Next Article in Journal
Photodynamic Therapy Review: Past, Present, Future, Opportunities and Challenges
Previous Article in Journal
Ultrafast Excited State Dynamics of a Verdazyl Diradical System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metallic Nanoparticles for Surface-Enhanced Raman Scattering Based Biosensing Applications

1
Department of Biosciences and Bioengineering, Indian Institute of Technology Dharwad, Dharwad 580011, India
2
Department of Chemistry, Gaya College, Gaya 823001, India
*
Author to whom correspondence should be addressed.
Photochem 2024, 4(4), 417-433; https://doi.org/10.3390/photochem4040026
Submission received: 8 August 2024 / Revised: 8 September 2024 / Accepted: 13 September 2024 / Published: 26 September 2024

Abstract

:
Surface-enhanced Raman scattering (SERS) is a powerful tool for biosensing with high sensitivity, selectivity, and capability of multiplex monitoring for both in vivo and in vitro studies. This has been applied for the identification and detection of different biological metabolites such as lipids, nucleic acids, and proteins. The present review article explores the vast applications of metallic nanoparticles for SERS-based biosensing. We have summarized and discussed the fundamental principles, theories, developments, challenges, and perspectives in the field of SERS-based biosensing using different metal nanoparticle substrates namely gold, silver, copper, and bimetallic nanoparticles.

1. Introduction

The rapid detection and identification of the biomolecule of interest in complex biological matrices comes with greater challenges such as sensitivity and reproducibility as biomolecules are complex and dynamic. One of the feasible label-free and non-destructive approaches for biosensing is aided by the advancement in Raman spectroscopy. Biological samples are natively surrounded by an aqueous environment. However, the Raman biomolecular fingerprint region is not masked by the presence of water. Raman spectroscopy is an analytical technique based on the Raman effect: “the inelastic scattering of photons”. When the light is incident on the sample, one in 106 photons undergoes inelastic scattering which makes it an inherently weak phenomenon. The Raman effect was discovered by Sir Chandrasekhara Venkata Raman in 1928 and published in a paper in Nature titled “A new type of secondary radiation” [1,2]. The conventional workflow of Raman spectroscopy involves the illumination of the sample with a monochromatic light source mostly in the visible to infrared range of the electromagnetic radiation. The inelastically scattered light is filtered and detected using a spectrometer. The output Raman spectra provide characteristic fingerprints of the sample giving out qualitative and quasi-quantitative information [3,4]. Additionally, it comes with versatility with its applicability to the physical state of the sample and the feasibility of multiplexing. However, the spontaneous Raman signals are often weak, which can negatively affect the detection limits. Further, these signals are also susceptible to fluorescence interference and background and system noise. The discovery of surface-enhanced Raman Scattering (SERS) has helped to enhance the sensitivity of Raman-spectroscopy-based biosensing for the detection of a wider range of biomolecules [5]. The SERS technique was first developed by Fleischmann et al. at the University of Southampton, UK. They reported that an adsorbed Pyridine monolayer on a roughened silver electrode showed an amplified Raman signal [2,6,7]. It is a well-known fact that the SERS effect can enhance the Raman signal intensity by 103 to 1012 times [2].
Two primary mechanisms contribute to the working principle of SERS, namely, electromagnetic enhancement (EME—108 to 1012 times) and chemical enhancement (CE—103 to 105 times) [2,8,9]. The SERS enhancement mechanism was explained and demonstrated by Van Duyne, Moskovits, Creighton, and their co-workers [7,10,11,12,13]. EME often occurs due to the excitation of localized surface plasmon resonance (LSPR), which involves the interaction of a strong electromagnetic field generated on the surface of the plasmonic nanostructure within the proximity of sample molecules. This LSPR excitation occurs at the resonating frequency of the incident light, making the EME enhancement mechanism light-wavelength-dependent [14]. Surface plasmons are the volume of oscillating conduction band electrons of metallic nanostructures [15]. When the frequency of incident light resonates with the plasmon frequency of a nanostructure, it leads to the oscillation of the surface plasmons, which induces a localized dipole in the nanostructure as shown in Figure 1. This induced dipole generates an enhanced electromagnetic field, which leads to another induced dipole on sample molecules present in its closed proximity, which in turn, results in the generation of a higher number of photons [2,7]. This whole phenomenon leads to the enhancement of Raman signal intensities, which is dependent on mainly two factors, the strength of the incident electric field and the polarizability of metallic nanostructures [2]. Another SERS enhancement mechanism is CE, which involves the transfer of electrons between the sample molecule and metallic nanostructures that are in close contact or a few angstroms apart. As it is a physio-chemical interaction, the effect is short-range and contributes much less enhancement compared to EME [7]. When the incident light falls on the nanostructure, an excited valence band electron jumps to the conduction band, leaving an empty hole in the valence band of the nanostructure. Due to resonant tunneling, this excited electron present in the conduction band of the nanostructure rapidly transfers itself to a similar energy level above the lowest unoccupied molecular orbital (LUMO) of the sample molecule. However, the electron will transfer back to its original position to fill the empty hole in the valence band in the nanostructure. This process will transfer energy exciting the sample molecules to a vibrational level, which results in the emergence of a Raman photon [7]. Due to this resonance charge transfer and radiated Raman photon, amplification of Raman scattered signal intensities and a noticeable shift in the spectra are observed. This shift occurs due to chemisorption on the geometrical and/or electronic structure of the adsorbed sample molecules [2,16,17]. The schematic diagram presented in Figure 1 represents a pictorial understanding of SERS enhancement mechanisms.

2. SERS-Based Biosensing

SERS-based biosensing is a rapid detection and identification technique that involves the interaction of biomolecules with minimally invasive biocompatible SERS substrates resulting in a multi-fold enhanced sensitivity of the Raman signal and an increased multiplex biomolecular limit of detection [7,14]. Most biological systems are inherently aqueous systems, but due to the very small Raman scattering cross-section of water, Raman signals are minimally perturbed by the background signal from water [14,15,16,17,18,19]. The most widely used SERS substrates for biosensing are metal nanostructures of gold (Au) and silver (Ag), which are noble metals. Another commonly used SERS substrate is copper (Cu), which is comparably reactive but higher in abundance and less expensive [7,14]. Au is chemically more stable and biocompatible, therefore, it is preferred for intracellular/in vivo studies, whereas Ag is less preferred as it is toxic to living organisms [7,20,21,22,23]. The concept of nanotechnology was first introduced by Norio Taniguchi in 1974 [24]. Metal nanostructures/nanoparticles (MNs) are metallic nanoscale particles which are of different sizes and shapes such as nanowires, nanorods, nanoprism, nanostars, nanoflowers, nanospheres, and others [7,10,14,25]. The shape and size of MNs play an important role in developing interaction between adsorbed sample molecules and the plasmonic surface of the nanostructure. Hence, fabricating and designing substrates of appropriate shape and size is crucial for SERS-based biosensing performance. When MN size is greater than 100 nm, multipole excitation occurs instead of an induced dipole, which results in redistribution of total charge within the substrate, leading to non-radiative mode charge transfer decreasing the SERS enhancement. Similarly, shape also plays an important role in the surface volume ratio for the substrate–sample interaction [2,7]. Another important aspect of SERS-based biosensing is the synthesis methods of the nanostructures. For the synthesis of gold nanoparticles (AUNPs), broadly performed methodologies are seed growth, one-step synthesis, and electron beam etching [26,27,28,29]. Like AuNPs, Silver nanoparticles (AgNPs) can be of different shapes and forms such as nanocubes [30,31], nanowires [32], flowered-shaped nanoparticles, and others [33,34]. The efficacy of SERS-based biosensors can be determined by assessing their sensitivity, specificity, and reproducibility. This approach has been used in a wide range of biosensing applications such as pesticide detection [35,36,37,38,39], the identification, characterization, and detection of bacteria and other pathogens [40,41,42,43,44], the detection of other biomolecules [45,46,47,48], and disease diagnosis [49,50,51,52,53]. In this review, we have discussed SERS substrates for biosensing applications where a SERS substrate refers to the metallic structure, which can be metallic nanoparticles in solution (such as water) also known as metallic colloids (colloidal solutions) and planar metallic structure such as metallic nanoparticles arranged in an array on a planar substrate [54]. In the laboratory, SERS is mostly carried out using metallic colloids as these are easy to produce; however, at the same time, a stabilizing agent is required to avoid aggregation. The simplest way to make a planar SERS substrate is by drying the metallic colloidal solution on a suitable substrate, however, this may result in large spatial inhomogeneities [54]. Potential biosensing applications of different metallic nanoparticles are shown in Table 1.

3. Gold-Based Nanomaterials for SERS and Their Applications in Biosensing

Plasmonic nanoparticles have the potential to enhance the Raman signal intensities of the Raman active mode of molecular bonds. Noble metal nanoparticles including Au, Ag, and Cu are typically used as SERS substrates [76,77,78,79,80]. The universality of SERS measurements is hindered by the expensive preparation and disposal problems of these substrates. Anisotropic gold, silver, and copper nanoparticles (NPs) with sharp corners and edges are beneficial for developing biosensors based on MEF and SERS due to the amplified plasmonic effect with incident light frequency in the visible to near-infrared regions [81]. Depending on reaction conditions, it is possible to produce gold nanoparticles (AuNPs) of various sizes and shapes. Based on the shape and size of the AuNPs, they exhibit a unique (LSPR) peak in the visible to NIR range (>500 nm). The redshift of the LSPR peak increments with the aspect ratio, i.e., the length-to-diameter ratio, of the AuNPs. Surface functionalization of metal nanostructures is essential for developing nanoparticle-based biosensors, also known as nano biosensors, as it enables target-specific detection. To prepare the bio-metal detecting nanoprobes, various biofunctionalization and assembling techniques are performed. Indeed, varied sizes, shapes, compositions, aggregations, and coatings can be used to create SERS-active nanostructures, opening the door to response customization for detection needs. The SERS substrate synthesized and fabricated for an incident light source in the near-infrared region minimizes the photodamage to biological samples and prevents auto-fluorescence, which can mask the SERS signal. When we compare AuNPs to other nanomaterial counterparts such as metal-oxide- or carbon-based nanomaterials, AuNPs are widely used for biosensing applications due to their biocompatibility, nontoxic nature, optical properties, electronic properties, and tunability. According to the literature, nanoparticles between 10 and 100 nm in size are more suited for SERS investigations as the size and shape of the nanoparticles play important roles, such as in the occurrence of a multipole, which is a non-radiative mode, rather than a dipole [82,83]. Another important factor to consider for enhancing SERS performance in suspensions of metal nanoparticles is the shape of the nanoparticle. Shapes like nanostars, nanorods, nanocubes, nanoplates, nanostars nano prisms, and others are used for biosensing [84,85].
AuNPs offer higher chemical stability and excellent biocompatibility and are well-suited for excitation in the visible and near-infrared regions [20,21]. Qian et al. reported that active targeting achieved more nanoparticle, specifically gold nanoparticle, accumulation in human tumor xenografts [86]. To confirm the nanoparticle accumulation in the tumor, gold nanoparticles functionalized with Raman labels were employed for SERS imaging in this investigation. Gold ion salts were used to synthesize the nanoparticles, which are functionalized with ScFv B10, which is an antibody fragment known to have specificity for human EGFR; the synthesis was carried out in such a way as to achieve the size that was best for SERS. The cells, namely, human non-small cell lung carcinoma NCIH520, which are EGFR-negative cancer cells, and Tu686, which are EGFR-positive cancer cells, were first incubated with the nanoparticles and the SERS analysis confirmed the uptake of nanoparticles by the EGFR-positive cells. However, there was no discernible SERS signal in the EGFR-negative cells. The authors switched to a mouse model after verifying the detection of the SERS signal from nanoparticles absorbed by the cancer cells. Injection of Tu686 cells into the flank of naked mice was performed, and strong SERS signals were visible in the resulting tumors after targeted nanoparticle injection, indicating that the probe had successfully reached the tumor, as shown in Figure 2. Meanwhile, tumors in the mice receiving injections of non-targeted nanoparticles did not exhibit any detectable SERS signal. The in vivo molecular imaging investigation of tumor biomarkers utilizing SERS was presented by the authors [86].
Fabris et al. presented a method for synthesizing a Au-nanostar (AuNS)-functionalized substrate, which can be applied for highly sensitive SERS-based sensing [87]. The AuNSs were fixed on the Au support using a Raman-silent organic tether, which allows analytes to be either chemisorbed or physisorbed on the AuNSs for sensing. The developed SERS substrate enables the quasi-quantitative and multiplexed detection and identification of analytes from mixtures. The SERS substrates can detect chemisorbed 4-mercaptobenzoic acid at a concentration as low as 10 fM with a SERS enhancement factor of 109 with high reproducibility. Most crucially, they offer a high signal-to-noise ratio when detecting physisorbed analytes, such as crystal violet [87]. Bhattacharya et al. [88] reported gold nanoflowers used as a biosensor in imaging applications. Raman signal amplification using the nanoflower (AuNf) petals is crucial for achieving signal improvements of the order of 106. The electromagnetic enhancement mechanism of metal nanoparticles can be used to explain this improvement. The toxicity of the nanostructures to the cells was evaluated in A549 human lung carcinoma cells by performing a colorimetric assay known as an MTT assay. Bamrungsap et al. [89] developed a layer-by-layer immersion technique that was used to create a plasmonic paper utilizing cellulose substrates and a composite of graphene oxide and gold nanorods (GO-AuNRs) with rhodamine 6G (R6G) was used as the probe molecule. When compared to a paper substrate coated with only AuNRs, the GO-AuNR plasmonic paper composite showed better signal intensity. The electromagnetic enhancement from the AuNRs and the chemical enhancement effect from the GO worked in concert to boost the signal. The optimal condition for developing plasmonic paper able to evaluate the highest SERS signals produced from R6G molecules was by using GO pretreatment (2 mg/mL) followed by two immersion cycles into the solution of AuNRs. The calculated enhancement factor of the plasmonic paper was reported to be 4.56 × 107, while a limit of detection of 0.1 nM R6G was achieved. The plasmonic paper was also evaluated for its ability to detect the anti-cancer medication Mitoxantrone (MTX). This straightforward and cost-effective plasmonic-paper-based SERS application is intended for trace analysis and medication monitoring.

4. Silver-Based Nanomaterials for SERS and Their Applications in Biosensing

The unique and distinctive optical properties of silver (Ag), which minimize inter-band transitions, favor plasmonic resonance and offer a high SERS enhancement factor, making silver a good SERS substrate [90]. Either chemical interactions or physical forces are used to fine-tune the AgNPs’ shape, size, and assembly. Tan et al. [91] have evaluated the beneficial use of AgNPs in clinical biosensing applications. In general, enhanced-performance modified AgNPs are proposed as biosensing platforms for conditions such as viral or bacterial infections and cancer. AgNPs are a reliable choice for improving the SERS signal and enhancing detection sensitivity. Robust SERS substrates should possess two main properties: (i) numerous “hot spots” in spaces between adjacent metal NPs, and (ii) a strong adsorption ability [92]. In this instance, a variety of hybrid nanomaterials, such as AuNPs@Q-Sepharose microsphere [93], AgNPs@polymethyl methacrylate film [94], AgNPs/agar/PAN electrospun nanofibers [95], AgNPs@graphene [96], AuNPs/WS2 nano-dome/graphene [97], and others have been developed and used as SERS-active substrates. However, the majority of these composite nanomaterials were unable to efficiently trap target molecules as they approached the surface of the metal NPs. A novel SERS biosensor based on AgNPs@MOFs capable of mimicking peroxidase was developed for the first time for blood cholesterol monitoring by Wu et al. [98]. The development of AgNPs on the surface of MOF produced hybrid AgNPs@MOF nanomaterials, which were used as the SERS substrate. A SERS biosensor for blood cholesterol detection with a limit of detection of 0.36 μM was demonstrated. Further, the sensor uses metal NPs@Fe-based MOFs, which act as both a SERS-active substrate and as a peroxidase-mimicking nanozyme. The monitoring of cholesterol with high sensitivity, specificity, and accuracy shows a significant and promising potential for a translational SERS-based blood cholesterol sensing approach [98]. Le et al. reported a SERS application based on an easy approach for developing a SERS-active substrate with a two-dimensional macroporous Ag film, which is made of silver nanosheet (AgNS)-coated inverted opal film. This substrate has good SERS reproducibility with a high enhancement factor of 6 × 107. The detection of rhodamine 6G and the detection of DNA in a label-free manner was achieved. The 2D AgNS-coated inverse opal film is a high-performance SERS-active substrate with potential for biosensing applications [99]. In 2013, Fang et al. [100] demonstrated how silver nanocubes (AgNC) self-assemble into dense bowl-shaped arrays using a polystyrene nanosphere (PSNS) template. The group discovered a facet-to-facet arrangement in most of the AgNCs and the SERS hot spot was localized to the cubes’ corners. This SERS substrate has strong amplification capability due to the densely organized AgNCs, enabling single molecule detection, which was demonstrated utilizing the bi-analyte Raman method [100]. Although SERS-based biosensing has shown significant promise in laboratory settings, its practical applicability is restricted due to the short life span of commonly used SERS-active substrates based on AgNPs. The development of a novel, cost-effective substrate made from AgNPs shielded by tiny nitrogen-doped graphene quantum dots (Ag NP@N-GQD), and its systematic evaluation for glucose sensing is an attempt at SERS-based field biosensing. The Raman amplification was significantly stronger on the new substrate in comparison to the pure Ag substrate. More significantly, unlike pure AgNPs, this substrate maintained its SERS capability for at least 30 days in a typical indoor environment, sustaining performance in both the wet and dry phases. The Ag NP@N-GQD thin film was proven to be an efficient SERS substrate for glucose detection in dried mouse blood samples details of which are mentioned in Table 2 [101].
The in-situ production of Ag colloids is another quick and highly sensitive method that in the future can be employed in point-of-care diagnostics [102,103,104]. This is gaining popularity as an adjunct approach in comparison to its widely known counterpart, which is carried out by exposing pre-made AgNPs to a biomass of bacteria, as the latter method’s limited spectrum consistency makes it challenging to probe the bacterial fingerprints. However, a direct interaction between AgNPs and the bacteria enables the acquisition of very strong SERS spectra due to the strong electrostatic attraction between the positively charged Ag ions and the negatively charged bacterial components. Most recently, induced adherence to the cell wall in cyanobacteria was also described [102,103,104,105,106]. For the simultaneous identification of multiple diseases, label-free detection uses the intrinsic SERS signals seen in cell walls. The label-free (direct) SERS-based approach is sensitive and depends heavily on the SERS substrate, and using this method in biofluids might be difficult. Additionally, the spectra of many bacterial species can be strikingly similar. Therefore, it is necessary to deploy appropriate chemometrics techniques and discrimination models.

5. Copper-Based Nanomaterials for SERS and Their Applications in Biosensing

In the visible region, metallic nanoparticles like Ag, Au, and Cu display LSPR. The two biggest drawbacks of these nanomaterials when used for SERS enhancement are their high cost and limited stability. Likewise, corrosion and oxidation are two of Copper Nanoparticles’ (CuNPs) primary drawbacks despite their low and reasonable pricing. Stable CuNPs can be made using several quick and simple chemical processes [107,108]. However, surface contamination is unavoidable when reducing agents are present. These contaminations disrupt applications that involve surfaces, like SERS, as well as those that are biological or catalytic. Copper oxide NPs are an important category of SERS substrates. In 1998, Kudelski et al. [109] reported the SERS activity of nanostructured Cu2O, making it the second metal oxide to be recognized as a SERS-active substrate [110]. Lin et al. demonstrated highly uniform cubic, rhombic dodecahedral, and octahedral Cu2O microcrystals with well-defined 100, 110, and 111 facets [111,112]. In 2019, Aghdam et al. performed a SERS experiment using spherical Cu/Cu2O core-shell NPs using pulsed laser ablation (nanosecond Ce: Nd YAG) in liquid. The SERS activity of Cu/Cu2O core-shell NP substrates was determined using crystal violet (CV) and methylene blue (MB) and a green laser (532 nm) excitation source. The high enhancement of the SERS signal from the adsorbed analyte molecule was due to the Cu core showing the LSPR effect, resulting in EE, and the Cu2O shell possessing a rough surface, resulting in chemical enhancement. Furthermore, the degree of charge transfer was assessed by determining the intensities of the fully and partially symmetric modes. This analysis showed that in the SERS of the Cu/Cu2O-CV and Cu/Cu2O-MB, the LSPR increase predominates the charge-transfer resonance contribution, with the substrates demonstrating a 28% variation in the strength of the SERS signals [113]. For biological sensing applications, the detection of bacteria with copper nanomaterials is challenging in SERS platforms. Surface-enhanced Raman spectrum of Staphylococcus aureus a bacterium found in the respiratory tract using 785 nm is shown in Figure 3 [114]. Kowalska et al. developed the platforms used to generate the spectrum shown in Figure 3. The platforms are made of copper instead of the customary, costly Au or Ag. The platforms were developed using a brand-new, straightforward, economical, and quick high-pressure approach that involves the breakdown of copper hydride. A malachite green isothiocyanate standard was used to verify the platform enhancement factors. Depending on the pressure applied, the platforms display extraordinarily high SERS enhancement factors. The computed improvement factors were discovered to be between 1.5 and 4.6 × 106, and the determined value of the average spectral correlation coefficient was 0.82. Staphylococcus aureus germs were detected using these SERS platforms, which are innovative platforms that have the potential to replace more traditional silver or gold SERS substrates as effective instruments for biological or medical diagnosis. N.E. Markina et al. reported the presence of cephalosporin antibiotics in urine samples, determined using CuNP-based SERS with a lower detection limit than using other approaches [70]. To speed up the reaction and boost the homogeneity of the CuNPs, the synthesis of the CuNPs was tuned to maximize the analytical signal. The analytes of interest were ceftriaxone (CTR), cefazolin (CZL), and cefoperazone (CPR). The determination tests were conducted on samples of actual human urine that had been intentionally contaminated with substances at levels that corresponded to therapeutic drug monitoring (TDM) limits (50–500 μg mL−1). The universality of the suggested technique was ensured by performing sample dilution as a pretreatment. The minimum concentrations needed for TDM were all below the limits of detection, which were 7.5 (CTR), 8.8 (CZL), and 36 (CPR) μg mL−1. CuNPs showed a competitively high Raman enhancement efficiency (for excitation at 638 nm) compared to Ag and Au nanoparticles.
For real-world SERS applications, Cu chips are more cost-effective compared to Ag and Au chips, but the poor stability and weak SERS of Cu substrates are a big concern. Dai et al. reported a process to develop Cu-based SERS chips that are highly sensitive as well as stable [115]. The Cu-coated fabric was used as a SERS chip to detect crystal violet with LOD as low as 10−8 and an enhancement factor of 2.0 × 106 M. Furthermore, the Cu membrane on the fabric, which is hydrophobic, also plays an important role by reducing background noise and providing stability. After air storage of nearly 2 months, the chip’s SERS intensity decreased to 18% of its initial level. A replacement reaction was used to introduce Ag onto the pristine Cu surface to improve the SERS performance of the Cu chips. The Ag-modified Cu chips showed a better enhancement factor of 7.6 × 106 and a three-orders-of-magnitude reduction in the LOD of CV to 10−11 M. This occurred due to the plasmonic coupling between Cu and Ag at the nanoscale. The SERS intensity of the Cu–Ag chip maintained 80% of its initial intensity after two months of storage because of the additional protection provided by the Ag shell [115].

6. Bimetallic Nanomaterials for SERS and Their Applications in Biosensing

The enhancement of SERS signal efficiency and reproducibility can be achieved by improving and developing effective substrates. As SERS is a phenomenon based on resonance, excitation condition matters [116]. Unlike monometallic nanoparticles (MNPs) such as gold, silver, copper, and others, as shown in Figure 4, bimetallic nanoparticles (BMNPs) are nanoparticles with two different metal elements such as gold–silver (Au–Ag), silver–copper (Ag–Cu), or gold–palladium (Au–Pd). Due to the presence of two metal elements, such nanoparticles possess better stability and better tunable plasmonic properties, which results in an enhanced SERS signal compared to their single-metal counterparts [2]. Liu et al. reported that, compared to the conventionally used AuNP substrates, nanoporous bimetallic gold and palladium nanostructures showed an enhanced SERS signal [117]. Two kinds of bimetallic nanoparticles are used to develop SERS substrates namely, alloyed nanoparticles and core–shell nanoparticles. In alloyed bimetallic nanoparticles, two metal atoms are homogenously mixed, resulting in the presence of both metals on the surface of the BMNPs. At room temperature, Au–Ag alloyed BMNPs have high spin electron relaxation, making them an effective substrate for room-temperature-based biosensing applications. Furthermore, Au–Ag alloys provide a significant increase in sensitivity, as this alloy is of lower density, increasing the signal-to-noise ratio and providing thermal and chemical stability. Conversely, the core–shell bimetallic configuration consists of one metal at the core and another metal surrounding the core known as the shell. The synthesis of core–shell BMNPs involves the synthesis of the core followed by the development of the shell. Au–Ag core–shell BMNPs of different shapes, such as spherical, cuboidal, and dumbbell-shaped, have been widely explored as SERS substrates, showing enhanced signal intensity [2,29,117,118]. Among Ag and Au, the plasmonic efficiency of Ag is higher, whereas the chemical stability of Au is higher. Studies have demonstrated that NPs made of either Au or Ag have lower SERS activity compared to Au–Ag BMNPs [117,119,120,121,122,123,124]. To synthesize the spherical Au–Ag BMNPs, the co-reduction method is commonly used in which Ag+ and Au3+ ions are reduced simultaneously. As reported earlier, this method has also been used for developing other nanostructures with rod, wire, and triangular shapes. Furthermore, for developing mostly hollow alloy Ag–Au BMNPs, the galvanic replacement method by controlling reaction time is used, which is based on the reduction potential of Au3+ and Ag+. BMNPs also combine the potentials of both metals in terms of their chemical and plasmonic properties relative to pure metals [2,85,125,126]. This also indicates that the chemical nature of the metals should also be considered when developing SERS substrates. Pinholes present in the Au nanoshell’s surface act as hot spots for SERS by enhancing the local electric field [127]. This electric field amplifies the Raman signal. Mucin-1, a breast-cancer-specific protein biomarker, was detected with a detection limit of 4.3 aM using BMNPs, as shown in Figure 4D–F. Moreover, the presence of Mucin-1 decreases the hotspots by separating the aptamers and their complementary strands, leading to a reduction in the SERS signal [128,129].
A SERS-based immune assay has been developed with high sensitivity to diagnose Hepatitis B virus in human blood plasma using an active SERS substrate on an antibody-carrying microfluidic chip [29,130,131]. Influenza A virus detection was achieved using a multilayered Au–Ag-SERS-substrate-based immune sensor [132]. Cao et al. reported that Au–Ag core–shell BMNPs enable highly sensitive and multiplexed detection of oligonucleotides captured on a microarray chip. They further demonstrated a capability to distinguish six dissimilar DNA targets with six Raman-labelled nanoparticles and two RNA targets with single nucleotide polymorphism with a detection limit on the 20 femtomolar scale [133]. Au-Ag-BMNP-nanowire-substrate-based SERS was used for the rapid detection of lung-cancer-related microRNA (mi-196a) by coupling it with hairpin DNA. Raman signals were enhanced by creating hot spots between the Ag nanowires and AuNPs. This method of detection was presented as an possible alternative to conventional techniques for microRNA detection i.e., RT-PCR (real-time polymerase chain reaction), with detection limits of 96.58 aM in the PBS and 130 aM in the serum, respectively [134]. Highly reproducible detection and discrimination with low power and short acquisition time of three bacteria, Escherichia coli, Salmonella typhimurium, and Bacillus subtilis was demonstrated using SERS and a Ag–Au BMNP substrate [135]. Other transition metals such as platinum, palladium, rhodium, ruthenium, cobalt, and nickel-coated Au core have also been reported [2].

7. Conclusions and Perspective

In this review, we have described the principles of SERS and different metallic nanoparticles and their application in the biological domain. The SERS technique has been widely explored and applied in many fields such as biomedical, drug and pharmaceutical, molecular, microbiological, and other research fields. SERS combines Raman spectroscopy with nanotechnology, turning it into a remarkable optoelectronics biosensor due to its fast detection with higher sensitivity. The goal of SERS-based biosensing is to improve the limit of detection. SERS probes have poor stability with respect to temperature fluctuations, making them of limited suitability. Storage is also an issue, as long-term storage makes the NPs susceptible to agglomeration. However, stability can be increased using encapsulation by polymers. Plasmonic metals are better in terms of SERS signal compared to semiconducting and dielectric materials but lag behind in terms of stability and storage. Challenges in terms of SERS-based biosensors come with developing highly sensitive and reusable SERS substrates. Many biosensing applications have already been developed and reported, but they still lack commercialization. The reproducibility of SERS substrates requires more improvement as their performance with narrow plasmons is influenced by even slight changes in the nanostructure, which leads to a significant LSPR shift. SERS-based biosensing also suffers from the presence of background molecules or impurities along with the biomolecule of interest in the physiological environment. Therefore, in terms of practical applications, much improvement is required. Optimization of sample preparation strategies is very crucial to SERS-based biosensing to ensure that the biomolecule of interest has a preference for the SERS-active substrate, minimizing the competition for binding sites and increasing the sensitivity. The conventional approach is monitoring and adjusting pH, ionic strength, dilution, and surface modification. Combining SERS biosensing with advanced chemometrics and machine learning algorithms for processing is one of the feasible approaches for determining the SERS signal of interest. For SERS-based intra-cellular biomolecule dynamic biosensing, the stability and toxicity of the SERS probe need to be extensively optimized and critically evaluated to understand biocompatibility. For analyzing the biocompatibility and choosing the SERS substrate, it is always better to analyze the size and chemical structure of the analyte and matrix complexity. The key factor influencing the optical setup of a SERS biosensor for analyte detection is the excitation source wavelength, which is highly dependent on the choice of metallic nanomaterial. For instance, the resonance frequency of silver nanoparticles is blue-shifted compared to a similar nanoparticle made of gold nanoparticles. Furthermore, the enhancement factor of the SERS signal is highly influenced by the shape and size of the colloidal metal nanoparticles. Therefore, different advanced strategies for improving SERS-based biosensing for both qualitative and quantitative analysis in the fields of bioimaging, disease diagnosis, microbiology, and molecular biology should be the focus of future research. Another limitation is the masking of the SERS signal by fluorescence, which has been mostly dealt with by the use of confocal SERS, an NIR excitation source, or quenching of the source of the fluorescence by metallic nanoparticles. Furthermore, Raman spectrometer instrumentation cost and the need for sophisticated instruments also limit SERS biosensor applicability. Rapid detection and diagnosis along with lower detection limits are one research interest in the biomedical field and must be improved for translation. In aiming for translation, the development of easy-to-use hardware and software, alongside higher sensitivity, selectivity, reproducibility, and cost-effective approaches, needs to be actively kept in mind. Further qualitative and quantitative multiplex analysis of biomolecules or analytes in complex mixture solutions can be the next step towards advanced in situ biosensing applications. SERS-based biosensing has promising potential with the developments in plasmonic nanoparticles and different machine-learning tools. The development of low-cost, easily operated, and miniaturized smart SERS biosensors could lead to these becoming point-of-care testing tools in the future.

Author Contributions

J.K.: outline, editing the original draft, writing—Abstract Introduction, Bimetallic nanoparticles for SERS and applications in biosensing, conclusion and perspective; S.S.R.: writing—Gold/Silver/Copper-based Nanomaterials for SERS and application in biosensing; M.S.P.: editing and reviewing the original draft. A.V.: editing and reviewing the original draft; S.P.S.: supervision, outline, writing and editing the original manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Raman, C.V.; Krishnan, K.S. A New Type of Secondary Radiation. Nature 1928, 121, 501–502. [Google Scholar] [CrossRef]
  2. Bora, T. Recent Developments on Metal Nanoparticles for SERS Applications. In Noble and Precious Metals—Properties, Nanoscale Effects and Applications; InTech Open: Rijeka, Croatia, 2018. [Google Scholar]
  3. Karlo, J.; Dhillon, A.K.; Siddhanta, S.; Singh, S.P. Reverse Stable Isotope Labelling with Raman Spectroscopy for Microbial Proteomics. J. Biophotonics 2024, 17, e202300341. [Google Scholar] [CrossRef]
  4. Karlo, J.; Dhillon, A.K.; Siddhanta, S.; Singh, S.P. Monitoring of Microbial Proteome Dynamics Using Raman Stable Isotope Probing. J. Biophotonics 2022, 6, e202200341. [Google Scholar] [CrossRef] [PubMed]
  5. Fornasaro, S.; Alsamad, F.; Baia, M.; Batista de Carvalho, L.A.E.; Beleites, C.; Byrne, H.J.; Chiadò, A.; Chis, M.; Chisanga, M.; Daniel, A.; et al. Surface Enhanced Raman Spectroscopy for Quantitative Analysis: Results of a Large-Scale European Multi-Instrument Interlaboratory Study. Anal. Chem. 2020, 92, 4053–4064. [Google Scholar] [CrossRef] [PubMed]
  6. Fleischmann, M.; Hendra, P.J.; McQuillan, A.J. Raman Spectra of Pyridine Adsorbed at a Silver Electrode. Chem. Phys. Lett. 1974, 26, 163–166. [Google Scholar] [CrossRef]
  7. Das, G.M.; Managò, S.; Mangini, M.; De Luca, A.C. Biosensing Using SERS Active Gold Nanostructures. Nanomaterials 2021, 11, 2679. [Google Scholar] [CrossRef] [PubMed]
  8. Le Ru, E.C.; Blackie, E.; Meyer, M.; Etchegoin, P.G. Surface Enhanced Raman Scattering Enhancement Factors: A Comprehensive Study. J. Phys. Chem. C 2007, 111, 13794–13803. [Google Scholar] [CrossRef]
  9. Karlo, J.; Dhillon, A.K.; Razi, S.S.; Siddhanta, S.; Singh, S.P. Imaging Based Raman Spectroscopy. In Raman Spectroscopy; Springer: Singapore, 2024; pp. 349–375. [Google Scholar]
  10. Moskovits, M. Surface Roughness and the Enhanced Intensity of Raman Scattering by Molecules Adsorbed on Metals. J. Chem. Phys. 1978, 69, 4159–4161. [Google Scholar] [CrossRef]
  11. Albrecht, M.G.; Creighton, J.A. Anomalously Intense Raman Spectra of Pyridine at a Silver Electrode. J. Am. Chem. Soc. 1977, 99, 5215–5217. [Google Scholar] [CrossRef]
  12. Yu, X.; Cai, H.; Zhang, W.; Li, X.; Pan, N.; Luo, Y.; Wang, X.; Hou, J.G. Tuning Chemical Enhancement of SERS by Controlling the Chemical Reduction of Graphene Oxide Nanosheets. ACS Nano 2011, 5, 952–958. [Google Scholar] [CrossRef]
  13. Jeanmaire, D.L.; Van Duyne, R.P. Surface Raman Spectroelectrochemistry. J. Electroanal. Chem. Interfacial Electrochem. 1977, 84, 1–20. [Google Scholar] [CrossRef]
  14. Bantz, K.C.; Meyer, A.F.; Wittenberg, N.J.; Im, H.; Kurtuluş, Ö.; Lee, S.H.; Lindquist, N.C.; Oh, S.-H.; Haynes, C.L. Recent Progress in SERS Biosensing. Phys. Chem. Chem. Phys. 2011, 13, 11551. [Google Scholar] [CrossRef]
  15. Zhan, C.; Chen, X.-J.; Yi, J.; Li, J.-F.; Wu, D.-Y.; Tian, Z.-Q. From Plasmon-Enhanced Molecular Spectroscopy to Plasmon-Mediated Chemical Reactions. Nat. Rev. Chem. 2018, 2, 216–230. [Google Scholar] [CrossRef]
  16. Jensen, L.; Aikens, C.M.; Schatz, G.C. Electronic Structure Methods for Studying Surface-Enhanced Raman Scattering. Chem. Soc. Rev. 2008, 37, 1061. [Google Scholar] [CrossRef] [PubMed]
  17. Ando, J.; Fujita, K.; Smith, N.I.; Kawata, S. Dynamic SERS Imaging of Cellular Transport Pathways with Endocytosed Gold Nanoparticles. Nano Lett. 2011, 11, 5344–5348. [Google Scholar] [CrossRef]
  18. Karlo, J.; Prasad, R.; Singh, S.P. Biophotonics in Food Technology: Quo Vadis? J. Agric. Food Res. 2023, 11, 100482. [Google Scholar] [CrossRef]
  19. Karlo, J.; Gupta, A.; Singh, S.P. In Situ Monitoring of the Shikimate Pathway: A Combinatorial Approach of Raman Reverse Stable Isotope Probing and Hyperspectral Imaging. Analyst 2024, 149, 2833–2841. [Google Scholar] [CrossRef]
  20. Yigit, M.V.; Medarova, Z. In Vivo and Ex Vivo Applications of Gold Nanoparticles for Biomedical SERS Imagingi. Am. J. Nucl. Med. Mol. Imaging 2012, 2, 232–241. [Google Scholar]
  21. Laing, S.; Jamieson, L.E.; Faulds, K.; Graham, D. Surface-Enhanced Raman Spectroscopy for in Vivo Biosensing. Nat. Rev. Chem. 2017, 1, 0060. [Google Scholar] [CrossRef]
  22. Stamplecoskie, K.G.; Scaiano, J.C.; Tiwari, V.S.; Anis, H. Optimal Size of Silver Nanoparticles for Surface-Enhanced Raman Spectroscopy. J. Phys. Chem. C 2011, 115, 1403–1409. [Google Scholar] [CrossRef]
  23. Guo, H.; Xing, B.; White, J.C.; Mukherjee, A.; He, L. Ultra-Sensitive Determination of Silver Nanoparticles by Surface-Enhanced Raman Spectroscopy (SERS) after Hydrophobization-Mediated Extraction. Analyst 2016, 141, 5261–5264. [Google Scholar] [CrossRef]
  24. Khodashenas, B.; Ghorbani, H.R. Synthesis of Silver Nanoparticles with Different Shapes. Arab. J. Chem. 2019, 12, 1823–1838. [Google Scholar] [CrossRef]
  25. Tian, F.; Bonnier, F.; Casey, A.; Shanahan, A.E.; Byrne, H.J. Surface Enhanced Raman Scattering with Gold Nanoparticles: Effect of Particle Shape. Anal. Methods 2014, 6, 9116–9123. [Google Scholar] [CrossRef]
  26. Niu, W.; Chua, Y.A.A.; Zhang, W.; Huang, H.; Lu, X. Highly Symmetric Gold Nanostars: Crystallographic Control and Surface-Enhanced Raman Scattering Property. J. Am. Chem. Soc. 2015, 137, 10460–10463. [Google Scholar] [CrossRef] [PubMed]
  27. Chatterjee, S.; Ringane, A.B.; Arya, A.; Das, G.M.; Dantham, V.R.; Laha, R.; Hussian, S. A High-Yield, One-Step Synthesis of Surfactant-Free Gold Nanostars and Numerical Study for Single-Molecule SERS Application. J. Nanoparticle Res. 2016, 18, 242. [Google Scholar] [CrossRef]
  28. Kedia, A.; Kumar, P.S. Halide Ion Induced Tuning and Self-Organization of Gold Nanostars. RSC Adv. 2014, 4, 4782–4790. [Google Scholar] [CrossRef]
  29. Awiaz, G.; Lin, J.; Wu, A. Recent Advances of Au@Ag Core–Shell SERS-based Biosensors. Exploration 2023, 3, 20220072. [Google Scholar] [CrossRef]
  30. Wiley, B.J.; Im, S.H.; Li, Z.-Y.; McLellan, J.; Siekkinen, A.; Xia, Y. Maneuvering the Surface Plasmon Resonance of Silver Nanostructures through Shape-Controlled Synthesis. J. Phys. Chem. B 2006, 110, 15666–15675. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, Q.; Li, W.; Wen, L.-P.; Chen, J.; Xia, Y. Facile Synthesis of Ag Nanocubes of 30 to 70 Nm in Edge Length with CF3 COOAg as a Precursor. Chem.-Eur. J. 2010, 16, 10234–10239. [Google Scholar] [CrossRef] [PubMed]
  32. Chen, H.M.; Liu, R.-S. Architecture of Metallic Nanostructures: Synthesis Strategy and Specific Applications. J. Phys. Chem. C 2011, 115, 3513–3527. [Google Scholar] [CrossRef]
  33. Zaheer, Z. Rafiuddin Multi-Branched Flower-like Silver Nanoparticles: Preparation and Characterization. Colloids Surf. A Physicochem. Eng. Asp. 2011, 384, 427–431. [Google Scholar] [CrossRef]
  34. Cai, X.; Zhai, A. Preparation of Microsized Silver Crystals with Different Morphologies by a Wet-Chemical Method. Rare Met. 2010, 29, 407–412. [Google Scholar] [CrossRef]
  35. Xu, M.-L.; Gao, Y.; Han, X.X.; Zhao, B. Detection of Pesticide Residues in Food Using Surface-Enhanced Raman Spectroscopy: A Review. J. Agric. Food Chem. 2017, 65, 6719–6726. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, K.; Gao, Y.; Fang, Z.; Li, H.; Feng, R.; Wang, Y.; Feng, Y.; Li, W.; Zhang, S.; Hu, L.; et al. SERS Detection for Pesticide Residue via a Single-Atom Sites Decoration Strategy. Appl. Surf. Sci. 2023, 621, 156832. [Google Scholar] [CrossRef]
  37. Moldovan, R.; Iacob, B.-C.; Farcău, C.; Bodoki, E.; Oprean, R. Strategies for SERS Detection of Organochlorine Pesticides. Nanomaterials 2021, 11, 304. [Google Scholar] [CrossRef]
  38. Augustine, S.; Sooraj, K.P.; Pachchigar, V.; Murali Krishna, C.; Ranjan, M. SERS Based Detection of Dichlorvos Pesticide Using Silver Nanoparticles Arrays: Influence of Array Wavelength/Amplitude. Appl. Surf. Sci. 2021, 544, 148878. [Google Scholar] [CrossRef]
  39. Sharma, V.; Krishnan, V. Fabrication of Highly Sensitive Biomimetic SERS Substrates for Detection of Herbicides in Trace Concentration. Sens. Actuators B Chem. 2018, 262, 710–719. [Google Scholar] [CrossRef]
  40. Allen, D.M.; Einarsson, G.G.; Tunney, M.M.; Bell, S.E.J. Characterization of Bacteria Using Surface-Enhanced Raman Spectroscopy (SERS): Influence of Microbiological Factors on the SERS Spectra. Anal. Chem. 2022, 94, 9327–9335. [Google Scholar] [CrossRef]
  41. Gukowsky, J.C.; He, L. Development of a Portable SERS Method for Testing the Antibiotic Sensitivity of Foodborne Bacteria. J. Microbiol. Methods 2022, 198, 106496. [Google Scholar] [CrossRef]
  42. Samek, O.; Bernatová, S.; Dohnal, F. The Potential of SERS as an AST Methodology in Clinical Settings. Nanophotonics 2021, 10, 2537–2561. [Google Scholar] [CrossRef]
  43. Lee, K.S.; Landry, Z.; Pereira, F.C.; Wagner, M.; Berry, D.; Huang, W.E.; Taylor, G.T.; Kneipp, J.; Popp, J.; Zhang, M.; et al. Raman Microspectroscopy for Microbiology. Nat. Rev. Methods Primers 2021, 1, 80. [Google Scholar] [CrossRef]
  44. Ahuja, T.; Ghosh, A.; Mondal, S.; Basuri, P.; Jenifer, S.K.; Srikrishnarka, P.; Mohanty, J.S.; Bose, S.; Pradeep, T. Ambient Electrospray Deposition Raman Spectroscopy (AESD RS) Using Soft Landed Preformed Silver Nanoparticles for Rapid and Sensitive Analysis. Analyst 2019, 144, 7412–7420. [Google Scholar] [CrossRef] [PubMed]
  45. Siddhanta, S.; Narayana, C. Surface Enhanced Raman Spectroscopy of Proteins: Implications for Drug Designing. Nanomater. Nanotechnol. 2012, 2, 1. [Google Scholar] [CrossRef]
  46. Li, Y.; Driver, M.; Decker, E.; He, L. Lipid and Lipid Oxidation Analysis Using Surface Enhanced Raman Spectroscopy (SERS) Coupled with Silver Dendrites. Food Res. Int. 2014, 58, 1–6. [Google Scholar] [CrossRef]
  47. Verma, M.; Naqvi, T.K.; Tripathi, S.K.; Kulkarni, M.M.; Prasad, N.E.; Dwivedi, P.K. Plasmonic Paper-Based Flexible SERS Biosensor for Highly Sensitive Detection of Lactic and Uric Acid. IEEE Trans. Nanobiosci. 2022, 21, 294–300. [Google Scholar] [CrossRef]
  48. Sharma, A.; Mondal, S.; Ahuja, T.; Karmakar, T.; Siddhanta, S. Ion-Mediated Protein Stabilization on Nanoscopic Surfaces. Langmuir 2023, 39, 1227–1237. [Google Scholar] [CrossRef] [PubMed]
  49. Sunil, N.; Unnathpadi, R.; Pullithadathil, B. Label-Free SERS Salivary Biosensor Based on Ni@Ag Core–Shell Nanoparticles Anchored on Carbon Nanofibers for Prediagnosis of Lung Cancer. ACS Appl. Nano Mater. 2023, 6, 11334–11350. [Google Scholar] [CrossRef]
  50. Beeram, R.; Vepa, K.R.; Soma, V.R. Recent Trends in SERS-Based Plasmonic Sensors for Disease Diagnostics, Biomolecules Detection, and Machine Learning Techniques. Biosensors 2023, 13, 328. [Google Scholar] [CrossRef]
  51. Carlomagno, C.; Cabinio, M.; Picciolini, S.; Gualerzi, A.; Baglio, F.; Bedoni, M. SERS-based Biosensor for Alzheimer Disease Evaluation through the Fast Analysis of Human Serum. J. Biophotonics 2020, 13, e201960033. [Google Scholar] [CrossRef] [PubMed]
  52. Beffara, F.; Perumal, J.; Puteri Mahyuddin, A.; Choolani, M.; Khan, S.A.; Auguste, J.; Vedraine, S.; Humbert, G.; Dinish, U.S.; Olivo, M. Development of Highly Reliable SERS-active Photonic Crystal Fiber Probe and Its Application in the Detection of Ovarian Cancer Biomarker in Cyst Fluid. J. Biophotonics 2020, 13, e201960120. [Google Scholar] [CrossRef] [PubMed]
  53. Kowalska, A.A.; Nowicka, A.B.; Szymborski, T.; Piecyk, P.; Kamińska, A. SERS-based Sensor for Direct L-selectin Level Determination in Plasma Samples as Alternative Method of Tumor Detection. J. Biophotonics 2021, 14, e202000318. [Google Scholar] [CrossRef]
  54. Le Ru, E.C.; Etchegoin, P.G. Metallic Colloids and Other SERS Substrates. In Principles of Surface-Enhanced Raman Spectroscopy; Elsevier: Amsterdam, The Netherlands, 2009; pp. 367–413. [Google Scholar]
  55. Vitol, E.A.; Brailoiu, E.; Orynbayeva, Z.; Dun, N.J.; Friedman, G.; Gogotsi, Y. Surface-Enhanced Raman Spectroscopy as a Tool for Detecting Ca2+ Mobilizing Second Messengers in Cell Extracts. Anal. Chem. 2010, 82, 6770–6774. [Google Scholar] [CrossRef] [PubMed]
  56. Kim, W.H.; Lee, J.U.; Song, S.; Kim, S.; Choi, Y.J.; Sim, S.J. A Label-Free, Ultra-Highly Sensitive and Multiplexed SERS Nanoplasmonic Biosensor for MiRNA Detection Using a Head-Flocked Gold Nanopillar. Analyst 2019, 144, 1768–1776. [Google Scholar] [CrossRef]
  57. Zengin, A.; Tamer, U.; Caykara, T. A SERS-Based Sandwich Assay for Ultrasensitive and Selective Detection of Alzheimer’s Tau Protein. Biomacromolecules 2013, 14, 3001–3009. [Google Scholar] [CrossRef] [PubMed]
  58. Kang, T.; Yoo, S.M.; Yoon, I.; Lee, S.Y.; Kim, B. Patterned Multiplex Pathogen DNA Detection by Au Particle-on-Wire SERS Sensor. Nano Lett. 2010, 10, 1189–1193. [Google Scholar] [CrossRef]
  59. Verde, A.; Mangini, M.; Managò, S.; Tramontano, C.; Rea, I.; Boraschi, D.; Italiani, P.; De Luca, A.C. SERS Sensing of Bacterial Endotoxin on Gold Nanoparticles. Front. Immunol. 2021, 12, 758410. [Google Scholar] [CrossRef]
  60. Kim, S.; Kim, T.G.; Lee, S.H.; Kim, W.; Bang, A.; Moon, S.W.; Song, J.; Shin, J.-H.; Yu, J.S.; Choi, S. Label-Free Surface-Enhanced Raman Spectroscopy Biosensor for On-Site Breast Cancer Detection Using Human Tears. ACS Appl. Mater. Interfaces 2020, 12, 7897–7904. [Google Scholar] [CrossRef] [PubMed]
  61. Huang, Y.-H.; Wei, H.; Santiago, P.J.; Thrift, W.J.; Ragan, R.; Jiang, S. Sensing Antibiotics in Wastewater Using Surface-Enhanced Raman Scattering. Environ. Sci. Technol. 2023, 57, 4880–4891. [Google Scholar] [CrossRef] [PubMed]
  62. Quarta, A.; Bettini, S.; Cuscunà, M.; Lorenzo, D.; Epifani, G.; Gigli, G.; Valli, L.; Aliyev, J.A.; Kazimov, E.E.; Bakhishova, M.J.; et al. Tailoring Gold Nanoisland-Based Biosensors for Ultrasensitive Detection of Doxorubicin in Biological Fluids. ACS Appl. Nano Mater. 2024, 7, 18724–18736. [Google Scholar] [CrossRef]
  63. Chen, H.; Park, S.-G.; Choi, N.; Kwon, H.-J.; Kang, T.; Lee, M.-K.; Choo, J. Sensitive Detection of SARS-CoV-2 Using a SERS-Based Aptasensor. ACS Sens. 2021, 6, 2378–2385. [Google Scholar] [CrossRef] [PubMed]
  64. Eremina, O.E.; Yarenkov, N.R.; Bikbaeva, G.I.; Kapitanova, O.O.; Samodelova, M.V.; Shekhovtsova, T.N.; Kolesnikov, I.E.; Syuy, A.V.; Arsenin, A.V.; Volkov, V.S.; et al. Silver Nanoparticle-Based SERS Sensors for Sensitive Detection of Amyloid-β Aggregates in Biological Fluids. Talanta 2024, 266, 124970. [Google Scholar] [CrossRef] [PubMed]
  65. Zeng, F.; Xu, D.; Zhan, C.; Liang, C.; Zhao, W.; Zhang, J.; Feng, H.; Ma, X. Surfactant-Free Synthesis of Graphene Oxide Coated Silver Nanoparticles for SERS Biosensing and Intracellular Drug Delivery. ACS Appl. Nano Mater. 2018, 1, 2748–2753. [Google Scholar] [CrossRef]
  66. Saridag, A.M.; Kahraman, M. Layer-by-Layer Coating of Natural Diatomite with Silver Nanoparticles for Identification of Circulating Cancer Protein Biomarkers Using SERS. Nanoscale 2023, 15, 13770–13783. [Google Scholar] [CrossRef]
  67. Jiang, G.; Xu, W.; Huang, X.; Dai, Z.; Zhou, C.; Hong, P.; Li, C. Detection of Bacillus cereus Spore Biomarkers Using SERS-Based Cuttlebone-Derived Organic Matrix/Silver Nanoparticles. ACS Sustain. Chem. Eng. 2023, 11, 4145–4154. [Google Scholar] [CrossRef]
  68. Butler, M.R.; Hrncirova, J.; Jacot, T.A.; Dutta, S.; Clark, M.R.; Doncel, G.F.; Cooper, J.B. Detection and Quantification of Antiviral Drug Tenofovir Using Silver Nanoparticles and Surface Enhanced Raman Spectroscopy (SERS) with Spatially Resolved Hotspot Selection. Front. Nanotechnol. 2023, 5, 1270474. [Google Scholar] [CrossRef]
  69. Ge, K.; Ni, R.; Tao, P.; Zhao, X.; Luo, Y.; Zhu, Y.; Song, B.; Zhang, W.; Dai, S.; Zhang, N.; et al. Low-Toxicity PEG-Coated SERS Silver Nanocubes for Diagnosing Esophageal Cancer. Opt. Commun. 2024, 554, 130183. [Google Scholar] [CrossRef]
  70. Markina, N.E.; Ustinov, S.N.; Zakharevich, A.M.; Markin, A.V. Copper Nanoparticles for SERS-Based Determination of Some Cephalosporin Antibiotics in Spiked Human Urine. Anal. Chim. Acta 2020, 1138, 9–17. [Google Scholar] [CrossRef]
  71. Halouzka, V.; Halouzkova, B.; Jirovsky, D.; Hemzal, D.; Ondra, P.; Siranidi, E.; Kontos, A.G.; Falaras, P.; Hrbac, J. Copper Nanowire Coated Carbon Fibers as Efficient Substrates for Detecting Designer Drugs Using SERS. Talanta 2017, 165, 384–390. [Google Scholar] [CrossRef] [PubMed]
  72. Fan, G.; Li, X.; Xu, S.; Dai, C.; Xue, Q.; Wang, H. SERS-Based Copper-Mediated Signal Amplification Strategy for Simple and Sensitive Detection of Telomerase Activity. Talanta 2021, 235, 122814. [Google Scholar] [CrossRef] [PubMed]
  73. Su, Y.; Wen, S.; Luo, X.; Xue, F.; Wu, S.; Yuan, B.; Lu, X.; Cai, C.; Jiang, L.-P.; Wu, P.; et al. Highly Biocompatible Plasmonically Encoded Raman Scattering Nanoparticles Aid Ultrabright and Accurate Bioimaging. ACS Appl. Mater. Interfaces 2021, 13, 135–147. [Google Scholar] [CrossRef] [PubMed]
  74. Lee, S.; Kim, S.; Choo, J.; Shin, S.Y.; Lee, Y.H.; Choi, H.Y.; Ha, S.; Kang, K.; Oh, C.H. Biological Imaging of HEK293 Cells Expressing PLCγ1 Using Surface-Enhanced Raman Microscopy. Anal. Chem. 2007, 79, 916–922. [Google Scholar] [CrossRef] [PubMed]
  75. Kaja, S.; Mathews, A.V.; Venuganti, V.V.K.; Nag, A. Bimetallic Ag–Cu Alloy SERS Substrates as Label-Free Biomedical Sensors: Femtomolar Detection of Anticancer Drug Mitoxantrone with Multiplexing. Langmuir 2023, 39, 5591–5601. [Google Scholar] [CrossRef]
  76. Sui, C.; Wang, K.; Wang, S.; Ren, J.; Bai, X.; Bai, J. SERS Activity with Tenfold Detection Limit Optimization on a Type of Nanoporous AAO-Based Complex Multilayer Substrate. Nanoscale 2016, 8, 5920–5927. [Google Scholar] [CrossRef] [PubMed]
  77. Choi, D.; Choi, Y.; Hong, S.; Kang, T.; Lee, L.P. Self-Organized Hexagonal-Nanopore SERS Array. Small 2010, 6, 1741–1744. [Google Scholar] [CrossRef]
  78. Shan, D.; Huang, L.; Li, X.; Zhang, W.; Wang, J.; Cheng, L.; Feng, X.; Liu, Y.; Zhu, J.; Zhang, Y. Surface Plasmon Resonance and Interference Coenhanced SERS Substrate of AAO/Al-Based Ag Nanostructure Arrays. J. Phys. Chem. C 2014, 118, 23930–23936. [Google Scholar] [CrossRef]
  79. Das, S.; Goswami, L.P.; Gayathri, J.; Tiwari, S.; Saxena, K.; Mehta, D.S. Fabrication of Low Cost Highly Structured Silver Capped Aluminium Nanorods as SERS Substrate for the Detection of Biological Pathogens. Nanotechnology 2021, 32, 495301. [Google Scholar] [CrossRef] [PubMed]
  80. Li, C.; Chen, M. Active Site-Dominated Electromagnetic Enhancement of Surface-Enhanced Raman Spectroscopy (SERS) on a Cu Triangle Plate. RSC Adv. 2020, 10, 42030–42037. [Google Scholar] [CrossRef] [PubMed]
  81. Bańkowska, M.; Krajczewski, J.; Dzięcielewski, I.; Kudelski, A.; Weyher, J.L. Au–Cu Alloyed Plasmonic Layer on Nanostructured GaN for SERS Application. J. Phys. Chem. C 2016, 120, 1841–1846. [Google Scholar] [CrossRef]
  82. Cong, S.; Liu, X.; Jiang, Y.; Zhang, W.; Zhao, Z. Surface Enhanced Raman Scattering Revealed by Interfacial Charge-Transfer Transitions. Innovation 2020, 1, 100051. [Google Scholar] [CrossRef]
  83. Terekhov, S.N.; Kachan, S.M.; Panarin, A.Y.; Mojzes, P. Surface-Enhanced Raman Scattering on Silvered Porous Alumina Templates: Role of Multipolar Surface Plasmon Resonant Modes. Phys. Chem. Chem. Phys. 2015, 17, 31780–31789. [Google Scholar] [CrossRef] [PubMed]
  84. Ham, J.; Yun, B.J.; Koh, W.-G. SERS-Based Biosensing Platform Using Shape-Coded Hydrogel Microparticles Incorporating Silver Nanoparticles. Sens. Actuators B Chem. 2021, 341, 129989. [Google Scholar] [CrossRef]
  85. Yaraki, M.T.; Tan, Y.N. Metal Nanoparticles-Enhanced Biosensors: Synthesis, Design and Applications in Fluorescence Enhancement and Surface-enhanced Raman Scattering. Chem. Asian J. 2020, 15, 3180–3208. [Google Scholar] [CrossRef]
  86. Qian, X.; Peng, X.-H.; Ansari, D.O.; Yin-Goen, Q.; Chen, G.Z.; Shin, D.M.; Yang, L.; Young, A.N.; Wang, M.D.; Nie, S. In Vivo Tumor Targeting and Spectroscopic Detection with Surface-Enhanced Raman Nanoparticle Tags. Nat. Biotechnol. 2008, 26, 83–90. [Google Scholar] [CrossRef] [PubMed]
  87. Indrasekara, A.S.D.S.; Meyers, S.; Shubeita, S.; Feldman, L.C.; Gustafsson, T.; Fabris, L. Gold Nanostar Substrates for SERS-Based Chemical Sensing in the Femtomolar Regime. Nanoscale 2014, 6, 8891–8899. [Google Scholar] [CrossRef] [PubMed]
  88. Patel, A.S.; Juneja, S.; Kanaujia, P.K.; Maurya, V.; Prakash, G.V.; Chakraborti, A.; Bhattacharya, J. Gold Nanoflowers as Efficient Hosts for SERS Based Sensing and Bio-Imaging. Nano-Struct. Nano-Objects 2018, 16, 329–336. [Google Scholar] [CrossRef]
  89. Ponlamuangdee, K.; Hornyak, G.L.; Bora, T.; Bamrungsap, S. Graphene Oxide/Gold Nanorod Plasmonic Paper—A Simple and Cost-Effective SERS Substrate for Anticancer Drug Analysis. New J. Chem. 2020, 44, 14087–14094. [Google Scholar] [CrossRef]
  90. Fernanda Cardinal, M.; Rodríguez-González, B.; Alvarez-Puebla, R.A.; Pérez-Juste, J.; Liz-Marzán, L.M. Modulation of Localized Surface Plasmons and SERS Response in Gold Dumbbells through Silver Coating. J. Phys. Chem. C 2010, 114, 10417–10423. [Google Scholar] [CrossRef]
  91. Tan, P.; Li, H.; Wang, J.; Gopinath, S.C.B. Silver Nanoparticle in Biosensor and Bioimaging: Clinical Perspectives. Biotechnol. Appl. Biochem. 2020, 68, 1236–1242. [Google Scholar] [CrossRef] [PubMed]
  92. Sun, Y.; Liu, K.; Miao, J.; Wang, Z.; Tian, B.; Zhang, L.; Li, Q.; Fan, S.; Jiang, K. Highly Sensitive Surface-Enhanced Raman Scattering Substrate Made from Superaligned Carbon Nanotubes. Nano Lett. 2010, 10, 1747–1753. [Google Scholar] [CrossRef]
  93. Ma, Z.; Zhai, Y.; Chen, Z.; Yang, H.; Wang, F. Preparation of QSS@AuNPs and Solvent Inducing Enhancement Strategy for Raman Determination of Salivary Thiocyanate. ACS Appl. Mater. Interfaces 2021, 13, 5966–5974. [Google Scholar] [CrossRef]
  94. Wang, T.-J.; Barveen, N.R.; Liu, Z.-Y.; Chen, C.-H.; Chou, M.-H. Transparent, Flexible Plasmonic Ag NP/PMMA Substrates Using Chemically Patterned Ferroelectric Crystals for Detecting Pesticides on Curved Surfaces. ACS Appl. Mater. Interfaces 2021, 13, 34910–34922. [Google Scholar] [CrossRef] [PubMed]
  95. Yang, T.; Yang, H.; Zhen, S.J.; Huang, C.Z. Hydrogen-Bond-Mediated In Situ Fabrication of AgNPs/Agar/PAN Electrospun Nanofibers as Reproducible SERS Substrates. ACS Appl. Mater. Interfaces 2015, 7, 1586–1594. [Google Scholar] [CrossRef] [PubMed]
  96. Cao, Z.; He, P.; Huang, T.; Yang, S.; Han, S.; Wang, X.; Ding, G. Plasmonic Coupling of AgNPs near Graphene Edges: A Cross-Section Strategy for High-Performance SERS Sensing. Chem. Mater. 2020, 32, 3813–3822. [Google Scholar] [CrossRef]
  97. Ghopry, S.A.; Alamri, M.; Goul, R.; Cook, B.; Sadeghi, S.M.; Gutha, R.R.; Sakidja, R.; Wu, J.Z. Au Nanoparticle/WS2 Nanodome/Graphene van Der Waals Heterostructure Substrates for Surface-Enhanced Raman Spectroscopy. ACS Appl. Nano Mater. 2020, 3, 2354–2363. [Google Scholar] [CrossRef]
  98. Wu, Y.; Chen, J.-Y.; He, W.-M. Surface-Enhanced Raman Spectroscopy Biosensor Based on Silver Nanoparticles@metal-Organic Frameworks with Peroxidase-Mimicking Activities for Ultrasensitive Monitoring of Blood Cholesterol. Sens. Actuators B Chem. 2022, 365, 131939. [Google Scholar] [CrossRef]
  99. He, L.; Huang, J.; Xu, T.; Chen, L.; Zhang, K.; Han, S.; He, Y.; Lee, S.T. Silver Nanosheet-Coated Inverse Opal Film as a Highly Active and Uniform SERS Substrate. J. Mater. Chem. 2012, 22, 1370–1374. [Google Scholar] [CrossRef]
  100. Fang, C.; Brodoceanu, D.; Kraus, T.; Voelcker, N.H. Templated Silver Nanocube Arrays for Single-Molecule SERS Detection. RSC Adv. 2013, 3, 4288. [Google Scholar] [CrossRef]
  101. Ju, J.; Liu, W.; Perlaki, C.M.; Chen, K.; Feng, C.; Liu, Q. Sustained and Cost Effective Silver Substrate for Surface Enhanced Raman Spectroscopy Based Biosensing. Sci. Rep. 2017, 7, 6917. [Google Scholar] [CrossRef] [PubMed]
  102. Chen, L.; Mungroo, N.; Daikuara, L.; Neethirajan, S. Label-Free NIR-SERS Discrimination and Detection of Foodborne Bacteria by in Situ Synthesis of Ag Colloids. J. Nanobiotechnol. 2015, 13, 45. [Google Scholar] [CrossRef] [PubMed]
  103. Dina, N.E.; Gherman, A.M.R.; Chiş, V.; Sârbu, C.; Wieser, A.; Bauer, D.; Haisch, C. Characterization of Clinically Relevant Fungi via SERS Fingerprinting Assisted by Novel Chemometric Models. Anal. Chem. 2018, 90, 2484–2492. [Google Scholar] [CrossRef]
  104. Zhou, H.; Yang, D.; Ivleva, N.P.; Mircescu, N.E.; Niessner, R.; Haisch, C. SERS Detection of Bacteria in Water by In Situ Coating with Ag Nanoparticles. Anal. Chem. 2014, 86, 1525–1533. [Google Scholar] [CrossRef] [PubMed]
  105. Hermann, D.-R.; Lilek, D.; Daffert, C.; Fritz, I.; Weinberger, S.; Rumpler, V.; Herbinger, B.; Prohaska, K. In Situ Based Surface-Enhanced Raman Spectroscopy (SERS) for the Fast and Reproducible Identification of PHB Producers in Cyanobacterial Cultures. Analyst 2020, 145, 5242–5251. [Google Scholar] [CrossRef]
  106. Wu, X.; Huang, Y.-W.; Park, B.; Tripp, R.A.; Zhao, Y. Differentiation and Classification of Bacteria Using Vancomycin Functionalized Silver Nanorods Array Based Surface-Enhanced Raman Spectroscopy and Chemometric Analysis. Talanta 2015, 139, 96–103. [Google Scholar] [CrossRef] [PubMed]
  107. Dang, T.M.D.; Le, T.T.T.; Fribourg-Blanc, E.; Dang, M.C. Synthesis and Optical Properties of Copper Nanoparticles Prepared by a Chemical Reduction Method. Adv. Nat. Sci. Nanosci. Nanotechnol. 2011, 2, 015009. [Google Scholar] [CrossRef]
  108. Khan, A.; Rashid, A.; Younas, R.; Chong, R. A Chemical Reduction Approach to the Synthesis of Copper Nanoparticles. Int. Nano Lett. 2016, 6, 21–26. [Google Scholar] [CrossRef]
  109. Kudelski, A.; Grochala, W.; Janik-Czachor, M.; Bukowska, J.; Szummer, A.; Dolata, M. Surface-Enhanced Raman Scattering (SERS) at Copper(I) Oxide. J. Raman Spectrosc. 1998, 29, 431–435. [Google Scholar] [CrossRef]
  110. Li, Y.-S. Surface-Enhanced Raman Scattering at Colloidal Silver Oxide Surfaces. J. Raman Spectrosc. 1994, 25, 795–797. [Google Scholar] [CrossRef]
  111. Lin, J.; Hao, W.; Shang, Y.; Wang, X.; Qiu, D.; Ma, G.; Chen, C.; Li, S.; Guo, L. Direct Experimental Observation of Facet-Dependent SERS of Cu2O Polyhedra. Small 2018, 14, 1703274. [Google Scholar] [CrossRef] [PubMed]
  112. Lin, J.; Shang, Y.; Li, X.; Yu, J.; Wang, X.; Guo, L. Ultrasensitive SERS Detection by Defect Engineering on Single Cu2O Superstructure Particle. Adv. Mater. 2017, 29, 1604797. [Google Scholar] [CrossRef]
  113. Dizajghorbani Aghdam, H.; Moemen Bellah, S.; Malekfar, R. Surface-Enhanced Raman Scattering Studies of Cu/Cu2O Core-Shell NPs Obtained by Laser Ablation. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 223, 117379. [Google Scholar] [CrossRef]
  114. Kowalska, A.A.; Kaminska, A.; Adamkiewicz, W.; Witkowska, E.; Tkacz, M. Novel Highly Sensitive Cu-Based SERS Platforms for Biosensing Applications. J. Raman Spectrosc. 2015, 46, 428–433. [Google Scholar] [CrossRef]
  115. Dai, P.; Li, H.; Huang, X.; Wang, N.; Zhu, L. Highly Sensitive and Stable Copper-Based SERS Chips Prepared by a Chemical Reduction Method. Nanomaterials 2021, 11, 2770. [Google Scholar] [CrossRef] [PubMed]
  116. Fan, M.; Lai, F.-J.; Chou, H.-L.; Lu, W.-T.; Hwang, B.-J.; Brolo, A.G. Surface-Enhanced Raman Scattering (SERS) from Au:Ag Bimetallic Nanoparticles: The Effect of the Molecular Probe. Chem. Sci. 2013, 4, 509–515. [Google Scholar] [CrossRef]
  117. Liu, H.; Yang, Q. Facile Fabrication of Nanoporous Au–Pd Bimetallic Foams with High Catalytic Activity for 2-Nitrophenol Reduction and SERS Property. J. Mater. Chem. 2011, 21, 11961. [Google Scholar] [CrossRef]
  118. Han, Q.; Zhang, C.; Gao, W.; Han, Z.; Liu, T.; Li, C.; Wang, Z.; He, E.; Zheng, H. Ag-Au Alloy Nanoparticles: Synthesis and in Situ Monitoring SERS of Plasmonic Catalysis. Sens. Actuators B Chem. 2016, 231, 609–614. [Google Scholar] [CrossRef]
  119. Jana, N.R. Silver Coated Gold Nanoparticles as New Surface Enhanced Raman Substrate at Low Analyte Concentration. Analyst 2003, 128, 954. [Google Scholar] [CrossRef]
  120. Cui, Y.; Ren, B.; Yao, J.-L.; Gu, R.-A.; Tian, Z.-Q. Synthesis of AgcoreAushell Bimetallic Nanoparticles for Immunoassay Based on Surface-Enhanced Raman Spectroscopy. J. Phys. Chem. B 2006, 110, 4002–4006. [Google Scholar] [CrossRef] [PubMed]
  121. Kumar, G.V.P.; Shruthi, S.; Vibha, B.; Reddy, B.A.A.; Kundu, T.K.; Narayana, C. Hot Spots in Ag Core–Au Shell Nanoparticles Potent for Surface-Enhanced Raman Scattering Studies of Biomolecules. J. Phys. Chem. C 2007, 111, 4388–4392. [Google Scholar] [CrossRef]
  122. Alvarez-Puebla, R.A.; Ross, D.J.; Nazri, G.-A.; Aroca, R.F. Surface-Enhanced Raman Scattering on Nanoshells with Tunable Surface Plasmon Resonance. Langmuir 2005, 21, 10504–10508. [Google Scholar] [CrossRef] [PubMed]
  123. Freeman, R.G.; Hommer, M.B.; Grabar, K.C.; Jackson, M.A.; Natan, M.J. Ag-Clad Au Nanoparticles: Novel Aggregation, Optical, and Surface-Enhanced Raman Scattering Properties. J. Phys. Chem. 1996, 100, 718–724. [Google Scholar] [CrossRef]
  124. Kim, K.; Kim, K.L.; Lee, S.J. Surface Enrichment of Ag Atoms in Au/Ag Alloy Nanoparticles Revealed by Surface Enhanced Raman Scattering Spectroscopy. Chem. Phys. Lett. 2005, 403, 77–82. [Google Scholar] [CrossRef]
  125. Russo, L.; Merkoçi, F.; Patarroyo, J.; Piella, J.; Merkoçi, A.; Bastús, N.G.; Puntes, V. Time- and Size-Resolved Plasmonic Evolution with Nm Resolution of Galvanic Replacement Reaction in AuAg Nanoshells Synthesis. Chem. Mater. 2018, 30, 5098–5107. [Google Scholar] [CrossRef]
  126. Daniel, J.R.; McCarthy, L.A.; Ringe, E.; Boudreau, D. Enhanced Control of Plasmonic Properties of Silver–Gold Hollow Nanoparticles via a Reduction-Assisted Galvanic Replacement Approach. RSC Adv. 2019, 9, 389–396. [Google Scholar] [CrossRef] [PubMed]
  127. Hao, E.; Li, S.; Bailey, R.C.; Zou, S.; Schatz, G.C.; Hupp, J.T. Optical Properties of Metal Nanoshells. J. Phys. Chem. B 2004, 108, 1224–1229. [Google Scholar] [CrossRef]
  128. Feng, J.; Wu, X.; Ma, W.; Kuang, H.; Xu, L.; Xu, C. A SERS Active Bimetallic Core–Satellite Nanostructure for the Ultrasensitive Detection of Mucin-1. Chem. Commun. 2015, 51, 14761–14763. [Google Scholar] [CrossRef]
  129. Wang, F.; Cao, S.; Yan, R.; Wang, Z.; Wang, D.; Yang, H. Selectivity/Specificity Improvement Strategies in Surface-Enhanced Raman Spectroscopy Analysis. Sensors 2017, 17, 2689. [Google Scholar] [CrossRef] [PubMed]
  130. Kamińska, A.; Witkowska, E.; Winkler, K.; Dzięcielewski, I.; Weyher, J.L.; Waluk, J. Detection of Hepatitis B Virus Antigen from Human Blood: SERS Immunoassay in a Microfluidic System. Biosens. Bioelectron. 2015, 66, 461–467. [Google Scholar] [CrossRef]
  131. Zengin, A.; Tamer, U.; Caykara, T. SERS Detection of Hepatitis B Virus DNA in a Temperature-Responsive Sandwich-Hybridization Assay. J. Raman Spectrosc. 2017, 48, 668–672. [Google Scholar] [CrossRef]
  132. Shanmukh, S.; Jones, L.; Driskell, J.; Zhao, Y.; Dluhy, R.; Tripp, R.A. Rapid and Sensitive Detection of Respiratory Virus Molecular Signatures Using a Silver Nanorod Array SERS Substrate. Nano Lett. 2006, 6, 2630–2636. [Google Scholar] [CrossRef]
  133. Cao, Y.C.; Jin, R.; Mirkin, C.A. Nanoparticles with Raman Spectroscopic Fingerprints for DNA and RNA Detection. Science 2002, 297, 1536–1540. [Google Scholar] [CrossRef]
  134. Xia, J.; Liu, Y.; Ran, M.; Lu, D.; Cao, X.; Wang, Y. SERS Platform Based on Bimetallic Au-Ag Nanowires-Decorated Filter Paper for Rapid Detection of MiR-196ain Lung Cancer Patients Serum. J. Chem. 2020, 2020, 5073451. [Google Scholar] [CrossRef]
  135. Prakash, O.; Sil, S.; Verma, T.; Umapathy, S. Direct Detection of Bacteria Using Positively Charged Ag/Au Bimetallic Nanoparticles: A Label-Free Surface-Enhanced Raman Scattering Study Coupled with Multivariate Analysis. J. Phys. Chem. C 2020, 124, 861–869. [Google Scholar] [CrossRef]
Figure 1. (A) Metal nanoparticles showing localized surface plasmon resonance when exposed to Electromagnetic waves of the specific wavelength resonating with plasmon frequency. (B) Spontaneous Raman Scattering and Surface-enhanced Raman scattering. (C) Electromagnetic enhancement of SERS signal (D). Chemical enhancement of SERS Signal due to SERS substrate and analyte complex formation. (Adapted from Refs. [2,7]).
Figure 1. (A) Metal nanoparticles showing localized surface plasmon resonance when exposed to Electromagnetic waves of the specific wavelength resonating with plasmon frequency. (B) Spontaneous Raman Scattering and Surface-enhanced Raman scattering. (C) Electromagnetic enhancement of SERS signal (D). Chemical enhancement of SERS Signal due to SERS substrate and analyte complex formation. (Adapted from Refs. [2,7]).
Photochem 04 00026 g001
Figure 2. ScFv-antibody-conjugated gold nanoparticles for tumor biomarker EGFR detection. In vivo SERS spectra from tumor site (red) and liver locations (blue) using (A) targeted and (B) nontargeted gold nanoparticles. Malachite green, which has unique spectral signatures, is the Raman reporter molecule and is indicated in (A,B). (C) Images of 785 nm laser beam (laser power = 20 mW) focused on the liver’s anatomical location and tumor’s site. The liver site (blue) and the tumor site (red) were used to obtain in vivo SERS spectra using 2 s signal integration and 785 nm excitation. (Adapted with permission from Ref. [86]. Copyright 2007, Springer Nature).
Figure 2. ScFv-antibody-conjugated gold nanoparticles for tumor biomarker EGFR detection. In vivo SERS spectra from tumor site (red) and liver locations (blue) using (A) targeted and (B) nontargeted gold nanoparticles. Malachite green, which has unique spectral signatures, is the Raman reporter molecule and is indicated in (A,B). (C) Images of 785 nm laser beam (laser power = 20 mW) focused on the liver’s anatomical location and tumor’s site. The liver site (blue) and the tumor site (red) were used to obtain in vivo SERS spectra using 2 s signal integration and 785 nm excitation. (Adapted with permission from Ref. [86]. Copyright 2007, Springer Nature).
Photochem 04 00026 g002
Figure 3. SERS spectrum of Staphylococcus aureus using novel highly sensitive Cu-based SERS platforms. (Adapted with permission from the Ref. [114]. Copyright 2015, John Wiley and Sons).
Figure 3. SERS spectrum of Staphylococcus aureus using novel highly sensitive Cu-based SERS platforms. (Adapted with permission from the Ref. [114]. Copyright 2015, John Wiley and Sons).
Photochem 04 00026 g003
Figure 4. Schematic representation of (A) Two different monometallic nanoparticles represented as yellow and green. (B) Alloy BMNP substrate surface. (C) Core (yellow) shell (green) BMNPs. (D) Detection of mucin1 via SERS with bimetallic gold nanorod core and AgNP satellite assemblies. (E) Enhanced SERS spectra of Mucin 1. (F) Peak intensity at 1142 cm−1 for Mucin-1 detection with detection limit of 4.3 aM. (DF are adapted with permission from [128,129]. Copyright 2015, Royal Society of Chemistry).
Figure 4. Schematic representation of (A) Two different monometallic nanoparticles represented as yellow and green. (B) Alloy BMNP substrate surface. (C) Core (yellow) shell (green) BMNPs. (D) Detection of mucin1 via SERS with bimetallic gold nanorod core and AgNP satellite assemblies. (E) Enhanced SERS spectra of Mucin 1. (F) Peak intensity at 1142 cm−1 for Mucin-1 detection with detection limit of 4.3 aM. (DF are adapted with permission from [128,129]. Copyright 2015, Royal Society of Chemistry).
Photochem 04 00026 g004
Table 1. SERS-based biosensing applications of metallic nanomaterials.
Table 1. SERS-based biosensing applications of metallic nanomaterials.
Metallic NanomaterialsApplicationsRef.
GoldDetection of calcium mobilizing second messenger nicotinic adenine dinucleotide phosphate (100 µM) in cell extract without sample purification and labeling.[55]
Label-free multiplex detection of miRNAs; miR– 10b, miR–21, and miR–373 relevant to cancer metastasis with LODs of 3.53 fM, 2.17 fM, and 2.16 fM.[56]
Ultrasensitive and selective detection of Alzheimer’s Tau protein using SERS-based sandwich assay.[57]
Quantitative multiplex sensing of DNAs with a limit of detection of 10 pM.[58]
Sensing of bacterial endotoxin (lipopolysaccharide) with the low detection limit of 5 lipopolysaccharide molecules per AuNP.[59]
SERS-based onsite detection of breast cancer using human tears and detection at femtomole scale[60]
Label-free sensing of quinoline antibiotic found in wastewater with a 5.01 ppb lower limit of detection.[61]
Drug (doxorubicin) biosensing in biological fluid using SERS. The experimental limit of detection in water is 1 nM and in human and bovine serum is 16 nM.[62]
SARS-CoV-2 detection using the SERS platform with a limit of detection of virus concentrations less than 10 PFU/ml.[63]
SilverDetection of amyloid-β (a known hallmark for Alzheimer’s disease pathogenesis) peptide aggregates in biological fluids with a nanomolar limit of detection using AgNP composite SERS sensors and a 532 nm laser.[64]
SERS biosensing and drug delivery of anticancerous doxorubicin using nanosized graphene-oxide-coated AgNPs.[65]
Label-free identification of circulating cancer protein biomarkers. Using SERS active strip natural diatomite coated with AgNPs via a layer-by-layer assembly method.[66]
Quantitative rapid SERS detection of Bacillus cereus spore biomarker 2,6-pyridine dicarboxylic acid with a limit of detection as 8.62 nM.[67]
Quantitative detection of tenofovir down to 25 ng/ml clinically relevant concentrations in an aqueous matrix using SERS.[68]
Detection of esophageal cancer using stable and low-toxicity polyethylene-glycol-coated silver nanocubes.[69]
CopperSERS-based determination of cephalosporins antibiotics in spiked human urine using CuNPs.[70]
SERS detection of designer drugs using copper nanowires coated carbon fibers.[71]
SERS-based sensitive detection of telomerase activity using copper oxide nanoparticles.[72]
BimetallicHighly reproducible and accurate single-cell Raman imaging and in vivo tumor imaging with 1 × 1 mm2 area can be quickly achieved within 35 s under open-air conditions using SERS Au core–Raman-active molecule–Ag shell–Au shell nanoparticles.[73]
Au/Ag core–shell nanoparticles, conjugated with monoclonal antibodies for highly sensitive SERS imaging of cancer biomarkers in live cells.[74]
Sensitive detection of Anticancer drug mitoxantrone using Bimetallic Ag–Au and Ag–Cu alloy microflowers SERS sensor with a limit of detection of 1 fM.[75]
Table 2. Detection of glucose and different parameters for cost analysis for SERS techniques with silver nanoparticles. Analysis of itemized cost for synthesizing 5 mg Ag NP@N-GQD, which requires 0.8 mg N-GQD and 10 mg AgNO3. The measured baseline glucose level (diluted by 10 times) is 0.21 mM. The first and second values in the column “Estimated increase in glucose concentration” give the mean value and the standard deviation, respectively. Note that the Mean Percent Error is calculated as (ΔEstimated − ΔExpected)/ΔExpected × 100%, where ΔEstimated represents the mean estimated increase in glucose concentration and ΔExpected represents the expected increase in glucose concentration. (Adapted from Ref. [101]).
Table 2. Detection of glucose and different parameters for cost analysis for SERS techniques with silver nanoparticles. Analysis of itemized cost for synthesizing 5 mg Ag NP@N-GQD, which requires 0.8 mg N-GQD and 10 mg AgNO3. The measured baseline glucose level (diluted by 10 times) is 0.21 mM. The first and second values in the column “Estimated increase in glucose concentration” give the mean value and the standard deviation, respectively. Note that the Mean Percent Error is calculated as (ΔEstimated − ΔExpected)/ΔExpected × 100%, where ΔEstimated represents the mean estimated increase in glucose concentration and ΔExpected represents the expected increase in glucose concentration. (Adapted from Ref. [101]).
ComponentChemical AgentACS NumberPack SizePrice (USD)Chemical Quantity for Synthesizing 5 mg Ag NP@N-GQDCost (USD)Cost Distribution
Ag NPAgNO37761-88-85 g24.910.0 mg4.98 × 10−299.67%
N-GQDCitric Acid77-92-9500 g90.30.8 mg1.45 × 10−40.29%
Dicyandiamid461-58-51000 g45.70.4 mg1.83 × 10−50.04%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Karlo, J.; Razi, S.S.; Phaneeswar, M.S.; Vijay, A.; Singh, S.P. Metallic Nanoparticles for Surface-Enhanced Raman Scattering Based Biosensing Applications. Photochem 2024, 4, 417-433. https://doi.org/10.3390/photochem4040026

AMA Style

Karlo J, Razi SS, Phaneeswar MS, Vijay A, Singh SP. Metallic Nanoparticles for Surface-Enhanced Raman Scattering Based Biosensing Applications. Photochem. 2024; 4(4):417-433. https://doi.org/10.3390/photochem4040026

Chicago/Turabian Style

Karlo, Jiro, Syed S. Razi, Mahamkali Sri Phaneeswar, Arunsree Vijay, and Surya Pratap Singh. 2024. "Metallic Nanoparticles for Surface-Enhanced Raman Scattering Based Biosensing Applications" Photochem 4, no. 4: 417-433. https://doi.org/10.3390/photochem4040026

Article Metrics

Back to TopTop