Hostname: page-component-68c7f8b79f-rgmxm Total loading time: 0 Render date: 2026-01-19T20:15:10.966Z Has data issue: false hasContentIssue false

Infrared nanoscopy for subcellular chemical imaging

Published online by Cambridge University Press:  19 January 2026

Katerina Kanevche
Affiliation:
Department of Chemistry, Princeton University , Princeton, NJ, USA
David Joll Burr
Affiliation:
Department of Physics, Experimental Biophysics and Space Sciences, Freie Universität Berlin , Berlin, Germany Institute for Biology – Microbiology, Freie Universität Berlin, Berlin, Germany
Janina Drauschke
Affiliation:
Department of Physics, Experimental Biophysics and Space Sciences, Freie Universität Berlin , Berlin, Germany
Jacek Kozuch
Affiliation:
Department of Physics, Experimental Molecular Biophysics, Freie Universität Berlin , Berlin, Germany
Carlos Baiz
Affiliation:
Department of Chemistry, University of Texas at Austin , Austin, TX, USA
Andreas Elsaesser
Affiliation:
Department of Physics, Experimental Biophysics and Space Sciences, Freie Universität Berlin , Berlin, Germany
Joachim Heberle*
Affiliation:
Department of Physics, Experimental Molecular Biophysics, Freie Universität Berlin , Berlin, Germany
*
Corresponding author: Joachim Heberle; Email: jheberle@zedat.fu-berlin.de
Rights & Permissions [Opens in a new window]

Abstract

Infrared (IR) nanoscopy represents a collection of imaging and spectroscopy techniques capable of resolving IR absorption on the nanometer scale. Chemical specificity is leveraged from vibrational spectroscopy, while light–matter interactions are detected by observing perturbations in the optical near field with an atomic force microscopy probe. Therefore, imaging is wavelength independent and has a spatial resolution on the nanometer scale, well beyond the classical diffraction limit. In this perspective, we outline the recent biological applications of scattering type scanning near-field optical microscopy and nanoscale Fourier-transform IR spectroscopy. These techniques are uniquely suited to resolving subcellular ultrastructure from a variety of cell types, as well as studying biological processes such as metabolic activity on the single-cell level. Furthermore, this review describes recent technical advances in IR nanoscopy, and emerging machine learning supported approaches to sampling, signal enhancement, and data processing. This emphasizes that label-free IR nanoscopy holds significant potential for ongoing and future biological applications.

Information

Type
Perspective
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2026. Published by Cambridge University Press

Introduction

Studying biomolecular processes in cells frequently relies on the ability to resolve chemical composition on a subcellular level, and thus requires nanoscale spatial resolution. Some established routes to subcellular imaging include super-resolution fluorescence microscopy (Sahl et al., Reference Sahl, Hell and Jakobs2017), or electron microscopy (EM) (Nogales and Mahamid, Reference Nogales and Mahamid2024). However, fluorescence-based microscopy requires labeling of the biomolecule of interest with a fluorescent tag. This carries certain drawbacks including prior knowledge of the sample composition, the development and availability of suitable fluorescent tag molecules, and the inclusion of fluorophores potentially introducing toxicity or perturbing the native state of the biological system being interrogated. Cryo-EM provides unprecedented spatial resolution down to the Ångstrom level and the cryogenic temperatures can mitigate sample damage from high-energy electrons. However, this technique is very costly to operate, involves complex computationally expensive data processing algorithms, and lacks direct information on the chemical composition.

Nondestructive techniques that are inherently sensitive to chemical composition, while allowing for mapping on the nanometer scale, are a promising alternative for cellular bioimaging. In the realm of tip-enhanced microscopy and spectroscopy, methodologies that combine scanning probe microscopy (SPM) and vibrational spectroscopy are particularly attractive. Atomic force microscopy (AFM), a type of SPM, is a surface-sensitive technique frequently employed for the characterization of soft matter and biological samples, which yields spatial resolution on the nanometer level (Allison et al., Reference Allison, Mortensen, Sullivan and Doktycz2010; Müller and Dufrêne, Reference Müller and Dufrêne2011). Combining this method with vibrational spectroscopy overcomes the diffraction limit by accessing the light–matter interaction in the optical near field via the scanning probe, thus, the molecular vibrations can be resolved with the spatial resolution of AFM.

Scattering type scanning near-field optical microscopy (sSNOM) (Figure 1) is one example that utilizes an AFM tip as an antenna to localize electromagnetic radiation to the immediate proximity of the tip. This approach provides routine spatial resolutions of 20 nm. sSNOM images are recorded using tunable laser sources, with the majority of its applications in the mid-IR and THz spectral range. When employing broadband laser sources, the spectra are recorded via a Fourier transform-based spectrometer, where the sample arm of the interferometer entails the AFM tip–sample interaction. This technique is known as nanometer Fourier transform infrared (nanoFTIR) (Huth et al., Reference Huth, Govyadinov, Amarie, Nuansing, Keilmann and Hillenbrand2012). Both approaches are encompassed under the term IR nanoscopy. With the emergence of novel IR laser sources (Faist et al., Reference Faist, Capasso, Sivco, Sirtori, Hutchinson and Cho1994; Keilmann and Amarie, Reference Keilmann and Amarie2012) and the development of a sophisticated detection scheme that allows for the extraction of optical material properties with nm resolution (Ocelic et al., Reference Ocelic, Huber and Hillenbrand2006), the applications of sSNOM and nanoFTIR have dramatically increased across fields. One complementary technique to sSNOM is AFM-IR, where the imaging contrast originates from photothermal-induced resonance in the sample (Dazzi et al., Reference Dazzi, Prater, Hu, Chase, Rabolt and Marcott2012). A recent review provides an excellent overview of developments and applications of nanoscopy (Hillenbrand et al., Reference Hillenbrand, Abate, Liu, Chen and Basov2025), while the key features of AFM-based mid-IR nanoscopy techniques are summarized in Table 1.

Figure 1. Technical overview of IR nanoscopy. Typical sSNOM (1) and nanoFTIR (2) setup based on an asymmetric Michelson interferometer. An IR light source is focused on an AFM tip, oscillating at frequency Ω, via parabolic mirror (PM). The backscattered light is collected via the same pathway, recombined with the reference beam at the beam splitter (BS) and focused on a mercury-cadmium-telluride (MCT) detector. This technique allows for recording the AFM topography, IR phase, and amplitude, which reveal the absorption and reflection of the specimen. Fourier transformation of the interferograms yields nanoFTIR spectra, thus capturing IR absorption across broad spectral range. Figure adapted from (Kanevche et al., Reference Kanevche, Burr, Nürnberg, Hass, Elsaesser and Heberle2021).

Table 1. Summary of AFM-based mid-IR nanoscopy techniques

The listed techniques can in principle operate in the visible and THz regime, as discussed in the review previously (Hillenbrand et al., Reference Hillenbrand, Abate, Liu, Chen and Basov2025).

Applying IR nanoscopy to soft matter and biological specimens presents additional challenges regarding adequate sample preparation and generally weaker signals. Successful nanoscale imaging and spectroscopy has been demonstrated on a variety of organic samples such as membrane proteins in lipids (Ballout et al., Reference Ballout, Krassen, Kopf, Ataka, Bründermann, Heberle and Havenith2011), individual protein complexes (Amenabar et al., Reference Amenabar, Poly, Nuansing, Hubrich, Govyadinov, Huth, Krutokhvostov, Zhang, Knez, Heberle, Bittner and Hillenbrand2013; Berweger et al., Reference Berweger, Nguyen, Muller, Bechtel, Perkins and Raschke2013), lipids (Kästner et al., Reference Kästner, Johnson, Hermann, Kruskopf, Pierz, Hoehl and Patoka2018), polypeptides and amyloids (Fernandes et al., Reference Fernandes, Chowdhary, Mikula, Saleh, Kanevche, Berlepsch and Böttcher2022; Paul et al., Reference Paul, Jeništová, Vosough, Berntsson, Mörman, Jarvet and Barth2023), and plant cell walls (Keplinger et al., Reference Keplinger, Konnerth, Aguié-Béghin, Rüggeberg, Gierlinger and Burgert2014; Veber et al., Reference Veber, Zancajo, Puskar, Schade and Kneipp2023) to name just a few.

Here, we focus on the various approaches taken for studying cellular structure and function, such as metabolic activity, using sSNOM and nanoFTIR in the mid-IR spectral range. We argue that IR nanoscopy can be utilized for correlative imaging in concert with other imaging techniques, as well as offering an alternative route for cellular bioimaging, and thus we highlight the potential of this technology. Specifically, we emphasize the developments in subcellular IR chemical imaging, outline applications utilizing nanoscale IR biomarker monitoring, and provide an overview of emerging technical advances, machine learning supported sampling, and postprocessing algorithms. We conclude with an outlook on how this technology may develop in the future.

Subcellular IR imaging

The ability of IR-nanoscopic techniques to move beyond the diffraction limit and resolve chemical composition on the nm level has resulted in community interest expanding from biological soft matter characterization to the study of whole model cellular organisms. Single-cell, label-free chemical imaging was achieved using sSNOM as early as 2016, wherein intracellular chemical mapping was performed on single human red blood cells (Amrania et al., Reference Amrania, Drummond, Coombes, Shousha, Woodley-Barker, Weir, Hart, Carter and Phillips2016). This provided a significant milestone toward applying IR nanoscopy in biomedical contexts.

As sSNOM is a surface-sensitive technique, subcellular IR nanoscopy requires a sample preparation protocol that provides the AFM probe access to the subcellular structure. To maintain high-quality sSNOM signal, it is crucial to avoid any tip damage or contamination through interaction with the samples. Operating the tip in tapping mode, minimizes these common challenges when using AFM for biological samples. To optimize the sSNOM signal, it is recommended to deposit samples on an atomically flat metallic substrate. Finally, it is best practice to use flat samples with a thickness similar to the IR near-field penetration depth. To achieve these requirements, cells or biological tissue can be thin sectioned using ultramicrotomy (Herrmann, Reference Herrmann2021), a procedure frequently employed in EM, and deposited on silicon or template-stripped gold (e.g., Horstmann et al., Reference Horstmann, Körber, Sätzler, Aydin and Kuner2012, and references therein). While robust samples, such as woody plant cells, can be sliced in this manner directly (Keplinger et al., Reference Keplinger, Konnerth, Aguié-Béghin, Rüggeberg, Gierlinger and Burgert2014), the imaging of ultrastructure or subcellular details in soft-matter samples typically requires cryopreservation or resin embedding prior to sectioning. Resin embedding poses less logistical challenges than cryosectioning and results in samples that can be indefinitely maintained at room temperature, with no measurement-induced loss in quality. However, it must be ensured that the IR features of the embedding resin itself do not obscure spectroscopic details of interest, and as chemical fixation or freeze substitution are necessary, this is not suitable for all sample types.

To our knowledge, the first instance of subcellular chemical imaging demonstrated melanin pigments within retina cells of zebrafish (Figure 2a) (Stanciu et al., Reference Stanciu, Tranca, Hristu and Stanciu2017). This initial application of subcellular sSNOM utilized a laser in the visible range (638 nm), but broader chemical information can be derived by probing in the mid-IR range. The first report of subcellular IR nanoscopy demonstrated the local absorption of subcellular membrane-bound organelles within the green alga Chlamydomonas reinhardtii (Figure 2b) (Kanevche et al., Reference Kanevche, Burr, Nürnberg, Hass, Elsaesser and Heberle2021). Absorption imaging at different wavenumbers (i.e., 1655 and 1540 cm−1) corresponded to unique cellular features. High absorption in the amide I and amide II regions demonstrated elevated protein concentrations within the pyrenoid, an organelle tightly packed with the enzyme Rubisco. Biomolecular condensates, such as nuclear bodies within the nucleus, were uniquely visible in the IR, but not via AFM topographic imaging. Furthermore, the remarkable resolving power of sSNOM revealed IR signatures of ultrafine structures within the cellular flagella which were as small as 30 nm as well as of the sub-20-nm-sized outer cell membrane. Being able to chemically resolve the detailed internal structure of C. reinhardtii on the nanoscale serves as a proof of concept for future applications that seek to answer outstanding questions in biology such as relating cell function to compartmentalization or how biochemical processes rely on the precise location of metabolites with cells.

Figure 2. Subcellular nanoscopy examples. A. AFM and sSNOM imaging (at 638 nm) of thin-sectioned zebrafish (Danio rerio) retina. B. AFM and sSNOM imaging (at wavelengths of interest) of thin-sectioned green algae (C. reinhardtii) cells, primarily revealing the protein distribution within the cell. C. sSNOM tomography constructed from ten sequential images of C. reinhardtii, imaged at 1655 cm-1. Video demonstrating the 3D structure of the green algae cell is available here: https://static-content.springer.com/esm/art%3A10.1038%2Fs42003-021-02876-7/MediaObjects/42003_2021_2876_MOESM4_ESM.mpg. D. Absorption spectrum of myeloma thin sections and sSNOM images at four wavelengths of interest, revealing the subcellular components. Figure adapted with permission from Stanciu et al., Reference Stanciu, Tranca, Hristu and Stanciu2017 © Optical Society of America, and (Greaves et al., Reference Greaves, Kiryushko, Auner, Porter and Phillips2023; Kanevche et al., Reference Kanevche, Burr, Nürnberg, Hass, Elsaesser and Heberle2021).

Performing IR-nanoscopic imaging on cellular thin sections can provide unprecedented subcellular chemical imaging detail in lateral dimensions, however, further information can be accessed in the z dimension. Other imaging techniques, including AFM (Chen et al., Reference Chen, Cai, Zhao, Wang, Dong, Luo and Chen2005) and serial block-face scanning EM (Denk and Horstmann, Reference Denk and Horstmann2004), have successfully employed a tomographic approach, sequentially examining consecutive cell sections to compile 3D cellular information. The first instance of IR cell tomography imaged the amide I absorption (1655 cm−1) of a single C. reinhardtii cell in 10 consecutive, 100-nm-thick sections (Kanevche et al., Reference Kanevche, Burr, Nürnberg, Hass, Elsaesser and Heberle2021). The resulting 2D images were superimposed, revealing the 3D amide I absorption of C. reinhardtii (Figure 2c).

Several contemporary studies have applied subcellular IR microscopy and sSNOM to investigate human disease and pathology (Freitas et al., Reference Freitas, Cernescu, Engdahl, Paulus, Levandoski, Martinsson and Prinz2021; Greaves et al., Reference Greaves, Kiryushko, Auner, Porter and Phillips2023, Reference Greaves, Allison, Machado, Morfill, Fleck, Porter and Phillips2024a,Reference Greaves, Pinna, Taylor, Porter and Phillipsb; Keogan et al., Reference Keogan, Nguyen, Bouzy, Stone, Jirstrom, Rahman and Meade2025). Intracellular IR imaging has been performed on human myeloma cells, identifying both the local protein content and isolated nucleic acid signatures through visualization at wavelengths corresponding to the amide regions, and the PO2 and C-O vibrational bands, respectively (Figure 2d) (Greaves et al., Reference Greaves, Kiryushko, Auner, Porter and Phillips2023). The sub-diffraction IR absorption of these cells revealed detailed subcellular structures; in the nuclear area, the nuclear membrane and nucleolus were prominent, and in contrast, dense fibrillar components within the nucleoli were characterized by low amide absorption. Furthermore, protein and nucleic acid absorption imaging at 1667 cm−1 revealed mitochondria and the 100-nm-sized cisternae structure of the endoplasmic reticulum. The observed binucleation with the fibrillar centers and components is a common phenotype for myeloma cells. Thus, the chemical content on nanometer scale obtained by sSNOM can be used in studying the mechanisms that drive the development and progression of diseases that are manifested through morphological and chemical changes on a sub- and interorganelle level. These studies emphasize IR nanoscopy as a mature technology uniquely suited for direct, label-free, subcellular bioimaging.

Nanoscale IR monitoring of cellular processes

IR imaging can be leveraged to study metabolic activity and nutrient uptake, and the single-cell capabilities of this technique allow for fine-scale monitoring beyond community-level analyses. A commonly employed technique in this context is stable isotope probing (SIP), which can provide highly detailed insights into cellular function (Alcolombri et al., Reference Alcolombri, Pioli, Stocker and Berry2022). During growth, cells can be cultured in the presence of distinct stable isotopes (commonly 15N, 2H, or 13C), which when metabolized are integrated into the molecular building blocks of the cells. As molecular vibrations are inherently affected by variations in atomic mass, isotopic labeling can be detected as characteristic peak shifts of the respective IR bands via the vibrational isotope effect (Figure 3a). Spectroscopic quantification of cellular isotope ratios can thus provide information on cellular metabolism and replication.

Figure 3. Applications of IR imaging for monitoring cellular processes. A. SIP-nanoFTIR as a means of deriving single-cell glucose uptake in E. coli. B. Correlative sSNOM absorption, sSNOM reflection (acquired at 1740 cm-1), and TEM imaging of AuNP interaction with hippocampal neurons. C. MSNP-treated malignant glioma cell imaged with sSNOM and AFM. Inset (ii) shows the MSNP-associated region, with nanoparticles visible at 1100 cm-1, but not at 1300 cm-1 nor AFM topography. Figure adapted from (Burr et al., Reference Burr, Drauschke, Kanevche, Kümmel, Stryhanyuk, Heberle and Elsaesser2024; Greaves, Allison, et al., Reference Greaves, Allison, Machado, Morfill, Fleck, Porter and Phillips2024a) and (Greaves, Pinna, et al., Reference Greaves, Pinna, Taylor, Porter and Phillips2024b).

Detection of SIP-labeled cells was pioneered using Raman spectroscopy (Wang et al., Reference Wang, Huang, Cui and Wagner2016) and mass spectrometry (i.e., nanoscale secondary ion mass spectrometry, nanoSIMS) (Musat et al., Reference Musat, Musat, Weber and Pett-Ridge2016). Recently, IR spectroscopy has been performed on individual SIP-labeled Escherichia coli bacteria, being detected by both optical photothermal IR spectroscopy (Lima et al., Reference Lima, Muhamadali, Xu, Kansiz and Goodacre2021) and nanoFTIR (Burr et al., Reference Burr, Drauschke, Kanevche, Kümmel, Stryhanyuk, Heberle and Elsaesser2024). These studies examined isotope-induced shifts in the amide region, with the former monitoring cellular uptake of 13C-glucose and 15N-ammonium chloride, while the latter study focused on the quantitation of single-cell 12C:13C ratios (Figure 3a). In addition to the proportional 13C-induced amide I peak shift (from 1655 to 1613 cm−1), the spectra were distinctly influenced by isotopic mode coupling, based on the distribution of isotopes within cellular proteins. This resulted in discernible amide I peak shapes, which could be attributed to cellular ‘age’ (i.e., cells harvested in early- or late-exponential growth phase). Such spectroscopic tracing of metabolic activity at the single-cell level provides insights into intercellular heterogeneity and microbial ecophysiological mechanisms. However, while isotopic labeling provides distinct IR spectral shifts and is a powerful tool to trace biomolecules, its application is limited to systems that can be cultured or experimentally manipulated. For tissue samples, patient-derived material, or complex organisms, isotopic labeling is often impractical or impossible, and IR nanoscopy of these samples must rely on endogenous molecular contrast.

IR imaging has also been utilized for the visualization of nanoparticles and their interaction with cells. Nanoparticles have attracted significant interest in the biomedical field due to their applications as drug delivery systems and in bioimaging (Doane and Burda, Reference Doane and Burda2012), however, their potential toxicological impact also requires careful consideration and assessment (Elsaesser and Howard, Reference Elsaesser and Howard2012). Quantifying the implementation of nanoparticles requires spatial tracking (through the cellular membrane to the intended target location), and the monitoring of their chemical activity and transformation upon biological interaction. Identification and tracking of nanoparticle uptake has been demonstrated using conventional EM-based methods (Elsaesser et al., Reference Elsaesser, Taylor, de Yanés, McKerr, Kim, O‘Hare and Howard2010, Reference Elsaesser, Barnes, McKerr, Salvati, Lynch, Dawson and Howard2011), however, for nanoparticles composed of low atomic number materials, EM struggles due to the lack of atomic contrast. An additional challenge is the formation of a so-called protein corona; a dynamic layer of biomolecules that adsorbs onto the nanoparticle surface in biological environments and effectively masks its physicochemical identity (Cedervall et al., Reference Cedervall, Lynch, Lindman, Berggård, Thulin, Nilsson and Linse2007). This phenomenon plays a critical role in governing how nanoparticles are recognized and processed by cells. Chemical imaging has long been suggested as a method to track nanoparticle uptake at the single-cell level (Elsaesser et al., Reference Elsaesser, Barnes, McKerr, Salvati, Lynch, Dawson and Howard2011). IR nanoscopy now enables direct visualization of not only nanoparticle distributions and chemical transformations, but also the protein corona itself, offering spatially resolved chemical insights into this critical bio–nano interface.

sSNOM imaging has been used to observe the interaction of Au nanoparticles (AuNP) within thin sections of hippocampal neurons (Greaves et al., Reference Greaves, Allison, Machado, Morfill, Fleck, Porter and Phillips2024a). In addition to identification of organelles such as the nuclei, mitochondria, plasma membranes, and the endoplasmic reticulum, this study used sSNOM phase and amplitude imaging to identify AuNP interactions with the neurons in three distinct scenarios: surface association, membrane embedded, and nanoparticle internalization (Figure 3b). Complementary studies have used sSNOM to image the cell volume and morphological features (such as lamellipodia and filopodia) of large (~50–100 μm) malignant glioma cells, and the internalized distribution of mesoporous silica nanoparticles (MSNP; Figure 3c) (Greaves et al., Reference Greaves, Pinna, Taylor, Porter and Phillips2024b). MSNP have a wide application in nanomedicine due to their broad biocompatibility, high drug-load capacity, and ease of functional modification. Topographic mapping using the unique absorption of MSNP (at 1100 cm−1) within whole glioblastoma cells demonstrates IR-nanoscopic imaging as a nondestructive and chemically informative method for quantifying subcellular nanoparticle distributions. This is crucial for evaluating efficacy in drug delivery and risks of nanotoxicology.

Emerging technical advances

Recent innovations in analytical instrumentation are enabling opportunities for IR nanoscopy in cellular and microbiological research. Preparing flat samples on reflective substrates now enable the integration of IR nanoscopy with multiple spectroscopic and imaging techniques. This creates opportunities for correlative spectroimaging workflows that greatly expand the breadth of chemical information derived from a single cell. For instance, correlative fluorescence in situ hybridization (FISH)–Raman–SEM-NanoSIMS has been performed on individual SIP-labeled microbial community samples (Schaible et al., Reference Schaible, Kohtz, Cliff and Hatzenpichler2022). Expanding multimodal techniques into the IR range, however, is still a new concept. Recent examples have focused on complementary photothermal IR methods, such as mid-IR photothermal-FISH (Bai et al., Reference Bai, Guo, Pereira, Wagner and Cheng2023) and optical photothermal IR Raman (Korona et al., Reference Korona, Orzechowska, Siąkała, Nowakowska, Pieczara, Buda and Baranska2025). As such, correlative IR nanoscopy is an exciting area of focus for future research.

Conventional IR nanoscopy systems are typically optimized for measurements in dry conditions. The advantage of operating in dry conditions is omitting water absorption contributions that overlaps with signals of biomolecules, a prevalent challenge in IR spectroscopy. However, many applications necessitate nanoscopy in aqueous conditions and recent technical developments are aimed at measurements in liquid; the native environment for many biological samples. Unwanted contributions from water absorption do pose a considerable challenge in bulk IR spectroscopic measurements where the majority of the volume occupied the solvent. Fortunately, in IR nanoscopy this issue is largely omitted as zeptoliter volumes (20 nm × 20 nm × 60 nm ≈ 24·10−21 L) of the material directly beneath the AFM tip are probed by the optical near field. Instead, a more considerable issue is the conventional top/side tip illumination geometry in IR nanoscopy, which requires the IR radiation to pass the bulk liquid. To overcome this, liquid sSNOM imaging has been performed on biomembranes and biomimetic peptoid sheets by immersing the AFM tip in an aqueous solution and focus the IR light via bottom illumination using an immersion lens (Pfitzner and Heberle, Reference Pfitzner and Heberle2020) or by illuminating the tip with an evanescent field generated by total internal reflection (O’Callahan et al., Reference O’Callahan, Park, Novikova, Jian, Chen, Muller and Lea2020). Alternatively, living cells have been investigated utilizing an invertible sample container filled with an aqueous solution covered with an IR transparent membrane. Live microorganisms can be adhered to a 10-nm-thick SiN membrane and after sample deposition, sSNOM imaging and/or nanoFTIR spectroscopy can be performed through the membrane, using the commonly used side-illumination geometry. In this configuration, the AFM tip is kept in dry conditions, which is characterized by better signal to noise, as compared to when the tip is immersed in liquid. This novel technology has been used to observe cell division and migration of both prokaryotic E. coli cells and eukaryotic lung carcinoma cells at a spatial resolution of ~150 nm (Kaltenecker et al., Reference Kaltenecker, Gölz, Bau and Keilmann2021). In addition to visualization over several hours, nanoFTIR spectra collected from the cell surface (characterized by lipid, protein, and tyrosine peaks) were chemically distinct from the extracellular fluid regions.

While many IR nanoscopy applications have used mid-IR, the spectral range can be expanded through the use of other sources. Examples include THz sSNOM (Hu et al., Reference Hu, Zhang, Qian, Lü, Zhu and Peng2024) or the broadband highly brilliant IR radiation generated by a synchrotron (Bechtel et al., Reference Bechtel, Johnson, Khatib, Muller and Raschke2020). Synchrotron IR nanospectroscopy (SINS) maintains the spatial resolution of nanoFTIR, while allowing for continuous spectral acquisition from 600 to 4000 cm−1, making this output comparable to conventional FTIR spectra. However, as the integrated power over this range is only ~1 mW, single-wavelength imaging is not feasible, and thus it is only possible to perform IR imaging via white light (translation of the interferogram centerburst intensity). While SINS is well established in solid-state physics, applications in biophysics are still in their infancy. However, a recent example of SINS demonstrated protein and lipid changes induced by chemotherapy on whole cancer cells (de Carvalho et al., Reference de Carvalho, Cinque, de Carvalho, Marques, Frogley, Vondracek and Marques2024). This is the first example of a nanoFTIR spectrum of the full fingerprint region of a human cancer cell.

Enhanced sampling and postprocessing algorithms

A key challenge in nanoscopy is the balance between spatial resolution, scan area, and sufficient signal-to-noise ratio. Biological samples, in particular, have inherently weak IR signals and involve large scan areas, if, for instance, the distribution of biomolecules within a cell is of interest. Weak signals require prolonged pixel integration times to provide sufficient contrast on weak absorption features. For instance, while recording of a single nanoFTIR spectrum acquisition takes a few seconds, long integration times and collecting tens to hundreds of co-additions to obtain a satisfactory signal-to-noise ratio results in total acquisition times in the order of minutes – thus comparable with conventional FTIR spectroscopy. The measurement time for hyperspectral nanoFTIR imaging of micron size areas can therefore be in the order of hours. The scanning speed for sSNOM and AFM-IR is equivalent to AFM scanning, which depending on the scan rate (commonly used 0.5-1 Hz/line), scan area (for cells 5–20 μm), and number of pixels (e.g., 256 × 256 px) is on the order of minutes. Recently, machine learning (ML) and enhanced sampling algorithms have begun to address this challenge by enhancing image quality, improving throughput, and reducing acquisition times without sacrificing essential spatial or molecular information.

Prior to the use of ML in sSNOM imaging, various algorithms have been developed for the denoising, image reconstruction, and resolution enhancement in AFM. Such approaches involved, for instance, common methods like wavelet transform based on the nonstationary character of images (Carmichael et al., Reference Carmichael, Vidu, Maksumov, Palazoglu and Stroeve2004; Kiwilszo et al., Reference Kiwilszo, Zieliński, Smulko and Darowicki2012; Schimmack and Mercorelli, Reference Schimmack and Mercorelli2018) or deconvolution by the tip geometry (Markiewicz and Goh, Reference Markiewicz and Goh1994; Udpa et al., Reference Udpa, Ayres, Fan, Chen and Kumar2006). However, parameters in these algorithms are user-defined, and as such, genuine objects in an AFM image may be removed or artefacts introduced. Such issues can be overcome by the use of ML, as demonstrated for a range of scanning probe-based techniques (Rahman Laskar and Celano, Reference Rahman Laskar and Celano2023).

Generative adversarial neural networks (GANs) have been effective across a wide range of scientific applications where denoising or feature enhancement is required (Li, Reference Li2023). GANs, initially developed for generating synthetic images (Isola et al., Reference Isola, Zhu, Zhou and Efros2017), have found broad utility across various scientific disciplines due to their powerful model-free capabilities (Ahmad et al., Reference Ahmad, Ali, Shah and Azmat2022). From initial implementations in biomedical imaging to astronomy and particle physics, GANs are increasingly used to enhance quality, fill gaps in incomplete datasets, and simulate complex systems where traditional modeling approaches are insufficient or computationally intensive (Dash et al., Reference Dash, Ye and Wang2023). However, traditional GAN approaches require paired image sets, which are difficult to obtain in sSNOM, due to sample drift. Cycle-consistent GANs (CycleGAN) sidestep this issue as the models are trained on unpaired datasets that do not require the same scan areas, or even the same samples (Zhu et al., Reference Zhu, Park, Isola and Efros2017).

Recently, CycleGANs have been applied to sSNOM images of C. reinhardtii cell slices, achieving a denoising performance equivalent to a fourfold increase in pixel integration time (Baiz et al., Reference Baiz, Kanevche, Kozuch and Heberle2025). The nm-scale absorption features from cellular proteins were significantly enhanced (Figure 4). Furthermore, the method operates on arbitrary image sizes via a patch-based reconstruction technique, preserving detailed spatial information on a subimage level, while enabling larger area scans.

Figure 4. Example machine learning processing of cell slices. The same cell is scanned with two different pixel integration times (left, center columns). The noisy image is then processed using a CycleGAN ML model (right column). The optical signals correspond to imaging at 1655 cm-1 in resonance with the amide I mode of proteins. Figure reprinted from Baiz, C. R., Kanevche, K., Kozuch, J., & Heberle, J. (2025) The Journal of Chemical Physics, 162(5), with the permission of AIP Publishing.

Sparse-sampling techniques are also being adapted from AFM into other probe microscopy methods for optimizing scans, in a way to maximize the information contained in each pixel (Das et al., Reference Das, Badal, Rahman, Islam, Sarker and Paul2019; Fu et al., Reference Fu, Xu, Zhang, Ruta, Pack and Mayer2024). Nonraster scanning and sparse-sampling approaches combined with ML-driven reconstruction algorithms facilitate image reconstruction from significantly reduced data points, further enhancing acquisition capabilities and reducing the data acquisition bottleneck (Liu et al., Reference Liu, Sun, Lu, Wang, Sun, Wang, Lu and Zeng2019). These methods exploit spatial and spectral correlations, decreasing overall scanning time. Such ML-based analyses of sSNOM images could further benefit from utilizing orthogonal AFM-based recording modes that can be operated simultaneously. For instance, sSNOM images can be subject to artefacts due to varying tip–sample interaction, because of a heterogeneous surface potential. Combining sSNOM with Kelvin probe force microscopy (where the AFM tip voltage is biased during a measurement) enabled compensation for such undesired effects (Nörenberg et al., Reference Nörenberg, Wehmeier, Lang, Kehr and Eng2021); however, such orthogonal data have not been subjected to ML-based analyses to our knowledge.

Hyperspectral IR nanoscopy combines aspects of both nanoFTIR and sSNOM, capturing full spectral information across multiple spatial locations. Although in practice, hyperspectral acquisition carries challenges related to sample drift, maintaining detector cooling, and AFM probe integrity, given the long data acquisition times (i.e., in the order of several hours, depending on the area of interest; Amenabar et al., Reference Amenabar, Poly, Goikoetxea, Nuansing, Lasch and Hillenbrand2017). Additionally, hyperspectral imaging yields inherently complex, high-dimensional datasets, making manual analysis challenging. However, automation and ML offer promising means to improve both hyperspectral acquisition and analysis. Autonomous adaptive data acquisition has been applied to scanning microscopy (Holman et al., Reference Holman, Krishnan, Holman, Holman and Sternberg2023), and thus could be employed in hyperspectral IR nanoscopy. Similarly, deep-learning models can greatly improve hyperspectral data acquisition (Gao et al., Reference Gao, Zheng, Meng, Wang, You, Zhong and He2024; Keogan et al., Reference Keogan, Nguyen, Bouzy, Stone, Jirstrom, Rahman and Meade2025), and when combined with ML denoising algorithms (Baiz et al., Reference Baiz, Kanevche, Kozuch and Heberle2025), the increased signal-to-noise ratio can consequently reduce data acquisition times. Furthermore, the ability of ML algorithms to recognize patterns, highlight correlations, and extract meaningful information from large datasets makes them particularly well suited for the analysis of the rich chemical datasets generated by hyperspectral IR nanoscopy (Chen et al., Reference Chen, Yao, Xu, McLeod, Gilbert Corder, Zhao and Carr2021, Reference Chen, Xu, Shabani, Zhao, Fu, Millis and Basov2023). Specifically, supervised learning techniques have been successfully employed for feature identification to facilitate spectral decomposition and have been shown to assist classification of microorganisms (Xue et al., Reference Xue, Wang, Tan, Ma, Liu and Xi2024).

The potential for integration of ML algorithms into sSNOM extends beyond mere noise reduction and accelerated imaging. ML can enable automated, hyperspectral analysis, transforming large datasets into biochemical insights. Similar to developments in far-field IR microscopy, the combination of advanced computation and data-rich high-resolution imaging further holds promise for biomedical applications, such as ‘digital staining’ for subcellular diagnostics (Amrania et al., Reference Amrania, Drummond, Coombes, Shousha, Woodley-Barker, Weir, Hart, Carter and Phillips2016; Bhargava, Reference Bhargava2023).

Outlook

The studies discussed in this perspective emphasize that IR nanoscopy is a mature technology, with a vast potential for future applications. With this outlook, we offer our insights on how novel instrumentation development, in combination with emerging technologies, can be a path forward to both biophysics and biomedical breakthroughs.

We envision a major focus will be the chemical imaging of subcellular components and single cells in aqueous conditions, utilizing invertible liquid containers (Kaltenecker et al., Reference Kaltenecker, Gölz, Bau and Keilmann2021) or liquid sSNOM setups (O’Callahan et al., Reference O’Callahan, Park, Novikova, Jian, Chen, Muller and Lea2020; Pfitzner and Heberle, Reference Pfitzner and Heberle2020). IR nanoscopy does not yet have the capability to capture the temporal dynamics of biomolecular systems. This is primarily due to the incompatibility of current IR nanoscopy instrumentation with high-speed AFM (HS-AFM). In HS-AFM, the scanning speed is sufficiently fast to spatially resolve dynamic processes in real-time, on millisecond to second timescales (Ando et al., Reference Ando, Uchihashi and Kodera2013). Therefore, developing an sSNOM system based on HS-AFM technology would provide a way to perform in vivo IR nanoscopic observations. An alternative approach to spatially visualize molecular dynamics is via cryogenically preserving samples at specific time points after the induction of a metabolic process, then analyzing by IR chemical imaging. Cryogenic ultramicrotomy can also yield high-quality cellular thin sections while avoiding invasive resin embedding and thus preserving the native chemical composition.

A route to accessing dynamics on sub-picosecond timescales is via ultrafast nanospectroscopy, an approach combining scanning probe microscopy with pump-probe spectroscopic methods (Jiang et al., Reference Jiang, Kravtsov, Tokman, Belyanin and Raschke2019). To date, applications of ultrafast nanoscopy are primarily focused on 2D materials, as summarized in the recent review (Zhao et al., Reference Zhao, Kravtsov, Wang, Zhou, Dou, Huang and Jiang2025). Similarly, pump-probe schemes employed in 2D spectroscopy (Xie et al., Reference Xie, Zhang, Janzen, Edgar and Xu2024; Xie and Xu, Reference Xie and Xu2025) offer information on the time evolution of vibrational bands and their correlation could be adapted to nanoscopy. Implementing this technology to address molecular dynamics on ultrafast timescales would provide unprecedented insight into the innerworkings of cellular building blocks.

The simultaneous multimodal data acquisition of AFM and IR nanoscopy allows sample topography to be recorded alongside chemical content. This is highly applicable for the characterization of biological samples, and can be extended to complementary techniques such as nanoscopy in the visible and THz spectral region with Kelvin probe force microscopy (Jakob et al., Reference Jakob, Li, Zhou and Xu2021) in order to capture the surface potential of the sample or via photo-induced force microscopy (Shcherbakov et al., Reference Shcherbakov, Potma, Sugawara, Nowak, Stepanova, Davies and Wickramasinghe2025) to access the dielectric properties of the sample. When applied to medical applications, such as tissue imaging, IR nanoscopy could be used correlatively with laser-based IR microscopy, allowing for rapid chemical imaging of a large field of view, followed by targeted IR-spectral analysis to resolve the chemical content on inter- and subcellular levels.

Due to the nondestructive nature of this technique, the broad range of compatible biological samples, the potential for correlative studies, and the unique capability to simultaneously derive morphological and chemical information, IR nanoscopic chemical imaging is a particularly exciting technology. The ongoing use of IR nanoscopy continues to advance the fields of biophysics and subcellular imaging.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/qrd.2025.10014.

Data availability statement

No new data were generated for this perspective.

Author contribution

J.H. conceived the article. K.K., D.B., J.D., J.K., C.B., and A.E. wrote the draft. All authors discussed and edited the manuscript.

Financial support

J.H. acknowledges funding from the German Research Foundation (DFG, project HE 2063/5-1, and from the SFB 1349 ‘Fluor-Specific Interactions: Fundamentals and Functions’, project number 387284271 – project C05). J.H. and J.K. acknowledge the support from the SupraFAB research building realized with funds from the federal government and the state of Berlin. J.K. acknowledges support from the DFG through the Individual Research Grant (Project No. 500707750). A.E. gratefully acknowledges financial support from the Volkswagen Foundation Freigeist Program and from the Bundesministerium für Wirtschaft und Energie (Projektträger Deutsches Zentrum für Luft- und Raumfahrt) grant numbers 50WB2023 and 50WB2323. C.B. acknowledges Funding from the National Institutes of Health (R35GM133359) as well as the Welch Foundation (F-1881) and a Fellowship from the Alexander von Humboldt Foundation.

Competing interests

The authors declare none.

References

Ahmad, W, Ali, H, Shah, Z and Azmat, S (2022) A new generative adversarial network for medical images super resolution. Scientific Reports 12(1), 9533.10.1038/s41598-022-13658-4CrossRefGoogle ScholarPubMed
Alcolombri, U, Pioli, R, Stocker, R and Berry, D (2022) Single-cell stable isotope probing in microbial ecology. ISME Communications 2(1), 55.10.1038/s43705-022-00142-3CrossRefGoogle ScholarPubMed
Allison, DP, Mortensen, NP, Sullivan, CJ and Doktycz, MJ (2010) Atomic force microscopy of biological samples. Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology 2(6), 618634.Google ScholarPubMed
Amenabar, I, Poly, S, Goikoetxea, M, Nuansing, W, Lasch, P and Hillenbrand, R (2017) Hyperspectral infrared nanoimaging of organic samples based on Fourier transform infrared nanospectroscopy. Nature Communications 8(1), 14402.CrossRefGoogle ScholarPubMed
Amenabar, I, Poly, S, Nuansing, W, Hubrich, EH, Govyadinov, AA, Huth, F, Krutokhvostov, R, Zhang, L, Knez, M, Heberle, J, Bittner, AM, Hillenbrand, R (2013) Structural analysis and mapping of individual protein complexes by infrared nanospectroscopy. Nature Communications 4(1), 2890.10.1038/ncomms3890CrossRefGoogle ScholarPubMed
Amrania, H, Drummond, L, Coombes, R, Shousha, S, Woodley-Barker, L, Weir, K, Hart, W, Carter, I and Phillips, C (2016) New IR imaging modalities for cancer detection and for intra-cell chemical mapping with a sub-diffraction mid-IR s-SNOM. Faraday Discussions 187, 539553.10.1039/C5FD00150ACrossRefGoogle ScholarPubMed
Ando, T, Uchihashi, T and Kodera, N (2013) High-speed AFM and applications to biomolecular systems. Annual Review of Biophysics 42(1), 393414.10.1146/annurev-biophys-083012-130324CrossRefGoogle Scholar
Bai, Y, Guo, Z, Pereira, FC, Wagner, M and Cheng, J-X (2023) Mid-infrared Photothermal–fluorescence in situ hybridization for functional analysis and genetic identification of single cells. Analytical Chemistry 95(4), 23982405.10.1021/acs.analchem.2c04474CrossRefGoogle ScholarPubMed
Baiz, CR, Kanevche, K, Kozuch, J and Heberle, J (2025) Data-driven signal-to-noise enhancement in scattering near-field infrared microscopy. The Journal of Chemical Physics 162(5).10.1063/5.0247251CrossRefGoogle ScholarPubMed
Ballout, F, Krassen, H, Kopf, I, Ataka, K, Bründermann, E, Heberle, J and Havenith, M (2011) Scanning near-field IR microscopy of proteins in lipid bilayers. Physical Chemistry Chemical Physics 13(48), 2143221436.CrossRefGoogle ScholarPubMed
Bechtel, HA, Johnson, SC, Khatib, O, Muller, EA and Raschke, MB (2020) Synchrotron infrared nano-spectroscopy and-imaging. Surface Science Reports 75(3), 100493.10.1016/j.surfrep.2020.100493CrossRefGoogle Scholar
Berweger, S, Nguyen, DM, Muller, EA, Bechtel, HA, Perkins, TT and Raschke, MB (2013) Nano-chemical infrared imaging of membrane proteins in lipid bilayers. Journal of the American Chemical Society 135(49), 1829218295.10.1021/ja409815gCrossRefGoogle ScholarPubMed
Bhargava, R (2023) Digital histopathology by infrared spectroscopic imaging. Annual Review of Analytical Chemistry 16(1), 205230.10.1146/annurev-anchem-101422-090956CrossRefGoogle ScholarPubMed
Burr, DJ, Drauschke, J, Kanevche, K, Kümmel, S, Stryhanyuk, H, Heberle, J and Elsaesser, A (2024) Stable isotope probing-nanoFTIR for quantitation of cellular metabolism and observation of growth-dependent spectral features. Small 20(36), 2400289.10.1002/smll.202400289CrossRefGoogle ScholarPubMed
Carmichael, M, Vidu, R, Maksumov, A, Palazoglu, A and Stroeve, P (2004) Using wavelets to analyze AFM images of thin films: Surface micelles and supported lipid bilayers. Langmuir 20(26), 1155711568.CrossRefGoogle ScholarPubMed
Cedervall, T, Lynch, I, Lindman, S, Berggård, T, Thulin, E, Nilsson, H and Linse, S (2007) Understanding the nanoparticle–protein corona using methods to quantify exchange rates and affinities of proteins for nanoparticles. Proceedings of the National Academy of Sciences 104(7), 20502055.10.1073/pnas.0608582104CrossRefGoogle ScholarPubMed
Chen, Y, Cai, J, Zhao, T, Wang, C, Dong, S, Luo, S and Chen, ZW (2005) Atomic force microscopy imaging and 3-D reconstructions of serial thin sections of a single cell and its interior structures. Ultramicroscopy 103(3), 173182.CrossRefGoogle ScholarPubMed
Chen, X, Xu, S, Shabani, S, Zhao, Y, Fu, M, Millis, AJ and Basov, D (2023) Machine learning for optical scanning probe nanoscopy. Advanced Materials 35(34), 2109171.CrossRefGoogle ScholarPubMed
Chen, X, Yao, Z, Xu, S, McLeod, AS, Gilbert Corder, SN, Zhao, Y and Carr, GL (2021) Hybrid machine learning for scanning near-field optical spectroscopy. ACS Photonics 8(10), 29872996.CrossRefGoogle Scholar
Das, SK, Badal, FR, Rahman, MA, Islam, MA, Sarker, SK and Paul, N (2019) Improvement of alternative non-raster scanning methods for high speed atomic force microscopy: A review. IEEE Access 7, 115603115624.10.1109/ACCESS.2019.2936471CrossRefGoogle Scholar
Dash, A, Ye, J and Wang, G (2023) A review of generative adversarial networks (GANs) and its applications in a wide variety of disciplines: From medical to remote sensing. IEEE Access 12, 1833018357.CrossRefGoogle Scholar
Dazzi, A, Prater, CB, Hu, Q, Chase, DB, Rabolt, JF and Marcott, C (2012) AFM–IR: Combining atomic force microscopy and infrared spectroscopy for nanoscale chemical characterization. Applied Spectroscopy 66(12), 13651384.10.1366/12-06804CrossRefGoogle ScholarPubMed
de Carvalho, LAB, Cinque, G, de Carvalho, ALB, Marques, J, Frogley, MD, Vondracek, H and Marques, MPM (2024) Synchrotron nano-FTIR spectroscopy for probing anticancer drugs at subcellular scale. Scientific Reports 14(1), 17166.10.1038/s41598-024-67386-yCrossRefGoogle ScholarPubMed
Denk, W and Horstmann, H (2004) Serial block-face scanning electron microscopy to reconstruct three-dimensional tissue nanostructure. PLoS Biology 2(11), e329.10.1371/journal.pbio.0020329CrossRefGoogle ScholarPubMed
Doane, TL and Burda, C (2012) The unique role of nanoparticles in nanomedicine: Imaging, drug delivery and therapy. Chemical Society Reviews 41(7), 28852911.10.1039/c2cs15260fCrossRefGoogle ScholarPubMed
Elsaesser, A, Barnes, CA, McKerr, G, Salvati, A, Lynch, I, Dawson, KA and Howard, CV (2011) Quantification of nanoparticle uptake by cells using an unbiased sampling method and electron microscopy. Nanomedicine 6(7), 11891198.CrossRefGoogle ScholarPubMed
Elsaesser, A and Howard, CV (2012) Toxicology of nanoparticles. Advanced Drug Delivery Reviews 64(2), 129137.10.1016/j.addr.2011.09.001CrossRefGoogle ScholarPubMed
Elsaesser, A, Taylor, A, de Yanés, GS, McKerr, G, Kim, EM, O‘Hare, E and Howard, CV (2010). Quantification of nanoparticle uptake by cells using microscopical and analytical techniques. Nanomedicine, 5(9), 14471457.10.2217/nnm.10.118CrossRefGoogle ScholarPubMed
Faist, J, Capasso, F, Sivco, DL, Sirtori, C, Hutchinson, AL and Cho, AY (1994) Quantum cascade laser. Science 264(5158), 553556.CrossRefGoogle ScholarPubMed
Fernandes, R, Chowdhary, S, Mikula, N, Saleh, N, Kanevche, K, Berlepsch, HV and Böttcher, C (2022) Cyanine dye coupling mediates self-assembly of a pH sensitive peptide into novel 3D architectures. Angewandte Chemie 134(48), e202208647.10.1002/ange.202208647CrossRefGoogle Scholar
Freitas, RO, Cernescu, A, Engdahl, A, Paulus, A, Levandoski, JE, Martinsson, I and Prinz, CN (2021) Nano-infrared imaging of primary neurons. Cells 10(10), 2559.CrossRefGoogle ScholarPubMed
Fu, M, Xu, S, Zhang, S, Ruta, FL, Pack, J, Mayer, RA, et al. (2024) Accelerated nano-optical imaging through sparse sampling. Nano Letters 24(7), 21492156.10.1021/acs.nanolett.3c03733CrossRefGoogle ScholarPubMed
Gao, Y, Zheng, P, Meng, Z-D, Wang, H-L, You, E-M, Zhong, J-H and He, H (2024) Fast Nano-IR hyperspectral imaging empowered by large-dataset-free miniaturized spatial–spectral network. Analytical Chemistry 96(23), 96109620.10.1021/acs.analchem.4c01211CrossRefGoogle ScholarPubMed
Greaves, GE, Allison, L, Machado, P, Morfill, C, Fleck, RA, Porter, AE and Phillips, CC (2024a) Infrared nanoimaging of neuronal ultrastructure and nanoparticle interaction with cells. Nanoscale 16(12), 61906198.CrossRefGoogle Scholar
Greaves, GE, Kiryushko, D, Auner, HW, Porter, AE and Phillips, CC (2023) Label-free nanoscale mapping of intracellular organelle chemistry. Communications Biology 6(1), 583.10.1038/s42003-023-04943-7CrossRefGoogle ScholarPubMed
Greaves, GE, Pinna, A, Taylor, JM, Porter, AE and Phillips, CC (2024b) In depth mapping of mesoporous silica nanoparticles in malignant glioma cells using scattering-type scanning near-field optical microscopy. Chemical & biomedical imaging 2(12), 842849.CrossRefGoogle Scholar
Herrmann, FC (2021) Easy ultrastructural insight into the internal morphology of biological specimens by atomic force microscopy. Scientific Reports 11(1), 10214.10.1038/s41598-021-89633-2CrossRefGoogle ScholarPubMed
Hillenbrand, R, Abate, Y, Liu, M, Chen, X and Basov, D (2025) Visible-to-THz near-field nanoscopy. Nature Reviews Materials, 126.Google Scholar
Holman, EA, Krishnan, H, Holman, DR, Holman, H-YN and Sternberg, PW (2023) Toward implementing autonomous adaptive data acquisition for scanning hyperspectral imaging of biological systems. Applied Physics Reviews 10(1), 011319.10.1063/5.0123278CrossRefGoogle Scholar
Horstmann, H, Körber, C, Sätzler, K, Aydin, D and Kuner, T (2012) Serial section scanning electron microscopy (S3EM) on silicon wafers for ultra-structural volume imaging of cells and tissues. PLoS One 7(4), e35172.CrossRefGoogle ScholarPubMed
Hu, X, Zhang, G, Qian, J, , J, Zhu, Y and Peng, Y (2024) Terahertz s-SNOM imaging of a single cell with nanoscale resolution. Nano Letters 24(25), 77577763.10.1021/acs.nanolett.4c01868CrossRefGoogle ScholarPubMed
Huth, F, Govyadinov, A, Amarie, S, Nuansing, W, Keilmann, F and Hillenbrand, R (2012) Nano-FTIR absorption spectroscopy of molecular fingerprints at 20 nm spatial resolution. Nano Letters 12(8), 39733978.10.1021/nl301159vCrossRefGoogle ScholarPubMed
Isola, P, Zhu, J-Y, Zhou, T and Efros, AA (2017) Image-to-image translation with conditional adversarial networks. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition. CVRP.Google Scholar
Jakob, DS, Li, N, Zhou, H and Xu, XG (2021) Integrated tapping mode kelvin probe force microscopy with photoinduced force microscopy for correlative chemical and surface potential mapping. Small 17(37), 2102495.10.1002/smll.202102495CrossRefGoogle ScholarPubMed
Jiang, T, Kravtsov, V, Tokman, M, Belyanin, A and Raschke, MB (2019) Ultrafast coherent nonlinear nanooptics and nanoimaging of graphene. Nature Nanotechnology 14(9), 838843.10.1038/s41565-019-0515-xCrossRefGoogle ScholarPubMed
Kaltenecker, KJ, Gölz, T, Bau, E and Keilmann, F (2021) Infrared-spectroscopic, dynamic near-field microscopy of living cells and nanoparticles in water. Scientific Reports 11(1), 21860.10.1038/s41598-021-01425-wCrossRefGoogle ScholarPubMed
Kanevche, K, Burr, DJ, Nürnberg, DJ, Hass, PK, Elsaesser, A and Heberle, J (2021) Infrared nanoscopy and tomography of intracellular structures. Communications Biology 4(1), 1341. https://doi.org/10.1038/s42003-021-02876-7.CrossRefGoogle ScholarPubMed
Kästner, B, Johnson, CM, Hermann, P, Kruskopf, M, Pierz, K, Hoehl, A and Patoka, P (2018) Infrared nanospectroscopy of phospholipid and surfactin monolayer domains. ACS Omega 3(4), 41414147.CrossRefGoogle ScholarPubMed
Keilmann, F and Amarie, S (2012) Mid-infrared frequency comb spanning an octave based on an Er fiber laser and difference-frequency generation. Journal of Infrared, Millimeter, and Terahertz Waves 33, 479484.10.1007/s10762-012-9894-xCrossRefGoogle Scholar
Keogan, A, Nguyen, TNQ, Bouzy, P, Stone, N, Jirstrom, K, Rahman, A and Meade, AD (2025) Prediction of post-treatment recurrence in early-stage breast cancer using deep-learning with mid-infrared chemical histopathological imaging. npj Precision Oncology 9(1), 18.10.1038/s41698-024-00772-xCrossRefGoogle ScholarPubMed
Keplinger, T, Konnerth, J, Aguié-Béghin, V, Rüggeberg, M, Gierlinger, N and Burgert, I (2014) A zoom into the nanoscale texture of secondary cell walls. Plant Methods 10, 17.10.1186/1746-4811-10-1CrossRefGoogle ScholarPubMed
Kiwilszo, M, Zieliński, A, Smulko, J and Darowicki, K (2012) Improving AFM images with harmonic interference by spectral analysis. Microscopy and Microanalysis 18(1), 186195.10.1017/S1431927611012281CrossRefGoogle ScholarPubMed
Korona, W, Orzechowska, B, Siąkała, K, Nowakowska, AM, Pieczara, A, Buda, S and Baranska, M (2025) Bioorthogonal Raman and IR probes for live cell metabolomics: A library. Sensors and Actuators B: Chemical, 137363.10.1016/j.snb.2025.137363CrossRefGoogle Scholar
Li, Y (2023) An application comparison of Gan-based image translation methods. In 2023 IEEE International Conference on Image Processing and Computer Applications (ICIPCA). IEEE.Google Scholar
Lima, C, Muhamadali, H, Xu, Y, Kansiz, M and Goodacre, R (2021) Imaging isotopically labeled bacteria at the single-cell level using high-resolution optical infrared photothermal spectroscopy. Analytical Chemistry 93(6), 30823088.CrossRefGoogle ScholarPubMed
Liu, Y, Sun, Q, Lu, W, Wang, H, Sun, Y, Wang, Z, Lu, X and Zeng, K (2019) General resolution enhancement method in atomic force microscopy using deep learning. Advanced Theory and Simulations 2(2), 1800137.CrossRefGoogle Scholar
Markiewicz, P and Goh, MC (1994) Atomic force microscopy probe tip visualization and improvement of images using a simple deconvolution procedure. Langmuir 10(1), 57.10.1021/la00013a002CrossRefGoogle Scholar
Müller, DJ and Dufrêne, YF (2011) Atomic force microscopy: A nanoscopic window on the cell surface. Trends in Cell Biology 21(8), 461469.10.1016/j.tcb.2011.04.008CrossRefGoogle ScholarPubMed
Musat, N, Musat, F, Weber, PK and Pett-Ridge, J (2016) Tracking microbial interactions with NanoSIMS. Current Opinion in Biotechnology 41, 114121.10.1016/j.copbio.2016.06.007CrossRefGoogle ScholarPubMed
Nogales, E and Mahamid, J (2024) Bridging structural and cell biology with cryo-electron microscopy. Nature 628(8006), 4756.10.1038/s41586-024-07198-2CrossRefGoogle ScholarPubMed
Nörenberg, T, Wehmeier, L, Lang, D, Kehr, SC and Eng, LM (2021) Compensating for artifacts in scanning near-field optical microscopy due to electrostatics. APL Photonics 6(3), 036102.10.1063/5.0031395CrossRefGoogle Scholar
O’Callahan, BT, Park, K-D, Novikova, IV, Jian, T, Chen, C-L, Muller, EA and Lea, AS (2020) In liquid infrared scattering scanning near-field optical microscopy for chemical and biological nanoimaging. Nano Letters 20(6), 44974504.10.1021/acs.nanolett.0c01291CrossRefGoogle ScholarPubMed
Ocelic, N, Huber, A and Hillenbrand, R (2006) Pseudoheterodyne detection for background-free near-field spectroscopy. Applied Physics Letters 89(10), 101124.10.1063/1.2348781CrossRefGoogle Scholar
Paul, S, Jeništová, A, Vosough, F, Berntsson, E, Mörman, C, Jarvet, J and Barth, A (2023) 13C-and 15N-labeling of amyloid-β and inhibitory peptides to study their interaction via nanoscale infrared spectroscopy. Communications Chemistry 6(1), 163.10.1038/s42004-023-00955-wCrossRefGoogle Scholar
Pfitzner, E and Heberle, J (2020) Infrared scattering-type scanning near-field optical microscopy of biomembranes in water. The Journal of Physical Chemistry Letters 11(19), 81838188.10.1021/acs.jpclett.0c01769CrossRefGoogle ScholarPubMed
Rahman Laskar, MA and Celano, U (2023) Scanning probe microscopy in the age of machine learning. APL Machine Learning 1(4), 041501.10.1063/5.0160568CrossRefGoogle Scholar
Sahl, SJ, Hell, SW and Jakobs, S (2017) Fluorescence nanoscopy in cell biology. Nature Reviews Molecular Cell Biology 18(11), 685701.10.1038/nrm.2017.71CrossRefGoogle ScholarPubMed
Schaible, GA, Kohtz, AJ, Cliff, J and Hatzenpichler, R (2022) Correlative SIP-FISH-Raman-SEM-NanoSIMS links identity, morphology, biochemistry, and physiology of environmental microbes. ISME Communications 2(1). https://doi.org/10.1038/s43705-022-00134-3.CrossRefGoogle ScholarPubMed
Schimmack, M and Mercorelli, P (2018) A wavelet packet tree Denoising algorithm for images of atomic-force microscopy. Asian Journal of Control 20(4), 13671378.10.1002/asjc.1718CrossRefGoogle Scholar
Shcherbakov, MR, Potma, EO, Sugawara, Y, Nowak, D, Stepanova, M, Davies, PR and Wickramasinghe, HK (2025) Photo-induced force microscopy. Nature Reviews Methods Primers 5(1), 34.10.1038/s43586-025-00403-0CrossRefGoogle Scholar
Stanciu, SG, Tranca, DE, Hristu, R and Stanciu, GA (2017) Correlative imaging of biological tissues with apertureless scanning near-field optical microscopy and confocal laser scanning microscopy. Biomedical Optics Express 8(12), 53745383.10.1364/BOE.8.005374CrossRefGoogle ScholarPubMed
Udpa, L, Ayres, VM, Fan, Y, Chen, Q and Kumar, SA (2006) Deconvolution of atomic force microscopy data for cellular and molecular imaging. IEEE Signal Processing Magazine 23(3), 7383.10.1109/MSP.2006.1628880CrossRefGoogle Scholar
Veber, A, Zancajo, VM, Puskar, L, Schade, U and Kneipp, J (2023) In situ infrared imaging of the local orientation of cellulose fibrils in plant secondary cell walls. Analyst 148(17), 41384147.10.1039/D3AN00897ECrossRefGoogle ScholarPubMed
Wang, Y, Huang, WE, Cui, L and Wagner, M (2016) Single cell stable isotope probing in microbiology using Raman microspectroscopy. Current Opinion in Biotechnology 41, 3442.10.1016/j.copbio.2016.04.018CrossRefGoogle ScholarPubMed
Xie, Q and Xu, XG (2025) Action-based two-dimensional infrared spectroscopy on the horizon. The Journal of Chemical Physics 162(10), 100901.10.1063/5.0244011CrossRefGoogle ScholarPubMed
Xie, Q, Zhang, Y, Janzen, E, Edgar, JH and Xu, XG (2024) Atomic-force-microscopy-based time-domain two-dimensional infrared nanospectroscopy. Nature Nanotechnology 19(8), 11081115.10.1038/s41565-024-01670-wCrossRefGoogle ScholarPubMed
Xue, Y, Wang, Y, Tan, W, Ma, L, Liu, X and Xi, N (2024) Microbial classification by multimodal atomic force microscopy with machine learning. In 2024 IEEE 17th International Conference on Nano/Molecular Medicine & Engineering (NANOMED). IEEE.Google Scholar
Zhao, Z, Kravtsov, V, Wang, Z, Zhou, Z, Dou, L, Huang, D and Jiang, T (2025) Applications of ultrafast nano-spectroscopy and nano-imaging with tip-based microscopy. eLight 5(1), 1.10.1186/s43593-024-00079-1CrossRefGoogle Scholar
Zhu, J-Y, Park, T, Isola, P and Efros, AA (2017) Unpaired image-to-image translation using cycle-consistent adversarial networks. In Proceedings of the IEEE International Conference on Computer Vision. CVPR.10.1109/ICCV.2017.244CrossRefGoogle Scholar
Figure 0

Figure 1. Technical overview of IR nanoscopy. Typical sSNOM (1) and nanoFTIR (2) setup based on an asymmetric Michelson interferometer. An IR light source is focused on an AFM tip, oscillating at frequency Ω, via parabolic mirror (PM). The backscattered light is collected via the same pathway, recombined with the reference beam at the beam splitter (BS) and focused on a mercury-cadmium-telluride (MCT) detector. This technique allows for recording the AFM topography, IR phase, and amplitude, which reveal the absorption and reflection of the specimen. Fourier transformation of the interferograms yields nanoFTIR spectra, thus capturing IR absorption across broad spectral range. Figure adapted from (Kanevche et al., 2021).

Figure 1

Table 1. Summary of AFM-based mid-IR nanoscopy techniques

Figure 2

Figure 2. Subcellular nanoscopy examples. A. AFM and sSNOM imaging (at 638 nm) of thin-sectioned zebrafish (Danio rerio) retina. B. AFM and sSNOM imaging (at wavelengths of interest) of thin-sectioned green algae (C. reinhardtii) cells, primarily revealing the protein distribution within the cell. C. sSNOM tomography constructed from ten sequential images of C. reinhardtii, imaged at 1655 cm-1. Video demonstrating the 3D structure of the green algae cell is available here: https://static-content.springer.com/esm/art%3A10.1038%2Fs42003-021-02876-7/MediaObjects/42003_2021_2876_MOESM4_ESM.mpg. D. Absorption spectrum of myeloma thin sections and sSNOM images at four wavelengths of interest, revealing the subcellular components. Figure adapted with permission from Stanciu et al., 2017 © Optical Society of America, and (Greaves et al., 2023; Kanevche et al., 2021).

Figure 3

Figure 3. Applications of IR imaging for monitoring cellular processes. A. SIP-nanoFTIR as a means of deriving single-cell glucose uptake in E. coli. B. Correlative sSNOM absorption, sSNOM reflection (acquired at 1740 cm-1), and TEM imaging of AuNP interaction with hippocampal neurons. C. MSNP-treated malignant glioma cell imaged with sSNOM and AFM. Inset (ii) shows the MSNP-associated region, with nanoparticles visible at 1100 cm-1, but not at 1300 cm-1 nor AFM topography. Figure adapted from (Burr et al., 2024; Greaves, Allison, et al., 2024a) and (Greaves, Pinna, et al., 2024b).

Figure 4

Figure 4. Example machine learning processing of cell slices. The same cell is scanned with two different pixel integration times (left, center columns). The noisy image is then processed using a CycleGAN ML model (right column). The optical signals correspond to imaging at 1655 cm-1 in resonance with the amide I mode of proteins. Figure reprinted from Baiz, C. R., Kanevche, K., Kozuch, J., & Heberle, J. (2025) The Journal of Chemical Physics, 162(5), with the permission of AIP Publishing.

Author comment: Infrared nanoscopy for subcellular chemical imaging — R0/PR1

Comments

Dear Dr. Paskins, dear Prof. Nordén,

The attached manuscript is being submitted in response to your invitation on March 9th to submit a perspective article to QRB Discovery. The title of our manuscript is

Infrared nanoscopy for subcellular chemical imaging

With the authors Katerina Kanevche, David J. Burr, Janina Drauschke, Jacek Kozuch, Carlos Baiz, Andreas Elsaesser, and myself.

Our manuscript provides an overview of a new technique that we believe is trans-formative for biophysical science. Infrared nanoscopy bridges the gap between vibrational spectroscopy and nanoscale imaging, providing molecular insights into cellular biophysics. Spatial resolution of 20 nm is routinely achieved, overcoming the diffraction limit of conventional infrared (IR) microscopy by far. Here, we focus on sSNOM (scattering-type scanning near-field optical microscopy), which uses the sharp tip of an atomic force microscope to scatter infrared radiation. IR absorption of a biological sample is detected in the optical near-field regime. We briefly discuss the complementary technique of AFM-IR, which uses pulsed IR lasers to detect thermal expansion from absorbed IR radiation. Both approaches correlate nanoscale morphology with chemical composition to yield non-destructive, label-free chemical images of biological cells. Our manuscript concludes with perspectives on future experiments.

I hope you share our excitement about this work and will send our work for in-depth review. We appreciate your interest in our work.

Sincerely yours

Joachim Heberle

Review: Infrared nanoscopy for subcellular chemical imaging — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

In this perspective, Kanevche et al. introduce IR nanoscopy, recent applications of the technology, and future opportunities. Overall, this is a well-balanced review that will educate biologists and vibrational spectroscopists outside of the field of IR nanoscopy and inspire them to enter the field. I particularly appreciate the discussions on how to prepare samples for imaging as this is a critical part of working with these technologies. The IR imaging field is rapidly developing; hence, it is timely to gain a perspective from several of the leaders in the field. I therefore recommend this perspective for publication.

I have a few minor suggestions:

1. A major advantage of IR nanoscopy over other vibrational approaches (and even many fluorescence approaches) is the high spatial resolution (~20 nm). It is unclear that many of the applications discussed required this level of spatial resolution. Why was sNOM the ideal approach to answer the question? I suggest more clearly highlighting where the spatial resolution provided details not available by other approaches or adding additional examples that required the high spatial resolution.

2. While there is a good discussion of strengths and limitations of the technique, two major limitations that could be more clearly addressed are the water background and time to collect images. Water is a problem for most mid-IR approaches. How does sNOM handle this? My understanding is that nanoscopy is quite slow. A transparent discussion of the imaging times would be approached. It may even be appropriate to put these items in context with other approaches – for example the issue of water in sNOM vs other IR approaches, imaging times vs other nanoscale measurements.

3. In the section title “Nanoscale IR Biomarker Monitoring” it was unclear what nanoparticle interactions with the cell had to do with the section title. I suggest creating a more inclusive title (perhaps matching the subtitle of Figure 3) or creating a new subsection.

4. IR shifts are mentioned in the section on “Nanoscale IR Biomarker Monitoring” and the origin of the effect is not explained. I suggest adding a sentence and reference about the shift being related to changes in reduced mass and the harmonic oscillator. This would assist a non-vibrational audience who might not understand the origin of the effect.

5. In a couple of places it is emphasized that nanoscopy is “non-destructive,” however, in the section on live cells it is discussed that live cells can’t be used in contact mode because it will be destructive to the sample. Workarounds are discussed. Therefore, I suggest using relatively non-destructive or acknowledging this limitation more clearly for soft samples.

6. A style suggestion is to capitalize the words in aconyms to distinguish words that make up the acronym.

Review: Infrared nanoscopy for subcellular chemical imaging — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

In general, the manuscript is very well written, particularly in providing an overview and comparison of several techniques such as s-SNOM, AFM-IR, nano-FTIR, and hyperspectral nano-IR. However, it is not always clear what the differences are between these methods. The review would benefit greatly if the authors could systematically compare the techniques—clarifying what each one measures, their respective limitations, and where they are complementary. For example, while the authors mention that AFM-IR and s-SNOM are complementary, a deeper explanation of this complementarity is missing.

The article would be significantly strengthened by including a comparative table summarizing the main features, advantages, limitations, and application areas of all the techniques discussed. In addition, a schematic or cartoon illustrating the key technical differences would make the review more accessible, especially for readers from the biological sciences. At present, the emphasis of the review seems somewhat weighted toward s-SNOM, and balancing the coverage across techniques would improve the overall scope.

It would also be valuable to more explicitly address the main bottlenecks in advancing biophotonics for biological imaging. These include challenges of data interpretation in complex cellular contexts as well as the practical advantages and disadvantages of operating in contact with biological samples using an AFM tip (e.g., issues of reproducibility, mechanical fragility, and the presence of sticky particles). To date, these methods have been successfully applied to isotope-labeled IR tags, but this is not always feasible, and such limitations should be acknowledged. In particular, it remains unclear to what extent these techniques can be applied to tissue samples and how nano-IR approaches can be combined with other modalities to enable true multimodal analysis. A discussion of these aspects would provide readers with a more realistic perspective on the current capabilities and future challenges of IR nanoscopy in biology.

Recommendation: Infrared nanoscopy for subcellular chemical imaging — R0/PR4

Comments

No accompanying comment.

Decision: Infrared nanoscopy for subcellular chemical imaging — R0/PR5

Comments

No accompanying comment.

Author comment: Infrared nanoscopy for subcellular chemical imaging — R1/PR6

Comments

see uploaded PDF as supplementary material

Review: Infrared nanoscopy for subcellular chemical imaging — R1/PR7

Conflict of interest statement

Reviewer declares none.

Comments

The authors have thoughtfully incorporated feedback from the reviewers.

Review: Infrared nanoscopy for subcellular chemical imaging — R1/PR8

Conflict of interest statement

Competing interests

Comments

Please check and correct if needed:

In Figures 1 and 2, the AFM topography of a cells is shown, but the corresponding IR image appears to represent most likepy a resin-embedded cell that was cut into an ultrathin section (most likely TEM-thin). In my understanding, in that case, the topography should appear relatively flat.

Recommendation: Infrared nanoscopy for subcellular chemical imaging — R1/PR9

Comments

One Reviewer had a comment that Authors may wish to consider:

In Figures 1 and 2, the AFM topography of a cells is shown, but the corresponding IR image appears to represent most likepy a resin-embedded cell that was cut into an ultrathin section (most likely TEM-thin). In my understanding, in that case, the topography should appear relatively flat.

Decision: Infrared nanoscopy for subcellular chemical imaging — R1/PR10

Comments

No accompanying comment.