Infrared Nanospectroscopy (AFM-IR)

Last updated
An atomic-force microscope with its controlling computer Atomic force microscope by Zureks.jpg
An atomic-force microscope with its controlling computer

AFM-IR (atomic force microscope-infrared spectroscopy) or infrared nanospectroscopy is one of a family of techniques [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] that are derived from a combination of two parent instrumental techniques. AFM-IR combines the chemical analysis power of infrared spectroscopy and the high-spatial resolution of scanning probe microscopy (SPM). The term was first used to denote a method that combined a tuneable free electron laser with an atomic force microscope (AFM, a type of SPM) equipped with a sharp probe that measured the local absorption of infrared light by a sample with nanoscale spatial resolution. [16] [17] [18]

Contents

Originally the technique required the sample to be deposited on an infrared-transparent prism and be less than 1μm thick. This early setup improved the spatial resolution and sensitivity of photothermal AFM-based techniques from microns [7] to circa 100 nm. [8] [9] [10] [16] [19] [20] Then, the use of modern pulsed optical parametric oscillators and quantum cascade lasers, in combination with top-illumination, have enabled to investigate samples on any substrate and with increase sensitivity and spatial resolution. As most recent advances, AFM-IR has been proved capable to acquire chemical maps and nanoscale resolved spectra at the single-molecule scale from macromolecular self-assemblies and biomolecules with circa 10 nm diameter, [18] [17] [21] [22] as well as to overcome limitations of IR spectroscopy and measure in aqueous liquid environments. [23]

Recording the amount of infrared absorption as a function of wavelength or wavenumber, AFM-IR creates an infrared absorption spectra that can be used to chemically characterize and even identify unknown samples. [12] [15] [24] Recording the infrared absorption as a function of position can be used to create chemical composition maps that show the spatial distribution of different chemical components. Novel extensions of the original AFM-IR technique [18] [17] and earlier techniques [1] [2] [3] [4] [6] [7] [24] have enabled the development of bench-top devices capable of nanometer spatial resolution, that do not require a prism and can work with thicker samples, and thereby greatly improving ease of use and expanding the range of samples that can be analysed. AFM-IR has achieved lateral spatial resolutions of ca. 10 nm, with a sensitivity down to the scale of molecular monolayer [25] and single protein molecules with molecular weight down to 400-600 kDa. [18] [17]

AFM-IR is related to techniques such as tip-enhanced Raman spectroscopy (TERS), scanning near-field optical microscopy (SNOM), [26] nano-FTIR and other methods of vibrational analysis with scanning probe microscopy.

History

Early history

Atomic force microscope inside a FTIR spectrometer with the optical interface AFM-FTIR-photograph.jpg
Atomic force microscope inside a FTIR spectrometer with the optical interface

The earliest measurements combining AFM with infrared spectroscopy were performed in 1999 by Hammiche et al. at the University of Lancaster in the United Kingdom, [1] in an EPSRC-funded project led by M Reading and H M Pollock. Separately, Anderson at the Jet Propulsion Laboratory in the United States made a related measurement in 2000. [2] Both groups used a conventional Fourier transform infrared spectrometer (FTIR) equipped with a broadband thermal source, the radiation was focused near the tip of a probe that was in contact with a sample. The Lancaster group obtained spectra by detecting the absorption of infrared radiation using a temperature sensitive thermal probe. Anderson [2] took the different approach of using a conventional AFM probe to detect the thermal expansion. He reported an interferogram but not a spectrum; the first infrared spectrum obtained in this way was reported by Hammiche et al. in 2004: [6] this represented the first proof that spectral information about a sample could be obtained using this approach.

Both of these early experiments used a broadband source in conjunction with an interferometer; these techniques could, therefore, be referred to as AFM-FTIR although Hammiche et al. coined the more general term photothermal microspectroscopy or PTMS in their first paper. [1] PTMS has various subgroups; [27] including techniques that measure temperature [1] [3] [4] [6] [7] [14] [28] measure thermal expansion [2] [6] [8] [9] [10] [11] [12] [13] use broadband sources. [1] [2] [3] [4] [6] [7] use lasers [8] [9] [10] [11] [12] [28] excite the sample using evanescent waves, [8] [9] [10] [11] [15] illuminate the sample directly from above [1] [2] [3] [4] [5] [6] [7] [12] [14] [25] [28] etc. and different combinations of these. Fundamentally, they all exploit the photothermal effect. Different combinations of sources, methods, methods of detection and methods of illumination have benefits for different applications. [6] Care should be taken to ensure that it is clear which form of PTMS is being used in each case. Currently there is no universally accepted nomenclature. The original technique dubbed AFM-IR that induced resonant motion in the probe using a Free Electron Laser has developed by exploiting the foregoing permutations so that it has evolved into various forms.

The pioneering experiments of Hammiche et al and Anderson had limited spatial resolution due to thermal diffusion - the spreading of heat away from the region where the infrared light was absorbed. The thermal diffusion length (the distance the heat spreads) is inversely proportional to the root of the modulation frequency. Consequently, the spatial resolution achieved by the early AFM-IR approaches was around one micron or more, due to the low modulation frequencies of the incident radiation created by the movement of the mirror in the interferometer. Also, the first thermal probes were Wollaston wire devices [1] [2] [3] [4] [5] [6] that were developed originally for Microthermal analysis [29] (in fact PTMS was originally considered to be one of a family of microthermal techniques [4] ). The comparatively large size of these probes also limited spatial resolution. Bozec et al. [3] and Reading et al. [7] used thermal probes with nanoscale dimensions and demonstrated higher spatial resolution. Ye et al [30] described a MEM-type thermal probe giving sub-100 nm spatial resolution, which they used for nanothermal analysis. The process of exploring laser sources began in 2001 by Hammiche et al when they acquired the first spectrum using a tuneable laser (see Resolution improvement with pulsed laser source ).

A significant development was the creation by Reading et al. in 2001 [4] of a custom interface that allowed measurements to be made while illuminating the sample from above; this interface focused the infrared beam to a spot of circa 500μm diameter, close to the theoretical maximum [Note 1] . The use of top-down or top-side illumination has the important benefit that samples of arbitrary thickness can be studied on arbitrary substrates. In many cases this can be done without any sample preparation. All subsequent experiments by Hammiche, Pollock, Reading and their co-workers were made using this type of interface including the instrument constructed by Hill et al. for nanoscale imaging using a pulsed laser. [12] The work of the University of Lancaster group in collaboration with workers from the University of East Anglia led to the formation of a company, Anasys Instruments, to exploit this and related technologies [31] (see Commercialization ).

Spatial resolution improvement with pulsed laser sources

An infrared optical parametric oscillator (OPO), 1997 IR Optical Parametric Oscillator.JPG
An infrared optical parametric oscillator (OPO), 1997
Schematic of the AFM-IR instrument using an OPO light source constructed at the University of East Anglia by Hill et al in 2007 AFM-IR using OPO schematic.jpg
Schematic of the AFM-IR instrument using an OPO light source constructed at the University of East Anglia by Hill et al in 2007

In the first paper on AFM-based infrared by Hammiche et al., [1] the relevant well-established theoretical considerations were outlined that predict that high spatial resolution can be achieved using rapid modulation frequencies because of the consequent reduction in the thermal diffusion length. They estimated that spatial resolutions in the range of 20 nm-30 nm should be achievable. [32] The most readily available sources that can achieve high modulation frequencies are pulsed lasers: even when the rapidity of the pulses is not high, the square wave form of a pulse contains very high modulation frequencies in Fourier space. In 2001, Hammiche et al. used a type of bench-top tuneable, pulsed infrared laser known as an optical parametric oscillator or OPO and obtained the first probe-based infrared spectrum with a pulsed laser, however, they did not report any images. [24]

Nanoscale spatial resolution AFM-IR imaging using a pulsed laser was first demonstrated by Dazzi et al [8] at the University of Paris-Sud, France. Dazzi and his colleagues used a wavelength-tuneable, free electron laser at the CLIO facility [Note 2] in Orsay, France to provide an infrared source with short pulses. Like earlier workers, [2] [6] they used a conventional AFM probe to measure thermal expansion but introduced a novel optical configuration: the sample was mounted on an IR-transparent prism so that it could be excited by an evanescent wave. Absorption of short infrared laser pulses by the sample caused rapid thermal expansion that created a force impulse at the tip of the AFM cantilever. The thermal expansion pulse induced transient resonant oscillations of the AFM cantilever probe. This has led to the technique being dubbed Photo-Thermal Induced Resonance (PTIR), by some workers in the field. [10] [24] Some prefer the terms PTIR or PTMS [1] [3] [5] [6] [7] to AFM-IR as the technique is not necessarily restricted to infrared wavelengths. The amplitude of the cantilever oscillation is directly related to the amount of infrared radiation absorbed by the sample. [33] [34] [35] [36] [37] [38] [39] By measuring the cantilever oscillation amplitude as a function of wavenumber, Dazzi's group was able to obtain absorption spectra from nanoscale regions of the sample. Compared to earlier work, this approach improved spatial resolution because the use of short laser pulses reduced the duration of the thermal expansion pulse to the point that the thermal diffusion lengths can be on the scale of nanometres rather than microns.

FFT catilever vibration after laser pulse.jpg
Fast Fourier Transform of cantilever vibrations after a laser pulse; the height of a characteristic peak measures the amount of infrared light absorbed by the sample
AFM-IR-FTIR.jpg
Spectrum obtained from the AFM measurement by changing the laser wavelength (below); it has good agreement with a conventional FTIR spectrum (above)

A key advantage of the use of a tuneable laser source, with a narrow wavelength range, is the ability to rapidly map the locations of specific chemical components on the sample surface. To achieve this, Dazzi's group tuned their free electron laser source to a wavelength corresponding to the molecular vibration of the chemical of interest, then mapped the cantilever oscillation amplitude as function of position across the sample. They demonstrated the ability to map chemical composition in E. coli bacteria. They could also visualize polyhydroxybutyrate (PHB) vesicles inside Rhodobacter capsulatus cells [35] and monitor the efficiency of PHB production by the cells.

At the University of East Anglia in the UK, as part of an EPSRC-funded project led by M. Reading and S. Meech, Hill and his co-workers [12] followed the earlier work of Reading et al. [4] and Hammiche et al. [6] and measured thermal expansion using an optical configuration that illuminated the sample from above [5] in contrast to Dazzi et al. who excited the sample with an evanescent wave from below. [8] Hill also made use of an optical parametric oscillator as the infrared source in the manner of Hammiche et al. [24] This novel combination of topside illumination, [4] OPO source [24] and measuring thermal expansion [2] [6] [8] proved capable of nanoscale spatial resolution for infrared imaging and spectroscopy (the figures show a schematic of the UEA apparatus and results obtained with it). The use by Hill and co-workers of illumination from above allowed a substantially wider range of samples to be studied than was possible using Dazzi's technique. By introducing the use of a bench top IR source and topdown illumination, the work of Hammiche, Hill and their coworkers made possible the first commercially viable SPM-based infrared instrument (see Commercialization).

Broadband pulsed laser sources

Reading et al. have explored the use of a broadband QCL combined with thermal expansion measurements. [40] Above, the inability of thermal broadband sources to achieve high spatial resolution is discussed (see history). In this case the frequency of modulation is limited by the mirror speed of the interferometer which, in turn, limits the lateral spatial resolution that can be achieved. When using a broadband QCL the resolution is limited not by the mirror speed but by the modulation frequency of the laser pulses (or other waveforms). [1] The benefit of using a broadband source is that an image can be acquired that comprises an entire spectrum or part of a spectrum for each pixel. This is much more powerful than acquiring images bases on a single wavelength. The preliminary results of Reading et al. [40] show that directing a broadband QCL though an interferometer can give an easily detectable response from a conventional AFM probe measuring thermal expansion.

Commercialization

The free-electron laser FELIX at the FOM Institute for Plasma Physics Rijnhuizen Nieuwegein, The Netherlands (2010); a large and uncommon piece of equipment FELIX.jpg
The free-electron laser FELIX at the FOM Institute for Plasma Physics Rijnhuizen Nieuwegein, The Netherlands (2010); a large and uncommon piece of equipment

The AFM-IR technique based on a pulsed infrared laser source was commercialized by Anasys Instruments, a company founded by Reading, Hammiche and Pollock in the United Kingdom in 2004; [31] [41] a sister, United States corporation was founded a year later. Anasys Instruments developed its product with support from the National Institute of Standards and Technology and the National Science Foundation. Since free electron lasers are rare and available only at select institutions, a key to enabling a commercial AFM-IR was to replace them with a more compact type of infrared source. Following the lead given by Hammiche et al in 2001 [24] and Hill et al in 2008, [12] Anasys Instruments introduced an AFM-IR product in early 2010, using a tabletop laser source based on a nanosecond optical parametric oscillator. [36] The OPO source enabled nanoscale infrared spectroscopy over a tuning range of roughly 1000–4000 cm−1 or 2.5-10 μm.

The initial product required samples to be mounted on infrared-transparent prisms, with the infrared light being directed from below in the manner of Dazzi et al. [Note 3] For best operation, this illumination scheme required thin samples, with optimal thickness of less than 1 μm, [24] prepared on the surface of the prism. In 2013, Anasys released an AFM-IR instrument based on the work of Hill et al. [12] [28] that supported top-side illumination. "By eliminating the need to prepare samples on infrared-transparent prisms and relaxing the restriction on sample thickness, the range of samples that could be studied was greatly expanded. The CEO of Anasys Instruments recognised this achievement by calling it " an exciting major advance" in a letter written to the university and included in the final report of EPSRC project EP/C007751/1. [42] The UEA technique went on to become Anasys Instruments' flagship product.

It is worth noting that the first infrared spectrum obtained by measuring thermal expansion using an AFM was obtained by Hammiche and co-workers [6] without inducing resonant motions in the probe cantilever. In this early example the modulation frequency was too low to achieve high spatial resolution but there is nothing, in principle, preventing the measurement of thermal expansion at higher frequencies without analysing or inducing resonant behaviour. [1] Possible options for measuring the displacement of the tip rather than the subsequent propagation of waves along the cantilever include; interferometry focused at the end of the cantilever where the tip is located, a torsional motion resulting from an offset probe (it would only be influenced by the motions of the cantilever as a second order effect) and exploiting the fact that the signal from a heated thermal probe is strongly influenced by the position of the tip relative to the surface thus this could provide a measurement of thermal expansion that wasn't strongly influenced by or dependent upon resonance. The advantages of a non-resonant method of detection is that any frequency of light modulation could be used thus depth information could be obtained in a controlled way (see below) whereas methods that rely on resonance are limited to harmonics. The thermal-probe based method of Hammiche et al. [1] has found a significant number of applications. [14] [28]

A unique application made possible by the top-down illumination combined with a thermal probe [4] is localized depth profiling, [28] this is not possible using either using the Dazzi et al. configuration of AFM-IR or that of Hill et al. despite the fact the latter uses top-down illumination. Obtaining linescans [4] [43] and images [28] with thermal probes has been shown to be possible, sub-diffraction limit spatial resolution can be achieved [4] and the resolution for delineating boundaries can be enhanced using chemometric techniques. [28] [43]

In all of these examples a spectrum is acquired that spans the entire mid-IR range for each pixel, this is considerably more powerful than measuring the absorption of a single wavelength as is the case for AFM-IR when using either the method of Dazzi et al. or Hill et al. Reading and his group demonstrated how, because thermal probes can be heated, localized thermal analysis [4] [28] [29] can be combined with photothermal infrared spectroscopy using a single probe. In this way local chemical information could be complemented with local physical properties such melting and glass transition temperatures. [29] This in turn led to the concept of thermally assisted nanosampling, [5] [28] where the heated tip performs a local thermal analysis experiment then the probe is retracted taking with it down to femtograms [Note 4] of softened material that adhere to the tip. [38] This material can then be manipulated and/or analysed by photothermal infrared spectroscopy or other techniques. [5] [44] [45] [46] [47] This considerably increases the analytical power of this type of SPM-based infrared instrument beyond anything that can be achieved with conventional AFM probes such as those used in AFM-IR when using either the Dazzi et al. or the Hill et al. version.

Thermal probe techniques have still not achieved the nanoscale spatial resolution that thermal expansion methods have attained though this is theoretically possible. For this, a robust thermal probe and a high intensity source is needed. Recently, the first images using a QCL and a thermal probe have been obtained by Reading et al. [40] A good signal to noise ratio enabled rapid imaging but sub-micron spatial resolution was not clearly demonstrated. Theory predicts improvements in spatial resolution could be achieved by confining data analysis to the early part of the thermal response to a step change increase in the intensity of the incident radiation. In this way pollution of the measurement from adjacent regions would be avoided, i.e. the measurement window could be confined to a suitable fraction of the time of flight of the thermal wave (using a Fourier analysis of the response could provide a similar outcome by using the high frequency components). This could be achieved by tapping the probe in synchrony with the laser. Similarly, lasers that provide very rapid modulations could further reduce thermal diffusion lengths.

Although most effort to date has been focused on thermal expansion measurements, this might change. Truly robust thermal probes have recently become available, [48] as have affordable compact QCL's that are tuneable over a broad frequency range. Consequently, it may soon be the case that thermal probe techniques will become as widely used as those based on thermal expansion. Ultimately, instruments that can easily switch between modes and even combine them using a single probe will certainly become available, for example, a single probe will eventually be able to measure both temperature and thermal expansion.

Recent improvements and single-molecule sensitivity

The original commercial AFM-IR instruments required most samples to be thicker than 50 nm to achieve sufficient sensitivity. Sensitivity improvements were achieved using specialized cantilever probes with an internal resonator [49] and by wavelet based signal processing techniques. [50] Sensitivity was further improved by Lu et al. [25] by using quantum cascade laser (QCL) sources. The high repetition rate of the QCL allows absorbed infrared light to continuously excite the AFM tip at a "contact resonance" [Note 5] of the AFM cantilever. This resonance-enhanced AFM-IR, in combination with electric field enhancement from metallic tips and substrates led to the demonstration of AFM-IR spectroscopy and compositional imaging of films as thin as single self-assembled monolayers. [25] AFM-IR has also been integrated with other sources including a picosecond OPO [24] offering a tuning range 1.55 μm to 16 μm (from 6450 cm−1 to 625 cm−1).

In its initial development, with samples deposited on transparent prisms and using OPO laser sources, the sensitivity of AFM-IR was limited to a minimal thickness of the sample of circa 50-100 nm as mentioned above. [8] [16] [33] [51] The advent of quantum cascade lasers (QCL) and the use of the electromagnetic field enhancement between metallic probes and substrates have improved the sensitivity and spatial resolution of AFM-IR down to the measurement of large (>0.3 μm) and flat (~2–10 nm) self-assembled monolayers, where still hundreds of molecules are present. [25] Ruggeri et al. have recently developed off-resonance, low power and short pulse AFM-IR (ORS-nanoIR) to prove the acquisition of infrared absorption spectra and chemical maps at the single molecule level, in the case of macromolecular assemblies [17] [22] [21] and large protein molecules with a spatial resolution of ca. 10 nm. [18]

Nanoscale chemical imaging and mapping

Nanoscale resolved chemical maps and spectra

AFM-IR enables nanoscale infrared spectroscopy, [52] i.e. the ability to obtain infrared absorption spectra from nanoscale regions of a sample.

Chemical compositional mapping AFM-IR can also be used to perform chemical imaging or compositional mapping with spatial resolution down to ~10-20 nm, [18] limited only by the radius of the AFM tip. In this case, the tuneable infrared source emits a single wavelength, corresponding to a specific molecular resonance, i.e. a specific infrared absorption band. By mapping the AFM cantilever oscillation amplitude as a function of position, it is possible to map out the distribution of specific chemical components. Compositional maps can be made at different absorption bands to reveal the distribution of difference chemical species.

Complementary morphological and mechanical mapping

Complementary elasticity mapping via simultaneous contact resonance measurements. Complementary mechanical mapping with AFM-IR.png
Complementary elasticity mapping via simultaneous contact resonance measurements.

The AFM-IR technique can simultaneously provide complementary measurements of the mechanical stiffness and dissipation of a sample surface. When infrared light is absorbed by the sample the resulting rapid thermal expansion excites a "contact resonance" of the AFM cantilever, i.e. a coupled resonance resulting from the properties of both the cantilever and the stiffness and damping of the sample surface. Specifically, the resonance frequency shifts to higher frequencies for stiffer materials and to lower frequencies for softer material. Additionally, the resonance becomes broader for materials with larger dissipation. These contact resonances have been studied extensively by the AFM community (see, for example, atomic force acoustic microscopy ). Traditional contact resonance AFM requires an external actuator to excite the cantilever contact resonances. In AFM-IR these contact resonances are automatically excited every time an infrared pulse is absorbed by the sample. So the AFM-IR technique can measure the infrared absorption by the amplitude of the cantilever oscillation response and the mechanical properties of the sample via the contact resonance frequency and quality factor. [53]

Applications

Applications of AFM-IR have include the characterisation of protein, [16] [17] [18] [19] [20] [21] [22] [54] polymers composites, [15] [36] [38] [39] [55] [56] bacteria, [37] [57] [58] [59] cells, [60] [61] [62] [63] [64] biominerals, [65] [66] pharmaceutical sciences, [17] [35] [67] [68] photonics/nanoantennas, [69] [70] [71] [72] fuel cells, [73] fibers, [39] [74] skin, [75] hair, [76] metal organic frameworks, [77] microdroplets, [51] self-assembled monolayers, [25] nanocrystals, [78] and semiconductors. [79]

Polymers

Polymers blends, composites, multilayer films and fibers AFM-IR has been used to identify and map polymer components in blends, [39] characterize interfaces in composites, [80] and even reverse engineer multilayer films [15] Additionally AFM-IR has been used to study chemical composition in Poly(3][4-ethylenedioxythiophene) (PEDOT) conducting polymers. [56] and vapor infiltration into polyethylene terephthalate PET fibers. [74]

Protein science

The chemical and structural properties of protein determine their interactions, and thus their functions, in a wide variety of biochemical processes. Since Ruggeri et al. pioneering work [16] on the aggregation pathways of the Josephin domain of ataxin-3, responsible for type-3 spinocerebellar ataxia, an inheritable protein-misfolding disease, AFM-IR was used to characterize molecular conformations in a wide spectrum of applications in protein and life sciences. [81] This approach has delivered new mechanistic insights into the behaviour of disease-related proteins and peptides, such as Aβ42, [17] huntingtin [21] and FUS, [53] which are involved in the onset of Alzheimer's, Huntington's and Amyotrophic lateral sclerosis (ALS). Similarly AFM-IR has been applied to study studying protein based functional biomaterials. [54]

Life sciences

AFM-IR has been used to characterise spectroscopically in detail chromosomes, [82] bacteria [59] and cells [60] with nanoscale resolution. For example, in the case of infection of bacteria by viruses [59] (Bacteriophages), and also the production of polyhydroxybutyrate (PHB) vesicles inside Rhodobacter capsulatus cells [58] and triglycerides [46] in Streptomyces bacteria (for biofuel applications). AFM-IR has also been used to evaluate and map mineral content, crystallinity, collagen maturity and acid phosphate content via ratiometric analysis of various absorption bands in bone. [66] AFM-IR has also been used to perform spectroscopy and chemical mapping of structural lipids in human skin, [75] cells [60] and hair [76]

Fuel cells

AFM-IR has been used to study hydrated Nafion membranes used as separators in fuel cells. The measurements revealed the distribution of free and ionically bound water on the Nafion surface. [73]

Photonic nanoantennas

AFM-IR has been used to study the surface plasmon resonance in heavily silicon-doped indium arsenide microparticles. [79] Gold split ring resonators have been studied for use with Surface-Enhanced Infrared Absorption Spectroscopy. In this case AFM-IR was used to measure the local field enhancement of the plasmonics structures (~30X) at 100 nm spatial resolution. [69] [80]

Pharmaceutical sciences

AFM-IR has been used to study miscibility and phase separation in drug polymer blends, [67] [68] the chemical analysis of nanocrystalline drug particles as small 90 nm across, [35] the interaction of chromosomes with chemotherapeutics drugs, [82] and of amyloids with pharmacological approaches to contrast neurodegeneration. [17]

Notes

  1. Graham Poulter, Research Director for Specac Instruments, "The energy available in an optical instrument is directly related the product of the area A of any point in the optical system, multiplied by the solid angle Ω filled by the beam at that point. This product, AΩ, is known as the étendue (also referred to as the "throughput" or "luminosity") and remains a constant at all points in the system. When focusing a beam down from say a 5mm diameter spot in a typical FTIR to a 0.5mm diameter spot, the area A is decreased by a factor of 100 and, therefore, the solid angle Ω has to be increased by the same factor. When illuminating something on a flat surface from one side there is a physical limitation that means Ω cannot exceed π steradians (it's illuminated from a complete hemisphere). Depending on the solid angle in the original instrument beam, this immediately puts a working limit on the minimum spot size that can be usefully obtained when focusing the beam down". Poulter designed the optics in the interface described by Reading et al. [4]
  2. Centre Laser Infrarouge d'Orsay, Orsay Infrared Laser Centre
  3. An arrangement is similar to attenuated total reflectance (ATR) schemes used in conventional infrared spectroscopy
  4. One femtogram is 10−15 grammes
  5. A contact resonance is a vibrational resonance frequency of an AFM cantilever that occurs when the tip of the AFM is in contact with a sample surface. When the QCL is pulsed synchronously with a contact resonance, the detection of the thermal expansion of the sample from infrared absorption is amplified by the quality factor Q of the contact resonance.

Related Research Articles

<span class="mw-page-title-main">Infrared spectroscopy</span> Measurement of infrared radiations interaction with matter

Infrared spectroscopy is the measurement of the interaction of infrared radiation with matter by absorption, emission, or reflection. It is used to study and identify chemical substances or functional groups in solid, liquid, or gaseous forms. It can be used to characterize new materials or identify and verify known and unknown samples. The method or technique of infrared spectroscopy is conducted with an instrument called an infrared spectrometer which produces an infrared spectrum. An IR spectrum can be visualized in a graph of infrared light absorbance on the vertical axis vs. frequency, wavenumber or wavelength on the horizontal axis. Typical units of wavenumber used in IR spectra are reciprocal centimeters, with the symbol cm−1. Units of IR wavelength are commonly given in micrometers, symbol μm, which are related to the wavenumber in a reciprocal way. A common laboratory instrument that uses this technique is a Fourier transform infrared (FTIR) spectrometer. Two-dimensional IR is also possible as discussed below.

<span class="mw-page-title-main">Microscopy</span> Viewing of objects which are too small to be seen with the naked eye

Microscopy is the technical field of using microscopes to view objects and areas of objects that cannot be seen with the naked eye. There are three well-known branches of microscopy: optical, electron, and scanning probe microscopy, along with the emerging field of X-ray microscopy.

<span class="mw-page-title-main">Spectroscopy</span> Study involving matter and electromagnetic radiation

Spectroscopy is the field of study that measures and interprets electromagnetic spectra. In narrower contexts, spectroscopy is the precise study of color as generalized from visible light to all bands of the electromagnetic spectrum.

<span class="mw-page-title-main">Raman spectroscopy</span> Spectroscopic technique

Raman spectroscopy is a spectroscopic technique typically used to determine vibrational modes of molecules, although rotational and other low-frequency modes of systems may also be observed. Raman spectroscopy is commonly used in chemistry to provide a structural fingerprint by which molecules can be identified.

<span class="mw-page-title-main">Atomic force microscopy</span> Type of microscopy

Atomic force microscopy (AFM) or scanning force microscopy (SFM) is a very-high-resolution type of scanning probe microscopy (SPM), with demonstrated resolution on the order of fractions of a nanometer, more than 1000 times better than the optical diffraction limit.

In physics and physical chemistry, time-resolved spectroscopy is the study of dynamic processes in materials or chemical compounds by means of spectroscopic techniques. Most often, processes are studied after the illumination of a material occurs, but in principle, the technique can be applied to any process that leads to a change in properties of a material. With the help of pulsed lasers, it is possible to study processes that occur on time scales as short as 10−16 seconds. All time-resolved spectra are suitable to be analyzed using the two-dimensional correlation method for a correlation map between the peaks.

Scanning probe microscopy (SPM) is a branch of microscopy that forms images of surfaces using a physical probe that scans the specimen. SPM was founded in 1981, with the invention of the scanning tunneling microscope, an instrument for imaging surfaces at the atomic level. The first successful scanning tunneling microscope experiment was done by Gerd Binnig and Heinrich Rohrer. The key to their success was using a feedback loop to regulate gap distance between the sample and the probe.

<span class="mw-page-title-main">Near-infrared spectroscopy</span> Analytical method

Near-infrared spectroscopy (NIRS) is a spectroscopic method that uses the near-infrared region of the electromagnetic spectrum. Typical applications include medical and physiological diagnostics and research including blood sugar, pulse oximetry, functional neuroimaging, sports medicine, elite sports training, ergonomics, rehabilitation, neonatal research, brain computer interface, urology, and neurology. There are also applications in other areas as well such as pharmaceutical, food and agrochemical quality control, atmospheric chemistry, combustion research and knowledge.

Infrared multiple photon dissociation (IRMPD) is a technique used in mass spectrometry to fragment molecules in the gas phase usually for structural analysis of the original (parent) molecule.

<span class="mw-page-title-main">Near-field scanning optical microscope</span>

Near-field scanning optical microscopy (NSOM) or scanning near-field optical microscopy (SNOM) is a microscopy technique for nanostructure investigation that breaks the far field resolution limit by exploiting the properties of evanescent waves. In SNOM, the excitation laser light is focused through an aperture with a diameter smaller than the excitation wavelength, resulting in an evanescent field on the far side of the aperture. When the sample is scanned at a small distance below the aperture, the optical resolution of transmitted or reflected light is limited only by the diameter of the aperture. In particular, lateral resolution of 6 nm and vertical resolution of 2–5 nm have been demonstrated.

<span class="mw-page-title-main">Characterization (materials science)</span> Study of material structure and properties

Characterization, when used in materials science, refers to the broad and general process by which a material's structure and properties are probed and measured. It is a fundamental process in the field of materials science, without which no scientific understanding of engineering materials could be ascertained. The scope of the term often differs; some definitions limit the term's use to techniques which study the microscopic structure and properties of materials, while others use the term to refer to any materials analysis process including macroscopic techniques such as mechanical testing, thermal analysis and density calculation. The scale of the structures observed in materials characterization ranges from angstroms, such as in the imaging of individual atoms and chemical bonds, up to centimeters, such as in the imaging of coarse grain structures in metals.

Chemical imaging is the analytical capability to create a visual image of components distribution from simultaneous measurement of spectra and spatial, time information. Hyperspectral imaging measures contiguous spectral bands, as opposed to multispectral imaging which measures spaced spectral bands.

Photothermal microspectroscopy (PTMS), alternatively known as photothermal temperature fluctuation (PTTF), is derived from two parent instrumental techniques: infrared spectroscopy and atomic force microscopy (AFM). In one particular type of AFM, known as scanning thermal microscopy (SThM), the imaging probe is a sub-miniature temperature sensor, which may be a thermocouple or a resistance thermometer. This same type of detector is employed in a PTMS instrument, enabling it to provide AFM/SThM images: However, the chief additional use of PTMS is to yield infrared spectra from sample regions below a micrometer, as outlined below.

Photothermal spectroscopy is a group of high sensitivity spectroscopy techniques used to measure optical absorption and thermal characteristics of a sample. The basis of photothermal spectroscopy is the change in thermal state of the sample resulting from the absorption of radiation. Light absorbed and not lost by emission results in heating. The heat raises temperature thereby influencing the thermodynamic properties of the sample or of a suitable material adjacent to it. Measurement of the temperature, pressure, or density changes that occur due to optical absorption are ultimately the basis for the photothermal spectroscopic measurements.

<span class="mw-page-title-main">Fourier-transform infrared spectroscopy</span> Technique to analyze the infrared spectrum of matter

Fourier-transform infrared spectroscopy (FTIR) is a technique used to obtain an infrared spectrum of absorption or emission of a solid, liquid, or gas. An FTIR spectrometer simultaneously collects high-resolution spectral data over a wide spectral range. This confers a significant advantage over a dispersive spectrometer, which measures intensity over a narrow range of wavelengths at a time.

<span class="mw-page-title-main">Scanning thermal microscopy</span>

Scanning thermal microscopy (SThM) is a type of scanning probe microscopy that maps the local temperature and thermal conductivity of an interface. The probe in a scanning thermal microscope is sensitive to local temperatures – providing a nano-scale thermometer. Thermal measurements at the nanometer scale are of both scientific and industrial interest. The technique was invented by Clayton C. Williams and H. Kumar Wickramasinghe in 1986.

<span class="mw-page-title-main">Raman microscope</span> Laser microscope used for Raman spectroscopy

The Raman microscope is a laser-based microscopic device used to perform Raman spectroscopy. The term MOLE is used to refer to the Raman-based microprobe. The technique used is named after C. V. Raman, who discovered the scattering properties in liquids.

The technique of vibrational analysis with scanning probe microscopy allows probing vibrational properties of materials at the submicrometer scale, and even of individual molecules. This is accomplished by integrating scanning probe microscopy (SPM) and vibrational spectroscopy. This combination allows for much higher spatial resolution than can be achieved with conventional Raman/FTIR instrumentation. The technique is also nondestructive, requires non-extensive sample preparation, and provides more contrast such as intensity contrast, polarization contrast and wavelength contrast, as well as providing specific chemical information and topography images simultaneously.

The photoacoustic effect or optoacoustic effect is the formation of sound waves following light absorption in a material sample. In order to obtain this effect the light intensity must vary, either periodically or as a single flash. The photoacoustic effect is quantified by measuring the formed sound with appropriate detectors, such as microphones or piezoelectric sensors. The time variation of the electric output from these detectors is the photoacoustic signal. These measurements are useful to determine certain properties of the studied sample. For example, in photoacoustic spectroscopy, the photoacoustic signal is used to obtain the actual absorption of light in either opaque or transparent objects. It is useful for substances in extremely low concentrations, because very strong pulses of light from a laser can be used to increase sensitivity and very narrow wavelengths can be used for specificity. Furthermore, photoacoustic measurements serve as a valuable research tool in the study of the heat evolved in photochemical reactions, particularly in the study of photosynthesis.

<span class="mw-page-title-main">Nano-FTIR</span> Infrared microscopy technique

Nano-FTIR is a scanning probe technique that utilizes as a combination of two techniques: Fourier transform infrared spectroscopy (FTIR) and scattering-type scanning near-field optical microscopy (s-SNOM). As s-SNOM, nano-FTIR is based on atomic-force microscopy (AFM), where a sharp tip is illuminated by an external light source and the tip-scattered light is detected as a function of tip position. A typical nano-FTIR setup thus consists of an atomic force microscope, a broadband infrared light source used for tip illumination, and a Michelson interferometer acting as Fourier-transform spectrometer. In nano-FTIR, the sample stage is placed in one of the interferometer arms, which allows for recording both amplitude and phase of the detected light. Scanning the tip allows for performing hyperspectral imaging with nanoscale spatial resolution determined by the tip apex size. The use of broadband infrared sources enables the acquisition of continuous spectra, which is a distinctive feature of nano-FTIR compared to s-SNOM. Nano-FTIR is capable of performing infrared (IR) spectroscopy of materials in ultrasmall quantities and with nanoscale spatial resolution. The detection of a single molecular complex and the sensitivity to a single monolayer has been shown. Recording infrared spectra as a function of position can be used for nanoscale mapping of the sample chemical composition, performing a local ultrafast IR spectroscopy and analyzing the nanoscale intermolecular coupling, among others. A spatial resolution of 10 nm to 20 nm is routinely achieved.

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 Hammiche, A.; Pollock, H. M.; Reading, M.; Claybourn, M.; Turner, P. H.; Jewkes, K. (1999). "Photothermal FT-IR Spectroscopy: A Step Towards FT-IR Microscopy at a Resolution Better Than the Diffraction Limit". Applied Spectroscopy. 53 (7): 810–815. Bibcode:1999ApSpe..53..810H. doi:10.1366/0003702991947379. S2CID   93359289.
  2. 1 2 3 4 5 6 7 8 9 10 Anderson, M. S. (2000). "Infrared Spectroscopy with an Atomic Force Microscope". Applied Spectroscopy. 54 (3): 349–352. Bibcode:2000ApSpe..54..349.. doi:10.1366/0003702001948538. S2CID   103205691.
  3. 1 2 3 4 5 6 7 8 Hammiche, A.; Bozec, L.; Conroy, M.; Pollock, H. M.; Mills, G.; Weaver, J. M. R.; Price, D. M.; Reading, M.; Hourston, D.J.; Song, M. (2000). "Highly localized thermal, mechanical and spectroscopic characterisation of polymers using miniaturized thermal probes". Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures. 18 (3): 1322–1332. Bibcode:2000JVSTB..18.1322H. doi:10.1116/1.591381. S2CID   55856483.
  4. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Reading, M.; Price, D. M.; Grandy, D.B.; Smith, R. M.; Bozec, L.; Conroy, M.; Hammiche, A.; Pollock, H. M. (2001). "Microthermal analysis of polymers: current capabilities and future prospects". Macromol. Symp. 167: 45–62. doi:10.1002/1521-3900(200103)167:1<45::aid-masy45>3.0.co;2-n.
  5. 1 2 3 4 5 6 7 Reading, M.; Grandy, D.; Hammiche, A.; Bozec, L.; Pollock, H. M. (2002). "Thermally assisted nanosampling and analysis using micro-IR spectroscopy and other analytical technciques". Vibrational Spectroscopy. 29 (1): 257–260. doi:10.1016/s0924-2031(01)00185-0.
  6. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Hammiche, A.; Bozec, L.; Pollock, H. M.; German, M.; Reading, M. (2004). "Progress in near-field photothermal infrared microspectroscopy". Journal of Microscopy. 213 (2): 129–134. doi:10.1111/j.1365-2818.2004.01292.x. PMID   14731294. S2CID   38880191.
  7. 1 2 3 4 5 6 7 8 Reading, M.; Grandy, D.; Pollock, H. M.; Hammiche, A. (2004). Micro-thermal analysis using a new high resolution thermal probe. United Kingdom SPM Meeting. Nottingham.
  8. 1 2 3 4 5 6 7 8 9 Dazzi, A.; Prazeres, R.; Glotin, F.; Ortega, J. M. (2005). "Local infrared microspectroscopy with subwavelength spatial resolution with an atomic force microscope tip used as a photothermal sensor". Optics Letters. 30 (18): 2388–2390. Bibcode:2005OptL...30.2388D. doi:10.1364/OL.30.002388. PMID   16196328.
  9. 1 2 3 4 5 Dazzi, A.; Glotin, F.; Ortega, J. M. (September 2006). "Subwavelength infrared spectromicroscopy using an AFM as a local absorption sensor". Infrared Physics and Technology. 49 (1–2): 113–121. Bibcode:2006InPhT..49..113D. doi:10.1016/j.infrared.2006.01.009.
  10. 1 2 3 4 5 6 Dazzi, A.; Prazeres, R.; Glotin, F.; Ortega, J. M. (2007). "Analysis of nano-chemical mapping performed by an AFM-based ("AFMIR") acousto-optic technique". Ultramicroscopy. 107 (12): 1194–1200. doi: 10.1016/j.ultramic.2007.01.018 . PMID   17382474.
  11. 1 2 3 4 Dazzi, A. (2008). "Sub-100-Nanometer Infrared Spectroscopy and Imaging Based on a Near-Field Photothermal Technique (PTIR)". In Kneipp, J.; Lasch, P. (eds.). Biomedical Vibrational Spectroscopy. pp. 291–312. doi:10.1002/9780470283172.ch13. ISBN   9780470283172.
  12. 1 2 3 4 5 6 7 8 9 10 Hill, G.A.; Rice, J.H.; Meech, S.R.; Craig, D.; Kuo, P.; Vodopyanov, K.; Reading, M. (2009). "Submicrometer infrared surface imaging using a scanning probe microscope and an optical parametric oscillator laser". Optics Letters. 34 (4): 433. Bibcode:2009OptL...34..431H. doi:10.1364/OL.34.000431. PMID   19373331. (published online, Feb 2008)
  13. 1 2 Vodopyanov, K.; Hill, G. A.; Rice, J. H.; Meech, S. R.; Craig, D. Q. M.; Reading, M. M.; Dazzi, A.; Kjoller, K.; Prater, C. (Fall 2009). Nano-Spectroscopy in the 2.5-10 Micron Wavelength Range Using Atomic Force Microscope. Frontiers in Optics Laser Science XXV.
  14. 1 2 3 4 Hammiche, A.; Bozec, L.; German, M.J.; Chalmers, J. M.; Everall, N. J.; Poulter, G.; Reading, M.; Grandy, D. B.; Martin, F.L.; Pollock, H.M. (2004). "Mid-infrared micro-spectroscopy of difficult samples using near-field photothermal micro-spectroscopy (PTMS)". Spectroscopy. 19 (2): 20–42. with erratum, 19(5), 14 May 2004
  15. 1 2 3 4 5 Eby, T.; Gundusharma, U.; Lo, M.; Sahagian, K.; Marcott, C.; Kjoller, K. (June 13, 2012). "Reverse engineering of polymeric multilayers using AFM-based nanoscale IR spectroscopy and thermal analysis". Spectroscopy Europe. 24 (3): 18–21.
  16. 1 2 3 4 5 Ruggeri, F. S.; Longo, G.; Faggiano, S.; Lipiec, E.; Pastore, A.; Dietler, G. (2015-07-28). "Infrared nanospectroscopy characterization of oligomeric and fibrillar aggregates during amyloid formation". Nature Communications. 6 (1): 7831. Bibcode:2015NatCo...6.7831R. doi:10.1038/ncomms8831. ISSN   2041-1723. PMC   4525161 . PMID   26215704.
  17. 1 2 3 4 5 6 7 8 9 Ruggeri, Francesco Simone; Habchi, Johnny; Chia, Sean; Horne, Robert I.; Vendruscolo, Michele; Knowles, Tuomas P. J. (2021-01-29). "Infrared nanospectroscopy reveals the molecular interaction fingerprint of an aggregation inhibitor with single Aβ42 oligomers". Nature Communications. 12 (1): 688. Bibcode:2021NatCo..12..688R. doi:10.1038/s41467-020-20782-0. ISSN   2041-1723. PMC   7846799 . PMID   33514697.
  18. 1 2 3 4 5 6 7 Ruggeri, Francesco Simone; Mannini, Benedetta; Schmid, Roman; Vendruscolo, Michele; Knowles, Tuomas P. J. (2020-06-10). "Single molecule secondary structure determination of proteins through infrared absorption nanospectroscopy". Nature Communications. 11 (1): 2945. Bibcode:2020NatCo..11.2945R. doi:10.1038/s41467-020-16728-1. ISSN   2041-1723. PMC   7287102 . PMID   32522983.
  19. 1 2 Müller, Thomas; Simone Ruggeri, Francesco; Kulik, Andrzej; Shimanovich, Ulyana; Mason, Thomas; Knowles, Tuomas; Dietler, Giovanni (2014). "Nanoscale spatially resolved infrared spectra from single microdroplets". Lab on a Chip. 14 (7): 1315–1319. arXiv: 1401.8204 . doi:10.1039/C3LC51219C. PMID   24519414. S2CID   16702240.
  20. 1 2 Ruggeri, Francesco Simone; Habchi, Johnny; Cerreta, Andrea; Dietler, Giovanni (2016). "AFM-Based Single Molecule Techniques: Unraveling the Amyloid Pathogenic Species". Current Pharmaceutical Design. 22 (26): 3950–3970. doi:10.2174/1381612822666160518141911. ISSN   1381-6128. PMC   5080865 . PMID   27189600.
  21. 1 2 3 4 Ruggeri, F. S.; Vieweg, S.; Cendrowska, U.; Longo, G.; Chiki, A.; Lashuel, H. A.; Dietler, G. (2016-08-08). "Nanoscale studies link amyloid maturity with polyglutamine diseases onset". Scientific Reports. 6 (1): 31155. Bibcode:2016NatSR...631155R. doi:10.1038/srep31155. ISSN   2045-2322. PMC   4976327 . PMID   27499269.
  22. 1 2 3 Adamcik, Jozef; Ruggeri, Francesco Simone; Berryman, Joshua T.; Zhang, Afang; Knowles, Tuomas P. J.; Mezzenga, Raffaele (2021). "Evolution of Conformation, Nanomechanics, and Infrared Nanospectroscopy of Single Amyloid Fibrils Converting into Microcrystals". Advanced Science. 8 (2): 2002182. doi:10.1002/advs.202002182. ISSN   2198-3844. PMC   7816722 . PMID   33511004.
  23. Ramer, Georg; Ruggeri, Francesco Simone; Levin, Aviad; Knowles, Tuomas P. J.; Centrone, Andrea (2018-07-24). "Determination of Polypeptide Conformation with Nanoscale Resolution in Water". ACS Nano. 12 (7): 6612–6619. doi:10.1021/acsnano.8b01425. ISSN   1936-0851. PMID   29932670. S2CID   49380687.
  24. 1 2 3 4 5 6 7 8 9 Bozec, L.; Hammiche, A.; Pollock, H.M.; Conroy, M.; Everall, N. J.; Turi, L. (2001). "Localized phtothermal infrared spectroscopy using a proximal probe". Journal of Applied Physics. 90 (10): 5159. Bibcode:2001JAP....90.5159B. doi:10.1063/1.1403671.
  25. 1 2 3 4 5 6 Lu, F.; Jin, M.; Belkin, M.A. (2014). "Tip-enhanced infrared nanospectroscopy via molecular expansion force detection". Nature Photonics. 8 (4): 307–312. Bibcode:2014NaPho...8..307L. doi:10.1038/nphoton.2013.373. S2CID   14353001.
  26. H M Pollock & D A Smith (2002). "The use of near-field probes for vibrational spectroscopy and photothermal imaging". In J M Chalmers & P R Griffiths (eds.). Handbook of vibrational spectroscopy vol. 2. pp. 1472–92.
  27. F L Martin & H M Pollock (2010). "Microspectroscopy as a tool to discriminate nano-molecular cellular alterations in biomedical research". In J A V Narlikar & Y Y Fu (eds.). Oxford Handbook of Nanoscience and Technology vol. 2. pp. 285–336.
  28. 1 2 3 4 5 6 7 8 9 10 Dai, X.; Moffat, J. G.; Wood, J.; Reading, M. (April 2012). "Thermal scanning probe microscopy in the development of pharmaceuticals". Advanced Drug Delivery Reviews. 64 (5): 449–460. doi:10.1016/j.addr.2011.07.008. PMID   21856345.
  29. 1 2 3 Gorbunov, V.V.; Grandy, D.; Reading, M.; Tsukruk, V.V. (2009). "7, Micro and Nanoscale Local Thermal Analysis". Thermal Analysis of Polymers, Fundamantals and Applications. John Wiley and Sons.
  30. J Ye; M Reading; N Gotzen & G van Assche (2007). "Scanning Thermal Probe Microscopy:NanoThermal Analysis with Raman Microscopy". Microscopy and Analysis. 21 (2): S5–S8.
  31. 1 2 "Impact case study (REF3b)". Research Excellence Framework.
  32. H M Pollock (2011). "Towards chemical mapping at sub-micron resolution: near-field spectroscopic delineation of interphase boundaries" (PDF). Materials Science Forum. 662: 1–11. doi:10.4028/www.scientific.net/msf.662.1. S2CID   43540112.
  33. 1 2 Lahiri, B.; Holland, G.; Centrone, A. (October 4, 2012). "Chemical Imaging Beyond the Diffraction Limit: Experimental Validation of the PTIR Technique". Small. 9 (3): 439–445. doi:10.1002/smll.201200788. PMID   23034929.
  34. Dazzi, A.; Glotin, F.; Carminati, R. (2010). "Theory of infrared nano-spectroscopy by Photo Thermal Induced Resonance". Journal of Applied Physics. 107 (12): 124519–124519–7. Bibcode:2010JAP...107l4519D. doi:10.1063/1.3429214.
  35. 1 2 3 4 Katzenmeyer, Aksyuk V.; Centrone, A. (2013). "Nanoscale Infrared Spectroscopy: Improving the Spectral Range of the Photothermal Induced Resonance Technique". Analytical Chemistry. 85 (4): 1972–1979. doi:10.1021/ac303620y. PMID   23363013.
  36. 1 2 3 Felts, J. R.; Kjoller, K.; Lo, M.; Prater, C. B.; King, W. P. (August 31, 2012). "Nanometer-scale infrared spectroscopy of heterogeneous polymer nanostructures fabricated by tip-based nanofabrication". ACS Nano. 6 (9): 8015–8021. doi:10.1021/nn302620f. PMID   22928657.
  37. 1 2 Mayet, A.; Deiset-Besseau, A.; Prazeres, R.; Ortega, J. M.; Dazzi, A. (2013). "Analysis of bacterial polyhydroxybutyrate production by multimodal nanoimaging". Biotechnology Advances. 31 (3): 369–374. doi:10.1016/j.biotechadv.2012.05.003. PMID   22634017.
  38. 1 2 3 Kjoller, K.; Prater, C.; Shetty, R. (November 1, 2010). "Polymer characterization using nanoscale infrared spectroscopy". American Laboratory. 42 (11).
  39. 1 2 3 4 Dazzi; Prater, C. B.; Hu, Q.; Chase, D. B.; Rabolt, J. F.; Marcott, C. (2012). "AFM-IR: combining atomic force microscopy and infrared spectroscopy for nanoscale chemical characterization". Applied Spectroscopy. 66 (12): 1365–1384. Bibcode:2012ApSpe..66.1365D. doi: 10.1366/12-06804 . PMID   23231899.
  40. 1 2 3 Reading, M.; Hammiche, A.; Pollock, H.M.; Rankl, C.; Rice, J.; Capponi, S.; Grandy, D. Two new scanning probe microscopy techniques for photothermal IR imaging and spectroscopy. Royal Society of Chemistry TAC. Cambridge, UK.30 March-1 April 2015
  41. "Anasys Instruments Limited". Company Check.
  42. Final report EPSRC grant EP/C007751/1 (PDF) (Report).
  43. 1 2 "Towards chemical mapping at sub-micron resolution : near-field spectroscopic delineation of interphase boundaries. Pollock, H.M.". Materials Science Forum (662): 1–11. November 2010.
  44. Dai, X.; Moffat, J. G.; Mayes, A.G.; Reading, M.; Craig, D.Q. M.; Belton, P.S.; Grandy, D. B. (2009). "Thermal Probe Based Analytical Microscopy: Thermal Analysis and Photothermal Fourier-Transform Infrared Microspectroscopy Together with Thermally Assisted Nanosampling Coupled with Capillary Electrophoresis". Analytical Chemistry. 81 (16): 6612–9. doi:10.1021/ac9004869. PMID   20337375.
  45. Harding, L.; Qi, S.; Hill, G.; Reading, M.; Craig, D. Q. M. (May 2008). "The development of microthermal analysis and photothermal microspectroscopy as novel approaches to drug-excipient compatibility studies". International Journal of Pharmaceutics. 354 (1–2): 149–157. doi:10.1016/j.ijpharm.2007.11.009. PMID   18162342.; 354(1-2)149-5.
  46. 1 2 Moffat, J. G.; Mayes, A. G.; Belton, P. S.; Craig, D. Q. M.; Reading, M. (2009). "Compositional Analysis of Metal Chelating Materials Using Near-Field Photothermal Fourier Transform Infrared Microspectroscopy". Analytical Chemistry. 82 (1): 91–7. doi:10.1021/ac800906t. PMID   19957959.
  47. Dai, X.; Belton, P.; DeCogan, D.; Moffat, J. G.; Reading, M. (2011). "Thermally induced movement of micro particles observed on a rough surface: A novel observation and its implications for high throughput analysis and synthesis". Thermochimica Acta. 517 (4): 121–125. doi:10.1016/j.tca.2011.01.037.
  48. "VertiSense Thermal Probes (VTR)". Appnano.com.
  49. Kjoller, K.; Felts, J. R.; Cook, D.; Prater, C. B.; King, W. P. (2010). "High-sensitivity nanometer-scale infrared spectroscopy using a contact mode microcantilever with an internal resonator paddle". Nanotechnology. 21 (18): 185705. Bibcode:2010Nanot..21r5705K. doi:10.1088/0957-4484/21/18/185705. PMID   20388971. S2CID   27042137. 185705
  50. Cho, H.; Felts, J. R.; Yu, M. F.; Bergman, L. A.; Vakakis, A. F.; King, W. P. (2013). "Improved Atomic Force microscope Infrared Spectroscopy for Rapid Nanometer-Scale Chemical Identification". Nanotechnology. 24 (44): 444007. Bibcode:2013Nanot..24R4007C. doi:10.1088/0957-4484/24/44/444007. PMID   24113150. S2CID   3857086.444007
  51. 1 2 Muller, T.; Ruggeri, F. S.; Kulik, A. J.; Shimanovich, U.; Mason, T. O.; Knowles, T. P. J.; Dietler, G. (2014). "Nanoscale spatially resolved infrared spectra from single microdroplets". Lab on a Chip. 14 (7): 1315–1319. arXiv: 1401.8204 . doi:10.1039/C3LC51219C. PMID   24519414. S2CID   16702240.
  52. Pollock, Hubert M.; Kazarian, Sergei G. (2006). "Microspectroscopy in the Mid-Infrared". Encyclopedia of Analytical Chemistry. pp. 1–26. doi:10.1002/9780470027318.a5609.pub2. ISBN   9780470027318.
  53. 1 2 Qamar, Seema; Wang, Guozhen; Randle, Suzanne J.; Ruggeri, Francesco Simone; Varela, Juan A.; Lin, Julie Qiaojin; Phillips, Emma C.; Miyashita, Akinori; Williams, Declan; Ströhl, Florian; Meadows, William; Ferry, Rodylyn; Dardov, Victoria J.; Tartaglia, Gian G.; Farrer, Lindsay A.; Kaminski Schierle, Gabriele S.; Kaminski, Clemens F.; Holt, Christine E.; Fraser, Paul E.; Schmitt-Ulms, Gerold; Klenerman, David; Knowles, Tuomas; Vendruscolo, Michele; St George-Hyslop, Peter (2018-04-19). "FUS Phase Separation Is Modulated by a Molecular Chaperone and Methylation of Arginine Cation-π Interactions". Cell. 173 (3): 720–734.e15. doi:10.1016/j.cell.2018.03.056. ISSN   0092-8674. PMC   5927716 . PMID   29677515.
  54. 1 2 Shen, Yi; Ruggeri, Francesco Simone; Vigolo, Daniele; Kamada, Ayaka; Qamar, Seema; Levin, Aviad; Iserman, Christiane; Alberti, Simon; George-Hyslop, Peter St; Knowles, Tuomas P. J. (2020). "Biomolecular condensates undergo a generic shear-mediated liquid-to-solid transition". Nature Nanotechnology. 15 (10): 841–847. Bibcode:2020NatNa..15..841S. doi:10.1038/s41565-020-0731-4. ISSN   1748-3395. PMC   7116851 . PMID   32661370.
  55. Marcott, C.; Lo, M.; Kjoller, K.; Prater, C.; Noda, I. (2011). "Spatial Differentiation of Sub-Micrometer Domains in a Poly(hydroxyalkanoate) Copolymer Using Instrumentation that Combines Atomic Force Microscopy (AFM) and Infrared (IR) Spectroscopy". Applied Spectroscopy. 65 (10): 1145–1150. Bibcode:2011ApSpe..65.1145M. doi:10.1366/11-06341. PMID   21986074. S2CID   207353084.
  56. 1 2 Ghosh, S.; Remita, H.; Ramos, L.; Dazzi, A.; Deiset-Besseau, A.; Beaunier, P.; Goubard, F.; Aubert, P. H.; Brisset, F.; Remita, S. (2014). "PEDOT nanostructures synthesized in hexagonal mesophases". New Journal of Chemistry. 38 (3): 1106–1115. doi:10.1039/c3nj01349a. S2CID   98578268.
  57. Deiset-Besseau, A.; Prater, C. B.; Virolle, M.J.; Dazzi, A. (2014). "Monitoring TriAcylGlycerols Accumulation by Atomic Force Microscopy Based Infrared Spectroscopy in Streptomyces Species for Biodiesel Applications". The Journal of Physical Chemistry Letters. 5 (4): 654–658. doi:10.1021/jz402393a. PMID   26270832.
  58. 1 2 Mayet, A.; Dazzi, A.; Prazeres, R.; Ortega, J. M.; Jaillard, D. (2010). "In situ identification and imaging of bacterial polymer nanogranules by infrared nanospectroscopy". Analyst. 135 (10): 2540–2545. Bibcode:2010Ana...135.2540M. doi:10.1039/c0an00290a. PMID   20820491.
  59. 1 2 3 Dazzi, A.; Prazeres, R.; Glotin, F.; Ortega, J. M.; Al-Sawaftah, M.; de Frutos, M. (2008). "Chemical mapping of the distribution of viruses into infected bacteria with a photothermal method". Ultramicroscopy. 108 (7): 635–641. doi:10.1016/j.ultramic.2007.10.008. PMID   18037564.
  60. 1 2 3 Ruggeri, Francesco S.; Marcott, Curtis; Dinarelli, Simone; Longo, Giovanni; Girasole, Marco; Dietler, Giovanni; Knowles, Tuomas P. J. (2018). "Identification of Oxidative Stress in Red Blood Cells with Nanoscale Chemical Resolution by Infrared Nanospectroscopy". International Journal of Molecular Sciences. 19 (9): 2582. doi: 10.3390/ijms19092582 . PMC   6163177 . PMID   30200270. S2CID   52185910.
  61. Clede, S.; Lambert, F.; Sandt, C.; Kascakova, S.; Unger, M.; Harte, E.M.; Plamont, A.; Saint-Fort, R.; Deiset-Besseau, A.; Gueroui, Z.; Hirschmugl, C.; Lecomte, S.; Dazzi, A.; Vessieres, A.; Policar, C. (2013). "Detection of an estrogen derivative in two breast cancer cell lines using a single core multimodal probe for imaging (SCoMPI) imaged by a panel of luminescent and vibrational techniques" (PDF). Analyst. 138 (19): 5627–5638. Bibcode:2013Ana...138.5627C. doi: 10.1039/c3an00807j . PMID   23897394.
  62. Policar, C.; Waern, J. B.; Plamont, M. A.; Clède, S.; Mayet, C.; Prazeres, R.; Ortega, J. M.; Vessières, A.; Dazzi, A. (2011). "Subcellular IR Imaging of a Metal–Carbonyl Moiety Using Photothermally Induced Resonance". Angewandte Chemie International Edition. 123 (4): 890–894. Bibcode:2011AngCh.123..890P. doi:10.1002/ange.201003161.
  63. Dazzi, A.; Policar, C. (2011). Chabal, C. M. P. J. (ed.). "Biointerface Characterization by Advanced IR Spectroscopy". Elsevier, Amsterdam: 245–278.{{cite journal}}: Cite journal requires |journal= (help)
  64. Mayet, C.; Dazzi, A.; Prazeres, R.; Allot, F.; Glotin, F.; Ortega, J. M. (2008). "Sub-100nm IR spectromicroscopy of living cells". Optics Letters. 33 (14): 1611–1613. Bibcode:2008OptL...33.1611M. doi:10.1364/OL.33.001611. PMID   18628814.
  65. Marcott, C.; Lo, M.; Hu, Q.; Kjoller, K.; Boskey, A.; Noda, I. (2014). "Using 2D correlation analysis to enhance spectral information available from highly spatially resolved AFM-IR spectra". Journal of Molecular Structure. 1069: 284–289. Bibcode:2014JMoSt1069..284M. doi:10.1016/j.molstruc.2014.01.036. PMC   4093835 . PMID   25024505.
  66. 1 2 Gourio-Arsiquaud, S; Marcott, C.; Hu, Q.; Boskey, A. (2014). "Studying variations in bone composition at nano-scale resolution: a preliminary report". Calcified Tissue International. 95 (5): 413–418. doi:10.1007/s00223-014-9909-9. PMC   4192085 . PMID   25155443.
  67. 1 2 Van Eerdenbrugh, B.; Lo, M.; Kjoller, K.; Marcott, C.; Taylor, L. S. (2012). "Nanoscale mid-infrared imaging of phase separation in a drug–polymer blend". Journal of Pharmaceutical Sciences. 101 (6): 2066–2073. doi: 10.1002/jps.23099 . PMID   22388948.
  68. 1 2 Van Eerdenbrugh, B.; Lo, M.; Kjoller, K.; Marcott, C.; Taylor, L. S. (2012). "Nanoscale Mid-Infrared Evaluation of the Miscibility Behavior of Blends of Dextran or Maltodextrin with Poly(vinylpyrrolidone)". Molecular Pharmaceutics. 9 (5): 1459–1469. doi:10.1021/mp300059z. PMID   22483035.
  69. 1 2 Lahiri, B.; Holland, G.; Aksyuk, V.; Centrone, A. (2013). "Nanoscale imaging of plasmonic hot spots and dark modes with the photothermal-induced resonance technique". Nano Letters. 13 (7): 3218–3224. Bibcode:2013NanoL..13.3218L. doi:10.1021/nl401284m. PMID   23777547.
  70. Felts, J. R.; Law, S.; Roberts, C. M.; Podolskiy, V.; Wasserman, D. M.; King, W. P. (2013). "Near-field infrared absorption of plasmonic semiconductor microparticles studied using atomic force microscope infrared spectroscopy". Applied Physics Letters. 102 (15): 152110. Bibcode:2013ApPhL.102o2110F. doi:10.1063/1.4802211.
  71. Katzenmeyer, A. M.; Chae, J.; Kasica, R.; Holland, G.; Lahiri, B.; Centrone, A. (2014). "Nanoscale Imaging and Spectroscopy of Plasmonic Modes with the PTIR Technique". Advanced Optical Materials. 2 (8): 718–722. doi:10.1002/adom.201400005. S2CID   54809198.
  72. Houel, J.; Sauvage, S.; Boucaud, P.; Dazzi, A.; Prazeres, R.; Glotin, F.; Ortega, J. M.; Miard, A.; Lemaitre, A. (2007). "Ultraweak-Absorption Microscopy of a Single Semiconductor Quantum Dot in the Midinfrared Range". Physical Review Letters. 99 (21): 217404. Bibcode:2007PhRvL..99u7404H. doi:10.1103/PhysRevLett.99.217404. PMID   18233255.217404
  73. 1 2 Awatani, T.; Midorikawa, H.; Kojima, N.; Ye, J.; Marcott, C. (2013). "Morphology of water transport channels and hydrophobic clusters in Nafion from high spatial resolution AFM-IR spectroscopy and imaging". Electrochemistry Communications. 30: 5–8. doi:10.1016/j.elecom.2013.01.021.
  74. 1 2 Akyildiz, H. I.; Lo, M.; Dillon, E.; Roberts, A. T.; Everitt, H. O.; Jur, J. S. (2014). "Formation of novel photoluminescent hybrid materials by sequential vapor infiltration into polyethylene terephthalate fibers". Journal of Materials Research. 29 (23): 2817–2826. Bibcode:2014JMatR..29.2817A. doi:10.1557/jmr.2014.333. S2CID   97838045.
  75. 1 2 Marcott, C.; Lo, M.; Kjoller, K.; Domanov, Y.; Balooch, G.; Luengo, G. S. (2013). "Nanoscale infrared (IR) spectroscopy and imaging of structural lipids in human stratum corneum using an atomic force microscope to directly detect absorbed light from a tunable IR laser source". Experimental Dermatology. 22 (6): 419–421. doi: 10.1111/exd.12144 . PMID   23651342. S2CID   11641941.
  76. 1 2 Marcott, C.; Lo, M.; Kjoller, K.; Fiat, F.; Baghdadli, N.; Balooch, G.; Luengo, G. S. (2014). "Localization of Human Hair Structural Lipids Using Nanoscale Infrared Spectroscopy and Imaging". Applied Spectroscopy. 68 (5): 564–569. Bibcode:2014ApSpe..68..564M. doi:10.1366/13-07328. PMID   25014600. S2CID   22916551.
  77. Katzenmeyer, A. M.; Canivet, J.; Holland, G.; Farrusseng, D.; Centrone, A. (2014). "Assessing Chemical Heterogeneity at the Nanoscale in Mixed-Ligand Metal–Organic Frameworks with the PTIR Technique". Angewandte Chemie International Edition. 53 (11): 2852–2856. doi:10.1002/anie.201309295. PMID   24615798.
  78. Rosen, E. L.; Buonsanti, R.; Llordes, A.; Sawvel, A. M.; Milliron, D. J.; Helms, B. A. (2012). "Exceptionally Mild Reactive Stripping of Native Ligands from Nanocrystal Surfaces by Using Meerwein's Salt". Angewandte Chemie International Edition. 51 (3): 684–689. doi:10.1002/anie.201105996. PMID   22147424.
  79. 1 2 Houel, J.; Homeyer, E.; Sauvage, S.; Boucaud, P.; Dazzi, A.; Prazeres, R.; Ortega, J. M. (2009). "Midinfrared absorption measured at a λ/400 resolution with an atomic force microscope". Opt Express. 17 (13): 10887–10894. Bibcode:2009OExpr..1710887H. doi: 10.1364/OE.17.010887 . PMID   19550489.
  80. 1 2 "Nanoscale Infrared Spectroscopy of Polymer Composites", americanlaboratory.com
  81. Kurouski, Dmitry; Dazzi, Alexandre; Zenobi, Renato; Centrone, Andrea (2020-06-08). "Infrared and Raman chemical imaging and spectroscopy at the nanoscale". Chemical Society Reviews. 49 (11): 3315–3347. doi:10.1039/C8CS00916C. ISSN   1460-4744. PMC   7675782 . PMID   32424384.
  82. 1 2 Lipiec, Ewelina; Ruggeri, Francesco S.; Benadiba, Carine; Borkowska, Anna M.; Kobierski, Jan D.; Miszczyk, Justyna; Wood, Bayden R.; Deacon, Glen B.; Kulik, Andrzej; Dietler, Giovanni; Kwiatek, Wojciech M. (2019-10-10). "Infrared nanospectroscopic mapping of a single metaphase chromosome". Nucleic Acids Research. 47 (18): e108. doi:10.1093/nar/gkz630. ISSN   1362-4962. PMC   6765102 . PMID   31562528.