Plasmonic silver and gold nanoparticles: shape- and structure-modulated plasmonic functionality for point-of-caring sensing, bio-imaging and medical therapy

Yingjie Hang , Anyang Wang and Nianqiang Wu *
Department of Chemical Engineering, University of Massachusetts Amherst, Amherst, MA 01003-9303, USA. E-mail: nianqiangwu@umass.edu; Tel: +1-413-545-6175

Received 20th September 2023

First published on 21st February 2024


Abstract

Silver and gold nanoparticles have found extensive biomedical applications due to their strong localized surface plasmon resonance (LSPR) and intriguing plasmonic properties. This review article focuses on the correlation among particle geometry, plasmon properties and biomedical applications. It discusses how particle shape and size are tailored via controllable synthetic approaches, and how plasmonic properties are tuned by particle shape and size, which are embodied by nanospheres, nanorods, nanocubes, nanocages, nanostars and core–shell composites. This article summarizes the design strategies for the use of silver and gold nanoparticles in plasmon-enhanced fluorescence, surface-enhanced Raman scattering (SERS), electroluminescence, and photoelectrochemistry. It especially discusses how to use plasmonic nanoparticles to construct optical probes including colorimetric, SERS and plasmonic fluorescence probes (labels/reporters). It also demonstrates the employment of Ag and Au nanoparticles in polymer- and paper-based microfluidic devices for point-of-care testing (POCT). In addition, this article highlights how to utilize plasmonic nanoparticles for in vitro and in vivo bio-imaging based on SERS, fluorescence, photoacoustic and dark-field models. Finally, this article shows perspectives in plasmon-enhanced photothermal and photodynamic therapy.


image file: d3cs00793f-p1.tif

Yingjie Hang

Yingjie Hang is currently a PhD candidate under the supervision of Prof. Nick Wu. in the Department of Chemical Engineering at University of Massachusettes Amherst, USA. She received her MS degree from Soochow University in 2019. Her research interests include plasmon-enhanced fluorescence and SERS for biosensors for point-of-care detection of protein and nucleic acids, as well as cellular sensors integrated with photodynamic therapy.

image file: d3cs00793f-p2.tif

Anyang Wang

Dr Anyang Wang is a Postdoctoral Fellow under the supervision of Prof. Nick Wu in the Department of Chemical Engineering at University of Massachusettes Amherst, USA. He received his PhD degree from University at Buffalo, State University of New York in 2020, and MS and BS degrees from Tohoku University, Japan, in 2014 and 2012, respectively. His research interests include optical devices, especially plasmonic nanostructures for biosensing and light management technologies, as well as microfluidics for point-of-care testing.

image file: d3cs00793f-p3.tif

Nianqiang Wu

Dr Nianqiang (Nick) Wu received his PhD degree in Materials Science & Engineering from Zhejiang University, China. Currently, he holds the position of Armstrong-Siadat Endowed Chair Professor in Materials Science at University of Massachusetts Amherst, USA. Dr Wu is a fellow of the Electrochemical Society (FECS), American Institute for Medical and Biological Engineering (F-AIMBE) and Royal Society of Chemistry (FRSC). He has received several honors and awards such as Highly Cited Researcher (Clarivate Analytics) and ECS Sensor Division Outstanding Achievement Award. He has authored or co-authored over 200 journal articles, 3 book chapters and 1 book entitled “Biosensors Based on Nanomaterials and Nanodevices”.


1. Introduction

Gold (AuNPs) and silver nanoparticles (AgNPs) possess many intriguing physicochemical properties and have been applied in diagnosis, cancer therapy, drug delivery, hygiene, catalysis, electronics and so on.1–3 This review article focuses on their plasmonic properties. Localized surface plasmon resonance (LSPR) refers to the collective oscillations of conduction electrons confined locally by metallic nanoparticles under the resonant excitation of incident light.4 Herein, we highlight five LSPR effects that are utilized in various applications of metal nanoparticles: (i) light absorption, (ii) light scattering, (iii) strongly enhanced electromagnetic (EM) field surrounding nanoparticles, (iv) hot carrier emission, and (v) heating.4–6 For example, light scattering is exploited for dark-field imaging.7 Light absorption and scattering are used in colorimetric probes in paper lateral flow assays.8 High-Q light absorption or reflection is used in refractive-index sensors.1 Strong EM field enhancement is used for SERS sensing and imaging.9,10 Hot electron emission is utilized for photoelectrochemical sensing. Plasmonic heating serves in photoacoustic imaging, photothermal and photothermal therapy.11 The abovementioned LSPR effects heavily depend on the particle shape and size of metal nanoparticles, offering ample opportunities to tune the performance of sensors, imaging and therapeutic agents.

A large amount of energy is stored in plasmonic metal nanoparticles during the light excitation of LSPR. When a plasmonic metal is coupled with other materials and media, plasmonic energy in metal nanoparticles can be transferred to adjacent materials and media via several processes such as4 (i) light trapping/scattering, (ii) hot electron injection, (iii) plasmon-induced resonance energy transfer (PIRET), (iv) near-field coupling and (v) Purcell effect. As a result, these plasmonic energy transfer processes are utilized in designing composite materials and hetero-structures to enhance sensing, bio-imaging and medical therapy. For example, plasmonic energy can transfer from metal to a semiconductor through light trapping, hot electron injection or PIRET in a metal–semiconductor heterojunction, which can be used for photoelectrochemical sensing and photodynamic therapy. Light scattering, Purcell effect and PIRET are employed for transferring energy from plasmonic metal nanoparticles to fluorophores to enhance fluorescence sensing and imaging. Hot electron injection is used for photoelectrochemical sensing.

Recently, several review articles on plasmonic silver and gold nanoparticles have been published.12–16 Those articles deal with particle synthesis, plasmon and applications, respectively. The present article is focused on the correlations among particle geometry, plasmonic properties and applications. This paper will start with a summary of Au and Ag nanoparticle synthesis and plasmonic functionalities such as scattering, hot electrons, heating, and EM field enhancement. Next, it will discuss how to utilize these properties to design materials for applications such as sensors, bioimaging and medical therapy. In the sensing section, it provides strategies for designing plasmon-enhanced colorimetry, fluorescence, SERS, electroluminescence and photoelectrochemistry, placing an emphasis on optical probe structure design, as well as probe integration with paper- and polymer-based microfluidic devices. In particular, it discusses how to use plasmonic nanoparticles to construct optical probes including colorimetric, SERS and plasmonic fluorescence probes (labels/reporters). Subsequently, it will describe plasmon-enhanced bio-imaging methods such as dark-field, fluorescence, SERS and photoacoustic methods, highlighting their working principle and applications under both in vitro and in vivo conditions. Finally, it will give an overview of plasmon-modulated photothermal and photodynamic therapy in terms of concepts, basic principles, therapeutic agent structures and case studies.

2. Shape modulation and plasmonic functionality of Au and Ag nanoparticles

Under irradiation of pulse light, strong LSPR is excited in nanoparticles, enhancing light absorption and the local field in a global nonequilibrium state.17 To restore a thermally equilibrium state, the engendered plasmon experiences dephasing through either radiative emission of photons or non-radiative generation of hot electron–hole pairs through Landau damping (on the time scale of 1–100 fs).11,18 Subsequently, the energy distribution of hot electrons relaxes to a Fermi–Dirac-like distribution via electron–electron scattering (on the time scale of 100–1000 fs), accompanied by concomitant electron–phonon interaction (on the time scale of 1–10 ps).19,20 This results in thermalization of hot carriers, consequently increasing the lattice temperature. Finally, thermal energy in the metallic nanoparticles is progressively dissipated into the ambient surroundings through phonon–phonon interaction. The temperature of the surrounding environment quickly elevates, while electrons in the conduction band of metallic nanoparticles regress to the original ground state before excitation (100 ps–10 ns).2 These processes are strongly dependent on the geometry and dimensions of gold and silver nanoparticles. In this section, we will discuss how the shapes and sizes of gold and silver nanoparticles are tailored, and how the shape and size influence plasmonic properties.

2.1. Nanospheres

Nanosphere synthesis. Gold nanospheres (AuNSs) can be synthesized by physical and chemical methods.16,21,22 Pulsed laser ablation and evaporation–condensation are common physical approaches. In the pulsed laser ablation method, AuNPs are developed from a bulk gold pellet using pulsed laser ablation, a technique first introduced by Cotton et al.23 It involves two steps: vigorous ablation for the generation of small gold seeds, and milder irradiation for the growth of NPs. The process is conducted by immersing bulk gold in an aqueous solution in the presence of polymeric or oligomeric capping ligands for stabilization and surface functionalization. Evaporation–condensation is conducted in a tube furnace at atmospheric pressure, transforming the gold source into a gas phase.24 Gold vapor condenses upon coming in contact with carrier gas, followed by nucleation and growth. Compared to chemical methods, physical methods involve no, or little hazardous reagents, resulting in the formation of high purity nanoparticles. However, the particles generated in physical methods are limited to a spherical shape.

Chemical methods are more commonly employed for Au nanosphere synthesis due to their low cost, and flexibility in controlling size, shape and post-modification. In the wet chemistry route, a gold salt precursor (chloroauric acid, HAuCl4) is reduced by agents such as trisodium citrate, sodium borohydride, formaldehyde, amines, and protected by stabilizers such as thiolates, polymer, trisodium citrate, and hexadecyltrimethylammonium bromide from aggregation through electrostatic and physical repulsion. The Turkevich–Frens method is widely accepted and has been further modified by several research groups for gold nanosphere synthesis.25 Briefly, trisodium citrate, serving as both a reducing agent and stabilizer, is quickly added to the boiling HAuCl4 solution under vigorous stirring until a wine-red colloidal suspension is obtained. Since trisodium citrate is a mild reducing reagent, the formed AuNSs are always larger than 10 nm. To control particle size down to 5 nm, a stronger reducing reagent, NaBH4, is used with the citrate stabilizer. Overall, the Turkevich–Frens method using citrate as a stabilizer can achieve 5–150 nm sized ranged nanoparticles but when the size exceeds 20 nm, particles tend to be polydispersed. The Brust–Schiffrin method further improved this situation based on the mechanism of nucleation and successive growth when using thiolate as a stabilizer.26 Thiolate binds to the surface of NPs through Au–S bonds, greatly inhibiting the growth process and leading to smaller and monodispersed NPs. The generated AuNSs show thermal and air stability with narrow dispersity for nanoparticles less than 5 nm. In addition, by the introduction of a functional thiolate stabilizer, it is easier to perform the functionalization of NSs. Despite the ease of operation of the above nucleation and growth method, the maximum size is around 200 nm, and it is hard to achieve uniform NSs for large-sized particles. The seed-mediated growth method is exploited by adding a “growth” solution to small-sized NSs formed using previous methods.27 Then, the newly reduced Au0 atoms assemble on the seed surface to form large-size AuNSs. Since the reducing agents used in the second step are always mild, the newly reduced Au atoms prefer to assemble on the surface of the Au seeds, instead of forming new particle nucleation in the solution. In this way, monodispersed AuNSs with large sizes (up to 300 nm) are achieved.

Silver nanospheres (AgNSs) are synthesized in a similar manner to Au nanospheres. AgNSs can also be synthesized by physical methods and chemical reduction methods.15 The citrate-based reduction and stabilization method for AgNSs synthesis was first reported in 1982.28 Similar to AuNSs, the citrate-based method usually shows poor control over size and shape. Multivalent alcohols, serving as reducing/solvent agents, have become more popular for producing monodispersed AgNSs, named the polyols method.29 The typical process involves AgNO3 as a precursor, ethylene glycol as a reducing/solvent agent, and poly(vinylpyrrolidone) (PVP) as a stabilizer.30

Plasmonic properties of nanospheres. Both AuNPs and AgNPs show a plasmonic dipole mode. A simple harmonic oscillator model or the Mie theory reveals the exact condition for LSPR of nanospheres.31 The extinction cross-section (Cext) is given by31
 
image file: d3cs00793f-t1.tif(1)
where ω refers to the frequency of light, V is the volume of the nanospheres given by 4/3πR3, εdiel is the dielectric constant of the surrounding medium, and ε′ and ε′′ are the real and imaginary parts of the relative permittivity of metals, respectively. Since the real part, ε′, of the dielectric function is negative for most of the visible range, a strong resonance occurs at the wavelength λ, where the condition ε′ + 2εdiel = 0 is satisfied, leading to the denominator becoming zero. The LSPR frequency is then expressed by32
 
image file: d3cs00793f-t2.tif(2)
where ωp is the plasma frequency of bulk metal. Based on the plasmon damping process, the ratio of absorption and scattering plays a vital role in determining the plasmon effect in the corresponding application. The Mie–Gans theory also provides the relationship between scattering (Csca), absorption (Cabs) and extinction cross-sections:33
 
image file: d3cs00793f-t3.tif(3)
 
image file: d3cs00793f-t4.tif(4)
 
Cext = Cabs + Csca = eqn (1) for a sphere(5)
where V is the volume of a nanosphere, given by image file: d3cs00793f-t5.tif. n(i) is the depolarization factor which is equal to 1/3 for a sphere. It can be seen from the above equations that the absorption scales with r3 for small particles (rλ), dominating over the scattering (Fig. 1(a)).34 Owing to the rapid scaling of scattering cross-sections with r6, scattering becomes dominant over absorption when particles become larger. When spheres are larger than 15 nm, plasmon tends to re-radiate energy with a large scattering cross-section. Light scattering is evident when the sphere size is 50 nm and becomes a major component in the light extinction spectrum when the size reaches 100 nm (Fig. 1(b)).3 In general, for spheres smaller than 15 nm, light absorption dominates and plasmon energy is likely to decay through the near-field or be absorbed as heat. When the size of spheres further decreases below 3 nm, the plasmon resonance behavior disappears due to high surface damping. In addition, the LSPR peak position experiences red shifts with the size of nanospheres due to the reduction of the depolarization field arising from the retardation effect.35 The retardation effect describes that the conduction electrons do not all move in phase, resulting in a diminished depolarization field at the center point, which originates from the peripheral polarized substance.35 In other words, larger nanoparticles experience less repulsion for electrons on the opposite surface, leading to greater red shifts of plasmon peaks. Larger nanoparticles also experience spectrum broadening resulting from radiation damping.

image file: d3cs00793f-f1.tif
Fig. 1 Plasmon behaviors of AuNPs and AgNPs. (a) (b) Tuning light absorption vs. light scattering through sphere size; (c) the difference in the LSPR band of Au and Ag nanoparticles. Reproduced with permission from (a) and (b) ref. 34 copyright 2013, The Electrochemical Society interface; (c) ref. 40, copyright 2010, John Wiley and Sons.

The plasmonic hot electron emission and injection processes are highly dependent on the size and shape of nanoparticles. Govorov et al. employed a quantum theory to evaluate the influence of shape modulation on hot electron energies and the electron mean free path.36 Their theories show that hot electrons from small nanospheres have high energy levels to overcome the typical Schottky barrier in metal–semiconductor junctions, while the majority of hot electrons from large particles exhibit low energetic levels. In addition, the number of injected carriers is affected by a finite mean free path, which indicates that hot electrons generated in small spheres have a higher possibility of transferring into the semiconductor. But if the particle size becomes less than 3 nm, light absorption capability is reduced because of surface damping. If the particle size is too small, the light absorption capability is low. Thus, the total number of hot electrons generated is low. Hence, the optimal size of gold nanoparticles is estimated to be 10–20 nm to maximize hot electron generation and injection efficiencies.

The plasmon heating effect usually occurs at plasmon resonance and arises from relaxation steps where lattice thermalization occurs, leading to the formation of a Fermi–Dirac-like distribution of electrons. The conversion efficiency (η) from light absorption to thermal energy is expressed as5

 
image file: d3cs00793f-t6.tif(6)
where Qext is the external heat input absorbed by nanoparticles, Q0 represents the heat dissipated by the surrounding environment, I is the intensity of the incident light impinging on the sample, and Aλ is the absorbance of the nanoparticles at the wavelength λ. Particularly, the heat input can be described as the incident light beam attenuation using the Beer–Lambert law37
 
I = I0 × 10Aλ(7)
 
Aλ = loptext(8)
where lopt is the beam optical path, C is the molar concentration and μ represents the molar extinction coefficient of nanoparticles at a given wavelength which is the sum of absorption (μabs) and scattering (μsca) coefficients described as follows:37
 
μext = μabs + μsca(9)

The conversion efficiency can then be expressed as:

 
image file: d3cs00793f-t7.tif(10)
where μs refers to the attenuation coefficient of the solvent. If there is no heat absorption in the solvent (μs = 0) and the optical length approaches 0, the efficiency could be approximated as:
 
image file: d3cs00793f-t8.tif(11)
El-Sayed et al. calculated the absorption and scattering properties of gold nanoparticles and found that Csca/Cabs increased with increasing size, resulting in a decrease in heat conversion efficiency accordingly.38 It is worth noting that the total heat generation by nanoparticles is directly related to the absorption crosssection, while efficiency is related to Csca/Cabs.37 In other words, the total heat generation could be higher in large particles since their absorption cross-section is larger than that of smaller particles, even though their absorption portion in total extinction is lower than that of smaller ones. Govorov et al., demonstrated that the local temperature around the nanoparticles is dependent on their size; thus larger AuNPs can have a higher total heat and heat generate rate, while their efficiency may be lower.6,39

Despite these similarities, AuNPs and AgNPs show different plasmon behaviors as exhibited in Fig. 1(c),40,41 including differences in the LSPR band position, local EM field enhancement, full width at half maximum (FWHM), and Q factor.40 First, for the plasmon peak position, it is determined from eqn (2), which is influenced by the plasma frequency of bulk materials and the dielectric constant of medium, leading to the appearance of AgNP peaks around 400 nm while AuNPs show shifts to over 500 nm. The high EM field and Q factor are attributed to the real and imaginary parts of their relative permittivity.42 The relative permittivity of gold and silver is complex, containing real parts and imaginary parts as mentioned before. Both the real parts of relative permittivity (ε′) are large and negative in most of the visible range, while their imaginary parts (ε′′) are smaller and positive in amplitude. Blaber et al. defined the figure-of-merit (FoM) as −ε′/ε′′ to assess the quality of a LSPR supported in a nanoparticle, where both Au and Ag show superior FoMs among noble metals.43ε′′ is related to the optical loss of materials; thus Au shows nonradiative damping compared to Ag for λ ≤ 600 nm due to interband transition, leading to a lower EM field and Q factor than its Ag counterparts. Additionally, AgNPs under resonance conditions generate almost 10 times greater heat that AuNPs due to their stronger plasmon enhancement.39

2.2. Nanorods

Nanorod synthesis. Gold nanorods (AuNRs) were first reported using the photochemical reduction method, where ion pairs were formed by auric acid bound to rod-like micelle surfactants when excited by UV light.44 Since then, various methods, such as X-ray irradiation, proton beam irradiation, template-method, bioreduction, solvothermal reduction, electrochemical method, have emerged to achieve high yield and mono-dispersity of AuNRs.45 The most popular is the seed-mediated growth method due to its simplicity, ease of size control, flexibility of modification as well as high yield and quality, which was reported by Jana et al., in 2001.46 The general procedure is similar to that used in large gold nanosphere synthesis, starting with citrate-stabilized gold seeds but using cetyltrimethylammonium bromide (CTAB) and silver ions to achieve the rod-like shape. The common principle for the formation of gold nanorods using CTAB is surfactant-preferential-binding-directed growth proposed by Murphy et al.47 For example, the 〈110〉 face, which has higher surface energy than other faces, is more likely to bind to CTAB for stabilization, leading to retarded growth on this face and continued growth in the end facet. Silver ions are used to control the aspect ratio. Nikoobakht and El-Sayed et al. further modified the process by using CTAB-stabilized seeds and ascorbic acid as the reducing agent in the growth step.48 This approach has greatly improved the yield to 99% with aspect ratios tuned from 1.5 to 4.5. Surfactants play a vital role in determining the aspect ratio of AuNRs.49 For example, benzyldimethylhexadecylammonium chloride (BDAC) as a co-surfactant can increase the aspect ratio to 10, while the Pluronic F-127 co-surfactant system can achieve aspect ratios up to 20.50 In addition, the continuous addition of the growth solution to the nanorod solution can result in a very long rod with an aspect ratio of up to 70.48

Silver nanorods have been synthesized using CTAB as the soft template, as reported by Murphy and co-workers.51 The aspect ratios can be adjusted by changing the seed concentration. Afterwards, multiple template-less, seed-less processes were reported by many researchers to improve the quality of Ag nanorods.52 In general, it is harder to control the shape of silver compared to gold, and to maintain that shape after purification. Silver nanorods with low aspect ratios are quite unstable under air and light, which is partially due to the release of Ag+ through a photooxidation process.47 In contrast, long aspect-ratio silver nanorods such as nanowires with a diameter of 30 nm and a length of dozen microns are very stable under air and light conditions.

Plasmonic properties of nanorods. Different from the single LSPR band of nanospheres, AuNRs and AgNRs have two plasmon modes: a weak transverse band similar to that of nanospheres in the visible range corresponding to surface plasmon oscillation along the short axis, and a strong longitudinal band in the near-infrared range related to oscillation along the long axis. The transverse band is not sensitive to the size of nanorods, while the longitudinal one is red shifted with an increase in the aspect ratio (R).53 For gold nanorods, the maximum resonance (λmax) of nanorods in an aqueous solution has a linear relationship with R, which is given by54
 
λmax = 95R + 420(12)
It shows that it is easier to tune the plasmon resonance of nanorods to longer wavelengths by varying their aspect ratio compared to nanospheres, which show slight red shifts with increasing size. The relationship among light absorption, scattering and extinction can be quantitatively described similar to nanospheres with the modification of n(i), which is defined as:33
 
image file: d3cs00793f-t9.tif(13)
 
n(b) = n(c) = (1 − n(a))/2(14)
where, a, b and c refer to the three axes of nanorods, a > b = c; and R is the aspect ratio equal to a/b. The LSPR occurs when image file: d3cs00793f-t10.tif the condition is satisfied. Specifically, i = a is used for longitudinal resonance while i = b and c is for transverse resonance. As shown in the formulas, Csca and Cabs are influenced not only by the wavelength of incident light and the particle size, but also by the aspect ratio. It has been found that at a fixed aspect ratio, absorption dominates over scattering for small particles, while scattering becomes dominant when particles are larger; this trend is consistent with that of nanospheres.55 With the increasing aspect ratio, scattering increases significantly but drops a little if R is further increased61 (Fig. 2).38,56 Scattering becomes less at a larger aspect ratio, at a more red-shifted longitudinal mode, due to increased absorption from the increasing imaginary part of Au. However, the scattering ratio is much higher than that of nanospheres because the strong longitudinal resonance band is far away from the interband absorption band. Therefore, the overall local EM field enhancement concentrated at the two ends of nanorods is on the scale of over 103 higher than that of spheres.57 Accordingly, the linewidth of gold nanorods is narrower than that of nanospheres, leading to a high quality factor.58 As for the plasmonic heating effect, gold nanorods show a very high absorption ratio (around 90%) at different aspect ratios according to calculations.38 El-Sayed's group further experimentally compared three different AuNRs and determined that the 28 × 8 nm one was the most effective for heat generation.59 Silver nanorods show similar trends but the plasmon peak position is slightly blue shifted at each aspect ratio compared to gold nanorods due to the intrinsic properties of silver.60 Also, owing to less interband damping in the resonance range, AgNRs result in stronger EM field enhancement than AuNRs.

image file: d3cs00793f-f2.tif
Fig. 2 Plasmons of gold nanorods. The LSPR band can be tuned by the aspect ratio of gold nanorods. Reproduced with permission from ref. 61, copyright 2010, Journal of Advanced Research.

2.3. Nanocubes

Nanocube synthesis. Expanding from gold nanorod synthesis, Murphy et al. have figured out that Au nanoparticles varied from rod-, rectangle-, hexagon-, cube-, and triangle- morphologies to star-like morphology by regulating the concentration of the seed, CTAB, ascorbic acid (AA) and precursors.62 They have studied the relationship among the particle shape, the facet affinity of surfactants and the growth kinetics. Given that CTAB exhibited higher affinity to 〈100〉 than 〈111〉 faces, a lower concentration of CTAB and a higher concentration of AA accelerated Au0 deposition on the 〈111〉 faces, producing cubic shapes bounded by predominantly 〈100〉 faces. In comparison, slightly reducing the AA concentration while keeping other components unchanged led to growth on both the 〈111〉 and 〈100〉 faces, leading to the formation of truncated octahedra. Huang et al. further systematically studied the shape evolution of cubic and cubic-related structures.63 They replaced CTAB with cetyltrimethylammonium chloride (CTAC) along with a very small amount of NaBr to achieve a high yield of gold nanocubes (AuNCs). Based on this reaction system, it was quite easy to control the shape of gold nanoparticles by adjusting the volume of AA; and changing the volume of seed introduced into the growth solution can lead to the generation of different sized nanocubes from 40 nm to 72 nm. The role of bromide ions was clearly discussed by Nam and co-workers.64 The bromide-concentration-dependent growth kinetics governed the corner sharpness of gold nanocubes. Low concentrations of bromide insufficiently blocked the 〈100〉 faces, resulting in a less significant growth-rate difference between 〈100〉 and 〈111〉/〈110〉, producing round-cornered nanocubes. When the bromide concentration was increased, effective and sufficient binding to 〈100〉 greatly reduced the reduction rate in 〈100〉, forming sharp-cornered nanocubes. Following the above principles, they synthesized AuNCs with highly controlled corner sharpness and tuned the cube size from 17 nm to 78 nm. Besides chemical reduction methods, Huang et al. also developed an electrochemical method to generate monodispersed AuNCs with highly uniform size.22

Silver nanocubes (AgNCs) are extensively studied and used; and they can serve as sacrificial templates for the generation of Ag–Au nanoshells, nanoboxes, and nanocages.65 The shape- and size-controlled synthesis of AgNCs has been well-developed by Xia's group. Different from AuNCs using CTAB or CTAC as surfactants, AgNCs are typically synthesized using poly(vinylpyrrolidone) (PVP) as a surfactant and silver nitrate as a precursor reduced in ethylene glycol.66,67 The geometric shape and size of AgNCs is determined by the molar ratio of PVP to the precursor. To improve the yield, purity, mono-dispersity and scale of synthesis, selective etching using HCl and oxygen is performed.68 For example, HNO3 formed from HCl and AgNO3 could selectively etch twinned seeds, leading to a high yield of single crystal nanocubes.68 Given that the whole reaction time for a typical polyol synthesis ranges from 16 to 26 h, a trace amount of sodium sulfide (Na2S) or sodium hydrosulfide (NaHS) was introduced into the reaction system for rapid synthesis (3–8 min in total).69 Sulfide species interact strongly with silver to form Ag2S, which could catalyze the reduction of Ag+ by reducing the reaction potential. In this way, they could not only speed up the reaction, but also minimize the size distribution by conducting a more simultaneous nucleation process. In addition, a new silver precursor, CF3COOAg, has been widely accepted since the nitrate group may decompose at high temperatures, especially during the polyol synthesis process.70 In the presence of CF3COOAg, a trace amount of NaHS, and HCl, successful production of high-quality AgNCs with a size range of 30–70 nm was achieved, making it more straightforward for scale-up production. The detailed protocol can be found in their publications.65,71,72

Hollow nanocubes are a special type of cube-shaped nanoparticles. When depositing Au on the surface of Ag nanocrystals, the interior Ag is oxidized and removed due to the galvanic replacement reaction, eventually producing hollow or even porous structures, named Au nanocages.73 By altering the morphology of the initial Ag template, this approach can be employed to achieve various shapes of Au nanocages including nanorings, nanocubes and prism-shaped nanoboxes.73 The evolution of porous Au nanocages from Ag nanocubes was studied systematically by Xia and colleagues by titrating HAuCl4 solution into a boiling suspension of Ag nanocubes.65,73 Briefly, a specific site with the highest surface energy, usually a surface step, point defect or hole in the capping layer, initiates the replacement reaction, forming a pinhole on one of faces of each tube. Such pinholes then act as the anode to oxide Ag and strip electrons which migrate to the nanocube faces and are captured by AuCl4. The generated Au0 epitaxially grows on the nanocube forming a Au layer, while the initial pinholes remain for Ag dissolution. As the reaction proceeds, the pinhole closes, forming nanoboxes with a hollow interior and a homogeneous wall composed of Au/Ag alloy. With the continuous addition of HAuCl4, nanoboxes will undergo dealloying and corner reconstruction, forming porous Au nanocages. The pore size and number can be controllably tuned by the volume of Au precursor added into the solution. Folloiwng the principle of Au nanocage formation, some special Matrioshka-like hollow nanostructures with multiple shells and movable solid cores are fabricated.74 For example, Au/Ag nanoshells are coated with another Ag layer, followed by the replacement reaction to generate another shell.75 As a result, multiple-walled nanoshells are formed.

Plasmonic properties of nanocubes. Different from the single LSPR band of nanospheres, AuNCs and AgNCs show higher order modes besides the dipole mode.76,77 When the dimensions of NCs are significantly less than the wavelength of the incoming light, the optical characteristics of the NCs are primarily influenced by the plasmon dipole mode. As a result, a peak in the optical extinction spectrum appears as the SPR peak. As the dimensions grow, larger NCs (over 140 nm) can stimulate higher order plasmonic wave modes such as electric quadrupole or magnetic dipole modes.78 This can lead to the emergence of multiple SPR peaks in the optical spectrum.

Furthermore, in comparison to AuNSs and AgNSs, AuNCs and AgNCs offer significant flexibility in tailoring resonance by varying particle size and shape (i.e., sharp vs. truncated) and local dielectric media.64,79 When increasing the dimensions of AgNCs, the plasmon resonance red-shifts and broadens because of the retardation effect and adiation damping (Fig. 3(a) and (b)).67 The sharpness of the corners provides additional flexibility in tuning the plasmon resonance and EM field enhancement. Sharpening the corner results in a red shift and a subtle broadening of the SPR band.80 The red-shift is again mainly attributed to the retardation effect, where the local EM field enhancement is not uniform across the entire surface, often exhibiting greater intensity near its sharp edges and corners.35


image file: d3cs00793f-f3.tif
Fig. 3 Plasmons of gold nanocubes. (a) The TEM image of different sized Ag nanocubes; (b) LSPR band position of Ag nanocubes can be tuned by regulating the particle size; (c) LSPR band of porous gold nanocubes can be tuned by regulating the pore size and number; (d) near-field distribution at the plasmon resonance wavelength. Reproduced with permission from (a) and (b) ref. 67, copyright 2010, Journal of Advanced Research. (c) and (d) Ref. 10, copyright 2022, Springer Nature.

It is worth noting that the LSPR band of Au nanocages can be easily adjusted from the near-infrared-I (NIR-I) region (around 700 nm) all the way to NIR-II range (1500 nm) (Fig. 3(c) and (d)).10 The remarkable NIR LSPR of Au nanocages results from the hollow and porous structure. Increasing the pore number and the particle size resulted in a red shift of the plasmon band. The numerical analysis showed that the pores formed in the Au nanocages intensified the local near-field and created high-density hotspots for EM field enhancements. However, since Au nanocages have less gold per unit volume compared to nanospheres, the radiation damping and the surface scattering from the thin walls are enhanced; thus the plasmon peaks become very broad especially at higher densities of porosity with long wavelengths.58 Also, the scattering fraction of Au nanocages increases with the cage size, reducing the heating conversion efficiency.73

2.4. Nanostars

Nanostar synthesis. From 2003, several seed-mediated or surfactant-less synthesis methods have been developed to synthesize gold nanostars (AuNSTs), including a representative method by Liz-Marzan et al., in which PVP serves as a surfactant and DMF serves as the solvent.81 The whole process is very simple, reproducible, and high-yield (almost 100% with no other shapes) and can be conducted in an ambient environment without an external energy source.82 Vo-Dinh and Liz-Marzan et al. further extended the protocol to tailor the size and the LSPR peaks of AuNSTs by tunning various parameters such as seed volume, size and types, PVP/Au molar ratios.81,83 Although the overall size can be tuned from around 40 nm to 116 nm, AuNSTs in this method generally exhibit multiple short, podgy and asymmetric spikes. Different from the common seeds used before, icosahedral seeds were used with the assistance of amine reagents to form highly symmetrical and uniform gold nanostars.84 The whole AuNSTs were enclosed with 〈321〉 high-index faces, which was attributed to the selective adsorption of dimethylamine on such faces. Owing to the specific seeds measuring over 50 nm in diameter, the formed nanostars usually are larger than 150 nm in diameter. On the other hand, a surfactant-free nanostar synthesis method has been developed by Vo-Dinh and co-workers.85 After achieving citrate-stabilized seeds, AgNO3 and ascorbic acid were added to the reaction system simultaneously to regulate the growth of spikes. Interestingly, this type of AuNSTs has long and slim spikes and shows strong SERS enhancement, presumably because of a significant amount of silver used in the reaction, along with the sharper tips.

Coral-like silver nanostars (AgNSTs) were first reported by Xia's group.86 The whole route was simple, fast and environmentally friendly, involving only AA and AgNO3 without the need for capping reagents. The regulation of AgNSTs depended on the ratio of AA and AgNO3 as well as the concentration of each reagent. The generated AgNSTs were highly branched although the branches were usually podgy. Some research groups further improved this method to obtain longer and sharper branches. For example, Sanchez-Cortes et al. used two reduction processes, first using neutral hydroxylamine followed by the citrate reagent.87 The former aimed to induce the growth of spikes, while the latter was used to accelerate and complete the reaction, serving as a soft capping surfactant. Although Ag nanostars can generate a higher EM field than gold, their overall size usually reaches up to hundreds of nanometers and is hard to reduce.

Plasmonic properties of nanostars. Gold nanostars show a plasmon hybridization mode associated with the core and tips. Nordlander and co-workers revealed that core plasmons interacted with tip plasmons to form bonding and anti-binding modes using a model consisting of a truncated spherical core and protruding prolate tips.88 The low-energy bonding mode was mostly contributed by tip plasmons, and it showed higher intensity than the antibonding mode. The core part served as an electron reservoir, and the core plasmons could significantly increase the excitation cross-section and further enhance the EM field of tips. Therefore, the LSPR peak position and the EM field enhancement were dependent on the tip length, tip thickness, tip angle and core diameter.89 Given that LSPR of nanostars is mainly confined to the tips, sharper tips could generate a more intense EM field. Wu's group compared the EM field enhancement between gold nanospheres, nanorods and nanostars and found that gold nanostars show highest “hot spots” due to the sharp tips of their nanostructures.9 Additionally, the heating effect of gold nanostars is also dependent on these parameters. Li's group synthesized a series of gold nanostars with different shapes and plasmon peak positions. It is interesting to find that the highest heat conversion efficiency occurs at an LSPR peak wavelength of 740 nm.90

As mentioned above, the plasmonic hot electron emission and injection processes are governed by the size and shape of nanoparticles. The experiments and simulations conducted by Wu's group have confirmed the energetics and population distribution of plasmonic hot electrons generated in gold nanospheres, nanorods and nanostars.91 The cut-off energies (the highest energetic levels) of the three types of nanoparticles were aligned with their LSPR band positions. That is, the highest energetic level of the nanosphere was the highest and that of nanostars was the lowest (Fig. 4(b)). After hot electrons transferred into the adjacent semiconductor, they remained “hot” and formed a nonthermal steady-state distribution in the semiconductor.91 They compared different shapes of gold nanoparticles and found that gold nanorods and nanostars exhibited hot electron injection rates several orders of magnitude higher than spherical ones. This is because nanoparticles with sharp corners and tips have intense “hot spots” which induce strong EM field enhancement and break the linear momentum of the electrons, thus enhancing hot electron generation. In addition, the nanorods produced the highest number of non-thermal electrons in the semiconductor while the nanospheres produced the least (Fig. 4(a)).


image file: d3cs00793f-f4.tif
Fig. 4 Hot electrons in gold nanoparticles. (a) Energetics of hot electrons in gold nanoparticles vary with shape modulation and they remain “hot” even upon transfer to TiO2; (b) energetics of hot electrons across the Au/TiO2 interface. Reproduced with permission from ref. 91, copyright 2018, American Chemistry Society.

2.5. Core–shell structures

Core–shell synthesis. Both gold and silver have a face-centered cubic (fcc) crystal structure and comparable lattice constants, ensuring epitaxial growth. Therefore, it is relatively facile to fabricate gold–silver alloys and one of the representative shapes is the core–shell structure. Given that the galvanic replacement reaction between Ag nanocrystals and Au ions is spontaneous and driven by the electrochemical potential difference between two metals, it is quite straightforward to form Au shells on the Ag core (Ag@Au) by titrating an Au precursor into a solution of Ag nanoparticles.92 The thickness of the Au shell can be controlled by the volume of the Au precursor. It may also be feasible to reduce the Au precursor using strong reducing agents to form Au nanocrystals on the surface of the Ag core.93 The galvanic replacement reaction still happens during the synthesis. In contrast, the formation of the Au core–Ag shell structure (Au@Ag) can only be achieved by reducing an Ag precursor to deposit Ag on the surface of Au nanocrystals. Generally, in Ag@Au nanocrystals, the shape of the Au shell is highly consistent with the shape of the Ag core, while the deposition of Ag can lead to the formation of various shell morphologies on an Au core such as Au nanosphere@Ag nanocube, Au nanotriangle@Ag nanobipyramid and Au nanodecahedron@ Ag nanorod.94 The shell shapes are highly dependent on the size, facet shape and capping agent of the Au seed, the precursor, and reducing and capping agents used for Ag shells, along with solvent, temperature, etc.95 For example, Huang et al. formed a spiky Au@Ag core–shell structure using different sized Au nanostars as seeds and the Ag shell thickness was tuned from 1 nm to over 10 nm.96 Xia's group precisely adjusted the edge lengths of Ag cubes from 13.4 nm to 50 nm using highly uniform single-crystal Au cuboctahedrons as seeds to form Au nanosphere@Ag nanocube.92 Besides the single shell structure, it is also possible to form multiple-shell nanoparticles. For instance, Karam et al. generated gold–silver–gold core–shell–shell nanoparticles via reducing Ag+ onto the gold core followed by depositing gold through a stronger reduction and galvanic replacement reaction.74 Th thickness of each shell was regulated by reaction agents (reducing agent and precursor) and the volume of the gold precursor, respectively.
Plasmonic properties of core–shell heterojunctions. Most of the abovementioned metallic nanoparticles are symmetric; and only the dipole mode can be excited under normal incident illumination, while higher-order multipole modes are hard to excite. Although generating a hybridization mode by coupling two dipole modes can lead to new resonance and enhance light management, achieving higher-order multipolar modes remains a great challenge for improving the plasmon effect.97,98 The core–shell structures, especially asymmetric structures, with variform core and shell shapes or with the core offset from the center, can excite higher-order multipole plasmon modes by coupling to the normal dipole mode due to symmetric breaking.99,100 The generation of higher order modes can significantly enhance their coupling efficiency with light. For example, silica@gold core–shell nano-eggs were found to exhibit multipole peaks and induced greater near-field enhancement when the core offset was increased. They also experienced a larger absorption-to-scattering ratio at the dipole resonance.100,101

2.6. Metasurfaces based on plasmonic nanoparticle assemblies/arrays

When multiple plasmonic nanoparticles are assembled to form a complex plasmonic system, their localized surface plasmon modes, characterized by a single-resonant nature, exhibit strong interaction, resulting in new hybridized modes across various resonant wavelengths along with spatial mode overlap. As such, besides LSPR modes at nanoscale metal–dielectric interfaces, there may exist propagating surface plasmons such as surface plasmon polariton (SPP) modes at continuous metal–dielectric interfaces and delocalized plasmonic modes (or lattice plasmon modes) in periodic metal–dielectric nanoarrays. Unlike LSPs, SPPs cannot be directly excited due to the lower momentum of a free-space photon compared to that of an SPP. To overcome this momentum mismatch, there are three techniques: prism coupling, scattering from a surface defect such as a subwavelength protrusion or hole and the introduction of periodic corrugation on the metal surface.98,102–105 Unlike LSPs, an SPP propagates over considerable distances, often reaching hundreds of micrometers along the surface. This distinctive feature offers a notable advantage as it permits the incident laser to avoid direct exposure to the measured sample.104 In contrast, lattice plasmon modes can be triggered when plasmonic nanoparticles are meticulously arranged in an ordered array pattern, where scattered light generated by the nanoparticles produces diffracted waves, which are coupled with the LSPR of individual nanoparticles. Lattice plasmon modes provide an unparalleled narrow plasmon resonance band and a high quality factor as compared to LSPR.102 The qualify factor (Q) is defined as the ratio of the central wavelength to the full width at half maximum (FWHM) of the resonance band, which reflects the maximum energy stored in the resonator to the maximum energy losses in a cycle.

Plasmonic modes in multiple nanostructures can undergo further coupling, resulting in the formation of new hybridized modes, resulting in the manifestation of multi-resonant plasmonic characteristics.106 This process enhances the flexibility in tailoring resonance, leading to a high Q value and elevated electromagnetic field properties.98,104 The fundamental concept of plasmonic mode hybridization can be elucidated as the amalgamation of plasmons that arise from basic geometric shapes, combining to form an interconnected system. The plasmonic features of metallic nanostructures are influenced by the electromagnetic interactions among these “free” plasmons. The interaction induces various phenomena like hybridization, splitting, and shifts in plasmons.106 The strength of coupling between two plasmonic modes determines the rate of energy exchange between them.98 To achieve robust coupling, it is crucial to satisfy conditions such as the non-orthogonality between the modes, spectral overlap of mode resonant energies, and congruence in mode profiles. In the strong coupling regime, the rate of energy exchange between the coupled oscillators significantly exceeds the rates of energy leakage. This results in a noticeable shift in their resonances away from their original modes, a phenomenon known as Rabi splitting.107 Even though strong coupling may not occur, other notable resonance modes may emerge, such as Fano resonance, electromagnetically induced transparency, and the Kerker effect.98 Generating such resonance modes can result in a high Q-factor, or/and high electromagnetic enhancement, or a combination of these benefits.104,108 More details can be found in our previous review article.104

To enhance the flexibility in designing metasurfaces, the creation of plasmonic metasurfaces extends beyond simply assembling nanoparticles on a film, which requires special nanofabrication techniques to create nanostructure arrays.104 A desirable nanofabrication method should be cost-effective and high throughput, offering excellent resolution, and allowing for control over nanostructure dimensions and configurations.104 Among various nanofabrication methods,104,109–115 electron beam lithography and focused-ion beam lithography feature high resolution and high controllability, but low throughput. Anodic aluminum oxide (AAO) and nanosphere lithography are known for their high throughput and high resolution, but offer less controllability and less flexibility. Nanoimprinting offers high throughput and high resolution, good controllability and moderate flexibility; however, creating high resolution nanoimprint templates is expensive and wearing-out of nanoimprint templates is of concern during long-term operation. Dip pen lithography and interference lithography have also been explored for nanofabrication. These nanofabrication approaches offer multiple choices and flexibility in engineering plasmonic “hot spots”, dipole–dipole coupling including in-plane plasmon coupling and out-of-plane coupling.116 The details of nanofabrication and the resulting plasmonic nano-arrays can be seen in our previous review article.104

3. Optical probes and point-of-care testing using Au and Ag nanoparticles

Gold and silver nanoparticles have been widely used in sensors and detection devices.1,117 This section is focused on the applications of gold and silver nanoparticles in lab-on-chips and microfluidic devices that can be directly operated by laypersons in point-of-care settings such as home, roadside, clinics, community, emergency departments and remote regions. For optical point-of-care testing, optical probes (optical reporters or labels), such as colorimetric, SERS and fluorescent probes/labels, play an important role in signal acquisition, transduction and amplification. The use of gold and silver nanoparticles in optical probes can improve the performance of optical probes significantly by utilizing their LSPR effects.

3.1. Colorimetric (plasmonic) probes based on Au and Ag nanoparticles

Probe structure and properties. The color of Au and Ag nanoparticles originates from LSPR, resulting in color intensity orders of magnitude stronger than that of traditional organic dyes. For example, monodispersed 15 nm sized Au NPs in an aqueous solution exhibit strong light absorption at around 520 nm, absorbing the green–blue portion and transmitting the remaining red light.16,118 For larger nanospheres or non-spherical nanoparticles, both absorption and scattering occur, resulting in a more complexcolor that may vary at different angles. The color intensity can be described as a local field intensity enhancement factor (LFIEF) or EM field enhancement, which is proportional to the square of the electric field amplitude,40
 
image file: d3cs00793f-t11.tif(15)
where E is the electric field amplitude at a specific point of nanoparticles, while E0 refers to the intensity of the incoming field at that point. When NPs aggregate or undergo etching to some extent, the color of the NPs will change caused by a shift in the LSPR peak.

To apply optical probes to detection systems, it is essential to functionalize the particle surface through the bio-conjugation of biomolecules.119 A variety of functionalization methods have been widely implemented, including (i) physical adsorption of antibodies onto the surface of citrate-stabilized NPs through electrostatic interactions between the negatively charged NP surface and the positively charged groups in the antibodies (i.e., positively charged amino acids and the N-terminal), as well as the hydrophobic interactions between the hydrophobic parts of the antibody and NP surfaces, and (ii) chemical adsorption of molecules with thiolate or dithiolane moieties onto the gold and silver surface through Au–S or Ag–S bonds, and (iii) chemical adsorption of molecules or proteins through carbodiimide cross-coupling (the amine groups of biomolecules can be linked to the carboxylate group-anchored on the NP surface through N-hydroxysuccinimide (NHS)/N-ethyl N-[3-dimethylaminopropyl]carbodiimide (EDC) activation). Other similar reactions for surface functionalization include maleimide–thiol reaction, click chemistry (i.e., azide–alkyne Huisgen cycloaddition) and streptavidin and biotin interaction.120

Because the color of optical probes arises from LSPR, nanoparticles with high EM field enhancement such as nanostars, nanorods and porous nanoparticles are typically used to achieve intense color.121,122 Additionally, compared to traditional gold nanospheres, preparing AuNPs aggregates as probes can enhance signal contrast due to formation of “hot spots” when multiple AuNPs are in proximity.123,124 For example, Hu et al. interconnected AuNPs through oligonucleotide hybridization. These AuNP aggregates not only ensured enhanced color intensity, but also allowed for improved capturing efficiency due to the relatively higher loading of capture ligands on conjugates (Fig. 5(a)).125 Additionally, charge-based aggregation provides an alternative strategy for protein detection.126,127 Compared to the pre-aggregate colorimetric probes, in situ aggregation of NPs provides another strategy for increasing the density of AuNPs. Dual AuNP–antibody conjugates used as labels were proposed by Choi et al. (Fig. 5(b)).128 They used small AuNPs coated with antibodies and bovine serum albumin (BSA) as the first label. After binding with analytes, big AuNPs conjugated with anti-BSA antibodies further bond to small AuNPs through antibody–antigen interaction. This design greatly improved the sensitivity of troponin I detection compared to single AuNP assays. Similar to in situ aggregation, in situ growth or etching of nanoparticles are alternative approaches to enhance and decrease the color intensity, respectively, as well as induce shifts in the plasmon band (Fig. 5(c)).121 For instance, the growth of Ag on gold nanoparticles greatly improved the color intensity due to the plasmon coupling between silver and gold.121 The capture antibody was labeled with alkaline phosphatase (ALP) as an enzyme to catalyze the formation of L-ascorbic acid (AA), which reduced Ag ions into Ag atoms and activated the growth of Ag inside Ag–Au nanocages. This assay had achieved a limit of detection (LOD) for the human carcinoembryonic antigen (CEA) in buffer approximately 3 times lower than a traditional ALP immunosorbent assay.


image file: d3cs00793f-f5.tif
Fig. 5 Signal enhancement mechanism of the colorimetric probe. (a) Aggregating AuNPs; (b) in situ aggregation of NPs, as probes can enhance signal contrast due to the formation of more hot spots when several AuNPs are in proximity; and (c) in situ growth or etching of nanoparticles are alternative approaches to enhance and decrease the color intensity, respectively, as well as induce shifts in the plasmon band. Reproduced with permission from (a) ref. 125, copyright 2013, Royal Society of Chemistry, (b) ref. 128, copyright 2010, Elsevier, (c) ref. 121, copyright 2021, American Chemistry Society.
Applications of colorimetric probes in paper-based microfluidics. Generally, human eyes are more sensitive to red than yellow, so gold nanospheres with an obvious red color are more suitable for use as probes in colorimetric paper-based lateral flow strips (PLFSs) compared to silver nanospheres, which appear yellow. The easy oxidation of AgNPs is also of concern. Currently, urine pregnancy test strips and COVID-19 antigen rapid detection kits are two successful examples for point-of-care testing; and both employ gold nanoparticles as colorimetric probes.130 The two paper lateral flow assays are inexpensive, user-friendly, and easy to manufacture. The scope of PLFSs is extended gradually from small molecules and proteins to metal ions and nucleic acids. Basic components of PLFSs include a sample pad to load liquid samples, a conjugation pad with the preloaded detection bioreceptor or optical probes, a nitrocellulose microfluidic membrane immobilized with at least one capture bioreceptor (test line) and a control line, and an absorbent/wick pad for ensuring complete flow and storage of the liquid sample.117 Hu et al. demonstrated a colorimetric lateral flow assay for the detection of nucleic acids.125 Taking advantage of the hybridization between two complementary oligonucleotides, they produced AuNPs with two different thiol labeled oligonucleotides for aggregation and detection, respectively. Such AuNPs were further reacted with other AuNPs labeled with complementary strands, leading to the formation of aggregations of AuNPs as the final probes on the conjugation pad. Unlike antibodies, which can be directly immobilized on the nitrocellulose membrane, it is hard to fix the capture DNA strand directly on the membrane. These capture strands needed to be labeled with biotin, which can react with streptavidin anchored on the nitrocellulose membrane. As a result, the aggregated AuNPs probes showed higher color intensity than conventional AuNPs, showing a LOD of 0.1 nM toward human immunodeficiency virus type 1 (HIV-1) under buffer conditions.

Developed from single analyte detection, multiplexed detection has become increasingly important due to its ability to save costs and time. One of the most established methods is to engineer multiple test zones on a single paper strip.131,132 Nadezhda et al. even created microarray spots containing different immobilized immunoreagents as the detection zone in a single 5-mm-wide strip (Fig. 6(a)). This device can measure up to 32 illicit drugs using gold nanoparticles with different antibodies that bind to different analytes selectively.133 Even though this method is high throughput, it is necessary to optimize the testing conditions to eliminate cross-selectivity and reduce false positive outcomes. An alternative approach is to directly integrate several single-target test strips into one cassette. These individual test strips share the same sample pad but the sample flow on each strip is separated, leading to low mutual interference and cross-contamination. A variety of integrated PLFSs have sprung up recently including “one-direction”, “two-direction”, “disc-sign”, etc.132,134 One typical example is a disc-like multi-branched paper device for the detection of copper ions using AgNPs.135 It has eight test zones, which are first loaded with analyte copper solutions, followed by the addition of AgNPs into the center loading zone. The loaded AgNPs then migrate to the test zone and aggregate in the presence of Cu2+ ions, leading to a different color change.


image file: d3cs00793f-f6.tif
Fig. 6 Applications of colorimetric probes. (a) Multiplexed detection in a single paper strip with microarray spots containing different immobilized immunoreagents as the detection zone; and (b) integration of different functional zones in one PDMS-based chip for visual detection. Reproduced with permission from (a) ref. 133, copyright 2013, Springer Nature, (b) ref. 122, copyright 2020, American Chemical Society.
Applications of colorimetric probes in PDMS-based microfluidics. Compared to paper-based microfluidics, polydimethylsiloxane (PDMS)-based microfluidic devices offer more flexibility in controlling the velocity, volume, direction and distance of liquid flow via pumps, valves, mixers and separation modules. It can also be engineered to allow for sequential programming of sample pre-treatment and detection. For instance, Liu et al. presented a colorimetric immunoassay built on a PDMS-based microfluidic device. It was used for the detection of human immunodeficiency virus (HIV) antigens, p24 (Fig. 6(b)).122 Initially, the aqueous reagents were stored in reservoirs separated by mineral oil. When the sample was loaded into the channel, the target antigen was captured by the antibody-labeled magnetic beads to form conjugates, which were dragged to the next two reservoirs to remove unbound proteins, followed by interaction with the HRP-labeled detection antibodies. The labeled HRP catalyzed the oxidation of tetramethylbenzidine (TMB). This device can further mediate the etching of gold nanorods to induce a color change, enabling the visual detection of p24 with a LOD of 0.5 ng mL−1.

3.2. Fluorescence probes based on gold nanoparticles

Plasmon-enhanced fluorescence. Compared to colorimetry, fluorescence involves the emission of relatively short-ranged light in response to narrow-wavelength excitation light so that the signal-to-noise ratio is higher. Moreover, NIR fluorophores in the biological transparent window show advantages over colorimetric probes when testing plasma, serum and blood samples. However, the quantum yield of NIR fluorophores is only 10% or much less in the NIR-I window and <1% in the NIR-II window. Hence, fluorophores are coupled to plasmonic gold nanoparticles and Ag nanoparticles to improve their quantum yield.117 The concept and principles of plasmon-enhanced fluorescence have been described in detail in our previous literature.117,136 In brief, when the LSPR band of plasmon overlaps with the absorption spectrum of a fluorophore, excitation enhancement of fluorescence can occur due to an increase in the excitation rate of the fluorophore at both short and long distances between the fluorophore and the plasmonic particle. Excitation enhancement may occur via different processes such as plasmonic light scattering, plasmon-induced resonance energy transfer (PIRET), and Purcell effect. When the LSPR band of plasmons overlaps with the emission spectrum of fluorophores, emission enhancement of fluorescence will take place at long distances via the Purcell effect, while emission quenching will occur through the Förster resonance energy transfer (FRET) process at short distances.

For organic fluorophores, the absorption and emission spectra are not far away from each other; the LSPR band of plasmonic metal nanoparticles may overlap with both the absorption and emission spectra. Hence, both excitation enhancement and emission enhancement may occur in plasmonic metal@fluorophores. Owing to the trade-off between excitation enhancement and emission enhancement, plasmonic fluorescence probes are typically formed as metal core@spacer@fluorophore (Fig. 7(a)).1,137 The spacer layer is usually 10–30 nm between metal and the fluorophore to maximize fluorescence enhancement. Materials used as the spacer layer varies from SiO2, polymers, double-stranded DNA and proteins. It is noticed that the thickness of the spacer layer should be tuned reproducibly and easily when choosing a material as the spacer layer. For example, the thickness of SiO2 can be regulated by adjusting the amount of precursor, while double strand DNA can be extended by increasing the number of base-pairs, resulting in an increase of 0.3 nm.117 To label biomarkers to the probe, another coating layer is employed, the metal-core@spacer@fluorophore@coating layer and it can also prevent the interaction between the fluorophore and solvent, which may lead to fluorescence quenching.


image file: d3cs00793f-f7.tif
Fig. 7 Plasmonic fluorescence probes. (a) Schematic and TEM images of the plasmon-based fluorescence probe structure; (b) emission enhancement of Cy5.5 can be tuned by adjusting the shape of probes such as gold nanorods, nanobipyramids, nanoprisms and nanostars. The figure (a) is further modified from ref. 137, copyright 2018, American Chemical Society. Reproduced with permission from (b) ref. 129, copyright 2021, American Chemical Society.

Plasmon-enhanced fluorescence depends on the shape and structure of particles. It was demonstrated that gold nanorods, nano-bipyramids, nano-prisms and nanostars showed different degrees of emission enhancement for Cy5.5 dye (Fig. 7(b)).129 Gold nanostars exhibited the highest enhancement, attributed to their greater number of “hot spots”. Even for the same shapes, such as gold nanorods, fluorescence intensity is also influenced by the aspect ratio, which determines the Q factor and the plasmon band position. Abadeer et al. synthesized a series of gold nanorods with aspect ratios ranging from 1 to 4.4, and coated them with a silica shell on the surface to illustrate the intensity changes of IRDye 800CW dye.138 The maximum fluorescence emission was achieved with an aspect ratio of 3.7 with the maximum overlap of the plasmon spectrum with the absorption of IRDye. When the silica layer was 17 nm, the highest enhancement was achieved to be around 10-fold.

Applications of fluorescence probes in paper-based microfluidics. Due to the nature of plasmon-enhanced fluorescence, its application can be categorized into “signal-on” and “signal-off” modes. In the former case, plasmon-enhanced fluorescence probes were employed into the paper-based microfluidics in a similar way to the colorimetric counterparts. For instance, a SiO2@Au@QDs structure was prepared using a polyethylenimine (PEI)-mediated electrostatic assembly method.139 Benefiting from the colorimetric gold and the plasmon-enhanced QDs, this method can achieve a dual-signal paper lateral flow immunoassay for the detection of SARS-COV-2 antibodies. It was noticed that fluorescence sensitivity was higher than the colorimetric signal. In addition, Wang et al. demonstrated the “signal-off” mode in paper strips based on the FRET mechanism (Fig. 8(a)).140 A fluorescein isothiocyanate (FITC)-labeled capture antibody was initially stabilized on the nitrocellulose membrane as the test line, while a gold nanoparticle with a detection antibody was precoated on the conjugation pad. When the samples containing carcinoembryonic antigens (CEA) were added to the sample pad, the sandwiched structure formed by the detection antibody, antigen and capture antibody realized FRET from fluorophore FITC to the gold, leading to the quenching of the fluorescence, namely, a lower signal represents a higher concentration.
image file: d3cs00793f-f8.tif
Fig. 8 Applications of fluorescence probes in microfluidics. (a) “Signal-off” mode in paper strips based on the FRET mechanism; and (b) integrated gold nanorod arrays into PDMS-based microfluidics chips for one-step immunoassay. Reproduced with permission from (a) ref. 140, copyright 2018, Elsevier (b) ref. 141, copyright 2021, Springer Nature.
Applications of fluorescence probes in PDMS-based microfluidics. As demonstrated by Zheng et al., a gold nanopyramid array pattern was fabricated by nanosphere lithography.110 This plasmonic array pattern was used not only as a biomodification platform for capture antibodies, but also as a fluorescence enhancer when using a silica layer to tune the distance from the fluorophore. It successfully enhanced the NIR fluorophore emission on the detection antibody, achieving a LOD of 0.6 pg mL−1 for glial fibrillary acidic protein (GFAP) in blood plasma. Wang et al. further integrated a gold nanorod array into a microfluidics chip for one-step immunoassay based on a similar enhancement mechanism by using a self-assembled monolayer (SAM) of 3,3′-dithiodipropionic acid di(N-hydroxysuccinimide ester) (DSP) to regulate the distance between the nanorod and the fluorophore (Fig. 8(b)).141 The microfluidic chips contained a detection antibody deposition zone, a reaction chamber with gold nanorod array substrates, and a capillary pump. After pipetting 20 μL of human serum samples into the inlet of the chip, the target analyte in the serum first interacted with the detection antibody with a fluorophore, and then it was captured by the detection antibody on the nanoarray located in the detection zone. A CCD camera was used to monitor the fluorescence strength in the detection zone and record the results.

3.3. SERS probes based on gold nanoparticles

Probe structure and properties. Because the scattering cross-section of Raman reporters is only 10−28–10−30 per molecule, the Raman signal is weak when detecting a low concentration of analytes in an assay. Therefore, plasmonic nanostructures have be widely used to generate SERS to amply signals. SERS signal amplification originates mainly from electromagnetic (EM) field enhancement and can achieve an enhancement factor in the range of 104–108.117,136 The energy of interaction between the oscillating electric field of a photon and the Raman reporter is dependent on the polarizability (α) of a molecule and EM field strength (E):1
 
image file: d3cs00793f-t12.tif(16)

Owing to a small Raman scattering cross-section, plasmon quenching is negligible. EM enhancement can be applied to both excitation and scattering processes, leading to an overall enhancement of |Eloc|4. It can be seen from the above equation that stronger polarization and EM fields can generate higher SERS signals. Thus, fabrication of plasmonic nanostructures with high EM field enhancement is the key to achieving high sensitivity and reproducibility of SERS probes.

Given that the decay of the LSPR-induced EM field is in the 10–30 nm range, Raman reporters should be near the metal surface to ensure the strongest EM field for signal amplification.1 Thus, metal core@Raman reporters are a typical form of SERS probes. Here, Raman reporters are usually required to have high polarizability, such as aromatic molecules, to ensure a relatively large scattering cross-section. Additionally, Raman reporters with SH– and NH2– groups are representative molecules for SERS since these groups show great affinity to the Au or Ag surface. Typical Raman reporters include 4-mercaptobenzoic acid (4-MBA), 4-aminothiophenol (4-ATP), 4-mercaptopyridine (4-MPY), malachite green isothiocyanate (MGITC), and 5,5-dithiobis-2-nitrobenzoic acid (DTNB).117 However, it is found that Raman reporters get detached from the metal surface during sensing or SERS probes aggregate in high ionic strength solutions. Also, it is difficult to bio-conjugate biomolecules such as antibodies onto the metal surface owing to the limited occupancy of the surface after coating with Raman reporters. Similar to fluorescence probes, another coating layer such as silica and polyethylene glycol (PEG) is typically applied to cover the metal-core@Raman reporters’ surface, forming a “metal-core@Raman reporters@coating layer”.117 This thin layer can not only eliminate the above leaking and limited space issues, but also provide excellent water-solubility and biocompatibility. The metal core can be tuned into different shapes to achieve higher EM enhancements (Fig. 9).


image file: d3cs00793f-f9.tif
Fig. 9 SERS probes. (a) Metallic nanospheres modified with Raman reporters (NSs@Raman reporters); (b) metallic nanospheres with Raman reporters covered with polymer shell (NSs@Raman reporters@polymer); (c) metallic nanospheres with Raman reporters covered with silica shells (NSs@Raman reporters@SiO2); (d) star-shaped metallic nanoparticles labeled with Raman reporters (NSTs@ Raman reporters); and (e) TEM images of star-shaped metal nanoparticles with Raman reporters encapsulated in a silica shell. (f) EM field enhancement can be modulated by the shape of gold nanoparticles. The figures are further modified from (e) ref. 142, copyright 2017, American Chemical Society, (f) ref. 9, copyright 2012, Institute of Physics.

Wu's group has developed the classic sandwich-structured SERS probes with a configuration of gold core@Raman reporter@silica, as shown in Fig. 9(e).9,142 They used Au nanospheres, Au nanorods, and Au nanostars as the metal cores, respectively, and compared the SERS enhancement factor of three different metal cores under 532 nm and 785 nm laser excitation, respectively. It was found that nanostars with abundant sharp tips generated the strongest EM field (Fig. 9(f)) among three metal cores, leading to the highest SERS enhancement.9 Besides engineering nanoparticle shapes with more sharp spikes, numerous plasmonic substrates have been designed and fabricated using electron beam lithography (EBL), nanoimprinting, nanosphere lithography, anode aluminum oxide (AAO) templates, etc. Wu's group demonstrated that coupling Au nanostars to Ag nanopyramids can significantly amplify SERS signals, resulting in the sensitive detection of nitrite in water.143 Briefly, the analyte nitrite reacts with 1-naphthylamine (1-NA) on AuNSTs and 4-aminothiophenol (4-ATP) modified on the Ag nanopyramid, forming the azo group, which is sandwiched by the AuNSTs and the nanopyramid array. The generated azo group showed a greatly enhanced SERS signal as the in situ chemical reaction took place, leading to intriguing sensitivity with a LOD of 0.6 pg mL−1 toward nitrite detection in water.

Applications of SERS probes in paper-based microfluidics. Commercial PLFS currently employs plasmonic gold nanoparticles as colorimetric probes for testing urine, saliva and nasal swap-transferred buffer samples. When applied to serum, plasma and whole blood samples, they suffer from low sensitivity, repeatability and poor performance due to severe optical interference of biomolecules in the complex sample matrix.144,145 The optical signals of SERS probes can be tuned to fall into the biological transparent window (NIR-I and NIR-II) under excitation of a 785 nm or 1064 nm laser, which made it suitable for testing blood-related samples in paper-based microfluidics.146 Wu's group has fabricated a SERS probe by sandwiching the Raman reporter (4-MBA) between the Au nanostar and a thin SiO2 layer functionalized with the detection antibody, and incorporated such probes into a PLFS (Fig. 10(a)).142 They compared the performance of the SERS–PLFS with that of the conventional colorimetric PLFS for the detection of neuron-specific enolase (NSE), a traumatic brain injury (TBI) biomarker in human plasma samples. The results showed that the SERS–PLFS achieved a LOD of 0.86 ng mL−1 toward NSE detection in a plasma sample matrix, which was almost 3 orders of magnitude lower than that of the colorimetric counterpart. To further explore the limits of sensitivity, an Au nanopyramid array substrate was inserted into the middle of the nitrocellulose membrane using a surfactant bridge to connect the neighboring paper parts (Fig. 10(b)).147 As capture antibodies were immobilized on top of Au nanopyramid array chips, the SERS probes conjugated with S-100β, a TBI biomarker, were captured on the substrate. The created hierarchical 3D plasmonic field between SERS probes and Au nanopyramids remarkably enhanced the SERS signal, resulting in a LOD down to 5.0 pg mL−1 toward S-100β detection in human plasma. The sensitivity was comparable to that of the most sensitive ELISA assay, but the testing time (around 30 min) was much shorter than the ELISA counterpart (2 h). When detecting whole blood samples, besides the strong visible interference from the red color of red blood cells, several-micrometer sized blood cells can severely disturb the flow rate of fluids in paper strips whose pore size is of the same order. This is why the generally used PLFS procedures are performed in blood plasma by centrifuging blood samples to get rid of blood cells instead of directly using whole blood samples. To reduce the sample treatment process and make PLFS testing more user-friendly, Wu's group integrated a plasma separation unit onto the PLFS based on the principle of salt-induced blood cell aggregation and created an “all-in-one” paper test strip.148 The apparent separation efficiency and yield can reach up to 90% and 30%, respectively. When a drop of whole blood (30 μL) was directly loaded on the plasma separation unit, blood cells were trapped by the salt treated filter, while plasma migrated to the bottom layer assisted by the surfactant on the filter membrane. Then, a buffer solution was applied to the strip to help plasma flow through the strip. The tumor-associated protein biomarker, CEA, spiked in the whole blood was captured on the test line and monitored using a portable Raman reader, achieving a LOD of 1.0 ng mL−1. This has demonstrated that the SERS-based paper test strip can be used as a POCT tool for testing finger-prick blood samples.
image file: d3cs00793f-f10.tif
Fig. 10 Applications of SERS probes in microfluidics. (a) SERS–PLFS shows a three orders of magnitude lower LOD than the conventional colorimetric PLFS; (b) Au nanopyramid substrate integrated PLFS shows remarkably enhanced SERS signals due to a hierarchical 3D plasmonic field between SERS probes and Au nanopyramids; (c) PDMS-based SERS biochips can directly identify the multiple point mutations in tumor cells; and (d) a wearable sweat sensor to recognize fingerprint information of targets in sweat with Ag nanomushroom array as SERS substrate in PDMS-based microfluidics. Reproduced with permission from (a) ref. 142, copyright 2017, American Chemical Society, (b) ref. 147, copyright 2021, Elsevier, (c) ref. 150, copyright 2020, American Chemical Society, (d) ref. 152, copyright 2022, Springer Nature.
Applications of SERS probes in PDMS-based microfluidics. When integrating SERS probes into PDMS-based microfluidics, a nanoarray pattern can also be employed as a detection substrate to couple with SERS probes and generate higher “hot spots” for signal amplification. Benefiting from their narrow spectral width and high multiplexing capabilities, SERS probes can be applied to the detection of multiple biomarkers in PDMS based microfluidics.149 Wu et al. have developed an SERS biochip to directly identify the multiple point mutations in tumor cells (Fig. 10(c)).150 The 3 stripes of Au@Ag nanorod array substrates in the biochip were patterned with KARS, BRAF, and PIK3CA molecular beacons, respectively, and each end of these molecule beacons was labeled with 3 different Raman reporters. When a group of samples containing a mixture of mutations was analyzed, a signal decrease in SERS probe intensity was observed due to the movement of Raman reporters away from the metal nanoarray substrate. The dependence of SERS signal intensity on the concentration of target DNA, and the ratio of three mutations was then investigated. Different from fluorescence-based approaches which are limited by the overlapping spectra and photobleaching issues of fluorophores, SERS-based microfluidics can even be used for simultaneously monitoring several targets in one substrate due to their efficient molecular fingerprint features. This method is necessary especially when it comes to extremely small volume and concentration samples, as well as one target (i.e., cell) containing multiple analytes. For example, subpopulations of extracellular vesicles (EVs) share similar morphological and physical properties, making it difficult to isolate them from bulk populations. To measure multiplex phenotyping of EVs simultaneously, Want et al. have developed a biochip composed of a single substrate modified with capture antibodies and AuNPs conjugated with target antibodies and different Raman reporters.151 Together with alternating current electrodynamics guiding lateral fluid movement, this biochip can ensure sufficient antibody–antigen collision and shear off nonspecific molecules, resulting in successfully presentation of EV phenotypic heterogeneity and patient treatment responses. Moreover, the label-free, multiplexing microfluidics can be realized by utilizing the unique biomolecule-specific Raman fingerprints for differentiation. As demonstrated by He et al., a PDMS microfluidic device was made into a wearable sweat sensor to recognize the fingerprint information of targets in sweat including urea, lactate and pH (Fig. 10(d)).152 The detection zone consisted of an Ag nanomushroom array as an SERS substrate, where the nanogap sites can generate a high electromagnetic enhancement effect, increasing the sensitivity of each biomarker. Driven by the capillary force in the microchannels, the sweat flew into the microchannel, and the targeted analytes in sweat showed distinct SERS signals when falling into nanogap sites. This made it possible to integrate SERS-based microfluidic devices with a dynamic biofluid sensing system for personalized medicine.

Label-free SERS is an effective way to collect a large amount of data and offers multiplexing capability due to its inherently narrow and fingerprint-specific peaks. However, it is a challenge to obtain quantitative and reproducible SERS results due to several factors.153 One of reasons is that the quantitative SERS signals are affected by variations in the adsorption of analytes to the substrate as well as its rich and complex spectroscopic features since the overall signal intensity in SERS is essentially the sum of signals from all the analytes in the whole detection area covered by the excitation laser beam. Furthermore, the random distribution of “hot spots” on the SERS substrate presents a challenge where only a fraction of analytes are exposed to “hot spots”. Laser intensity and laser beam size (diameter) may also affect the reproducibility of SERS signals. To address these challenges, internal calibration standards are employed, including isotopes of analytes for correction, molecular Raman tag-based internal standards, and background scattering.154–156

In addition to internal standards, machine learning (ML) analysis stands out as a valuable approach for the analysis of label-free and labeled-SERS spectra. ML techniques, including various algorithms and models, enable the identification of patterns, relationships, and trends within the data, thereby helping interpretation of SERS signals.157 It has been applied to biosensing, including cancer screening, drug testing, food safety monitoring, pathogen and toxic substance detection.158 Some popular ML methods include principal component analysis combined with linear discriminant analysis (PCA-LDA), partial least-squares discriminant analysis (PLSDA), classification tree (CT), support vector machine (SVM), k-nearest neighbor (KNN) and artificial neural network (ANN) method.159 Wonil and colleagues conducted a comparison between five machine learning methods for multi-class classification of cellular responses to drugs.159 Their findings indicated that PCA-LDA and PLSDA had high misclassification rates, particularly noticeable at drug dosages lower than IC50. In contrast, the nonlinear-based ANN method demonstrated a markedly higher statistical sensitivity and specificity (ranging from 90% to 99%), effectively classifying the control, two sub-IC50, and IC50 treated living cancer cells. Meanwhile, CT and KNN were found to be ineffective for classifying the drug responses of living cancer cells. The performance of SVM was noted to be comparable to that of ANN, albeit slightly less effective.

3.4. Electroluminescence (ECL) sensors based on gold nanoparticles

The principle and structure of ECL. Electrochemiluminescence (ECL) is a process wherein stable precursors at the electrode's surface generate reactive intermediates upon the application of a voltage. Subsequently, an electron transfer process occurs, leading to the formation of excited states that emit light. The reaction mechanisms include annihilation, co-reactant ECL and cathodic luminescence. As shown in Fig. 11, annihilation involves an energy transfer reaction between an electrogenerated radical anion and cation, both of which react with each other, forming an excited state.160 Then the excited state species return to the ground state and emit light. Different from annihilation which is a double-potential step, co-reactant ECL is a single-potential step. It refers to a process where both the luminophore and co-reactant are oxidized or reduced to generate intermediates through a single-potential step. These intermediates react with each other to form an excited state of luminophores and emit light. Annihilation and co-reactant ECL are electron transfer chemiluminescent reactions, while cathodic luminescence is the direct formation of the excited state at the oxide-covered metal (i.e. aluminum and tantalum) electrode. For instance, at oxide-covered aluminum electrodes, the emission from Dy(III), Sm(III), and Tb(III) has been detected during the reduction of hydrogen peroxide, persulfate, or oxylate in an aqueous solution. This form of cathodic luminescence at high voltages is suggested to be a consequence of injecting hot electrons into the electrolyte solution, which may lead to the formation of excited states. The available chromophores used in ECL include small chromophores (i.e., Ru(bpy)32+), polymers, and quantum dots, realizing wavelength-tunable light emission. In some ECL sensors, a capture element (i.e., antibodies, aptamers, and enzymes with high specificity of the analyte) is incorporated on the working electrode's surface to selectively capture the target analyte of interest.161 After the analyte is captured on the surface of working electrode, a sandwich structure is formed by the capture and detection elements and the reactants on the detection element are then excited to emit light when the voltage is applied to the electrode.
image file: d3cs00793f-f11.tif
Fig. 11 Electrochemiluminescence detection. (a) Annihilation ECL mechanism; (b) ECL from Ru(bpy)32+; (c) ECL enhancement with AuNPs. (a) and (b) These figures are further modified from ref. 160, copyright 2004, American Chemical Society, (c) reproduced with permission from ref. 167, copyright 2009, Royal Society of Chemistry.

As electrochemiluminescence (ECL) does not rely on external light excitation, it circumvents the potential issues of auto-photoluminescence and light scattering in analytical measurements, leading to a higher signal-to-noise ratio and improved detection accuracy.162 The redox reaction is controlled by voltage; thus, ECL reactions are generally fast and have short reaction times, enabling real-time and rapid measurements in various applications.163 Benefiting from the development of thin-film and thick-film technology, it is feasible to miniaturize ECL sensors and integrate them into portable devices for on-site and point-of-care testing applications.164,165

Although ECL has lower background interference than photoluminescence due to its excitation by voltage, ECL's quantum efficiency is lower than that of the photoluminescence process because it involves highly reactive intermediates, and the population of the excited state is also restricted due to the mass-transfer process of redox species in the solution. For instance, the ECL efficiency (defined as photons generated per redox event) of Ru(bpy)32+ is around 5%.160 Lately, several signal amplification techniques have been suggested to enhance the analytical capabilities of ECL sensors including enzyme-mediated amplification, DNA-mediated amplification and plasmon-enhanced ECL. Biological methods are sensitive to the environment such as pH, temperature, and ionic concentrations. In contrast, plasmonic approaches are more robust and efficient for the ECL enhancement.

The excited state in ECL is typically formed through an electrochemical reaction involving the oxidation and reduction of molecules in the presence of an electric potential. It is unclear that the plasmonic nanostructures do not directly enhance the chemical reaction responsible for the formation of the excited state in ECL. On the other hand, ECL enhancement can be achieved through the mechanism of plasmon-enhanced emission. Similar to the emission enhancement in the plasmon-enhanced fluorescence section, ECL emission is likely to be regulated by the FRET and Purcell effect, depending upon the spectra overlap between the ECL emission peak and plasmon band, and upon the distance between the ECL chromophore and plasmonic nanoparticle.166 Utilizing the conformation change of nucleic acids, the distance between ECL emitters and plasmonic structures is regulated, leading to the quenching and enhancement of ECL emission, thus modulating luminescence signals. For example, Chen's group has constructed a specific DNA strand for analyte recognition and detection with one end labeled with CdS:Mn QDs as the ECL luminophores, while AuNPs are modified at another end, serving the dual role of an ECL quencher and an enhancer. Initially, AuNPs got close to the QDs forced by the hairpin structure of DNA strands, leading to the quenching of ECL signals. After hybridization with target DNA, the strand unfolded, separating AnNPs and QDs and causing ECL enhancement (Fig. 11(c)).167

Applications of ECL sensors in paper-based microfluidics. The electrodes are important parts in ECL-based paper microfluidics.168,169 Benefiting from the roughness and porosity of cellulose paper, coating electrodes directly on paper can be realized via thick-film and thin film technologies, which are also suitable for other substrates such as glass, silicon, polymer, and ceramic.165 Thick-film technology depends on the printing process (i.e., screen, inkjet or aerosol jet printing) used to pattern the electrode structure by using conductive inks including carbon, metals and conducting polymers. In contrast, thin-film technology can realize a thickness of nanometers through photolithography and metal deposition. Based on these technologies, one can fabricate film-based three electrodes, and bipolar electrodes on paper connected with conductive paperboard or wires.163,170,171 For example, carbon electrodes can be firmly attached to paper surfaces through screening-printing or ink-writing methods. Other parts such as the ECL reaction zones, electrolyte, and microfluidic channels can be generated through sputtering, screening-printing, ink-writing, soaking, cutting, stacking, origami, etc.172

The basic structure of ECL paper sensors includes planar paper-based devices, and 3D origami paper-based devices.172 Chen's group has fabricated the first ECL planar paper-based devices by screen printing bipolar electrodes on cellulose paper.173 After connecting electrodes with two independent hydrophilic channels using hydrophobic wax boundaries, the anode part was loaded with Ru(bpy)32+, while the cathode was modified with multi-walled carbon nanotubes (MWCNTs) to immobilize the capture anti-prostate specific antigen (anti-PSA) on it. When the analyte PSA was present in the system, PSA was sandwiched between the detection anti-PSA labeled with glucose oxidase and the capture anti-PSA, followed by washing with PBS buffer to remove unreacted species. Subsequently, glucose was introduced into the cathode cell to undergo oxidization by glucose oxidase and to form H2O2in situ. The ECL light output can be initiated by the electrochemical reduction of H2O2 at the cathode, arising due to balance between the anode and cathode of bipolar electrodes. Inspired by lateral flow immunosensors, Hong et al. have developed ECL-based paper lateral flow strips to avoid multiple washing procedures and simplify operation (Fig. 12(a)).171 Instead of using a film backing layer as conventional optical lateral flow strips, they directly located a screen-printed electrode beneath the test line of nitrocellulose membranes to introduce the ECL reaction in paper strips. Also, owing to the use of NC membranes, the capture antibody could be directly immobilized on it through electrostatic interaction rather than covalent bonding using cellulose membranes. Specifically, an ECL probe was synthesized by first covalently bonding Ru(bpy)32+-NH2 to the carboxylated surface of nonporous silica nanoparticles, and then further conjugating antibodies to the probe through the primary amine groups of the antibodies. A sandwiched structure was formed after adding the sample solution in the sample pad with the remaining unreacted probes continuing to flow to the absorption pad. After 15 min, tripropylamine (TPA) was then added to the strip to perform ECL measurement at a given potential. The assay operation took 20 minutes to finish and achieved a LOD of 0.81 pg mL−1 toward troponin I, a biomarker of acute myocardial infarction, detection in human serum, which was three orders of magnitude less than that of the fluorescence assay. Zhan et al. further improved the device through secondary addition of the co-reagent (TPA) after a time interval of immobilizing the Ru(II)–L-Cys complex on the antibody.170 In this way, the assay time was shortened to 7 min for troponin I detection at a LOD of 0.44 pg mL−1.


image file: d3cs00793f-f12.tif
Fig. 12 Electrochemiluminescence devices. (a) ECL-based paper lateral flow strips to avoid multiple washing procedures and simplify operations; and (b) integrated PDMS-based ECL for H-FABP without sample pretreatment. Reproduced with permission from (a) ref. 171, copyright 2020, John Wiley and Sons, (b) ref. 174, copyright 2023, Elsevier.
Applications of ECL sensors in PDMS-based microfluidics. The integration of electrodes into PDMS-based microfluidics is more flexible than in paper microfluidics. The electrodes can be fabricated in PDMS or glass/silicon substrates using thick or thin film technology, and even directly integrated onto the channels, which are enclosed with PDMS sheets.174,175 PDMS is an insulation material and the whole sensor parts enclosed in PDMS can prevent electric leakage and ensure operation safety. Besides electrodes, other parts such as batteries, microcontroller and reader can be integrated as a whole operation-reading system. Additionally, benefiting from the flexible design of PDMS channels, it is feasible to create multiple inlets or solution storage reservoirs for the introduction of probes, co-reagents and washing buffers. As reported by Zhu et al., the electrodes were screen-printed on ITO glass, followed by sealing and aligning with PDMS using soft-lithography technique (Fig. 12(b)).174 This integrated system can directly detect heart-type fatty acid binding protein (H-FABP), a biomarker for acute myocardial infarction, in human serum without any pretreatment.

3.5. Photoelectrochemical (PEC) sensors based on gold nanoparticles

The principle and structure of PEC sensors. In contrast to ECL, which converts electrical energy to light energy, PEC is a reverse process that converts light into electrical energy. It is a light excitation-dependent electrochemical process, in which molecules or semiconductors are excited under light illumination, generating electron–hole pairs, followed by transferring these photogenerated charge carriers to the electrode, resulting in electric signals.176,177 The target analyte in the PEC system can be oxidized (as electron donors) or reduced (as electron acceptors), leading to a change in the electric signal. A general PEC cell includes a photoactive material as the working electrode (WE) for light-harvesting and signal transduction, a counter electrode (CE), and an electrolyte where analytes are present, a light source and signal reading system. Taking direct oxidation of target analytes in a PEC as an example in Fig. 13(a), the photoactive material such as a semiconductor is excited under light irradiation, generating electrons in the conduction band (CB) and holes in the valence band (VB) at the photoanode.176 The photogenerated electrons transfer to the photoanode and holes are injected to the electrolyte. In the presence of a reducing target in the electrolyte, holes in the VB can be further scavenged by the target (DD*), contributing one electron to the semiconductor. Thus, the generated photocurrent is proportional to the concentration of target molecules. In a different case, when oxidant analytes are present in the electrolyte, the photogenerated electrons are scavenged by the target, modulating the photocurrent. Importantly, the prerequisite of the above redox reaction is that the oxidation potential of the target analyte should be more negative than the VB, while the potential of oxidizing species should be more positive than the CB.
image file: d3cs00793f-f13.tif
Fig. 13 Photoelectrochemical cells. (a) PEC mechanism with an example of oxidation of biomolecules; (b) decoration of AuNPs on the graphene–WO3 membrane increased the visible light absorption and improved the catalytic activity of labeled enzyme; and (c) AuNPs not only aid in plasmonic resonance enhancement but also play a role in element recognition in the detection. Reproduced with permission from (a) ref. 176, copyright 2015, Elsevier, (b) ref. 181, copyright 2014, American Chemical Society (c) ref. 182, copyright 2018, Elsevier.

Electrochemical sensors are recognized for their remarkable sensitivity and rapid response. However, this heightened sensitivity comes at a cost since they are susceptible to background interference. Even slight absorption can lead to signal fluctuations, posing a challenge for accurate measurements. As compared to electrochemical sensors, non-specific adsorption on electrodes generally has negligible interference in the sensing signal of PEC. On the other hand, fluorescence sensors are sometimes susceptible to autofluorescence of sample matrix. In contrast to fluorescence sensors, the auto-fluorescence of the sample matrix will not interfere with the sensing signals. Hence, PEC sensors theoretically exhibit a high signal-to-noise ratio and strong resistance to interference of complex sample matrix.178

Plasmon enhanced PEC sensor. Plasmonic metal nanostructures can serve as light antennae, and plasmon energy can transfer from the plasmonic metal to the semiconductor nearby. As a result, coupling of a plasmonic metal nanostructure to a semiconductor can be used to “turn on” or “turn off” the electric signal, or enhance photocurrent. If the presence of an analyte in the electrolyte can modulate this coupling process, a plasmonic PEC can be constructed based on the coupling process. Coupling of a plasmonic metal nanostructure to a semiconductor will result energy transfer via three major mechanisms: (i) light scattering/trapping, (ii) hot electron injection, and (iii) plasmon-induced resonance energy transfer (PIRET). Wu et al. have revealed the balance between these mechanisms, which depends on the dephasing time, and proposed a rational design of plasmonic metal–semiconductor heterojunctions according to these principles.3,4,179 In brief, in the process of light scattering, incident light is scattered by the large plasmonic nanoparticles, allowing it to penetrate the semiconductor, resulting in an increase in photon flux in the semiconductor, which further increases light absorption in the semiconductor. It is worth noting that light scattering cannot extend the light absorption spectral range of semiconductors; instead, it is used to increase the optical path length. Hot electrons are generated by dephasing plasmon non-radiatively via Landau damping and they possess higher energy than those generated through thermal excitation. Plasmonic hot electrons can be directly injected into the conduction band of semiconductors owing to their higher energetics than the Schottky barrier at the interface between metal and semiconductors. PIRET is a dipole–diploe interaction through which plasmonic energy in the metal can be non-radiatively transferred to the semiconductor, generating electron–hole pairs in the semiconductor. Its efficiency is related to the spectral overlap and distance between the metal and semiconductors. Generally, PIRET can achieve the highest photoconversion efficiency among these three mechanisms. Besides the above three mechanisms employed for common metal–semiconductor heterojunctions, quantum dots (QDs) as an emission semiconductor can couple with metal plasmonic nanomaterials. Not only the absorption process is influenced, but the radiative and non-radiative decay after excitation can also be tailored by plasmons.180

To realize signal transduction in plasmon-enhanced PEC sensors, plasmonic metal nanoparticles can modulate the photocurrent response by enhancing photoconversion or as a photoconversion mediator to generate “signal-on” or “signal-off” modes. For example, Paik and co-workers generated a graphene–WO3–Au hybrid membrane for glucose detection (Fig. 13(b)).181 Decoration of AuNPs on the graphene–WO3 membrane increased the visible light absorption and improved the catalytic activity of the labeled enzyme. This design can realize a linear range for glucose detection from 0.5 to 7 mM. Another example was a “signal-off” PEC based on WO3/Au composites.182 Initially, this sensor showed a higher photocurrent due to plasmon-enhanced photoconversion than WO3 alone. In the presence of Hg2+, Au–amalgam was formed by interaction of Au with Hg2+, which suppressed the hot electron transfer and PIRET, thus weakening the photocurrent of WO3/Au NPs. Here, AuNPs played a significant role in element recognition in the detection (Fig. 13(c)).

Applications of PEC sensors in paper-based microfluidics. Yu and co-workers have reported a PEC-based paper microfluidics for the detection of cancer biomarker CEA.183 In their design, the PEC sensor consisted of a sample zone, a PEC collection zone, and two rectangular legs coated with graphite films and gel electrolyte. Especially, the sample zone was assembled with positively charged carbon nanotubes and the negatively charged CdS NPs were further modified with capture antibodies through carbodiimide chemistry. The sample zone and the collection zone were screen-printed with a carbon electrode as working electrode and the counter electrode, respectively, after which they were folded to form a PEC reaction zone with buffer solution. The two rectangular legs were stacked to form a supercapacitor. After incubation with the antigen and capture antibody labeled with AuNPs, signal transduction and amplification were realized through internal chemiluminescence-triggered photocurrent from CdS and the photocurrent generated in the paper supercapacitor (Fig. 14).
image file: d3cs00793f-f14.tif
Fig. 14 Photoelectrochemical devices. (a) Photocurrent generation mechanism in paper-based PEC sensors; (b) PEC reaction disk and the collection pad; and (c) pictures of the assembled device. Reproduced with permission from ref. 183, copyright 2013, Royal Society of Chemistry.
Applications of PEC sensors in PDMS-based microfluidics. Thanks to the flexible structure, integration of multiple hardware components, and stretchable nature of PDMS-based microfluidics, PEC sensors can be made into wearable devices for health monitoring.177,184 This allows for user-friendly and comfortable health tracking in real-time. Liu et al. have demonstrated a wearable PEC sensor for the detection of glucose in sweat.184 Briefly, after fabricating three-electrode arrays on the PDMS substrate, the photoactive materials (BiVO4 loaded on multiwalled carbon nanotubes), the SiO2 NP layer, the redox center (glucose oxidase), and a Nafion layer were assembled on the working electrode to form the reaction zone, where glucose oxidase was used for glucose oxidation, and the Nafion layer was employed to protect glucose oxidase. A PDMS sweat collector aligned with a three-electrode array was also engineered by a laser engraving method. When the glucose present in sweat came in contact with glucose oxidase, the generated H2O2 was then oxidized by the photoactive BiVO4 layer under light illumination, generating photocurrent. This design showed a linear glucose response ranging from 0.1 nM to 1 mM at a LOD of 22.2 pM.

4. Bio-imaging using Au and Ag nanoparticles

Medical imaging techniques can help diagnose, localize, and monitor disorder. Common bio-imaging modalities include computed tomography (CT), positron emission tomography (PET), magnetic resonance imaging (MRI) and ultrasonography. However, these techniques cannot provide target-specific results, thus requiring professionals to distinguish disordered areas. And it is hard to achieve long-time and real-time monitoring. In contrast, metallic nanoparticle assisted bioimaging has a great potential to achieve specific, long-term, and real-time imaging.

When applying nanoparticles into bio-imaging applications, it is essential to perform thorough characterization and evaluation of the structure and properties of the nanoparticles for quality assurance.14 Some key factors, such as size, water-solubility and biocompatibility of nanoparticles, should be taken into account. Small nanoparticles are known for their enhanced permeability and retention (EPR) effects in tumors, leading to higher local concentrations of the imaging agent in these areas. However, if nanoparticles are smaller than 10 nm, they are quickly eliminated through the renal excretion system, as the average pore size in renal filtration is 10 nm. In contrast, nanoparticles larger than 100 nm are prone to detection by macrophages and tend to accumulate in organs associated with the mononuclear phagocyte system (MPS), such as the lymph nodes, liver, spleen, and lungs. The particle size distribution needs to be controlled because particle size significantly affects many aspects like biodistribution, retention time in blood circulation, cellular uptake, tumor penetration, and targeting.185 A second crucial aspect is the water-solubility and stability of nanoparticles in in vitro and in vivo experiments. Uncoated metallic nanoparticles are prone to aggregation in a high ionic strength environment, such as under in vivo conditions. Aggregation can alter their original characteristics, including their size, shape, plasmonic resonance, and signal strength. Furthermore, specific nanoparticle shapes, like gold nanorods, can undergo fusion or deformation under intense pulsed laser exposure.186a A water-soluble coating is essential for enhancing the stability of these metallic nanoparticles. An additional key factor is biocompatibility, especially when various nanoparticles are introduced into the human body. Prior to commercialization, it is imperative to assess and gain understanding of toxicity of nanoparticle-based bio-imaging agents.117 It is worth noting that oxidative stress has been implicated as one of underlying mechanisms for toxicity of nanoparticles. Cellular reactive oxygen species (ROS) generation is an indication of oxidative stress. Intracellular ROS generation can be triggered by cellular uptake of nanoparticles and intracellular metal ion release.186b Coating nanoparticle surfaces is an effective way to improve biocompatibility.186c

4.1. Dark-field imaging

Concept and operating principle. Dark field imaging is an optical microscopy technique that consists of a light source, a condenser, an objective lens, and a camera as illustrated in Fig. 15(a).187 Specifically, the specialized dark field condenser plays a pivotal role by obstructing the central portion of the vertically incident light, permitting only oblique or grazing-angle light to illuminate the specimen within the focal plane.7 As a result, light scattered by the specimen enters the objective lens, giving rise to a vivid image against a backdrop of darkness. For bio-imaging, achieving high resolution is imperative using a high numerical aperture. However, an excessively high numerical aperture permits incident light to pass through, diminishing the overall contrast within the image. Hence, employing a high numerical aperture is a trade-off between resolution and contrast. To mitigate this issue, total internal reflection microscopy has been introduced as a method for dark-field bio-imaging. This technique employs an evanescent wave that penetrates solely into the optically less dense medium up to a depth corresponding to the observed particle. The wave is not captured in the far-field, leading to a dark background. Hence, the implementation of total internal reflection microscopy significantly reduces incident and stray light, leading to a substantial reduction of the imaging background and enhancing the contrast in dark-field images. Building upon this technology, various other modes have emerged, encompassing nonlinear dark-field microscopy, and waveguide scattering microscopy.188–190 Besides generating an evanescent wave, controlling angular emission is another direction. For instance, Chazot et al. employed a flat Bragg mirror, utilizing its spectrally selective and angle-dependent transmission properties, to dictate the angular emission profile of the surface.191 By combining this Bragg mirror with quantum dots (QDs) and a micro-patterned metallic bottom surface, they successfully constructed an integrated luminescent photonic substrate. These innovative modes have expanded the capabilities of dark-field imaging, and offered research opportunities in bio-imaging. In short, dark-field imaging does not need any label, providing an advantage over fluorescence. However, it requires a close-up objective lens, which limits its application in in vivo imaging. Hence, it is typically used for in vitro imaging.
image file: d3cs00793f-f15.tif
Fig. 15 Optical microscope for dark-field imaging. (a) Dark-field imaging principle; (b) intracellular transport process and transport speed of DNA strand-tailored gold nanoparticles; (c) dynamic mechanical force transduction in live cells, which can be seen as alterations in plasmon intensity and peak position due to changes in the distance between two AuNPs; and (d) monitoring the dynamic intracellular ˙OH levels in single tumor cells. Reproduced with permission from (a) ref. 187, copyright 2022, MDPI, (b) ref. 192, copyright 2017, Springer Nature, (c) ref. 194, copyright 2017, American Chemical Society and (d) ref. 195, copyright 2020, American Chemical Society.

Plasmonic metal nanoparticles, particularly large Au and Ag nanoparticles, exhibit predominant light scattering over absorption. Their scattering cross-section is around 5 orders stronger than that of conventional organic fluorophores, rendering them highly amenable to capture and imaging using a dark-field microscope.188 The significant scattering cross-section, combined with their remarkable sensitivity to structural changes and the surrounding environment, enables the enhanced visualization and analysis of nanoscale phenomena and biological processes. This makes plasmonic metal nanoparticles valuable probes for dark-field imaging.

In vitro imaging. Prior to employing Ag and Au nanoparticles in bio-imaging and subsequent therapeutic applications, it is essential to understand the interactions between metal nanoparticles and cells. This includes the processes of nanoparticle uptake, intracellular transport, and distribution following internalization. Distinguished from transmission electron microscope (TEM), which is a static method for the visualization of NPs in cells, dark-field imaging presents a direct and dynamic means to monitor the fate of nanoparticles in cells with time, offering valuable insights into their behaviors and localization. Fan's work has shed light on the intracellular transport process and transport speed of DNA strand-tailored gold nanoparticles, as shown in Fig. 15(b).192 By exploiting the dependence of the LSPR band on nanoparticle aggregation, they were able to visually and, in real-time, observe the clustering state at various stages of endocytosis. To gain a comprehensive understanding of the interplay between cells and nanoparticles (NPs), Fajardo et al. employed microfluidics to enable rapid screening of circulating cells and tracking of internalized NPs within cells.193 This technique has the potential to explore multiple factors, such as metal nanoparticle shape, composition, surface chemistry, cell type, and culture medium, allowing for the systematic investigation of the intricate dynamics of cellular–NP interactions.

Specific chemical or physical-induced changes in the plasmon properties have been devised to visualize and unravel complex biological processes. As demonstrated by Xiong et al., a unique approach involved connecting two gold nanoparticles (AuNPs) using an elastic molecular linkage (PEG75) (Fig. 15(c)).194 This AuNP pair, acting as a plasmonic nano-spring, exhibited distance-dependent plasmon coupling, as evidenced by changes in intensity and spectral shifts, in response to varying forces applied to the AuNPs. One AuNP was anchored to the glass substrate, while the other was tethered to the HeLa cell surface through Arg–Gly–Asp (RGD)–integrin binding. When the cells were exposed to reactive oxygen species (ROS), the subsequent reorganization of the cellular actin cytoskeleton induced changes in the distance between two AuNPs, resulting in alterations in plasmon intensity and peak position. This dynamic mechanical force transduction capability enabled the sensing of forces ranging from sub-piconewtons to tens of piconewtons, showcasing its potential for investigating diverse biological functions and physiological processes associated with mechanical force signaling and regulation. In addition to changes in plasmon coupling, altering the shape, size or even composition of single metal nanoparticles provides an alternative way for visualizing specific biological process. For example, the oxidation of endogenous hydroxyl radicals (˙OH) on Ag nanoparticles was employed to monitor the dynamic intracellular ˙OH levels in single tumor cells (Fig. 15(d)).195 They synthesized AgAu nanocages with Ag confined within the frame of the nanocage, followed by the attachment of PEG/RGD to the nanocage surface. The steric hindrance effect of polyethylene glycol (PEG)/RGD on Ag significantly slowed down reactions with superoxide anions and hydrogen peroxide; but PEG/RGD exhibited a higher reaction rate with ˙OH as compared to other ROS, mitigating its steric hindrance effect and facilitating the further oxidation of Ag by ˙OH. This method allowed for the visualization of the ROS process and provided specific distinction of ˙OH from other radicals, showing a great promise in monitoring cellular homeostasis and injury research.

4.2. Fluorescence imaging

Concept and operating principle. Fluorescence microscopy is performed by employing collimated light from the light source through the objective to the specimen, and the emitted light is collected through the eyepiece for observations. In a fluorescence microscope, both excitation light and emission light exhibit a constrained spectral range; thus filters are indispensable components,117 which effectively segregate the excitation and emission light, and separate the emitted light from the excitation light, especially when both may pass through the same objective lens. By performing these functions, filters effectively mitigate the influence of background light and significantly enhance the efficiency of emission collection. Specifically, for in vitro fluorescence imaging, a sample is placed on the designated sample holder, and an optimal combination of light sources (such as mercury, LEDs, and metal halide lamps) is selected. Hence, filters are essential for fluorescence imaging.196 Furthermore, to satisfy different measurement conditions and resolution requirements, specific types of objectives, including air, water-immersed, and oil-immersed objectives, are employed. These different objective types accommodate diverse imaging scenarios, ensuring accurate and high-resolution imaging in various experimental setups. Due to the diffraction-limit of light, the resolution (R) of traditional optical microscope is around 200–250 nm, calculated as197
 
image file: d3cs00793f-t13.tif(17)
where λ is the wavelength of incident light, NA represents the numerical aperture of the objective lens. To achieve higher resolution, confocal laser scanning microscopy (CLSM) has been developed to improve xy plane resolution to180–250 nm.198 The confocal microscopes also offer distinctive features, including controllable depth scanning, serial optical sectioning, and spatial filtering, which can be used to construct 3D images. Moreover, other super-resolution techniques, including 4Pi and I5M microscopy, stimulated emission depletion (STED) microscopy, saturated structured illumination microscopy and tip-enhanced fluorescence (TEF) microscopy, have been introduced.199–201 These cutting-edge techniques leverage bio-imaging to an unprecedented level of resolution, and enhance the visualization of intricate cellular structures and dynamic processes.

Fluorescence microscopes can be adapted to accommodate longer optical path length required for in vivo imaging of living organisms. This can be achieved through the utilization of optical fibers, enabling minimally invasive insertion into tissues and flexible navigation within hollow tissue cavities. Additionally, given that near-infrared (NIR) light is commonly employed for in vivo imaging, the low quantum yield of NIR fluorophores and restricted light penetration necessitate high detector sensitivity. Thus most in vivo microscopes are equipped with a charge-coupled device (CCD) camera or a complementary metal oxide semiconductor (CMOS) sensor.196 However, this spatial resolution of in vivo microscopes is notably lower than that of in vitro microscopes due to the above restrictions. The maximum spatial resolution ranges from 50 μm to 35 μm based on the imaging mode for a Peltier-cooled 1024 × 1024 pixel CCD camera used for NIR-I imaging, while the deep cooling InGaAs camera used for NIR-II imaging has an even lower pixel resolution of 640 × 512 pixels.202 Like in vitro microscopes, in vivo imaging can also incorporate 3D imaging capabilities through tomographic imaging systems. Tomographic systems utilize computational algorithms to generate 3D images from a series of 2D fluorescence images, thus capturing the sample's topography. This process involves calibrating the correlation between the fluorophore concentration and the fluorescence intensity, enabling the reconstruction of detailed 3D representations of the specimen under investigation.

The spatial and time resolution stand as prominent advantages of fluorescence imaging, hinging on the brightness and stability of fluorescent probes.203 Notably, NIR organic dyes suffer from low quantum yields, significantly impairing the effectiveness of in vivo imaging. Plasmon-enhanced fluorescence, on the other hand, can enhance the fluorescence intensity through absorption and/or emission enhancement mechanisms. Furthermore, the encapsulation of numerous fluorophores within a single probe can markedly reduce the likelihood of encountering dim, blinking, or photobleaching phenomena.

In vitro imaging. Plasmon-enhanced fluorescence probes often incorporate recognition molecules to enable target-specific imaging. Ligand-assisted imaging offers the dual benefits of providing specific information on biological events and enabling the targeted accumulation of imaging probes at designated sites, thereby enabling heightened sensitivity.204 Additionally, nanoparticles can be taken up directly by cells for non-targeting imaging. The efficiency of endocytosis is influenced by factors such as the shape, size, hydrophobicity, and electrostatic properties of the nanoparticles. Gao et al. explored the use of different shapes of Au cores and found that gold nanostars exhibited the highest fluorescence enhancement among the various shapes tested.129 Besides regulating the core shape of the probes, additional enhancements for in vitro imaging can be achieved through substrate modification with plasmonic nanostructures. Theodorou et al. have demonstrated this concept using gold nanostars as self-assembled monolayers on glass slides, employing thiolate–gold coordinative bonding (Fig. 16(a)).205 Then HeLa cells were cultured on the plasmonic array surface, followed by incubation with dye AF680 bonded antibodies. The plasmonic substrate, coupled with the fluorophores on cells, enabled a remarkable 19-fold fluorescence enhancement for cellular NIR imaging. With the development of tip-enhanced fluorescence microscopy, Garcia-Parajo's group has created a gold nanoparticle tip, functioning as an optical antenna, to generate a localized and enhanced near-field.201 The use of the aperture fiber tip not only amplified the fluorescence signal but achieved an impressive spatial resolution of 26 nm.
image file: d3cs00793f-f16.tif
Fig. 16 Fluorescence imaging. (a) Tumor cells using a pair of assistant probes on the surface of silica-coated Au nanostars; (b) NIR imaging with 19-fold fluorescence enhancement due to coupling between the plasmonic substrate and the fluorophores on cells. Reproduced with permission from (a) ref. 205, copyright 2019, Royal Society of Chemistry, (b) ref. 129, copyright 2021, American Chemical Society.

In addition to the “always on” imaging modes discussed earlier, triggering modes offer an alternative signal transduction approach to detect changes in fluorescence intensity induced by specific biological parameters.206,207 This mode allows for higher sensitivity, accuracy, and reduced interference from background signals. The parameters under investigation encompass environmental factors, such as pH, viscosity, and temperature, as well as specific biomarkers like small molecules, proteins, and nucleic acids within cells.208,209 For instance, Gao et al. implemented a triggering mode by employing a pair of assistant probes on the surface of silica-coated Au nanostars (Fig. 16(b)).129 These assistant probes consisted of two single-stranded DNA (ssDNA) strands, each labeled with a fluorophore and a quencher, respectively. In the presence of the quencher, the probes were initially quenched. Upon uptake by tumor cells, miRNA-21 in the target cells hybridized with the fluorophore-labeled SSDNA, causing the replacement of the quencher-labeled strand. This activated fluorescence signals, distinguishing tumor cells from normal cells.

In vivo imaging. In vivo imaging, primarily based on the permeability of tumor vasculature or the enhanced permeability and retention (EPR) effect, finds significant application in tumor angiography and fluorescence-guided surgery.206 This imaging approach offers a relatively straightforward means of directly localizing tumor sites from surrounding normal tissues through real-time fluorescence imaging. Notably, in the case of brain tumors, which are often irregular and indistinct, accurate estimation of tumor margins is critical to avoid irreparable harm to patients. Leveraging the capabilities of fluorescence imaging, brain tumor imaging can be achieved non-invasively through the skull, allowing for cerebral perfusion assessment in acute stroke patients. A typical example of this application was the use of near-infrared (NIR-II) PbS/CdS quantum dots (QDs) for tumor imaging.210 These QDs exhibited long-term retention and photostability, providing ample brightness and time for imaging-guided surgery. Moreover, they were excreted out of the human body after a suitable interval, ensuring the safety of patients during imaging and surgical procedures.

To overcome the limited penetration capabilities of fluorescence imaging, researchers have turned to multimodal imaging, which combines NIR fluorescence imaging with other optical modalities with higher or even unlimited tissue-penetration depths. Some of these modalities include magnetic resonance imaging (MRI), computed tomography (CT), photoacoustic imaging (PA), and positron emission tomography (PET). Although these methods offer deeper imaging capabilities, they often suffer from lower sensitivity and lack real-time functionality. Thus, multimodal imaging approaches allow for the exploitation of each modality's strength and compensation of their respective drawbacks. For example, Li et al. have presented a fluorescent and CT bimodal method for CL1-5 tumor imaging.211 They conjugated gold nanoparticles with diatrizoic acid for CT imaging and with the nucleolin-targeted AS1411aptamer for fluorescence imaging. During measurement, CT imaging was used to locate the tumor before surgery; and fluorescence at the tumor site can be easily seen by the naked eye, facilitating the resection in surgery with improved accuracy. Furthermore, a “triggering mode” has emerged as an enhancement to the conventional “always on” multimodal imaging approach. Yan et al. have developed bimodal probes that activate NIR fluorescence and MRI, showing improved sensitivity.212 Fluorescence of these probes was triggered by an alkaline phosphatase (ALP) reaction. Upon delivery to ALP-positive tumor sites, the NIR dye's phosphate group underwent dephosphorylation, leading to re-emission of fluorescence and elevating sensitivity of imaging.

4.3. SERS imaging

Concept and operating principle. The basic structure of a Raman microscope resembles that of the traditional optical ones, but it employs lasers fitted with laser rejection filters as incident light. Commonly used lasers have wavelengths of 532 nm, 633 nm, 785 nm, and 1064 nm. As Raman signals carry spectral information, both CCD cameras and spectrometers (or monochromators) are typically integrated to capture spatial and spectral data, revealing the distribution of components and material properties, respectively.213 Similar to conventional optical microscopes, the spatial resolution of Raman microscopy is constrained by the diffraction of light, while sensitivity is impacted by background signals. Thus, confocal Raman microscopy, total internal reflection Raman spectroscopy and other super-resolution methods combined with Raman microscopes, such as tip-enhanced Raman spectroscopy (TERS), have been employed to improve its performance.214 TERS merges Raman microscopy with an atomic force microscope (AFM), utilizing an AFM tip coated with gold or silver to significantly enhance the Raman signal.215 This integration allows TERS to vastly improve signal sensitivity, achieve nanometer spatial resolution, and reveal surface topography.

Raman microscopy, utilizing 785 nm and 1064 nm lasers within the biological transparent window, is suitable for in vivo imaging. Optical fibers are subsequently employed in Raman endoscopy to overcome tissue penetration limitations. Raman endoscopy has demonstrated the capability to detect lesions ranging from 0.5 mm to 1.0 mm in size, a significant improvement compared to traditional white-light endoscopy, which can only visualize lesions ranging from 5–7 mm in size.216

In vitro imaging. In vitro SERS imaging is categorized into label-based and label-free methods. Label-based methods resemble fluorescence imaging, including “always on” direct modes and “on–off” triggering modes, which is exemplified by the measurement of intracellular pH in living cells. Cellular pH is a crucial indicator for proliferation, metastasis, drug resistance, apoptosis and especially for identification of cancer cells whose extracellular milieu is acid. Puppulin et al. utilized SERS to visualize the location of pH induced by proton extrusion in cancer cells Fig. 17(a).217 They first anchored biotin on the cell membrane and AuNPs. Then streptavidin was used to bind AuNPs and the cell surface. 4-MBA, serving as a Raman reporter, was sensitive to pH, which was preloaded on AuNPs to measure cellular pH changes. This method not only enabled sensitive pH measurements, but also realized relatively high resolution imaging for the visualization of pH gradience on a single cell.
image file: d3cs00793f-f17.tif
Fig. 17 SERS imaging. (a) In vitro imaging to visualize the location of pH induced by proton extrusion in cancer cells; (b) in vivo imaging of sub-millimeter microtumors with porous AuAg nanocages; and (c) imaging with multiplexed SERS nanoprobes. Reproduced with permission from (a) ref. 217, copyright 2018, Springer Nature, (b) ref. 10, copyright 2022, Springer Nature, (c) ref. 219, copyright 2023, Elsevier.

Label-based SERS imaging shows notable specificity, while label-free methods tend to visualize multiple targets in the sample matrix or discover unknown components when combined with statistical methods such as principal component analysis. Label-free SERS imaging relies on intrinsic molecular Raman signals present in cells and tissues. By incorporating plasmonic nanoparticles or substrates, the Raman signals can be greatly amplified, enabling SERS imaging. For example, He's group showcased SERS mapping of pesticide penetration in edible leaves.218 Balancing the use of pesticides to enhance productivity with the need to safeguard human health and the environment poses a dilemma. To address this, researchers employed infiltrated gold nanoparticles (AuNPs) as probes for in situ mapping of pesticide penetration. It was reported that 1 ppb of analytes can be detected using this method and the mapping depth can be up to 300 μm using a confocal Raman microscope.

In vivo imaging. Direct “always on” modes and “on–off” triggering modes have been developed for in vivo SERS imaging. Their applicability has been limited to the visible (400–700 nm) and NIR I regions (700–900 nm), leading to restrictions in tissue penetrations. Thus, the principal challenge in in vivo SERS imaging lies to achieve access to the NIR II region (1000–1400 nm) to enable deeper tissue penetration. Given that SERS imaging is primarily utilized for tumor identification, the plasmonic particle's size should be kept below 200 nm to ensure the EPR effect, posing further complexities in nanoparticle design and synthesis.10 Fortunately, porous AuAg nanocages showed a tunable spectral range from 787 nm to 1540 nm by increasing the pore size and number, with an overall size of around 56 nm.10 Li’ s group has further coated the AuAg nanocages with IR-1061 dye as the Raman reporter, encapsulated them with a polyethyleneimine (PEI) layer, and subsequently modified them with hyaluronic acid (HA) to target the CD44 receptor on the cell membrane of 4T1 mouse breast cancer cells (Fig. 17(b)).10 This reported method located the tumor margin in a living mouse model successfully, and the retention time of these NIR-II SERS probes was sufficient for intravenous administration. Typically, the identification of cancer biomarkers involves more than one marker, necessitating the use of multiple SERS probes for effective imaging. For instance, immune checkpoints, which are crucial inhibition pathways for maintaining self-tolerance and modulating the immune response in peripheral tissues, are increasingly being employed in cancer imaging and immunotherapy. Ye's group has developed a set of 32 resolvable Raman reporters whose spectrum features at various frequencies ranging from 1800–2800 cm−1.219 From this set, they selected 10 SERS probes to target 10 different immune checkpoint inhibitors in breast cancer biopsy samples. The findings from SERS imaging were found to align closely with those obtained through immunohistochemistry (Fig. 17(c)).

SERS imaging has also been combined with other optical techniques to realize multimodal imaging, thereby compensating for their respective drawbacks. For example, a dual SERS and photoacoustic (PA) imaging approach has been demonstrated using a single core–satellite configuration of gold nanoparticles to detect inflammation/cancer-related H2O2.220 A double-ratio response of H2O2 was utilized for quantitatively monitoring H2O2. In brief, a silica-coated gold nanorod embedded with reference Raman reporters, such as 2-naphthalenethiol (NAT). This nanorod was then labeled with gold nanoparticles modified with 4-mercaptobenzonitrile (reported Raman molecules, MBN) and D-(+)-galactose. Additionally, 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) serving as a PA reagent and horseradish peroxidase (HRP) were embedded in the silica layer. In the presence of H2O2, AuNPs were released and the MBN signal decreased, while the reference Raman signal of NAT within the nanorod structures remained unchanged. For PA imaging, HRP, in the presence of H2O2, catalyzed ABTS, leading to its oxidization and an increase in the PA signal at 750 nm, while the PA signal at 1250 nm from the nanorods did not change.

4.4. Photoacoustic imaging

Concept and operating principle. Different from the above light-to-light conversion imaging modalities, photoacoustic (PA) imaging is a novel technique that converts light into thermal energy (temperature rise), leading to the generation of sound waves due to thermoelastic expansion. In contrast to ultrasound imaging that provides anatomy information, PA can give functional information such as the total and distribution of biomarkers, and even molecular information with the assistance of exogenous contrast agents. While traditional optical imaging is hampered by limited depth penetration due to tissue scattering, PA imaging retains optical imaging's benefits like heightened contrast and multiplexing possibilities. Additionally, it offers spatial resolution superior to that of ultrasonic imaging, effective even at greater tissue depths.221 The lateral resolution and image depth values are 20–50 μm and ∼5 mm for acoustic-resolution photoacoustic microscopy (AR-PAM) and 100–400 μm and ∼8 cm for photoacoustic computed tomography (PACT), respectively.

The PA contrast agents include organic materials such as Cy5, ICG, polymers, inorganic nanoparticles such as gold NPs, silica NPs and copper sulfide NPs.222 Among them, plasmonic nanoparticles are notably effective as PA imaging agents due to their high optical absorption capacity and adjustable absorption peaks. It is worth noting that the thermalization of metallic nanoparticles is in the picosecond range as mentioned above, significantly shorter than a nanosecond laser pulse. Consequently, the PA signal predominantly originates from the heat transferred to the surrounding environment and not directly from the nanoparticle. In essence, photoacoustic imaging contrast hinges on the efficiency of optical-to-acoustic conversion, encompassing the photon-to-heat transfer, heat diffusion from plasmonic nanoparticles to their surroundings, and the thermal expansion capability of those surroundings. Combining the above parameters, the pressure difference generated by thermal expansion (ρ0) is then given as follows:223

 
image file: d3cs00793f-t14.tif(18)
 
image file: d3cs00793f-t15.tif(19)
 
A = uaF(20)
where β is the thermal expansion coefficient of the surrounding tissue; c is the speed of sound in the medium; Cp is the specific heat capacity of the tissue at constant pressure; μa represents the optical absorption coefficient of the photo-absorber; and F is the laser fluence. Γ is the Grüneisen parameter representing the thermoacoustic conversion efficiency, and A is the local energy deposition density.

The above equation demonstrates that the criteria for selecting plasmonic nanoparticles for photoacoustic imaging align with the principles of plasmon heating. The use of plasmonic nanoparticles can enhance the absorption cross-section, improving the efficiency of photon conversion to heat. Repenko and colleagues conducted a comparative analysis of the photoacoustic contrast in gold–melanin core–shell structures with various shapes, such as spheres, nanostars, nanorods, and nanoshells (Fig. 18(a)). Their findings revealed that gold nanorods coated with melanin exhibited the most significant photoacoustic enhancement, which was attributed to their superior absorption ratio.224 Besides, photoacoustic imaging involves thermal energy transfer from the nanoparticle to the surrounding environment. Emelianov et al. pointed out that a thin silica coating on the surface of gold nanorods can boost the amplitude of the generated photoacoustic signal due to the reduction of the gold interfacial thermal resistance with the solvent by the silica coating.186


image file: d3cs00793f-f18.tif
Fig. 18 Photoacoustic image. (a) Photoacoustic performance of nanoparticles with different morphologies and coatings; (b) schematic of the plasmon peak shift induced by gold nanoparticle fusion triggered by H2O2 (i), as well as the PA imaging to distinguish tumor tissue from normal tissue (ii); and (c) the schematic of the “on–off” mode of PA imaging (i) and in vivo photoacoustic oxidative stress sensing (ii). Reproduced with permission from (a) ref. 224, copyright 2017, John Wiley and Sons, (b) ref. 229, copyright 2021, John Wiley and Sons, (c) ref. 230, copyright 2020, Royal Society of Chemistry.
In vitro and in vivo imaging. The contrast and efficiency of PA are heavily dependent on colloidal size, shapes as well as the coating materials used to stabilize metallic nanoparticles, enhance photoacoustic intensity and introduce bioactive layer.225,226 Gold nanorods have high absorption that creates strong imaging contrast and their spectrum can be adjusted to the NIR window. However, their anisotropic structure is prone to deformation under NIR-I and NIR-II high-pulsed lasers used in photoacoustic imaging. Exposure to laser energy can transform these rods into spheres, leading to a weakened PA signal and a blue shift in absorption. To prevent this deformation under laser irradiation, Jokerst and his team coated the surface of gold nanorods with a protective polydopamine layer, enhancing their laser stability.227 The study revealed that gold nanorods with a coating could retain 87% of their initial PA signal in vivo after 10 minutes of illumination, in contrast to uncoated gold nanorods, which experienced a 75% decrease.

Besides preserving the shape of plasmonic nanoparticles for consistent “always on” imaging, enhancing PA imaging sensitivity through biochemical triggers, such as “activable” imaging, is also effective. An example of this is the spectrum red-shift and signal amplification triggered by aggregation, offering additional benefits for PA imaging. Emelianov and co-workers engineered plasmonic gold nanospheres with a matrix metalloproteinase-9 aptamer and complementary sequences.228 This modification caused the aggregation of gold nanospheres through DNA hybridization and displacement in the presence of the target protein. They have found that AuNSs were aggregated with a red shift of the plasmon peak to the NIR-I window only when incubated with the target protein. The PA signal at 700 nm was also 10 times higher than that of non-aggregated ones. This aggregation-induced PA imaging showed high sensitivity and selectivity, proving effective in both in vitro cell cultures and in vivo xenograft murine models of human breast cancer. Similarly, Zhou et al. developed an activable PA imaging technique for tumor diagnosis using the silica-encapsulated self-assembled gold nanochains (Fig. 18(b)).229 This approach hinged on the redox reaction between the over-expressed H2O2 in a tumor microenvironment and citrate ligands on particles, inducing particle fusion and resulting in a high PA signal and spectral red shift. With the high absorption of fused gold nanochains, they combined PA imaging with photothermal therapy at 1064 nm, achieving effective tumor diagnosis and significant tumor growth inhibition in vivo (Fig. 18(b)). In addition, reactive oxygen and nitrogen species (RONS) can activate PA imaging through metal etching. Mantri et al. have reported that the plasmon peak of gold nanorods exhibited a blue shift after coating with a silver shell, indicating an “off” mode of PA imaging.230 In the presence of RONS, silver ions were selectively etched out, restoring the NIR resonance of the AuNRs. By doping the silver shells with iodide, the reduction potential of silver was aligned with the oxidation potential of RONS, enhancing the detection limit by 3–5 logarithmic orders of magnitude as compared to the undoped counterpart and increasing imaging sensitivity (Fig. 18(c)).

5. Therapy using Au and Ag nanoparticles

Cancer is a significant health problem and one of the most common causes of death in the world. Traditional treatment modalities include tumor resection or resection combined with radiotherapy and/or chemotherapy depending on the types of tumors. Tumor resection works best for localized tumors, while radiotherapy and chemotherapy are suitable for killing the residual cancer cells and metastatic tumor cells to ensure complete removal of cancer. However, chemotherapy and radiotherapy have various side effects, such as killing normal cells due to their non-specificity, along with cancer cells becoming drug resistant. Photothermal therapy (PTT) and photodynamic therapy (PDT) revolutionize cancer treatment as a non-invasive treatment modality. They normally utilize NIR light to trigger therapy and immobilize biomarkers, drugs, and imaging agents into probes, realizing safe and specific cancer treatment. Besides oncology, PDT and PTT are finding increasing applications in other diseases, such as dermatology, cardiovascular and ophthalmology, inflammatory disorders, plastic and reconstructive surgery.

5.1. Photothermal therapy

Concept and principle. Utilization of heat for tumor therapy can be traced back to 1700 BC, during which a heated tip from a fire drill was used to treat breast cancer.231 Later, heating treatment was further extended to include radiofrequency, microwaves to ultrasound waves. The increased temperature in a target region using these techniques is named hyperthermia. Specifically, hyperthermia is defined as heating tissue to 41–47 °C for sustained periods. Hyperthermia damage to tumors is derived from the intrinsic vascular deficiencies of tumors, which render them more susceptible to this temperature range compared to normal tissues. However, it is difficult to avoid the surrounding healthy issues from heating damage using the hyperthermia method. Therefore, photothermal therapy (PTT) represents a heating treatment revolution aimed at killing malignant cells through a light-to-heat process to confine thermal damage to tumor tissue. PTT employs photothermal transaction agents (PTA) under external light irradiation. The observed irreversible cellular destruction in tumors stems from a combination of processes including protein denaturation and perturbation of cell membranes, leading to their structural disruption, and coagulation events. Additionally, the localized formation of vapor bubbles around certain strong PTAs (such as AuNPs) upon fluid evaporation introduces mechanical stress, further exacerbating the cellular damage.232 The effectiveness of PTT hinges predominantly on the attributes of the employed PTAs. These agents range from endogenous chromophores in tissues to exogenous organic dyes (i.e., indocyanine green, naphthalocyanines, and porphyrins), semiconductors and metallic nanostructures, where indocyanine green (ICG) is an FDA approved NIR photothermal agent.233 An optimal PTA should possess several key characteristics: (i) high light-harvesting ability and photothermal conversion efficiency (PCE); (ii) preferential accumulation within tumor sites and effective heat confinement to minimize thermal impact on adjacent normal tissues; and (iii) NIR range absorption to enable enhanced tissue penetration while minimizing interference with tumor background. As compared to traditional organic molecules such as ICG, plasmonic metallic PTAs have adjustable light absorption ranging from the visible-light to NIR-II region, and are much more stable during operation. Therefore, plasmonic PTAs, within the array of available PTA options, exhibit substantial promise in augmenting PTT's efficacy for cancer therapy.

Photothermal conversion within plasmonic nanoparticles arises due to electron–phonon interactions. As described in Section 3, incident light excites plasmons, which generate hot electrons, followed by electron–electron scattering and electron–phonon interactions, leading to the thermalization of energy within metallic nanoparticles. Hence, thermal energy is dissipated from the nanoparticles into the adjacent medium, thereby increasing the temperature of the surroundings. Due to the comparatively modest heat capacity of conduction electrons, electron temperature of 1000–5000 K in plasmonic nanoparticles can be easily achieved through utilization of pulse energies as low as 100 nJ.234 This capability can notably increase the thermal contrast between the target tumor region and background, a vital aspect ensuring the efficacy of PTT while simultaneously safeguarding the integrity of normal tissues. Furthermore, the temperature distribution created by a photothermally converted particle diminishes proportionally to the reciprocal of distance (1/d) from the particle surface, namely, the environmental heating is confined to a nanometer scale region around the particle, which mitigates damage to healthy tissues.39 Moreover, plasmonic PTAs exhibit distinct significantly amplified absorption cross-sections (often spanning four to five orders of magnitude higher than their counterparts), inherent photostability, pronounced tumor site accumulation through the enhanced permeability and retention (EPR) effect, and facile adaptability for targeted, activated, and even multimodal therapeutic strategies, underscore their potential in advancing the field of photothermal treatment.

Therapeutic agent structure and features. Plasmonic nanoparticles should fulfill two optical prerequisites: (i) modulation of plasmon resonance within the NIR region to facilitate proficient penetration through deep-seated tissues; and (ii) manipulation of the scattering-to-absorption ratio (Csca/Cabs) to minimize the scattering portion, thereby maximizing photothermal conversion efficiency. Among a variety of Au or composite nanoparticles, Au@SiO2 core–shell structures, Au nanostars, AuAg nanocages, and Au nanorods can be easily tuned to the NIR range, across in vitro and in vivo models.231,235 Experimental and theoretical approaches have been adopted to discern single particle photothermal conversion efficiency and bulk photothermal conversion efficiency. Halas and co-workers documented that Au@SiO2, nanorods and Au/Au2S nanoshells had superior conversion efficiency.236 However, achieving a given heating requirement necessitated a higher dosage of smaller-sized nanorods and Au/Au2S nanoshells compared to Au@SiO2. Additionally, Bucharskaya et al. have further compared conversion efficiency among Au@SiO2, AuAg nanocages and Au nanorods using an optical density at 808 nm.237 It was found that the three nanostructures displayed similar heating kinetics, but AuAg nanocages had the highest efficiency per unit mass of gold, followed by Au nanorods, with Au@SiO2 being the least efficient.

Bare metallic nanoparticles typically undergo surface modification to improve biocompatibility, biostability and potency for tumor-targeting under in vivo conditions. The common passivation strategies include the application of protective layers, such as SiO2 coatings, bovine serum albumin (BSA), melanin, and polymer layers (i.e., poly(ethylene)glycol (PEG) and polydopamine (PDA)).2,224,227,238 These biomolecules prevent nanoparticles from aggregating in physiological environments and increase the hydrophilicity of nanoparticles. These kinds of PTAs can be directly injected into tumor sites for PTT therapy or can accumulate at the tumor site through the EPR effect. To achieve concentrated accumulation of PTAs at the tumor sites, modification of these nanoparticles with target ligands has been exploited. Ligands such as antibodies, peptides, hormones, and glycopolymers that target the tumor cell surface are widely utilized for the labeling of metallic nanoparticles.239 In addition, certain tumor cells which elude detection have been incorporated onto the particle surface for “on-demand” therapy by utilizing conventional biomarkers or environmental sensitive molecules such as amino acids which are responsive to pH and N-isopropylacrylamide which are responsive to temperature.240 In addition to these approaches, biomimetic-engineered metallic nanoparticles leverage the intrinsic targeting sensitivity of specific cell types towards tumor cells.

In vivo therapy. As a safe and effective therapeutic approach, PTT has been conducted in clinical trials. One typical clinical pilot device study was executed by Halas et al., focusing on the direct focal ablation of prostate tumors using Au@SiO2 nanoshell structures (Fig. 19(a)).241 These nanoshells were introduced intravenously, selectively accumulating within solid tumor tissues while maintaining a lower nanoparticle burden in healthy tissues. Upon illumination with NIR laser irradiation at a sub-ablation power, cancerous lesions featuring substantial Au@SiO2 loading underwent coagulative necrosis induced by the resulting photothermal energy, while preserving the integrity of the neighboring healthy tissue milieu. After laser ablation for 3 and 12 months, the post-biopsies through MRI/ultrasound confirmed a success ratio of 95% for Au@SiO2 mediated focal laser ablation.
image file: d3cs00793f-f19.tif
Fig. 19 Photothermal therapy. (a) In vivo therapy with direct focal ablation of prostate tumors using Au@SiO2; (b) photothermally controlled drug release. Reproduced with permission from (a) ref. 241, copyright 2019, National Academy of Sciences, (b) ref. 244, copyright 2016, John Wiley and Sons.

However, the delivery of PTAs to the intended tumor sites remains a challenge in photothermal therapy. The efficiency of passive delivery is merely 0.6% of the injected dose.242 Even when employing target ligand-mediated delivery mechanisms, which aim to facilitate the accumulation of PTAs, the efficiency is only moderately elevated to 0.9% of the injected dose. It is partially attributed to the intrinsic hypoxic condition inside solid tumors. Thus, less distribution of blood vessels in tumors impedes the accessibility of PTAs to the tumor milieu. A recent strategy involving the encapsulation of metallic nanoparticles within specific cell types has exhibited promise in augmenting delivery efficiency. For example, macrophages can serve as a cellular carrier by phagocytosing PTAs. Also, macrophages can be recruited by tumors to achieve the homing of PTAs to the tumor sites, realizing deep tumor penetration.243 This immune cell-based PTT can significantly increase the targeting efficiency and improve transport to the desired locations.243

Besides the direct photothermal effect, PTT-generated heat can synergistically enhance other therapeutic modalities, including alleviating hypoxia for photodynamic therapy and stimulating drug release for chemotherapy. An illustrative example involved the creation of a photothermally controlled drug release system (Fig. 19(b)).244 They devised a system that embedded poly(methacryloxyethyl trimethyl ammonium chloride) (P(METAC)) functionalized AuNPs and bevacizumab within an agarose hydrogel matrix. The temperature-dependent gelation behavior of agarose facilitated reversible softening upon a temperature increase induced by the photothermal effect of AuNPs. Consequently, the released bevacizumab within the hydrogel contributed to the therapeutic regimen. The dynamics of drug release were modulated by various factors such as AuNP concentration, agarose content, light intensity, and exposure duration.

5.2. Photodynamic therapy

Concept and principle. Photodynamic therapy (PDT) is another light-triggered non-invasive therapy. It utilizes non-toxic photosensitizers (PS) that can turn oxygen molecules into reactive oxygen species (ROS) under incident light irradiation. The generated ROS cause damage to cellular constituents, including mitochondria, endoplasmic, peroxisomes, reticulum, and nucleus, leading to cell apoptosis. As illustrated in Fig. 20(a), a PS is initially excited to the singlet state (S1). S1 may initiate photochemistry or undergo intersystem crossing to form the excited triplet state (T1) through nonradiative transition.233 It can also relax back to its ground state (S0) through non-radiative decay to generate heat or emit fluorescence through radiative decay. These pathways are dependent on the types of PS and wavelength of incident light used. The generated T1 can also participate in a photochemical reaction, emit phosphorescence through radiative decay or relax back to S0via non-radiative decay. Generally, T1 has a longer lifetime than S1; thus photochemistry is often favored by T1. The biological event-related photochemistry or production of ROS involves two types of reaction processes, type I and type II, respectively. In the type I process, photoinduced electron in the excited triplet state is transferred to the biomolecules or oxygen molecules to form ROS radicals (such as O2˙), which further undergo a redox reaction and produce other intermediate ROS including hydrogen peroxide (H2O2) and hydroxyl radicals (OH˙). In contrast, type II is an energy transfer process between T1 and triplet oxygen (3O2), where T1 can convert triplet oxygen (3O2) into singlet oxygen (1O2). Both type I and type II processes can occur simultaneously in a competitive manner. It is generally assumed that type II is the most important process, conditioning the efficiency of PDT, while type I begins to prevail when oxygen runs out. The ratio of two contributions is determined by the oxygen concentration, PS types, the tissue dielectric constant, and pH.
image file: d3cs00793f-f20.tif
Fig. 20 Photodynamic therapy. (a) PDT mechanism and associated band diagram for designing PDT reactions; (b) PIRET-enhanced PDT with metal/semiconductor composites; and (c) imaging-guided synergistic PDT/PTT cancer treatment. Reproduced with permission from (a) ref. 233, copyright 2021, American Chemical Society and ref. 249, copyright 2012, American Chemical Society, (b) ref. 250, copyright 2018, American Chemical Society, (c) ref. 253, copyright 2013, American Chemical Society.

The performance of PDT is mostly dependent on the efficiency of the photosensitizer. Commonly used PSs include natural compounds (e.g., hypericin, hypocrellin, riboflavin, and curcumin), tetrapyrrole structures (e.g., porphyrins, chlorins, bacteriochlorins, and phthalocyanines), boron-dipyrromethene (BODIPY), squaraine, transition metal complexes, and inorganic materials (e.g., TiO2, Cu2O and quantum dots).233,245 Tetrapyrrole structures fall into the largest groups used in PDT. Currently, indocyanine green (ICG) and porphyrins are FDA-approved PDT agents.233 However, like other photosensitizers used in catalysis, the NIR PS agents exhibit a very low quantum yield, which is defined as the ratio between the number of generated 1O2 and the number of absorbed photons. In addition, organic PSs suffer from photo-bleaching and photo-induced instability during operation, which makes it impossible for a single dose of an organic PS to perform multiple rounds of PDT per day for a long duration.246 To overcome these shortcomings, plasmonic metals have been integrated with semiconductors to form plasmonic PDT agents.247 Inorganic plasmonic PDT agents are highly stable during long-term operation, exhibit the extended light absorption region to NIR-II region, and can achieve high quantum yield and light-to-chemical energy conversion efficiency.248 In short, plasmon-enhanced PDT has the potential to improve the performance of PDT. On the other hand, gold and or silver materials themselves can act as photosensitizers due to their high-energy electrons; however their short lifetime of excited intraband transitions hinders the production of singlet oxygen. Therefore, metal–semiconductor heterojunctions are more common as plasmonic PDT agents and photocatalysts.

Therapeutic agent structure and features. Essentially, PDT is a photocatalysis process and can be understood as the application of photocatalysis in the medical field. One can borrow the concepts from plasmonic photocatalysis to design plasmonic-enhanced PDT agents. It is necessary to consider the band diagram when designing PDT reactions. Taking semiconducting PSs as an example as shown in Fig. 20(a), the 1O/O2 redox potential is 1.88 eV and it can be realized with TiO2, SiO2, Al2O3, Fe2O3, and ZnO nanoparticle suspensions, while H2O/OH˙ is 2.2 eV. TiO2 and CeO2 were durable for reaction of O2/O2˙ (−0.2 eV), while O2˙ can be generated in ZnO2 and Fe2O3 due to their n-type band up-bending.249 As discussed in the PEC section, plasmon energy transfer mechanisms include light trapping, hot electron injection and PIRET; thus plasmonic PDT agents should be designed based on these mechanisms to maximize the energy conversion efficiency, that is, “materials-by-design”.

For example, the PIRET mechanism was used to design the plasmonic metal/semiconductor Au@SiO2@Cu2O materials as a plasmonic PDT agent (Fig. 20(b))250 The conduction band of Cu2O is 1.92 eV, higher than the redox potential of 1O2/O2 (1.88 eV), and Cu2O was reported to absorb O2 to consume the photogenerated electrons during water photocatalysis, indicating the application potential of Cu2O in PDT. To enable the PIRET mechanism, which theoretically can achieve the highest energy transfer efficiency, a large spectral overlap and a short separation distance should be achieved based on the PIRET theory. It was found that Au@SiO2@Cu2O exhibited the largest overlap integral compared to Au@Cu2O and extended light absorption to the longer red-to-NIR region (670 nm). Also, Au@SiO2@Cu2O exhibited higher photocatalytic activity than Au@Cu2O and Cu2O alone. Considering a hypoxia microenvironment in solid tumors, perfluorohexane (PFH) with high oxygen capacity was used to maintain a high oxygen content at a given oxygen partial pressure. Thus, the overall PS structure consisted of Au@SiO2@Cu2O coated with PFH and liposome nanocomposites. When excited by incident light (670 nm), energy of incident photons was converted into LSPR by the Au core and then the generated plasmonic energy was transferred to Cu2O via the PIRET process, leading to the generation of electron–hole pairs in Cu2O. The electrons generated in Cu2O then reacted with O2 provided by PFH to produce 1O2 for PDT. Besides PIRET, the hot electron injection mechanism has also been used to design plasmonic PDT agents. The TiO2 PS was only photoactive in the ultraviolet region due to its wide band gap (around 3.2 eV). Hence, gold nanoparticles were incorporated with TiO2 to form AuNPs–TiO2 heterojunctions to extend the light absorption region to the visible-light region. In the AuNPs–TiO2 heterojunction system, hot electron injection is the only mechanism that can transfer plasmonic energy from Au to TiO2. Fang et al. illustrated a gold nanorod@TiO2 core–shell structure as a PDT agent. This core–shell structure also formed a Schottky barrier between metal and semiconductors.251 Plasmon-generated hot electrons with higher energy than the Schottky barrier were injected into the conduction band of TiO2, leading to ROS generation. Light absorption of this plasmon-enhanced reaction could be extended to beyond 800 nm, facilitating the deep tissue penetration for PDT applications.

In vivo therapy. As discussed above, Au@SiO2@Cu2O designed through the PIRET mechanism had an average diameter of 193 nm after coating with PFH and liposome nanodrops, which met the requirements of the EPR effect for prolonged blood circulation and accumulation at tumor sites.250 After injecting such reagents into MCF-7 tumor-bearing mice, nanoparticles could be passively homed to the tumor through the EPR effect. Under 670 nm laser irradiation, these nanoparticles showed the most effective tumor growth suppression compared to other groups, creating a new platform for cancer therapy using PIRET-enabled PDT.

PDT and PTT can be coordinated to obtain the therapeutic index for synergistic effects. Wang et al. have synthesized gold nanoparticles that were covalently bound to the PEG-labeled Chlorin e6 (a photosensitizer), GNS-PEG-Ce6.252 Although PDT efficiency of Ce6 decreased over time with an increase in the PTT effect upon irradiation, PDT worked well in the early phase while PTT became dominant in the late phase. The process of PDT and PTT could be modulated by adjusting the irradiation time based on the difference in photostability between the photosensitizer and gold nanoparticles. The coordinated PDT/PTT treatment was superior to either PDT or PTT alone, improving its anticancer effectiveness and simplifying the treatment process. Given Ce6 can also emit fluorescence and release heat when excited by light, the participation of Ce6 in the system could further achieve an imaging-guided PDT process via fluorescence and photoacoustic imaging. Lin et al. have produced gold vesicles consisting of a monolayer of assembled AuNPs and encapsulated Ce6 in the nanoparticles, thus realizing fluorescence, thermal, and photoacoustic (PA) trimodal imaging guided synergistic PTT/PDT for cancer treatment (Fig. 20(c)).253 While fluorescence emission, heat release and photocatalysis of Ce6 were competitive processes, all of them could be influenced by plasmonic materials; thus the plasmon-enhanced mechanism was complicated and needed further clarification. Given that PDT and PTT are competitive processes based on their working principle, further research should be conducted to understand the underlying mechanisms and to determine the optimal conditions for the synergistic effect.

6. Remarks and perspective

Owing to tremendous efforts in the past decades, the synthetic protocols for Au and Ag nanoparticles have been well established, which have enabled systematic tailoring of particle composition, shape and size with good reproducibility, high purity, and high yield. The combination of experimental and theoretical work sheds light on plasmonic properties, including light absorption, scattering, EM field enhancement, hot electron emission and heating processes. The correlations between nanoparticles and their plasmonic properties have also been well-developed, laying a solid foundation for future applications of AuNPs and AgNPs.

Based on their plasmonic properties, AuNPs and AgNPs have been harnessed for the development of optical probes, including colorimetric, SERS and plasmonic fluorescence labels/reporters. These optical probes are finding increasing applications in point-of-care testing (POCT). Notably, colorimetric probes have emerged as highly successful tools in diverse applications, particularly in commercial paper lateral flow assays, which are exemplified by pregnancy and COVID-19 test strips. The success is underpinned by large-scale manufacturing, stringent quality control, and the ability to facilitate long-term storage and transportation at ambient temperatures. However, colorimetric probes are limited by their reduced sensitivity and susceptibility to interference from blood matrices; thus it is necessary to develop new optical probes to expand the horizons of sensors. In contrast, fluorescence and SERS probes have the potential to achieve high sensitivity and to avoid the interference of blood matrices. Applications of fluorescence probes have burgeoned in both research and commercialization fields, capitalizing on the untapped levels of sensitivity beyond those achievable with colorimetric counterparts. Nevertheless, the utilization of conventional fluorescence probes is hampered by their inherently low quantum yields, particularly in the NIR range, which continues to pose challenges in various applications. Hence, it is necessary to develop plasmon-enhanced fluorescence for signal enhancement. Presently, existing methodologies have yielded enhancement factors that, while valuable, remain relatively modest, typically on the order of a few-fold. Therefore, there is a need to further improve the quantum yield of plasmonic fluorescence probes, especially in the NIR spectral range, through a “materials-by-design” approach based on a complete understanding of the energy transfer mechanisms between plasmon and fluorophores.

Compared to plasmonic fluorescence probes, SERS probes have been well-developed with a high enhancement factor of 105 to 108. Integrating SERS probes into detection platforms has enabled the development of ultrasensitive sensors and POCT tools. SERS sensors have a high signal-to-noise ratio and strong resistance to interference of sample matrices when applying to blood samples. However, SERS-based sensors and POCT tools have not yet seen extensive adoption in homes, communities, clinics, and hospitals due to several limiting factors. First, SERS-based sensors and POCT tools have received limited awareness and recognition at clinics and hospitals although fluorescence and chemiluminescence instruments are well accepted by professionals and laypersons. Second, the high price of handheld Raman readers poses a significant impediment. Currently, portable Raman readers can command prices typically over $20[thin space (1/6-em)]000, substantially exceeding the price of colorimetric and fluorescence readers that typically range from $35 to $2000. Therefore, to promote commercialization and wide application of SERS-based sensors and POCT tools, it is essential to make low-cost Raman readers available in the market, establish standard detection method, and compare performance against colorimetric and fluorescence counterparts.

So far, optical probes have achieved great success in in vitro diagnostics (IVD). In contrast, the applications of optical probes in in vitro bio-imaging remain unexploited. Given the tissue penetration limitations, it is imperative to tune fluorescence probes toward three biological transparency windows in the NIR range.117 Conventional fluorophores typically exhibit quantum yields in the vicinity of 10% in the NIR-I spectral region, and this yield further diminishes to a range of 0.01% to 1.4% in the NIR-II region. Consequently, there is a pressing need to engineer plasmon-enhanced NIR fluorescence approaches to amplify signal intensity and quantum efficiency. Another challenge in bioimaging pertains to the site-specific delivery of optical probes. For example, in tumor monitoring, intravenous injection is a typical practice at this stage, and nanoparticles larger than 8 nm (between 8–100 nm) could passively locate to the target sites via the enhanced permeability and retention (EPR) effect. But the delivery efficiency, precision, and specificity are still uncertain and lack control. In addition, attention should be paid to biocompatibility and biodegradability of plasmon-enhanced fluorescence probes. It is highly recommended to comprehensively explore the physiological or pathological processes associated with established optical probes in cells, tissues, and organs. Such endeavors will be instrumental in shaping guidelines for material selection and establishing evaluation standards for both materials and methodologies. Also, optical probes can be further engineered with ligand modification and biologically mimetic coatings to achieve more precise delivery and increase the accuracy of optical diagnostics. SERS imaging shares similar challenges as the abovediscussed fluorescence counterparts. Additionally, operation is more complicated for SERS imaging. Because the intrinsic signal from SERS is scattering light, an optical fiber needs be used to collect this scattered light. Hence, specific microscopes or optical units need to be well fabricated for light collection and observation.

To date, FDA has approved several photothermal therapy (PTT) agents such as indocyanine green (ICG). Clinical applications of PTT have been successfully employed for the treatment of various diseases, including head and neck cancer, prostate cancer, and diabetic macular abnormalities. In addition, one of the plasmon probes (gold@silica core@shell particles) is currently undergoing clinical pilot trials. In parallel with the challenges encountered in bioimaging probes, site-specific delivery, biocompatibility, metabolism of nanoparticles are major challenges for the clinical implementation of PTT agents. Additionally, the thermal effects generated by different plasmonic structures exhibit variation, underscoring the necessity to quantitatively measure temperature at both target sites and in adjacent normal tissues, as well as establish evaluation standards to prevent hyperthermia in the surrounding normal environment. To achieve deeper tissue penetration, plasmonic PTT probes should be engineered into the NIR-II window. Concurrently, biological modifications should be tailored to facilitate precise delivery, addressing the challenges posed by tissue depth and specificity.

Several photodynamic therapy (PDT) agents have been approved by the FDA to treat certain cancers or precancers. For example, photofrin is approved to treat patients with esophagus and lung cancers, while aminolaevulinic acid serves as a drug for skin cancer treatment. However, achieving effective PDT in deep tissues remains a challenge, primarily attributed to the inherently low conversion efficiency of PDT agents, particularly within the NIR-II range. Plasmon enhanced PDT exhibits potential for increasing photoconversion efficiency in the NIR range. Currently, the prevailing PDT agents are organic photosensitizers, which are not stable and can be degraded in vivo with minimal health risks. In turn, their operation instability weakens the therapy performance. Multiple injections may be needed for a long-duration treatment process. In contrast, inorganic photosensitizers are much more stable than organic counterparts, while their biodegradability within the body remains a subject of uncertainty. Another concern is the hypoxic microenvironment in solid tumors, which limits the generation of reactive oxygen species (ROS) due to insufficient oxygen supply. To address this issue, several strategies have been proposed such as the incorporation of oxygen in the probes, in situ generation of oxygen by catalyzing H2O2 inside the tumor, transportation of oxygen from other sites to the tumor.

Taking advantages of plasmonic materials in imaging and therapy, it is feasible to combine PTT and PDT therapy with imaging and drug delivery, resulting in multi-modal therapeutic agents. Since multiply pathways may be involved in plasmon enhancement processes, the underlying mechanisms should be comprehensively discovered to enable better manipulation of each process and maximize each effect.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was partially supported by the National Institute of Biomedical Imaging And Bioengineering  (3U54EB007958-15S1) and the National Institute of Allergy and Infectious Diseases of the National Institutes of Health (2R01AI114495-06). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. The resources provided by the Armstrong-Siadat Endowed Chair Professorship Funds are greatly appreciated.

References

  1. M. Li, S. K. Cushing and N. Wu, Analyst, 2015, 140, 386–406 RSC .
  2. X. Cui, Q. Ruan, X. Zhuo, X. Xia, J. Hu, R. Fu, Y. Li, J. Wang and H. Xu, Chem. Rev., 2023, 123, 6891–6952 CrossRef CAS PubMed .
  3. S. K. Cushing and N. Wu, J. Phys. Chem. Lett., 2016, 7, 666–675 CrossRef CAS PubMed .
  4. N. Wu, Nanoscale, 2018, 10, 2679–2696 RSC .
  5. D. K. Roper, W. Ahn and M. Hoepfner, J. Phys. Chem. C, 2007, 111, 3636–3641 CrossRef CAS PubMed .
  6. H. H. Richardson, M. T. Carlson, P. J. Tandler, P. Hernandez and A. O. Govorov, Nano Lett., 2009, 9, 1139–1146 CrossRef CAS PubMed .
  7. P. F. Gao, G. Lei and C. Z. Huang, Anal. Chem., 2021, 93, 4707–4726 CrossRef CAS PubMed .
  8. D. M. Cate, J. A. Adkins, J. Mettakoonpitak and C. S. Henry, Anal. Chem., 2015, 87, 19–41 CrossRef CAS PubMed .
  9. M. Li, S. K. Cushing, J. Zhang, J. Lankford, Z. P. Aguilar, D. Ma and N. Wu, Nanotechnology, 2012, 23, 115501 CrossRef PubMed .
  10. L. Li, R. Jiang, B. Shan, Y. Lu, C. Zheng and M. Li, Nat. Commun., 2022, 13, 5249 CrossRef CAS PubMed .
  11. H. Tang, C. J. Chen, Z. Huang, J. Bright, G. Meng, R. S. Liu and N. Wu, J. Chem. Phys., 2020, 152, 220901 CrossRef CAS PubMed .
  12. Y. Liu, P. Bhattarai, Z. Dai and X. Chen, Chem. Soc. Rev., 2019, 48, 2053–2108 RSC .
  13. M.-C. Daniel and D. Astruc, Chem. Rev., 2004, 104, 293–346 CrossRef CAS PubMed .
  14. K. Saha, S. S. Agasti, C. Kim, X. Li and V. M. Rotello, Chem. Rev., 2012, 112, 2739–2779 CrossRef CAS PubMed .
  15. S. Iravani, H. Korbekandi, S. V. Mirmohammadi and B. Zolfaghari, Res. Pharm. Sci., 2014, 9, 385–406 CAS .
  16. E. C. Dreaden, A. M. Alkilany, X. Huang, C. J. Murphy and M. A. El-Sayed, Chem. Soc. Rev., 2012, 41, 2740–2779 RSC .
  17. C. F. Bohren, Am. J. Phys., 1983, 51, 323–327 CrossRef CAS .
  18. C. Voisin, N. Del Fatti, D. Christofilos and F. Vallée, J. Phys. Chem. B, 2001, 105, 2264–2280 CrossRef CAS .
  19. M. Lisowski, P. A. Loukakos, U. Bovensiepen, J. Stähler, C. Gahl and M. Wolf, Appl. Phys. A, 2004, 78, 165–176 CrossRef CAS .
  20. H. Inouye, K. Tanaka, I. Tanahashi and K. Hirao, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 57, 11334–11340 CrossRef CAS .
  21. S. Besner, A. V. Kabashin, F. M. Winnik and M. Meunier, J. Phys. Chem. C, 2009, 113, 9526–9531 CrossRef CAS .
  22. C.-J. Huang, Y.-H. Wang, P.-H. Chiu, M.-C. Shih and T.-H. Meen, Mater. Lett., 2006, 60, 1896–1900 CrossRef CAS .
  23. J. Neddersen, G. Chumanov and T. M. Cotton, Appl. Spectrosc., 1993, 47, 1959–1964 CrossRef CAS .
  24. M. H. Magnusson, K. Deppert, J.-O. Malm, J.-O. Bovin and L. Samuelson, J. Nanopart. Res, 1999, 1, 243–251 CrossRef CAS .
  25. P. Zhao, N. Li and D. Astruc, Coord. Chem. Rev., 2013, 257, 638–665 CrossRef CAS .
  26. Y. Li, O. Zaluzhna, B. Xu, Y. Gao, J. M. Modest and Y. J. Tong, J. Am. Chem. Soc., 2011, 133, 2092–2095 CrossRef CAS PubMed .
  27. C. Ziegler and A. Eychmüller, J. Phys. Chem. C, 2011, 115, 4502–4506 CrossRef CAS .
  28. P. C. Lee and D. Meisel, J. Phys. Chem., 1982, 86, 3391–3395 CrossRef CAS .
  29. K. TekaiaáElhsissen, J. Mater. Chem., 1996, 6, 573–577 RSC .
  30. P.-Y. Silvert, R. Herrera-Urbina and K. Tekaia-Elhsissen, J. Mater. Chem., 1997, 7, 293–299 RSC .
  31. K. M. Mayer and J. H. Hafner, Chem. Rev., 2011, 111, 3828–3857 CrossRef CAS PubMed .
  32. P. L. Stiles, J. A. Dieringer, N. C. Shah and R. P. Van Duyne, Annu. Rev. Anal. Chem., 2008, 1, 601–626 CrossRef CAS PubMed .
  33. L. D. Landau, J. S. Bell, M. Kearsley, L. Pitaevskii, E. Lifshitz and J. Sykes, Electrodynamics of continuous media, Elsevier, 2013 Search PubMed .
  34. S. K. Cushing and N. Q. Wu, Electrochem. Soc. Interface, 2013, 22, 63–67 CrossRef CAS .
  35. S. A. Maier and H. A. Atwater, J. Appl. Phys., 2005, 98, 011101 CrossRef .
  36. A. O. Govorov, H. Zhang and Y. K. Gun'ko, J. Phys. Chem. C, 2013, 117, 16616–16631 CrossRef CAS .
  37. R. Marin, A. Skripka, L. V. Besteiro, A. Benayas, Z. Wang, A. O. Govorov, P. Canton and F. Vetrone, Small, 2018, 14, 1803282 CrossRef PubMed .
  38. P. K. Jain, K. S. Lee, I. H. El-Sayed and M. A. El-Sayed, J. Phys. Chem. B, 2006, 110, 7238–7248 CrossRef CAS PubMed .
  39. A. O. Govorov and H. H. Richardson, Nano Today, 2007, 2, 30–38 CrossRef .
  40. S. E. J. Bell and A. Stewart, Surface Enhanced Raman Spectroscopy, 2010, pp. 71–86 Search PubMed .
  41. P. G. Etchegoin and E. C. Le Ru, Surface Enhanced Raman Spectroscopy: analytical, biophysical and life science applications, 2010, pp. 1–37 Search PubMed .
  42. Y. Gutiérrez, A. S. Brown, F. Moreno and M. Losurdo, J. Appl. Phys., 2020, 128, 080901 CrossRef .
  43. M. G. Blaber, M. D. Arnold and M. J. Ford, J. Phys. Condens. Matter., 2010, 22, 143201 CrossRef CAS PubMed .
  44. K. Esumi, K. Matsuhisa and K. Torigoe, Langmuir, 1995, 11, 3285–3287 CrossRef CAS .
  45. J. Pérez-Juste, I. Pastoriza-Santos, L. M. Liz-Marzán and P. Mulvaney, Coord. Chem. Rev., 2005, 249, 1870–1901 CrossRef .
  46. N. R. Jana, L. Gearheart and C. J. Murphy, J. Chem. Mater., 2001, 13, 2313–2322 CrossRef CAS .
  47. C. J. Murphy, T. K. Sau, A. M. Gole, C. J. Orendorff, J. Gao, L. Gou, S. E. Hunyadi and T. Li, J. Phys. Chem. B, 2005, 109, 13857–13870 CrossRef CAS PubMed .
  48. B. Nikoobakht and M. A. El-Sayed, Chem. Mater., 2003, 15, 1957–1962 CrossRef CAS .
  49. X. Huang, S. Neretina and M. A. El-Sayed, Adv. Mater., 2009, 21, 4880–4910 CrossRef CAS PubMed .
  50. M. Iqbal, Y.-I. Chung and G. Tae, J. Mater. Chem. B, 2007, 17, 335–342 RSC .
  51. N. R. Jana, L. Gearheart and C. J. Murphy, Chem. Commun., 2001, 617–618,  10.1039/b100521i .
  52. J. Q. Hu, Q. Chen, Z. X. Xie, G. B. Han, R. H. Wang, B. Ren, Y. Zhang, Z. L. Yang and Z. Q. Tian, Adv. Funct. Mater., 2004, 14, 183–189 CrossRef CAS .
  53. R. Gans, AnP, 1915, 352, 270–284 Search PubMed .
  54. S. Link, M. B. Mohamed and M. A. El-Sayed, J. Phys. Chem. B, 1999, 103, 3073–3077 CrossRef CAS .
  55. K. S. Lee and M. A. El-Sayed, J. Phys. Chem. B, 2005, 109, 20331–20338 CrossRef CAS PubMed .
  56. X. Huang and M. A. El-Sayed, J. Adv. Res., 2010, 1, 13–28 CrossRef .
  57. E. Hao and G. C. Schatz, J. Chem. Phys., 2003, 120, 357–366 CrossRef PubMed .
  58. M. Hu, C. Novo, A. Funston, H. Wang, H. Staleva, S. Zou, P. Mulvaney, Y. Xia and G. V. Hartland, J. Mater. Chem. B, 2008, 18, 1949–1960 RSC .
  59. M. A. Mackey, M. R. Ali, L. A. Austin, R. D. Near and M. A. El-Sayed, J. Phys. Chem. B, 2014, 118, 1319–1326 CrossRef CAS PubMed .
  60. M. A. Mahmoud and M. A. El-Sayed, J. Phys. Chem. Lett., 2013, 4, 1541–1545 CrossRef CAS PubMed .
  61. M. L. Taylor, R. E. Wilson, K. D. Amrhein and X. Huang, Bioengineering, 2022, 9, 200 CrossRef CAS PubMed .
  62. T. K. Sau and C. J. Murphy, J. Am. Chem. Soc., 2004, 126, 8648–8649 CrossRef CAS PubMed .
  63. H.-L. Wu, C.-H. Kuo and M. H. Huang, Langmuir, 2010, 26, 12307–12313 CrossRef CAS PubMed .
  64. J.-E. Park, Y. Lee and J.-M. Nam, Nano Lett., 2018, 18, 6475–6482 CrossRef CAS PubMed .
  65. S. E. Skrabalak, L. Au, X. Li and Y. Xia, Nat. Protoc., 2007, 2, 2182–2190 CrossRef CAS PubMed .
  66. Y. Sun and Y. Xia, Science, 2002, 298, 2176–2179 CrossRef CAS PubMed .
  67. Q. Zhang, W. Li, C. Moran, J. Zeng, J. Chen, L. P. Wen and Y. Xia, J. Am. Chem. Soc., 2010, 132, 11372–11378 CrossRef CAS PubMed .
  68. S. H. Im, Y. T. Lee, B. Wiley and Y. Xia, Angew. Chem., Int. Ed., 2005, 44, 2154–2157 CrossRef CAS PubMed .
  69. A. R. Siekkinen, J. M. McLellan, J. Chen and Y. Xia, Chem. Phys. Lett., 2006, 432, 491–496 CrossRef CAS PubMed .
  70. Q. Zhang, W. Li, L. P. Wen, J. Chen and Y. Xia, Chemistry, 2010, 16, 10234–10239 CrossRef CAS PubMed .
  71. S. V. Jenkins, T. D. Gohman, E. K. Miller and J. Chen, J. Chem. Educ., 2015, 92, 1056–1060 CrossRef CAS .
  72. I. I. Niyonshuti, V. R. Krishnamurthi, D. Okyere, L. Song, M. Benamara, X. Tong, Y. Wang and J. Chen, ACS Appl. Mater. Interfaces, 2020, 12, 40067–40077 CrossRef CAS PubMed .
  73. S. E. Skrabalak, J. Chen, Y. Sun, X. Lu, L. Au, C. M. Cobley and Y. Xia, Acc. Chem. Res., 2008, 41, 1587–1595 CrossRef CAS PubMed .
  74. T. E. Karam, H. T. Smith and L. H. Haber, J. Phys. Chem. C, 2015, 119, 18573–18580 CrossRef CAS .
  75. Y. Sun, B. Wiley, Z.-Y. Li and Y. Xia, J. Am. Chem. Soc., 2004, 126, 9399–9406 CrossRef CAS PubMed .
  76. M. Tavakkoli Yaraki, S. Daqiqeh Rezaei and Y. N. Tan, Phys. Chem. Chem. Phys., 2020, 22, 5673–5687 RSC .
  77. E. Ringe, J. M. McMahon, K. Sohn, C. Cobley, Y. Xia, J. Huang, G. C. Schatz, L. D. Marks and R. P. Van Duyne, J. Phys. Chem. C, 2010, 114, 12511–12516 CrossRef CAS .
  78. F. Zhou, Z. Y. Li, Y. Liu and Y. N. Xia, J. Phys. Chem. C, 2008, 112, 20233–20240 CrossRef CAS .
  79. M. Rycenga, X. Xia, C. H. Moran, F. Zhou, D. Qin, Z. Y. Li and Y. Xia, Angew. Chem., Int. Ed., 2011, 50, 5473–5477 CrossRef CAS PubMed .
  80. J. M. McLellan, A. Siekkinen, J. Chen and Y. Xia, Chem. Phys. Lett., 2006, 427, 122–126 CrossRef CAS .
  81. S. Barbosa, A. Agrawal, L. Rodriguez-Lorenzo, I. Pastoriza-Santos, R. A. Alvarez-Puebla, A. Kornowski, H. Weller and L. M. Liz-Marzan, Langmuir, 2010, 26, 14943–14950 CrossRef CAS PubMed .
  82. P. Senthil Kumar, I. Pastoriza-Santos, B. Rodríguez-González, F. Javier García de Abajo and L. M. Liz-Marzán, Nanotechnology, 2008, 19, 015606 CrossRef PubMed .
  83. C. G. Khoury and T. Vo-Dinh, J. Phys. Chem. C, 2008, 112, 18849–18859 CrossRef CAS PubMed .
  84. W. Niu, Y. A. A. Chua, W. Zhang, H. Huang and X. Lu, J. Am. Chem. Soc., 2015, 137, 10460–10463 CrossRef CAS PubMed .
  85. A. S. De Silva Indrasekara, S. F. Johnson, R. A. Odion and T. Vo-Dinh, ACS Omega, 2018, 3, 2202–2210 CrossRef CAS PubMed .
  86. Y. Wang, P. H. C. Camargo, S. E. Skrabalak, H. Gu and Y. Xia, Langmuir, 2008, 24, 12042–12046 CrossRef CAS PubMed .
  87. A. Garcia-Leis, J. V. Garcia-Ramos and S. Sanchez-Cortes, J. Phys. Chem. C, 2013, 117, 7791–7795 CrossRef CAS .
  88. F. Hao, C. L. Nehl, J. H. Hafner and P. Nordlander, Nano Lett., 2007, 7, 729–732 CrossRef CAS PubMed .
  89. N. Pazos-Perez, L. Guerrini and R. A. Alvarez-Puebla, ACS Omega, 2018, 3, 17173–17179 CrossRef CAS PubMed .
  90. Y. Pu, Y. Zhao, P. Zheng and M. Li, Inorg. Chem., 2018, 57, 8599–8607 CrossRef CAS PubMed .
  91. S. K. Cushing, C.-J. Chen, C. L. Dong, X.-T. Kong, A. O. Govorov, R.-S. Liu and N. Wu, ACS Nano, 2018, 12, 7117–7126 CrossRef CAS PubMed .
  92. Y. Ma, W. Li, E. C. Cho, Z. Li, T. Yu, J. Zeng, Z. Xie and Y. Xia, ACS Nano, 2010, 4, 6725–6734 CrossRef CAS PubMed .
  93. L. Lu, G. Burkey, I. Halaciuga and D. V. Goia, J. Colloid Interface Sci., 2013, 392, 90–95 CrossRef CAS PubMed .
  94. F. R. Fan, D. Y. Liu, Y. F. Wu, S. Duan, Z. X. Xie, Z. Y. Jiang and Z. Q. Tian, J. Am. Chem. Soc., 2008, 130, 6949–6951 CrossRef CAS PubMed .
  95. Y. G. Sun, B. Wiley, Z. Y. Li and Y. N. Xia, J. Am. Chem. Soc., 2004, 126, 9399–9406 CrossRef CAS PubMed .
  96. Z. Huang, G. Meng, X. Hu, Q. Pan, D. Huo, H. Zhou, Y. Ke and N. Wu, Nano Res., 2018, 12, 449–455 CrossRef .
  97. H. Wang and M. Li, J. Phys. Chem. Lett., 2021, 12, 4928–4935 CrossRef CAS PubMed .
  98. S. A. S. Tali and W. Zhou, Nanophotonics, 2019, 8, 1199–1225 CrossRef .
  99. J. Gong, F. Zhou, Z. Li and Z. Tang, Langmuir, 2012, 28, 8959–8964 CrossRef CAS PubMed .
  100. M. W. Knight and N. J. Halas, New J. Phys., 2008, 10, 105006 CrossRef .
  101. Y. Hu, S. J. Noelck and R. A. Drezek, ACS Nano, 2010, 4, 1521–1528 CrossRef CAS PubMed .
  102. V. G. Kravets, A. V. Kabashin, W. L. Barnes and A. N. Grigorenko, Chem. Rev., 2018, 118, 5912–5951 CrossRef CAS PubMed .
  103. W. L. Barnes, A. Dereux and T. W. Ebbesen, Nature, 2003, 424, 824–830 CrossRef CAS PubMed .
  104. S. Kasani, K. Curtin and N. Wu, Nanophotonics, 2019, 8, 2065–2089 CrossRef CAS .
  105. T. Maurer, P.-M. Adam and G. Lévêque, Nanophotonics, 2015, 4, 363–382 CAS .
  106. E. Prodan, C. Radloff, N. J. Halas and P. Nordlander, Science, 2003, 302, 419–422 CrossRef CAS PubMed .
  107. W. Zhou, J. Y. Suh, Y. Hua and T. W. Odom, J. Phys. Chem. C, 2012, 117, 2541–2546 CrossRef .
  108. W. Zhou, Y. Hua, M. D. Huntington and T. W. Odom, J. Phys. Chem. Lett., 2012, 3, 1381–1385 CrossRef CAS PubMed .
  109. S. Yan, C. Zhu, A. Wang, J. Sun, Y. Hang, B. Chen, X. Wang, N. Wu, H. Tang, L. Wen and G. Meng, Adv. Opt. Mater., 2023, 11, 2300508 CrossRef CAS .
  110. P. Zheng, S. Kasani, W. Tan, J. Boryczka, X. Gao, F. Yang and N. Wu, Anal. Chim. Acta, 2022, 1203, 339721 CrossRef CAS PubMed .
  111. M. Horak, K. Bukvisova, V. Svarc, J. Jaskowiec, V. Krapek and T. Sikola, Sci. Rep., 2018, 8, 9640 CrossRef PubMed .
  112. D. K. Oh, T. Lee, B. Ko, T. Badloe, J. G. Ok and J. Rho, Front. Optoelectron., 2021, 14, 229–251 CrossRef PubMed .
  113. X. Chen, H. R. Park, M. Pelton, X. Piao, N. C. Lindquist, H. Im, Y. J. Kim, J. S. Ahn, K. J. Ahn, N. Park, D. S. Kim and S. H. Oh, Nat. Commun., 2013, 4, 2361 CrossRef PubMed .
  114. W. Zhou and T. W. Odom, Nat. Nanotechnol., 2011, 6, 423–427 CrossRef CAS PubMed .
  115. G. Liu, S. H. Petrosko, Z. Zheng and C. A. Mirkin, Chem. Rev., 2020, 120, 6009–6047 CrossRef CAS PubMed .
  116. W. Nam, J. Song, S. A. Safiabadi Tali, H. J. Lezec, A. Agrawal and W. Zhou, ACS Appl. Nano Mater., 2021, 4, 3175–3184 CrossRef CAS .
  117. Y. Hang, J. Boryczka and N. Wu, Chem. Soc. Rev., 2022, 51, 329–375 RSC .
  118. T. Jennings and G. Strouse, Bio-Appl. Nanopart., 2007, 620, 34–47 Search PubMed .
  119. M. H. Jazayeri, H. Amani, A. A. Pourfatollah, H. Pazoki-Toroudi and B. Sedighimoghaddam, Sens. Biosens. Res., 2016, 9, 17–22 Search PubMed .
  120. J. O. Tam, H. de Puig, C. W. Yen, I. Bosch, J. Gomez-Marquez, C. Clavet, K. Hamad-Schifferli and L. Gehrke, J. Immunoassay Immunochem., 2017, 38, 355–377 CrossRef CAS PubMed .
  121. Z. Gao, S. Shao, W. Gao, D. Tang, D. Tang, S. Zou, M. J. Kim and X. Xia, ACS Nano, 2021, 15, 2428–2438 CrossRef CAS PubMed .
  122. D. Liu, Y. Zhang, M. Zhu, Z. Yu, X. Ma, Y. Song, S. Zhou and C. Yang, Anal. Chem., 2020, 92, 11826–11833 CrossRef CAS PubMed .
  123. P. Teengam, W. Siangproh, A. Tuantranont, T. Vilaivan, O. Chailapakul and C. S. Henry, Anal. Chem., 2017, 89, 5428–5435 CrossRef CAS PubMed .
  124. Y. J. Kim, R. C. Johnson and J. T. Hupp, Nano Lett., 2001, 1, 165–167 CrossRef CAS .
  125. J. Hu, L. Wang, F. Li, Y. L. Han, M. Lin, T. J. Lu and F. Xu, Lab Chip, 2013, 13, 4352–4357 RSC .
  126. Z. Jin, Y. Mantri, M. Retout, Y. Cheng, J. Zhou, A. Jorns, P. Fajtova, W. Yim, C. Moore, M. Xu, M. N. Creyer, R. M. Borum, J. Zhou, Z. Wu, T. He, W. F. Penny, A. J. O'Donoghue and J. V. Jokerst, Angew. Chem., Int. Ed., 2022, 61, e202112995 CrossRef CAS PubMed .
  127. A. Maity, D. Bagchi, S. K. De and A. Chakraborty, Langmuir, 2023, 39, 4881–4894 CrossRef CAS PubMed .
  128. D. H. Choi, S. K. Lee, Y. K. Oh, B. W. Bae, S. D. Lee, S. Kim, Y. B. Shin and M. G. Kim, Biosens. Bioelectron., 2010, 25, 1999–2002 CrossRef CAS PubMed .
  129. Y. Gao, J. Wang, W. Wang, T. Zhao, Y. Cui, P. Liu, S. Xu and X. Luo, Anal. Chem., 2021, 93, 2480–2489 CrossRef CAS PubMed .
  130. X. Yang, Q. Yu, X. Cheng, H. Wei, X. Zhang, Z. Rong, C. Wang and S. Wang, ACS Appl. Mater. Interfaces, 2023, 15, 12327–12338 CrossRef CAS PubMed .
  131. T. T. Tsai, T. H. Huang, N. Y. Ho, Y. P. Chen, C. A. Chen and C. F. Chen, Sci. Rep., 2019, 9, 15679 CrossRef PubMed .
  132. Y. Wu, Y. Zhou, Y. Leng, W. Lai, X. Huang and Y. Xiong, Biosens. Bioelectron., 2020, 157, 112168 CrossRef CAS PubMed .
  133. N. A. Taranova, N. A. Byzova, V. V. Zaiko, T. A. Starovoitova, Y. Y. Vengerov, A. V. Zherdev and B. B. Dzantiev, Mikrochim. Acta, 2013, 180, 1165–1172 CrossRef CAS .
  134. Y. Zhao, H. Wang, P. Zhang, C. Sun, X. Wang, X. Wang, R. Yang, C. Wang and L. Zhou, Sci. Rep., 2016, 6, 21342 CrossRef CAS PubMed .
  135. N. Ratnarathorn, O. Chailapakul, C. S. Henry and W. Dungchai, Talanta, 2012, 99, 552–557 CrossRef CAS PubMed .
  136. Y. Liu, J. R. Ashton, E. J. Moding, H. Yuan, J. K. Register, A. M. Fales, J. Choi, M. J. Whitley, X. Zhao, Y. Qi, Y. Ma, G. Vaidyanathan, M. R. Zalutsky, D. G. Kirsch, C. T. Badea and T. Vo-Dinh, Theranostics, 2015, 5, 946–960 CrossRef CAS PubMed .
  137. T. J. Wax, S. Dey, S. Chen, Y. Luo, S. Zou and J. Zhao, ACS Omega, 2018, 3, 14151–14156 CrossRef CAS PubMed .
  138. N. S. Abadeer, M. R. Brennan, W. L. Wilson and C. J. Murphy, ACS Nano, 2014, 8, 8392–8406 CrossRef CAS PubMed .
  139. C. Wang, X. Yang, B. Gu, H. Liu, Z. Zhou, L. Shi, X. Cheng and S. Wang, Anal. Chem., 2020, 92, 15542–15549 CrossRef CAS PubMed .
  140. J. Wang, F. Cao, S. He, Y. Xia, X. Liu, W. Jiang, Y. Yu, H. Zhang and W. Chen, Talanta, 2018, 176, 444–449 CrossRef CAS PubMed .
  141. Y. Wang, J. Zhao, Y. Zhu, S. Dong, Y. Liu, Y. Sun, L. Qian, W. Yang and Z. Cao, Microsyst. Nanoeng., 2021, 7, 65 CrossRef CAS PubMed .
  142. X. Gao, P. Zheng, S. Kasani, S. Wu, F. Yang, S. Lewis, S. Nayeem, E. B. Engler-Chiurazzi, J. G. Wigginton, J. W. Simpkins and N. Wu, Anal. Chem., 2017, 89, 10104–10110 CrossRef CAS PubMed .
  143. P. Zheng, S. Kasani, X. Shi, A. E. Boryczka, F. Yang, H. Tang, M. Li, W. Zheng, D. E. Elswick and N. Wu, Anal. Chim. Acta, 2018, 1040, 158–165 CrossRef CAS PubMed .
  144. O. O. Abugo, R. Nair and J. R. Lakowicz, Anal. Biochem., 2000, 279, 142–150 CrossRef CAS PubMed .
  145. C. Swanson and A. D'Andrea, Clin. Chem., 2013, 59, 641–648 CrossRef CAS PubMed .
  146. E. Hemmer, N. Venkatachalam, H. Hyodo, A. Hattori, Y. Ebina, H. Kishimoto and K. Soga, Nanoscale, 2013, 5, 11339–11361 RSC .
  147. X. Gao, J. Boryczka, P. Zheng, S. Kasani, F. Yang, E. B. Engler-Chiurazzi, J. W. Simpkins, J. G. Wigginton and N. Wu, Biosens. Bioelectron., 2021, 177, 112967 CrossRef CAS PubMed .
  148. X. Gao, J. Boryczka, S. Kasani and N. Wu, Anal. Chem., 2021, 93, 1326–1332 CrossRef CAS PubMed .
  149. K. B. Shanmugasundaram, J. Li, A. I. Sina, A. Wuethrich and M. Trau, Mater. Adv., 2022, 3, 1459–1471 RSC .
  150. L. Wu, A. Teixeira, A. Garrido-Maestu, L. Muinelo-Romay, L. Lima, L. L. Santos, M. Prado and L. Diéguez, Biosens. Bioelectron., 2020, 165, 112392 CrossRef CAS PubMed .
  151. J. Wang, A. Wuethrich, A. A. Sina, R. E. Lane, L. L. Lin, Y. Wang, J. Cebon, A. Behren and M. Trau, Sci. Adv., 2020, 6, eaax3223 CrossRef CAS PubMed .
  152. L. P. Ge, X. Ye, Z. P. Yu, B. Chen, C. J. Liu, H. Guo, S. Y. Zhang, F. Sassa and K. Hayashi, Npj Flexible Electron., 2022, 6, 40 CrossRef CAS .
  153. H. Wei, W. Leng, J. Song, M. R. Willner, L. C. Marr, W. Zhou and P. J. Vikesland, Anal. Chem., 2018, 90, 3227–3237 CrossRef CAS PubMed .
  154. A. Loren, J. Engelbrektsson, C. Eliasson, M. Josefson, J. Abrahamsson, M. Johansson and K. Abrahamsson, Anal. Chem., 2004, 76, 7391–7395 CrossRef CAS PubMed .
  155. W. Nam, Y. Zhao, J. Song, S. Ali Safiabadi Tali, S. Kang, W. Zhu, H. J. Lezec, A. Agrawal, P. J. Vikesland and W. Zhou, J. Phys. Chem. Lett., 2020, 11, 9543–9551 CrossRef CAS PubMed .
  156. A. Subaihi, Y. Xu, H. Muhamadali, S. T. Mutter, E. W. Blanch, D. I. Ellis and R. Goodacre, Anal. Methods, 2017, 9, 6636–6644 RSC .
  157. C. L. M. Morais, K. M. G. Lima, M. Singh and F. L. Martin, Nat. Protoc., 2020, 15, 2143–2162 CrossRef CAS PubMed .
  158. (a) H. Tang, C. Zhu, G. Meng and N. Q. Wu, J. Electrochem. Soc., 2018, 165(8), B3098–B3118 CrossRef CAS ; (b) C. Zhu, X. Wang, X. Shi, F. Yang, G. Meng, Q. Xiong, Y. Ke, H. Wang, Y. Lu and N. Q. Wu, ACS Appl. Mater. Interfaces, 2017, 9, 39618–39625 CrossRef CAS PubMed ; (c) F. Lussier, D. Missirlis, J. P. Spatz and J. F. Masson, ACS Nano, 2019, 13, 1403–1411 CAS .
  159. W. Nam, H. Chen, X. Ren, M. Agah, I. Kim and W. Zhou, ACS Appl. Nano Mater., 2022, 5, 10358–10368 CrossRef CAS .
  160. M. M. Richter, Chem. Rev., 2004, 104, 3003–3036 CrossRef CAS PubMed .
  161. X. Ying, L. Zhou, W. Fu, Y. Wang and B. Su, Sens. Diagn., 2023, 2, 480–491 RSC .
  162. C. Ma, Z. Zhang, T. Tan and J. J. Zhu, Biosensors, 2023, 13, 200 CrossRef CAS PubMed .
  163. E. M. Gross, H. E. Durant, K. N. Hipp and R. Y. Lai, ChemElectroChem, 2017, 4, 1594–1603 CrossRef CAS .
  164. Y. Uludag, Z. Olcer and M. S. Sagiroglu, Biosens. Bioelectron., 2014, 57, 85–90 CrossRef CAS PubMed .
  165. A. Fernández-la-Villa, D. F. Pozo-Ayuso and M. Castaño-Álvarez, Curr. Opin. Electrochem., 2019, 15, 175–185 CrossRef .
  166. Y. Chen, K. Munechika and D. S. Ginger, Nano Lett., 2007, 7, 690–696 CrossRef CAS PubMed .
  167. Y. Shan, J. J. Xu and H. Y. Chen, Chem. Commun., 2009, 905–907,  10.1039/b821049g .
  168. S. Jose, R. Rajeev, D. A. Thadathil, A. Varghese and G. Hegde, J. Sci.: Adv. Mater. Devices, 2022, 7, 100460 CAS .
  169. L. L. Shen, G. R. Zhang and B. J. M. Etzold, ChemElectroChem, 2020, 7, 10–30 CrossRef CAS PubMed .
  170. T. Zhan, Y. Su, W. Lai, Z. Chen and C. Zhang, Biosens. Bioelectron., 2022, 214, 114494 CrossRef CAS PubMed .
  171. D. Hong, E. J. Jo, K. Kim, M. B. Song and M. G. Kim, Small, 2020, 16, e2004535 CrossRef PubMed .
  172. S. R. Chinnadayyala, J. Park, H. T. N. Le, M. Santhosh, A. N. Kadam and S. Cho, Biosens. Bioelectron., 2019, 126, 68–81 CrossRef CAS PubMed .
  173. Q. M. Feng, J. B. Pan, H. R. Zhang, J. J. Xu and H. Y. Chen, Chem. Commun., 2014, 50, 10949–10951 RSC .
  174. L. Zhu, W. Fu, B. Zhu, Q. Feng, X. Ying, S. Li, J. Chen, X. Xie, C. Pan, J. Liu, C. Chen, X. Chen and D. Zhu, Talanta, 2023, 262, 124626 CrossRef CAS PubMed .
  175. A. Krishnakumar, R. K. Mishra, S. Kadian, A. Zareei, U. H. Rivera and R. Rahimi, Anal. Chim. Acta, 2022, 1229, 340332 CrossRef CAS PubMed .
  176. A. Devadoss, P. Sudhagar, C. Terashima, K. Nakata and A. Fujishima, J. Photochem. Photobiol., C, 2015, 24, 43–63 CrossRef CAS .
  177. J. Shu and D. Tang, Anal. Chem., 2020, 92, 363–377 CrossRef CAS PubMed .
  178. Z. Yu, H. Gong, Y. Li, J. Xu, J. Zhang, Y. Zeng, X. Liu and D. Tang, Anal. Chem., 2021, 93, 13389–13397 CrossRef CAS PubMed .
  179. S. K. Cushing, J. Li, F. Meng, T. R. Senty, S. Suri, M. Zhi, M. Li, A. D. Bristow and N. Wu, J. Am. Chem. Soc., 2012, 134, 15033–15041 CrossRef CAS PubMed .
  180. J. Shu and D. Tang, Chem. – Asian J., 2017, 12, 2780–2789 CrossRef CAS PubMed .
  181. A. Devadoss, P. Sudhagar, S. Das, S. Y. Lee, C. Terashima, K. Nakata, A. Fujishima, W. Choi, Y. S. Kang and U. Paik, ACS Appl. Mater. Interfaces, 2014, 6, 4864–4871 CrossRef CAS PubMed .
  182. B. Zhang, H. Wang, H. Ye, B. Xu, F. Zhao and B. Zeng, Sens. Actuators, B, 2018, 273, 1435–1441 CrossRef CAS .
  183. P. Wang, L. Ge, S. Ge, J. Yu, M. Yan and J. Huang, Chem. Commun., 2013, 49, 3294–3296 RSC .
  184. Y. Liu, L. Zhong, S. Zhang, J. Wang and Z. Liu, Sens. Actuators, B, 2022, 354, 131204 CrossRef CAS .
  185. X. Han, K. Xu, O. Taratula and K. Farsad, Nanoscale, 2019, 11, 799–819 RSC .
  186. (a) Y. S. Chen, W. Frey, S. Kim, P. Kruizinga, K. Homan and S. Emelianov, Nano Lett., 2011, 11, 348–354 CrossRef CAS PubMed ; (b) M. Horie and Y. Tabei, Free Radic. Res., 2021, 55(4), 331–342 CrossRef CAS PubMed ; (c) R. F. Hamilton, N. Wu, C. Xiang, M. Li, F. Yang, M. Wolfarth, D. W. Porter and A. Holian, Part. Fibre Toxicol., 2014, 11, 43 CrossRef PubMed .
  187. J. Shang, J. S. Fan, W. W. Qin and K. Li, Catalysts, 2022, 12, 764 CrossRef CAS .
  188. M. Liu, J. Chao, S. Deng, K. Wang, K. Li and C. Fan, Colloids Surf., B, 2014, 124, 111–117 CrossRef CAS PubMed .
  189. J. Renger, R. Quidant and L. Novotny, Opt. Express, 2011, 19, 1777–1785 CrossRef CAS PubMed .
  190. J. W. Jarrett, T. Zhao, J. S. Johnson and K. L. Knappenberger, J. Phys. Chem. C, 2015, 119, 15779–15800 CrossRef CAS .
  191. C. A. C. Chazot, S. Nagelberg, C. J. Rowlands, M. R. J. Scherer, I. Coropceanu, K. Broderick, Y. Kim, M. G. Bawendi, P. So and M. Kolle, Nat. Photonics, 2020, 14, 310–315 CrossRef CAS PubMed .
  192. M. Liu, Q. Li, L. Liang, J. Li, K. Wang, J. Li, M. Lv, N. Chen, H. Song, J. Lee, J. Shi, L. Wang, R. Lal and C. Fan, Nat. Commun., 2017, 8, 15646 CrossRef CAS PubMed .
  193. V. Rodriguez-Fajardo, V. Sanz, I. de Miguel, J. Berthelot, S. S. Acimovic, R. Porcar-Guezenec and R. Quidant, Nanoscale, 2018, 10, 4019–4027 RSC .
  194. B. Xiong, Z. Huang, H. Zou, C. Qiao, Y. He and E. S. Yeung, ACS Nano, 2017, 11, 541–548 CrossRef CAS PubMed .
  195. S. Wu, C. Ma, Y. Gao, Y. Su, Q. Xia, Z. Chen and J. J. Zhu, Anal. Chem., 2020, 92, 9940–9947 CrossRef CAS PubMed .
  196. S. K. Cool, K. Breyne, E. Meyer, S. C. De Smedt and N. N. Sanders, J. Fluoresc., 2013, 23, 909–920 CrossRef CAS PubMed .
  197. J. F. Li, C. Y. Li and R. F. Aroca, Chem. Soc. Rev., 2017, 46, 3962–3979 RSC .
  198. A. S. Maddox and P. S. Maddox, in Methods Cell Biology, ed. J. H. Rothman and A. Singson, Academic Press, 2012, vol. 107, pp. 1–34 Search PubMed .
  199. K. I. Willig, S. O. Rizzoli, V. Westphal, R. Jahn and S. W. Hell, Nature, 2006, 440, 935–939 CrossRef CAS PubMed .
  200. G. Perinetti, T. Müller, A. Spaar, R. Polishchuk, A. Luini and A. Egner, Traffic, 2009, 10, 379–391 CrossRef CAS PubMed .
  201. T. S. van Zanten, M. J. Lopez-Bosque and M. F. Garcia-Parajo, Small, 2010, 6, 270–275 CrossRef CAS PubMed .
  202. A. Refaat, M. L. Yap, G. Pietersz, A. P. G. Walsh, J. Zeller, B. Del Rosal, X. Wang and K. Peter, J. Nanobiotechnol., 2022, 20, 450 CrossRef PubMed .
  203. J. E. Donehue, E. Wertz, C. N. Talicska and J. S. Biteen, J. Phys. Chem. C, 2014, 118, 15027–15035 CrossRef CAS .
  204. C. Joyce, S. M. Fothergill and F. Xie, Mater. Today Adv., 2020, 7, 100073 CrossRef .
  205. I. G. Theodorou, P. Ruenraroengsak, D. A. Gonzalez-Carter, Q. Jiang, E. Yague, E. O. Aboagye, R. C. Coombes, A. E. Porter, M. P. Ryan and F. Xie, Nanoscale, 2019, 11, 2079–2088 RSC .
  206. A. L. Vahrmeijer, M. Hutteman, J. R. van der Vorst, C. J. van de Velde and J. V. Frangioni, Nat. Rev. Clin. Oncol., 2013, 10, 507–518 CrossRef CAS PubMed .
  207. Q. Sun, W. Wang, Z. Chen, Y. Yao, W. Zhang, L. Duan and J. Qian, Chem. Commun., 2017, 53, 6432–6435 RSC .
  208. Y. Song, H. Zhang, X. Wang, X. Geng, Y. Sun, J. Liu and Z. Li, Anal. Chem., 2021, 93, 1786–1791 CrossRef CAS PubMed .
  209. J. Yin, X. Kong and W. Lin, Anal. Chem., 2021, 93, 2072–2081 CrossRef CAS PubMed .
  210. M. Zhang, J. Yue, R. Cui, Z. Ma, H. Wan, F. Wang, S. Zhu, Y. Zhou, Y. Kuang, Y. Zhong, D.-W. Pang and H. Dai, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 6590–6595 CrossRef CAS PubMed .
  211. C. H. Li, T. R. Kuo, H. J. Su, W. Y. Lai, P. C. Yang, J. S. Chen, D. Y. Wang, Y. C. Wu and C. C. Chen, Sci. Rep., 2015, 5, 15675 CrossRef CAS PubMed .
  212. R. Yan, Y. Hu, F. Liu, S. Wei, D. Fang, A. J. Shuhendler, H. Liu, H.-Y. Chen and D. Ye, J. Am. Chem. Soc., 2019, 141, 10331–10341 CrossRef CAS PubMed .
  213. W. Wang, J. Zhao, M. Short and H. Zeng, J. Biophotonics, 2015, 8, 527–545 CrossRef PubMed .
  214. S. Hosaka, Advances in Solid State Circuits Technologies, InTech China, Shanghai, 2010, pp. 431–446 Search PubMed .
  215. J. S. Lin, X. D. Tian, G. Li, F. L. Zhang, Y. Wang and J. F. Li, Chem. Soc. Rev., 2022, 51, 9445–9468 RSC .
  216. S. Harmsen, S. Rogalla, R. Huang, M. Spaliviero, V. Neuschmelting, Y. Hayakawa, Y. Lee, Y. Tailor, R. Toledo-Crow, J. W. Kang, J. M. Samii, H. Karabeber, R. M. Davis, J. R. White, M. van de Rijn, S. S. Gambhir, C. H. Contag, T. C. Wang and M. F. Kircher, ACS Nano, 2019, 13, 1354–1364 CAS .
  217. L. Puppulin, S. Hosogi, H. Sun, K. Matsuo, T. Inui, Y. Kumamoto, T. Suzaki, H. Tanaka and Y. Marunaka, Nat. Commun., 2018, 9, 5278 CrossRef CAS PubMed .
  218. T. Yang, Z. Zhang, B. Zhao, R. Hou, A. Kinchla, J. M. Clark and L. He, Anal. Chem., 2016, 88, 5243–5250 CrossRef CAS PubMed .
  219. J. Li, F. Liu, X. Bi and J. Ye, Biomaterials, 2023, 302, 122327 CrossRef CAS PubMed .
  220. Q. Li, X. Ge, J. Ye, Z. Li, L. Su, Y. Wu, H. Yang and J. Song, Angew. Chem., Int. Ed., 2021, 60, 7323–7332 CrossRef CAS PubMed .
  221. Y. S. Chen, W. Frey, S. Aglyamov and S. Emelianov, Small, 2012, 8, 47–52 CrossRef CAS PubMed .
  222. C. L. Bayer, S. Y. Nam, Y. S. Chen and S. Y. Emelianov, J. Biomed. Opt., 2013, 18, 16001 CrossRef PubMed .
  223. Y. Mantri and J. V. Jokerst, ACS Nano, 2020, 14, 9408–9422 CrossRef CAS PubMed .
  224. T. Repenko, A. Rix, A. Nedilko, J. Rose, A. Hermann, R. Vinokur, S. Moli, R. Cao-Milàn, M. Mayer, G. von Plessen, A. Fery, L. De Laporte, W. Lederle, D. N. Chigrin and A. J. C. Kuehne, Adv. Funct. Mater., 2017, 28, 1705607 CrossRef .
  225. V.-P. Nguyen, Y. Li, J. Henry, W. Zhang, M. Aaberg, S. Jones, T. Qian, X. Wang and Y. M. Paulus, ACS Sens., 2020, 5, 3070–3081 CrossRef CAS PubMed .
  226. C. Tian, W. Qian, X. Shao, Z. Xie, X. Cheng, S. Liu, Q. Cheng, B. Liu and X. Wang, Adv. Sci., 2016, 3, 1600237 CrossRef PubMed .
  227. W. Yim, J. Zhou, Y. Mantri, M. N. Creyer, C. A. Moore and J. V. Jokerst, ACS Appl. Mater. Interfaces, 2021, 13, 14974–14984 CrossRef CAS PubMed .
  228. J. Kim, A. M. Yu, K. P. Kubelick and S. Y. Emelianov, Photoacoustics, 2022, 25, 100307 CrossRef PubMed .
  229. C. Zhou, L. Zhang, T. Sun, Y. Zhang, Y. Liu, M. Gong, Z. Xu, M. Du, Y. Liu, G. Liu and D. Zhang, Adv. Mater., 2021, 33, e2006532 CrossRef PubMed .
  230. Y. Mantri, B. Davidi, J. E. Lemaster, A. Hariri and J. V. Jokerst, Nanoscale, 2020, 12, 10511–10520 RSC .
  231. X. Huang, P. K. Jain, I. H. El-Sayed and M. A. El-Sayed, Lasers Med. Sci., 2008, 23, 217–228 CrossRef PubMed .
  232. V. P. Zharov, E. N. Galitovskaya, C. Johnson and T. Kelly, Lasers Surg. Med., 2005, 37, 219–226 CrossRef PubMed .
  233. T. C. Pham, V. N. Nguyen, Y. Choi, S. Lee and J. Yoon, Chem. Rev., 2021, 121, 13454–13619 CrossRef CAS PubMed .
  234. S. Link and M. A. El-Sayed, Int. Rev. Phys. Chem., 2000, 19, 409–453 Search PubMed .
  235. J. Chen, C. Glaus, R. Laforest, Q. Zhang, M. Yang, M. Gidding, M. J. Welch and Y. Xia, Small, 2010, 6, 811–817 CrossRef CAS PubMed .
  236. J. R. Cole, N. A. Mirin, M. W. Knight, G. P. Goodrich and N. J. Halas, J. Phys. Chem. C, 2009, 113, 12090–12094 CrossRef CAS .
  237. A. B. Bucharskaya, N. G. Khlebtsov, B. N. Khlebtsov, G. N. Maslyakova, N. A. Navolokin, V. D. Genin, E. A. Genina and V. V. Tuchin, Materials, 2022, 15, 1606 CrossRef CAS PubMed .
  238. X. Li, Y. Zhang, G. Liu, Z. Luo, L. Zhou, Y. Xue and M. Liu, RSC Adv., 2022, 12, 7635–7651 RSC .
  239. X. Deng, Z. Shao and Y. Zhao, Adv. Sci., 2021, 8, 2002504 CrossRef CAS PubMed .
  240. J. Wu, Y. Zheng, S. Jiang, Y. Qu, T. Wei, W. Zhan, L. Wang, Q. Yu and H. Chen, ACS Appl. Mater. Interfaces, 2019, 11, 12357–12366 CrossRef CAS PubMed .
  241. A. R. Rastinehad, H. Anastos, E. Wajswol, J. S. Winoker, J. P. Sfakianos, S. K. Doppalapudi, M. R. Carrick, C. J. Knauer, B. Taouli, S. C. Lewis, A. K. Tewari, J. A. Schwartz, S. E. Canfield, A. K. George, J. L. West and N. J. Halas, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 18590–18596 CrossRef CAS PubMed .
  242. S. Wilhelm, A. J. Tavares, Q. Dai, S. Ohta, J. Audet, H. F. Dvorak and W. C. W. Chan, Nat. Rev. Mater., 2016, 1, 16014 CrossRef CAS .
  243. V. D. Nguyen, H. K. Min, D. H. Kim, C. S. Kim, J. Han, J. O. Park and E. Choi, ACS Appl. Mater. Interfaces, 2020, 12, 10130–10141 CrossRef CAS PubMed .
  244. J. S. Basuki, F. Qie, X. Mulet, R. Suryadinata, A. V. Vashi, Y. Y. Peng, L. Li, X. Hao, T. Tan and T. C. Hughes, Angew. Chem., Int. Ed., 2017, 56, 966–971 CrossRef CAS PubMed .
  245. S. Kwiatkowski, B. Knap, D. Przystupski, J. Saczko, E. Kedzierska, K. Knap-Czop, J. Kotlinska, O. Michel, K. Kotowski and J. Kulbacka, Biomed. Pharmacother., 2018, 106, 1098–1107 CrossRef PubMed .
  246. S. K. Singh, S. Mazumder, A. Vincy, N. Hiremath, R. Kumar, I. Banerjee and R. Vankayala, ACS Appl. Nano Mater., 2023, 6, 1508–1521 CrossRef CAS .
  247. N. Zhou, V. López-Puente, Q. Wang, L. Polavarapu, I. Pastoriza-Santos and Q.-H. Xu, RSC Adv., 2015, 5, 29076–29097 RSC .
  248. B. P. George, A. Chota, P. Sarbadhikary and H. Abrahamse, Front. Chem., 2022, 10, 964674 CrossRef CAS PubMed .
  249. Y. Li, W. Zhang, J. Niu and Y. Chen, ACS Nano, 2012, 6, 5164–5173 CrossRef CAS PubMed .
  250. C. Liu, H. Dong, N. Wu, Y. Cao and X. Zhang, ACS Appl. Mater. Interfaces, 2018, 10, 6991–7002 CrossRef CAS PubMed .
  251. C. Fang, H. Jia, S. Chang, Q. Ruan, P. Wang, T. Chen and J. Wang, Energy Environ. Sci., 2014, 7, 3431–3438 RSC .
  252. S. Wang, P. Huang, L. Nie, R. Xing, D. Liu, Z. Wang, J. Lin, S. Chen, G. Niu, G. Lu and X. Chen, Adv. Mater., 2013, 25, 3055–3061 CrossRef CAS PubMed .
  253. J. Lin, S. Wang, P. Huang, Z. Wang, S. Chen, G. Niu, W. Li, J. He, D. Cui, G. Lu, X. Chen and Z. Nie, ACS Nano, 2013, 7, 5320–5329 CrossRef CAS PubMed .

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.