- R. Ghasemi, S. Z. Mirahmadi-zare, A. Allafchian and M. Behmanesh, Fast fluorescent screening assay and dual electrochemical sensing of bacterial infection agent (Streptococcus agalactiae) based on a fluorescent-immune nanofibers, Sens. Actuators, B, 2022, 352, 130968 CrossRef CAS.
- P. Jolly, V. Tamboli, R. L. Harniman, P. Estrela, C. J. Allender and J. L. Bowen, Aptamer–MIP hybrid receptor for highly sensitive electrochemical detection of prostate specific antigen, Biosens. Bioelectron., 2016, 75, 188–195 CrossRef CAS PubMed.
- M. M. El-Wekil, H. M. Halby, M. Darweesh, M. E. Ali and R. Ali, An innovative dual recognition aptasensor for specific detection of Staphylococcus aureus based on Au/Fe3O4 binary hybrid, Sci. Rep., 2022, 12(1), 12502 CrossRef CAS PubMed.
- X. Jin, L. Gong, J. Liang, Z. Wang, K. Wang, T. Yang and H. Zeng, Polydopamine-enhanced vertically-ordered mesoporous silica film anti-fouling electrochemical aptasensor for indicator-free Vibrio parahaemolyticus discrimination using stable inherent Au signal, Sens. Actuators, B, 2024, 407, 135485 CrossRef CAS.
- N. Baskaran, R. Sakthivel, C. S. Karthik, Y.-C. Lin, X. Liu, H.-W. Wen, W. Yang and R.-J. Chung, Polydopamine-modified 3D flower-like ZnMoO4 integrated MXene-based label-free electrochemical immunosensor for the food-borne pathogen Listeria monocytogenes detection in milk and seafood, Talanta, 2025, 282, 127008 CrossRef CAS PubMed.
- S. Ghosh Dastider, A. Abdullah, I. Jasim, N. S. Yuksek, M. Dweik and M. Almasri, Low concentration E. coli O157:H7 bacteria sensing using microfluidic MEMS biosensor, Rev. Sci. Instrum., 2018, 89(12), 125009 CrossRef PubMed.
- I. Jasim, Z. Shen, Z. Mlaji, N. S. Yuksek, A. Abdullah, J. Liu, S. G. Dastider, M. El-Dweik, S. Zhang and M. Almasri, An impedance biosensor for simultaneous detection of low concentration of Salmonella serogroups in poultry and fresh produce samples, Biosens. Bioelectron., 2019, 126, 292–300 CrossRef CAS PubMed.
- Z. Babaie, G. Kibar, H. Yeşilkaya, Y. Amrani, S. Doğan, B. G. Tuna, V. C. Özalp and B. Çetin, Microfluidic rapid isolation and electrochemical detection of S. pneumonia via aptamer-decorated surfaces, Anal. Chim. Acta, 2025, 1345, 343726 CrossRef CAS PubMed.
- S. Fande, K. Amreen, D. Sriram and S. Goel, Microfluidic electrochemical device for real-time culturing and interference-free detection of Escherichia coli, Anal. Chim. Acta, 2023, 1237, 340591 CrossRef CAS PubMed.
- S. Gopalakrishnan, D. Mall, S. Pushpavanam and R. Karmakar, Rapid antimicrobial susceptibility testing using carbon screen printed electrodes in a microfluidic device, Sci. Rep., 2025, 15(1), 5133 CrossRef CAS PubMed.
- T. T.-Q. Nguyen and M. B. Gu, An ultrasensitive electrochemical aptasensor using Tyramide-assisted enzyme multiplication for the detection of Staphylococcus aureus, Biosens. Bioelectron., 2023, 228, 115199 CrossRef CAS PubMed.
- W. Li, X. Bai, F. Xiao, J. Huang, X. Zeng, Q. Xu, Y. Song, X. Xu and H. Xu, MXene@Au based electrochemical biosensor with pretreatment by magnetic nanoparticles for determination of MRSA from clinical samples, J. Hazard. Mater., 2023, 457, 131823 CrossRef CAS PubMed.
- S. Xiao, C. Cao, T. Ming, Y. Cao, Z. Yu and N. Gan, Simultaneous and rapid screening of live and dead E. coli O157:H7 with three signal outputs: An all-in-one biosensor using phage-apoferritin@CuO2 signal tags on MXenes-modified electrode platform, J. Hazard. Mater., 2023, 458, 131875 CrossRef CAS PubMed.
- X. Lin, P. P. Liu, J. Yan, D. Luan, T. Sun and X. Bian, Dual Synthetic Receptor-Based Sandwich Electrochemical Sensor for Highly Selective and Ultrasensitive Detection of Pathogenic Bacteria at the Single-Cell Level, Anal. Chem., 2023, 95(13), 5561–5567 CrossRef CAS PubMed.
- L. Huang, N. Yuan, W. Guo, Y. Zhang and W. Zhang, An electrochemical biosensor for the highly sensitive detection of Staphylococcus aureus based on SRCA-CRISPR/Cas12a, Talanta, 2023, 252, 123821 CrossRef CAS PubMed.
- H. Zhou, W. Guo, S. Wang, T. Hao, Z. Wang, Y. Hu, S. Wang, J. Xie, X. Jiang and Z. Guo, Electrochemical aptasensor for Staphylococcus aureus by stepwise signal amplification, Microchim. Acta, 2022, 189(9), 353 CrossRef CAS PubMed.
- J. Cui, Q. Luo, C. Wei, X. Deng, H. Liang, J. Wei, Y. Gong, Q. Tang, K. Zhang and X. Liao, Electrochemical biosensing for E.coli detection based on triple helix DNA inhibition of CRISPR/Cas12a cleavage activity, Anal. Chim. Acta, 2024, 1285, 342028 CrossRef CAS PubMed.
- S. Y. Jo, J. H. Shin and J. P. Park, CRISPR/Cas12a-mediated electrochemical biosensor for the sensitive detection of Vibrio parahaemolyticus, Biotechnol. Bioprocess Eng., 2025, 1–12 Search PubMed.
- C. Wang, Y. Zhang, S. Liu, Y. Yin, G.-C. Fan, Y. Shen, H. Han and W. Wang, Allosteric probe-triggered isothermal amplification to activate CRISPR/Cas12a for sensitive electrochemiluminescence detection of Salmonella, Food Chem., 2023, 425, 136382 CrossRef CAS PubMed.
- D. C. Nwobodo, M. C. Ugwu, C. Oliseloke Anie, M. T. S. Al-Ouqaili, J. Chinedu Ikem, U. Victor Chigozie and M. Saki, Antibiotic resistance: The challenges and some emerging strategies for tackling a global menace, J. Clin. Lab. Anal., 2022, 36(9), e24655 CrossRef PubMed.
- H. Yousefi, S. E. Samani, S. Khan, A. Prasad, A. Shakeri, Y. Li, C. D. M. Filipe and T. F. Didar, LISzyme Biosensors: DNAzymes Embedded in an Anti-biofouling Platform for Hands-free Real-Time Detection of Bacterial Contamination in Milk, ACS Nano, 2022, 16(1), 29–37 CrossRef CAS PubMed.
- S. Sakib, Z. Zhang, E. Osman, F. Kanji, F. Bakhshandeh, C. Pattyn, Y. Li, I. Zhitomirsky and L. Soleymani, Programming synthetic DNA molecules for building and breaking heterostructures between TiO2 nanomaterials for photoelectrochemical bacterial detection, Nano Today, 2023, 49, 101783 CrossRef CAS.
- X. Shan, Q. Feng, J. Yang and P. Wang, Dual-mode detection and efficient annihilation of pathogenic bacteria based on the construction of a ratiometric electrochemiluminescent/electrochemical sensor, Sens. Actuators, B, 2022, 372, 132633 CrossRef CAS.
- X. Shan, H. Xie, T. Zhou, M. Wu and J. Yang, Dual DNA recycling amplifications coupled with Au NPs@ZIF-MOF accelerator for enhanced electrochemical ratiometric sensing of pathogenic bacteria, Talanta, 2023, 263, 124751 CrossRef CAS PubMed.
- T. Wu, C. Wang, M. Wu, P. Wang and Q. Feng, Novel integrating polymethylene blue nanoparticles with dumbbell hybridization chain reaction for electrochemical detection of pathogenic bacteria, Food Chem., 2022, 382, 132501 CrossRef CAS PubMed.
- Y. Liu, X.-Z. Meng, X. Luo, H.-W. Gu, X.-L. Yin, W.-L. Han, H.-C. Yi and Y. Chen, Molecularly imprinted polymer combined with MOF-assisted redox recycling amplification: A powerful electrochemical sensing strategy for pathogenic bacteria, Sens. Actuators, B, 2024, 410, 135682 CrossRef CAS.
- H. Wang, L. Yang, Y. Tan, W. Deng and Q. Xie, Ag2S quantum dots loaded dendritic mesoporous silica nanospheres as signal amplification labels for ultrasensitive electrochemical immuno-biosensor for Staphylococcus aureus, J. Electroanal. Chem., 2022, 919, 116496 CrossRef CAS.
- R. Abedi, J. Bakhsh Raoof, M. Mohseni and A. Bagheri Hashkavayi, Sandwich-Type Electrochemical Aptasensor for Highly Sensitive and Selective Detection of Pseudomonas Aeruginosa Bacteria Using a Dual Signal Amplification Strategy, Bioelectrochemistry, 2023, 150, 108332 CrossRef CAS PubMed.
- L. Liu, Z. Yang, J. Zhang, L. Wang, J. Pang, A. Wang, L. Ding, H. Liu and X. Yu, Research progress in the application of MXene in bacterial detection and eradication, Mater. Today Phys., 2024, 43, 101412 CrossRef CAS.
- S. Wang, J. Hu, S. Xiao, M. Wang, J. Yu, Z. Jia, Z. Yu and N. Gan, Fluorescent/electrochemical dual-signal response biosensing strategy mediated by DNAzyme-ferrocene-triggered click chemistry for simultaneous rapid screening and quantitative detection of Vibrio parahaemolyticus, Sens. Actuators, B, 2023, 380, 133393 CrossRef CAS.
- X. Shan, D. Kuang, Q. Feng, M. Wu and J. Yang, A dual-mode ratiometric aptasensor for accurate detection of pathogenic bacteria based on recycling of DNAzyme activation, Food Chem., 2023, 423, 136287 CrossRef CAS PubMed.
- Y. Liu, X. Meng, Z. Ma, H. Gu, X. Luo, X. Yin, H. Yi and Y. Chen, Hybrid recognition-enabled ratiometric electrochemical sensing of Staphylococcus aureus via in situ growth of MOF/Ti3C2Tx-MXene and a self-reporting bacterial imprinted polymer, Food Chem., 2025, 463, 141496 CrossRef CAS PubMed.
- L. Yang, Y. Ding, Y. Ma, J. Wen, J. Wang, G. Dai and F. Mo, An electrochemical sensor based on 2D Zn-MOFs and 2D C-Ti3C2Tx composite materials for rapid and direct detection of various foodborne pathogens, Food Chem., 2025, 462, 140922 CrossRef CAS PubMed.
- G. Dai, Y. Li, Z. Li, J. Zhang, X. Geng, F. Zhang, Q. Wang and P. He, Zirconium-Based Metal–Organic Framework and Ti 3 C 2 T x Nanosheet-Based Faraday Cage-Type Electrochemical Aptasensor for Escherichia coli Detection, ACS Appl. Nano Mater., 2022, 5, 9201–9208 CrossRef CAS.
- M. Duan, B. Li, Y. He, Y. Zhao, Y. Liu, B. Zou, Y. Liu, J. Chen, R. Dai, X. Li and F. Jia, A CG@MXene nanocomposite-driven E-CRISPR biosensor for the rapid and sensitive detection of Salmonella Typhimurium in food, Talanta, 2024, 266, 125011 CrossRef CAS PubMed.
- O. Wang, X. Jia, J. Liu, M. Sun and J. Wu, Rapid and simple preparation of an MXene/polypyrrole-based bacteria imprinted sensor for ultrasensitive Salmonella detection, J. Electroanal. Chem., 2022, 918, 116513 CrossRef CAS.
- Z. Novakovic, M. Khalife, V. Costache, M. J. Camacho, S. Cardoso, V. Martins, I. Gadjanski, M. Radovic and J. Vidic, Rapid Detection and Identification of Vancomycin-Sensitive Bacteria Using an Electrochemical Apta-Sensor, ACS Omega, 2024, 9(2), 2841–2849 CrossRef CAS PubMed.
- R. B. Jamal, U. Bay Gosewinkel and E. E. Ferapontova, Electrocatalytic aptasensor for bacterial detection exploiting ferricyanide reduction by methylene blue on mixed PEG/aptamer monolayers, Bioelectrochemistry, 2024, 156, 108620 CrossRef CAS PubMed.
- Q. Tao, N. Tang, Y. Jiang, B. Chen, Y. Liu, X. Xiong and S. Liu, Double bipolar electrode electrochemiluminescence color switch for food-borne pathogens detection, Biosens. Bioelectron., 2023, 237, 115452 CrossRef CAS PubMed.
- Q. Feng, C. Wang, X. Miao and M. Wu, A novel paper-based electrochemiluminescence biosensor for non-destructive detection of pathogenic bacteria in real samples, Talanta, 2024, 267, 125224 CrossRef CAS PubMed.
- W. Wang, S. Xiao, M. Zeng, H. Xie and N. Gan, Dual-mode colorimetric-electrochemical biosensor for Vibrio parahaemolyticus detection based on CuO2 nanodot-encapsulated metal-organic framework nanozymes, Sens. Actuators, B, 2023, 387, 133835 CrossRef CAS.
- H. Morsalpour, H. R. Zare, Z. Shekari and M. Mirbagheri-Firousabad, Development of a highly sensitive and selective electrochemical aptasensor for the detection of Staphylococcus aureus in various samples, Microchem. J., 2024, 207, 111940 CrossRef CAS.
- L. Zhao, G. Rosati, A. Piper, C. de Carvalho Castro e Silva, L. Hu, Q. Yang, F. Della Pelle, R. R. Alvarez-Diduk and A. Merkoci, Laser Reduced Graphene Oxide Electrode for Pathogenic Escherichia coli Detection, ACS Appl. Mater. Interfaces, 2023, 15(7), 9024–9033 CrossRef CAS PubMed.
- M. Angelopoulou, D. Kourti, M. Mertiri, P. Petrou, S. Kakabakos and C. Kokkinos, A 3D-Printed Electrochemical Immunosensor Employing Cd/Se ZnS QDs as Labels for the Rapid and Ultrasensitive Detection of Salmonella typhimurium in Poultry Samples, Chemosensors, 2023, 11(9), 475 CrossRef CAS.
- Q. Sun, X. Liu, H. Tang, Y. Qian, H. Gu and H. He, A Sandwich-type Electrochemical Immunosensor for the Sensitive Determination of Salmonella Typhimurium in Food, Electroanalysis, 2022, 34(5), 911–918 CrossRef CAS.
- M. Fleischmann, P. J. Hendra and A. J. McQuillan, Raman spectra of pyridine adsorbed at a silver electrode, Chem. Phys. Lett., 1974, 26(2), 163–166 CrossRef CAS.
- J. Langer, D. Jimenez de Aberasturi, J. Aizpurua, R. A. Alvarez-Puebla, B. Auguie, J. J. Baumberg, G. C. Bazan, S. E. J. Bell, A. Boisen, A. G. Brolo, J. Choo, D. Cialla-May, V. Deckert, L. Fabris, K. Faulds, F. J. Garcia de Abajo, R. Goodacre, D. Graham, A. J. Haes, C. L. Haynes, C. Huck, T. Itoh, M. Kall, J. Kneipp, N. A. Kotov, H. Kuang, E. C. Le Ru, H. K. Lee, J. F. Li, X. Y. Ling, S. A. Maier, T. Mayerhöfer, M. Moskovits, K. Murakoshi, J. M. Nam, S. Nie, Y. Ozaki, I. Pastoriza-Santos, J. Perez-Juste, J. Popp, A. Pucci, S. Reich, B. Ren, G. C. Schatz, T. Shegai, S. Schlucker, L. L. Tay, K. G. Thomas, Z. Q. Tian, R. P. Van Duyne, T. Vo-Dinh, Y. Wang, K. A. Willets, C. Xu, H. Xu, Y. Xu, Y. S. Yamamoto, B. Zhao and L. M. Liz-Marzan, Present and Future of Surface-Enhanced Raman Scattering, ACS Nano, 2020, 14(1), 28–117 CrossRef CAS PubMed.
- Samriti, V. Rajput, R. K. Gupta and J. Prakash, Engineering metal oxide semiconductor nanostructures for enhanced charge transfer: fundamentals and emerging SERS applications, J. Mater. Chem. C, 2022, 10(1), 73–95 RSC.
- J. Plou, P. S. Valera, I. Garcia, C. D. L. de Albuquerque, A. Carracedo and L. M. Liz-Marzan, Prospects of Surface-Enhanced Raman Spectroscopy for Biomarker Monitoring toward Precision Medicine, ACS Photonics, 2022, 9(2), 333–350 CrossRef CAS PubMed.
- M. Wang, H. Diao, C. Dou, Z. Wu, L. Ruan, Z. Wang, Z. Wang, W. Ye, J. Duan and Y. Li, Label-free detection of pathogenic microorganism using Ag NPs@PDMS sponge SERS substrate and machine learning, Sens. Actuators, B, 2024, 413, 135811 CrossRef CAS.
- R. M. Jarvis and R. Goodacre, Discrimination of bacteria using surface-enhanced Raman spectroscopy, Anal. Chem., 2004, 76(1), 40–47 CrossRef CAS PubMed.
- S. Y. Lee, S.-J. Kim and J.-H. Ha, Quantification of Pectobacterium carotovorum subsp. carotovorum in kimchi cabbage using a surface-enhanced Raman scattering platform with silver nanostructures, Biosens. Bioelectron., 2025, 267, 116766 CrossRef CAS PubMed.
- S. Husain, C. Mutalik, S. Yougbaré, C.-Y. Chen and T.-R. Kuo, Plasmonic Au@Ag Core–Shell Nanoisland Film for Photothermal Inactivation and Surface-Enhanced Raman Scattering Detection of Bacteria, Nanomaterials, 2024, 695 CrossRef CAS PubMed.
- S. Zhou, X. Guo, H. Huang, X. Huang, X. Zhou, Z. Zhang, G. Sun, H. Cai, H. Zhou and P. Sun, Triple-Function Au-Ag-Stuffed Nanopancakes for SERS Detection, Discrimination, and Inactivation of Multiple Bacteria, Anal. Chem., 2022, 94(15), 5785–5796 CrossRef CAS PubMed.
- P. Wen, F. Yang, H. Zhao, Y. Xu, S. Li and L. Chen, Novel Digital SERS-Microfluidic Chip for Rapid and Accurate Quantification of Microorganisms, Anal. Chem., 2024, 96(4), 1454–1461 CrossRef CAS PubMed.
- X. Huang, L. Chen, W. Zhi, R. Zeng, G. Ji, H. Cai, J. Xu, J. Wang, S. Chen, Y. Tang, J. Zhang, H. Zhou and P. Sun, Urchin-Shaped Au-Ag@Pt Sensor Integrated Lateral Flow Immunoassay for Multimodal Detection and Specific Discrimination of Clinical Multiple Bacterial Infections, Anal. Chem., 2023, 95(35), 13101–13112 CrossRef CAS PubMed.
- E. Garcia-Rico, R. A. Alvarez-Puebla and L. Guerrini, Direct surface-enhanced Raman scattering (SERS) spectroscopy of nucleic acids: from fundamental studies to real-life applications, Chem. Soc. Rev., 2018, 47(13), 4909–4923 RSC.
- K. Hadjigeorgiou, E. Kastanos, C. Pitris and C. Andreou, Surface Enhanced Raman Spectroscopy as a Sensitive Method for UTI Diagnosis, IEEE Sens. J., 2022, 22(11), 10063–10074 CAS.
- J.-W. Tang, J.-W. Lyu, J.-X. Lai, X.-D. Zhang, Y.-G. Du, X.-Q. Zhang, Y.-D. Zhang, B. Gu, X. Zhang, B. Gu and L. Wang, Determination of Shigella spp. via label-free SERS spectra coupled with deep learning, Microchem. J., 2023, 189, 108539 CrossRef CAS.
- Y. Liu, Y. Gao, R. Niu, Z. Zhang, G.-W. Lu, H. Hu, T. Liu and Z. Cheng, Rapid and accurate bacteria identification through deep-learning-based two-dimensional Raman spectroscopy, Anal. Chim. Acta, 2024, 1332, 343376 CrossRef CAS PubMed.
- L. Shang, Y. Wang, F. Chen, H. Peng, X. Bao, X. Tang, K. Liu, L. Xu, D. Xiao, P. Liang and B. Li, Ultra-fast identification of lactic acid bacteria colonies based on droplet microcavity label-free SERS, LWT, 2024, 204, 116435 CrossRef CAS.
- Y. Zhao, Z. Li, Y. Yan, Y. Wen, Y. Ning, Z. Li and H. Wang, Discriminant analysis of a mixture of pathogenic bacteria into different types and proportions by surface-enhanced Raman scattering spectroscopy combined with chemometric methods, Vib. Spectrosc., 2024, 132, 103692 CrossRef CAS.
- H. Zuo, Y. Sun, M. Huang, S. M. Fowler, J. Liu, Y. Zhang and Y. Mao, Classification and Identification of Foodborne Bacteria in Beef by Utilising Surface-Enhanced Raman Spectroscopy Coupled with Chemometric Methods, Foods, 2024, 3688 CrossRef CAS PubMed.
- W. Liu, J.-W. Tang, J.-W. Lyu, J.-J. Wang, Y.-C. Pan, X.-Y. Shi, Q.-H. Liu, X. Zhang, B. Gu and L. Wang, Discrimination between Carbapenem-Resistant and Carbapenem-Sensitive Klebsiella pneumoniae Strains through Computational Analysis of Surface-Enhanced Raman Spectra: a Pilot Study, Microbiol. Spectrum, 2022, 10(1), e02409–e02421 CAS.
- J.-W. Lyu, X. Zhang, J.-W. Tang, Y.-H. Zhao, S.-L. Liu, Y. Zhao, N. Zhang, D. Wang, L. Ye, X.-L. Chen, L. Wang and B. Gu, Rapid Prediction of Multidrug-Resistant Klebsiella pneumoniae through Deep Learning Analysis of SERS Spectra, Microbiol. Spectrum, 2023, 11(2), e04126 Search PubMed.
- W. Wang, X. Wang, Y. Huang, Y. Zhao, X. Fang, Y. Cong, Z. Tang, L. Chen, J. Zhong, R. Li, Z. Guo, Y. Zhang and S. Li, Raman spectrum combined with deep learning for precise recognition of Carbapenem-resistant Enterobacteriaceae, Anal. Bioanal. Chem., 2024, 416(10), 2465–2478 CrossRef CAS PubMed.
- Z. Al-Shaebi, F. Uysal Ciloglu, M. Nasser, M. Kahraman and O. Aydin, Staphylococcus Aureus-Related antibiotic resistance detection using synergy of Surface-Enhanced Raman spectroscopy and deep learning, Biomed. Signal Process. Control, 2024, 91, 105933 CrossRef CAS.
- N. Mehmood, M. W. Akram, M. I. Majeed, H. Nawaz, M. A. Aslam, A. Naman, M. Wasim, U. Ghaffar, A. Kamran, S. Nadeem, N. Kanwal and M. Imran, Surface-enhanced Raman spectroscopy for the characterization of bacterial pellets of Staphylococcus aureus infected by bacteriophage, RSC Adv., 2024, 14(8), 5425–5434 RSC.
- H. Tahseen, N. Ul Huda, H. Nawaz, M. I. Majeed, N. Alwadie, N. Rashid, M. A. Aslam, N. Zafar, M. Asghar, A. Anwar, A. Ashraf and R. Umer, Surface-enhanced Raman spectroscopy for comparison of biochemical profile of bacteriophage sensitive and resistant methicillin-resistant Staphylococcus aureus (MRSA) strains, Spectrochim. Acta, Part A, 2024, 310, 123968 CrossRef CAS PubMed.
- B. Krafft, A. Tycova, R. D. Urban, C. Dusny and D. Belder, Microfluidic device for concentration and SERS-based detection of bacteria in drinking water, Electrophoresis, 2021, 42(1–2), 86–94 CrossRef CAS PubMed.
- L. Shang, P. Liang, L. Xu, Y. Xue, K. Liu, Y. Wang, X. Bao, F. Chen, H. Peng, Y. Wang, J. Ju and B. Li, Stable SERS Detection of Lactobacillus fermentum Using Optical Tweezers in a Microfluidic Environment, Anal. Chem., 2024, 96(1), 248–255 CrossRef PubMed.
- R. Hunter, A. N. Sohi, Z. Khatoon, V. R. Berthiaume, E. I. Alarcon, M. Godin and H. Anis, Optofluidic label-free SERS platform for rapid bacteria detection in serum, Sens. Actuators, B, 2019, 300, 126907 CrossRef CAS.
- H. Jayan, R. Zhou, Y. Zheng, S. Xue, L. Yin, H. R. El-Seedi, X. Zou and Z. Guo, Microfluidic-SERS platform with in situ nanoparticle synthesis for rapid E. coli detection in food, Food Chem., 2025, 471, 142800 CrossRef CAS PubMed.
- A. S. Rourke-Funderburg, A. Mahadevan-Jansen and A. K. Locke, Characterization of vaginal Lactobacillus in biologically relevant fluid using surface-enhanced Raman spectroscopy, Analyst, 2024, 149(19), 4862–4871 RSC.
- A. Yadav, A. K. Yadav and N. Tarannum, Fabrication of Aluminum Foil Integrated Pegylated Gold Nanoparticle Surface-Enhanced Raman Scattering Substrate for the Detection and Classification of Uropathogenic Bacteria, ACS Appl. Bio Mater., 2024, 7(9), 6127–6137 CrossRef CAS PubMed.
- A. Anwer, A. Shahzadi, H. Nawaz, M. I. Majeed, A. Alshammari, N. A. Albekairi, M. U. Hussain, I. Amin, A. Bano, A. Ashraf, N. Rehman, R. M. Pallares and N. Akhtar, Differentiation of different dibenzothiophene (DBT) desulfurizing bacteria via surface-enhanced Raman spectroscopy (SERS), RSC Adv., 2024, 14(28), 20290–20299 RSC.
- Z. Ishfaq, L. A. Almutairi, M. Y. Ali, S. H. Alrefaee, M. A. Fahmy, E. A. Shokralla, L. G. Alharbe, A. Ali, A. Ashfaq and A. R. Abd-Elwahed, Enhanced surface functionalization of 2D molybdenum/tin chalcogenide nanostructures for effective SERS detection of Escherichia coli, RSC Adv., 2024, 14(47), 35021–35034 RSC.
- L. Jin, X. Cai, F. Ren and J. Yang, An aptamer-based SERS method for rapid screening and identification of pathogens assisted by machine learning technique with robustness evaluation, Sens. Actuators, B, 2024, 405, 135356 CrossRef CAS.
- S. Peng, Z. Zhang, M. Xin and D. Liu, SERS-based Ag NCs@PDMS flexible substrate combined with chemometrics for rapid detection of foodborne pathogens on egg surface, Microchim. Acta, 2024, 191(10), 612 CrossRef CAS PubMed.
- S. Senapati, M. Kaur, N. Singh, S. Kulkarni and J. Singh, Affordable Paper-Based Surface-Enhanced Raman Scattering Substrates Containing Silver Nanorods Using Glancing-Angle Deposition for Nosocomial Infection Detection, ACS Appl. Nano Mater., 2024, 7, 6736–6748 CrossRef CAS.
- S. Das, K. Saxena, J.-C. Tinguely, A. Pal, N. Wickramasinghe, A. Khezri, V. Dubey, A. Ahmad, V. Perumal, R. Ahmad, D. Wadduwage, B. Ahluwalia and D. Mehta, SERS Nanowire Chip and Machine Learning-Enabled Classification of Wild-Type and Antibiotic-Resistant Bacteria at Species and Strain Levels, ACS Appl. Mater. Interfaces, 2023, 15, 24047–24058 CrossRef CAS PubMed.
- G. Su, Y. Liu, Y. Hou, R. Zhang, W. Wang, J. Zhang and L. Dang, Surface-Enhanced Raman Spectroscopy Sensor Integrated with Ag@ZIF-8@Au Core-Shell-Shell Nanowire Membrane for Enrichment, Ultrasensitive Detection, and Inactivation of Bacteria in the Environment, ACS Appl. Mater. Interfaces, 2024, 16(22), 28080–28092 CrossRef CAS PubMed.
- K. W. Chang, H. W. Cheng, J. Shiue, J. K. Wang, Y. L. Wang and N. T. Huang, Antibiotic Susceptibility Test with Surface-Enhanced Raman Scattering in a Microfluidic System, Anal. Chem., 2019, 91(17), 10988–10995 CrossRef CAS PubMed.
- H. K. Huang, H. W. Cheng, C. C. Liao, S. J. Lin, Y. Z. Chen, J. K. Wang, Y. L. Wang and N. T. Huang, Bacteria encapsulation and rapid antibiotic susceptibility test using a microfluidic microwell device integrating surface-enhanced Raman scattering, Lab Chip, 2020, 20(14), 2520–2528 RSC.
- C.-C. Liao, Y.-Z. Chen, S.-J. Lin, H.-W. Cheng, J.-K. Wang, Y.-L. Wang, Y.-Y. Han and N.-T. Huang, A microfluidic microwell device operated by the automated microfluidic control system for surface-enhanced Raman scattering-based antimicrobial susceptibility testing, Biosens. Bioelectron., 2021, 191, 113483 CrossRef CAS PubMed.
- P. Wen, F. Yang, H. Zhao, S. Li, Y. Xu and L. Chen, Microcavity Array-Based Digital SERS Chip for Rapid and Accurate Label-free Quantitative Detection of Live Bacteria, ACS Sens., 2024, 9(11), 6167–6173 CrossRef CAS PubMed.
- Y. F. Chen, M. C. Lu, C. J. Lee and C. W. Chiu, Flexible nanohybrid substrates utilizing gold nanocubes/nano mica platelets with 3D lightning-rod effect for highly efficient bacterial biosensors based on surface-enhanced Raman scattering, J. Mater. Chem. B, 2024, 12(13), 3226–3239 RSC.
- C. Wang, X. Mu, J. Huo, B. Zhang and K. Zhang, Highly-efficient SERS detection for E. coli using a microfluidic chip with integrated NaYF4:Yb,Er@SiO2@Au under near-infrared laser excitation, Microsyst. Technol., 2021, 27(9), 3285–3291 CrossRef CAS.
- B. Shi, L. Jiang, R. Ma, W. Zhao, Y. Zheng, W. Pan, M. Liu, S. Jin and Y. Zhou, Ti3C2Tx-AuNP based paper substrates for label-free SERS detection of bacteria and multimodal antibacterials, RSC Adv., 2024, 14(26), 18739–18749 RSC.
- W. Zhao, L. Jiang, W. Pan, D. Li, M. Duan, Y. Zhou and L. Chen, V2CTx-Au Nanostructures for Rapid Surface-Enhanced Raman Scattering Detection of Bacteria and Synergistic Photothermal/Photodynamic Therapy, ACS Appl. Nano Mater., 2024, 14673–14683 CrossRef CAS.
- P. Zeng, Q. Guan, Q. Zhang, L. Yu, X. Yan, Y. Hong, L. Duan and C. Wang, SERS detection of foodborne pathogens in beverage with Au nanostars, Microchim. Acta, 2023, 191(1), 28 CrossRef PubMed.
- K. Li, Y. Yang, C. Xu, Y. Ye, L. Huang, L. Sun, Y. Cai, W. Zhou, Y. Ge, Y. Li, Q. Zhang, Y. Wang and X. Liu, Vertical gold nanowires-based surface-enhanced Raman scattering for direct detection of ocular bacteria, Sens. Actuators, B, 2023, 380, 133381 CrossRef CAS.
- Kenry, F. Nicolson, L. Clark, S. R. Panikkanvalappil, B. Andreiuk and C. Andreou, Advances in Surface Enhanced Raman Spectroscopy for in Vivo Imaging in Oncology, Nanotheranostics, 2022, 6(1), 31–49 CrossRef CAS PubMed.
- X. Song, H. Wang, X. Shao, X. Yu and X. Xu, Au@Ag core-shell nanorods enhance surface-enhanced Raman scattering aptasensor for ultrasensitive detection of Salmonella Typhimurium, Food Control, 2024, 161, 110379 CrossRef CAS.
- P. Wang, Y. Liu, X. Li, C. Li and G. Li, A SERS aptasensor based on HCR mediated signal conversion strategy for the rapid and reliable determination of Escherichia coli O157:H7, Microchem. J., 2024, 199, 109951 CrossRef CAS.
- F. Mi, M. Guan, Y. Wang, G. Chen, P. Geng and C. Hu, Integration of three non-interfering SERS probes combined with ConA-functionalized magnetic nanoparticles for extraction and detection of multiple foodborne pathogens, Microchim. Acta, 2023, 190(3), 103 CrossRef CAS PubMed.
- Y. Xiao, W. Liu, Y. Zhang, S. Zheng, J. Liao, H. Shan, B. Tian, T. Wu, L. Zhang, Z. Tu, Y. Hua, B. Gu and X. Hu, Simple and rapid co-freezing construction of SERS signal probes for the sensitive detection of pathogens, Chem. Eng. J., 2023, 466, 143066 CrossRef CAS.
- C. Y. Effah, L. Ding, L. Tan, S. He, X. Li, H. Yuan, Y. Li, S. Liu, T. Sun and Y. Wu, A SERS bioassay based on vancomycin-modified PEI-interlayered nanocomposite and aptamer-functionalized SERS tags for synchronous detection of Acinetobacter baumannii and Klebsiella pneumoniae, Food Chem., 2023, 423, 136242 CrossRef CAS PubMed.
- X. Wang, W. Li, S. Dai, M. Dou, S. Jiao, J. Yang, W. Li, Y. Su, Q. Li and J. Li, High-throughput, highly sensitive and rapid SERS detection of Escherichia coli O157:H7 using aptamer-modified Au@macroporous silica magnetic photonic microsphere array, Food Chem., 2023, 424, 136433 CrossRef CAS PubMed.
- W. Shen, J. Liu, C. Wang, C. Wang, Y. Nie, B. Gu, R. Xiao and R. Hao, Development of a dual-Raman signal lateral flow immunoassay with self-assembling nanogapped SERS tags for the simultaneous monitoring of respiratory bacteria and virus, Sens. Actuators, B, 2024, 417, 136078 CrossRef CAS.
- Ü. Dogan, F. Sucularlı, E. Yildirim, D. Cetin, Z. Suludere, I. H. Boyaci and U. Tamer, Escherichia coli Enumeration in a Capillary-Driven Microfluidic Chip with SERS, Biosensors, 2022, 12(9), 765 CrossRef PubMed.
- B. Huo, L. Xia, Y. Hu and G. Li, Flexible microfluidic co-recognition coupled with magnetic enrichment and silent SERS sensing for simultaneous analysis of bacteria in food, Biosens. Bioelectron., 2024, 255, 116227 CrossRef CAS PubMed.
- L. Rodriguez-Lorenzo, A. Garrido-Maestu, A. Bhunia, B. Espiña, M. Prado, L. Dieguez and S. Abalde-Cela, Gold Nanostars for the Detection of Foodborne Pathogens via Surface-Enhanced Raman Scattering Combined with Microfluidics, ACS Appl. Nano Mater., 2019, 2019, 6081–6086 CrossRef.
- S. Asgari, R. Dhital, A. Mustapha and M. Lin, Duplex detection of foodborne pathogens using a SERS optofluidic sensor coupled with immunoassay, Int. J. Food Microbiol., 2022, 383, 109947 CrossRef CAS PubMed.
- S. Asgari, R. Dhital, S. A. Aghvami, A. Mustapha, Y. Zhang and M. Lin, Separation and detection of E. coli O157:H7 using a SERS-based microfluidic immunosensor, Microchim. Acta, 2022, 189(3), 111 CrossRef CAS PubMed.
- J. Zhuang, Z. Zhao, K. Lian, L. Yin, J. Wang, S. Man, G. Liu and L. Ma, SERS-based CRISPR/Cas assay on microfluidic paper analytical devices for supersensitive detection of pathogenic bacteria in foods, Biosens. Bioelectron., 2022, 207, 114167 CrossRef CAS PubMed.
- A. V. Markin, A. I. Arzhanukhina, N. E. Markina and I. Y. Goryacheva, Analytical performance of electrochemical surface-enhanced Raman spectroscopy: A critical review, TrAC, Trends Anal. Chem., 2022, 157, 116776 CrossRef CAS.
- R. A. Karaballi, A. Nel, S. Krishnan, J. Blackburn and C. L. Brosseau, Development of an electrochemical surface-enhanced Raman spectroscopy (EC-SERS) aptasensor for direct detection of DNA hybridization, Phys. Chem. Chem. Phys., 2015, 17(33), 21356–21363 RSC.
- C. L. Brosseau, A. Colina, J. V. Perales-Rondon, A. J. Wilson, P. B. Joshi, B. Ren and X. Wang, Electrochemical surface-enhanced Raman spectroscopy, Nat. Rev. Methods Primers, 2023, 3(1), 79 CrossRef CAS.
- D. Y. Wu, J. F. Li, B. Ren and Z. Q. Tian, Electrochemical surface-enhanced Raman spectroscopy of nanostructures, Chem. Soc. Rev., 2008, 37(5), 1025–1041 RSC.
- A. J. Wain and M. A. O'Connell, Advances in surface-enhanced vibrational spectroscopy at electrochemical interfaces, Adv. Phys.:X, 2017, 2(1), 188–209 CAS.
- K. A. Willets, Probing nanoscale interfaces with electrochemical surface-enhanced Raman scattering, Curr. Opin. Electrochem., 2019, 13, 18–24 CrossRef CAS.
- T. P. Lynk, C. S. Sit and C. L. Brosseau, Electrochemical Surface-Enhanced Raman Spectroscopy as a Platform for Bacterial Detection and Identification, Anal. Chem., 2018, 90(21), 12639–12646 CrossRef CAS PubMed.
- X. Tang, T. Wu, F. Pan, G. Lu, W. Zeng, W. Fan and L. Wu, An applied potential powered screen-printed electrode for sensitive EC-SERS detection of acetamiprid, Sens. Actuators, B, 2024, 404, 135279 CrossRef CAS.
- S. H. Lee, I. B. Ansah, C. Mun, J.-Y. Yang, H. S. Jung, J. Park, S.-Y. Nam, M.-Y. Lee, S. Lee, D.-H. Kim and S.-G. Park, 3D plasmonic hotspot engineering toward ultrasensitive and rapid EC-SERS recognition of plasticizers, Microchem. J., 2024, 196, 109599 CrossRef CAS.
- I. Drozdowski and C. Brosseau, Nonaqueous Electrochemical Surface-Enhanced Raman Spectroscopy (EC-SERS) Using Screen-Printed Electrodes, J. Phys. Chem. C, 2024, 128, 14734–14741 CrossRef CAS.
- T. Wu, X. Tang, W. Zeng, Y. Han, S. Zhang, J. Wei and L. Wu, Potential powered EC-SERS for sensitive detection of acetamiprid, J. Food Eng., 2024, 378, 112109 CrossRef CAS.
- N. Albarghouthi, M. Eisnor, C. Pye and C. Brosseau, Electrochemical Surface-Enhanced Raman Spectroscopy (EC-SERS) and Computational Study of Atrazine: Toward Point-of-Need Detection of Prevalent Herbicides, J. Phys. Chem. C, 2022, 126, 9836–9842 CrossRef CAS.
- A. Falamas, D. Cuibus, N. Tosa, I. Brezestean, C. M. Muntean, K. Milenko, E. Vereshchagina, R. Moldovan, E. Bodoki and C. Farcau, Toward microfluidic SERS and EC-SERS applications via tunable gold films over nanospheres, Discover Nano, 2023, 18(1), 73 CrossRef CAS PubMed.
- D. Ibáñez, M. B. González-García, D. Hernández-Santos and P. Fanjul-Bolado, Detection of dithiocarbamate, chloronicotinyl and organophosphate pesticides by electrochemical activation of SERS features of screen-printed electrodes, Spectrochim. Acta, Part A, 2021, 248, 119174 CrossRef PubMed.
- S. Li, Y. Ling, J. Chen, X. Yuan and Z. Zhang, Portable Copper-Based Electrochemical SERS Sensor for Point-of-Care Testing of Paraquat and Diquat by On-Site Electrostatic Preconcentration, Langmuir, 2024, 15677–15687 CAS.
- J. Liu, R. Siavash Moakhar, S. Mahshid, F. Vasefi and S. Wachsmann-Hogiu, Multimodal electrochemical and SERS platform for chlorfenapyr detection, Appl. Surf. Sci., 2021, 566, 150617 CrossRef CAS.
- R. Moldovan, M. Perez-Estebanez, A. Heras, E. Bodoki and A. Colina, Activating the SERS features of screen-printed electrodes with thiocyanate for sensitive and robust EC-SERS analysis, Sens. Actuators, B, 2024, 407, 135468 CrossRef CAS.
- M. Velička, E. Zacharovas, S. Adomavičiūtė and V. Šablinskas, Detection of caffeine intake by means of EC-SERS spectroscopy of human saliva, Spectrochim. Acta, Part A, 2021, 246, 118956 CrossRef PubMed.
- C. E. Ott, A. Burns, E. Sisco and L. E. Arroyo, Targeted fentanyl screening utilizing electrochemical surface-enhanced Raman spectroscopy (EC-SERS) applied to authentic seized drug casework samples, Forensic Chem., 2023, 34, 100492 CrossRef CAS.
- M. Haroon, M. Ashraf, N. Ullah, M. N. Tahir and A. A. Al-Saadi, SERS and EC-SERS detection of local anesthetic procaine using Pd loaded highly reduced graphene oxide nanocomposite substrate, Spectrochim. Acta, Part A, 2022, 278, 121381 CrossRef CAS PubMed.
- Y. Mu, J. Islam, R. Murray, C. Larrigy, A. Russo, X. Zhang, A. J. Quinn and D. Iacopino, Silver nanoparticles - laser induced graphene (Ag NPs - LIG) hybrid electrodes for sensitive electrochemical-surface enhanced Raman spectroscopy (EC-SERS) detection, Analyst, 2023, 148(13), 3087–3096 RSC.
- C. E. Ott, M. Perez-Estebanez, S. Hernandez, K. Kelly, K. A. Dalzell, M. J. Arcos-Martinez, A. Heras, A. Colina and L. E. Arroyo, Forensic Identification of Fentanyl and its Analogs by Electrochemical-Surface Enhanced Raman Spectroscopy (EC-SERS) for the Screening of Seized Drugs of Abuse, Front. Anal. Sci., 2022, 2, 834820 CrossRef.
- T. Cooman, C. E. Ott and L. E. Arroyo, Evaluation and classification of fentanyl-related compounds using EC-SERS and machine learning, J. Forensic Sci., 2023, 68(5), 1520–1526 CrossRef CAS PubMed.
- C. E. Ott, A. Wilcox, S. Kalb, A. McConnell, E. Sisco and L. E. Arroyo, Development and evaluation of a nontargeted electrochemical surface-enhanced Raman spectroscopy (EC-SERS) screening method applied to forensic seized drug casework samples, Forensic Chem., 2024, 41, 100621 CrossRef CAS.
- L. Lu, Y. Zhou, T. Zheng and Y. Tian, SERS and EC dual-mode detection for dopamine based on WO3-SnO2 nanoflake arrays, Nano Res., 2023, 16(3), 4049–4054 CrossRef CAS.
- M. Haroon and A. A. Al-Saadi, The Role of Gold-Loaded Copper Oxide Nanoplates in the Formulation of Sensitive EC-SERS Substrates for the Detection of Anti-Cancer Drugs: Spectro–Electro and DFT Studies, Arabian J. Sci. Eng., 2024, 49(1), 673–684 CrossRef CAS.
- C. M. Muntean, D. Cuibus, S. Boca, A. Falamas, N. Tosa, I. A. Brezeştean, A. Bende, L. Barbu-Tudoran, R. Moldovan, E. Bodoki and C. Farcǎu, Gold vs. Silver Colloidal Nanoparticle Films for Optimized SERS Detection of Propranolol and Electrochemical-SERS Analyses, Biosensors, 2023, 530 CrossRef CAS PubMed.
- M. M. Eisnor, K. E. R. McLeod, S. Bindesri, S. A. Svoboda, K. L. Wustholz and C. L. Brosseau, Electrochemical surface-enhanced Raman spectroscopy (EC-SERS): a tool for the identification of polyphenolic components in natural lake pigments, Phys. Chem. Chem. Phys., 2021, 24(1), 347–356 RSC.
- Z. Zhang, J. Liu, Y. Dai, M. Ye, Y. Sun, K. Zhang and J.-J. Xu, Dual-mode SERS-ECL biosensor for robust detection of circulating miRNAs based on in situ synthesized probes, Chem. Eng. J., 2024, 495, 153607 CrossRef CAS.
- I. B. Ansah, S. H. Lee, J.-Y. Yang, C. Mun, S. Jung, H. S. Jung, M.-Y. Lee, T. Kang, S. Lee, D.-H. Kim and S.-G. Park, In situ fabrication of 3D interior hotspots templated with a protein@Au core–shell structure for label-free and on-site SERS detection of viral diseases, Biosens. Bioelectron., 2023, 220, 114930 CrossRef CAS PubMed.
- Y. Sun, A. Feng, R. Zhang, K. Quan, Z. Hua and K. Zhang, SERS and ECL dual-mode sensing of bacterial lipopolysaccharide based on bifunctional DNA strands mediated transformation from Au@Ag nanocubes to silver nanoclusters, J. Hazard. Mater., 2025, 488, 137283 CrossRef CAS PubMed.
- M. Haroon, M. Tahir, H. Nawaz, M. I. Majeed and A. A. Al-Saadi, Surface-enhanced Raman scattering (SERS) spectroscopy for prostate cancer diagnosis: A review, Photodiagn. Photodyn. Ther., 2022, 37, 102690 CrossRef CAS PubMed.
- L. Jiang, L. Luo, Z. Zhang, C. Kang, Z. Zhao, D. Chen and Y. Long, Rapid detection of Pseudomonas syringae pv. actinidiae by electrochemical surface-enhanced Raman spectroscopy, Talanta, 2024, 268, 125336 CrossRef CAS PubMed.
- K. E. R. McLeod, T. P. Lynk, C. S. Sit and C. L. Brosseau, On the origin of electrochemical surface-enhanced Raman spectroscopy (EC-SERS) signals for bacterial samples: the importance of filtered control studies in the development of new bacterial screening platforms, Anal. Methods, 2019, 11(7), 924–929 RSC.
- H. Do, S. R. Kwon, K. Fu, N. Morales-Soto, J. D. Shrout and P. W. Bohn, Electrochemical Surface-Enhanced Raman Spectroscopy of Pyocyanin Secreted by Pseudomonas aeruginosa Communities, Langmuir, 2019, 35(21), 7043–7049 CrossRef CAS PubMed.
- S. H. Lee, W.-C. Lee, E. H. Koh, I. B. Ansah, J.-Y. Yang, C. Mun, S. Lee, D.-H. Kim, H. S. Jung and S.-G. Park, Organometallic hotspot engineering for ultrasensitive EC-SERS detection of pathogenic bacteria-derived DNAs, Biosens. Bioelectron., 2022, 210, 114325 CrossRef CAS PubMed.
- R.-S. Juang, K.-S. Wang, T.-Y. Kuan, Y.-J. Chu, R.-J. Jeng, A. Hardiansyah, S.-H. Liu and T.-Y. Liu, Electric field-stimulated Raman scattering enhancing biochips fabricated by Au nano-islands deposited on laser-scribed 3D graphene for uremic toxins detection, J. Taiwan Inst. Chem. Eng., 2024, 154, 105115 CrossRef CAS.
- Y. Lai, G. Jiang, T. Liang, X. Huang, W. Jiang, W. Xu, R. Sun, Z. Dai and C. Li, Rapid analysis of Bacillus cereus spore biomarkers based on porous channel cuttlebone SERS substrate, Anal. Chim. Acta, 2024, 1320, 343034 CrossRef CAS PubMed.
- J. Jia, S. R. Kwon, S. Baek, V. Sundaresan, T. Cao, A. R. Cutri, K. Fu, B. Roberts, J. D. Shrout and P. W. Bohn, Actively Controllable Solid-Phase Microextraction in a Hierarchically Organized Block Copolymer-Nanopore Electrode Array Sensor for Charge-Selective Detection of Bacterial Metabolites, Anal. Chem., 2021, 93(43), 14481–14488 CrossRef CAS PubMed.
- N. S. Chandel, Nucleotide Metabolism, Cold Spring Harbor Perspect. Biol., 2021, 13(7), a040592 CrossRef CAS PubMed.
- M. A. Rouf and R. F. Lomprey Jr., Degradation of uric acid by certain aerobic bacteria, J. Bacteriol., 1968, 96(3), 617–622 CrossRef CAS PubMed.
- L. Zhao, J. Blackburn and C. L. Brosseau, Quantitative detection of uric acid by electrochemical-surface enhanced Raman spectroscopy using a multilayered Au/Ag substrate, Anal. Chem., 2015, 87(1), 441–447 CrossRef CAS PubMed.
- S. Hernandez, J. V. Perales-Rondon, A. Heras and A. Colina, Determination of uric acid in synthetic urine by using electrochemical surface oxidation enhanced Raman scattering, Anal. Chim. Acta, 2019, 1085, 61–67 CrossRef CAS PubMed.
- C.-Y. Huang and H.-C. Hsiao, Integrated EC-SERS Chip with Uniform Nanostructured EC-SERS Active Working Electrode for Rapid Detection of Uric Acid, Sensors, 2020, 20(24), 7066 CrossRef CAS PubMed.
- N. R. Hendricks-Leukes, M. R. Jonas, Z. C. Mlamla, M. Smith and J. M. Blackburn, Dual-Approach Electrochemical Surface-Enhanced Raman Scattering Detection of Mycobacterium tuberculosis in Patient-Derived Biological Specimens: Proof of Concept for a Generalizable Method to Detect and Identify Bacterial Pathogens, ACS Sens., 2022, 7(5), 1403–1418 CrossRef CAS PubMed.
- C. Meng, S. Knežević, F. Du, Y. Guan, F. Kanoufi, N. Sojic and G. Xu, Recent advances in electrochemiluminescence imaging analysis, eScience, 2022, 2(6), 591–605 CrossRef.
- W. Wei, H. Lin, T. Hao, X. Su, X. Jiang, S. Wang, Y. Hu and Z. Guo, Dual-mode ECL/SERS immunoassay for ultrasensitive determination of Vibrio vulnificus based on multifunctional MXene, Sens. Actuators, B, 2021, 332, 129525 CrossRef CAS.
- M. Tian, J. Wang, C. Li, Z. Wang, G. Liu, E. Lv, X. Zhao, Z. Li, D. Cao, H. Liu, C. Zhang, S. Xu and B. Man, Qualitative and quantitative detection of microcystin-LR based on SERS-FET dual-mode biosensor, Biosens. Bioelectron., 2022, 212, 114434 CrossRef CAS PubMed.
- J.-A. Chou, C.-L. Chung, P.-C. Ho, C.-H. Luo, Y.-H. Tsai, C.-K. Wu, C.-W. Kuo, Y.-S. Hsiao, H.-H. Yu and P. Chen, Organic Electrochemical Transistors/SERS-Active Hybrid Biosensors Featuring Gold Nanoparticles Immobilized on Thiol-Functionalized PEDOT Films, Front. Chem., 2019, 7, 281 CrossRef CAS PubMed.
- A. M. Münzer, W. Seo, G. J. Morgan, Z. P. Michael, Y. Zhao, K. Melzer, G. Scarpa and A. Star, Sensing Reversible Protein-Ligand Interactions with Single-Walled Carbon Nanotube Field-Effect Transistors, J. Phys. Chem. C, 2014, 118(31), 17193–17199 CrossRef PubMed.
- K. Yuan, Q. Qian, M. Wu, B. Wang, S. Zeng, D. Chen, M. D. Birowosuto, D. S. Ang and C. Gu, A WO(x)/MoO(x) hybrid oxide based SERS FET and investigation on its tunable SERS performance, Phys. Chem. Chem. Phys., 2024, 26(14), 10814–10823 RSC.
- A. B. Nowicka, M. Czaplicka, T. Szymborski and A. Kamińska, Combined negative dielectrophoresis with a flexible SERS platform as a novel strategy for rapid detection and identification of bacteria, Anal. Bioanal. Chem., 2021, 413(7), 2007–2020 CrossRef CAS PubMed.
- C. L. Nogueira, J. C. Sousa, R. Afonso, D. R. Santos, P. Rodrigues, L. Albuquerque, A. Faustino, S. Abalde-Cela, P. P. Freitas, S. Cardoso and C. M. Carvalho, Combining droplet microfluidics and magnetoresistive sensors for the rapid and quantitative detection of Klebsiella pneumoniae in urinary tract infections, Sens. Actuators, B, 2024, 404, 135175 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.