Nanoscale chemical characterization of materials and interfaces by tip-enhanced Raman spectroscopy

Yi-Fan Bao a, Meng-Yuan Zhu a, Xiao-Jiao Zhao a, Hong-Xuan Chen a, Xiang Wang *ab and Bin Ren *ab
aCollaborative Innovation Center of Chemistry for Energy Materials, State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China. E-mail: wangxiang@xmu.edu.cn; bren@xmu.edu.cn
bInnovation Laboratory for Sciences and Technologies of Energy Materials of Fujian Province (IKKEM), Xiamen 361005, China

Received 11th June 2024

First published on 4th September 2024


Abstract

Materials and their interfaces are the core for the development of a large variety of fields, including catalysis, energy storage and conversion. In this case, tip-enhanced Raman spectroscopy (TERS), which combines scanning probe microscopy with plasmon-enhanced Raman spectroscopy, is a powerful technique that can simultaneously obtain the morphological information and chemical fingerprint of target samples at nanometer spatial resolution. It is an ideal tool for the nanoscale chemical characterization of materials and interfaces, correlating their structures with chemical performances. In this review, we begin with a brief introduction to the nanoscale characterization of materials and interfaces, followed by a detailed discussion on the recent theoretical understanding and technical improvements of TERS, including the origin of enhancement, TERS instruments, TERS tips and the application of algorithms in TERS. Subsequently, we list the key experimental issues that need to be addressed to conduct successful TERS measurements. Next, we focus on the recent progress of TERS in the study of various materials, especially the novel low-dimensional materials, and the progresses of TERS in studying different interfaces, including both solid–gas and solid–liquid interfaces. Finally, we provide an outlook on the future developments of TERS in the study of materials and interfaces.


image file: d4cs00588k-p1.tif

Yi-Fan Bao

Yi-Fan Bao received his BSc Degree in Chemistry from Xiamen University in 2019. Currently, he is a PhD student under the supervision of Professor Xiang Wang and Professor Bin Ren at the Collaborative Innovation Center of Chemistry for Energy Materials, State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Chemistry and Chemical Engineering, Xiamen University. His current research focuses on the design and fabrication of highly active TERS tips and electrochemical-tip enhanced Raman spectroscopy (EC–TERS) study of electrocatalytic processes.

image file: d4cs00588k-p2.tif

Meng-Yuan Zhu

Meng-Yuan Zhu received her BSc Degree in Chemistry from Lanzhou University in 2020. Currently, she is a PhD student under the supervision of Professor Xiang Wang and Professor Bin Ren at the Collaborative Innovation Center of Chemistry for Energy Materials, State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Chemistry and Chemical Engineering, Xiamen University. Her current research focuses on the EC–TERS study of the electrocatalysis.

image file: d4cs00588k-p3.tif

Xiao-Jiao Zhao

Xiao-Jiao Zhao received her BSc Degree in Materials Chemistry from Shanxi Normal University in 2020. Currently, she is a PhD student under the supervision of Professor Xiang Wang and Professor Bin Ren at the Collaborative Innovation Center of Chemistry for Energy Materials, State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Chemistry and Chemical Engineering, Xiamen University. Her current research focuses on the TERS study of catalytic processes.

image file: d4cs00588k-p4.tif

Hong-Xuan Chen

Hong-Xuan Chen received his BSc Degree in Chemistry from Xiamen University in 2023. Currently, he is a PhD student under the supervision of Professor Xiang Wang and Professor Bin Ren at the Collaborative Innovation Center of Chemistry for Energy Materials, State Key Laboratory of Physical Chemistry of Solid Surfaces, College of Chemistry and Chemical Engineering, Xiamen University. His current research focuses on the EC–TERS study of electrocatalytic processes.

image file: d4cs00588k-p5.tif

Xiang Wang

Professor Xiang Wang received both his BSc Degree (2007) and PhD Degree (2013) from Xiamen University supervised by Professor Bin Ren. He joined Xiamen University as an Associate Professor in 2020. His research interests focus on developing plasmon-enhanced Raman spectroscopy (PERS), including TERS and SERS, to correlate the structure and activity of surfaces and interfaces.

image file: d4cs00588k-p6.tif

Bin Ren

Professor Bin Ren received his PhD Degree from Xiamen University in 1998 supervised by Professor Zhong-Qun Tian. He spent his sabbatical year at the Fritz-Haber Institute, Germany as an AvH Fellow during 2002 and 2003. He was promoted to a Full Professor at Xiamen University in 2004. He was awarded the Young Chemist Award from Chinese Chemical Society, the Chinese Young Electrochemist Award and the Electrochemistry Award of Analytical Chemistry Division, ACS. He was elected as a Fellow of The International Society of Electrochemistry in 2021. The research interests of Professor Bin Ren include the development of high spatial and temporal resolution Raman methods and instruments as well as deep learning-based methods for studying surfaces and interfaces as well as biological systems. He is a now an Associate Editor of Analytical Chemistry (ACS).


1. Introduction

The development of a large variety of fields highly relies on high-performance materials, including catalysis, energy storage and conversion. With the fast development of nanotechnology, it is generally recognized that when the size of materials is reduced to the nanometer scale, owing to the dramatic increase in their surface area and the quantum confinement effect, nanomaterials exhibit unique physical and chemical properties, especially with impressive optical behaviors and catalytic activities.1–3 In particular, the surfaces and the interfaces are at the heart of the materials, which connect the material itself with the outside environment and are the core for their applications. Different from their bulk material, the surfaces of materials possess unsaturated coordination sites with unique structural and electronic properties, leading to strong interactions with surface molecular species, and thus a higher activity. Some special sites (the so-called active sites) at nanoscale or even atomic scale on the surfaces or interfaces of the materials may determine their performance.4–6 Therefore, to clearly correlate the structure with the activity of both the materials and interfaces requires advanced characterization techniques that are capable of providing nanometer scale resolved structural and chemical information at the molecular level.7,8

In recent years, various advanced techniques with nanoscale or even atomic spatial resolution have started to flourish, offering great opportunities to gain a deeper understanding of the structure–activity relation of both materials and interfaces. Basically, these techniques can be categorized by the probe utilized, as follows: (1) scanning electron microscopy and transmission electron microscopy (TEM), which utilize an electron beam as the incident source with a wavelength shorter than 1 nm, can obtain fine structural information of the target sample. Recently, the fast development of environmental TEM9,10 and liquid or even electrochemical TEM11,12 has made it possible to visualize structural changes under the reaction conditions. Combined with electron energy loss spectroscopy, the morphological and chemical information of the target material with a spatial resolution better than 5 nm can be obtained.13,14 (2) Scanning probe microscopy (SPM) techniques such as scanning tunneling microscopy (STM) and atomic force microscopy (AFM), which employ a physical tip to raster scan the sample, can probe a variety of properties of the sample according to the specific interaction between the tip and surface, such as morphology, electronic and mechanical properties. Since the early stage, in situ SPM techniques have been developed to characterize materials and interfaces even at atomic resolution,15–18 enabling the visualization of dynamic structural information of materials and the molecular orientation of surface species under different environmental conditions.19,20 Recently, the development of noise-STM has made it possible to directly identify the active sites on the surface of Pt during the hydrogen evolution reaction (HER) by analyzing the noise of the tunneling current at different sites.21 In particular, the invention and rapid development of the qPlus sensor significantly improved the sensitivity of AFM,22 allowing the visualization of the chemical structure inside a molecule, intermolecular interactions and the fine structure of the monolayer water on a metal surface.23–26 (3) With the development of the synchrotron-based light source, X-ray-based techniques, such as X-ray diffraction, X-ray adsorption (XAS), X-ray scattering and X-ray photoelectron spectroscopy, have become a popular method in probing the crystal structure, valence state and local atomic structure of materials.12,27–33 In a recent reported work that combined XAS with STM (SX-STM), even the elemental and chemical states of just one atom were obtained based on the X-ray-excited electrons detected by STM, paving the way for using X-ray-based techniques to study materials at the single-atom limit.34 (4) Benefitting from their non-destructive analysis with abundant information, optical techniques are widely used for characterizing materials and their interfaces. However, the spatial resolution is limited by the Abbe diffraction limit to around hundreds of nanometers. Thus, different strategies such as scanning near-field optical microscopy35–38 and single-molecule fluorescence imaging (SMFI)39,40 have been developed to overcome this limitation. For example, SMFI enabled by super-localization was able to identify the active sites during chemical reactions and measure the kinetics of chemical reactions at different sites.41,42

Vibrational spectroscopy, including infrared spectroscopy, Raman spectroscopy, and sum-frequency spectroscopy, are the most commonly used methods to obtain chemical information. Among them, Raman spectroscopy is unique, because it can present vibrational modes across a wide frequency region. As a result, abundant information can be extracted from a Raman spectrum, including weak interactions (such as hydrogen bonds43–45 and interlayer interactions) in the ultralow-frequency region (below 100 cm−1), the lattice vibration,46–50 the molecule–metal interaction51–57 in the low frequency region (hundreds of cm−1) and the molecular vibrational modes in the high frequency region (above 1000 cm−1). It is worth mentioning that the interference from water can be neglected because of its small Raman cross section, making Raman spectroscopy a powerful tool for the in situ characterization of chemical and biological processes in aqueous samples. However, Raman spectroscopy suffers from ultralow sensitivity owing to the very small Raman scattering cross section, where approximately one Raman photon is produced with 106–1010 incident photons.58,59

The emergence of surface-enhanced Raman spectroscopy (SERS)60–62 in the 1970s dramatically increased the sensitivity of Raman spectroscopy up to single-molecules level,63,64 which significantly broadened its applications. The enhancement of SERS mainly originates from the surface plasmon resonance (SPR) effect. Briefly, SPR effect is the collective oscillation of free electrons in metal nanostructures under the excitation of incident light with a suitable wavelength.65 Consequently, the electromagnetic field on the surface of the metal nanostructures will be significantly enhanced, thus amplifying the Raman signal of the species located on or in the vicinity of the surface. After 50 years of development, SERS technique has become a routine analytical method in fields of materials science, biology,66–68 surface science,18,69–71 etc.72–74 In particular, the spatial resolution of SERS limited by diffraction was overcome by the development of the tip-enhanced Raman spectroscopy (TERS) technique in 2000.75–78 TERS is an apertureless near-field optical technique, combining SPM and SERS. In principle, a sharp tip with plasmonic activity is controlled by the SPM system to approach close to the vicinity of the target sample. Then, with the illumination of an incident laser with a suitable wavelength and polarization, a highly enhanced and localized electromagnetic field will be generated underneath the tip apex owing to SPR and lightning-rod effects. This enhanced and confined field serves as a nanoscale “amplifier” to enhance the Raman scattering of the sample beneath the tip. By raster scanning over the sample surface, TERS can simultaneously obtain morphological and the chemical information at nanometer scale spatial resolution with high sensitivity even down to the single-molecule level, providing a promising tool to comprehensively investigate materials79–82 and their surfaces.83–86

Recently, a review highlighted the new era of SERS and TERS, especially the understanding and application of angstrom-resolved TERS.74 Additionally, some excellent reviews have discussed the general advances of TERS in recent years.87–94 There are also reviews focusing on TERS tips,95–97 the characterization of bio-materials,98,99 ultrahigh-vacuum (UHV) TERS,100 application of TERS in catalysis,101,102 TERS in single-molecule chemistry103 and electrochemical (EC)-TERS.104–106 Alternatively, in this review, we mainly focus on the applications of TERS in the nanoscale chemical characterization of materials and interfaces. The way nano-resolved Raman spectroscopy helps to reveal the unique properties and activities of materials and interfaces is discussed in detail. In Section 2, we discuss the theoretical understanding and technical improvements of TERS in recent years, which are the basis to optimize the performance of the TERS technique, including the general interpretation of the enhancement and the spatial resolution of TERS, the development of TERS instruments, and the fabrication of TERS tips. In particular, we also note the applications of algorithms in TERS, which is rapidly growing. In Section 3, we present several experimental issues that need to be addressed to conduct successful and reliable TERS measurements. This seldomly discussed topic is believed to be helpful for beginners in TERS. In Sections 4 and 5, we present an overview of the progress in TERS for characterizing materials and interfaces. The applications of TERS in studying carbon-based materials, two-dimensional materials and other materials are discussed, while the investigation of solid–gas interfaces under either ambient condition or UHV condition and solid–liquid interfaces (especially the electrochemical interface) are included. In the final section, we propose the future developments of TERS to make it a more powerful technique for broader applications.

2. Theoretical understanding and technical improvements of TERS in recent years

After development for more than two decades, it is time to clarify the understanding of enhancement when TERS measurement is conducted on substrates with different dielectric properties. This forms the basis to design optical path and tips to achieve an optimized TERS performance. In addition to the hardware, recently the utilization of artificial intelligence (AI) has introduced a new opportunity to further improve the sensitivity of TERS.

2.1. General understanding of TERS

The enhancement in TERS originates from the lightning-rod effect and the SPR effect. The lightning-rod effect comes from the accumulation of free electrons at the sharp apex of the tips driven by an external electromagnetic field. Therefore, the lightning-rod effect is a non-resonant effect, which exists in all types of TERS tips. Incident light with polarization parallel to the tip shaft will increase the lightning-rod effect. In contrast, the SPR effect is a resonance effect, where the collective oscillation of conduction electrons leads to much more effective accumulation of free electrons at the tip apex, and thus a higher surface charge density than that of the lightning-rod effect. Therefore, to obtain a strong enhancement, the SPR effect needs to be efficiently excited. The dielectric function of metals will dramatically influence the efficiency of the SPR effect. In the visible range, where TERS generally works, coinage metals including Au, Ag and Cu are the best materials to support the SPR effect. In terms of the TERS configuration, there are different ways to excite the SPR effect, including gap plasmon, localized surface plasmon resonance (LSPR) and surface plasmon polariton (SPP).

When the substrate used in the TESR measurement is a coinage metal (so-called gap mode TERS) (Fig. 1a), a so-called gap plasmon will form from the electromagnetic coupling between the tip and the coinage metallic substrate with the illumination of light polarized along the tip shaft. Gap plasmons are the most efficient way to excite the SPR effect and exhibit a strongly enhanced electric field. Consequently, the Raman spectrum of even a single molecule can be obtained in gap mode TERS. In principle, both the tip and substrate made of coinage metals can support the generation of gap plasmons. Considering the good chemical stability of Au in air and liquid, the most widely used configuration in experiment is a Au tip coupled with a Au substrate. Gap plasmons are quite sensitive to the distance between the tip and the substrate (so-called gap distance), where the enhancement exponentially decreases with an increase in the gap distance.107,108 As a result, only ultra-thin materials can be studied in gap mode TERS. However, a too small gap distance, for example, the gap distance in the tunneling region may reduce the plasmon enhancement owing to the tunneling and nonlocal screening.109 In addition, the morphology of the TERS tip, including the radius and the cone angle of the tips as well as the illumination configuration will influence the efficiency of gap plasmons. For example, according to the simulation result of side illumination TERS under 633 nm light (Au tip coupled with Au substrate), the enhancement increases when the tip radius increases from 15 nm to 50 nm and does not grow with a further increase in the radius. An interesting result showed that the TERS enhancement will sharply increase when the tip radius decreases from 15 nm to 5 nm, which may originate from an enhancement in the lightning-rod effect.107 Besides the ultra-high enhancement, gap mode TERS possesses a spatial resolution that is much smaller than the radius of the TERS tip. It is well accepted that the lateral confinement (w) of the electromagnetic field in the gap is in the form of image file: d4cs00588k-t1.tif,110 where R is the radius of the tip and d is the gap distance. Consequently, TERS images with a spatial resolution better than 5 nm can be obtained under ambient conditions.83,111 The spatial resolution and the enhancement in gap mode TERS can be further improved by introducing some atomistic protrusions at the tip apex,112–114 which serve as picocavities.115 These picocavities can be stabilized at cryogenic temperatures to confine the electromagnetic field to the atomic scale,113,114,116 enabling chemical bond-resolved TERS imaging117,118 with a spatial resolution of up to 0.15 nm.118


image file: d4cs00588k-f1.tif
Fig. 1 Schematic of the origin of the TERS enhancement. (a) Gap plasmon. (b) Lightning-rod effect. (c) Localized surface plasmon resonance (LSPR). (d) Surface plasmon polariton (SPP). (e) Atomic point contact (APC).

When the substrate used in TERS is not a coinage metal such as dielectric materials and transition metals, efficient coupling between the tip and the substrate will not occur. Thus, only the TERS tip offers the full enhancement. This configuration is called non-gap mode TERS (conventional TERS). Here, it should be noted that for a semi-infinite metallic tip alone, as shown Fig. 1b, an optical near-field measurement near the tip apex within 1 nm based on rescattering of the photoemitted electrons showed that the field enhancement factor for tungsten and gold nanotips is governed by a geometric effect, i.e. the lighting-rod effect, while the SPR effect plays a minor role.89,97,119 However, the lighting-rod effect alone cannot produce sufficient enhancement to study most of the samples. Therefore, the key for effective non-gap mode TERS is to excite efficient SPR effect on the tip.

Given that LSPR can only be excited on metallic nanostructures with a sub-wavelength size, it is necessary to fabricate TERS tips with finite nanoscale features. As illustrated in Fig. 1c, this finite tip can be either isolated nanoparticles attached on an Si tip, groove-based nanotip or size-adjustable nanopyramid. In the case of the nanoparticle-based finite tip, the size and number of nanoparticles will influence the LSPR efficiency.120,121 Besides nanoparticles, the introduction of a groove with a suitable distance away from the tip apex can turn the semi-infinite metallic tip into a finite nanotip.121,122 The nanotip can serve as a dipole to support LSPR. As shown in Fig. 1c, the LSPR wavelength of the nanotip is dominated by the distance (L) between the groove and the tip apex. Besides, the cut length (LC) and the cut depth (D) will also influence the LSPR efficiency. The simulation suggests that the enhancement will first be improved with an increase in both Lc and D, and then reach a plateau when Lc is around 100 nm, while D is around 20 nm.121 The best parameter for the groove structure varies in different illumination configurations. Compared with the groove-based structure, the LSPR effect can be achieved in a more efficient way using a monopole antenna, where the flat plateau provides the image of the monopole, and thus the field pattern generated at the apex can be regarded as the overlap of half of a fictitious freestanding dipole with the other half from the mirror plane.123 The monopole antenna consists of a size-adjustable nanopyramid grounded on a flat surface. The so-called plasmon-tunable tip pyramid (PTTP) exhibits a stronger enhancement than the groove-based nanotip.123,124 By utilizing these specially designed tips to efficiently excite the LSPR effect, non-gap mode TERS can also produce a relatively high enhancement to characterize materials, especially two-dimensional materials with a spatial resolution of around 10 nm.123

It should be noted that because of the large penetration depth of light for dielectric materials, which produce strong far-field Raman signals, it is a great challenge to characterize the bulk materials employing non-gap mode TERS. In this case, an effective strategy is to remotely excite the LSPR of the tip apex with SPP (Fig. 1d). Owing to the large momentum mismatch between the incident light and the plasmon, SPP effect cannot be directly excited. Consequently, the core to excite SPP is the design of a suitable coupler (Fig. 1d), such as a grating,125,126 which converts the incident light into plasmons. During the propagation of plasmons to the apex of the tip, some loss will occur, which depends on the slope of the tip127,128 and the surface roughness.129 Therefore, the coupling efficiency and the loss during propagation need to be considered when designing SPP-based TERS tips.130 One of the advantages of SPP-based TERS is that the incident laser is not directly irradiated on the sample but on the coupler (so-called remote-excitation TERS), while the enhanced Raman signal near the tip apex is collected. This process produces a pure TERS signal without a far-field background.125 Technically, the setup for remote-excitation TERS will be more complicated than the conventional TERS setup, given that the excitation and collection paths need be separated, such as two separate objectives for excitation and collection, respectively,131 or noncolinear optical path to separate the excitation and collection with one objective.125,132,133

Besides the above-mentioned electromagnetic enhancement, there is also a chemical enhancement in TERS.134–137 Although the chemical enhancement is less discussed in TERS, it may dominate the TERS signal in some special situations. The Kumagai group observed a dramatic enhancement when a so-called atomic point contact (APC) was formed between the tip and the sample (Fig. 1e).138–140 This dramatic enhancement appeared over the ZnO layer but disappeared over the inert NaCl layer, indicating the sensitivity to the surface reactivity. According to this result, the enhancement was attributed to the hybridization between the sample and both the tip and substrate in the STM configuration, which made charge transfer enhancement operative through the metal-sample-metal system.138 It is surprising that this chemical enhancement could even enable the observation of the phonons on the Si(111)-7[thin space (1/6-em)]×[thin space (1/6-em)]7 surface over a bulk Si sample in the non-gap mode configuration.139 Recently, the Dong group investigated in detail the mechanism of the APC-based enhancement for molecules in well-defined single-molecule TERS configurations under UHV and ultralow temperature (LT) conditions. The resonance charge transfer mechanism, which is the resonance between the excitation wavelength and the molecule–metal charge transfer transition was ruled out according to the slightly changed TERS signals in the bias voltage-dependent experiment. Then, a combined physicochemical mechanism was proposed, instead of a pure chemical effect to explain the enhancement during the APC. It was found that the weak interaction between the tip and the molecule will enable ground-state charge transfer (GSCT), which will induce vertical Raman polarizability. Subsequently, the vertical Raman polarizability will be further enhanced by the vertical electromagnetic field in the plasmonic nanogap. This GSCT mechanism was also corroborated in wavelength-dependent TERS experiments under different resonance conditions.141

2.2. TERS instruments

From an instrumental viewpoint, a TERS setup generally consists of a Raman microscopy coupled with an SPM system, while a control module is required for communication between the two parts. STM, AFM and shear force microscopy (SFM) have been successfully used in TERS. Among them, STM-based TERS can provide Raman signals with a high signal-to-noise ratio and high-quality morphological images. However, the conductivity of the sample should be sufficiently high to establish the tunneling current-based feedback. Alternatively, working with the atomic force as feedback, AFM- and SFM-based TERS can be used to study either conductive samples, semi-conductive samples or insulated samples. It is important to note that all the SPM techniques used in TERS are ultra-sensitive to the vibrational noise transmitted from the working environment. A large noise makes it difficult to keep the TERS tip in close vicinity of the sample to produce a stable enhancement. In this regard, one of the biggest challenges in the TERS technique is how to realize a high optical excitation/collection efficiency in the limited space of SPM without introducing too much noise. Accordingly, four optical configurations have been employed in TERS setups, including bottom illumination, side illumination, top illumination and parabolic mirror-based illumination.

In the bottom illumination configuration (Fig. 2a), the optical part is mounted at the opposite side of the TERS tip, which makes it possible to use an objective with a high numerical aperture (NA) and a short working distance, such as an oil-immersion objective (NA[thin space (1/6-em)]=[thin space (1/6-em)]1.4) and water-immersion objective (NA[thin space (1/6-em)]=[thin space (1/6-em)]1.0), to guarantee both high excitation and collection efficiencies. The further introduction of radial polarization can efficiently excite the SPR effect to produce a higher TERS enhancement.142 However, given that the laser and the back-scattered Raman signals need to pass through the sample, only transparent samples and substrates can be studied in the bottom illumination configuration.


image file: d4cs00588k-f2.tif
Fig. 2 Schematic illustration of the optical configurations of TERS. (a) Bottom illumination. (b) Side illumination. (c) Top illumination. (d) Parabolic mirror-based illumination.

In the side illumination configuration (Fig. 2b), a tilted mounted objective is used for the excitation and collection. Limited by the open space of SPM, the working distance of the objective used in side illumination should be long enough. The recent available objective with both a long working distance (6 mm) and a high NA (0.7) can help to improve both the excitation and collection efficiency. The p-polarized laser is used to make the polarization of the incident light parallel to the tip shaft to produce a higher TERS enhancement. The tilted angle should be carefully designed given that it will not only influence the component of the polarized laser parallel to the tip shaft but also determine the amount of Raman signals to be blocked by the sample, especially when an objective with a high NA is employed. Side illumination TERS can be employed to study both transparent and opaque samples. However, the quality of the optical image is deteriorated, which makes it difficult to coarsely mount the tip in the interested region on the sample. Another objective mounted in the top illumination configuration is now available in some commercial AFM-based TERS instruments and quite helpful for obtaining optical images.

Similar to side illumination, top illumination can also be applied to study both transparent and opaque samples. In addition, the top illumination configuration provides both a high collection efficiency and high optical image quality. As shown in Fig. 2c, the objective is mounted on the top side of the tip to focus the laser on the tip apex and collect the back-scattered Raman signals. As a result, the TERS tips used in this configuration need to be tilted to maintain a high optical efficiency. However, part of the incident laser and the Raman signal are still blocked by the tip, which leads to a loss in both the excitation and collection efficiency. In addition, the image quality of SPM, especially STM may decrease with the use of tilted tips.

A very special configuration based on parabolic mirror-based illumination (Fig. 2d) is also used in TERS, where a hole is drilled at the center of the parabolic mirror. This configuration possesses a high NA of up to 1.0, producing both a high excitation and collection efficiency. It was demonstrated that the introduction of radial polarization in this configuration will also improve the excitation efficiency,143,144 thus improving the TERS enhancement. It should be noted that very accurate alignment between the tip and the parabolic mirror is required in this configuration.

With the development of the above-mentioned basic configurations, various progress in instrumentation has been realized in recent years, including the improvement of sensitivity and the strict control of the working environment.

2.2.1. Strategies to improve the sensitivity of TERS. The sensitivity of the TERS technique is always the core. Experimentally, the TERS sensitivity depends on the accurate alignment between the laser and the tip, e.g. a misalignment of 1 μm will lead to a dramatic loss in the TERS enhancement.145 However, if the light path is too long, there may be drift of the incident laser in both the vertical direction (z direction, the focal plane of the laser) and horizontal directions (xy directions, the relative position of the laser and the tip) when the acquisition time is long (e.g. TERS imaging). Thus, to tackle this issue, the Verma group introduced a guide laser together with a positional sensor and a galvanometer mirror in the TERS instrument.146 The guide laser was focused on the sample, and then reflected back to the positional sensor, where feedback was set between the positional sensor and the piezo scanner-controlled objective (Fig. 3a). Then, the focus shift could be observed from the positional sensor and compensated by moving the objective in the z direction. The drift in the xy directions could be compensated by performing high-speed laser scanning and finding the tip position from the scattering image (Fig. 3b). Then, the laser moved automatically to the tip apex. This improvement in optical stability makes it possible to perform TERS imaging of a large area for more than 6 h, which is beneficial to reveal the various randomly distributed defects on two-dimensional materials.
image file: d4cs00588k-f3.tif
Fig. 3 Strategies to improve the sensitivity of TERS. (a) Schematic illustration of the z direction compensation system based on a guide laser. (b) Schematic illustration of the xy direction compensation system based on high-speed laser scanning. Reprinted with permission from ref. 146. Copyright 2022, American Association for the Advancement of Science. (c) Schematic illustration of the adaptive-TERS (a-TERS) setup. (d) Evolution of the TEPL signal during the wavefront shaping with a stepwise sequential algorithm. Reprinted with permission from ref. 147. Copyright 2021, Springer Nature.

In addition, the quality of the excitation beam will influence the sensitivity of TERS. The Domke group148 and the Park group147 introduced a spatial light modulator (SLM) in the TERS light path to manipulate the excitation beam. The quality of the excitation beam, including its focusing, spatial coherence and polarization would be improved after wavefront modulation, and thus the beam could be better coupled with the tip to produce a higher SPR efficiency. In their experiment, the Park group developed an adaptive optics algorithm-based TERS system (Fig. 3c). In this system, a stepwise sequence algorithm was established via the feedback between the TERS (or tip-enhanced photoluminescence, TEPL) intensity and the wavefront shaping (the phase mask of the SLM). Then, this algorithm started with a random phase mask and ran automatically and adaptively until the highest TERS (or TEPL) intensity was obtained (Fig. 3d). The TERS intensity could be increased by ∼2.5 times through this adaptive-TERS (a-TERS) compared with the normal TERS intensity.

2.2.1. Strategies to strictly control the working environment of TERS. Given that the properties of materials and their interfaces are dramatically influenced by their working environment and evolve with reactions, it is significant to develop TERS instruments with a controlled environment. In this case, UHV provides an ideally controlled and clean environment. Further combined with LT, the stability of the TERS system can be obviously improved with an almost negligible thermal drift down to 50 pm h−1 at around 7 K.118 To realize UHV–TERS, the optical system needs to be integrated with an UHV-SPM system, either inside or outside the UHV chamber. In the first demonstration of UHV–TERS by the Pettinger group in 2007, most of the Raman optics were mounted on an optical platform inside the UHV chamber.149 To make the optics more flexible, the Dong group placed an aspheric lens inside the UHV chamber for the excitation and collection, while mounting the other optics outside of the UHV chamber.150 Alternatively, the Jiang group and the Kim group mounted two aspheric lenses inside the chamber for excitation and collection, respectively, while leaving the rest of the optics outside the chamber.151,152 With these configurations, it is easy to change the wavelength of the incident laser by just changing the long-pass filter in the collection path. For these configurations, attention should be paid to avoid the interference of the vibration from the piezo, where the aspheric lens is mounted, to the stability of the SPM during aligning the laser and the tip. To fully avoid interference from the Raman optics, the Van Duyne group mounted all the optics outside of the UHV chamber with two aspheric lens for excitation and collection, respectively.153 This ultra-stable and clean environment from UHV–TERS makes it possible to characterize the materials79,86 and their interfaces154 with a sub-nanometer resolution.

Different from the ultra-clean environment of UHV, a glovebox provides a working environment without oxygen and water, which makes it possible to study samples that are unstable in air. A TERS setup in a glovebox was achieved by placing the SPM part of the TERS instrument inside the glovebox, while leaving the optical part outside of the glovebox.155,156 This controlled environment helped to reveal the nanoscale heterogeneity of the solid electrolyte interphase (SEI) in either lithium ion batteries155 or sodium ion batteries.156

Given that an aqueous solution can isolate materials/interfaces from oxygen in air, thus improving the chemical stability of surface species, the liquid condition, especially the electrochemical condition can also provide a controllable environment. In particular, the surface state, charge state and molecular state can be well controlled by the applied potential in the electrochemical condition. However, extending TERS from the air to electrochemical condition is quite challenging given that EC–TERS is a complex combination of SPM, electrochemistry and SERS. The main challenge is that when light passes from the air to the electrolyte (or from the electrolyte to the air), it will be severely distorted owing to the mismatch of the refractive indices between different media. This distortion will lead to the severe loss of the Raman signals, which significantly decreases the effective NA.157 In addition, the distortion leads to poor focus of the incident laser, making the accurate alignment of the laser and the tip apex difficult. Therefore, the core in the development of EC–TERS is the optimization of the optical path to minimize the light distortion. The first side illumination EC–TERS system utilized a horizontal illumination configuration,158 where the influence of the curved liquid surface in the optical path can be avoided. The thickness of the liquid layer is controlled to be around 1 mm to reduce the optical distortion. Thereafter, side illumination EC–TERS systems using an air objective were developed by different groups.159–161 An SLM was adapted in the side illumination EC–TERS setup by the Domke group to modulate the wavefront of the incident light to minimize the optical distortion.162 However, optical distortion still existed in these configurations.

To fully solve this problem, a water-immersion objective-based side illumination EC–TERS setup was developed recently by our group.163 In this new configuration, a droplet of water was added between the objective and the optical window to replace the air (Fig. 4a). As a result, the mismatch among air, the optical window and electrolyte could be eliminated, providing an increased sensitivity by around six times. Recently, with the use of a water-immersion objective, the Maisonhaute group developed the STM-based top illumination EC–TERS setup (Fig. 4b),145 and later our group developed the AFM-based top illumination EC–TERS setup (Fig. 4c).164 The water-immersion objective was directly immersed into the electrolyte to avoid the optical distortion in the top illumination configuration. It should be noted that a thin optically transparent film was covered on the objective to prevent the possible etching and contamination from the electrolyte. Alternatively, for a transparent sample, it is more convenient to use the bottom illumination configuration (Fig. 4d). In this way, the light will not pass from the air to liquid (or from liquid to air), fully avoiding light distortion.165


image file: d4cs00588k-f4.tif
Fig. 4 Schematic illustration of optical configurations of EC–TERS without light distortion. (a) STM-based side illumination. (b) STM-based top illumination. (c) AFM-based top illumination. (d) AFM-based bottom illumination.

2.3. TERS tips

The TERS tip serves as the probe to characterize the morphology of the sample as well as the electromagnetic field amplifier to enhance the Raman signal, which is the key to determine the performance of TERS measurements. In the early stage of TERS, the reproducible fabrication of highly active TERS tips was one of the bottlenecks of this technique. After around 20 years, many methods have been developed to fabricate TERS tips,95–97,166 and commercial TERS tips have become available recently.167 The improvement in the fabrication of TERS tips extends the applications of TERS and makes it possible for TERS to become a routine analytical technique. In addition, a variety of special TERS tips was designed and fabricated to modulate the TERS performance for specific samples.122,123,131,133,168 In the following parts, we discuss the fabrication of TERS tips for different purposes, including gap mode TERS, non-gap mode TERS and remote-excitation TERS.
2.3.1. Tips for gap mode TERS. As mentioned before, when the target materials or the species are “sandwiched” between a coinage metallic substrate and the TERS tip, the lightning-rod effect and the gap plasmon produce an enhanced electromagnetic field to amplify the Raman signals. Thus, the morphology, including the radius and the cone angle of the tip, should be considered for efficient gap mode TERS. From an experimental viewpoint, different methods have been developed to fabricate tips for gap mode TERS in both STM and AFM-based setups.

(1) Electrochemical etching of metal wire based on the drop-off method76,78,169 is the most frequently used method to produce pure metallic tips, including Au and Ag tips, for STM170,171 and SFM-based TERS.172 In the drop-off method, a sharp tip is obtained when the portion of the wire immersed in the solution drops off as its weight exceeds the tensile strength of the etched region of the wire at the air/electrolyte interface.173 The core of this method is the selection of suitable etchants and suitable applied voltage (DC voltage or AC voltage). In addition, to avoid over-etching, a suitable cut-off threshold needs to be used during etching. Using a mixed solution of fuming HCl and anhydrous ethanol (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) as the etchant and applying a constant potential, our group developed the most widely used electrochemical etching method to fabricate Au STM (SFM)–TERS tips,170 showing the best reproducibility to date. With a defined etchant, the potential applied during the etching process will influence both the radius and the cone angle of the Au tips.171 Au tips with a radius of around 10–30 nm can be obtained by applying a DC voltage of 2.2–2.4 V (Fig. 5a). Recently, a KCl solution was used as the etchant to produce Au STM–TERS tips, which could avoid the use of corrosive fuming HCl (Fig. 5b).174 Compared with Au tips, Ag tips produce a higher TERS enhancement. Various etchants have been developed to fabricate Ag tips, where ammonia solution,175 mixed solution of perchloric acid and methanol176,177 or ethanol178 are the most frequently used etchants. Unfortunately, the reproducibility of the fabrication of Ag tips is not as high as that of Au tips. Usually, the Ag tips obtained by electrochemical etching suffer from a rough surface and uncontrollable radius. Thus, to solve this problem, some post-smoothing and sharpening processes were conducted on electrochemically etched Ag tips. For example, the Ag tip fabricated by first electrochemical etching, and then focused ion beam (FIB) milling possessed a controllable morphology (Fig. 5c).179 Besides FIB milling, Ar ion sputtering followed by annealing through electron bombarding180 and field-directed sputter sharpening181 methods was conducted to produce sharp and smooth Ag tips for UHV–TERS.


image file: d4cs00588k-f5.tif
Fig. 5 TERS tips for gap mode TERS. (a) Au STM–TERS tip obtained by electrochemical etching in fuming HCl and anhydrous ethanol solution. Reprinted with permission from ref. 171. Copyright 2007, AIP Publishing. (b) Au STM–TERS tip obtained by electrochemical etching in KCl solution. Reprinted with permission from ref. 174. Copyright 2018, the American Chemical Society. (c) Ag STM–TERS tip fabricated by electrochemical etching and FIB milling. Reprinted with permission from ref. 179. Copyright 2019, the American Chemical Society. (d) Au-coated AFM–TERS tip prepared by vacuum deposition, insert: Si AFM tip. Reprinted with permission from ref. 167. Copyright 2021, the American Chemical Society. (e) Au-coated AFM–TERS tip fabricated by electrochemical deposition. Reproduced from ref. 182 with permission from The Royal Society of Chemistry. (f) Au pyramid-based TERS tip fabricated by template-stripping. Reprinted with permission from ref. 183. Copyright 2012, the American Chemical Society. (g) Ag-nanowire-based AFM–TERS tip. Reprinted with permission from ref. 184. Copyright 2023, the American Chemical Society.

(2) Another type of commonly used TERS tips for gap mode TERS is metal-coated tips, which are usually used in AFM–TERS. This type of TERS tip is fabricated by depositing a layer of Au or Ag on commercial Si or Si3N4 AFM tips. The most prevalent method to fabricate metal-coated AFM–TERS tips, including the commercial AFM–TERS tips, is vacuum evaporation,167,185–187 where the morphology of the tips can be controlled by the temperature of the substrate, the evaporation rate, etc. Fig. 5d shows a typical SEM image of an Au-coated AFM–TERS tip fabricated by vacuum evaporation, where the insert shows the morphology of a commercial Si AFM tip. One of the advantages of the vacuum evaporation method is the possibility to deposit a metallic layer on a specific side of Si AFM tips by controlling the position or angle of the tip.188 Besides vacuum deposition, our group developed an electrochemical deposition method to fabricate either Au or Ag-coated AFM–TERS tips within minutes in normal chemical labs.182,189 The radius of the AFM–TERS tips could be reproducibly controlled by the applied potential and the deposition time (Fig. 5e). In particular, the electrochemically deposited metal film showed a smooth surface and had a strong adhesive strength to the Si AFM tip, which could improve the chemical stability, and thus the lifetime of the as-prepared Ag tips. Recently, a template-stripping-based method was developed to fabricate Au AFM–TERS tips (Fig. 5f).183 In this method, high-quality inverted pyramidal molds in silicon were produced by photolithography, followed by anisotropic etching. Then, after the deposition of Au, tips with a controlled morphology and radius could be obtained by template-stripping from these Au-coated Si molds. This method makes it possible to batch-scale fabricate TERS tips, given that over one million tips with a similar radius and cone angle can be obtained in a four in. wafer.

It is worth mentioning that owing to their relatively large radius, the spatial resolution in morphological imaging using metal-coated AFM–TERS tips is lower than that using sharp Si AFM tips. In this regard, AFM–TERS tips with a chemical-synthesized Ag nanowire attached to commercial Si AFM tips (Fig. 5g) were fabricated.190 These tips even showed a higher spatial resolution in morphological imaging than the Si AFM tips. The mechanical stability of the Ag nanowire-based TERS tips could be improved by controlling the length of the Ag nanowire.184 The lifetime could be extended by coating an Au layer on the Ag nanowire.191

2.3.2. Tips for non-gap mode TERS. When the target materials or the species are not loaded on a coinage metallic substrate, the gap plasmon cannot be efficiently excited.107,108 Given that the lightning-rod effect alone cannot produce sufficient enhancement for TERS, the size of the TERS tips needs to become finite to match the incident laser to excite the LSPR effect to offer sufficient enhancement. A straightforward way is to deposit isolated Ag nanoparticles on Si AFM tips (Fig. 6a).120,186,192 It was demonstrated in both simulation and experiment that isolated nanoparticle-coated AFM–TERS tips produced a higher TERS enhancement than the continuous film-coated tips in non-gap mode TERS. The enhancement in this type of TERS tips could be further improved by oxidizing Si into SiO2, which modified the reflective index for control of the plasmon resonance wavelength.186
image file: d4cs00588k-f6.tif
Fig. 6 TERS tips for non-gap mode TERS and remote-excitation TERS. (a) Isolated Ag-NP-based AFM–TERS tip. Reproduced from ref. 120 with permission from The Royal Society of Chemistry. (b) Groove-based SFM-TERS tip fabricated by electrochemical etching and FIB. Reprinted with permission from ref. 122. Copyright 2015, the American Chemical Society. (c) Plasmon-tunable tip pyramid (PTTP). Reprinted with permission from ref. 123. Copyright 2018, John Wiley and Sons. (d) Au-TERS tip with a grating on the tip shaft. Reprinted with permission from ref. 125. Copyright 2010, the American Chemical Society. (e) Remote-excitation AFM–TERS tip based on Kretschmann-Raether prism configuration. Reprinted with permission from ref. 131. Copyright 2021, the American Chemical Society. (f) Nanowire-assisted selective-coupling STM–TERS tip. Reprinted with permission from ref. 130. Copyright 2019, Springer Nature.

The development of advanced micro- and nano-processing techniques makes it possible to accurately fabricate metal structures and control the LSPR wavelength. A groove structure could be introduced on either the Si AFM tip (Ag is deposited after FIB milling)168 or the metallic Au tip (Fig. 6b),122 which turned the semi-infinite tip into a finite tip to excite the LSPR effect. The LSPR wavelength could be accurately controlled by the distance between the groove and the apex of the tip. However, FIB milling may damage the Au layer during the nanofabrication, reducing the efficiency of the LSPR.122 Thus, to avoid this damage, PTTP was developed via some modifications of the above-mentioned template-stripping method. In the fabrication of PTTP, a two-step lithography process produces an Si mold with an inverted micropyramidal body and a nanopyramidal end. After the deposition of Au, TERS tips could be obtained by the template-stripping process (Fig. 6c).123,193 The core of PTTP is the size-adjustable nanopyramid, where the LSPR wavelength can be accurately controlled by the length of the nanopyramid to provide a highly efficient enhancement.124

2.3.3. Tips for remote-excitation TERS. When a thick material is studied in TERS (either in gap mode TERS or non-gap mode TERS), it is better to use remote-excitation TERS to minimize the interference from the far-field background Raman signal. Various specially designed tips have been fabricated for remote-excitation TERS to date.166 The pioneering experimental work was achieved by the Raschke group. A grating was introduced on the shaft of a full metallic Au tip, which served as the coupler to excite the SPP effect (Fig. 6d). Given that the incident laser illuminated at the grating structure far away from the tip apex, while the Raman signal was collected near the tip apex, a TERS signal without far-field background could be obtained.125 The grating structure could also be introduced onto the metal-coated AFM–TERS tip to realize remote-excitation.126 Afterwards, other couplers such as a nano-slit,194 Ag nanocube on Ag nanowire132 and two parallel assembled Ag nanowires133 were designed to achieve remote-excitation TERS. Recently, Zhang et al. designed and fabricated another remote-excitation TERS tip based on the Kretschmann-Raether prism configuration, where FIB milling was used to remove part of the AFM–TERS tip to produce a flat Si/Au interface (insert of Fig. 6e).131 When the Si part is illuminated by the incident light with a suitable incident angle, SPP will be generated at the interface between the Au layer and air, and then propagated to the tip apex to enhance the Raman signal (Fig. 6f). To achieve a higher coupling efficiency, a nanowire-assisted selective-coupling tip was carefully designed.130 This type of tip utilized a tapered fiber as a coupler to couple the incident light to a nanowire (Fig. 6f), while the Raman signals could be coupled from the nanowire back to the optical fiber. Therefore, both a high excitation efficiency and collection efficiency were achieved (72% for 633 nm excitation). This configuration also made it possible to achieve a lens-free TERS technique with a spatial resolution of 1 nm.

2.4. AI-assisted TERS

Given that TERS only enhances the Raman signals of the samples underneath the tip apex, only a small amount of samples has been studied via TERS, which requires a strong enhancement to produce a good signal-to-noise ratio (SNR). However, simply extending the acquisition time cannot improve SNR in TERS because a long acquisition time will enlarge the influence of the thermal drift, especially in TERS imaging.

Thus, to solve this problem and speed up the TERS imaging, some denoising algorithms have been developed and applied in TERS. The Wang group developed both the adaptive low-rank matrix approximation (ALRMA)195 and collaborative low-rank matrix approximation (CLRMA)196 methods to improve the SNR in TERS. In ALRMA, the TERS imaging data were decomposed into a variety of submatrices by singular value decomposition (SVD). According to the contribution to the SNR, the submatrices were divided into a positive group and negative group. Then, the denoised data were reconstructed from the positive group after discarding the negative group reflecting the noise. However, it was found that it is difficult for ALRMA to distinguish the signal and noise when the SNR is relatively low. Thus, to improve the performance in extracting weak signals, high-SNR reference data was introduced to perform collaborative matrix factorization in CLRMA. The introduction of high-SNR reference data forces the components of factorization to approximate the true signal components or noise components, and thus a better denoising performance was obtained by CLRMA than ALRMA (Fig. 7a). Unfortunately, it may be difficult to obtain the high-SNR reference data in TERS. In addition, both ALRMA and CLRMA depend on a relatively large data volume, making them unavailable to denoise a single TERS spectrum. In this regard, our group developed deep learning-based algorithms for denoising TERS signals.197 An AUnet neural network was pre-trained to learn the noise characteristics of instruments, thereby predicting and subtracting the instrument noise from the raw spectrum with a low SNR. As a result, a denoised spectrum with a high-SNR will be obtained (Fig. 7b). This denoising method has the advantage of being independent of the data volume, making it suitable for both a single TERS spectrum and TERS imaging. With the dramatically increased SNR, it is possible to realize fast-speed TERS imaging. Fig. 7c shows the TERS imaging obtained by the raw data (fast-speed TERS imaging, 0.1 s pixel−1), AUnet-treated data (fast-speed TERS imaging, 0.1 s pixel−1) and long-time acquisition data (1 s pixel−1) from the same region in the STM image reconstructed from the peak position of the NC triple bond of phenyl isocyanide (PIC) molecules adsorbed on a Pd/Au surface. The TERS image obtained with the data treated by AUnet fits better with the topography compared with the raw data owing to the improved SNR. Furthermore, the influence of the thermal drift was significantly reduced compared with the long-time acquisition TERS image.


image file: d4cs00588k-f7.tif
Fig. 7 AI-assisted TERS. (a) Comparison of the TERS spectra and the TERS images reconstructed from noisy data, adaptive low-rank matrix approximation (ALRMA)-treated data and collaborative low-rank matrix approximation (CLRAM)-treated data. Reprinted with permission from ref. 196. Copyright 2021, the American Chemical Society. (b) Pipeline of the AUnet-based denoising method. (c) STM image and the corresponding TERS images obtained from raw data, AUnet-treated data and long-time-acquisition data. Reprinted with permission from ref. 197. Copyright 2024, Springer Nature. (d) Workflow of the study the a-Si by TERS imaging and the multi-resolution manifold (MML) algorithm-based analysis. (e) Mean TERS spectra of the child clusters. Reprinted with permission from ref. 198. Copyright 2021, Springer Nature.

Similar to other hyperspectral techniques, two-dimensional (2D) TERS imaging usually contains a large number of spectra, which makes it nearly impossible to manually analyze each spectrum. In this regard, some algorithms are introduced in TERS to help analyze the hyperspectral TERS data. The Lagugne-Labarthet group applied an artificial neural network (ANN) algorithm to distinguish different types of spectra in TERS images. The trained ANN models could identify the location of carbon nanotubes (CNTs) and distinguish different types of spectra at each pixel, enabling the analysis of the properties of different CNTs.199 The Nanda group used a multi-resolution manifold (MML) learning algorithm to classify the TERS spectra of amorphous silicon (a-Si).198 The high-dimensional TERS spectra (around 2500 spectra) obtained from the 2D TERS image were first projected to a low-dimensional (i.e. 2D) manifold space by the MML algorithm (Fig. 7d). Then, straightforward data visualizations and structural categorization were performed to reveal the underlying properties of a-Si. As a result, the 2500 TERS spectra of a-Si were first categorized into seven clusters according to their peak intensity and peak position, which determined the degree of Si–Si bond distortion in various regions of a-Si. Further dividing these seven clusters at a finer resolution grid enabled the discovery of a small region with a characteristic peak at 2435 cm−1 (Fig. 7e), which may come from trace amounts of defects (≤0.3%) in a-Si. It was demonstrated that with the help of the AI algorithms, analysis of the TERS data can be more efficient and precise. Some subtle features that are difficult for human to notice can be revealed using AI algorithms to assist in the analysis of TERS imaging data.

3. Implementation for TERS in studying materials and interfaces

After more than 20 years of development, TERS technique has been successfully applied in various fields, such as the characterization of materials and chemical reactions. Recently, commercial TERS instruments and TERS tips have become available, making it possible for more researchers to use the TERS technique. However, there are still a lot of challenges in conducting successful and reliable TERS measurements. As illustrated in Fig. 8, we discuss some experimental issues for TERS in studying materials and interfaces in the following parts.
image file: d4cs00588k-f8.tif
Fig. 8 Procedures for performing convincing TERS measurements.

3.1. Evaluation before experiment

Careful evaluation should be carried out for targeting samples before performing TERS measurements. Here, we summarize some basic principles that can be applied, as follows: (1) the first thing to consider is which mode of TERS is appropriate for the targeted samples, i.e. gap mode TERS or non-gap mode (conventional) TERS. Gap mode TERS possesses a higher enhancement and a higher spatial resolution. However, the interaction between the sample and the coinage metallic substrate (Au, Ag or Cu) may change the intrinsic property of the sample. In this case, a possible solution is to introduce an ultra-thin isolation layer, such as an ultra-thin h-BN layer, in ambient TERS or NaCl with atomic thickness underneath the sample in UHV–TERS to prevent the influence from the metallic substrate but enable the generation of highly efficient gap plasmons.141 Besides this, the sample studied in gap mode TERS should be thin enough (thinner than 5 nm) to maintain a short distance between the tip and the substrate for achieving efficient gap plasmons. Non-gap mode TERS provides a platform to study the pristine property of the sample (without interference from the metallic substrate) at the nanoscale, but specially designed tips with either high LSPR activity or SPP activity are required. (2) Then, a suitable SPM technique should be selected according to the property of the sample. STM-based TERS can only be used to study conductive samples with a relatively low roughness, while usually a relatively higher spatial resolution can be obtained. SFM- and AFM-based TERS can be employed to study both conductive and insulated samples. The development of modern AFM makes AFM-based TERS more powerful, where abundant information beyond morphology and vibrational spectroscopy, such as electrical and mechanical information, can also be obtained. (3) Finally, the working environment should be selected according to the research interests. UHV, liquid and electrochemical TERS provide a much more controllable environment than the ambient TERS but require more sophisticated instrumentation.

3.2. Preparations for TERS

Considering the above-mentioned evaluations, experimental preparations for different specific conditions are required. The sample needs to be well prepared on suitable substrates before the TERS measurements. In the case of gap mode TERS, the most frequently used substrate in ambient or liquid condition is an Au substrate. The Au substrate should be smooth enough for SPM to work properly and avoid the SERS enhancement from the substrate. Among the Au substrates, Au(111)200 or Au(111) on mica provides a clean and atomically smooth surface, which can be regarded as a model substrate for TERS. Another frequently used Au substrate is atomically smooth Au films fabricated by the template-stripping method,201 if there is no need to use a single crystal surface. Besides Au substrates, Ag114,138,152 and Cu117,202 substrates are also used in UHV–TERS, where the oxidation of the substrate can be avoided. Under the UHV condition, an in situ cleaning process is required to produce a clean surface. In the case of non-gap mode TERS, either a glass slide or Si with a relatively low roughness can be used as the substrate, depending on the illumination configuration.

Then, suitable methods should be selected to prepare the samples on the substrate for TERS measurements. For example, immersion, drop-casting and spin coating can be used to prepare self-assembled monolayer (SAM) samples on Au substrates under ambient conditions. It is important to optimize the procedures for assembling different molecules, including the concentration and the immersion time, which determines the molecular state on the surface.203 In the case of 2D materials, mechanical exfoliation, which produces 2D materials with high-quality, followed by dry transfer is still the most frequently used method for preparing samples for TERS. In UHV–TERS, the sample is usually prepared inside a UHV chamber, either via vacuum deposition,117,150,204,205 dry contact transfer (DCT)206,207 or epitaxial growth79,86 can be used depending on the sample. The coverage of the sample can be well controlled by optimizing the experimental details.

With well prepared samples, the next step will be the fabrication of suitable TERS tips. As discussed before in this review, different methods have been developed to fabricate STM-, AFM- and SFM-based TERS tips for either gap mode or non-gap mode TERS. Special pretreatment, such as insulation and coating with a protection layer, is necessary before conducting liquid TERS, especially EC–TERS, which has been discussed in detail in the literature.106 However, although the fabrication of TERS tips has become mature, different TERS tips may still show varying TERS performances. In this regard, screening tips in terms of enhancement before measuring the target sample will be helpful for successful TERS measurements. When evaluating the enhancement of TERS tips, a highly uniform standard sample is required to confirm that the differences in TERS intensity obtained with different tips originate from the varying enhancement of the tip instead of the sample. In experiments, SAMs on Au(111) will be a good choice for gap mode TERS, while some flat 2D materials assembled on a glass or Si substrate will be a good choice for non-gap mode TERS. It should be note that some defect sites of 2D materials may present stronger Raman signals than that of the basal plane,81,146 which should be careful when performing evaluation experiments. In addition, given that a Stokes shift exists between the Raman scattering and incident laser, TERS tips with different SPR resonance wavelengths may produce varying enhancements for peaks with different Raman shifts.208–210 Therefore, it will be much better to choose a standard sample that has Raman peaks close to that of the target samples. Besides the enhancement, spatial resolution is another important parameter of TERS. Currently, the spatial resolution of TERS can be evaluated by either the full width at half maximum (FWHM) or the 10–90% contrast of the spatial distribution of the intensity of the individual Raman peaks in TERS images.83,111,117,118 Considering that the spatial profile of the experimentally measured intensity is a convolution of the TERS spatial resolution and the distribution of the sample, it is straightforward and convincing to utilize the signal of one-dimensional materials such as carbon nanotubes or a sharp edge to evaluate the spatial resolution.211

3.3. Experimental keys for convincing TERS

3.3.1. Tip contamination. In principle, to achieve a high enhancement, the TERS tip needs to be very close to the substrate during measurement. As a result, the targeted sample, especially surface molecular species may adsorb on the TERS tip to produce a Raman signal, leading to so-called “tip contamination”. Once the tip is contaminated, it is difficult to determine whether the variation in Raman signals comes from the different behaviors of the target sample or from the samples adsorbed on the TERS tip. Thus, to make the TERS result convincing, contamination-free tests should be performed on a clean Au substrate after the TERS measurement, and if the Raman signal of the targets appears, it indicates tip contamination.
3.3.2. Sample degradation. When acquiring TERS spectra or imaging, the tip–sample interaction, thermal effect of the laser, plasmonic thermal effect, etc. may lead to sample degradation, especially in the case of some soft materials.212 Therefore, the morphology before and after TERS measurements should be recorded to check whether the sample is destroyed. Alternatively, time-series TERS spectral acquisition can be conducted to monitor the variation in samples during measurement. Reducing the tip–sample interaction by using TERS tips with a smaller force constant and decreasing the setpoint value (AFM-based TERS), decreasing the laser power and shortening the acquisition time may help avoid sample degradation. Here, we also emphasize that a controlled working environment that can keep samples away from the oxygen and water in air is important to improve the stability of the sample.158
3.3.3. Thermal drift. In TERS measurements at room temperature (either under air condition or liquid condition), thermal drift always exists, which leads to a mismatch between the TERS image and the morphology, given that it always takes a much longer time to perform TERS imaging than morphology imaging. Thus, to make the TERS result more convincing, the morphology of the same location before and after TERS imaging should be recorded to evaluate the thermal drift during TERS imaging. The sample needs to be well fixed on the stage to reduce the influence from thermal drift. In addition, usually a stabilization time is required to release the mechanical strain to minimize the thermal drift before the TERS imaging. Alternatively, fast TERS imaging, which can reduce the total time required for the measurement, can also alleviate the influence of thermal drift. However, speeding up the TERS imaging means reducing the acquisition time at each pixel, which deteriorates the SNR. Thus, a balance needs to be achieved between the SNR and the thermal drift.

4. Nanoscale chemical characterization of materials by TERS

4.1. TERS study of CNTs and carbyne

As one-dimensional carbon-based materials, CNTs have been widely used in electronic devices and industry.213–216 Given that the electronic and optical properties of CNTs are intimately related to their structure,217 characterization of CNTs at the nanoscale is the core to understanding their intrinsic structure–property relationship. Owing to their one-dimensional nature with negligible far-field background, CNTs are ideal samples for TERS. In 2003, the Novotny group first characterized single-wall CNTs (SWCNTs) by TERS,218 observing the variation inside a single SWCNT with a spatial resolution better than 30 nm. This variation was attributed to the change in the tube structure, which resulted in the modification of the electronic properties, and thus influenced the resonance enhancement. Thereafter, TERS has shown ability for studying localized vibrational modes, chirality changes, the detection of dopants and local deformation in SWCNTs.219–222 For example, the Jorio and the Novotny group utilized TERS to spatially resolve the charged defect in a single SWCNT. A new peak whose frequency is near the G′ band was observed at the defect sites, marked as image file: d4cs00588k-t2.tif. This new peak originated from the existence of charged defects (dopants), which lead to the renormalization of the electron and phonon energies. Furthermore, image file: d4cs00588k-t3.tif exhibited a blue-shift for p-doping, while a red-shift for n-doping relative to the G′ peak, which makes the behavior of the image file: d4cs00588k-t4.tif band a sensitive probe to reveal the local effect of a single defect on the electronic properties of SWCNT.223

To further improve the spatial resolution for finer characterization of the localized property, the Kawata group performed STM-based gap mode TERS in a sealed chamber under an N2 atmosphere (Fig. 9a). The nanogap formed by the Au tip and Au substrate provided a strong plasmonic enhancement, enabling a spatial resolution of 1.7 nm under ambient conditions (Fig. 9d).111 The simultaneously obtained high-resolution STM image and corresponding TERS imaging reconstructed from different vibrational modes of CNT (Fig. 9c and e) enabled visualization of the lateral confinement of each Raman mode in real spaces, respectively. The underlying physical chemistry, including the diameter of tubes, local defects and bundling effect can be studied using the spatially resolved Raman spectra corroborated with topographies. With the further improvement of the spatial resolution to 0.7 nm in UHV–TERS, the Dong group directly visualized the heterogeneity inside a single SWCNT. A stronger D band signal at the end of the tube and overall uniform distribution of the G band signal was observed, providing more structural information than that of the STM image. The D band signals were also observed far away from the tube end and extended to around 3 nm, indicating the presence of defects in the SWCNT and a relaxation length of the D band near a defect. In addition, the spectral evolution of the G band helps to reveal the strain inside a single SWCNT.207 According to the above-mentioned results, we can conclude that the well-correlated topography and chemical information obtained from TERS are beneficial to reveal the properties of CNTs. Recently, the analysis of the interwall interactions in multiwalled carbon nanotubes (MWNTs) was achieved through D band imaging by the Verma group.224 Nowadays, SWCNTs have been used as standard materials for the evaluation of the spatial resolution in TERS.122,123,130,132,190


image file: d4cs00588k-f9.tif
Fig. 9 TERS study of CNTs and carbyne. (a) Schematic illustration of the STM-based TERS characterization of CNTs on an Au substrate. (b) High-resolution STM images of three CNTs. Simultaneously obtained STM images (c) and TERS images (e) reconstructed from different vibrational modes of CNT. (d) Intensity distribution of the D band across the gray line in TERS image reconstructed from the intensity of D band. Reprinted with permission from ref. 111. Copyright 2014, Springer Nature. (f) Illustration of the carbyne chain encapsulated in a DWCNT. Simultaneously obtained TERS image (g) and AFM image (h) of a single confined carbyne. (i) Typical TERS and micro-Raman spectrum of confined carbyne. Reprinted with permission from ref. 225. Copyright 2021, the American Chemical Society.

As another one-dimensional carbon material, carbyne was successfully synthesized inside MWCNTs or double-wall carbon nanotubes (DWCNTs) (Fig. 9f).226,227 This new material is predicted to possess a higher stiffness and elastic modulus than any known materials.228 In addition, the electronic property of carbyne can be tuned by the length of the carbon line as well as the external perturbations.229–231 However, exploring the properties of carbyne via experiment is quite challenging given that the nanoscale and even atomic scale carbon line is encapsulated in a DWCNT. In this regard, TERS offers nanoscale spatial resolution to directly measure the Raman spectra of the carbyne inside a DWCNT (Fig. 9g–i),225,227,232–234 which is a convincing way to ensure that all the spectral features come from single carbyne. Moreover, TERS was also used to investigate the interaction between the inner carbyne within DWCNTs. For example, it was found that the chirality and the radius of the tube will determine the properties of the encapsulated carbyne, and thus could be reflected by the Raman shift. A linear scaling of the C-mode frequency with the diameter of the CNT exists, indicating that the dominant interaction between the nanotube and the encapsulated chain is van der Waals force.232

4.2. TERS study of graphene

After the successful mechanical exfoliation of graphene in 2004,235 it soon became a rapidly rising star in both materials science and condensed-matter physics236 owing to its quite unique properties, such as “massive” electrons, ultra-high electron mobility at room temperature, and unexpectedly high opacity.237–240 Raman spectroscopy is a versatile tool for the study of graphene,48 where its number of layers,48,49,241 defects, doping and strain242–244 can be reflected by its Raman peaks, respectively. However, limited by the spatial resolution of far-field Raman spectroscopy, it is impossible to deliver the local distributions of properties, such as strain, doping, and interlayer interaction. Moreover, the defects will dramatically influence the property of graphene, but the accurate characterization of the defect sites via far-field Raman spectroscopy is quite challenging given that the signals will be easily obscured by the average signal from the pristine material.

Thus, to overcome this limitation, TERS was utilized to obtain localized Raman spectra from the near-field regime. It is important to note that all the vibrational modes of single-layer graphene (SLG) are in-plane modes, while mainly out-plane modes are effectively enhanced in gap mode TERS according to the selection rule. The Jorio group compared the enhancement of graphene in TERS performed on different substrates and obtained a stronger enhancement on a glass substrate than that on a Au substrate.245 In this regard, most of the TERS studies on graphene were performed in the non-gap mode configuration.246–252

The Su group studied the edge defect of SLG by TERS in detail,253,254 where an obvious D band at the edge site was observed (Fig. 10a–c) and the spatial distribution of the D band near the edge was analyzed to be around 4.2 nm.250 Recently, the Cançado group directly measured the localization of the D band experimentally to be around 7.8[thin space (1/6-em)]±[thin space (1/6-em)]3.2 nm near the edge by deconvoluting the excitation field of the tip used for the experiment from the TERS image.252 This directly measured value agrees with the previous indirect far-field Raman measurement, which is 4[thin space (1/6-em)]±[thin space (1/6-em)]1 nm.255,256 These results showed that beyond direct visualization of the specific sites, TERS can be a powerful tool to obtain the spatial distribution of the unique property induced by these sites, such as the different electronic structures compared with the pristine material. In addition, TERS can even visualize some atomic sites such as point defects. The Pollard and the Roy group successfully probed the point defects in flat graphene (Fig. 10d) by TERS using the TERS image reconstructed from the intensity of the D band (Fig. 10e). It can be concluded that although there are no specific defects in the topography, numerous point defects are observed from the appearance of the D peak on the defect site (Fig. 10f).250


image file: d4cs00588k-f10.tif
Fig. 10 TERS characterization of defect sites in graphene. (a) AFM image of single-layer graphene (SLG). (b) Overlay image of TERS maps using the intensity of the 2D, D and G peaks. (c) TERS spectra and far-field Raman spectra obtained at the positions marked in (b). Reproduced from ref. 254 with permission from The Royal Society of Chemistry. (d) AFM image of SLG. (e) TERS image reconstructed from the intensity of the D band. (f) Typical far-field Raman spectrum and near-field Raman spectra with and without defects. Reproduced from ref. 250 with permission from The Royal Society of Chemistry.

Since the discovery of strong correlations and superconductivity in twisted bilayer graphene (TBG),237 it has become a great challenge to access the modulations and understand the related effects from the superlattice in TBG, given that the modulations are so small that numerous experimental techniques failed in the accurate characterization. Recently, the Jorio group utilized TERS to measure the localization of the lattice dynamics in TBG (Fig. 11a).82 The superlattice image in TBG could be clearly visualized in the TERS image compared with the featureless topography, where the spatially resolved Raman spectra could reflect the distribution of the local vibrational state and the electronic structure (Fig. 11b). The nanoscale spatial resolution and high sensitivity of TERS enabled the observation of the different G peaks at different stacking regions of various positions (Fig. 11e). These vibrational states were demonstrated to be localized in specific regions based on the TERS image (Fig. 11d), which agreed well with the predication of the theory (Fig. 11c). In addition, the electron localization was also measured from the analysis of the G′ peak, whose line shape is sensitive to the electronic structure. This TERS-based nano-Raman characterization successfully pushed the understanding of the phonon-related effects to the nanometer and even atomic scales from experiment. With the successful synthesis of a large variety of graphene-related materials, the TERS technique was also used to characterize the local properties of these materials, such as graphene oxide,80 functional graphene257,258 and graphene nanoribbons.259,260


image file: d4cs00588k-f11.tif
Fig. 11 Characterization of twisted bilayer graphene (TBG) by TERS. (a) Schematic of the TERS configuration. (b) Simultaneously obtained TERS image and topography image of the TBG. (c) Prediction of the spatial distributions of the phonon density of states (DOS) (left) and Raman intensity of the lower-frequency G-band based on theoretical simulation. (d) Hyperspectral mapping obtained with the contribution of AA stacking. (e) Typical G band spectra of different regions inside the TBG. Reprinted with permission from ref. 82. Copyright 2021, Springer Nature.

4.3. TERS study of transition metal dichalcogenides (TMDCs)

TMDCs are a variety of 2D materials containing a sandwich-like structure. Owing to their original band gap and high carrier mobility at atomic thickness, this type of 2D material is regarded as the candidate for next-generation semiconductor transistors.261–263 However, structural defects, which are generated during the production process, always exist in most TMDCs. Although these defects are usually in the nanometer and even atomic scale, they can dramatically influence the electronic properties of TMDC materials, and thus affect their application.264 In this case, the nanoscale spatial resolution under ambient condition of TERS, which is similar to the real working environment of TMDC-based devices, offers an opportunity to achieve the accurate characterization of the defect sites and the defect-induced transition region.81,146,265–268

The Jeong group successfully unveiled the defect-related Raman mode by TERS, where the D and D′ modes were observed during TERS imaging of monolayer chemical vapor deposition (CVD)-grown WS2. These two modes were assigned to the S vacancies according to the DFT calculation, and thus could be regarded as a probe to evaluate the quality of 2D materials.265,269 However, we note that owing to the limited spatial resolution of TERS under ambient conditions, direct characterization of single-point defects is quite challenging, which can be probably achieved with UHV–TERS. The Verma group probed the nanoscale defects and wrinkles on MoS2 by TERS, observing plenty of nanoscale inhomogeneities, which may come from the preparation process.266 Recently, our group probed the edge-related properties, including the edge-induced reconstruction of band structure and electron density in bilayer MoS2 (Fig. 12a) by TERS. Compared with the TERS spectra obtained at the basal plane, a new peak at 396 cm−1 was observed at the edge site (Fig. 12b). This new peak was assigned to the LA[thin space (1/6-em)]+[thin space (1/6-em)]TA mode, which originated from the double resonance Raman (DRR) process activated by the edge-induced band bending, and thus sensitive to the electronic band structure. Then, the intensity profile of the LA[thin space (1/6-em)]+[thin space (1/6-em)]TA mode and the A1g mode (whose intensity is sensitive to the electron density) as a function of the position over the edge site could be obtained from the TERS line-trace (Fig. 12c), which clearly showed the defect-induced variation in electron density and the electronic band structure. After deconvoluting the spatial resolution of TERS, the edge-induced band bending region and the electron transition region (ETR) could be measured experimentally. This quantitative measurement of the defect-induced property will be helpful for the design of the TMDC-based devices in the future.81


image file: d4cs00588k-f12.tif
Fig. 12 TERS characterization of transition metal dichalcogenide (TMDC) materials. (a) Typical AFM image of MoS2 with different types of defects. (b) TERS spectra obtained at the 2L edge and 2L basal plane of MoS2. (c) TERS line-trace with the intensity of the 396 cm−1 peak and 406 cm−1 peak over the 2L edge of MoS2. Reprinted with permission from ref. 81. Copyright 2019, Springer Nature. (d) AFM image of a nanobubble over the heterostructure. (e) Combined TERS image reconstructed from the intensity of the 22 cm−1 peak (blue) and 420 cm−1 peak (red). (f) Average TERS spectra of the marked areas in the TERS image. Reprinted with permission from ref. 270. Copyright 2022, the American Chemical Society. (g) AFM image of a chemical vapor deposition (CVD)-grown lateral heterostructure. (h) and (i) TERS images of two types of lateral heterostructures. Reprinted with permission from ref. 271. Copyright 2021, the American Chemical Society.

Besides defects, the strain inside the TMDCs will also influence the electronic property of materials. By loading MoS2 on Au nanotriangles, the Rahaman and Rodriguez group revealed the localized strain by TERS, where around 1.4% of biaxial strain was introduced in MoS2 according to the frequency shift of up to around 4.2 cm−1 of the strain-sensitive E2g mode.272 This strain-related property can also be probed by the TEPL technique.273–275 Based on nanoscale imaging, the strain distribution inside a single wrinkle (the center and the boundary)274 or individual nanobubble275 could be obtained by TEPL, which shed light on the measurement of the strain in the materials.

In addition to the characterization of the properties of a material, TERS can also be employed to study the interaction between TMDC materials and the substrate or the interactions between different TMDC materials. The Frank group investigated the strong interaction between monolayer MoS2 and Au by TERS. A dramatic frequency shift in the E2g peak and splitting of the image file: d4cs00588k-t5.tif peak in MoS2 was observed when the substrate was changed from SiO2/Si to Au. The frequency shift of the E2g mode originated from the lattice mismatch between MoS2 and the Au substrate, while the splitting of the image file: d4cs00588k-t6.tif mode originated from the electron doping owing to the strong interaction between MoS2 and the Au substrate.276 The Jorio group investigated the property of the in-plane homojunctions consisting of graphene on talc and graphene on SiO2. According to the TERS image over the interface, an abrupt change in the 2D band and the G band in terms of both intensity and frequency was observed, which was then assigned to the local doping effect from the talc substrate rather than strain effect. In addition, an oscillation of the 2D/G intensity ratio was observed across the interface, indicating the formation of a charge depletion region.277

Recently, the Frank and the El-Khoury group investigated the interlayer interactions in WSe2/WS2 vertical heterostructures (Fig. 12d) by ultralow-frequency TERS (ULF-TERS), which were very sensitive to the interlayer interaction. The TERS image of a nanobubble reconstructed from the intensity of the 22 cm−1 peak (blue) and the 420 cm−1 peak (red) showed a heterogeneous interlayer interaction over the heterostructure (Fig. 12e). The average spectra from three different regions (flat region, peripheral region of the bubble and center of the bubble) (Fig. 12d) showed that the 22 cm−1 peak could be observed in both the flat region and the peripheral region of the bubble, but this peak disappeared and the peak intensity at 420 cm−1 (assigned to single-layer WS2) increased in the center region of the bubble. This result was attributed to the decoupling of WS2 and WSe2 at the center of the bubble.270 The above-mentioned results clearly show the power of TERS in the investigation of the interaction between the material and the substrate as well as the interlayer interactions, which can help to better understand the role of the substrate and the role of the interlayer interaction on the property of the material, and thus help the future application of the material.

Recently, the successful synthesis of an in-plane (lateral) heterostructure,278–280 which is an atomically thin p–n junction, showed promising applications as semiconductor transistors. The TERS and TEPL techniques were utilized to characterize the transition region of the in-plane heterostructure under ambient condition, which could influence the optoelectronic properties of the heterostructure.271,281,282 The Su group utilized TERS to characterize the atomic diffusion in CVD-grown lateral heterostructures (Fig. 12g), observing two types of lateral heterostructures according to the size of the transition region. A clear interface with an abrupt change was observed in one type of heterostructure (Fig. 12h), while a broad interface was obtained in the other type of heterostructure (Fig. 12i), which was assigned to the alloy region originating from the CVD process.271 A similar alloy region range was observed in a heterostructure by the Kung group.282 These observations demonstrate that TERS can be a promising tool in ambient condition for the fast evaluation of the quality of an interface, which is usually done by electron microscopy under vacuum conditions.

4.4. TERS study of other 2D materials

Beyond graphene and TMDCs, numerous other 2D materials have also been successfully synthesized, such as silicene283,284 and borophene,285,286 offering unique properties and possible applications in many fields.287,288 Thus, TERS has been utilized to probe these materials to investigate their vibrational properties at nanoscale. Owing to the easy oxidation of these 2D materials, UHV–TERS is usually used to provide a clean and oxygen-free environment.

The Wu group studied the vibrational properties of monolayer silicene on an Ag(111) surface by UHV–TERS.79 Compared with far-field Raman spectroscopy, the ultra-high spatial resolution of UHV–TERS enabled the distinction of different phases of silicene on Ag(111). The high-resolution STM images of silicene with different phases on Ag(111) and the corresponding TERS spectra clearly showed the difference in the Raman spectra (Fig. 13a–c). Furthermore, TERS makes it possible to distinguish the local vibrational properties induced by domain boundaries in silicene. A 4[thin space (1/6-em)]×[thin space (1/6-em)]4 – β phase stabilized by the strain effect existed at the boundary of different normal 4[thin space (1/6-em)]×[thin space (1/6-em)]4 – α phase domains (Fig. 13d). The TERS spectra at these two regions and the edge site of the α phase showed a slight red-shift of around 4 cm−1 for the A1 mode (located at 170 cm−1) and a decrease in the intensity of the ZO modes in the β phase compared with that in the α phase (Fig. 13e). An even lower intensity of the A1 and A2 peaks was observed at the edge site of the α phase. Considering the sensitivity of the A2 and ZO peaks to the interaction between the silicene and the substrate, these changes in the Raman spectra in the β phase and at the edge site were attributed to the decreased interaction between the silicene and the substrate at these special sites.


image file: d4cs00588k-f13.tif
Fig. 13 TERS characterization of silicene on Ag(111). (a)–(c) High-resolution STM images of silicene of different phases and the corresponding TERS spectra. (d) STM image of silicene 4[thin space (1/6-em)]×[thin space (1/6-em)]4 – α phase and β phase. (e) TERS spectra obtained from the silicene 4[thin space (1/6-em)]×[thin space (1/6-em)]4 – α phase, 4[thin space (1/6-em)]×[thin space (1/6-em)]4 – β phase and domain edge. Reprinted with permission from ref. 79. Copyright 2017, American Physical Society.

Besides silicene, TERS was also utilized to study the vibrational properties of 2D borophene to clearly determine its structure with β12 and χ3 phases. The local vibrational properties from commensuration and strain in the smooth β12 phase and striped β12 phase were also observed by TERS.289 Recently, the Jiang group studied the interaction between tetraphenyldibenzoperiflanthene (DBP) and borophene by TERS. It was found that the adsorption of DBP on borophene introduced subtle tensile strain in borophene, and the strain was found to be highly localized in the range of 1–1.5 nm according to the TERS image.290

4.5. TERS study of other materials

Besides the above-mentioned materials, TERS was also used in the nanoscale characterization of some other important materials, including perovskite BaTiO3,291 strontium titanate (SrTiO3),292 CdSe nanowire,293 Si nanoribbon,294 a-Si,155,198 and 2D polymers (2DPs).295–297 The Zenobi group characterized 2DPs in detail at the nanoscale by TERS. Recently, the in situ characterization of the plasmon-driven polymerization process on Au(111) was achieved. Combining TERS images with DFT calculations, the whole mechanism of polymerization including a hot electron tunneling process and a self-stimulating growth process was proposed.297 The Wu group and the Jiang group investigated Si nanoribbons on the Ag(110) surface, combining noncontact (nc)-AFM and TERS. Both the Si nanoribbon and Si clusters were demonstrated to be composed of purely pentagon rings, instead of the predicted hexagonal model.294 Recently, the Nanda and the Yang group utilized TERS to characterize the nanoscale heterogeneity of the SEI after different electrochemical cycling over a-Si. The composition of the SEI gradually evolved with different species dominating the chemical composition of the SEI. A multilayer structure with different chemical compositions in each layer of the SEI of a-Si was proposed, while the carboxylates with various molecular conformations dominated the outer layer of the SEI.155

5. Nanoscale chemical characterization of interfaces by TERS

Surfaces and interfaces are bridges between materials and their applications, where numerous key processes such as charge transfer and energy conversion occur. The performance of a material is sometimes dominated by the processes occurring on its surface and interface. Taking chemical reactions as an example, given that almost all the key steps of a chemical reaction including adsorption, reaction, and desorption occur on the surface and interface, the surface structure, molecule–surface interaction and intermolecular interaction will influence the above-mentioned processes, and then result in either different activity or selectivity. Therefore, a complete understanding of the surface and interfacial processes, including the molecular level and nanometer scale information of the surface and interface, becomes the key to understand the origin of the performance and design of highly efficient processes. Since the first demonstration of TERS, this technique has been utilized to characterize surfaces and interfaces benefitting from its nanoscale spatial resolution and abundant chemical information. In the past 20 years, TERS possesses broad application in the nanoscale chemical characterization of the solid–gas interfaces, solid–liquid interfaces and electrochemical interfaces, where the molecular structures, the detailed reaction process and the nanoscale active sites have been probed at the nanoscale and even atomic scale by TERS.

5.1. TERS for solid–gas interface

5.1.1. Determining the molecular structure of surface species. The molecular structure, including the chemical structure, orientation, and intermolecular interaction, will influence the property and reactivity of these surface species. Therefore, the accurate determination of the molecular structure of surface species is regarded as one of the key steps to elucidate surface processes, such as chemical reactions. A large variety of chemical information, including chemical composition and variation in bond length or bond strength can be extracted from the frequency, relative intensity and peak width of Raman peaks. This abundant chemical information makes Raman spectroscopy a powerful tool for characterizing surface species. In TERS, the selection rule, which describes that only the vibrational modes with a component that is perpendicular to the surface can be enhanced, further allowing TERS to reveal the orientation of the surface species. Since the first demonstration of TERS, this technique has been used to characterize the molecular structure of surface species at the solid–gas interface.

The Zenobi group reported the TERS results of brilliant cresyl blue (BCB) molecules and a C60 thin film on a glass substrate.76 The Pettinger group demonstrated that the Raman signal from monolayer CN ions adsorbed on rough Au and Ag surfaces could be obtained by TERS.298 However, the complicated environments of rough surfaces make it difficult to clearly understand the structure of surface species. Subsequently, the TERS signal of CN and malachite green isothiocyanate (MGITC) adsorbed on an atomically smooth single crystal surface was obtained for the first time.299 In addition, TERS spectra of thiophenol molecules with varying Raman shifts and relative intensities adsorbed on Au and Pt single crystal surfaces were observed, indicating different molecule-surface interactions.300 Thereafter, the Zenobi and the Deckert group investigated some large biomolecules on an Au or Ag substrate, such as cytochrome c,301 glutathione302 and histidine.303 These early works demonstrated from the proof-of-principle, TERS has the ability to determine the molecular structure of surface species from nano-resolved Raman spectra, especially on the well-defined single crystal surfaces.

Recently, with the development of the TERS technique, it is now possible to directly visualize the nanoscale heterogeneity of molecular structures by TERS imaging. To suppress the spectral diffusions and inhibit chemical reactions under ambient conditions at room temperature, which may interference the analysis of the results, the Park group coated an ultra-thin Al2O3 layer over a few BCB molecules adsorbed on an Au substrate.304 The TERS image reconstructed from the intensity of the ∼580 cm−1 peak (upper panel in Fig. 14a) showed obvious variations in peak intensity at different sites (lower panel of Fig. 14a). These sites could be divided into two groups according to the Raman shift as well as the peak width of the peak at around 580 cm−1. One group showed a similar Raman shift and less variation and the other group showed a large variation in both the Raman shift and peak width, indicating the nanoscale conformational heterogeneity of the molecules. The Zenobi group utilized 4-mercapopyridene (4-PySH) molecules as a buffer layer, and then adsorbed cobalt tetraphenyl-porphyrin (CoTPP) molecules to build coordination species as a model system for TERS to characterize the metal–organic coordination structures at the solid–gas interface (Fig. 14d).305 Different from the TERS spectra obtained from CoTPP on Au(111), which only showed some broad peaks, CoTPP/4PySH on Au(111) clearly showed the characteristic peaks of the CoTPP molecules. This variation may come from the weakened molecule–metal interaction with the existence of a buffer layer, but the detailed origin of this variation still requires further studies. The TERS image of CoTPP/4PySH on Au(111) showed three types of TERS spectra, reflecting the differences in the orientation of the CoTPP/4PySH coordination species (Fig. 14e).


image file: d4cs00588k-f14.tif
Fig. 14 Elucidating the molecular adsorption configuration at the solid–gas interface. (a) Upper panel: TERS image reconstructed from the intensity of the ∼580 cm−1 peak of a few BCB molecules. Lower panel: TERS spectra recorded at the marked spots in the TERS image. (b) and (c) Two groups of TERS spectra recorded in the TERS image with different Raman shifts and peak widths. Reprinted with permission from ref. 304. Copyright 2022, Springer Nature. (d) Schematic illustration of the TERS characterization of the metal–organic coordination structures at the solid–gas interface. (e) Three types of TERS spectra of CoTPP/4PySH extracted from the TERS image. Reprinted with permission from ref. 305. Copyright 2021, the American Chemical Society.

Besides the molecule itself and its interaction with the substrate, TERS can also elucidate the intermolecular interaction, which will influence the molecular structure at the interface. Our group utilized the aromatic C[double bond, length as m-dash]C bond stretching vibration of the 4′-(pyridin-4-yl)biphenyl-4-yl-methanethiol (PBT) molecule as a marker to monitor the strength of the intermolecular interaction of SAMs on the Au(111) surface.203 The aromatic C[double bond, length as m-dash]C bond stretching peak first showed a red-shift, and then a blue-shift. The first red-shift originated from the stronger π–π interaction between molecules with increasing surface coverage, while surface restructuring occurred as the molecular coverage further increased owing to the strong repulsive interactions, leading to the latter blue-shift. It is interesting to note that in addition to a single type of molecules on the surface, benefitting from the ability to offer chemical information, TERS can also be employed to investigate the molecular structure of mixed components of molecules on a surface.306,307 Therefore, the coadsorption, displacement and separation processes in SAMs, and even restructuring of the underlying Au surface can be visualized by TERS, providing insight into the surface processes at the molecular level.

Compared with ambient condition, the UHV condition provides a more controllable and cleaner environment, which has shown power in studying surface science. In 2007, the Pettinger group first reported a room temperature UHV–TERS apparatus.149 The correlated UHV–TERS and STM imaging of a single BCB molecule was achieved at a spatial resolution of around 15 nm.308 Furthermore, the LT-UHV condition dramatically reduced the thermal drift of the system, enabling TERS characterization at sub-nanometer spatial resolution to directly determine the chemical structure at the single-molecule and even single-chemical-bond levels.117,118,150,152,153,205,309–311

In 2013, the Dong group first convincingly performed TERS imaging of a single-molecule with a spatial resolution of around 0.5 nm.150 This milestone work clearly showed the unique power of TERS in chemical imaging at the sub-nanometer level, making it possible to distinguish two adjacent molecules with similar chemical structures. In addition, combining the TERS spectra of single molecules adsorbed on different positions and calculation, their molecular structure, especially their adsorption configurations and orientations can be clearly determined.150,204 For example, the Dong group successfully distinguished adjacent individual DNA bases in a network by TERS. The site-dependent TERS spectra revealed that the co-deposited thymine (T) was located outside the boundary of the adenine (A)-networks. Some of the T molecules formed repeated trimer-like structures.312 In 2019, the spatial resolution of TERS was further improved to impressively the single-chemical-bond level by the Apkarian group117 and the Dong group,118 where distinct spectral features can be obtained at different positions inside a single-molecule. With this powerful tool, the Dong group developed the so-called “scanning Raman picoscopy (SRP)”, which can visualize the distribution of each vibrational mode inside a single-molecule, and then assemble the chemical structure of this molecule through a Lego-like building process, providing a new way to identify chemical structures at the chemical bond level.118 With the further combination of nc-AFM, the molecular structures of surface species could be accurately determined at the single-bond level by the Wang group.310 The single-molecule resolved STM images indicated that on the Ag(110) surface, besides the intact pentacene molecules (α), two new species with different shapes (β and γ) existed after applying a highly positive bias (Fig. 15a). Subsequently, more structural information was obtained from the nc-AFM image, where two bright protrusions located around the C–H bonds were observed in α, while one of them disappeared in β, and both of them disappeared in γ (Fig. 15b). These bright spots were attributed to the C–H bonds according to the strongest C–H stretching mode in α and the absence of the C–H stretching mode in γ in the TERS spectra obtained at the marked positions (Fig. 15c). This could be further corroborated by the TERS images reconstructed from the intensity of the C–H stretching mode (Fig. 15d), where two bright spots existed in α, one bright spot in β and no bright spot in γ. In this regard, the molecular structure is now accurately determined by the combination of modern tip based techniques.


image file: d4cs00588k-f15.tif
Fig. 15 Determining molecular structure at the single-chemical-bond level. (a) and (b) STM and constant-height AFM images of individual intact pentacene molecules (α) and two transformed species (β and γ). (c) Typical TERS spectra obtained at the middle and the end site of the three species. (d) Simultaneously obtained STM images and TERS images reconstructed from the intensity of the C–H stretching mode of the α, β and γ species. Reprinted with permission from ref. 310. Copyright 2021, The American Association for the Advancement of Science.

More recently, the Dong group tracked the dynamic variations in a single-molecule, including the tip-induced bond weakening, tilting and hopping process by TERS.202 Taking a single CO molecule adsorbed on the Cu(100) surface as a model system, a dramatic red-shift in the C–O stretching vibrational mode was observed with the tip approaching the molecule, indicating an obvious tilt in the CO molecule. Then the further approaching of the tip-induced a two-step hopping according to the variation in the C–O stretching mode, initiated by the first hopping to the nearest Cu atom via the bridge site, followed by the second hopping process to the diagonal Cu atom via the bridge site.

5.1.2. Monitoring the chemical reaction at nanoscale and atomic scale. After revealing the nano-resolved molecular structure, TERS has been applied to monitor chemical reactions at the nanoscale.313–316 The pioneering work was done in 2012 by the Deckert group,313 where an Ag-coated AFM tip was used as the catalyst to initiate the dimerization of para-nitrothiophenol (pNTP) adsorbed on an Au nanoplate to produce p,p′-dimercaptoazobisbenzene (DMAB) under the illumination of a 532 nm laser (Fig. 16a). Then, the reaction was monitored by TERS excited with another 632.8 nm laser. The reaction can only occur with illumination by a 532 nm laser, which matches the SPR of the Ag tip with the presence of the characteristic Raman spectrum of DMAB (Fig. 16b). Subsequently, this plasmon-driven coupling reaction was proven to be influenced by the adsorption configuration. Our group found that on the Au(111) surface, the oxidative coupling of p-aminothiophenol (PATP) can take place under illumination with a 632.8 nm laser. However, this reaction cannot occur on the Ag(111) surface either under illumination with a 632.8 nm laser or a 532 nm laser. DFT calculation suggested that PATP adopts a tilted orientation on Au(111), while a relatively vertical orientation on Ag(111), which results in different reaction activities.317
image file: d4cs00588k-f16.tif
Fig. 16 Chemical reaction monitored by ambient TERS. (a) Schematic illustration of the TERS setup used for monitoring the chemical reaction. (b) Two TERS spectra obtained before and after illumination with a 532 nm laser. Reprinted with permission from ref. 313. Copyright 2012, Springer Nature. (c) and (d) left panel: high-resolution STM image of the pNTP-adsorbed Au(111) prepared by the immersion protocol and drop-cast protocol. Middle panel: TERS images reconstructed from the intensity ratio between the characteristic peaks of DMAB and pNPT. Right panel: TERS spectra of the marked sites in the TERS images. Reprinted with permission from ref. 318. Copyright 2022, the American Chemical Society.

It is interesting to note that the way how the sample is prepared will influence its adsorption configuration, and thus reaction activity. The Zenobi group compared the activity of pNTP adsorbed on an Au surface prepared by the drop-cast and immersion protocols. The molecular-resolved STM images showed that the drop-cast protocol produced a less ordered adlayer than that of the immersion protocol (Fig. 16c and d), respectively. The TERS images reconstructed from the intensity ratio between the characteristic peaks of DMAB and pNPT showed that more than 72.3% of the drop-cast sample was covered by DMAB, while this value decreased to less than 7.3% on the immersion sample, indicating a higher efficiency in a less ordered surface (drop-cast sample).318 Recently, the study of plasmon-driven reaction by TERS was extended to a series of monometallic and bimetallic surfaces by the Kurouski group.316,319–321 For example, under the illumination of light, the TERS spectra revealed that 4-mercapto-phenyl-methanol (MPM) could be oxidized to 4-mercaptobenzoic acid (MBA) on Au@Pt nanoplates, while MBA could be reduced to MPM on Au@Pd nanoplates. In addition, both MBA and MPM would be transformed to thiophenol (TP) on Au nanoplates. A higher catalytic activity was observed at the edges of these nanoplate catalysts.321

With the dramatic improvement in stability and spatial resolution, the chemical reaction can be probed at the single-molecule even the single-chemical-bond level by UHV–TERS. Recently, taking a single melamine molecule chemically adsorbed on the Cu(100) surface as a model system, the Dong group revealed the H tautomerization process by TERS at the single-chemical-bond level.154 H tautomerization rapidly occurred under laser illumination and at a positive bias. The TERS spectra of melamine before (C0) and after (C1/C2) H tautomerization showed that no symmetric and antisymmetric vibrations of an –NH2 group could be observed simultaneously in C1/C2, indicating the break of one N–H bond in the top –NH2 group (Fig. 17a). In addition, another new N–H bond (blue peak obtained at position a′ on C1/C2) could be observed at one side of the cyanuric ring, suggesting the formation of a new N–H bond. Then, the whole H tautomerization can be clearly obtained, including a bond breaking, H atom transferring and new bond formation process (Fig. 17b).


image file: d4cs00588k-f17.tif
Fig. 17 Monitoring the chemical reaction at the single-chemical bond level. (a) Comparison of the TERS spectra before and after H tautomerization. (b) Schematic illustration of the N–H stretching mode at the top and side of the cyanuric ring. Reprinted with permission from ref. 154. Copyright 2021, the American Chemical Society. STM image (c) and TERS spectra (d) of borophene after exposure to atomic oxygen. STM image (e) and TERS (f) spectra of borophene after exposure to molecular oxygen. (g) STM image of oxidized borophene and 2D TERS image of the white-rectangle marked region reconstructed from the intensity of the 205 cm−1 mode. Reprinted with permission from ref. 86. Copyright 2022, Springer Nature.

Furthermore, more chemical reactions can be initiated by introducing a small amount of gas in the UHV system, such as O2, which serve as the reactive species. Thus, the chemical reaction can be accurately studied in this well-defined system with a much higher spatial resolution. The Van Duyne group investigated the chemical adsorption of O on cobalt phthalocyanine (CoPc) molecules on Ag(111), where two adsorption configurations, assigned to O2/CoPc/Ag(111) and O/CoPc/Ag(111), were observed after introducing O2 according to the molecule-resolved STM images and the TERS spectra.322 The Jiang group studied the oxidation reactions of borophene with single-bond sensitivity in UHV–TERS.86 After introducing atomic oxygen, homogeneous chemically adsorbed oxygens were observed in the basal plane of borophene from both the STM image (Fig. 17c) and TERS spectra obtained from borophene and oxygen adatom (Fig. 17d), indicating the homogeneous oxidation of borophene in the presence of atomic oxygen. After exposure to a high molecular oxygen dose, a different oxidation process was observed, where the edge started to degrade and some inhomogeneous oxide species formed (Fig. 17e and f). The TERS image of two types of adsorbates with different heights reconstructed from the intensity of 205 cm−1 (Fig. 17g) clearly distinguished these two types of species, with three oxygen adatoms identified as bright pixels, while one higher cluster appeared featureless. This difference was proposed to originate from the varying adsorption configurations of molecular O2 on borophene. Furthermore, phase-dependent oxidation reactivity was also observed, where the ν1/6 phase possesses a higher reactivity than the ν1/5 phase.

5.1.3. Deciphering the active sites. Revealing the active sites and building the accurate structure–activity relationship are always the goal in heterogeneous catalysis. With the ability to provide the correlated structural and chemical information at the nanoscale, the TERS technique shows great advantages.

In 2015, the Roy group and the Wain group utilized TERS to image the different activities on nanostructured Ag surfaces.315 The TERS image reconstructed from the intensity of the Raman peak assigned to DMAB (Fig. 18b) over an Ag film (Fig. 18a) showed that the plasmon-driven coupling reaction of pATP to DMAB only occurs on a few locations. The active and inactive sites were distinguished with a spatial resolution of 20 nm. This is the first work showing the ability of TERS in revealing the active sites; however, the use of a polycrystal substrate makes it difficult to build a clear structure–activity relationship. Subsequently, our group fabricated a well-defined sub-monolayer Pd/Au(111) bimetallic catalyst via electrochemical underpotential deposition, which was then applied as a model system for TERS characterization. Then, the phenyl isocyanide (PIC) molecule was selected as the probe molecule, where the Raman shift of the NC triple bond of PIC is sensitive to the electronic properties of the metal surface, and thus served as a marker to reveal the site-dependent electronic property (Fig. 18c). According to the TERS line-trace with a spatial resolution of 3 nm, the NC triple bond red shifted at the Pd step edge compared with that at the terrace site, indicating the enhanced reactivity of the PIC molecule owing to the higher electronic d band profile of the step edge of the Pd site.83 Subsequently, a similar strategy was applied to reveal the unique electronic properties of different Pt sites containing varying coordination numbers over the Pt/Au(111) bimetallic surface.323


image file: d4cs00588k-f18.tif
Fig. 18 Deciphering the varying electronic properties and activities by TERS at the nanoscale. (a) AFM image of an Ag substrate. (b) TERS image measured in the dashed rectangle marked in (a) reconstructed from the intensity of the Raman peak assigned to DMAB. Reproduced from ref. 315 with permission from The Royal Society of Chemistry. (c) Schematic illustration of deciphering the nanoscale electronic property of a bimetallic surface with the probe molecule by TERS. (d) Plots of TERS intensities of three peaks as a function of tip position and topographical height profile of the Pd/Au(111) bimetallic surface. (e) TERS spectra obtained during the line-trace from the Pd terrace to the Au terrace region. Reprinted with permission from ref. 83. Copyright 2016, Springer Nature.

With the successful visualization of the unique electronic properties of different sites in a well-defined catalyst by TERS, it is more interesting to directly compare the different activities of different sites. Recently, our group used TERS to probe the generation and diffusion of active oxygen species (AOS) on a Pd/Au(111) bimetallic catalyst.324 The PBT molecule, which has a large Raman cross section, was used as the probe. In principle, the adsorbed PBT will be oxidized at the site where AOS is generated. As a result, the varying TERS signal of PBT can be a sign of the presence of AOS. After immersion in H2O2 for 30 min, there were no PBT molecules on the Pd site, while the PBT molecules on the Au remained unchanged (Fig. 19a). However, the TERS line-trace across the Pd step edge revealed that some of the PBT molecules on the Au site near the step edge of Pd were also oxidized, indicating that the AOS generated at the step edge of the Pd site diffused to the Au site, and thus oxidized the PBT molecule on Au (Fig. 19b and c). After deconvolution of the TERS spatial resolution from the distribution of PBT molecules after immersion in H2O2, the diffusion length was estimated to be around 5.4 nm (Fig. 19d). Subsequently, the Zenobi group used TERS to study the selective hydrogenation reaction on a Pd/Au(111) bimetallic catalyst.85 The hydrogenation of chloronitrobenzenethiol (CNBT) to chloroaminobenzenethiol (CABT) mainly occurred on the Pd site after exposure to an H2 atmosphere (Fig. 19e) because H2 only dissociated on the Pd site instead of the Au site. It is interesting to note that some of the CNBT molecules adsorbed on the Au sites near the Pd sites were reduced to CABT according to the TERS intensity of the characteristic Raman peak of CNBT and the correlated STM height (Fig. 19f and g). This phenomenon was attributed to the hydrogen spillover from the Pd site to the Au site with the help of calculation.


image file: d4cs00588k-f19.tif
Fig. 19 Visualization of the generation and diffusion of the reactive species by TERS. (a) Typical ambient TERS spectra of PBT on Pd/Au(111) and Au(111) before and after immersion in H2O2 solution for 30 min. (b) Top panel: schematic illustration of the diffusion of OH radicals and STM image of PBT-adsorbed Pd/Au(111) bimetallic catalyst. Bottom panel: topographical height profile along the white dash line in STM image. (c) Ambient line-trace TERS spectra from the terrace of Au(111) to the terrace of Pd obtained after immersion in low-concertation H2O2 solution for 2 min, where a weaker PBT signal is observed near the edge of the Pd site. (d) Analysis of the diffusion of OH radicals (generated at the edge of the Pd site) according to the distribution of the PBT molecules. Reprinted with permission from ref. 323. Copyright 2020, the American Chemical Society. (e) Ambient TERS spectra obtained from the CNBT-adsorbed Pd/Au(111) (top panel) and CNBT-adsorbed Au(111) (bottom panel) surface before and after exposure to an H2 atmosphere and TERS spectrum of CABT adsorbed on Pd/Au(111) (middle panel). (f) and (g) Plot on intensity of characteristic Raman peak of CNBT as a function of position and the correlated topographical height profile on a Pd nanoisland on Au(111), where the distribution of CNBT with a weaker signal is narrower than the size of the Pd nanoisland, indicating the diffusion of H from Pd to Au. Reprinted with permission from ref. 85. Copyright 2020, Springer Nature.

Although the TERS technique has demonstrated power in studying heterogeneous catalysis, the sensitivity of TERS needs to be further dramatically improved to enable the characterization of non-resonant molecules that possess small Raman cross sections but are important in catalysis, such as CO and H. In addition, environment-controlled TERS systems, including temperature, humility, and gas control, are urgently required to perform in situ characterization of real catalytic processes.

5.2. TERS for solid–liquid interfaces

Solid–liquid interfaces are omnipresent in nature and at the heart of numerous fields, including biology, surface science and energy science. Since the development of TERS, there has been growing interest in utilizing TERS to characterize solid–liquid interfaces. In 2009, the Zenobi group performed the first liquid TERS. With the use of a SAM-protected Ag tip, they successfully avoided the contamination of the tip and obtained the TERS signal of thiophenol (PhS) on an Au surface in liquid.325 The Domke group investigated the influence of the STM parameters on the TERS intensity in detail, revealing that the variations in TERS intensity in water were more moderate than that under argon condition.160,326 Thereafter, the Weckhuysen group investigated the plasmon-driven coupling reaction on an Ag substrate in liquid, where heterogeneous activities were observed according to the hyperspectral TERS imaging.327 It is worth mentioning that hyperspectral TERS imaging was also successfully performed in an organic liquid,328 implying the bright future of using TERS to study various interfaces.

As a crucial type of solid–liquid interface, electrochemical interfaces provide a convenient approach to control the surface state by modulating the applied potential, and thus have received increasing attention due to the growing demand for energy storage and conversion. EC–TERS can shed light on characterizing electrochemical interfaces in situ benefitting from its nanometer-resolved chemical information. The first EC–TERS work reported the potential-dependent behavior of PBT molecules at the electrochemical interface (Fig. 20a). The doublet peaks at around 1592 cm−1 and 1600 cm−1 at 0.3 V weakened and became a shoulder peak when the potential negatively shifted to −0.7 V, originating from the partial protonation of the PBT molecules at a negative potential (Fig. 20b and c),158 and this phenomenon was not observed by electrochemical-SERS (EC-SERS).


image file: d4cs00588k-f20.tif
Fig. 20 EC–TERS characterization of the molecular structure at the electrochemical interface. (a) EC–TERS spectra of 4-PBT adsorbed on the Au(111) surface at different potentials. (b) and (c) Schematic illustration of the EC–TERS systems at positive and negative potentials, respectively. Reprinted with permission from ref. 158. Copyright 2015, the American Chemical Society. (d) Potential-dependent EC–TERS spectra of adenine adsorbed on Au(111) surface. (e) Variation in the Raman shift of three peaks upon a positive shift of applied potential. (f) Proposed model for the adsorption and reaction of adenine on Au(111). Reprinted with permission from ref. 329. Copyright 2017, John Wiley and Sons.

The Domke group investigated the potential-dependent variation of adenine adsorbed on the Au(111) surface (Fig. 20d). The intensity of the 734 cm−1 peak decreased and the 260 cm−1 peak showed a blue-shift (top panel of Fig. 20e, assigned to the Au–N bond) with a positive shift in potential, indicating the reorientation of the molecules from tilted to upright to flat relative to Au(111). In addition, the deprotonation process during the positive shift of the potential was revealed by the blue-shift of the peaks at 736 cm−1 (middle panel of Fig. 20e) and 1464 cm−1 (bottom panel of Fig. 20e). Thus, the whole potential-dependent adsorption and reaction process of adenine on Au(111) was proposed (Fig. 20f), including the reorientation and the deprotonation process at different potentials.329 These results clearly show the advantages of EC–TERS as a model system with well-defined configuration and rich chemical information in revealing the molecular adsorption and molecular structure at the electrochemical interface.

When the potential reaches the onset potential of the electrochemical reaction, there will be charge transfer between the electrode and the reactive species at the electrochemical interface. Using a single crystal as a modern system, EC–TERS provides a platform to monitor the reaction process and reveal the reaction mechanism. Our group investigated the redox process of a non-resonant molecule, thiolacetyl-terminated-phenylene ethynylene-substituted anthraquinone (2-AQ) adsorbed on the Au(111) surface. A partially irreversible reaction was observed according to the potential-dependent EC–TERS spectra, which originated from the synergistic effect of the potential and the laser illumination.163 Subsequently, we successfully observed the intermediate of the electrochemical reduction of a nitrobenzene derivative by EC–TERS. A negatively charged intermediate could be captured in neutral solution, while no intermediate could be observed in acidic solution, indicating different reaction processes in acidic and neutral solution.330 Recently, the Lucas and the Noël group also elucidated the electrochemical reduction of a nitrobenzene derivative by EC–TERS. The prevailing product was the azoxybenzene dimer according to the EC–TERS image and the dynamic EC–TERS results. The absence of azoxybenzene dimer-related bonds under an open circuit potential (OCP) with a higher laser power indicated that the dimerization was initiated by the electrochemical potential instead of the laser illumination.331

Owing to the nano-heterogeneity on the electrode, it will be very interesting for EC–TERS to distinguish the activity of different sites, which is significant in electrochemistry, especially electrocatalysis. The Van Duyne group studied the redox reaction of Nile blue (NB) molecules on an ITO electrode (Fig. 21a), where the potential-dependent TERS signals of the NB molecules were obtained to form TERS voltammetry (TERS-CV).165,332 A shift in the redox potential acquired from TERS-CV was observed in comparison with that from the conventional CV (Fig. 21b), which was attributed to the perturbation of the electric double layer by the TERS tip.165 In addition, some step-like behaviors were observed from TERS-CV at different locations, indicating the reduction and oxidation of few molecules and even a single-molecule underneath the tip.332 Subsequently, the EC–TERS image of the NB molecules was recorded on an Au plate on an ITO electrode by EC–AFM–TERS at a spatial solution of around 80 nm.333 By fitting the TERS spectra recorded at each position under different potentials, the site-dependent formal potential of the redox reaction over Au and ITO electrodes can be extracted (Fig. 21c). A more negative formal potential at Au than ITO and a heterogeneous distribution of formal potential on ITO electrode were observed, indicating the nanoscale heterogeneity of the electrode surface. From a technical viewpoint, this work showed the possibility of EC–TERS in revealing the heterogeneous activities in electrochemistry for the first time.


image file: d4cs00588k-f21.tif
Fig. 21 (a) Potential-dependent TERS spectra of Nile blue (NB) molecules. (b) Conventional CV and TERS-CV of NB molecules. Reprinted with permission from ref. 165. Copyright 2015, the American Chemical Society. (c) Distribution of the formal potential extracted from EC–TERS imaging. Reprinted with permission from ref. 333. Copyright 2019, the American Chemical Society.

Thereafter, with the development of instrumentation, the spatial resolution of EC–TERS has been improved to be better than 10 nm,84,145,334 enabling the characterization of the electrocatalysis process at a spatial resolution level of several nanometers. The Van Duyne group in situ characterized the variation in the iron phthalocyanine (FePc) during the oxygen reduction reaction (ORR) by molecule-resolved EC-STM imaging and EC–TERS spectra. The irreversible demetallation of the FePc molecules was observed during electrocatalytic reactions, which led to the loss of activity in ORR. This correlation also indicated that the active site of FePc for the ORR process was the Fe site.335 In 2019, the Domke group in situ characterized the site-dependent variation on Au(111) during the oxygen evolution reaction (OER) by EC–TERS.84 A broad peak located at around 560 cm−1 to 580 cm−1 assigned to AuOx (Fig. 22b) appeared over the defect sites when the OER process was initiated on these sites (potential over 1.45 V vs. Pd–H, Fig. 22a). The hyperspectral TERS imaging (Fig. 22d) performed in the region of Fig. 22c showed that AuOx could only be detected at some of the defect sites on the Au(111) electrode, indicating the varying reactivity of different defect sites. This variation was attributed to the atomic active site heterogeneities, which results in site-specific surface charge and work function. In addition, by analyzing the EC–TERS spectra obtained at different positions, two types of Au oxides (Au2O3: >565 cm−1 and Au2O: <565 cm−1) were observed. The relationship between the height and the identity of the Au oxides was proposed by correlating the height from EC-STM and EC–TERS spectra, where Au2O3 was generated on the flatter defect-terrace-like region, while Au2O formed sharper protrusions over the nanodefects.


image file: d4cs00588k-f22.tif
Fig. 22 In situ characterization of the nano-defect sites by EC–TERS. (a) Cyclic voltammogram of Au(111) in 0.1 M H2SO4. (b) EC–TERS spectra acquired at the defect site for the on and off states. (c) EC-STM image of Au(111) in 0.1 M H2SO4. (d) Corresponding EC–TERS image reconstructed from the peak intensity of AuOx in the region of (c). Reprinted with permission from ref. 84. Copyright 2019, Springer Nature.

Besides identification of the active sites, it is crucial to monitor the evolution of the geometric and electronic structures of the active sites, which indeed determine the catalytic performance and establish accurate structure–activity relationships. Our group utilized EC–AFM–TERS to monitor the evolution of the active sites of MoS2 during the HER process. It is interesting to mention that the variation in the EC–TERS spectra including the peak shift and relative intensity can only be observed under the reaction conditions (Fig. 23a and b). The potential-dependent TERS line-trace (Fig. 23c) over both the basal plane and edge site of MoS2 showed an obvious change in peak intensity from the basal plane to the edge site. The site-dependent variation in both electron density and the lattice deformation at different potentials can be visualized from the variation in the intensity of the peaks at 406 cm−1 (A1g, sensitive to the electron density) and 455 cm−1 (2LA, sensitive to the lattice deformation) (Fig. 23d). It was found that the edge site induces an ETR and lattice reconstruction region (LRR). After comparing these two regions obtained at different potentials, the whole evolution of the active site could be visualized (Fig. 23e). Before HER, the edge-induced an ETR of around 10 nm and LRR of around 2 nm in the pristine MoS2. During the activation process, the adsorption of H atoms on the edge sites will induce the lattice strain and electron density variation at the edge sites. After stabilization, the ETR extends to around 25 nm and the LRR extends to around 21 nm. This reconstruction induced by the HER process can be regarded as a self-optimization step, which improved the activity of the edge sites. Then, during the HER process after activation, the H atoms continuously adsorbed and desorbed at the edge sites, which led to the angle distortion in MoS2, and thus further increased the ETR and LRR to around 40 nm. This in situ characterization revealed the progressive generation and the dynamic variation of the active sites during the electrocatalysis, which provided new insights into designing highly active electrocatalysts.336


image file: d4cs00588k-f23.tif
Fig. 23 Monitoring the evolution of the active sites by EC–TERS. (a) EC–TERS spectra of the edge and basal plane before and during HER. (b) Ex situ TERS spectra of the edge and basal plane before and after HER. (c) EC–TERS line-trace over the edge of MoS2. (d) Plots of the peak intensity of A1g peak and 2LA(K–M) peak with the tip position. (e) Schematic illustration of the structural evolution of the active site at different stages. Reprinted with permission from ref. 336. Copyright 2024, Springer Nature.

It is interesting to note that similar to that under ambient condition, the surface plasmon (SP) in EC–TERS can also induce plasmon-driven photochemical reactions. Furthermore, by modulating the potential of the substrate, the energy of the photo-induced hot carriers can be easily modulated, which makes EC–TERS a flexible platform to study SP-based photoelectrochemical reactions at the electrochemical interface at the nanoscale. Recently, our group used EC–TERS to successfully visualize the spatial distribution of SP-induced hot carriers in real space (Fig. 24a).334 After the TERS tip approached the 4-mercaptobenzoic acid (4-MBA)-adsorbed Au(111) electrode, the feature peaks (998 cm−1 and 1020 cm−1) of TP were observed (Fig. 24b), indicating the conversion of 4-MBA molecules to TP. This irreversible decarboxylation reaction could be controlled by both the potential and SP (Fig. 24c and d), which makes it possible to turn on (before TERS imaging) or turn off (during TERS imaging) the reaction by modulating the potential, respectively. Subsequently, the SP-induced reaction region was visualized from the TERS line-trace when the reaction was off. After deconvolution of the spatial resolution and the initial distribution of the hot carriers, the transport distance of the reactive plasmonic hot carriers was measured to be around 20 nm in real space (Fig. 24e). In addition, a narrower distribution was observed at the negative potential (Fig. 24f), indicating the shorter transport distance of hot carriers with higher energies. With further improvements in sensitivity, we believe that EC–TERS will be able to address more interesting and challenging issues in electrochemical and photoelectrocatalytic systems.


image file: d4cs00588k-f24.tif
Fig. 24 Photoelectrochemical reaction probed by EC–TERS. (a) Schematic illustration of the EC–TERS setup used for the photoelectrochemical reaction. (b) Time-dependent EC–TERS spectra of 4-MBA in 0.1 M NaClO4. EC–TERS spectra recorded with different powers (c) and at different potentials (d). (e) Schematic of the generation and transport of hot carriers on a surface. (f) Profile of product obtained at different potentials. Reprinted with permission from ref. 334. Copyright 2020, Springer Nature.

6. Conclusion and perspective

After more than two decades since the first demonstration of TERS in 2000, great efforts have been devoted to making this powerful but challenging technique mature. The progress in theoretical aspects is beneficial to understand the origin of the strong enhancement and high spatial resolution of TERS, which guides experiments to find the optimized parameters for the best performance. Several methods have been developed to fabricate different TERS tips for characterizing different samples. It has become routine in most TERS labs to fabricate normal TERS tips, while the recently introduced nanofabrication method in tip fabrication makes the accurate fabrication of specially designed TERS tips possible. The TERS instruments have also become mature. Commercial ambient TERS and UHV–TERS instruments are available, while the glovebox-based TERS and EC–TERS instruments also experience fast development. These developments have synergistically boosted the TERS technique to become powerful nanoscopy, and thus broaden the applications of TERS in many research fields. However, there are still challenges to further broaden the application of TERS.

6.1. Design and fabrication of TERS tips with stronger enhancement

The TERS tip is the core of the TERS technique, where the tips need to be sharp to produce a morphology image with a high spatial resolution, while they also need to possess effective plasmonic activity to enhance the Raman signal. Although the development in the fabrication of TERS tips has already made it possible to design and fabricate tips with strong enhancement, the pursuit of higher enhancement never stops. With the reported tips, it is still quite challenging to characterize the surface of a bulk material, where the near-field Raman signal will be easily obscured in the large far-field background signal. In addition, it is also quite difficult to probe the reactive species on non-coinage metal surfaces (such as Pt and Pd) owing to the relatively low TERS enhancement on these surfaces as well as the low coverage of the reactive species. The design of highly active TERS tips from simulation, including the optimized radii of the curvature, cone angle and some special structures, combined with accurate nanofabrication strategy, will be significant to fabricate more powerful TERS tips.

6.2. High-speed TERS imaging

TERS imaging provides 2D Raman imaging of a sample at a nanoscale spatial resolution. Analysis of the image with different Raman modes can visualize the spatial distribution of different properties. However, owing to the small cross section of Raman scattering, a long-time (from several minutes to hours) is usually required to obtain a TERS image. This time-consuming TERS imaging not only enlarges the influence of thermal drift, leading to a mismatch between morphology and the TERS image but also prevents the application of TERS to investigate dynamic processes. In this case, the development of high-speed TERS imaging techniques will further broaden the applications of TERS. High-speed TERS imaging may be achieved by combining the improvement of sensitivity and the help of AI algorithms. A major improvement in sensitivity can be achieved by the design and fabrication of TERS tips with a much higher enhancement. In addition, the sensitivity can also be improved from an optical aspect by introducing radially polarized light,337 modulating the phase of the laser using a SLM147,148 or supporting the sample with a Fabry–Perot cavity.338 Some early attempts in the development of AI algorithm-based methods show the possibility to speed up TERS imaging without the loss of SNR.197

6.3. TERS indirect detection technique

Although the TERS technique has high sensitivity up to the single-molecule level, it is still quite challenging to comprehensively study reaction processes. This challenge is not only attributed to the small Raman cross section, the low coverage and the short lifetime of the reactive species, but also the fact that some important properties of the interface cannot be directly probed by TERS, such as interfacial pH. In this regard, TERS with indirect detection was developed, wherein molecules sensitive to the reactive species or properties (such as the pH) are adsorbed on the tip. As a result, the distribution and dynamic of the reactive species or properties on material or the interface can be measured by the change in the Raman signal of the probe molecules. Pioneering works have been conducted by the Ozaki group using MBA and pATP as the probe to visualized the nanoscale pH distribution at the liquid–solid interface.339 The core of TERS indirect detection is the choice of suitable probe molecules that are sensitive enough to the change in the target properties and are highly reversible. For example, using an achiral para-mercaptopyridine (pMPY)-modified tip, the Ozaki group recently distinguished the enantiomers according to the variation of the relative intensities of the pMPY molecules.340 In addition, the way to obtain a high spatial resolution needs to be carefully considered in TERS indirect detection given that all the molecules adsorbed on the tip may contribute to the final TERS signal. Furthermore, given that the migration of the reactive species over the interface always occurs, TERS indirect detection may obtain the distribution of the reactive species including the influence of migration, hindering the investigation of the nanoscale heterogeneity, and thus reducing the spatial resolution.

6.4. Characterization of the materials and interface under the working condition

With the fast development of in situ characterization techniques in recent years, it is generally recognized that the structure and properties of materials and interfaces experience dynamic evolution under the working conditions.12 Therefore, it will be significant to perform operando TERS characterization to visualize the changes in materials and interfaces. Although glovebox-based TERS or EC–TERS can provide a controlled environment for in situ characterization, we believe some further developments in optical path and the design of cells will achieve operando TERS characterization in the future.

6.5. Correlation of the activity and the property of a single active site

Establishing accurate structure–activity relationships at the single-site level is one of the goals in catalysis, which will guide the design of highly active catalysts. The TERS technique, especially EC–TERS can simultaneously obtain the topography and chemical information with nanoscale spatial resolution under the reaction conditions. However, there is still a lack of information on the electrochemical activity of individual sites in TERS, which makes it difficult to determine the real active site. If TERS can be further combined with some techniques that can obtain nano-resolved activity, such as scanning electrochemical microscopy (SECM),341 a systematic understanding of the material or the interface will be obtained to establish an accurate structure–activity relationship. However, a lot of challenges need to be solved to achieve this goal, including the design of functional TERS tips, which can be used for EC-SPM, EC–TERS and SECM, and the design of special working modes to obtain all the information from the same site.

6.6. Characterization with a high spatiotemporal resolution

The TERS technique has enabled the characterization of materials and interfaces with a high spatial resolution up to the chemical bond level. However, the temporal resolution of TERS is still in the time scale of seconds. If the temporal resolution can be improved to the picosecond (ps) or femtosecond (fs) level, it will enable the visualization of the dynamics of some physical or chemical processes with a nanoscale spatial resolution. To achieve this goal, the complicated ultra-fast optical system needs to be well coupled with the TERS system without interfering with the stability of the TERS instrument. Some early attempts have been done in either the ambient TERS system342,343 or UHV–TERS system,344,345 demonstrating the successful measurement of TERS with ultrashort laser pulses. We believe that in the future, after solving the challenges in the stability (the tip, the plasmonic gap and the sample) as well as the strong background signal, which comes from the metallic tip and substrate,346 TERS measurement with a high spatiotemporal resolution will be available. There is no doubt that this development will deepen the understanding of physical and chemical processes.

Data availability

No primary research results, software or code have been included and no new data were generated or analyzed as part of this review.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors acknowledge the financial supports from the National Natural Science Foundation of China (Grant No: 22227802, 22021001, 22393901 and 22372141); the Fundamental Research Funds for the Central Universities (20720220018).

References

  1. R. Vajtai, Springer Handbook of Nanomaterials, Springer-Verlag, Berlin, 2013 Search PubMed .
  2. D. Vollath, Nanomaterials: An Introduction to Synthesis, Properties and Applications, Wiley-VCH, Weinheim, 2013 Search PubMed .
  3. M. Liu, N. Yazdani, M. Yarema, M. Jansen, V. Wood and E. H. Sargent, Nat. Electron., 2021, 4, 548–558 CrossRef .
  4. B. Garlyyev, J. Fichtner, O. Piqué, O. Schneider, A. S. Bandarenka and F. Calle-Vallejo, Chem. Sci., 2019, 10, 8060–8075 RSC .
  5. J. T. Mefford, A. R. Akbashev, M. Kang, C. L. Bentley, W. E. Gent, H. D. Deng, D. H. Alsem, Y.-S. Yu, N. J. Salmon, D. A. Shapiro, P. R. Unwin and W. C. Chueh, Nature, 2021, 593, 67–73 CrossRef CAS PubMed .
  6. C. Vogt and B. M. Weckhuysen, Nat. Rev. Chem., 2022, 6, 89–111 CrossRef PubMed .
  7. Z. W. Seh, J. Kibsgaard, C. F. Dickens, I. Chorkendorff, J. K. Nørskov and T. F. Jaramillo, Science, 2017, 355, eaad4998 CrossRef PubMed .
  8. E. Pastor, Z. Lian, L. Xia, D. Ecija, J. R. Galán-Mascarós, S. Barja, S. Giménez, J. Arbiol, N. López and F. P. García de Arquer, Nat. Rev. Chem., 2024, 8, 159–178 CrossRef PubMed .
  9. H. Frey, A. Beck, X. Huang, J. A. van Bokhoven and M. G. Willinger, Science, 2022, 376, 982–987 CrossRef CAS PubMed .
  10. X. Sun, D. Wu, L. Zou, S. D. House, X. Chen, M. Li, D. N. Zakharov, J. C. Yang and G. Zhou, Nature, 2022, 607, 708–713 CrossRef CAS PubMed .
  11. S. Zhou, J. Shi, S. Liu, G. Li, F. Pei, Y. Chen, J. Deng, Q. Zheng, J. Li, C. Zhao, I. Hwang, C.-J. Sun, Y. Liu, Y. Deng, L. Huang, Y. Qiao, G.-L. Xu, J.-F. Chen, K. Amine, S.-G. Sun and H.-G. Liao, Nature, 2023, 621, 75–81 CrossRef CAS PubMed .
  12. Y. Yang, S. Louisia, S. Yu, J. Jin, I. Roh, C. Chen, M. V. Fonseca Guzman, J. Feijóo, P.-C. Chen, H. Wang, C. J. Pollock, X. Huang, Y.-T. Shao, C. Wang, D. A. Muller, H. D. Abruña and P. Yang, Nature, 2023, 614, 262–269 CrossRef CAS PubMed .
  13. X. Yan, C. Liu, C. A. Gadre, L. Gu, T. Aoki, T. C. Lovejoy, N. Dellby, O. L. Krivanek, D. G. Schlom, R. Wu and X. Pan, Nature, 2021, 589, 65–69 CrossRef CAS PubMed .
  14. C. A. Gadre, X. Yan, Q. Song, J. Li, L. Gu, H. Huyan, T. Aoki, S.-W. Lee, G. Chen, R. Wu and X. Pan, Nature, 2022, 606, 292–297 CrossRef CAS PubMed .
  15. S. Yoshimoto and K. Itaya, Annu. Rev. Anal. Chem., 2013, 6, 213–235 CrossRef CAS PubMed .
  16. J. Wei, R. Amirbeigiarab, Y.-X. Chen, S. Sakong, A. Gross and O. M. Magnussen, Angew. Chem., Int. Ed., 2020, 59, 6182–6186 CrossRef CAS PubMed .
  17. X. Wang, Y.-Q. Wang, Y.-C. Feng, D. Wang and L.-J. Wan, Chem. Soc. Rev., 2021, 50, 5832–5849 RSC .
  18. R. Amirbeigiarab, J. Tian, A. Herzog, C. Qiu, A. Bergmann, B. Roldan Cuenya and O. M. Magnussen, Nat. Catal., 2023, 6, 837–846 CrossRef CAS .
  19. Y. Liu, F. Yang, Y. Zhang, J. Xiao, L. Yu, Q. Liu, Y. Ning, Z. Zhou, H. Chen, W. Huang, P. Liu and X. Bao, Nat. Commun., 2017, 8, 14459 CrossRef CAS PubMed .
  20. Y.-Q. Wang, X.-H. Dan, X. Wang, Z.-Y. Yi, J. Fu, Y.-C. Feng, J.-S. Hu, D. Wang and L.-J. Wan, J. Am. Chem. Soc., 2022, 144, 20126–20133 CrossRef CAS PubMed .
  21. J. H. K. Pfisterer, Y. Liang, O. Schneider and A. S. Bandarenka, Nature, 2017, 549, 74–77 CrossRef CAS PubMed .
  22. F. J. Giessibl, Rev. Mod. Phys., 2003, 75, 949–983 CrossRef CAS .
  23. L. Gross, F. Mohn, N. Moll, P. Liljeroth and G. Meyer, Science, 2009, 325, 1110–1114 CrossRef CAS PubMed .
  24. J. Zhang, P. Chen, B. Yuan, W. Ji, Z. Cheng and X. Qiu, Science, 2013, 342, 611–614 CrossRef CAS PubMed .
  25. R. Ma, D. Cao, C. Zhu, Y. Tian, J. Peng, J. Guo, J. Chen, X.-Z. Li, J. S. Francisco, X. C. Zeng, L.-M. Xu, E.-G. Wang and Y. Jiang, Nature, 2020, 577, 60–63 CrossRef CAS PubMed .
  26. Y. Tian, J. Hong, D. Cao, S. You, Y. Song, B. Cheng, Z. Wang, D. Guan, X. Liu, Z. Zhao, X.-Z. Li, L.-M. Xu, J. Guo, J. Chen, E.-G. Wang and Y. Jiang, Science, 2022, 377, 315–319 CrossRef CAS PubMed .
  27. O. M. Magnussen, J. Drnec, C. Qiu, I. Martens, J. J. Huang, R. Chattot and A. Singer, Chem. Rev., 2024, 124, 629–721 CrossRef CAS PubMed .
  28. Y. Zheng, Y. Jiao, Y. Zhu, L. H. Li, Y. Han, Y. Chen, A. Du, M. Jaroniec and S. Z. Qiao, Nat. Commun., 2014, 5, 3783 CrossRef PubMed .
  29. A. Zitolo, V. Goellner, V. Armel, M.-T. Sougrati, T. Mineva, L. Stievano, E. Fonda and F. Jaouen, Nat. Mater., 2015, 14, 937–942 CrossRef CAS PubMed .
  30. D. Ren, B. S.-H. Ang and B. S. Yeo, ACS Catal., 2016, 6, 8239–8247 CrossRef CAS .
  31. W. Karim, C. Spreafico, A. Kleibert, J. Gobrecht, J. VandeVondele, Y. Ekinci and J. A. van Bokhoven, Nature, 2017, 541, 68–71 CrossRef CAS PubMed .
  32. Y. Zhou, F. Che, M. Liu, C. Zou, Z. Liang, P. De Luna, H. Yuan, J. Li, Z. Wang, H. Xie, H. Li, P. Chen, E. Bladt, R. Quintero-Bermudez, T.-K. Sham, S. Bals, J. Hofkens, D. Sinton, G. Chen and E. H. Sargent, Nat. Chem., 2018, 10, 974–980 CrossRef CAS PubMed .
  33. D. Wang, Q. Li, C. Han, Q. Lu, Z. Xing and X. Yang, Nat. Commun., 2019, 10, 3899 CrossRef PubMed .
  34. T. M. Ajayi, N. Shirato, T. Rojas, S. Wieghold, X. Cheng, K. Z. Latt, D. J. Trainer, N. K. Dandu, Y. Li, S. Premarathna, S. Sarkar, D. Rosenmann, Y. Liu, N. Kyritsakas, S. Wang, E. Masson, V. Rose, X. Li, A. T. Ngo and S.-W. Hla, Nature, 2023, 618, 69–73 CrossRef CAS PubMed .
  35. W. Ma, P. Alonso-González, S. Li, A. Y. Nikitin, J. Yuan, J. Martín-Sánchez, J. Taboada-Gutiérrez, I. Amenabar, P. Li, S. Vélez, C. Tollan, Z. Dai, Y. Zhang, S. Sriram, K. Kalantar-Zadeh, S.-T. Lee, R. Hillenbrand and Q. Bao, Nature, 2018, 562, 557–562 CrossRef CAS PubMed .
  36. S. Zhang, B. Li, X. Chen, F. L. Ruta, Y. Shao, A. J. Sternbach, A. S. McLeod, Z. Sun, L. Xiong, S. L. Moore, X. Xu, W. Wu, S. Shabani, L. Zhou, Z. Wang, F. Mooshammer, E. Ray, N. Wilson, P. J. Schuck, C. R. Dean, A. N. Pasupathy, M. Lipson, X. Xu, X. Zhu, A. J. Millis, M. Liu, J. C. Hone and D. N. Basov, Nat. Commun., 2022, 13, 542 CrossRef CAS PubMed .
  37. T. L. Cocker, V. Jelic, R. Hillenbrand and F. A. Hegmann, Nat. Photonics, 2021, 15, 558–569 CrossRef CAS .
  38. S. Chen, A. Bylinkin, Z. Wang, M. Schnell, G. Chandan, P. Li, A. Y. Nikitin, S. Law and R. Hillenbrand, Nat. Commun., 2022, 13, 1374 CrossRef CAS PubMed .
  39. S. C. M. Reinhardt, L. A. Masullo, I. Baudrexel, P. R. Steen, R. Kowalewski, A. S. Eklund, S. Strauss, E. M. Unterauer, T. Schlichthaerle, M. T. Strauss, C. Klein and R. Jungmann, Nature, 2023, 617, 711–716 CrossRef CAS PubMed .
  40. B. Dong, Y. Pei, N. Mansour, X. Lu, K. Yang, W. Huang and N. Fang, Nat. Commun., 2019, 10, 4815 CrossRef PubMed .
  41. B. Dong, Y. Pei, F. Zhao, T. W. Goh, Z. Qi, C. Xiao, K. Chen, W. Huang and N. Fang, Nat. Catal., 2018, 1, 135–140 CrossRef .
  42. T.-X. Huang, B. Dong, S. L. Filbrun, A. A. Okmi, X. Cheng, M. Yang, N. Mansour, S. Lei and N. Fang, Sci. Adv., 2021, 7, eabj4452 CrossRef CAS PubMed .
  43. G. E. Walrafen, M. R. Fisher, M. S. Hokmabadi and W. H. Yang, J. Chem. Phys., 1986, 85, 6970–6982 CrossRef CAS .
  44. T. Fukasawa, T. Sato, J. Watanabe, Y. Hama, W. Kunz and R. Buchner, Phys. Rev. Lett., 2005, 95, 197802 CrossRef PubMed .
  45. Y. Amo and Y. Tominaga, Chem. Phys. Lett., 2000, 320, 703–706 CrossRef CAS .
  46. M. Ziemba, C. Schilling, M. V. Ganduglia-Pirovano and C. Hess, Acc. Chem. Res., 2021, 54, 2884–2893 CrossRef CAS PubMed .
  47. X. Zhang, X.-F. Qiao, W. Shi, J.-B. Wu, D.-S. Jiang and P.-H. Tan, Chem. Soc. Rev., 2015, 44, 2757–2785 RSC .
  48. A. C. Ferrari and D. M. Basko, Nat. Nanotechnol., 2013, 8, 235–246 CrossRef CAS PubMed .
  49. P. H. Tan, W. P. Han, W. J. Zhao, Z. H. Wu, K. Chang, H. Wang, Y. F. Wang, N. Bonini, N. Marzari, N. Pugno, G. Savini, A. Lombardo and A. C. Ferrari, Nat. Mater., 2012, 11, 294–300 CrossRef CAS PubMed .
  50. H. Wu, M.-L. Lin, Y.-C. Leng, X. Chen, Y. Zhou, J. Zhang and P.-H. Tan, npj 2D Mater. Appl., 2022, 6, 87 CrossRef CAS .
  51. Z.-Q. Tian, B. Ren and D.-Y. Wu, J. Phys. Chem. B, 2002, 106, 9463–9483 CrossRef CAS .
  52. S. C. Chang and M. J. Weaver, J. Chem. Phys., 1990, 92, 4582–4594 CrossRef CAS .
  53. Z.-Q. Tian and B. Ren, Annu. Rev. Phys. Chem., 2004, 55, 197–229 CrossRef CAS PubMed .
  54. M. C. Figueiredo, I. Ledezma-Yanez and M. T. M. Koper, ACS Catal., 2016, 6, 2382–2392 CrossRef CAS .
  55. C. Zhan, F. Dattila, C. Rettenmaier, A. Bergmann, S. Kühl, R. García-Muelas, N. López and B. R. Cuenya, ACS Catal., 2021, 11, 7694–7701 CrossRef CAS PubMed .
  56. M. Su, J.-C. Dong, J.-B. Le, Y. Zhao, W.-M. Yang, Z.-L. Yang, G. Attard, G.-K. Liu, J. Cheng, Y.-M. Wei, Z.-Q. Tian and J.-F. Li, Angew. Chem., Int. Ed., 2020, 59, 23554–23558 CrossRef CAS PubMed .
  57. Z. Lu, Y. Huang, N. Chen, C. Liu, X. Wang and B. Ren, ACS Catal., 2024, 14, 6204–6210 CrossRef CAS .
  58. E. C. Le Ru, E. Blackie, M. Meyer and P. G. Etchegoin, J. Phys. Chem. C, 2007, 111, 13794–13803 CrossRef CAS .
  59. P. G. Etchegoin and E. C. Le Ru, Phys. Chem. Chem. Phys., 2008, 10, 6079–6089 RSC .
  60. M. Fleischmann, P. J. Hendra and A. J. McQuillan, Chem. Phys. Lett., 1974, 26, 163–166 CrossRef CAS .
  61. D. L. Jeanmaire and R. P. Van Duyne, J. Electroanal. Chem., 1977, 84, 1–20 CrossRef CAS .
  62. M. G. Albrecht and J. A. Creighton, J. Am. Chem. Soc., 1977, 99, 5215–5217 CrossRef CAS .
  63. K. Kneipp, Y. Wang, H. Kneipp, L. T. Perelman, I. Itzkan, R. R. Dasari and M. S. Feld, Phys. Rev. Lett., 1997, 78, 1667–1670 CrossRef CAS .
  64. S. Nie and S. R. Emory, Science, 1997, 275, 1102–1106 CrossRef CAS PubMed .
  65. M. Moskovits, J. Chem. Phys., 1978, 69, 4159–4161 CrossRef CAS .
  66. S. Laing, L. E. Jamieson, K. Faulds and D. Graham, Nat. Rev. Chem., 2017, 1, 0060 CrossRef CAS .
  67. D. Cialla-May, X. S. Zheng, K. Weber and J. Popp, Chem. Soc. Rev., 2017, 46, 3945–3961 RSC .
  68. C. Zong, M. Xu, L.-J. Xu, T. Wei, X. Ma, X.-S. Zheng, R. Hu and B. Ren, Chem. Rev., 2018, 118, 4946–4980 CrossRef CAS PubMed .
  69. Y.-H. Wang, S. Zheng, W.-M. Yang, R.-Y. Zhou, Q.-F. He, P. Radjenovic, J.-C. Dong, S. Li, J. Zheng, Z.-L. Yang, G. Attard, F. Pan, Z.-Q. Tian and J.-F. Li, Nature, 2021, 600, 81–85 CrossRef CAS PubMed .
  70. H. An, L. Wu, L. D. B. Mandemaker, S. Yang, J. de Ruiter, J. H. J. Wijten, J. C. L. Janssens, T. Hartman, W. van der Stam and B. M. Weckhuysen, Angew. Chem., Int. Ed., 2021, 60, 16576–16584 CrossRef CAS PubMed .
  71. Y. Gu, E.-M. You, J.-D. Lin, J.-H. Wang, S.-H. Luo, R.-Y. Zhou, C.-J. Zhang, J.-L. Yao, H.-Y. Li, G. Li, W.-W. Wang, Y. Qiao, J.-W. Yan, D.-Y. Wu, G.-K. Liu, L. Zhang, J.-F. Li, R. Xu, Z.-Q. Tian, Y. Cui and B.-W. Mao, Nat. Commun., 2023, 14, 3536 CrossRef CAS PubMed .
  72. M. Fan, G. F. S. Andrade and A. G. Brolo, Anal. Chim. Acta, 2020, 1097, 1–29 CrossRef CAS PubMed .
  73. H.-L. Wang, E.-M. You, R. Panneerselvam, S.-Y. Ding and Z.-Q. Tian, Light: Sci. Appl., 2021, 10, 161 CrossRef CAS PubMed .
  74. T. Itoh, M. Procházka, Z.-C. Dong, W. Ji, Y. S. Yamamoto, Y. Zhang and Y. Ozaki, Chem. Rev., 2023, 123, 1552–1634 CrossRef CAS PubMed .
  75. M. S. Anderson, Appl. Phys. Lett., 2000, 76, 3130–3132 CrossRef CAS .
  76. R. M. Stöckle, Y. D. Suh, V. Deckert and R. Zenobi, Chem. Phys. Lett., 2000, 318, 131–136 CrossRef .
  77. N. Hayazawa, Y. Inouye, Z. Sekkat and S. Kawata, Opt. Commun., 2000, 183, 333–336 CrossRef CAS .
  78. B. Pettinger, G. Picardi, R. Schuster and G. Ertl, Electrochemistry, 2000, 68, 942–949 CrossRef CAS .
  79. S. Sheng, J.-b Wu, X. Cong, W. Li, J. Gou, Q. Zhong, P. Cheng, P.-h Tan, L. Chen and K. Wu, Phys. Rev. Lett., 2017, 119, 196803 CrossRef PubMed .
  80. W. Su, N. Kumar, A. Krayev and M. Chaigneau, Nat. Commun., 2018, 9, 2891 CrossRef PubMed .
  81. T.-X. Huang, X. Cong, S.-S. Wu, K.-Q. Lin, X. Yao, Y.-H. He, J.-B. Wu, Y.-F. Bao, S.-C. Huang, X. Wang, P.-H. Tan and B. Ren, Nat. Commun., 2019, 10, 5544 CrossRef CAS PubMed .
  82. A. C. Gadelha, D. A. A. Ohlberg, C. Rabelo, E. G. S. Neto, T. L. Vasconcelos, J. L. Campos, J. S. Lemos, V. Ornelas, D. Miranda, R. Nadas, F. C. Santana, K. Watanabe, T. Taniguchi, B. van Troeye, M. Lamparski, V. Meunier, V.-H. Nguyen, D. Paszko, J.-C. Charlier, L. C. Campos, L. G. Cançado, G. Medeiros-Ribeiro and A. Jorio, Nature, 2021, 590, 405–409 CrossRef CAS PubMed .
  83. J.-H. Zhong, X. Jin, L. Meng, X. Wang, H.-S. Su, Z.-L. Yang, C. T. Williams and B. Ren, Nat. Nanotechnol., 2017, 12, 132–136 CrossRef CAS PubMed .
  84. J. H. K. Pfisterer, M. Baghernejad, G. Giuzio and K. F. Domke, Nat. Commun., 2019, 10, 5702 CrossRef CAS PubMed .
  85. H. Yin, L.-Q. Zheng, W. Fang, Y.-H. Lai, N. Porenta, G. Goubert, H. Zhang, H.-S. Su, B. Ren, J. O. Richardson, J.-F. Li and R. Zenobi, Nat. Catal., 2020, 3, 834–842 CrossRef CAS .
  86. L. Li, J. F. Schultz, S. Mahapatra, Z. Lu, X. Zhang and N. Jiang, Nat. Commun., 2022, 13, 1796 CrossRef CAS PubMed .
  87. P. Verma, Chem. Rev., 2017, 117, 6447–6466 CrossRef CAS PubMed .
  88. M. Richard-Lacroix, Y. Zhang, Z. Dong and V. Deckert, Chem. Soc. Rev., 2017, 46, 3922–3944 RSC .
  89. T. Deckert-Gaudig, A. Taguchi, S. Kawata and V. Deckert, Chem. Soc. Rev., 2017, 46, 4077–4110 RSC .
  90. X. Wang, S.-C. Huang, T.-X. Huang, H.-S. Su, J.-H. Zhong, Z.-C. Zeng, M.-H. Li and B. Ren, Chem. Soc. Rev., 2017, 46, 4020–4041 RSC .
  91. Z.-F. Cai, N. Kumar and R. Zenobi, CCS Chem., 2023, 5, 55–71 CrossRef CAS .
  92. P. Pienpinijtham, Y. Kitahama and Y. Ozaki, Nanoscale, 2022, 14, 5265–5288 RSC .
  93. N. Kumar, B. M. Weckhuysen, A. J. Wain and A. J. Pollard, Nat. Protoc., 2019, 14, 1169–1193 CrossRef CAS PubMed .
  94. C. Höppener, J. Aizpurua, H. Chen, S. Gräfe, A. Jorio, S. Kupfer, Z. Zhang and V. Deckert, Nat. Rev. Methods Primers, 2024, 4, 47 CrossRef .
  95. Y. Fujita, P. Walke, S. De Feyter and H. Uji-i, Jpn. J. Appl. Phys., 2016, 55, 08NA02 CrossRef .
  96. X. Shi, N. Coca-López, J. Janik and A. Hartschuh, Chem. Rev., 2017, 117, 4945–4960 CrossRef CAS PubMed .
  97. T. L. Vasconcelos, B. S. Archanjo, B. S. Oliveira, W. F. Silva, R. S. Alencar, C. Rabelo, C. A. Achete, A. Jorio and L. G. Cançado, IEEE J. Sel. Top. Quantum Electron., 2021, 27, 1–11 Search PubMed .
  98. S. Bonhommeau, G. S. Cooney and Y. Huang, Chem. Soc. Rev., 2022, 51, 2416–2430 RSC .
  99. T. Stepanenko, K. Sofińska, N. Wilkosz, J. Dybas, E. Wiercigroch, K. Bulat, E. Szczesny-Malysiak, K. Skirlińska-Nosek, S. Seweryn, J. Chwiej, E. Lipiec and K. M. Marzec, Analyst, 2024, 149, 778–788 RSC .
  100. E. A. Pozzi, G. Goubert, N. Chiang, N. Jiang, C. T. Chapman, M. O. McAnally, A.-I. Henry, T. Seideman, G. C. Schatz, M. C. Hersam and R. P. V. Duyne, Chem. Rev., 2017, 117, 4961–4982 CrossRef CAS PubMed .
  101. T. Hartman, C. S. Wondergem, N. Kumar, A. van den Berg and B. M. Weckhuysen, J. Phys. Chem. Lett., 2016, 7, 1570–1584 CrossRef CAS PubMed .
  102. C. Hess, Chem. Soc. Rev., 2021, 50, 3519–3564 RSC .
  103. A. B. Zrimsek, N. Chiang, M. Mattei, S. Zaleski, M. O. McAnally, C. T. Chapman, A.-I. Henry, G. C. Schatz and R. P. Van Duyne, Chem. Rev., 2017, 117, 7583–7613 CrossRef CAS PubMed .
  104. S. Zaleski, A. J. Wilson, M. Mattei, X. Chen, G. Goubert, M. F. Cardinal, K. A. Willets and R. P. Van Duyne, Acc. Chem. Res., 2016, 49, 2023–2030 CrossRef CAS PubMed .
  105. J. H. K. Pfisterer and K. F. Domke, Curr. Opin. Electrochem., 2018, 8, 96–102 CrossRef CAS .
  106. S.-C. Huang, Y.-F. Bao, S.-S. Wu, T.-X. Huang, M. M. Sartin, X. Wang and B. Ren, Annu. Rev. Phys. Chem., 2021, 72, 213–234 CrossRef CAS PubMed .
  107. Z. Yang, J. Aizpurua and H. Xu, J. Raman Spectrosc., 2009, 40, 1343–1348 CrossRef CAS .
  108. S. F. Becker, M. Esmann, K. Yoo, P. Gross, R. Vogelgesang, N. Park and C. Lienau, ACS Photonics, 2016, 3, 223–232 CrossRef CAS .
  109. W. Zhu, R. Esteban, A. G. Borisov, J. J. Baumberg, P. Nordlander, H. J. Lezec, J. Aizpurua and K. B. Crozier, Nat. Commun., 2016, 7, 11495 CrossRef CAS PubMed .
  110. I. Romero, J. Aizpurua, G. W. Bryant and F. J. G. D. Abajo, Opt. Express, 2006, 14, 9988–9999 CrossRef PubMed .
  111. C. Chen, N. Hayazawa and S. Kawata, Nat. Commun., 2014, 5, 3312 CrossRef PubMed .
  112. M. Barbry, P. Koval, F. Marchesin, R. Esteban, A. G. Borisov, J. Aizpurua and D. Sánchez-Portal, Nano Lett., 2015, 15, 3410–3419 CrossRef CAS PubMed .
  113. J.-Z. Zhu, G. Chen, T. Ijaz, X.-G. Li and Z.-C. Dong, J. Chem. Phys., 2021, 154, 214706 CrossRef CAS PubMed .
  114. B. Yang, G. Chen, A. Ghafoor, Y. Zhang, Y. Zhang, Y. Zhang, Y. Luo, J. Yang, V. Sandoghdar, J. Aizpurua, Z. Dong and J. G. Hou, Nat. Photonics, 2020, 14, 693–699 CrossRef CAS .
  115. F. Benz, M. K. Schmidt, A. Dreismann, R. Chikkaraddy, Y. Zhang, A. Demetriadou, C. Carnegie, H. Ohadi, B. de Nijs, R. Esteban, J. Aizpurua and J. J. Baumberg, Science, 2016, 354, 726–729 CrossRef CAS PubMed .
  116. X. Chen, P. Liu, Z. Hu and L. Jensen, Nat. Commun., 2019, 10, 2567 CrossRef PubMed .
  117. J. Lee, K. T. Crampton, N. Tallarida and V. A. Apkarian, Nature, 2019, 568, 78–82 CrossRef CAS PubMed .
  118. Y. Zhang, B. Yang, A. Ghafoor, Y. Zhang, Y.-F. Zhang, R.-P. Wang, J.-L. Yang, Y. Luo, Z.-C. Dong and J. G. Hou, Natl. Sci. Rev., 2019, 6, 1169–1175 CrossRef CAS PubMed .
  119. S. Thomas, M. Krüger, M. Förster, M. Schenk and P. Hommelhoff, Nano Lett., 2013, 13, 4790–4794 CrossRef CAS PubMed .
  120. A. Taguchi, J. Yu, P. Verma and S. Kawata, Nanoscale, 2015, 7, 17424–17433 RSC .
  121. R. J. Hermann and M. J. Gordon, Opt. Express, 2018, 26, 27668–27682 CrossRef CAS PubMed .
  122. T. L. Vasconcelos, B. S. Archanjo, B. Fragneaud, B. S. Oliveira, J. Riikonen, C. Li, D. S. Ribeiro, C. Rabelo, W. N. Rodrigues, A. Jorio, C. A. Achete and L. G. Cançado, ACS Nano, 2015, 9, 6297–6304 CrossRef CAS PubMed .
  123. T. L. Vasconcelos, B. S. Archanjo, B. S. Oliveira, R. Valaski, R. C. Cordeiro, H. G. Medeiros, C. Rabelo, A. Ribeiro, P. Ercius, C. A. Achete, A. Jorio and L. G. Cançado, Adv. Opt. Mater., 2018, 6, 1800528 CrossRef .
  124. H. Miranda, C. Rabelo, T. L. Vasconcelos, L. G. Cançado and A. Jorio, Phys. Status Solidi RRL, 2020, 14, 2000212 CrossRef CAS .
  125. S. Berweger, J. M. Atkin, R. L. Olmon and M. B. Raschke, J. Phys. Chem. Lett., 2010, 1, 3427–3432 CrossRef CAS .
  126. T. Umakoshi, Y. Saito and P. Verma, Nanoscale, 2016, 8, 5634–5640 RSC .
  127. N. A. Issa and R. Guckenberger, Plasmonics, 2007, 2, 31–37 CrossRef CAS .
  128. D. K. Gramotnev, M. W. Vogel and M. I. Stockman, J. Appl. Phys., 2008, 104, 034311 CrossRef .
  129. P. Nagpal, N. C. Lindquist, S.-H. Oh and D. J. Norris, Science, 2009, 325, 594–597 CrossRef CAS PubMed .
  130. S. Kim, N. Yu, X. Ma, Y. Zhu, Q. Liu, M. Liu and R. Yan, Nat. Photonics, 2019, 13, 636–643 CrossRef CAS .
  131. K. Zhang, Y. Bao, M. Cao, S.-I. Taniguchi, M. Watanabe, T. Kambayashi, T. Okamoto, M. Haraguchi, X. Wang, K. Kobayashi, H. Yamada, B. Ren and T. Tachizaki, Anal. Chem., 2021, 93, 7699–7706 CrossRef CAS PubMed .
  132. X. Ma, Y. Zhu, N. Yu, S. Kim, Q. Liu, L. Apontti, D. Xu, R. Yan and M. Liu, Nano Lett., 2019, 19, 100–107 CrossRef CAS PubMed .
  133. W. Peeters, S. Toyouchi, Y. Fujita, M. Wolf, B. Fortuni, E. Fron, T. Inose, J. Hofkens, T. Endo, Y. Miyata and H. Uji-i, ACS Omega, 2023, 8, 38386–38393 CrossRef CAS PubMed .
  134. H. Watanabe, N. Hayazawa, Y. Inouye and S. Kawata, J. Phys. Chem. B, 2005, 109, 5012–5020 CrossRef CAS PubMed .
  135. M. Sun, Y. Fang, Z. Yang and H. Xu, Phys. Chem. Chem. Phys., 2009, 11, 9412–9419 RSC .
  136. P. Singh, T. Deckert-Gaudig, H. Schneidewind, K. Kirsch, E. M. van Schrojenstein Lantman, B. M. Weckhuysen and V. Deckert, Phys. Chem. Chem. Phys., 2015, 17, 2991–2995 RSC .
  137. R. L. M. Gieseking, J. Lee, N. Tallarida, V. A. Apkarian and G. C. Schatz, J. Phys. Chem. Lett., 2018, 9, 3074–3080 CrossRef CAS PubMed .
  138. S. Liu, B. Cirera, Y. Sun, I. Hamada, M. Müller, A. Hammud, M. Wolf and T. Kumagai, Nano Lett., 2020, 20, 5879–5884 CrossRef CAS PubMed .
  139. S. Liu, A. Hammud, M. Wolf and T. Kumagai, Nano Lett., 2021, 21, 4057–4061 CrossRef CAS PubMed .
  140. B. Cirera, Y. Litman, C. Lin, A. Akkoush, A. Hammud, M. Wolf, M. Rossi and T. Kumagai, Nano Lett., 2022, 22, 2170–2176 CrossRef CAS PubMed .
  141. B. Yang, G. Chen, A. Ghafoor, Y.-F. Zhang, X.-B. Zhang, H. Li, X.-R. Dong, R.-P. Wang, Y. Zhang, Y. Zhang and Z.-C. Dong, Angew. Chem., Int. Ed., 2023, 62, e202218799 CrossRef CAS PubMed .
  142. N. Kazemi-Zanjani, S. Vedraine and F. Lagugné-Labarthet, Opt. Express, 2013, 21, 25271–25276 CrossRef PubMed .
  143. J. Mihaljevic, C. Hafner and A. J. Meixner, Opt. Express, 2013, 21, 25926–25943 CrossRef PubMed .
  144. J. Stadler, C. Stanciu, C. Stupperich and A. J. Meixner, Opt. Lett., 2008, 33, 681–683 CrossRef CAS PubMed .
  145. T. Touzalin, S. Joiret, I. T. Lucas and E. Maisonhaute, Electrochem. Commun., 2019, 108, 106557 CrossRef CAS .
  146. R. Kato, T. Moriyama, T. Umakoshi, T.-A. Yano and P. Verma, Sci. Adv., 2022, 8, eabo4021 CrossRef CAS PubMed .
  147. D. Y. Lee, C. Park, J. Choi, Y. Koo, M. Kang, M. S. Jeong, M. B. Raschke and K.-D. Park, Nat. Commun., 2021, 12, 3465 CrossRef CAS PubMed .
  148. G. Giuzio, N. Martín Sabanés and K. F. Domke, Appl. Spectrosc., 2020, 74, 1407–1413 CrossRef CAS PubMed .
  149. J. Steidtner and B. Pettinger, Rev. Sci. Instrum., 2007, 78, 103104 CrossRef PubMed .
  150. R. Zhang, Y. Zhang, Z. C. Dong, S. Jiang, C. Zhang, L. G. Chen, L. Zhang, Y. Liao, J. Aizpurua, Y. Luo, J. L. Yang and J. G. Hou, Nature, 2013, 498, 82–86 CrossRef CAS PubMed .
  151. P. J. Whiteman, J. F. Schultz, Z. D. Porach, H. Chen and N. Jiang, J. Phys. Chem. C, 2018, 122, 5489–5495 CrossRef CAS .
  152. R. B. Jaculbia, H. Imada, K. Miwa, T. Iwasa, M. Takenaka, B. Yang, E. Kazuma, N. Hayazawa, T. Taketsugu and Y. Kim, Nat. Nanotechnol., 2020, 15, 105–110 CrossRef CAS PubMed .
  153. N. Jiang, E. T. Foley, J. M. Klingsporn, M. D. Sonntag, N. A. Valley, J. A. Dieringer, T. Seideman, G. C. Schatz, M. C. Hersam and R. P. Van Duyne, Nano Lett., 2012, 12, 5061–5067 CrossRef CAS PubMed .
  154. R.-P. Wang, B. Yang, Q. Fu, Y. Zhang, R. Zhu, X.-R. Dong, Y. Zhang, B. Wang, J.-L. Yang, Y. Luo, Z.-C. Dong and J. G. Hou, J. Phys. Chem. Lett., 2021, 12, 1961–1968 CrossRef CAS PubMed .
  155. J. Nanda, G. Yang, T. Hou, D. N. Voylov, X. Li, R. E. Ruther, M. Naguib, K. Persson, G. M. Veith and A. P. Sokolov, Joule, 2019, 3, 2001–2019 CrossRef CAS .
  156. S. Dinda, S. Trivedi, A. Roy, F. D. Pammer and M. Fichtner, Adv. Energy Mater., 2023, 13, 2302176 CrossRef CAS .
  157. Z.-C. Zeng, S. Hu, S.-C. Huang, Y.-J. Zhang, W.-X. Zhao, J.-F. Li, C. Jiang and B. Ren, Anal. Chem., 2016, 88, 9381–9385 CrossRef CAS PubMed .
  158. Z.-C. Zeng, S.-C. Huang, D.-Y. Wu, L.-Y. Meng, M.-H. Li, T.-X. Huang, J.-H. Zhong, X. Wang, Z.-L. Yang and B. Ren, J. Am. Chem. Soc., 2015, 137, 11928–11931 CrossRef CAS PubMed .
  159. Y. Yokota, N. Hayazawa, B. Yang, E. Kazuma, F. C. I. Catalan and Y. Kim, J. Phys. Chem. C, 2019, 123, 2953–2963 CrossRef CAS .
  160. N. Martín Sabanés, L. M. A. Driessen and K. F. Domke, Anal. Chem., 2016, 88, 7108–7114 CrossRef PubMed .
  161. X. Chen, G. Goubert, S. Jiang and R. P. Van Duyne, J. Phys. Chem. C, 2018, 122, 11586–11590 CrossRef CAS .
  162. B. Gjonaj, P. Johnson, M. Bonn and K. F. Domke, Appl. Opt., 2012, 51, 8034–8040 CrossRef PubMed .
  163. S.-C. Huang, J.-Z. Ye, X.-R. Shen, Q.-Q. Zhao, Z.-C. Zeng, M.-H. Li, D.-Y. Wu, X. Wang and B. Ren, Anal. Chem., 2019, 91, 11092–11097 CrossRef CAS PubMed .
  164. Y.-F. Bao, M.-F. Cao, S.-S. Wu, T.-X. Huang, Z.-C. Zeng, M.-H. Li, X. Wang and B. Ren, Anal. Chem., 2020, 92, 12548–12555 CrossRef CAS PubMed .
  165. D. Kurouski, M. Mattei and R. P. Van Duyne, Nano Lett., 2015, 15, 7956–7962 CrossRef CAS PubMed .
  166. D. Xu, B. Liang, Y. Xu and M. Liu, Nano Res., 2023, 16, 5555–5571 CrossRef .
  167. P. Z. El-Khoury, Acc. Chem. Res., 2021, 54, 4576–4583 CrossRef CAS PubMed .
  168. I. Maouli, A. Taguchi, Y. Saito, S. Kawata and P. Verma, Appl. Phys. Express, 2015, 8, 032401 CrossRef CAS .
  169. A. Hartschuh, N. Anderson and L. Novotny, J. Microsc., 2003, 210, 234–240 CrossRef CAS PubMed .
  170. B. Ren, G. Picardi and B. Pettinger, Rev. Sci. Instrum., 2004, 75, 837–841 CrossRef CAS .
  171. X. Wang, Z. Liu, M.-D. Zhuang, H.-M. Zhang, X. Wang, Z.-X. Xie, D.-Y. Wu, B. Ren and Z.-Q. Tian, Appl. Phys. Lett., 2007, 91, 101105 CrossRef .
  172. S. S. Kharintsev, A. I. Noskov, G. G. Hoffmann and J. Loos, Nanotechnology, 2011, 22, 025202 CrossRef PubMed .
  173. J. P. Ibe, P. P. Bey Jr., S. L. Brandow, R. A. Brizzolara, N. A. Burnham, D. P. DiLella, K. P. Lee, C. R. K. Marrian and R. J. Colton, J. Vac. Sci. Technol., A, 1990, 8, 3570–3575 CrossRef CAS .
  174. B. Yang, E. Kazuma, Y. Yokota and Y. Kim, J. Phys. Chem. C, 2018, 122, 16950–16955 CrossRef CAS .
  175. S. S. Sasaki, S. M. Perdue, A. R. Perez, N. Tallarida, J. H. Majors, V. A. Apkarian and J. Lee, Rev. Sci. Instrum., 2013, 84, 096109 CrossRef PubMed .
  176. C. Zhang, B. Gao, L. G. Chen, Q. S. Meng, H. Yang, R. Zhang, X. Tao, H. Y. Gao, Y. Liao and Z. C. Dong, Rev. Sci. Instrum., 2011, 82, 083101 CrossRef CAS PubMed .
  177. J. Stadler, T. Schmid and R. Zenobi, Nano Lett., 2010, 10, 4514–4520 CrossRef CAS PubMed .
  178. M. Li, R. Lv, S. Huang, Y. Dai, Z. Zeng, L. Wang and B. Ren, J. Raman Spectrosc., 2016, 47, 808–812 CrossRef CAS .
  179. H. Böckmann, S. Liu, M. Müller, A. Hammud, M. Wolf and T. Kumagai, Nano Lett., 2019, 19, 3597–3602 CrossRef PubMed .
  180. X.-B. Zhang, Y.-F. Zhang, H. Li, J. Cui, S. Jiang, B. Yang, Y. Zhang, Y. Zhang and Z.-C. Dong, Chin. J. Chem. Phys., 2022, 35, 713–719 CrossRef CAS .
  181. N. Tallarida, J. Lee and V. A. Apkarian, ACS Nano, 2017, 11, 11393–11401 CrossRef CAS PubMed .
  182. L.-K. Yang, T.-X. Huang, Z.-C. Zeng, M.-H. Li, X. Wang, F.-Z. Yang and B. Ren, Nanoscale, 2015, 7, 18225–18231 RSC .
  183. T. W. Johnson, Z. J. Lapin, R. Beams, N. C. Lindquist, S. G. Rodrigo, L. Novotny and S.-H. Oh, ACS Nano, 2012, 6, 9168–9174 CrossRef CAS PubMed .
  184. H. Wen, J. Li, Q. Zhang, T. Inose, W. Peeters, B. Fortuni, H. Asakawa, A. Masuhara, K. Hirai, S. Toyouchi, Y. Fujita and H. Uji-i, Nano Lett., 2023, 23, 1615–1621 CrossRef CAS PubMed .
  185. B.-S. Yeo, W. Zhang, C. Vannier and R. Zenobi, Appl. Spectrosc., 2006, 60, 1142–1147 CrossRef CAS PubMed .
  186. A. Taguchi, N. Hayazawa, Y. Saito, H. Ishitobi, A. Tarun and S. Kawata, Opt. Express, 2009, 17, 6509–6518 CrossRef CAS PubMed .
  187. G. L. Bartolomeo, Y. Zhang, N. Kumar and R. Zenobi, Anal. Chem., 2021, 93, 15358–15364 CrossRef CAS PubMed .
  188. R. Kato, T. Umakoshi and P. Verma, J. Phys. Chem. C, 2021, 125, 20397–20404 CrossRef CAS .
  189. T.-X. Huang, C.-W. Li, L.-K. Yang, J.-F. Zhu, X. Yao, C. Liu, K.-Q. Lin, Z.-C. Zeng, S.-S. Wu, X. Wang, F.-Z. Yang and B. Ren, Nanoscale, 2018, 10, 4398–4405 RSC .
  190. P. Walke, Y. Fujita, W. Peeters, S. Toyouchi, W. Frederickx, S. De Feyter and H. Uji-i, Nanoscale, 2018, 10, 7556–7565 RSC .
  191. H. Wen, T. Inose, K. Hirai, T. Akashi, S. Sugioka, J. Li, W. Peeters, E. Fron, B. Fortuni, Y. Nakata, S. Rocha, S. Toyouchi, Y. Fujita and H. Uji-i, Nanoscale, 2022, 14, 5439–5446 RSC .
  192. V. Deckert, T. Deckert-Gaudig, M. Diegel, I. Götz, L. Langelüddecke, H. Schneidewind, G. Sharma, P. Singh, P. Singh, S. Trautmann, M. Zeisberger and Z. Zhang, Faraday Discuss., 2015, 177, 9–20 RSC .
  193. B. S. Oliveira, B. S. Archanjo, R. Valaski, C. A. Achete, L. G. Cançado, A. Jorio and T. L. Vasconcelos, J. Chem. Phys., 2020, 153, 114201 CrossRef CAS PubMed .
  194. T. Umakoshi, M. Tanaka, Y. Saito and P. Verma, Sci. Adv., 2020, 6, eaba4179 CrossRef CAS PubMed .
  195. H. He, C. Lin, C. Zong, M. Xu, P. Zheng, R. Ye, L. Wang and B. Ren, J. Raman Spectrosc., 2020, 51, 2552–2561 CrossRef CAS .
  196. H. He, M. Cao, X. Yue, M. Xu, L. Wang and B. Ren, Anal. Chem., 2021, 93, 14609–14617 CrossRef CAS PubMed .
  197. H. He, M. Cao, Y. Gao, P. Zheng, S. Yan, J.-H. Zhong, L. Wang, D. Jin and B. Ren, Nat. Commun., 2024, 15, 754 CrossRef CAS PubMed .
  198. G. Yang, X. Li, Y. Cheng, M. Wang, D. Ma, A. P. Sokolov, S. V. Kalinin, G. M. Veith and J. Nanda, Nat. Commun., 2021, 12, 578 CrossRef CAS PubMed .
  199. U. Kajendirarajah, M. Olivia Avilés and F. Lagugné-Labarthet, Phys. Chem. Chem. Phys., 2020, 22, 17857–17866 RSC .
  200. J. Clavilier, R. Faure, G. Guinet and R. Durand, J. Electroanal. Chem., 1980, 107, 205–209 CrossRef CAS .
  201. M. Hegner, P. Wagner and G. Semenza, Surf. Sci., 1993, 291, 39–46 CrossRef CAS .
  202. D. Xiaoru, Y. Ben, Z. Rui, W. Ruipu, Z. Yang, Z. Yao and D. Zhenchao, Light Adv. Manuf., 2022, 3, 52 Search PubMed .
  203. X. Wang, J.-H. Zhong, M. Zhang, Z. Liu, D.-Y. Wu and B. Ren, Anal. Chem., 2016, 88, 915–921 CrossRef CAS PubMed .
  204. S. Jiang, Y. Zhang, R. Zhang, C. Hu, M. Liao, Y. Luo, J. Yang, Z. Dong and J. G. Hou, Nat. Nanotechnol., 2015, 10, 865–869 CrossRef CAS PubMed .
  205. C. Zhang, R. B. Jaculbia, Y. Tanaka, E. Kazuma, H. Imada, N. Hayazawa, A. Muranaka, M. Uchiyama and Y. Kim, J. Am. Chem. Soc., 2021, 143, 9461–9467 CrossRef CAS PubMed .
  206. P. M. Albrecht and J. W. Lyding, Appl. Phys. Lett., 2003, 83, 5029–5031 CrossRef CAS .
  207. M. Liao, S. Jiang, C. Hu, R. Zhang, Y. Kuang, J. Zhu, Y. Zhang and Z. Dong, Nano Lett., 2016, 16, 4040–4046 CrossRef CAS PubMed .
  208. K.-Q. Lin, J. Yi, J.-H. Zhong, S. Hu, B.-J. Liu, J.-Y. Liu, C. Zong, Z.-C. Lei, X. Wang, J. Aizpurua, R. Esteban and B. Ren, Nat. Commun., 2017, 8, 14891 CrossRef CAS PubMed .
  209. A. D. McFarland, M. A. Young, J. A. Dieringer and R. P. Van Duyne, J. Phys. Chem. B, 2005, 109, 11279–11285 CrossRef CAS PubMed .
  210. C. L. Haynes and R. P. Van Duyne, J. Phys. Chem. B, 2003, 107, 7426–7433 CrossRef CAS .
  211. X. Wang, S.-C. Huang, S. Hu, S. Yan and B. Ren, Nat. Rev. Phys., 2020, 2, 253–271 CrossRef .
  212. J. Szczerbiński, J. B. Metternich, G. Goubert and R. Zenobi, Small, 2020, 16, 1905197 CrossRef PubMed .
  213. H. Ago, K. Petritsch, M. S. P. Shaffer, A. H. Windle and R. H. Friend, Adv. Mater., 1999, 11, 1281–1285 CrossRef CAS .
  214. J. Lefebvre, R. D. Antonov, M. Radosavljević, J. F. Lynch, M. Llaguno and A. T. Johnson, Carbon, 2000, 38, 1745–1749 CrossRef CAS .
  215. S. Ravindran, S. Chaudhary, B. Colburn, M. Ozkan and C. S. Ozkan, Nano Lett., 2003, 3, 447–453 CrossRef CAS .
  216. J. L. Blackburn, ACS Energy Lett., 2017, 2, 1598–1613 CrossRef CAS .
  217. P. J. Harris, Carbon Nanotube Science: Synthesis, Properties and Applications, Cambridge University Press, Cambridge, 2009 Search PubMed .
  218. A. Hartschuh, E. J. Sánchez, X. S. Xie and L. Novotny, Phys. Rev. Lett., 2003, 90, 095503 CrossRef PubMed .
  219. N. Anderson, A. Hartschuh, S. Cronin and L. Novotny, J. Am. Chem. Soc., 2005, 127, 2533–2537 CrossRef CAS PubMed .
  220. N. Anderson, A. Hartschuh and L. Novotny, Nano Lett., 2007, 7, 577–582 CrossRef CAS PubMed .
  221. L. G. Cançado, A. Hartschuh and L. Novotny, J. Raman Spectrosc., 2009, 40, 1420–1426 CrossRef .
  222. T.-a Yano, P. Verma, Y. Saito, T. Ichimura and S. Kawata, Nat. Photonics, 2009, 3, 473–477 CrossRef CAS .
  223. I. O. Maciel, N. Anderson, M. A. Pimenta, A. Hartschuh, H. Qian, M. Terrones, H. Terrones, J. Campos-Delgado, A. M. Rao, L. Novotny and A. Jorio, Nat. Mater., 2008, 7, 878–883 CrossRef CAS PubMed .
  224. R. Kato, S. Igarashi, T. Umakoshi and P. Verma, ACS Appl. Nano Mater., 2020, 3, 6001–6008 CrossRef CAS .
  225. C. D. Tschannen, M. Frimmer, G. Gordeev, T. L. Vasconcelos, L. Shi, T. Pichler, S. Reich, S. Heeg and L. Novotny, ACS Nano, 2021, 15, 12249–12255 CrossRef CAS PubMed .
  226. X. Zhao, Y. Ando, Y. Liu, M. Jinno and T. Suzuki, Phys. Rev. Lett., 2003, 90, 187401 CrossRef PubMed .
  227. L. Shi, P. Rohringer, K. Suenaga, Y. Niimi, J. Kotakoski, J. C. Meyer, H. Peterlik, M. Wanko, S. Cahangirov, A. Rubio, Z. J. Lapin, L. Novotny, P. Ayala and T. Pichler, Nat. Mater., 2016, 15, 634–639 CrossRef CAS PubMed .
  228. M. Liu, V. I. Artyukhov, H. Lee, F. Xu and B. I. Yakobson, ACS Nano, 2013, 7, 10075–10082 CrossRef CAS PubMed .
  229. J. Kūrti, C. Magyar, A. Balázs and P. Rajczy, Synth. Met., 1995, 71, 1865–1866 CrossRef .
  230. A. Milani, M. Tommasini, D. Fazzi, C. Castiglioni, M. D. Zoppo and G. Zerbi, J. Raman Spectrosc., 2008, 39, 164–168 CrossRef CAS .
  231. S. Yang and M. Kertesz, J. Phys. Chem. A, 2006, 110, 9771–9774 CrossRef CAS PubMed .
  232. S. Heeg, L. Shi, L. V. Poulikakos, T. Pichler and L. Novotny, Nano Lett., 2018, 18, 5426–5431 CrossRef CAS PubMed .
  233. C. D. Tschannen, T. L. Vasconcelos and L. Novotny, J. Chem. Phys., 2022, 156, 044203 CrossRef CAS PubMed .
  234. C. D. Tschannen, M. Frimmer, T. L. Vasconcelos, L. Shi, T. Pichler and L. Novotny, Nano Lett., 2022, 22, 3260–3265 CrossRef CAS PubMed .
  235. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666–669 CrossRef CAS PubMed .
  236. A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183–191 CrossRef CAS PubMed .
  237. Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras and P. Jarillo-Herrero, Nature, 2018, 556, 43–50 CrossRef CAS PubMed .
  238. Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y. Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, E. Kaxiras, R. C. Ashoori and P. Jarillo-Herrero, Nature, 2018, 556, 80–84 CrossRef CAS PubMed .
  239. Y. Cao, D. Rodan-Legrain, O. Rubies-Bigorda, J. M. Park, K. Watanabe, T. Taniguchi and P. Jarillo-Herrero, Nature, 2020, 583, 215–220 CrossRef CAS PubMed .
  240. A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov and A. K. Geim, Rev. Mod. Phys., 2009, 81, 109–162 CrossRef CAS .
  241. M.-L. Lin, J.-B. Wu, X.-L. Liu and P.-H. Tan, J. Raman Spectrosc., 2018, 49, 19–30 CrossRef CAS .
  242. J. E. Lee, G. Ahn, J. Shim, Y. S. Lee and S. Ryu, Nat. Commun., 2012, 3, 1024 CrossRef PubMed .
  243. N. S. Mueller, S. Heeg, M. P. Alvarez, P. Kusch, S. Wasserroth, N. Clark, F. Schedin, J. Parthenios, K. Papagelis, C. Galiotis, M. Kalbáč, A. Vijayaraghavan, U. Huebner, R. Gorbachev, O. Frank and S. Reich, 2D Mater., 2018, 5, 015016 CrossRef .
  244. G. Ahn and S. Ryu, Carbon, 2018, 138, 257–263 CrossRef CAS .
  245. H. Miranda, C. Rabelo, L. G. Cançado, T. L. Vasconcelos, B. S. Oliveira, F. Schulz, H. Lange, S. Reich, P. Kusch and A. Jorio, Phys. Rev. Res., 2020, 2, 023408 CrossRef CAS .
  246. M. Ghislandi, G. G. Hoffmann, E. Tkalya, L. Xue and G. D. With, Appl. Spectrosc. Rev., 2012, 47, 371–381 CrossRef CAS .
  247. R. H. Rickman and P. R. Dunstan, J. Raman Spectrosc., 2014, 45, 15–21 CrossRef CAS .
  248. Z. J. Lapin, R. Beams, L. G. Cançado and L. Novotny, Faraday Discuss., 2015, 184, 193–206 RSC .
  249. R. Beams, L. G. Cançado, A. Jorio, A. N. Vamivakas and L. Novotny, Nanotechnology, 2015, 26, 175702 CrossRef PubMed .
  250. S. Mignuzzi, N. Kumar, B. Brennan, I. S. Gilmore, D. Richards, A. J. Pollard and D. Roy, Nanoscale, 2015, 7, 19413–19418 RSC .
  251. R. Beams, L. G. Cançado, S.-H. Oh, A. Jorio and L. Novotny, Phys. Rev. Lett., 2014, 113, 186101 CrossRef PubMed .
  252. H. Miranda, V. Monken, J. L. E. Campos, T. L. Vasconcelos, C. Rabelo, B. S. Archanjo, C. M. Almeida, S. Grieger, C. Backes, A. Jorio and L. G. Cançado, 2D Mater., 2023, 10, 015002 CrossRef CAS .
  253. W. Su and D. Roy, J. Vac. Sci. Technol., B, 2013, 31, 041808 CrossRef .
  254. W. Su, N. Kumar, N. Dai and D. Roy, Chem. Commun., 2016, 52, 8227–8230 RSC .
  255. R. Beams, L. G. Cançado and L. Novotny, Nano Lett., 2011, 11, 1177–1181 CrossRef CAS PubMed .
  256. L. Gustavo Cançado, M. Gomes da Silva, E. H. Martins Ferreira, F. Hof, K. Kampioti, K. Huang, A. Pénicaud, C. Alberto Achete, R. B. Capaz and A. Jorio, 2D Mater., 2017, 4, 025039 CrossRef .
  257. E. J. Legge, K. R. Paton, M. Wywijas, G. McMahon, R. Pemberton, N. Kumar, A. P. Aranga Raju, C. P. Dawson, A. J. Strudwick, J. W. Bradley, V. Stolojan, S. R. P. Silva, S. A. Hodge, B. Brennan and A. J. Pollard, ACS Appl. Mater. Interfaces, 2020, 12, 13481–13493 CrossRef CAS PubMed .
  258. N. Kumar, S. Marchesini, T. Howe, L. Edwards, B. Brennan and A. J. Pollard, J. Chem. Phys., 2020, 153, 184708 CrossRef CAS PubMed .
  259. T. Inose, S. Toyouchi, S. Hara, S. Sugioka, P. Walke, R. Oyabu, B. Fortuni, W. Peeters, Y. Usami, K. Hirai, S. De Feyter, H. Uji-i, Y. Fujita and H. Tanaka, Small, 2024, 20, 2301841 CrossRef CAS PubMed .
  260. W. Su, A. Esfandiar, O. Lancry, J. Shao, N. Kumar and M. Chaigneau, Chem. Commun., 2021, 57, 6895–6898 RSC .
  261. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis, Nat. Nanotechnol., 2011, 6, 147–150 CrossRef CAS PubMed .
  262. M. Chhowalla, D. Jena and H. Zhang, Nat. Rev. Mater., 2016, 1, 16052 CrossRef CAS .
  263. Y. Liu, X. Duan, H.-J. Shin, S. Park, Y. Huang and X. Duan, Nature, 2021, 591, 43–53 CrossRef CAS PubMed .
  264. Z. Lin, B. R. Carvalho, E. Kahn, R. Lv, R. Rao, H. Terrones, M. A. Pimenta and M. Terrones, 2D Mater., 2016, 3, 022002 CrossRef .
  265. C. Lee, B. G. Jeong, S. J. Yun, Y. H. Lee, S. M. Lee and M. S. Jeong, ACS Nano, 2018, 12, 9982–9990 CrossRef CAS PubMed .
  266. R. Kato, T. Umakoshi, R. T. Sam and P. Verma, Appl. Phys. Lett., 2019, 114, 073105 CrossRef .
  267. D. V. Voronine, G. Lu, D. Zhu and A. Krayev, IEEE J. Sel. Top. Quantum Electron., 2017, 23, 138–143 Search PubMed .
  268. D. Jariwala, A. Krayev, J. Wong, A. E. Robinson, M. C. Sherrott, S. Wang, G.-Y. Liu, M. Terrones and H. A. Atwater, 2D Mater., 2018, 5, 035003 CrossRef .
  269. C. Lee, B. G. Jeong, S. H. Kim, D. H. Kim, S. J. Yun, W. Choi, S.-J. An, D. Lee, Y.-M. Kim, K. K. Kim, S. M. Lee and M. S. Jeong, npj 2D Mater. Appl., 2022, 6, 67 CrossRef CAS .
  270. A. Rodriguez, A. Krayev, M. Velický, O. Frank and P. Z. El-Khoury, J. Phys. Chem. Lett., 2022, 13, 5854–5859 CrossRef CAS PubMed .
  271. J. Shao, F. Chen, W. Su, Y. Zeng and H.-W. Lu, ACS Appl. Mater. Interfaces, 2021, 13, 20361–20370 CrossRef CAS PubMed .
  272. M. Rahaman, R. D. Rodriguez, G. Plechinger, S. Moras, C. Schüller, T. Korn and D. R. T. Zahn, Nano Lett., 2017, 17, 6027–6033 CrossRef CAS PubMed .
  273. K.-D. Park, O. Khatib, V. Kravtsov, G. Clark, X. Xu and M. B. Raschke, Nano Lett., 2016, 16, 2621–2627 CrossRef CAS PubMed .
  274. S.-S. Wu, T.-X. Huang, X. Xu, Y.-F. Bao, X.-D. Pei, X. Yao, M.-F. Cao, K.-Q. Lin, X. Wang, D. Wang and B. Ren, ACS Nano, 2022, 16, 4786–4794 CrossRef CAS PubMed .
  275. T. P. Darlington, C. Carmesin, M. Florian, E. Yanev, O. Ajayi, J. Ardelean, D. A. Rhodes, A. Ghiotto, A. Krayev, K. Watanabe, T. Taniguchi, J. W. Kysar, A. N. Pasupathy, J. C. Hone, F. Jahnke, N. J. Borys and P. J. Schuck, Nat. Nanotechnol., 2020, 15, 854–860 CrossRef CAS PubMed .
  276. M. Velický, A. Rodriguez, M. Bouša, A. V. Krayev, M. Vondráček, J. Honolka, M. Ahmadi, G. E. Donnelly, F. Huang, H. D. Abruña, K. S. Novoselov and O. Frank, J. Phys. Chem. Lett., 2020, 11, 6112–6118 CrossRef PubMed .
  277. A. C. Gadelha, T. L. Vasconcelos, L. G. Cançado and A. Jorio, J. Phys. Chem. Lett., 2021, 12, 7625–7631 CrossRef CAS PubMed .
  278. Y. Gong, J. Lin, X. Wang, G. Shi, S. Lei, Z. Lin, X. Zou, G. Ye, R. Vajtai, B. I. Yakobson, H. Terrones, M. Terrones, B. K. Tay, J. Lou, S. T. Pantelides, Z. Liu, W. Zhou and P. M. Ajayan, Nat. Mater., 2014, 13, 1135–1142 CrossRef CAS PubMed .
  279. Y. Gong, S. Lei, G. Ye, B. Li, Y. He, K. Keyshar, X. Zhang, Q. Wang, J. Lou, Z. Liu, R. Vajtai, W. Zhou and P. M. Ajayan, Nano Lett., 2015, 15, 6135–6141 CrossRef CAS PubMed .
  280. M.-Y. Li, Y. Shi, C.-C. Cheng, L.-S. Lu, Y.-C. Lin, H.-L. Tang, M.-L. Tsai, C.-W. Chu, K.-H. Wei, J.-H. He, W.-H. Chang, K. Suenaga and L.-J. Li, Science, 2015, 349, 524–528 CrossRef CAS PubMed .
  281. W. Xue, P. K. Sahoo, J. Liu, H. Zong, X. Lai, S. Ambardar and D. V. Voronine, J. Vac. Sci. Technol., A, 2018, 36, 05G502 CrossRef .
  282. S. Garg, J. P. Fix, A. V. Krayev, C. Flanery, M. Colgrove, A. R. Sulkanen, M. Wang, G.-Y. Liu, N. J. Borys and P. Kung, ACS Nano, 2022, 16, 340–350 CrossRef CAS PubMed .
  283. N. Takagi, C.-L. Lin, K. Kawahara, E. Minamitani, N. Tsukahara, M. Kawai and R. Arafune, Prog. Surf. Sci., 2015, 90, 1–20 CrossRef CAS .
  284. L. Tao, E. Cinquanta, D. Chiappe, C. Grazianetti, M. Fanciulli, M. Dubey, A. Molle and D. Akinwande, Nat. Nanotechnol., 2015, 10, 227–231 CrossRef CAS PubMed .
  285. A. J. Mannix, X.-F. Zhou, B. Kiraly, J. D. Wood, D. Alducin, B. D. Myers, X. Liu, B. L. Fisher, U. Santiago, J. R. Guest, M. J. Yacaman, A. Ponce, A. R. Oganov, M. C. Hersam and N. P. Guisinger, Science, 2015, 350, 1513–1516 CrossRef CAS PubMed .
  286. B. Feng, J. Zhang, Q. Zhong, W. Li, S. Li, H. Li, P. Cheng, S. Meng, L. Chen and K. Wu, Nat. Chem., 2016, 8, 563–568 CrossRef CAS PubMed .
  287. C.-C. Liu, W. Feng and Y. Yao, Phys. Rev. Lett., 2011, 107, 076802 CrossRef PubMed .
  288. B. Feng, O. Sugino, R.-Y. Liu, J. Zhang, R. Yukawa, M. Kawamura, T. Iimori, H. Kim, Y. Hasegawa, H. Li, L. Chen, K. Wu, H. Kumigashira, F. Komori, T.-C. Chiang, S. Meng and I. Matsuda, Phys. Rev. Lett., 2017, 118, 096401 CrossRef PubMed .
  289. S. Sheng, J.-B. Wu, X. Cong, Q. Zhong, W. Li, W. Hu, J. Gou, P. Cheng, P.-H. Tan, L. Chen and K. Wu, ACS Nano, 2019, 13, 4133–4139 CrossRef CAS PubMed .
  290. L. Li, J. F. Schultz, S. Mahapatra, X. Liu, C. Shaw, X. Zhang, M. C. Hersam and N. Jiang, J. Am. Chem. Soc., 2021, 143, 15624–15634 CrossRef CAS PubMed .
  291. S. Berweger, C. C. Neacsu, Y. Mao, H. Zhou, S. S. Wong and M. B. Raschke, Nat. Nanotechnol., 2009, 4, 496–499 CrossRef CAS PubMed .
  292. A. Hadj Youssef, J. Zhang, A. Ehteshami, G. Kolhatkar, C. Dab, D. Berthomieu, A. Merlen, F. Légaré and A. Ruediger, J. Phys. Chem. C, 2021, 125, 6200–6208 CrossRef CAS .
  293. M. Böhmler, Z. Wang, A. Myalitsin, A. Mews and A. Hartschuh, Angew. Chem., Int. Ed., 2011, 50, 11536–11538 CrossRef PubMed .
  294. S. Sheng, R. Ma, J.-B. Wu, W. Li, L. Kong, X. Cong, D. Cao, W. Hu, J. Gou, J.-W. Luo, P. Cheng, P.-H. Tan, Y. Jiang, L. Chen and K. Wu, Nano Lett., 2018, 18, 2937–2942 CrossRef CAS PubMed .
  295. F. Shao, V. Müller, Y. Zhang, A. D. Schlüter and R. Zenobi, Angew. Chem., Int. Ed., 2017, 56, 9361–9366 CrossRef CAS PubMed .
  296. F. Shao, W. Dai, Y. Zhang, W. Zhang, A. D. Schlüter and R. Zenobi, ACS Nano, 2018, 12, 5021–5029 CrossRef CAS PubMed .
  297. F. Shao, W. Wang, W. Yang, Z. Yang, Y. Zhang, J. Lan, A. Dieter Schlüter and R. Zenobi, Nat. Commun., 2021, 12, 4557 CrossRef CAS PubMed .
  298. B. Pettinger, G. Picardi, R. Schuster and G. Ertl, J. Electroanal. Chem., 2003, 554–555, 293–299 CrossRef CAS .
  299. B. Pettinger, B. Ren, G. Picardi, R. Schuster and G. Ertl, Phys. Rev. Lett., 2004, 92, 096101 CrossRef PubMed .
  300. B. Ren, G. Picardi, B. Pettinger, R. Schuster and G. Ertl, Angew. Chem., Int. Ed., 2005, 44, 139–142 CrossRef CAS PubMed .
  301. B.-S. Yeo, S. Mädler, T. Schmid, W. Zhang and R. Zenobi, J. Phys. Chem. C, 2008, 112, 4867–4873 CrossRef CAS .
  302. T. Deckert-Gaudig, E. Bailo and V. Deckert, Phys. Chem. Chem. Phys., 2009, 11, 7360–7362 RSC .
  303. T. Deckert-Gaudig and V. Deckert, J. Raman Spectrosc., 2009, 40, 1446–1451 CrossRef CAS .
  304. M. Kang, H. Kim, E. Oleiki, Y. Koo, H. Lee, H. Joo, J. Choi, T. Eom, G. Lee, Y. D. Suh and K.-D. Park, Nat. Commun., 2022, 13, 4133 CrossRef CAS PubMed .
  305. Z.-F. Cai, L.-Q. Zheng, Y. Zhang and R. Zenobi, J. Am. Chem. Soc., 2021, 143, 12380–12386 CrossRef CAS PubMed .
  306. M. Hong, Y. Yokota, N. Hayazawa, E. Kazuma and Y. Kim, J. Phys. Chem. C, 2020, 124, 13141–13149 CrossRef CAS .
  307. F. Shao, L. Zheng, J. Lan and R. Zenobi, Anal. Chem., 2022, 94, 1645–1653 CrossRef CAS PubMed .
  308. J. Steidtner and B. Pettinger, Phys. Rev. Lett., 2008, 100, 236101 CrossRef PubMed .
  309. N. Chiang, N. Jiang, L. R. Madison, E. A. Pozzi, M. R. Wasielewski, M. A. Ratner, M. C. Hersam, T. Seideman, G. C. Schatz and R. P. Van Duyne, J. Am. Chem. Soc., 2017, 139, 18664–18669 CrossRef CAS PubMed .
  310. J. Xu, X. Zhu, S. Tan, Y. Zhang, B. Li, Y. Tian, H. Shan, X. Cui, A. Zhao, Z. Dong, J. Yang, Y. Luo, B. Wang and J. G. Hou, Science, 2021, 371, 818–822 CrossRef CAS PubMed .
  311. S. Mahapatra, Y. Ning, J. F. Schultz, L. Li, J.-L. Zhang and N. Jiang, Nano Lett., 2019, 19, 3267–3272 CrossRef CAS PubMed .
  312. R. Zhang, X. Zhang, H. Wang, Y. Zhang, S. Jiang, C. Hu, Y. Zhang, Y. Luo and Z. Dong, Angew. Chem., Int. Ed., 2017, 56, 5561–5564 CrossRef CAS PubMed .
  313. E. M. van Schrojenstein Lantman, T. Deckert-Gaudig, A. J. G. Mank, V. Deckert and B. M. Weckhuysen, Nat. Nanotechnol., 2012, 7, 583–586 CrossRef CAS PubMed .
  314. L.-Q. Zheng, S. Yang, J. Lan, L. Gyr, G. Goubert, H. Qian, I. Aprahamian and R. Zenobi, J. Am. Chem. Soc., 2019, 141, 17637–17645 CrossRef CAS PubMed .
  315. N. Kumar, B. Stephanidis, R. Zenobi, A. J. Wain and D. Roy, Nanoscale, 2015, 7, 7133–7137 RSC .
  316. Z. Li and D. Kurouski, Acc. Chem. Res., 2021, 54, 2477–2487 CrossRef CAS PubMed .
  317. J.-J. Sun, H.-S. Su, H.-L. Yue, S.-C. Huang, T.-X. Huang, S. Hu, M. M. Sartin, J. Cheng and B. Ren, J. Phys. Chem. Lett., 2019, 10, 2306–2312 CrossRef CAS PubMed .
  318. Z.-F. Cai, J. P. Merino, W. Fang, N. Kumar, J. O. Richardson, S. De Feyter and R. Zenobi, J. Am. Chem. Soc., 2022, 144, 538–546 CrossRef CAS PubMed .
  319. Z. Li and D. Kurouski, Nano Lett., 2022, 22, 7484–7491 CrossRef CAS PubMed .
  320. Z. Li, S. Ehtesabi, S. Gojare, M. Richter, S. Kupfer, S. Gräfe and D. Kurouski, ACS Photonics, 2023, 10, 3390–3400 CrossRef CAS PubMed .
  321. Z. Li and D. Kurouski, ACS Photonics, 2021, 8, 2112–2119 CrossRef CAS .
  322. D. Nguyen, G. Kang, N. Chiang, X. Chen, T. Seideman, M. C. Hersam, G. C. Schatz and R. P. Van Duyne, J. Am. Chem. Soc., 2018, 140, 5948–5954 CrossRef CAS PubMed .
  323. H.-S. Su, X.-G. Zhang, J.-J. Sun, X. Jin, D.-Y. Wu, X.-B. Lian, J.-H. Zhong and B. Ren, Angew. Chem., Int. Ed., 2018, 57, 13177–13181 CrossRef CAS PubMed .
  324. H.-S. Su, H.-S. Feng, Q.-Q. Zhao, X.-G. Zhang, J.-J. Sun, Y. He, S.-C. Huang, T.-X. Huang, J.-H. Zhong, D.-Y. Wu and B. Ren, J. Am. Chem. Soc., 2020, 142, 1341–1347 CrossRef CAS PubMed .
  325. T. Schmid, B.-S. Yeo, G. Leong, J. Stadler and R. Zenobi, J. Raman Spectrosc., 2009, 40, 1392–1399 CrossRef CAS .
  326. N. M. Sabanés, A. Elizabeth, J. H. K. Pfisterer and K. F. Domke, Faraday Discuss., 2017, 205, 233–243 RSC .
  327. N. Kumar, C. S. Wondergem, A. J. Wain and B. M. Weckhuysen, J. Phys. Chem. Lett., 2019, 10, 1669–1675 CrossRef CAS PubMed .
  328. T. Touzalin, A. L. Dauphin, S. Joiret, I. T. Lucas and E. Maisonhaute, Phys. Chem. Chem. Phys., 2016, 18, 15510–15513 RSC .
  329. N. M. Sabanés, T. Ohto, D. Andrienko, Y. Nagata and K. F. Domke, Angew. Chem., Int. Ed., 2017, 56, 9796–9801 CrossRef PubMed .
  330. S.-C. Huang, Q.-Q. Zhao, H.-S. Feng, H. Ma, L.-B. Zhao, X. Wang and B. Ren, J. Phys. Chem. C, 2023, 127, 12568–12575 CrossRef CAS .
  331. A. Fiocco, A. A. Pavlic, F. Kanoufi, E. Maisonhaute, J.-M. Noël and I. T. Lucas, Anal. Chem., 2024, 96, 2791–2798 CAS .
  332. M. Mattei, G. Kang, G. Goubert, D. V. Chulhai, G. C. Schatz, L. Jensen and R. P. Van Duyne, Nano Lett., 2017, 17, 590–596 CrossRef CAS PubMed .
  333. G. Kang, M. Yang, M. S. Mattei, G. C. Schatz and R. P. Van Duyne, Nano Lett., 2019, 19, 2106–2113 CrossRef CAS PubMed .
  334. S.-C. Huang, X. Wang, Q.-Q. Zhao, J.-F. Zhu, C.-W. Li, Y.-H. He, S. Hu, M. M. Sartin, S. Yan and B. Ren, Nat. Commun., 2020, 11, 4211 CrossRef CAS PubMed .
  335. Z. Chen, S. Jiang, G. Kang, D. Nguyen, G. C. Schatz and R. P. Van Duyne, J. Am. Chem. Soc., 2019, 141, 15684–15692 CrossRef CAS PubMed .
  336. T.-X. Huang, X. Cong, S.-S. Wu, J.-B. Wu, Y.-F. Bao, M.-F. Cao, L. Wu, M.-L. Lin, X. Wang, P.-H. Tan and B. Ren, Nat. Catal., 2024, 7, 646–654 CrossRef CAS .
  337. N. Hayazawa, Y. Saito and S. Kawata, Appl. Phys. Lett., 2004, 85, 6239–6241 CrossRef CAS .
  338. Y. Guo, S. Jiang, X. Chen, M. Mattei, J. A. Dieringer, J. P. Ciraldo and R. P. Van Duyne, J. Phys. Chem. C, 2018, 122, 14865–14871 CrossRef CAS .
  339. P. Pienpinijtham, S. Vantasin, Y. Kitahama, S. Ekgasit and Y. Ozaki, J. Phys. Chem. C, 2016, 120, 14663–14668 CrossRef CAS .
  340. T. Sukmanee, K. Wongravee, Y. Kitahama, S. Ekgasit, T. Itoh, P. Pienpinijtham and Y. Ozaki, Angew. Chem., Int. Ed., 2020, 59, 14564–14569 CrossRef CAS PubMed .
  341. H.-Y. Du, Y.-F. Huang, D. Wong, M.-F. Tseng, Y.-H. Lee, C.-H. Wang, C.-L. Lin, G. Hoffmann, K.-H. Chen and L.-C. Chen, Nat. Commun., 2021, 12, 1321 CrossRef CAS PubMed .
  342. H. K. Wickramasinghe, M. Chaigneau, R. Yasukuni, G. Picardi and R. Ossikovski, ACS Nano, 2014, 8, 3421–3426 CrossRef CAS PubMed .
  343. T. Ichimura, N. Hayazawa, M. Hashimoto, Y. Inouye and S. Kawata, Phys. Rev. Lett., 2004, 92, 220801 CrossRef PubMed .
  344. E. A. Pozzi, M. D. Sonntag, N. Jiang, N. Chiang, T. Seideman, M. C. Hersam and R. P. Van Duyne, J. Phys. Chem. Lett., 2014, 5, 2657–2661 CrossRef CAS PubMed .
  345. Y. Luo, A. Martin-Jimenez, R. Gutzler, M. Garg and K. Kern, Nano Lett., 2022, 22, 5100–5106 CrossRef CAS PubMed .
  346. C. Zong and J.-X. Cheng, Nanophotonics, 2021, 10, 617–625 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2024