Quantitative active super-resolution thermal imaging: The melanoma case study
-
Mario Marini
, Margaux Bouzin, Riccardo Scodellaro
, Laura D’Alfonso , Laura Sironi , Francesca Granucci , Francesca Mingozzi , Giuseppe Chirico and Maddalena Collini
Abstract
Super-resolution image acquisition has turned photo-activated far-infrared thermal imaging into a promising tool for the characterization of biological tissues. By the sub-diffraction localization of sparse temperature increments primed by the sample absorption of modulated focused laser light, the distribution of (endogenous or exogenous) photo-thermal biomarkers can be reconstructed at tunable ∼10−50 μm resolution. We focus here on the theoretical modeling of laser-primed temperature variations and provide the guidelines to convert super-resolved temperature-based images into quantitative maps of the absolute molar concentration of photo-thermal probes. We start from camera-based temperature detection via Stefan–Boltzmann’s law, and elucidate the interplay of the camera point-spread-function and pixelated sensor size with the excitation beam waist in defining the amplitude of the measured temperature variations. This can be accomplished by the numerical solution of the three-dimensional heat equation in the presence of modulated laser illumination on the sample, which is characterized in terms of thermal diffusivity, conductivity, thickness, and concentration of photo-thermal species. We apply our data-analysis protocol to murine B16 melanoma biopsies, where melanin is mapped and quantified in label-free configuration at sub-diffraction 40 µm resolution. Our results, validated by an unsupervised machine-learning analysis of hematoxylin-and-eosin images of the same sections, suggest potential impact of super-resolved thermography in complementing standard histopathological analyses of melanocytic lesions.
Introduction
Temperature and externally induced temperature variations provide a powerful imaging contrast agent. A number of well-established techniques indeed rely on temperature for the quantitative investigation of the morphology and functional state of biological samples [1,2,3,4,5,6,7]. Temperature is typically probed indirectly, either through the response of optical properties, such as reflectance [8] and index of refraction [4,5,6], to a modulated change in the sample temperature, or by the quantification of the sample thermal radiation in the infrared spectral band [9,10]. The latter approach is at the basis of (far) infrared thermography, and enables a conceptually straightforward retrieval of absolute temperature values: a thermal camera senses the sample surface infrared radiance R, which is subsequently converted into an absolute temperature T by Stefan–Boltzmann’s law (
With the main advantages of quantitative and non-contact temperature measurements over variably extended (>mm2 sized) sample areas, infrared thermography is nowadays a well-established technology and finds promising applications in the life sciences [2,11]. In fact, information on the alterations and structural properties of biological tissues can be recovered by exploiting heat transfer processes [1,2,11,12]. For example, the spatial distribution of photo-thermal biomarkers (e.g., melanin in melanoma skin lesions [5]) could be obtained by photo-activated thermal imaging experiments on biopsy specimens, thereby leading to the development of novel protocols aimed at the screening and early-stage detection of pathologies. Similarly, in the field of precision medicine, where light-responsive (nano-) regulators have gained increased attention [13] to manipulate cell functions, active thermography can map the biodistribution of photothermal probes in both superficial tissues and biopsies of the main organs involved in the accumulation process [14]. For all these applications, (i) imaging at high (
The spatial resolution of commercially available far-infrared microbolometer-based cameras is theoretically limited to
However, in the simplest implementation of super-resolution thermography, while the peak coordinates of laser-primed temperature increments assign the location of photo-thermal entities in the reconstructed image of the sample, the amplitude of temperature variations is only exploited to define the image false-color code [14]. The sample thermal conductivity can be measured by exploiting the temperature rise-and-decay temporal kinetics under modulated laser illumination [15]; by contrast, no quantitative information on the concentration of absorbing and heat releasing entities is retrieved [14].
In a view of fully exploiting the information encoded in the super-resolved temperature-based image of a biological tissue, we complement our super-resolution imaging approach [14,15] with the necessary theoretical framework to model experimentally detected temperature increments as a function of both instrumental parameters and the concentration of laser-excited photo-thermal probes. Provided that temperature heterogeneities in super-resolution imaging experiments typically occur close to (or even below) the spatial scale of individual camera pixels, we especially focus on the tight interplay of the camera response function (the diffraction-limited widening of a point source point-spread-function [PSF]), the camera pixelated sensor size, and the excitation laser beam waist in defining the amplitude of the measured temperature variations. Based on the solution of the three-dimensional heat transfer model in the presence of laser light illumination [20] and camera-based temperature detection, we outline the general guidelines and formalism to convert an exclusively morphological thermal map into a quantitative image of the concentration of laser-excited photo-thermal probes.
We validate our results and demonstrate a biologically relevant exemplary application with the non-destructive characterization of the spatial distribution and absolute molar concentration of melanin pigments in excised murine melanoma biopsies. By taking advantage of the endogenous photo-thermal effect primed by melanin absorption of focused 514 nm laser light, we spatially map melanin pigments in label-free configuration at sub-diffraction 40 µm resolution, and provide the data analysis protocol to retrieve the absolute pigment concentration with 10−4 M sensitivity. The space-resolved quantification of the melanin concentration is further strengthened with a lightness-based [21] k-means clustering algorithm [22] of hematoxylin-and-eosin (H&E) images [23] of the same tissue sections, and retains biological relevance in complementing the immuno-histochemical and morphological features that are traditionally extracted from the standard histopathological visual inspection of H&E-stained sections. Provided the still debated but crucial role played by melanin and altered melanogenesis in the overall development and evolution of melanoma [24,25,26,27], the melanin concentration is regarded as a promising marker for melanoma diagnosis by pathologists [25,28]. We expect therefore the proposed super-resolution photo-thermal imaging method to quantitatively complement traditional histopathology in the characterization of pigmented lesions ex-vivo.
Materials and methods
Super-resolution far-infrared thermography: principles and experimental setup
Our recently developed super-resolution photo-thermal imaging scheme [14,15] shares some of the working principles of the fluorescence-based super-resolution techniques of Photo-Activated Localization Microscopy (PALM) [29] and Stochastic Optical Reconstruction Microscopy (STORM) [30]. We take advantage of a photo-activated configuration of thermal imaging, where a focused low-power visible/near-infrared laser beam is raster-scanned and modulated in time to sparsely prime the photo-thermal effect (light absorption and thermal relaxation) inside the sample. The resulting sequence of isolated temperature increments is imaged by a far-infrared thermal camera, and the automated non-linear surface Gaussian fit of the individual thermal camera frames is exploited to retrieve the peak amplitude
Super-resolution thermography experiments are performed here on a custom-made benchtop optical setup [14]. The photo-thermal effect is primed by a continuous-wave Argon laser beam (
An uncooled microbolometer-based thermal camera (FLIR T650sc, FLIR Systems Inc., OR, USA; 7.5–13 μm spectral detection band, 640 × 480 sensor format, and 30 Hz frame rate) detects the thermal radiation in reflection configuration with numerical aperture N.A. = 0.023 and
For melanin pigments in murine melanoma biopsies, the emissivity has been set to
Transmission microscopy setup
Transmitted-light images of murine B16 melanoma biopsies have been acquired on a TCS SP5 STED-CW confocal microscope (Leica Microsystems, D). A 633 nm He–Ne beam with 10 μW power on the sample plane is focused by a 20× 0.5-N.A. air objective (HCX PL Fluotar, Leica Microsystems, D), and transmitted light is collected by a non-spectral dedicated photo-multiplier tube with no confocal pinhole along the detection optical path. Whole biopsies (up to cm2 in size) have been imaged in mosaic (tile-scan) mode with 400 Hz raster scan frequency per line.
B16 biopsies
The murine tumor B16 cell line has been cultured in IMDM-10 complete medium (10% heat-inactivated FBS, 2 mM l-glutamine, 100 U/mL penicillin, and 100 μg/mL streptomycin), and a tumorigenic dose of 2 × 106 cells at 70% confluence has been injected in the deep derma of the left flank of C57bl/6 mice at 7–12 weeks of age. Tumors have been collected from euthanized mice 14 days after cells injection, and explanted tissues have been cut in 20 μm sections on a cryostat upon embedding in OCT freezing medium (Bio-Optica, I). Sections have been adhered to 1 mm glass slides (Superfrost Plus, Thermo Fisher Scientific, MA, USA). Upon completion of super-resolution photo-thermal imaging, the same sections have been stained with H&E. Sections have been immersed in Meyer’s hematoxylin solution for 8 min and washed for 5 min in running water; they have been immersed in Eosin Y solution for 1 min, washed in running water, rinsed in distilled water, and dehydrated with passages in 95% and absolute alcohol. Slides have been cleared in xylene and mounted with Eukitt. A NanoZoomer scanner (Hamamatsu, JP) has been employed for the acquisition of H&E images.
All the experiments have been performed under protocols approved by the Institutional Animal Care and Use Committee of Università degli Studi di Milano-Bicocca and the Italian Ministry of Health.
Finite-element simulations
Finite-element simulations have been performed by COMSOL Multiphysics (Comsol AB, SE) based on the built-in heat transfer module. The Gaussian-shaped excitation laser source has been implemented as a square-wave pulse lasting

B16 melanoma biopsy imaged by super-resolution photo-activated thermography. (a) Super-resolved photo-thermal image of an unstained murine B16 melanoma biopsy;

Finite-element simulations. Green: temporal evolution of the amplitude of the temperature increment primed on a 20 μm skin tissue section (
Due to the overall axisymmetric properties of the system, cylindrical samples have been modeled as a 2D rectangle and the 3D temperature profile has been generated by rotation with significant reduction in the computational time. A triangular mesh has been employed for all the simulations with minimum element size of 80 nm and maximum element size of 10 µm.
k-means clustering
A color-based algorithm has been developed to automatically perform a semantic segmentation of the melanin content in B16 melanoma biopsies based on the corresponding 8-bit RGB-format H&E images. At first, background pixels in the image get identified by selecting a white-based threshold: each pixel is classified as a background pixel whenever all its red, green, and blue
Data acquisition and analysis software
The software FLIR Tools + (FLIR Systems Inc., OR, USA) has been employed for the acquisition of raw thermo-camera images. Image sequences have been exported in .csv file format and entirely processed by a custom-written Python code. Simulated datasets in Figures 3 and 4 have been generated by a custom Python code, which performs the numerical integration of the solution of the three-dimensional heat equation (equations (1–2)) by the built-in Scipy integration routine. For Figure 5, the 2-means clustering on H&E-stained sections has been carried out on a custom-written MATLAB code.

Quantification of the concentration of photo-thermal probes by super-resolved photo-activated thermography. (a) Temperature profile

Camera-based detection of temperature increments. (a)–(d) Thermal-camera frame simulated by spatially averaging a PSF-convoluted (and fourth-power elevated) temperature distribution over a fixed grid of adjacent pixels with

Quantification of the melanin concentration. (a) Super-resolution image of the absolute melanin molar concentration
Results
Super-resolution melanin-based photo-thermal imaging on B16 murine melanoma biopsies
Melanogenesis mainly occurs in melanocytes in the epidermal–dermal junction and provides the skin with ultraviolet-absorbing photo-protective agents: melanin shields the DNA of epidermal cells from the direct action of UV light, and scavenges the reactive oxygen species (ROS) associated with UV-induced oxidative stress [33]. Such a photo-protective ability of melanin pigments originates from a monotonically decreasing absorption spectrum in the 300–1,100 nm range [34,35]. Importantly, with a relatively low fluorescence quantum yield [34] upon one- and two-photon excitation, both eu- and pheo-melanins display a highly efficient thermal relaxation upon visible to near-infrared light absorption [5,36], and can be conveniently detected by photo-activated infrared thermography in label-free configuration.
We take advantage here of unstained excised murine B16 melanoma biopsies, and we prime the photo-thermal effect of melanin pigments at the 514 nm wavelength of a low-power Argon laser beam (
It is finally worth remarking that the adopted super-resolution image acquisition scheme provides a nearly twenty times resolution enhancement [14] relative to the mm-sized resolution of our thermal camera in conventional operation. structures that are clearly resolved in the super-resolution image of Figure 1a would appear therefore indiscernible in the conventional thermal image of the same sample.
From super-resolution photo-thermal imaging to the space-resolved quantification of photo-thermal probes: theoretical framework and experimental guidelines
Besides providing the necessary protection against ultraviolet radiation and reactive oxygen species in the skin, melanin acts as a marker for melanocytes differentiation [26]. It also plays a highly debated regulatory role in cancer evolution, by both decreasing the efficacy of antitumor treatments [27] (radio- and chemo-therapy) and by favoring tumor progression with immunosuppressive and mutagenic properties [25,26]. Recent findings [24,37] have also suggested that melanin granules may change cell elasticity, thereby modifying cell migration parameters which in turn play a key role in the process of metastasis of malignant melanoma. The quantification of melanin pigments in skin tissues and melanoma biopsies is regarded therefore as a potential diagnostic tool, and the melanin concentration has been suggested [25,28] as a useful parameter to be added to the standard reports of histopathological analyses to complement the morphological/immunohistochemical features traditionally extracted from the visual inspection of (H&E stained) melanocytic lesions ex-vivo.
In this framework, we outline the theoretical formalism to extract the absolute molar concentration of melanin pigments from super-resolved photo-thermal images of melanoma biopsies. While we exploit the concentration of melanin as a biologically relevant case study, we provide the readers with theoretical and experimental guidelines that have general validity in the conversion of temperature-based maps into quantitative images of the concentration C of photo-thermal laser-excited probes.
In order to extract quantitative information from detected
Under the single-layer approximation, the temperature
where
The spatial temperature profile predicted by the numerical integration of equation (1) on a
where d is the camera pixel size on the sample plane and
When the temperature profile predicted by equation (1) in Figure 3a is convoluted with our thermal-camera PSF, having
By the substitution of equation (1) into equation (2) and by the numerical integration over both space and time, the dependence of
According to the results of Figure 3d, the beam waist of the excitation laser beam has been adjusted to
Besides offering the advantage of a linear
We finally remark that equation (2) (and the conclusions drawn in Figure 3) assumes that the laser-induced temperature increment (convoluted with the camera PSF) is spatially averaged across a square pixel that is centered on the temperature peak itself. We now consider the general case of a possible shift between the temperature peak and the pixel center, and evaluate the effect of such a shift on the measured amplitude
We focus on three possible positions of the temperature peak, as marked and color-coded in the inset of Figure 4e: we consider (i) the extreme case of a temperature peak centered on the pixel corner, (ii) the intermediate case of a temperature peak centered on the pixel edge, and (iii) the case of a temperature peak centered midway between the pixel center and the pixel corner. For each condition, we generate a simulated camera frame by spatially averaging the PSF-convoluted (and fourth-power elevated) temperature profile over a fixed grid of adjacent pixels having 420 μm side (Figure 4b–d), we perform a Gaussian fit of the temperature distribution in the resulting image, and we compare the retrieved maximum amplitude
Quantification of the melanin absolute concentration by photo-thermal imaging
According to the results of Figure 3d, the super-resolution photo-thermal image of Figure 1a is converted into a map of the melanin molar concentration
Melanin specificity in the photo-thermal image of Figures 1a and 5a is ensured by the fact that melanin is the most prominent photo-thermal light absorber in melanoma and skin tissues along with hemoglobin, but the blood content and the impact of blood chromophores are expected to be significantly lower in excised biopsies than in vascularized tissues in vivo [42]. Furthermore, provided the 0.99 heat-release efficiency of melanin [36] and that the melanin absorption coefficient is about four times larger than the absorption coefficient of hemoglobin at 514 nm, the contribution of melanin dominates the detected photo-thermal signal [43].
Further confirmation of our melanin specificity can be achieved by an unsupervised machine learning analysis of the H&E image of the same tissue section. The H&E stain provides a comprehensive picture of the anatomy of a biological tissue with cellular nuclei being stained blue and the cells’ cytoplasm/extracellular matrix being stained pink [23]. Melanin is not specifically stained, and instead it appears in B16 melanoma as optically contrasted eumelanosome-like brown to black granules [23,27,44]. The pixel color content has already been exploited in the literature to segment melanin in histopathological images [45]. Here we take advantage of the pixel lightness value, the lightness
Results are shown in Figure 5c, where the H&E image of Figure 5b is segmented to highlight the result of three consecutive clustering iterations (clusters 1 to 3, respectively: cluster 3 corresponds to the cluster identified at the (n − 1)th iteration before convergence; cluster 2 corresponds to the cluster identified at the (n − 2)th iteration, and cluster 1 corresponds to the one identified at the (n − 3)th iteration). The average color of each group
If the clustered image of Figure 5c is overlaid to the super-resolution photo-thermal image of the same region (Figure 1a), pixels identified as clusters 1–3, respectively, appear to be associated with increasingly higher average detected temperature increments
Of note, our machine learning analysis of the H&E image qualitatively selects melanin-rich regions in the tissue but cannot provide any quantitative information regarding the melanin absolute concentration. The exact functional dependence of the color lightness (or of any other combination of the R, G, B color components) on the pigment concentration is not known a priori, preventing the conversion of clustered maps (like Figure 5c) into images of the melanin relative or absolute concentration. Furthermore, the comparison of Figure 5a and c suggests that the minimum melanin concentration that can be detected is lower for super-resolved thermography, with the pixels corresponding to extremely low temperature increments (
Discussion
In summary, we have reported the theoretical foundations of a data acquisition and analysis protocol that enables the space-resolved quantification of the absolute molar concentration of photo-thermal biomarkers in biological tissues by means of super-resolution photo-activated thermal imaging. We have further provided the demonstration of the experimental applicability of the proposed approach on murine B16 melanoma biopsies, where melanin pigments have been spatially mapped at sub-diffraction resolutions over whole (mm2–cm2) tissue sections and quantified with 10−4 M concentration sensitivity.
Our data acquisition and analysis protocol does not necessarily require single-layer thermally thick slabs, and can be applied to opaque thermally thin samples with surface light absorption by the only adaption of the solution of the 3D heat equation in Equation 1 [20]. Furthermore, super-resolved thermography can be applied to endogenous photo-thermal entities in label-free configuration, as demonstrated here with melanin pigments in melanoma biopsies, as well as to exogenous photo-thermal markers, these can be quantified given the knowledge of their heat release efficiency and molar extinction coefficient. Reflectance properties should also be accounted for in equation (1) whenever light scattering is not negligible with respect to light absorption. The approximation of negligible scattering by melanin pigments is suggested in the literature for dilute melanin solutions [46], and might represent a major source of uncertainty on recovered concentration values in the presence of pigment aggregation.
A general source of uncertainty on measured concentrations can be identified in the assumption of uniform absorption coefficient of photo-thermal probes inside the excitation laser spot. In practical terms, the detected laser-primed photo-thermal signal is interpreted here (equations (1) and (2)) as originating from a spatially uniform distribution of photo-thermal entities with effective concentration
Our application to melanocytic lesions exemplifies the biological and biomedical relevance of our work, but also implies a broader technological and instrumental impact. From the biomedical point of view, melanin has already been suggested as an important marker for melanoma diagnosis [25,28] but its quantification is not routinely performed in standard histopathological and immunohistochemical analyses of pigmented lesions ex-vivo. Existing techniques either rely on cell extracts to quantify melanin pigments with no spatial resolution [44,47,48], or spatially map melanin with no quantification of its concentration [3,5,28,49,50,51]. Difficulties in tackling the pigment concentration also affect the well-established Fontana-Masson method [52,53] and formalin-fixed paraffin-embedded melanin specific exogenous stains (e.g., melanoma-black-45, S-100, and Melan-A). The thermal imaging approach presented here combines tissue imaging at sub-diffraction resolution with the quantification of the melanin molar content and, by operating in label-free configuration without sample photo-damage, it can be efficiently combined with a subsequent H&E analysis of the same tissue sections. We expect therefore melanin-based super-resolution active thermal imaging to significantly expand the capability of state-of-the-art thermography in complementing standard histopathology in the characterization of biological tissues ex-vivo.
From the instrumental and technological point of view, super-resolution thermography can be exploited even in the presence of thermal cameras with smaller pixel size and higher nominal resolution. Commercially available cameras achieving
We conclude by remarking that a final technological improvement over the results presented here would involve a reduction in the imaging time, which is currently limited in the hours range due to the intrinsically time-consuming procedure of sample illumination and temperature localization. Possible strategies to reduce the total data acquisition time include the implementation of multi-spot laser illumination and the exploitation of sparsity-based data analysis algorithms recently proposed for super-resolution infrared thermography and capable of yielding faster results than the Gaussian image fitting exploited here [54].
-
Funding information: The authors acknowledge funding from the European Union’s Horizon 2020 research and innovation programme under grant agreement No. 964481 (FET Open project “IN2SIGHT”). The authors also acknowledge funding (2018-ATE-0070 and 2019-ATE-0126) from Università degli Studi di Milano-Bicocca.
-
Conflict of interest: The authors declare no conflicts of interest.
-
Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author upon reasonable request.
References
[1] Okabe K, Inada N, Gota C, Harada Y, Funatsu T, Uchiyama S. Intracellular temperature mapping with a fluorescent polymeric thermometer and fluorescence lifetime imaging microscopy. Nat Commun. 2012;3:705.10.1038/ncomms1714Search in Google Scholar PubMed PubMed Central
[2] Herman C. The role of dynamic infrared imaging in melanoma diagnosis. Expert Rev Dermatol. 2013;8:177–84.10.1586/edm.13.15Search in Google Scholar PubMed PubMed Central
[3] He J, Wang N, Tsurui H, Kato M, Iida M, Kobayashi T. Noninvasive, label-free, three-dimensional imaging of melanoma with confocal photothermal microscopy: Differentiate malignant melanoma from benign tumor tissue. Sci Rep. 2016;6:30209.10.1038/srep30209Search in Google Scholar PubMed PubMed Central
[4] Lapierre-Landry M, Gordon AY, Penn JS, Skala MC. In vivo photothermal optical coherence tomography of endogenous and exogenous contrast agents in the eye. Sci Rep. 2017;7:9228.10.1038/s41598-017-10050-5Search in Google Scholar PubMed PubMed Central
[5] Kobayashi T, Nakata K, Yajima I, Kato M, Tsurui H. Label-free imaging of melanoma with confocal photothermal microscopy: Differentiation between malignant and benign tissue. Bioengineering. 2018;5:67.10.3390/bioengineering5030067Search in Google Scholar PubMed PubMed Central
[6] Miyazaki J, Kobayahsi T. Photothermal microscopy for high sensitivity and high resolution absorption contrast imaging of biological tissues. Photonics. 2017;4(32):1–12.10.3390/photonics4020032Search in Google Scholar
[7] Donner JS, Thompson SA, Kreuzer MP, Baffou G, Quidant R. Mapping intracellular temperature using green fluorescent protein. Nano Lett. 2012;12:2107–11.10.1021/nl300389ySearch in Google Scholar PubMed
[8] Farzaneh M, Maize K, Lüeren D, Summers JA, Mayer PM, Raad PE, et al. CCD-based thermoreflectance microscopy: Principles and applications. J Phys D Appl Phys. 2009;42:143001.10.1088/0022-3727/42/14/143001Search in Google Scholar
[9] Vollmer M, Möllmann KP. Infrared thermal imaging: Fundamentals, research and applications. NJ, USA: Wiley-Vch; 2017.10.1002/9783527693306Search in Google Scholar
[10] Usamentiaga R, Venegas P, Guerediaga J, Vega L, Molleda J, Bulnes FG. Infrared thermography for temperature measurement and non-destructive testing. Sensors. 2014;14:12305–48.10.3390/s140712305Search in Google Scholar PubMed PubMed Central
[11] Pirtiniçetingl M, Herman C. Quantification of the thermal signature of a melanoma lesion. Int J Therm Sci. 2011;50:421–31.10.1016/j.ijthermalsci.2010.10.019Search in Google Scholar
[12] Okabe T, Fujimura T, Okajima J, Kambayashi Y, Aiba S, Maruyama S. First-in-human clinical study of novel technique to diagnose malignant melanoma via thermal conductivity measurements. Sci Rep. 2019;9:3853.10.1038/s41598-019-40444-6Search in Google Scholar PubMed PubMed Central
[13] Thang DC, Wang Z, Lu X, Xing B. Precise cell behaviors manipulation through light-responsive nano-regulators: Recent advance and perspective. Theranostics. 2019;9:3308–40.10.7150/thno.33888Search in Google Scholar PubMed PubMed Central
[14] Bouzin M, Marini M, Zeynali A, Borzenkov M, Sironi L, D’Alfonso L, et al. Photo-activated raster scanning thermal imaging at sub-diffraction resolution. Nat Commun. 2019;10:1–9.10.1038/s41467-019-13447-0Search in Google Scholar PubMed PubMed Central
[15] Marini M, Bouzin M, Sironi L, D’Alfonso L, Colombo R, Di Martino D, et al. A novel method for spatially-resolved thermal conductivity measurement by super-resolution photo-activated infrared imaging. Mater Today Phys. 2021;18:100375.10.1016/j.mtphys.2021.100375Search in Google Scholar
[16] Bouzin M, Marini M, Chirico G, Granucci F, Mingozzi F, Colombo R, et al. Melanin concentration maps by label-free super-resolution photo-thermal imaging on melanoma biopsies. Biomed Opt Exp. 2022;13:1173.10.1364/BOE.445945Search in Google Scholar PubMed PubMed Central
[17] Hecht E. Optics. MA, USA: Addison-Wesley; 2001.Search in Google Scholar
[18] Abbe E. Beiträge zur. Theorie des. Mikrosk. und der mikroskopischen Wahrnem. Arch für Mikroskopische Anat. 1873;9:413–68.10.1007/BF02956173Search in Google Scholar
[19] Graciani G, Amblard F. Super-resolution provided by the arbitrarily strong superlinearity of the blackbody radiation. Nat Commun. 2019;10.10.1038/s41467-019-13780-4Search in Google Scholar PubMed PubMed Central
[20] Bäuerle D. Laser processing and chemistry. Germany (Springer Berlin Heidelberg): Springer; 2011.10.1007/978-3-642-17613-5Search in Google Scholar
[21] Saravanan G, Yamuna G, Nandhini S. Real time implementation of RGB to HSV/HSI/HSL and its reverse color space models. In 2016 International Conference on Communication and Signal Processing (ICCSP). 2016. p. 462–6.10.1109/ICCSP.2016.7754179Search in Google Scholar
[22] Hartigan JA, Wong MA. Algorithm AS 136: A K-means clustering algorithm. J. R. Stat. Soc. Ser. C. (Appl. Stat). 1979;28:100–8.10.2307/2346830Search in Google Scholar
[23] Chan JKC. The wonderful colors of the hematoxylin-eosin stain in diagnostic surgical pathology. Int J Surg Pathol. 2014;22:12–32.10.1177/1066896913517939Search in Google Scholar PubMed
[24] Sarna M, Krzykawska-Serda M, Jakubowska M, Zadlo A, Urbanska K. Melanin presence inhibits melanoma cell spread in mice in a unique mechanical fashion. Sci Rep. 2019;9:1–9.10.1038/s41598-019-45643-9Search in Google Scholar PubMed PubMed Central
[25] Slominski RM, Zmijewski MA, Slominski AT. The role of melanin pigment in melanoma. Exp Dermatol. 2015;24:258–9.10.1111/exd.12618Search in Google Scholar PubMed PubMed Central
[26] Brozyna AA, Jóźwicki W, Carlson JA, Slominski AT. Melanogenesis affects overall and disease-free survival in patients with stage III and IV melanoma. Hum Pathol. 2013;44:2071–4.10.1016/j.humpath.2013.02.022Search in Google Scholar PubMed PubMed Central
[27] Brozyna AA, Józwicki W, Roszkowski K, Filipiak J, Slominski AT. Melanin content in melanoma metastases affects the outcome of radiotherapy. Oncotarget 2016;7:17844–53.10.18632/oncotarget.7528Search in Google Scholar PubMed PubMed Central
[28] Matthews TE, Piletic IR, Selim MA, Simpson MJ, Warren WS. Pump-probe imaging differentiates melanoma from melanocytic nevi. Sci Transl Med. 2011;3.10.1126/scitranslmed.3001604Search in Google Scholar PubMed PubMed Central
[29] Betzig E, Patterson GH, Sougrat R, Lindwasser OW, Olenych S, Bonifacino JS, et al. Imaging intracellular fluorescent proteins at nanometer resolution. Sci. (80-). 2006;313:1642–5.10.1126/science.1127344Search in Google Scholar PubMed
[30] Rust MJ, Bates M, Zhuang X. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM). Nat Methods. 2006;3:793–5.10.1038/nmeth929Search in Google Scholar PubMed PubMed Central
[31] Bernard V, Staffa E, Mornstein V, Bourek A. Infrared camera assessment of skin surface temperature - Effect of emissivity. Phys Medica. 2013;29:583–91.10.1016/j.ejmp.2012.09.003Search in Google Scholar PubMed
[32] Ogi H, Ishihara T, Ishida H, Nagakubo A, Nakamura N, Hirao M. Thermal Mode Spectroscopy for Thermal Diffusivity of Millimeter-Size Solids. Phys Rev Lett. 2016;117:195901.10.1103/PhysRevLett.117.195901Search in Google Scholar
[33] Solano F. Photoprotection and skin pigmentation: melanin-related molecules and some other new agents obtained from natural sources. Molecules. 2020;25(1537):1–18.10.3390/molecules25071537Search in Google Scholar
[34] Fu D, Ye T, Matthews TE, Yurtsever G, Warren WS. Two-color, two-photon, and excited-state absorption microscopy. J Biomed Opt. 2007;12:054004.10.1117/1.2780173Search in Google Scholar
[35] Sun C-K, Wu P-J, Chen S-T, Su Y-H, Wei M-L, Wang C-Y, et al. Slide-free clinical imaging of melanin with absolute quantities using label-free third-harmonic-generation enhancement-ratio microscopy. Biomed Opt Exp. 2020;11:3009.10.1364/BOE.391451Search in Google Scholar
[36] Meredith P, Riesz J. Radiative relaxation quantum yields for synthetic Eumelanin. Photochem Photobiol. 2004;79:211.10.1562/0031-8655(2004)079<0211:RCRQYF>2.0.CO;2Search in Google Scholar
[37] Sarna M, Zadlo A, Hermanowicz P, Madeja Z, Burda K, Sarna T. Cell elasticity is an important indicator of the metastatic phenotype of melanoma cells. Exp Dermatol. 2014;23:813–8.10.1111/exd.12535Search in Google Scholar
[38] Ansari MA, Mohajerani E. Mechanisms of laser-tissue interaction: I. Optical properties of tissue. J Lasers Med Sci. 2011;2:119–25.Search in Google Scholar
[39] Piletic IR, Matthews TE, Warren WS. Estimation of molar absorptivities and pigment sizes for eumelanin and pheomelanin using femtosecond transient absorption spectroscopy. J Chem Phys. 2009;131:1–4.10.1063/1.3265861Search in Google Scholar
[40] Shu X, Li H, Dong B, Sun C, Zhang HF. Quantifying melanin concentration in retinal pigment epithelium using broadband photoacoustic microscopy. Biomed Opt Exp. 2017;8:2851.10.1364/BOE.8.002851Search in Google Scholar
[41] Morishima T, Fukuda E. Quantitative analysis of eumelanin and pheomelanin in human malignant-melanoma tissues. Arch Dermatol Res. 1985;277:248–50.10.1007/BF00404326Search in Google Scholar
[42] Hult J, Merdasa A, Pekar-Lukacs A, Tordengren Stridh M, Khodaverdi A, Albinsson J, et al. Comparison of photoacoustic imaging and histopathological examination in determining the dimensions of 52 human melanomas and nevi ex vivo. Biomed Opt Exp. 2021;12:4097.10.1117/12.2603561Search in Google Scholar
[43] Meglinski I V, Matcher SJ. Quantitative assessment of skin layers absorption and skin reflectance spectra simulation in the visible and near-infrared spectral regions. Physiol Meas. 2002;23:741–53.10.1088/0967-3334/23/4/312Search in Google Scholar PubMed
[44] Ito S, Jimbow K. Quantitative analysis of eumelanin and pheomelanin in hair and melanomas. J Invest Dermatol. 1983;80:268–72.10.1111/1523-1747.ep12534616Search in Google Scholar PubMed
[45] Kłeczek P, Lech M, Jaworek-Korjakowska J, Dyduch G, Tadeusiewicz R. Segmentation of black ink and melanin in skin histopathological images. Proc SPIE. 2018;10581:335–42. 10.1117/12.2292859.Search in Google Scholar
[46] Riesz J, Gilmore J, Meredith P. Quantitative scattering of melanin solutions. Biophys J. 2006;90:4137–44.10.1529/biophysj.105.075713Search in Google Scholar PubMed PubMed Central
[47] Chung S, Lim GJ, Lee JY. Quantitative analysis of melanin content in a three-dimensional melanoma cell culture. Sci Rep. 2019;9:1–9.10.1038/s41598-018-37055-ySearch in Google Scholar PubMed PubMed Central
[48] Fernandes B, Matamá T, Guimarães D, Gomes A, Cavaco-Paulo A. Fluorescent quantification of melanin. Pigment Cell Melanoma Res. 2016;29:707–12.10.1111/pcmr.12535Search in Google Scholar PubMed
[49] Dimitrow E, Riemann I, Ehlers A, Koehler MJ, Norgauer J, Elsner P, et al. Spectral fluorescence lifetime detection and selective melanin imaging by multiphoton laser tomography for melanoma diagnosis. Exp Dermatol. 2009;18:509–15.10.1111/j.1600-0625.2008.00815.xSearch in Google Scholar PubMed
[50] Fu D, Ye T, Matthews TE, Grichnik J, Hong L, Simon JD, et al. Probing skin pigmentation changes with transient absorption imaging of eumelanin and pheomelanin. J Biomed Opt. 2008;13:054036.10.1117/1.2976424Search in Google Scholar PubMed
[51] He J, Miyazaki J, Wang N, Kobayashi T. Label-free imaging of melanoma with nonlinear photothermal microscopy. Opt Lett. 2015;40:1141–4.10.1364/OL.40.001141Search in Google Scholar PubMed
[52] Joly-Tonetti N, Wibawa JID, Bell M, Tobin D. Melanin fate in the human epidermis: a reassessment of how best to detect and analyse histologically. Exp Dermatol. 2016;25:501–4.10.1111/exd.13016Search in Google Scholar PubMed
[53] Carriel VS, Aneiros-Fernandez J, Arias-Santiago S, Garzón IJ, Alaminos M, Campos A. A novel histochemical method for a simultaneous staining of melanin and collagen fibers. J Histochem Cytochem. 2011;59:270–7.10.1369/0022155410398001Search in Google Scholar PubMed PubMed Central
[54] Burgholzer P, Berer T, Gruber J, Mayr G. Super-resolution thermographic imaging using blind structured illumination. Appl Phys Lett. 2017;111:031908.10.1063/1.4995410Search in Google Scholar
© 2022 Mario Marini et al., published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Diagnostic accuracy of genetic markers for identification of the Lr46/Yr29 “slow rusting” locus in wheat (Triticum aestivum L.)
- NADPH-derived ROS generation drives fibrosis and endothelial-to-mesenchymal transition in systemic sclerosis: Potential cross talk with circulating miRNAs
- Effect of omega-3 fatty acids on the telomere length: A mini meta-analysis of clinical trials
- Analysis of differentially expressed genes and signaling pathways involved in atherosclerosis and chronic obstructive pulmonary disease
- The epigenetic dimension of protein structure
- Gamma-induced mutants of Bacillus and Streptomyces display enhanced antagonistic activities and suppression of the root rot and wilt diseases in pulses
- Corticosterone potentiates ochratoxin A-induced microglial activation
- Supercomplex supercomplexes: Raison d’etre and functional significance of supramolecular organization in oxidative phosphorylation
- Insights into functional connectivity in mammalian signal transduction pathways by pairwise comparison of protein interaction partners of critical signaling hubs
- The effects of supplementation of Nannochloropsis oculata microalgae on biochemical, inflammatory and antioxidant responses in diabetic rats
- Molecular epidemiology of human papillomavirus in pregnant women in Burkina Faso
- Review Articles
- Interaction of cervical microbiome with epigenome of epithelial cells: Significance of inflammation to primary healthcare
- Seaweeds’ pigments and phenolic compounds with antimicrobial potential
- The capture of host cell’s resources: The role of heat shock proteins and polyamines in SARS-COV-2 (COVID-19) pathway to viral infection
- Erratum
- Erratum to “Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents”
- Special Issue on XXV Congress of the Italian Society for Pure and Applied Biophysics
- Low-temperature librations and dynamical transition in proteins at differing hydration levels
- The phosphoinositide PI(3,5)P2 inhibits the activity of plant NHX proton/potassium antiporters: Advantages of a novel electrophysiological approach
- Targeted photoimmunotherapy for cancer
- Calorimetry of extracellular vesicles fusion to single phospholipid membrane
- Calcium signaling in prostate cancer cells of increasing malignancy
- Oxygen diffusion pathways in mutated forms of a LOV photoreceptor from Methylobacterium radiotolerans: A molecular dynamics study
- A photosensitizing fusion protein with targeting capabilities
- Ion channels and neuronal excitability in polyglutamine neurodegenerative diseases
- Is styrene competitive for dopamine receptor binding?
- Diffusion of molecules through nanopores under confinement: Time-scale bridging and crowding effects via Markov state model
- Quantitative active super-resolution thermal imaging: The melanoma case study
- Innovative light sources for phototherapy
- Electrophysiological study of the effects of side products of RuBi-GABA uncaging on GABAA receptors in cerebellar granule cells
- Subcellular elements responsive to the biomechanical activity of triple-negative breast cancer-derived small extracellular vesicles
Articles in the same Issue
- Research Articles
- Diagnostic accuracy of genetic markers for identification of the Lr46/Yr29 “slow rusting” locus in wheat (Triticum aestivum L.)
- NADPH-derived ROS generation drives fibrosis and endothelial-to-mesenchymal transition in systemic sclerosis: Potential cross talk with circulating miRNAs
- Effect of omega-3 fatty acids on the telomere length: A mini meta-analysis of clinical trials
- Analysis of differentially expressed genes and signaling pathways involved in atherosclerosis and chronic obstructive pulmonary disease
- The epigenetic dimension of protein structure
- Gamma-induced mutants of Bacillus and Streptomyces display enhanced antagonistic activities and suppression of the root rot and wilt diseases in pulses
- Corticosterone potentiates ochratoxin A-induced microglial activation
- Supercomplex supercomplexes: Raison d’etre and functional significance of supramolecular organization in oxidative phosphorylation
- Insights into functional connectivity in mammalian signal transduction pathways by pairwise comparison of protein interaction partners of critical signaling hubs
- The effects of supplementation of Nannochloropsis oculata microalgae on biochemical, inflammatory and antioxidant responses in diabetic rats
- Molecular epidemiology of human papillomavirus in pregnant women in Burkina Faso
- Review Articles
- Interaction of cervical microbiome with epigenome of epithelial cells: Significance of inflammation to primary healthcare
- Seaweeds’ pigments and phenolic compounds with antimicrobial potential
- The capture of host cell’s resources: The role of heat shock proteins and polyamines in SARS-COV-2 (COVID-19) pathway to viral infection
- Erratum
- Erratum to “Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents”
- Special Issue on XXV Congress of the Italian Society for Pure and Applied Biophysics
- Low-temperature librations and dynamical transition in proteins at differing hydration levels
- The phosphoinositide PI(3,5)P2 inhibits the activity of plant NHX proton/potassium antiporters: Advantages of a novel electrophysiological approach
- Targeted photoimmunotherapy for cancer
- Calorimetry of extracellular vesicles fusion to single phospholipid membrane
- Calcium signaling in prostate cancer cells of increasing malignancy
- Oxygen diffusion pathways in mutated forms of a LOV photoreceptor from Methylobacterium radiotolerans: A molecular dynamics study
- A photosensitizing fusion protein with targeting capabilities
- Ion channels and neuronal excitability in polyglutamine neurodegenerative diseases
- Is styrene competitive for dopamine receptor binding?
- Diffusion of molecules through nanopores under confinement: Time-scale bridging and crowding effects via Markov state model
- Quantitative active super-resolution thermal imaging: The melanoma case study
- Innovative light sources for phototherapy
- Electrophysiological study of the effects of side products of RuBi-GABA uncaging on GABAA receptors in cerebellar granule cells
- Subcellular elements responsive to the biomechanical activity of triple-negative breast cancer-derived small extracellular vesicles