Next Article in Journal
Centrifugation-Induced Stable Colloidal Silver Nanoparticle Aggregates for Reproducible Surface-Enhanced Raman Scattering Detection
Previous Article in Journal
Carbon Nanotube-Based Field-Effect Transistor Biosensors for Biomedical Applications: Decadal Developments and Advancements (2016–2025)
Previous Article in Special Issue
Detection of Exosomal microRNA from Malignant Lung Cells Using a Surface Plasmon Resonance Imaging Biosensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Non-Invasive Imaging Technologies for Cutaneous Melanoma

1
Department of Dermatology, University of California Irvine, Irvine, CA 92617, USA
2
Department of Dermatology, Wayne State University School of Medicine, Wayne State University, Detroit, MI 48202, USA
3
The Richard and Loan Hill Department of Biomedical Engineering, University of Illinois at Chicago, Chicago, IL 60607, USA
4
Department of Dermatology and Cutaneous Surgery, University of Miami Miller School of Medicine, Miami, FL 33136, USA
5
Department of Medicine, University of Chicago, Chicago, IL 60637, USA
6
Department of Dermatology, Memorial Sloan Kettering Cancer Center, New York, NY 10065, USA
7
Department of Dermatology, University of Illinois at Chicago College of Medicine, Chicago, IL 60607, USA
*
Author to whom correspondence should be addressed.
Biosensors 2025, 15(5), 297; https://doi.org/10.3390/bios15050297
Submission received: 13 January 2025 / Revised: 7 April 2025 / Accepted: 17 April 2025 / Published: 7 May 2025
(This article belongs to the Special Issue Advanced Optical Methods for Biosensing)

Abstract

:
Imaging technologies are constantly being developed to improve not only melanoma diagnosis, but also staging, treatment planning, and disease progression. We start with a description of how melanoma is characterized using histology, and then continue by discussing nearly two dozen different technologies, including systems currently used in medical practice and those in development. For each technology, we describe its method of operation, how it is or would be projected to be most commonly used in diagnosing and managing melanoma, and for systems in current use, we identify at least one current manufacturer. We also provide a table including the biomarkers identified by and main limitations associated with each technology and conclude by offering suggestions on specific characteristics that might best enhance a technology’s potential for widespread clinical adoption.

1. Introduction

The worldwide incidence of cutaneous melanoma has increased over the last two decades [1,2,3,4,5,6]. Notably, the annual incidence has risen 4–6% in light skin populations, with a lifetime risk of one in thirty-three (3%) for Caucasians [7,8,9,10]. The increased detection of melanoma has likely played a role in the rise in diagnoses over the past 40 years [11,12]. While survival from the most common cancer types in the United States has increased from improvements in early detection and treatment, melanoma is still increasing in both morbidity and mortality each year [13,14]. Since 1975, the death rate has increased by 16%, with men being afflicted more than women [14]. Genetics and ultraviolet light exposure are known risk factors for melanoma [15], but the lack of availability of dermatologists in the county of residence is another independent risk factor [16,17]. Additionally, while the risk of melanoma diagnosis increases with age, with 65 being the average age of diagnosis, melanoma is still one of the most frequent cancer types diagnosed in individuals under the age of 30 [10,18]. Melanoma can appear anywhere on the body, but the most common locations are on the head and neck, back, and lower extremities [18].
Melanoma has the highest mortality of any skin cancer type, accounting for 75% of skin cancer deaths annually, despite making up only 1% of diagnosed skin cancers [10,19,20]. Lesion thickness is one the most important predictors of mortality of melanoma; therefore, early diagnosis is crucial for a better prognosis [21]. The current predicted five year survival rate for patients diagnosed with melanoma is greater than 99% for tumors excised before metastasis, decreasing to 75% with nodal metastasis, and to 49% for patients with distant metastasis [22].
The most widely accepted and utilized method of diagnosis relies on a skin exam adhering to ABCDE criteria, followed by biopsy and histopathologic analysis of suspicious lesions. This simple procedure relies heavily on dermatologist expertise; thus, its efficacy is variable [23]. Furthermore, since access to dermatologists varies widely by place of residence, patient demographics, and insurance status, the responsibility of performing full-body skin exams often falls upon primary care physicians who may or may not have recent training for melanoma detection [24]. These factors lead to both underdiagnosis and overdiagnosis. Underdiagnosis by primary care physicians causes melanomas detected by non-dermatologists to be thicker and later stage melanomas than those detected by dermatologists. Overdiagnosis leads to very high biopsy rates, with some centers reporting as many as 287 skin biopsies to diagnose one melanoma [25]. The increasing incidence of melanoma, the need to reduce costs of melanoma detection, as well as to form objective methods in skin cancer screening, has created a demand for an accurate, non-invasive, cost-effective, diagnostic imaging modality for melanoma. To that end, an explosion of detection techniques has been developed and adapted to the skin [26,27,28,29,30,31,32,33,34]. However, many challenges remain with regards to identifying and implementing prevention and early detection recommendations, understanding what drives the differences between dormancy and metastasis, and developing targeted therapy recommendations [35]. Appropriate imaging modalities may play a role in meeting these challenges. The connection between better screening and diagnostic information and prevention and early detection is straightforward. In addition, emerging imaging modalities may also play a part in identifying biomarkers of metastasis and assessing treatment efficacy sooner and non-invasively.
The histological subtype is important in the staging of melanoma, although it is not the only consideration. In melanoma, relevant prognostic biomarkers for staging include the tumor thickness (Breslow depth), Clark level, dermal mitotic rate per square millimeter, presence of lymphocytic invasion, degree of atypia, ulceration, tumor regression, presence of perineural or angiolymphatic invasion (tumor invasion of the dermis microvasculature), neurotropism, microsatellitosis, and margin assessment [36,37,38]. Tumor thickness and ulceration are important prognostic factors, with thickness being the most accurate tumor characteristic predictive of survival [39,40,41]. The tumor stage (T) is determined by the tumor thickness and presence of ulceration [38]. The nodal stage (N) is determined by the number of lymph nodes involved. This can be determined via sentinel lymph node biopsy (SNLB) and physical exam [42]. Although not included in the staging of melanoma, a high mitotic rate is associated with an increased risk of SLN metastasis [41,43,44]. The metastatic stage (M) of the disease is determined by the presence or absence of metastasis and the site of metastasis (skin, lymph nodes, viscera, lungs, etc.). Melanoma without metastasis is defined as either stage I or II, depending on the extent of vertical invasion. Either microscopic or gross metastasis to lymph nodes defines stage III melanoma. Stage IV melanoma is characterized by distant metastasis and elevated levels of serum lactate dehydrogenase (LDH) [38,45,46]. Stage 0 melanoma (in situ) is defined by the tumor cells being microscopically identified but confined to the epidermis, with no evidence that cancer has spread to the lymph nodes or distant sites [47].
Beyond lower morbidity and mortality for the patient, targeted early detection screening programs can be cost effective [48,49]. The advent of adjuvant treatments for advanced melanoma (e.g., one year of adjuvant pembrolizumab is estimated to cost over USD 160,000) emphasizes the potential cost savings to the healthcare system, should a comprehensive early detection screening program be implemented. Imaging modalities could also lead to non-invasive monitoring of the course of melanoma progression from diagnostics to differentiation of metastatic dormancy and progression, and even monitor response to therapy. In this review, we thoroughly describe the operation and strengths and shortcomings of over 20 imaging and spectroscopic modalities for melanoma screening, staging, treatment planning, and disease tracking by analyzing each modality for accuracy, reproducibility, cost, and current technology readiness level. This is not intended to serve as a scoping review of every clinical trial that has been performed using each of the described technologies but rather as an entry for familiarization with the breadth of technologies available for this challenging disease. For technologies still in development, we also assess their potential for widespread adoption.

2. Melanoma Characterization (Without Imaging)

2.1. Melanoma Formation and Subtypes

Melanoma begins in melanocytes, which are primarily found in the basal layer of the epidermis. Malignant melanocytes, that is, melanocytes with uncontrolled growth, form from genetic mutations. Melanomas typically follow a radial growth phase in the epidermis, followed by a vertical growth phase which leads to invasive spread, with different forms of melanoma progressing at different rates, many never going beyond radial growth, and some rapidly progressing to vertical growth [50].
Melanoma has a wide variety of presentations, and histopathological features vary by type of melanoma [51,52]. Histopathological features of superficial spreading melanoma (SSM) include asymmetry, poor circumscription, and lack of cellular maturation. In SSM, malignant melanocytes, large, atypical epitheliods with large nuclei, can spread as single cells or nests. The presence of melanocytes above the basal layer is known as Pagetoid spread; this is common in SSM but is also seen, to a lesser extent, in benign nevi. Melanocytic nests in SSM will display dyscohesion; they will look like they are falling apart. Nodular melanoma (NM) often displays a thinning of the epidermis and dermis, and histopathological features can include a nodular confluence of atypical melanocytes that are epitheloid or spindled with frequent and often atypical mitoses. Balloon cells are also seen. In the vertical growth phase, the dermal component of NM looks very similar to the dermal component of SSM. By contrast, lentigo maligna melanoma (LMM) is a slow-growing form of melanoma that can be identified by the proliferation of atypical melanocytes in the epidermal basal layer, where the atypical melanocytes are polygonal with atypical nuclei and the epidermis is often described as atrophic. In the dermis, melanocytes are hyperchromatic, typically small, and may be spindle-shaped or multinucleated. As LMM progresses, nodules are formed in the dermis and the atypical melanocytes proliferate along the dermal epidermal junction (DEJ) and down cutaneous appendages. Unlike in SSM and NM, in LMM, Pagetoid spread is not common, but the dermis shows solar elastosis (yellowing of skin due to sun damage). Acral lentiginous melanoma (ALM) is usually found within nail beds. Melanocytes present as nests and single cells along the DEJ. Pagetoid upward migration is widespread and melanocytes in the epidermis resemble those seen in LMM. Dermal invasion often tracks down eccrine structures and aggregates around blood vessels. Large DEJ nests of atypical melanocytes can be found.

2.2. Pathological and Histopathological Analysis of Melanoma

The analysis of a biopsied skin sample starts with fixing the tissue in formalin, embedding it in paraffin, and thinly slicing it. The slices are mounted on glass slides and stained, usually by hematoxylin and eosin (H&E). This staining enables pathologists to analyze tissue at the cellular level to identify the presence, type, and stage of melanoma.
All of features identified above can be seen by H&E staining, but melanoma is heterogeneous and there are histological mimics of melanoma [53]. In fact, differences in the interpretation of morphological features can lead to high levels of interobserver variation: in one study, 17% of diagnoses were recommended to be reclassified when reviewed by a specialist panel (both false positives and false negatives identified) [54]. This has led to incorporation of immunohistochemical (IHC) stains of melanocytic markers and proliferative markers. Melanocytic markers include S-100, Melan-A, and HMB-45 [55]. S-100 is a calcium-binding protein expressed by melanoma cells; Melan-A is a melanoma-associated antigen, and HMB-45 is a monoclonal antibody that targets the premelanosome protein gp100 [56]. The most commonly used proliferation marker is Ki-67, which is highly elevated in the most aggressive melanomas [57]. The combination of morphological features and molecular features increases the melanoma detection accuracy.

3. Review Method

To investigate the emerging non-invasive techniques in melanoma diagnosis, staging, and monitoring, we utilized PubMed and Google Scholar and searched “melanoma” with each of the following more than 20 modalities (Figure 1): digital photography (including total body photography), dermoscopy, hyperspectral imaging, multispectral imaging, electrical impedance spectroscopy, electron paramagnetic resonance spectroscopy, reflectance confocal microscopy, photoacoustic imaging (and optoacoustic imaging), optical coherence tomography, non-interferometric photoacoustic remote sensing microscopy, Raman spectroscopy, elastic scattering spectroscopy, real-time elastography, terahertz pulsed imaging, multiphoton imaging, ultrasound (including high-frequency ultrasound), magnetic resonance imaging, positron emission tomography, single photon emission computed tomography, fiber diffraction, and Fourier transform infrared spectroscopy (and microspectroscopy), restricting our search to studies published between 1995 and 2024. It should be noted that as we delved into the subject, we realized the value in including sensing technologies and spectroscopy methods not commonly considered “imaging”. These were incorporated into the review in order to cover the full range of non-invasive techniques involved in melanoma diagnosis, staging, and monitoring.
Our searches (Figure 2) generated over 400,000 articles. For each search, we identified a selection of articles from high-impact journals and from leaders in the field, with the highest citation counts and/or that otherwise best exemplified the use of the technology for melanoma staging, screening, and monitoring, in our opinion. We discuss the principles of each modality, followed by a summary of their strengths and limitations. The first section describes imaging modalities commonly used in daily medical practice and those with more robust data in melanoma diagnosis and management; the second section describes technologies still under development.

4. Imaging Modalities Currently Used in Medical Practice

4.1. Photography

Methods of photography, including total body digital photography (TBDP) [58], mole mapping, various forms of 3D photography, and smartphone photography [59], are being increasingly utilized to diagnose melanoma [60,61,62,63,64]. Photography allows clinicians to study lesions visually over time and is a useful method to help diagnose melanoma in high-risk patients, including those with multiple dysplastic nevi, a family history of melanoma, or both [63]. Recent studies have described the utility of photography in detecting early melanomas [65,66], which have greater accuracy when combined with dermoscopy [67,68]. Digital photography only allows for the naked eye evaluation of lesions [69]. Moreover, photography can play an essential role in telemedicine, especially in rural settings [70]. A shortcoming of smartphone photography is that the cameras produce inaccurate colors when operating in automatic mode and use a filter which intensifies edges. These deficiencies can be eliminated through adjusting camera parameters manually or during post-production [64]. An important limitation of photography is that it only images the surface features of the lesions [65]. Photography may be more useful in older patients, as one study showed that less than 1% of new lesions were histologically confirmed to be melanoma in patients younger than 50 years old, while 30% of new lesions were melanoma for those older than 50 [71]. A recent review of TBDP studies found the method enhances the surveillance and detection of new lesions, and is less time-consuming, but did not outperform dermatologists and should be integrated with dermoscopy and dermatologist expertise [72]. Examples of image assessments by smartphone (Figure 3a) and by TBDP (Figure 3b) are included below. TBDP systems such as 3D Vectra are offered by Canfield Scientific (Parsippany, NJ, USA) and FotoFinder Systems GmbH (Bad Birnbach, Germany), among others.

4.2. Dermoscopy

Dermoscopy is performed with a relatively inexpensive device, the dermatoscope. This simple operator-dependent modality is well suited for the diagnosis of pigmented lesions [73,74]. Dermoscopy relies on color and structure to differentiate between melanoma and benign nevi [75]. Dermoscopy offers magnified (6×–100×) images of the lesion in the horizontal plane [76,77]. There are two main dermoscopy systems: (i) immersion, non-polarized contact dermoscopy, which uses fluid to improve contact between the lens and the lesion and decrease the light reflected by the stratum corneum, and (ii) polarized light dermoscopy, which uses a filter to block reflected light. Dual-mode dermoscopy incorporates both systems, allowing the physician to examine the superficial components of the skin in immersion mode and the deeper structures in the polarized mode [78]. A study by Benvenuto-Andrade et al. found that under polarized light, melanin looked sharper and darker and vessels were better visualized, making it more useful in identifying malignancies compared to non-polarized light [79].
With dermoscopy, the physician can apply a variety of algorithms to determine if a suspicious lesion should undergo biopsy, including the three-point method, seven-point checklist, Menzies method, pattern analysis, and others [69,80,81]. General features of melanoma on dermoscopy include asymmetry, numerous colors, negative pigment network, chrysalis structures, blue–white veil, and regression structures, among others [77,82].
Numerous studies have described a significant improvement in the diagnostic accuracy of melanoma through the use of dermoscopy [69,83,84,85,86]. Sensitivities of 85–95% and specificities of 73–86% for the dermoscopic identification of melanoma were reported in a recent scoping review article [80]. A retrospective study showed that the excision ratio of benign–malignant lesions decreased from 18:1 in the pre-dermoscopy era to 4:1 with the use of dermoscopy in a specialist setting [87]. Super-high-resolution dermoscopy has recently been implemented with 400× magnification (typical dermoscopy utilizes 20× magnification). Super-high-resolution dermoscopy was able to identify individual pigmented cells in malignant melanoma and other features that could help differentiate melanomas from benign nevi [88]. A limitation of this modality (super-high-resolution dermoscopy) is the small field of view and the potential for inter-operator variability in analysis.
Dermoscopy is a quick, non-invasive method to increase diagnostic accuracy in real-time for clinicians, but it relies on biopsy and histologic analysis to confirm a diagnosis of melanoma [89,90]. Further, dermoscopy remains limited as a diagnostic tool, as certain melanoma subtypes, such as desmoplastic melanoma, often present devoid of characteristic dermoscopic features. Moreover, because dermoscopic features vary widely on different body locations, its efficacy depends heavily on the skill of the practitioner [89]. To resolve this problem, computer-aided digital dermoscopic image analysis is being refined to improve diagnostic performance for all practitioners [81,90,91,92,93]. Sequential digital dermoscopic imaging (SDDI) is a technique used to store and retrieve dermoscopic images for comparative analysis over time. SDDI monitors lesion evolution, but it requires frequent visits, often performed every three months. Encouragingly, SDDI has been shown to detect melanoma earlier in its progression [94,95,96,97]. Dermoscopic images of lentigo maligna are shown in Figure 3c. Manufacturers of dermoscopic equipment include DermLite (Aliso Viejo, CA, USA) and Canfield Scientific, (Parsippany, NJ, USA) among others.

4.3. Electrical Impedance Spectroscopy

Skin lesions can be analyzed by measuring the electrical impedance of cutaneous structures and comparing the value to nearby healthy skin, since pathogenic alterations in tissue influence the ability of cells to conduct electricity [98]. Electrical impedance spectroscopy (EIS) applies an alternating electric current to the skin and can detect changes in cell size, orientation, shape, and structure of the cell membrane. EIS does not create images of the skin; rather, it uses algorithms to provide a score based on the resistance [99]. Electrical resistance is measured over a range of four different depths (different colors) and ten permutations at different frequencies (1.0 kHz to 2.5 MHz), utilizing a safe voltage (i.e., 150 mV and 75 micro-A) [100]. The device is intended for lesions measuring between 2 mm and 20 mm in diameter [98].
EIS is a simple, fast, and safe procedure used to increase diagnostic accuracy. Figure 3d shows a typical implementation of EIS on a suspected lesion. EIS can assess lesions deemed suspicious on clinical and dermoscopic examination, with high reported sensitivity in detecting melanoma [101]. Various classification algorithms of EIS imaging have been developed [100,102]. A recent review article found reported sensitivities of 100–95% and specificities of 69.5% to 58.6% [103]. While EIS may be useful in detecting early melanomas and monitoring lesions over time, it has a high false-positive rate, and inflammation, ulceration, or scar tissue may further limit its validity [99,101,104]. EIS should be performed by physicians who are trained to clinically detect skin cancer because benign lesions, such as seborrheic keratoses, are frequently inaccurately classified as malignant by EIS [99,101]. One of the most widely known EIS devices is the Nevisense system (SciBase AB, Stockholm, Sweden). The manufacturer noted that an update fee schedule for the use of Nevisense for melanoma detection was published in 2023 [105].

4.4. Reflectance Confocal Microscopy

Reflectance confocal microscopy (RCM) provides cellular-level high-resolution images of the epidermis and papillary dermis in real-time [106,107,108]. Superficial spreading melanoma in situ (dermoscopic and RCM images) is shown in Figure 3e [109]. RCM has shown diagnostic potential in differentiating benign melanocytic lesions (nevi) from melanoma. RCM operates by directing a focused laser beam at the skin and capturing backscattered light from different tissue depths, using a spatial pinhole to eliminate out-of-focus signals, thus producing high-resolution, enface images of cellular structures in the epidermis and dermis with micrometer-scale optical sectioning. [89]. RCM is capable of presenting two-dimensional (en face) horizontal images of the skin up to a depth of a few hundred to several hundred μm, depending on the system configuration, including the light source wavelength. It is typically used to image from the top-most layer of the stratum corneum into the superficial (papillary) dermis [76,106,110,111,112]. These sequential depth images can be stacked into a three-dimensional rendering of the imaged area [106].
RCM relies on the intrinsic reflective structures found in the skin, such as free cytoplasmic melanin, melanosomes, and keratin, which provide a sharp contrast for near infrared (NIR) light sources [108,110]. RCM allows for near histological cellular resolution (0.5–1 µm lateral and 3–5 µm axial), providing the shape, distribution, and morphology of cells, as well as the visualization of the dermal–epidermal junction and vessels [107,113]. Additionally, this technique shows a good correlation with dermoscopic and histologic findings of malignancies such as basal cell carcinoma (BCC), squamous cell carcinoma (SCC), and melanoma, especially the LMM subtype [114]. It is particularly well suited for the examination of flat skin lesions and is used to define tumor margins and monitor responses to therapies [112,115].
Because RCM has higher specificity than dermoscopy, it leads to fewer unnecessary biopsies, which is particularly useful for the in vivo imaging of cosmetically sensitive areas such as the face and genital region [116]. Additionally, because the biomarkers for melanoma that RCM relies on are not pigment-based (e.g., atypical keratinocytes, pagetoid cells, changes in skin architecture), RCM has been used to detect both pigmented and amelanotic melanoma that are challenging to diagnose with dermoscopy [110,111,112]. Other uses of RCM include assessing margins of slow-growing LMM [117] and as an adjunct modality for lesions requiring re-excision [118]. Because RCM optical sectioning provides cellular-level image interpretation if widely used, it could significantly reduce the number of biopsies performed for in situ melanoma. In addition, RCM can be also used ex vivo on freshly excised tissue with slight laboratory processing (gentle flattening, moistening, with no need for tissue fixation or staining), and can be used in Mohs micrographic surgery to accelerate the definition of surgical margins [116].
A recent meta-analysis of 32 studies shows a pooled sensitivity and specificity of 92% and 70% for RCM [119]. Although RCM provides excellent resolution, significant dermatological expertise is needed to interpret RCM images for melanoma detection. For example, it is difficult to distinguish dendritic melanoma cells (pagetoid cells) from dendritic benign Langerhans cells, which can be seen in pigmented actinic keratosis and traumatized nevi, often leading to the overdiagnosis of melanoma [116,120]. Consequently, the diagnostic accuracy of RCM depends upon the skill of the user, where expertise is acquired through extensive training [98]. A major factor in the implementation of RCM is that individual RCM systems can cost >USD 100 k. This makes them too costly for most independent dermatology practices. Caliber ID (Rochester, NY, USA) is a manufacturer of RCM devices. RCM obtained billing codes in 2018 and is being integrated in clinics across the US.

4.5. Optical Coherence Tomography

Different configurations of optical coherence tomography (OCT) have been used in dermatology. These include swept source (SS) OCT, spectral domain (SD) OCT, dynamic (D) OCT, line-field confocal (LC) OCT, full-field (FF) OCT, high-definition (HD) OCT, and optical coherence microscopy (OCM) [121]. Depending on the configuration, OCT generates 2D and 3D images from backscattered light from within the tissue [122,123,124,125,126,127,128]. OCT imaging is based on endogenous scatterers and can image more significant volumes of skin than RCM [18,73]. OCT has been used to image the stratum corneum of glabrous skin, the epidermis, papillary dermis, dermo–epidermal junction, blood vessels, sweat glands, and hair follicles [89,122,129]. SS-OCT has a deeper penetration depth but a lower resolution than OCM, HD-OCT, or LC-OCT. A typical implementation of SS-OCT images uses a field of view of 6.0 × 6.0 mm2, an axial resolution of 5–10 µm, a lateral resolution of 7.5–15 µm, and a penetration depth of up to 2 mm [18,124,130,131,132]. This places SS-OCT resolution between RCM and high-frequency ultrasound. SS-OCT provides architectural details within tissue with near-cellular clarity, superior to high-frequency ultrasound and with better penetration depth than RCM, yet SS-OCT resolution is not sufficient to distinguish individual cells, limiting its usefulness for distinguishing between pigmented benign and malignant lesions such as dysplastic nevi and melanoma [124,133]. SS-OCT cannot characterize melanoma from visually identifiable features [133]. The low penetration depth of under 2 mm prevents OCT from imaging the breadth of more advanced tumors. Further, SS-OCT cannot delineate cellular features and relies on distinct architectural pattern recognition; its usefulness is therefore limited in the diagnosis of cutaneous melanoma [89,134]. To overcome this limitation, Turani et al. proposed a computational method for the analysis of OCT images. One study reported a high sensitivity (97%) and specificity (98%) in differentiating melanoma from benign nevi [133]. Another research group utilized deep learning to train a computational kernel to differentiate melanoma and benign nevi in mice, achieving 98/99% specificity and sensitivity with their model [135].
D-OCT is based on speckle variance and can visualize skin microvasculature and detect blood vessels within specific lesions by detecting motion and blood flow—these images may enhance the diagnostic accuracy of melanoma [136]. HD-OCT has a resolution of 3 μm for both lateral and axial imaging, allowing for the visualization of individual cells [137]. The penetration depth is 0.5 to 1.0 mm and the field of view is 1.8 × 1.5 mm2. For example, HD-OCT has been used to measure melanoma tumor thickness for shallow tumors relatively accurately, within an average error of 0.08 mm compared with histologic measurement [138]. A 2014 study of 64 patients, evaluating 93 melanocytic lesions (27 melanomas), indicated a sensitivity of 74.1% and specificity of 92.4% for HD-OCT [137]. Gambichler et al. indicated the accuracy of a melanoma diagnosis with HD-OCT depends upon tumor thickness and the existence of other suspicious lesions, as thin melanomas had a high false-negative rate and dysplastic nevi had a high false-positive rate [137].
LC-OCT utilizes a supercontinuum fiber laser as a broadband spatially coherent light source, typically with a central wavelength of ~800 nm [139]. This configuration enhances resolution to ~1.5 μm but reduces penetration depth. However, it has been implemented with a three-dimensional cube that provides fully cellular resolution [139,140,141,142,143]. LC-OCT can visualize melanoma directly, including providing horizontal images similar to RCM in clarity [144]. OCM is an extension of OCT that integrates microscopy-level spatial resolution by combining OCT with high-numerical-aperture (NA) objectives [145,146]. OCM achieves high lateral resolution (1–2 μm) due to the use of microscope objectives: resolution can approach the cellular level. The penetration depth is 400–700 μm and the field of view is around 1 to 2 mm. OCM fused with pump–probe spectroscopy has demonstrated the ability to detect melanoma in a human skin sample [147]. Compared to SS-OCT, HD-OCT, LC-OCT, and OCM can differentiate between malignant and benign melanocytic lesions (nevi). One study of LC-OCT reported 93% sensitivity and 100% specificity [146]; studies are underway to validate their utility [148,149].
Each version of OCT has difficulty visualizing deeper structures as the lesion thickness increases, limiting the ability to image deep tumor invasion [149]. Further, OCT scanning devices are expensive, and the technology is considered experimental by most insurers and therefore not reimbursed. With the similarities in cost and design between RCM and OCT devices, there is hope that OCT will soon be reimbursed. Other configurations of OCT can be learned from more established disciplines that currently use OCT, including ophthalmology [121,150,151,152,153,154]. Conventional and LC-OCT images are shown in Figure 3f. There are several commercial OCT systems available, such as the SS-OCT system VivoSight (Michelson Diagnostics, Maidstone, UK), which is equipped with D-OCT capability, an HD-OCT system SkinTell (Agfa, Mortsel, Belgium), and the LC-OCT system DeepLive (Damae Medical Paris, France). Aquyre Biosciences has an FF-OCT system (CelTivity System) for histological applications.

4.6. Multispectral Imaging

Multispectral imaging (MSI) creates images of the epidermis and papillary dermis by using multispectral illumination to illuminate sub-surface pigmentation in lesions up to 2 mm thick [155,156]. MSI uses multiple wavelengths of visible and near-infrared light to illuminate a lesion. Chromophores in hemoglobin, melanin, and collagen absorb and the transmit energy that can be measured/imaged [157]. These images have been used to analyze melanoma, pigmented skin lesions, basal cell carcinoma, and skin color [156,158,159,160]. In 2012, Bekina et al. used MSI to analyze lesions at four different wavelengths: 450 nm to evaluate superficial layers, 545 nm to evaluate blood distribution, 660 nm to detect melanin, and 940 nm for deeper skin structures [161]. Figure 3g shows a comparative overview of the multispectral image analysis of a melanoma and a benign nevus [160].
In a 2011 multicenter, blinded study, Monheit et al. [162] investigated the effectiveness of an MSI system on 127 melanoma lesions (in vivo) and compared the results to an independent biopsy reader study performed by 39 dermatologists. They demonstrated a 98.4% sensitivity in comparison to the 78% sensitivity of the dermatologists. Additionally, on pigmented lesions biopsied to rule out melanoma, MSI demonstrated a 9.9% specificity in contrast to the 3.7% specificity by dermatologists. The study concluded MSI to be a safe and effective aid in the diagnosis of melanomas [162]. Similarly, a 2017 study by Fink et al. [163] analyzed the performance of MSI in the clinical setting. In the study, 360 pigmented skin lesions were observed by dermatologists using an MSI system called Melafind, which produces a score based on the probability of melanoma. Lesions with scores > 2 were considered suspicious of malignancy, but the decision to biopsy was made by the dermatologist. Of the 113 lesions biopsied, the sensitivity and specificity of MelaFind were 100% and 5.5%, respectively (68.5% specificity for the entire set of 360 lesions) [163]. Overall, both studies demonstrate the high sensitivity of MSI but low specificity. Although this imaging method is a useful tool to decide whether to biopsy or not, MSI does not have depth sectioning capability and as such depth information is limited; further, is it not designed to evaluate colorless amelanotic melanomas [155,157]. MSI is not indicated for use in the eyes, mucosal, subungual, palmar, or plantar (acral) anatomical areas [61].
MSI has also been extensively evaluated as a tool for improving pathology screening [164]. While it has demonstrated efficacy, frameworks for skin tissue classification and segmentation are not standardized and the method has not been widely adopted clinically. MelaFind (Strata Skin Sciences, Horsham, PA, USA) was one of the manufacturers of MSI which has discontinued the development and sales of its MSI product line, effective on 30 September 2017 [165].

4.7. Ultrasound and High-Frequency Ultrasound

Ultrasound transmits ultrasound waves into the tissue and uses the reflected sound wave to reconstruct an image of the internal structures. Conventional ultrasound (3.5–14 MHz) is used for measuring lymph nodes and classifying them as benign, suspicious, or malignant prior to fine needle aspiration (FNAC) biopsy [166]. Ultrasound analysis often relies on the Berlin morphological criteria to predict metastasis to sentinel nodes. The criteria include the presence of peripheral perfusion, loss of central echoes, and balloon-shaped lymph nodes [166]. Conventional ultrasound has greater depth penetration but less spatial resolution than ultrahigh and high-frequency ultrasound (HFUS) and is therefore not typically used for analyzing primary melanomas [167]. HFUS provides real-time, non-invasive, non-ionizing images of cross-section slices through the skin, in a similar orientation to histology or OCT [106,168]. Transducers of 20, 75, or 100 MHz have been developed, offering a resolution of 10–200 µm and a penetration depth of 1.5 to 10 mm, with higher frequencies having lower penetration depths, but higher resolution, up to histological resolution [169,170]. HFUS has been utilized to image benign and malignant tumors, inflammatory diseases, and nail and scalp entities [171,172]. At its highest resolution, HFUS can image individual layers of the epidermis and dermis, cutaneous appendages, blood vessels and blood flow characteristics (with the color Doppler capability), and the stage of the disease (proliferation or regression) [172,173], and can be used for evaluating primary melanoma lesions, satellite/in-transit metastasis, and lymph node metastasis [174]. Melanoma typically presents as hypoechoic and or heterogeneous oval-shaped, fusiform, and hyper-vascularized structures with sharp margins and infiltration to the dermis [168]. The Doppler imaging technique shows increased and anarchic vascularization [168,174].
Importantly, HFUS does not rely on melanin as a contrast agent and is useful for detecting amelanotic melanomas [175]. HFUS with color Doppler imaging was used to investigate intralesional vascularization in 107 pigmented lesions and provided 100% specificity (albeit 34% sensitivity) in distinguishing pigmented melanomas from non-melanoma lesions [176]. Beyond its use for tumor screening, HFUS’s potential to provide deep penetration is a distinct advantage over other imaging modalities such as RCM (<0.3 mm) and OCT (~1–2 mm) and is important in estimating tumor (Breslow) thickness [169,172,177,178,179]. In one study, HFUS’s capability to determine tumor thickness showed a 99.4% correlation with histologic analysis for superficial spreading melanoma and a 98.4% correlation for nodular melanoma [180]. A study of 25 patients with cutaneous melanoma by Kunte et al. used B-mode HFUS for the preoperative identification and characterization of sentinel lymph nodes (SLN) to correctly identify two of the six positive SLNs, with a sensitivity of 33.3% and specificity of 100%, concluding that HFUS cannot replace SLN biopsy in the detection of micrometastis [181]. In a separate study evaluating melanoma surveillance and regional lymph node involvement, HFUS achieved a specificity of 85% to 99% and sensitivity of 95% to 100% [182]. Ultrasound coupled with the Berlin US Morphology Criteria and combined with fine needle aspiration cytology can significantly improve sensitivity [183].
The main advantage of HFUS is the capability to visualize the skin layers, deep structures, and perfusion patterns in real-time, allowing for the pre- and post-operative assessment of melanoma with a 3D image. However, HFUS cannot provide information about cellular morphological features due to its low resolution [184]. Additionally, HFUS is unable to differentiate between melanoma and inflammatory infiltrates as they are both hypoechoic, potentially resulting in an overestimation of tumor thickness when compared to histological section [169,185], although more recently, greater accuracy has been obtained [178,186]. Moreover, HFUS cannot detect pigment differences such as melanin content, as there are no specific pathogenic ultrasound features that are capable of distinguishing melanoma from nevi [170,187]. Additional limitations include measuring lesions that are less than 0.1 mm in depth, the detection of melanin, and the detection of flat epidermal lesions [187]. Finally, HFUS requires an operator extensively trained in its use and interpretation [172,188]. Images of US-guided fine needle aspiration and HFUS are included in Figure 3h,i.
Longport Inc., (Episcan I-200, Chadds Ford, PA, USA), FujiFilm VisualSonics (VevoMD, Bothell, WA, USA), and Cortex Technology (Aalborg, Denmark) are manufacturers of HFUS systems, among others.

4.8. Magnetic Resonance Imaging

Magnetic resonance imaging (MRI) has traditionally been utilized to image internal body structures, but in the last 20 years, it has been adapted in dermatology due to the development of specialized surface coils that enable higher image resolution than standard MRI coils [110,189]. With the use of contrast agents, MRI can differentiate the epidermis, dermis, and subcutaneous layers [190]. MRI has been used to image melanoma metastasis in vivo [191,192,193], as well as to image large skin cancers and their surrounding anatomy [194,195]. The significant advantage of MRI over most other imaging modalities is the exquisite soft-tissue contrast. The major disadvantages, however, are the low sensitivity, long scanning time, and requirement of exogenous contrast agents [196]. Furthermore, MRI cannot reliably differentiate between malignant and benign tumors. Nonetheless, a study by Jouvet et al. showed MRI to have a sensitivity of 84% and specificity of 87.1% for staging of melanoma [197]. Moreover, whole-body MRI is the imaging modality of choice for the detection of intracerebral and other distant metastases [198,199]. MRI images of a metastatic bone lesion in the left ilium are shown in Figure 3j, along with PET and CT images (discussed below). Manufacturers of MRI with dermatologic surveillance capabilities include GE Healthcare (Waukesha, WI, USA), Philips (Amsterdam, The Netherlands), and Siemens Healthcare (Munich, Germany), among others.

4.9. Raman Spectroscopy

Raman spectroscopy uses a laser to irradiate samples, which results in light scattering that varies depending on the molecular vibrations of the proteins and lipids making up the tissue [200,201]. A plot of the intensity of Raman scattered radiation as a function of its frequency difference from the incident radiation is called a Raman spectrum or the Raman shift [202]. Through inelastic scattering, the molecules within the tissue absorb photons from the light source and re-emit photons at a lower frequency [203]. Raman spectroscopy is a non-destructive and non-invasive method to image melanoma either in vivo or in excised tissue [98,200,204,205,206,207,208]. In a study by Lui et al., 518 skin lesions from 453 patients were imaged in vivo using Raman spectroscopy and the method was shown to distinguish malignant from healthy skin with a sensitivity of 95–99%, and a specificity of 15–54% [207]. Another study reported the correct classification of all melanomas with a specificity of 43.8%, sensitivity of 100%, and a number needed to treat of 2.7 [209]. For melanoma detection, Raman spectroscopy can differentiate melanomas, pigmented nevi, basal cell carcinomas, seborrheic keratoses, and healthy skin with a sensitivity and specificity of 85% and 99%, respectively; however, its imaging depth is limited, up to a few hundred micrometers, depending on the spectrometer setup [200,207]. Coherent anti-Stokes Raman scattering (CARS) microscopy has been utilized to detect pheomelanin signals in human tissue with amelanotic melanoma [210]. Another advance is the coupling of Raman spectroscopy with deep learning for melanoma detection, particularly for biopsied tissue analysis [200,204,205,206,209] (Figure 3k).
The Verisante Aura (Verisante Inc., Richmond, BC, Canada) is a commercially available handheld Raman spectroscopy device [98].

4.10. Elastic Scattering Spectroscopy

Elastic light single-scattering spectroscopy (ESS) is a technique that directs white light through a probe and measured the intensity at different spectra of the reflected light [211,212]. In this way, it operates similarly to Raman scattering spectroscopy. There is an FDA-cleared device by DermaSensor that includes an algorithm with an AI core, trained using over 11,000 spectral scans from 3500 skin lesions [213]. A multicenter trial was recently performed on the device involving 311 participants, including 44 melanomas. The device achieved 95.5% sensitivity and 32.5% specificity in a high-risk population with lesions selected for biopsy [214]. The technique has also been tested in a rodent model for its ability to monitor cancerous tissue response to laser therapy [215].

4.11. PET/CT and SPECT/CT

Positron emission tomography/computed tomography (PET/CT) is a whole-body imaging technique commonly used in the diagnosis of metastatic cancer. It relies on the introduction of a tracer, most commonly 18F-flouro-deoxy-glucose (18F-FDG), into the body. This tracer is a glucose analog with a positron-emitting radioisotope fluorine-18. Due to the increased glucose uptake of cancer cells, the tracer is visible more rapidly in malignant tissue than in healthy tissue [216]. Figure 3l shows a malignant lymph node detected by PET/CT. PET lacks optimal resolution to visualize early-stage cutaneous melanoma, and thus its use is limited to advanced metastatic melanoma and for staging clinically apparent nodal or distant metastasis (Figure 3j). PET has been tested for its ability for staging melanoma and to influence the treatment plan in melanoma patients with satellite or in-transit metastases, but showed only moderate sensitivity and specificity for this purpose [217]. Another application is for predicting or monitoring the therapeutic response in patients with metastatic melanoma [218,219]. 18F-FDG PET/CT parameters are promising predictors of the final response of metastatic melanoma patients to immunotherapy. In one study, 57 patients with metastatic melanoma were treated with ipilimumab or PD-1 inhibitors and received 18F-FDG PET/CT scans before treatment and 12–18 weeks later. The percent change in metabolic tumor volume and total lesion glycolysis were assessed [220]. The accuracy of parameters for therapy response were 96% in group 1 and 97% in group 2.
While PET/CT utilizes positron-emitting tracers and detects gamma rays produced by positron–electron interactions, single photon emission computed tomography/computed tomography (SPECT/CT) uses single photon-emitting tracers (detects gamma rays directly emitted by the tracers). While PET/CT is commonly used for diagnosing metastatic cancer, the most widespread use of SPECT/CT is to improve the accuracy of sentinel node biopsies for staging melanoma [221,222,223]. In a recent analysis of 1522 primary cutaneous melanoma patients, SPECT/CT was able to detect 50% more sentinel nodes than planar lymphoscintigraphy, which translated to a significantly reduced risk of death from melanoma. These findings have been replicated by others [224,225]. Figure 3m shows detection of SLNs by SPECT/CT.
Manufacturers of PET/CT and SPECT/CT systems include GE Healthcare (Waukesha, WI, USA), Philips (Amsterdam, The Netherlands), and Siemens Healthcare (Munich, Germany).

4.12. Lymphoscintigraphy

Lymphoscintigraphy for sentinel lymph node detection involves injecting a radiotracer, typically technetium-99m labeled colloids, into tissue near the identified melanoma. Then, the skin is massaged to accelerate the distribution of the tracer toward lymph nodes to identify a sentinel lymph node. The region is imaged using a gamma camera to detect the radiation emitted by the tracer. This process, which has been used in clinical practice for over 30 years, is often performed before sentinel lymph node biopsy [226,227]. More recently, a dye or another tracer is injected subsequently and combined with a handheld gamma camera for intraoperative monitoring during SLN biopsy. A typical image is shown in Figure 3n. There has been a concern regarding the possibility of false-negative sentinel node biopsies, and the quality of information from lymphoscintigraphy can vary from center to center [228]. Gamma cameras are available from many manufacturers, including Siemens Healthineers, GE Healthcare, and Philips Healthcare.

4.13. Non-Imaging (Genetic) Melanoma Detection and Disease Management

Genetic information from either skin cells or plasma are being employed to assist in melanoma detection and disease management. Epidermal genetic information retrieval (EGIR) is a melanoma diagnosis technology in which cells from the surface of pigmented melanocytic lesions are obtained via an adhesive patch and then sent to a lab for analysis [229,230]. A genomic signature is derived from the extracted mRNA, specifically analyzing the expression of two genes known to increase in melanoma: PRAME (preferentially expressed antigen in melanoma) and LINC (LINC00518, long intergenic noncoding RNA 518) [231]. EGIR is a diagnostic test intended to guide the decision on whether to biopsy or not and is especially useful for patients with multiple nevi or those that have a suspicious lesion on a cosmetically sensitive area [232,233]. Post biopsy, the gene expression profile (GEP) test is a prognostic test that utilizes reverse transcription polymerase chain reaction (RT-PCR) in primary tissue to determine the expression of thirty-one genes (three control) to predict metastatic risk [234,235]. For metastatic melanoma, liquid biopsies contain information on circulating melanoma cells (CMCs) and circulating tumor DNA (ctDNA). CMCs can be analyzed through the FDA-cleared Cell-Search platform, and ctDNA can be profiled using next generation sequencing (NGS) droplet digital PCR and customized genetic analysis [236]. Given the diversity of presentations of melanoma, these genetic assays can assist in disease management but have not yet delivered definitive information in most cases.
A 2017 validation study using the EGIR test examined 398 biopsied lesions, concluding the two-gene signature identified melanomas with a 91% sensitivity and 69% specificity [233]. Another study indicated that molecular expression profiling was able to differentiate melanoma from nevi with a sensitivity of 97.6% and specificity of 72.7%, suggesting this may be a useful method to reduce unnecessary biopsies [237]. In a cohort study of 205 stage I and II melanomas by Ferris et al., the post-biopsy GEP was able to predict distant metastasis with a sensitivity of 79% and specificity of 68% [235]. However, a 2023 panel from the Melanoma Prevention Working Group did not reach consensus on the role of gene expression profile testing for melanoma screening [238].
Genomic testing offers a method to possibly forego a skin biopsy, or SLNB, and can improve accuracy when diagnosing melanoma and determining the likelihood of metastasis [233,239]. Genomic testing may also aid in targeted therapy for stage III melanoma and higher, as cases displaying the BRAF-V600 mutation have proved to be somewhat treatable [240], especially through combination therapies [241]. However, the use of liquid biopsies for monitoring therapy response, cancer progression, and the onset of resistance is still experimental [236,242]. One of the companies that produces EGIR is Dermtech, Inc. (La Jolla, CA, USA). Two molecular tests for melanoma based on gene expression have been developed that acquire RNA from formalin-fixed paraffin-embedded sections rather than in situ lesions. GEP is offered by Castle Biosciences (Friendswood, TX, USA) and Myriad Genetics (Salt Lake City, UT, USA) offers myPath to distinguish nevi from melanoma [229]. CellSearch (Menarini Silicon Biosystems, Bologna, Italy) is the only FDA-approved system for detecting CMCs and several ctDNA based diagnostic assays have been FDA-approved, including Signatera (Natera, Austin, TX, USA).
Figure 3. Images from technologies currently used in medical practice. (a) Smart phone image analysis, courtesy of [59]. (b) Total body digital photography (iiii), courtesy of [58]. (c) Dermoscopy of lentigo maligna (ivi), courtesy of [74]. (d) Electrical impedance spectroscopy protocol. Image courtesy of [243]. (e) Superficial spreading melanoma in situ shown by (i) dermoscopic image, (ii) RCM mosaic image, and (iii) RCM individual image. White arrows show biomarkers of melanoma, courtesy of [109]. (f) OCT images of melanoma: (i) vertical image by conventional OCT, (ii) verticle image of melanoma by LC-OCT, and (iii) horizontal image showing irregular honeycomb pattern with atypical melanocytic cells, courtesy of [144]. (g) Hyperspectral imaging of a melanoma (i-iv) and a benign nevus (vviii). Left to right: photographic image for lesion localization in RGB, raw mosaic image, image after preprocessing, and classification result (red, malignant melanoma), courtesy of [160]. (h) Fine needle biopsy under ultrasound guidance, courtesy of [244]. (i) Measurement of primary cutaneous melanoma by HFUS (i,ii), courtesy of [174]. (j) Comparison of (i) CT (lesion is not evident), (ii) MRI (rapid acquisition sequency) -lesion is evident, (iii) PET imaging of metastatic bone lesion is evident (iv) MRI (diffuse weighted imaging sequence) -lesion is evident, courtesy of [193]. (k) Raman spectroscopy (i,ii) of biopsied tissue generates (iii) Raman shift spectra. Image courtesy of [209]. (l) Malignant lymph node detected by PET/CT. Coronal (i), transaxial (ii), and PET/CT fusion (iii) images of foot after ingesting tracer. Image courtesy of [245]. (m) SPECT/CT image showing injection site (shoulder) and two sentinel lymph nodes. Static images acquired after 10 min (i) and 2 h (ii) post injection and (iii)volume-rendered SPECT-CT, courtesy of [223]. (n) Radiography images of tracer dye movement to identify sentinel lymph nodes imaged shown from calf to popliteal (i) and groin (ii) 10 min and one hour (iii,iv) after injection of tracer dye, courtesy of [226]. Key biomarkers associated with each modality are described in Table 1.
Figure 3. Images from technologies currently used in medical practice. (a) Smart phone image analysis, courtesy of [59]. (b) Total body digital photography (iiii), courtesy of [58]. (c) Dermoscopy of lentigo maligna (ivi), courtesy of [74]. (d) Electrical impedance spectroscopy protocol. Image courtesy of [243]. (e) Superficial spreading melanoma in situ shown by (i) dermoscopic image, (ii) RCM mosaic image, and (iii) RCM individual image. White arrows show biomarkers of melanoma, courtesy of [109]. (f) OCT images of melanoma: (i) vertical image by conventional OCT, (ii) verticle image of melanoma by LC-OCT, and (iii) horizontal image showing irregular honeycomb pattern with atypical melanocytic cells, courtesy of [144]. (g) Hyperspectral imaging of a melanoma (i-iv) and a benign nevus (vviii). Left to right: photographic image for lesion localization in RGB, raw mosaic image, image after preprocessing, and classification result (red, malignant melanoma), courtesy of [160]. (h) Fine needle biopsy under ultrasound guidance, courtesy of [244]. (i) Measurement of primary cutaneous melanoma by HFUS (i,ii), courtesy of [174]. (j) Comparison of (i) CT (lesion is not evident), (ii) MRI (rapid acquisition sequency) -lesion is evident, (iii) PET imaging of metastatic bone lesion is evident (iv) MRI (diffuse weighted imaging sequence) -lesion is evident, courtesy of [193]. (k) Raman spectroscopy (i,ii) of biopsied tissue generates (iii) Raman shift spectra. Image courtesy of [209]. (l) Malignant lymph node detected by PET/CT. Coronal (i), transaxial (ii), and PET/CT fusion (iii) images of foot after ingesting tracer. Image courtesy of [245]. (m) SPECT/CT image showing injection site (shoulder) and two sentinel lymph nodes. Static images acquired after 10 min (i) and 2 h (ii) post injection and (iii)volume-rendered SPECT-CT, courtesy of [223]. (n) Radiography images of tracer dye movement to identify sentinel lymph nodes imaged shown from calf to popliteal (i) and groin (ii) 10 min and one hour (iii,iv) after injection of tracer dye, courtesy of [226]. Key biomarkers associated with each modality are described in Table 1.
Biosensors 15 00297 g003

4.14. The Role of Artificial Intelligence in Melanoma Technologies

AI has distinct and complementary applications for improving melanoma management which can be grouped as follows: image enhancement, image interpretation, and multimodal data integration. Image enhancement can be accomplished through AI-based super-resolution methods that use deep learning models to intelligently add details, textures, and sharper edges while minimizing noise [246,247]. These methods can improve resolution or reduce the time needed to acquire high-resolution images [248]. Image enhancement also has the potential to overcome difficulties in inter-operator variability [249]. AI can compensate for different conditions including lighting, steadiness of the hand of the operator, angle of probe used, and distance between imaging probe and skin. Enhancements could be accomplished by advances in signal and image processing, leading to standardization after relatively minimal training has been accomplished. However, such compensation models are in their infancy and would require a significant commitment from the research community. AI-based automatic image interpretation can boost melanoma detection, screening, and monitoring through methods such as pattern recognition, classification, object detection, segmentation, textural analysis, and feature extraction [250]. Pattern recognition and classification using methods such as convolutional neural networks [251] and transformers [252] are most effective for modalities with very large datasets, such as dermoscopy [253,254,255] and histology [256]. A shortcoming of the AI-based analysis of dermoscopy images that applies to rarer forms of melanoma and for melanoma in persons of color (where melanoma is far less prevalent) is that most studies have access to an insufficient number of training images for accurate model-building in these conditions [257]. Newer or emerging technologies generally implement methods that are leaner and can train from fewer samples, such as feature extraction [135] and textural analysis [258,259]. Finally, multimodal data integration techniques such as feature-level fusion, decision-level fusion, mixture of experts, and attention-based methods can be employed to incorporate dermoscopy, histopathology, ultrasound, or other imaging modalities together and/or to combine their results with genetic, molecular, and electronic health record data [260,261,262]. For example, one combination of modalities involves the use of RGB (dermoscopy) images (such as dermoscopy images) and extended near infrared images. A recent article demonstrated that training deep learning algorithms using both sets of images improved skin cancer diagnosis over RGB images alone [263]. Others have used AI to extract HSI information from RGB images, improving melanoma detection accuracy [264,265].

5. Technologies in Development for Non-Invasive Imaging of Cutaneous Melanoma

5.1. Photoacoustic Imaging (Optoacoustic Imaging)

In photoacoustic imaging (PAI), also known as optoacoustic imaging, the thermoelastic expansion of tissue chromophores occurs when they are irradiated by a nanosecond pulsed laser—resulting in the emission of acoustic waves that are detected by ultrasound transducers and translated into an image by a reconstruction algorithm [107]. To increase the specificity of PAI, disease-specific biomarkers have been utilized with endogenous absorbers, such as myoglobin, melanin, water, DNA, RNA, lipid, carboxyhemoglobin, bilirubin, cytochrome C, and methemoglobin [266]. Additionally, PAI has been used in conjunction with exogenous contrast agents such as nanoparticles and dyes [267,268,269,270,271,272]. With the use of a near-infrared light source, one can increase the penetration depth since the light absorption and scattering properties of hemoglobin and deep tissues are low at these wavelengths [271,273,274]. Because melanin is an endogenous contrast agent, PAI has great potential for melanoma assessment [34,268,275,276,277,278,279,280,281,282,283]. In fact, PAI has been used for detection and depth measurement [284], lymph node metastasis detection (both in vivo [285,286] and ex vivo [287,288]), tumor angiogenesis for monitoring the spread of metastasis and treatment efficacy [289,290], the detection of circulating tumor cells either in vivo [291,292] or by using photoacoustic flow cytometry [293,294], and virtual histology, using a technique called photoacoustic remote sensing [295,296]. PAI has been utilized to detect blood vessels within the skin as well as circulating tumor cells (CTCs) in the blood and for locating sentinel lymph nodes for biopsy. A PAT cross-sectional image of a melanoma is shown in Figure 4a.
PAI has been developed in both tomography (larger penetration depth with coarser resolution) and microscopy (shallow penetration depth, higher resolution) configurations [297,298]. While tomography is often employed for in vivo applications, recently, handheld photoacoustic microscopy devices have been developed that measure tumor thickness in patients with suspected melanoma lesions up to 7 mm [299,300]. In a study involving 27 patients with pigmented cutaneous lesions, the measured lesion thickness gave a high correlation to the resected surgical samples (r = 0.99, p < 0.001 for melanomas, r = 0.98, p < 0.001 for nevi), showing linear-array PAI is reliable in measuring the thickness of cutaneous lesions in vivo [299]. The exquisite ability of PAI to detect endogenous melanin with high accuracy is the reason PAI has been tested for so many melanoma-related applications.
PAI typically utilizes a transponder in contact with the medium to record and analyze acoustic waves. Non-interferometric PA remote sensing (PARS) microscopy analyzes the modulation of the elasto-optical refractive index caused by photoacoustic transients, when the absorbing interface has an appreciable refractive index difference from baseline [301]. This enables non-contact photoacoustic microscopy [301,302,303]. A non-interferometric approach with a low-coherence probe beam can be used to detect intensity variations unaffected by phase modulations. The system’s use has been demonstrated on a chicken model of melanoma [301]. PARS microscopy has also been applied on histology slides to enable the direct visualization of subcellular morphology on unprocessed, excised tissue, obviating the need for tissue freezing and histochemical staining [296]. A more in-depth discussion of PAI for melanoma detection and assessment can be found in Fakhoury et al. [34].
The main challenges of using melanin as the endogenous marker for melanoma (for detection and depth measurement, SLN analysis, and CTC assays) is that the use of such imaging modalities requires additional processing steps to differentiate melanoma from benign nevi (as both can have significant melanin), and they have no effective method to detect amelanotic melanoma. Additionally, PAI does not have adequate resolution to visualize cellular structures. One commercially available PA device is the Acuity Echo (iThera Medical, Munich, Germany); no commercially available devices have been developed for PARS.

5.2. Hyperspectral Imaging

Hyperspectral imaging (HSI) combines optical spectroscopy with optical imaging to analyze and record a larger optical spectrum for every pixel in the field of view (Figure 4b) [304]. Each pixel of the image contains spectral information, which is added to the two-dimensional spatial image, generating a three-dimensional data volume [305]. HSI identifies spectral characteristics, defined as the relationship between the wavelength and the physical properties of an object. This system typically uses a light source and a spectrometer, which records the scattered photons from a tissue’s surface up to 2 mm deep and presents an image. The advantage of HSI over MSI systems is that it can distinguish hemoglobin from melanin, providing more accurate information in pigmented lesions and darker skin types [306,307]. A recent study analyzed 325 lesions from 285 patients. Lesions were imaged prior to excision and a deep neural network algorithm was trained to distinguish between histopathologically verified malignant and benign lesions, to classify subgroups, and to delineate the extent of tumor [308]. Using the “majority vote” classification method, a sensitivity of 95% and specificity of 92% were achieved in differentiating malignant from benign lesions. HSI has also been used ex vivo to analyze pathological sections. Combining spectral features in the 500–675 nm band, a classification procedure achieved 98% accuracy [309]. HSI is most often used in combination with machine learning/deep learning techniques [310]. Another implementation of HSI does not require a spectrometer, but instead has been implemented by extracting color bands from calibrated RGB images via snapshot hyperspectral conversion [311] or by a CNN-based approach [265]. This technique shows significantly improved accuracy over the analysis of the RGB images alone at detecting melanoma and differentiating melanoma subtypes. Using the snapshot conversion method, one model, implemented with machine-learning analysis, detected nodular melanomas with a 0.9 sensitivity and 0.851 precision [311].
A commercially available SkinSpect™ dermatoscope (Spectral Molecular Imaging Inc., 201 N. Robertson Blvd, Beverly Hills, CA, USA) has been introduced that combines polarization and HSI to map the distribution of skin melanin and hemoglobin [307]. Another HSI system is the GaiaMicro-G-V10E-AZ4 (Dualix Spectral Imaging, Wuxi, Jiangsu, China), which is only approved for research. Any photographic system could be adapted for HSI using one of the methods described above, but based on their wide spectral acquisition set, smartphones have also been proposed as instruments for HSI-based melanoma detection [312].

5.3. Quantitative Dynamic Infrared Imaging (Thermographic Imaging)

Under certain conditions, the thermal radiation emitted by the skin in the infrared domain can be measured [62,132,313,314]. Due to increased blood supply, cancerous lesions, including melanoma, are warmer than healthy skin, and this difference in temperature can be detected by thermographic imaging [62,132,315]. Infrared imaging can image large surface areas and multiple lesions. Infrared thermography alone cannot distinguish between different types of cutaneous cancers; the purpose of this technology is to identify potential abnormalities. While passive thermography can provide only qualitative results, by adding a cooling system, the quantitative analysis of the efficiency with which an object radiates thermal energy (emissivity) can be determined (Figure 4c). Quantitative dynamic infrared imaging can possibly stage melanomas based on temperature differences and other thermal characteristics [62,316].
In a study of 74 patients, 251 palpable lesions were imaged with infrared thermography to determine if they were melanoma or non-melanoma, with a sensitivity ranging from 39 to 95% and specificity from 89 to 100%. Larger melanoma lesions (<15 mm) were more accurately detected via infrared thermography as hyperthermic [317]. Another study of 140 patients to detect skin cancer using quantitative dynamic infrared imaging demonstrated >99% sensitivity and >90% specificity (fifty-eight subjects were diagnosed with a cutaneous malignancy via biopsy, six of which had melanoma) [318]. A further refinement analyzed temperature recovery curves of suspicious lesions, achieving a sensitivity of 98% and specificity of 95% on a dataset of the same 116 subjects [319]. One of the manufacturers of infrared cameras used for thermography is QmagiQ, LLC (Nashua, NH, USA). Differences in skin temperature between cancerous and non-cancerous lesions can also be detected without the use of infrared cameras. One study showed that absolute skin surface temperatures and thermal conductivity (measured with a pen-shaped, guard-heated thermistor probe) differ between invasive melanoma vs. healthy skin or melanoma in situ [320].

5.4. Terahertz Pulsed Imaging

Terahertz pulsed imaging (TPI) is a non-invasive optical imaging modality using terahertz (THz) radiation [321,322]. The THz frequency range is 0.1 THz to 10 THz, corresponding to a wavelength range of 3 mm to 30 µm [323]. The frequency excites molecules, leading to vibrations which provide spectroscopic feedback and imaging contrast. THz radiation is highly absorbed by liquid water, which allows for imaging based on differences in water content between pathologic and non-pathologic skin lesions. The potential for imaging is limited by the water absorption coefficient of 80–350 cm−1 at 0.1–2.0 THz to a depth of 0.2–0.3 mm [322]. Sim et al. utilized TPI for the diagnosis of oral melanoma ex vivo in a frozen section. This research found significantly less water content in the melanoma versus healthy skin, allowing for differentiation between benign and malignant mucosal cells. Due to the increased protein and cell density in melanoma, the absorption coefficient and refractive index of the tumor are higher in diseased tissue than in healthy tissue [323]. TPI has been used for the in vivo imaging of dysplastic and non-dysplastic nevi, as it is effective at defining rough tumor margins (Figure 4d); however, penetration is better with frozen sections [322,324,325]. Promising terahertz imaging techniques include THz pulse imaging, continuous-wave terahertz, THz time-domain spectroscopy [326], and coherent THz confocal imaging [327]. The Terasense Group, Inc. (San Jose, CA, USA) produces Terahertz imaging devices.

5.5. Multiphoton Imaging

Multiphoton imaging is a spectrum of modalities that includes two-photon excitation microscopy (2PE), multiphoton microscopy (MPM), and second harmonic generation microscopy (SHG) [328,329]. Two-photon excitation signals are induced from endogenous fluorophores such as nicotinamide adenine dinucleotide (NADH), melanin, elastin, porphyrins, and collagen using two near-infrared lasers: the returning signal is captured by multichannel detectors and allows for subcellular resolution (0.5 µm lateral and 1–2 µm axial) [329,330,331] and results in minimal scattering. The multiphoton absorption also helps to suppress the background signal. This modality can be used to image tissue at a depth of up to 200 µm [330]. In SHG microscopy, the laser source generates a second harmonic signal when a laser passes through certain organic materials. The generated signal is twice the frequency (half the wavelength) of the incident laser and can be picked up by a detector. The signal varies depending on the characteristics of the tissue specimen. Many biologic components, including collagen, myosin, and microtubules, generate strong SHG signals [332]. MPM and SHG images of benign nevi and melanoma at different depths are shown in Figure 4e. A study conducted by Dimitrow et al. using multiphoton laser tomography in vivo to investigate eighty-three melanocytic skin lesions revealed distinct morphological differences in melanoma compared with melanocytic nevi. The study showed sensitivity values ranging from 71 to 95% and specificity values ranging from 69 to 97% [329]. Elagin et al. combined 2PE with optical coherence angiography to study equivocal melanocytic lesions [333]. They utilized a multiphoton tomograph, the MPTflex (JenLab, Berlin, Germany), and evaluated malignancy features from the images. They also acquired images by optical coherence angiography (OCA) to identify microvascular networks. Two-photon excitation imaging was able to discriminate benign lesions from melanoma in situ and invasive melanomas but could not differentiate between in situ and invasive melanomas. However, OCA detected significant differences in the vascular networks of melanoma in situ compared with invasive melanomas. Combining results from both modalities by a discriminant function analysis enabled perfect separation on their small dataset (49 lesions).
Chernyavskiy et al. used 2PE and SHG in addition to 1PE fluorescence and 1PE reflectance to image melanoma in mice. The resulting images revealed the collagen network, which is useful to identify invasive tumor cells. The authors concluded that these techniques are suitable to evaluate tissue modifications secondary to clinical interventions [332]. Multiphoton microscopy detects NADH without contrast to provide information about the mitochondrial organization within skin cells. The fluorescent patterns differ between normal skin and skin cancer: this has been used to differentiate skin cancer from healthy skin [334]. SHG is currently most widely used for non-invasive assessments of mouse melanoma growth [335]. A recent study showed the use of 2PE with an appropriate photosensitizer that targets melanomas for photodynamic therapy and in vivo melanoma ablation, which was tested in a mouse model [336].
Limitations of this imaging modality include laser damage for melanin rich lesions, slow speed of imaging, and small field of view. Manufacturers of multiphoton imaging devices with dermatologic capabilities include Olympus Corporation (Tokyo, Japan), Sutter Instrument (Novato, CA USA), Thorlabs Inc. (Newton, NJ, USA), and MPTFlex (Jena, Germany), among others.

5.6. Fiber Diffraction

Fiber diffraction for the detection of melanoma uses X-ray diffraction to acquire structural information from the fibrous content of the lesion (Figure 4f). Because biological fibers are composed of elongated molecules that are naturally aligned in parallel, identifying changes in fiber diffraction patterns of skin can indicate melanoma [337]. Interestingly, fiber diffraction can uncover physical changes in the biopsied tissue of fingernails and skin of cancer patients [337,338,339]. Each cancer type has a specific ring pattern, and the presence of a particular pattern indicates there is cancer somewhere on the body [338]. However, these ring changes only indicate the presence of carcinoma, and the actual location must then be determined. A study conducted by James and Kirby on 296 blinded skin samples (including 52 from melanoma patients) in subjects aged from 18 to 90 indicated that the fiber diffraction of skin could be a diagnostic test for melanoma even if other types of cancers are present [337]. Encouragingly, fiber diffraction’s potential for any-type cancer detection was reinforced in a small study of 30 samples, which showed the presence of prostate cancer with 100% sensitivity and 99.2% specificity [339]. With further development, fiber diffraction could be a low-cost, minimally invasive method to detect melanoma using small-angle X-ray beamlines at synchrotrons or a rotating anode X-ray generator [337,338]. However, this technique has not reported recent developments, suggesting translational or other difficulties.

5.7. Fourier Transform Infrared (FTIR) Spectroscopy and Microspectroscopy

Fourier transform infrared (FTIR) spectroscopy starts with a source with a wide spectrum of IR light, which passes through an interferometer before reaching the sample, and then measures how much of the beam is absorbed by the sample. The interferometer sequentially blocks and transmits different wavelengths to acquire multiple absorption spectra, and the data from the different acquisitions are combined and processed to determine the specific absorption across the IR spectrum, typically 4000–500 cm−1 or 780 nm–1 mm. This spectral range is able to extract information on intramolecular bonds in organic molecules, which enables the extraction of unique molecular fingerprints of biochemical information from fluids or tissue samples. FTIR is most often performed ex vivo using one of three sampling modes: transmission, transflection, and attenuated total reflection. The choice of sampling mode depends on the tissue or fluid being evaluated. FTIR can be used to construct images of tissue or cell architecture [340,341]. Because cancerous tissue has a different vibrational mode compared to healthy tissue, FTIR spectroscopy has been shown to rapidly detect melanoma in tissue samples [341]. Interestingly, FTIR can detect metabolic discharges from cancerous tissue into body fluids (blood, saliva, etc.), this has been used to provide guidance for clinical assessment and may aid in the rapid detection of melanoma [341,342]. FTIR can be combined with a microscope device to provide spatially resolved information on biochemical content (Figure 4g). This is often referred to as FTIR microspectroscopy [343]. Wald et al. used FTIR spectroscopy to image 34 melanoma biopsies from sentinel lymph nodes and was able to recognize melanoma cells with 87.1% sensitivity and 85.7% specificity [344]. Tosi et al. analyzed paraffin-embedded skin samples of benign nevi and melanoma and was able to differentiate them through principle component analysis [343]. FTIR microspectroscopy has a modest lateral resolution (~20 μm). Recently, FTIR is more often used to measure changes in melanoma tissues treated with anti-tumor molecules [345]. For example, the metastatic potential of different cell lines from the same genetic material could be differentiated by attenuated total reflection FTIR [346]. Manufacturers of FTIR spectroscopy systems include Thermo Fisher Scientific (Waltham, MA, USA), among others.

5.8. Real-Time Elastography

Real-time elastography (RTE) is an ultrasound imaging technique that provides images from the relative elasticity or stiffness of a lesion in comparison to adjacent healthy tissue [347,348]. Rigid tissue has less deformation than elastic tissue, and malignancy is indicated by a predominance of stiff tissue (Figure 4h) [348], which has been suggested for use for identifying tumor margins [348] and for differentiating between reactive and metastatic lymph nodes [349]. There are two diagnostic methods: elasticity score (ES) and strain ratio (SR). The limitation of RTE is that the technique is time-consuming and labor-intensive [350]. At a frequency of 14 MHz, RTE can penetrate up to 40 mm, which will allow the epidermis, dermis, and subcutaneous tissues to be imaged [351]. A metanalysis by Ying et al., which analyzed 835 superficial lymph nodes (LNs) for the diagnosis of malignant LNs, had an ES sensitivity of 74% and specificity of 90%, and an SR sensitivity of 88% and specificity of 81% [350]. In a study by Ogata et al., lymph nodes of 12 patients with cutaneous melanoma were imaged with RTE resulted in 100% sensitivity and 71% specificity when detecting metastasis with an ES cutoff score of three [349]. Hitachi, Ltd. (Tokyo, Japan) is one of the companies that produces a commercially available ultrasound platform with real-time elastography.

5.9. Electron Paramagnetic Resonance Imaging

Electron paramagnetic resonance (EPR) spectroscopy or electron spin resonance spectroscopy is a method for studying materials that have unpaired electrons (free radicals and paramagnetic transition metal ions). The basic concepts are analogous to NMR, but instead of measuring nuclear transitions in the nuclei, the spins excited are of the electrons. Melanin is one of very few organic compounds found in nature that contains unpaired electrons [352]; melanomas have strong EPR signals [353]. This has led to the suggestion that clinical low frequency EPR spectroscopy could be used as an in vivo technique for melanoma detection [354,355]. An EPR image of a melanoma is shown in Figure 4i. A first clinical trial of healthy volunteers and patients suspected of melanoma was performed, with subjects analyzed using a whole-body EPR scanner. The system was not sensitive enough to measure melanin differences in skin pigmentation, but was able to show the melanin signal was significantly higher in melanoma lesions than in benign nevi (p < 0.0001) [356]. However, this did not translate to high specificity. In addition, the system was not sensitive enough to detect Breslow depth with confidence. A shortcoming of this technique is it cannot see amelanotic melanomas. The system used was a whole-body clinical EPR system (Clin-EPR LLC, Lyme, NH, USA), operating at 1.15 GHz.

5.10. Multimodal Screening Technologies

The combination of two or more imaging systems into a single device can improve screening accuracy by leveraging the capabilities of both modalities. For example, there are reports of devices in which a Raman probe is incorporated onto a spectral domain OCT system [357,358], or a trimodal system including OCT, Raman, and PAI [359], or an ultrasound, OCT, and Raman combined system [360]. There is also a four-modal system combining OCT, Raman, ultrasound, and PAI [361]. In these cases, OCT and ultrasound are used for structural imaging (including tumor depth measurement), and PAI and Raman techniques are used for acquiring functional information [362,363,364,365]. We look forward to publications reporting observational trials of such multimodal systems to fully assess their capabilities compared with single-modal systems.
Figure 4. Images from technologies in development for cutaneous melanoma. (a) Photoacoustic imaging of in situ melanoma using linear array-based PAT, courtesy of [299]. (b) Hyperspectral imaging output for melanoma screening, courtesy of [304]. (c) Example of quantitative dynamic infrared imaging (i,ii). At steady state (iii), all images show similar heating, but at 2 s post recovery (iv), the melanoma has already cooled (iv,v), courtesy of [132]. (d) Terahertz pulse spectroscopy image of melanoma from mouse skin. Visual image (i), THz images (ii,iii) where red area corresponds to the normal region and yellow area corresponds to high THZ power-loss (indicative of melanoma), courtesy of [326]. (e) Multiphoton microscopy at different depths (rows) of junctional nevus (iiv), compound nevus (vviii), and melanoma (ixxii). Cellular autofluorescence excited by MPM in red and collagen (from second harmonic generation signal) in green. Bottom row shows histology, courtesy of [333]. (f) Fiber diffraction of skin samples from (i) healthy control and (ii) patient with melanoma. M ring indicates additional ring not seen in healthy control, courtesy of [337]. (g) FTIR analysis of skin tissue compared with H&E staining (i) of insert, selected ROI for FTIR imaging (ii), FTIR reconstructed image using 9 classes of wavelengths reveal histological features (iii), courtesy of [345] (h) Real-time elastography shows stiff skin lesion (iiiv), which on biopsy (i) was found to be melanoma, courtesy of [348]. (i) EPR analysis (i) Image of melanoma shown, cut in three, (ii) samples measured by EPR and images joined together. Boxes show highly pigmented, moderately pigmented, and nonpigmented areas (left to right), courtesy of [355].
Figure 4. Images from technologies in development for cutaneous melanoma. (a) Photoacoustic imaging of in situ melanoma using linear array-based PAT, courtesy of [299]. (b) Hyperspectral imaging output for melanoma screening, courtesy of [304]. (c) Example of quantitative dynamic infrared imaging (i,ii). At steady state (iii), all images show similar heating, but at 2 s post recovery (iv), the melanoma has already cooled (iv,v), courtesy of [132]. (d) Terahertz pulse spectroscopy image of melanoma from mouse skin. Visual image (i), THz images (ii,iii) where red area corresponds to the normal region and yellow area corresponds to high THZ power-loss (indicative of melanoma), courtesy of [326]. (e) Multiphoton microscopy at different depths (rows) of junctional nevus (iiv), compound nevus (vviii), and melanoma (ixxii). Cellular autofluorescence excited by MPM in red and collagen (from second harmonic generation signal) in green. Bottom row shows histology, courtesy of [333]. (f) Fiber diffraction of skin samples from (i) healthy control and (ii) patient with melanoma. M ring indicates additional ring not seen in healthy control, courtesy of [337]. (g) FTIR analysis of skin tissue compared with H&E staining (i) of insert, selected ROI for FTIR imaging (ii), FTIR reconstructed image using 9 classes of wavelengths reveal histological features (iii), courtesy of [345] (h) Real-time elastography shows stiff skin lesion (iiiv), which on biopsy (i) was found to be melanoma, courtesy of [348]. (i) EPR analysis (i) Image of melanoma shown, cut in three, (ii) samples measured by EPR and images joined together. Boxes show highly pigmented, moderately pigmented, and nonpigmented areas (left to right), courtesy of [355].
Biosensors 15 00297 g004

6. Remarks on Barriers to Technological Adoption

Barriers exist to introducing new screening technology to the dermatology clinic. So far, no melanoma screening technology has broken through and been embraced by the dermatology community. To accomplish this, new technology would need to achieve an appropriate sensitivity and specificity. Other practical considerations include whether the new system would take up valuable space in the clinic, perhaps needing a dedicated room, whether it would slow down clinical workflow, whether it is sufficiently durable (not requiring significant maintenance/calibration), and whether it would have burdensome training requirements for operation or interpretation of its results. Moreover, FDA approval of any new technology is essential before bringing a new device into clinical use, and insurance coverage is necessary to enable the clinic to recover the costs of purchasing and maintaining the equipment in a working condition (whether it needs daily or weekly calibration, for example, or whether it is costly to repair). Finally, some technologies are just not suited to melanoma detection and can never achieve sufficiently accurate specificity and sensitivity because they do not significantly assess the fundamental melanoma mechanism and therefore are not capable of analyzing the pathophysiology with sufficient clarity. These factors cause clinicians to be generally reluctant to replace or supplement dermoscopy and their own clinical expertise with new technology. In summary, key factors limiting the clinical use of these emerging modalities include cost (both capital cost and cost per use including specialist time for data interpretation), reimbursement, portability/lack of handheld devices, and lack of experience with interpretation of images.
Reviewing existing technologies with regard to these potential barriers to adoption, photography is a relatively inexpensive imaging method but can be cumbersome, lacks standardization with regards to color matching, which can be valuable for automated image analysis, has limited resolution without dermoscopy, and likely does not have the capability to analyze the pathophysiology sufficiently. Nevertheless, whole-body photographic imaging for screening, through companies such as MelanoScan, are marketed to patients at risk for skin cancer. The funding model is out of pocket payments for peace of mind. The use of photography coupled with smartphone apps has been described as lacking in consistent validation and transparent user communication [366]. Dermoscopy is the most inexpensive and prevalent technology in the clinic for assessing lesions, but its sensitivity, which is never very high, depends on the expertise of the user, and early and amelanotic melanomas are challenging to diagnose via dermoscopy, as they lack well-known superficial features of melanoma. In fact, there is a significant concern that dermoscopy will never have the capability to fully assess the pathophysiology of all forms of cutaneous melanoma.
The EIS system Nevisense is FDA-approved, reimbursement coverage is improving (covered by some private insurers and Medicare patients in about one-third of US states), and the device is fast, easy to use, and low-cost. One drawback is that while it enhances sensitivity, it may produce false positives, leading to unnecessary biopsies and patient anxiety. Another drawback is that it primarily assesses superficial lesions. SS-OCT provides around ten micron-resolution images to the depth of 2 mm, allowing for the assessment of tumors. However, although its use in dermatology is FDA-approved, it cannot reliably differentiate melanomas from benign pigmented lesions without the use of a sophisticated computational method, which has not yet been FDA-approved, so it is not currently marketed for melanoma detection. In addition, equipment costs are relatively high, and insurance does not reimburse for OCT skin imaging. Interestingly, insurance does not reimburse for OCT ocular imaging, but there, ophthalmologists appreciate the highly detailed images of the retina, optic nerve, and other eye structures, and offer the service to their patients (at increased cost). MSI was commercialized through the MelaFind device but failed commercially due to lack of clinical adoption. Moreover, its FDA approval was limited to “dermatologists experienced in melanoma detection”, narrowing its market potential, and it was never able to secure insurance reimbursement. HFUS has a good penetration depth and can visualize the size and shape of a tumor, but its poor resolution and low specificity make melanoma diagnosis difficult and there are no current efforts to commercialize it for melanoma screening. The Aura, which uses Raman spectroscopy, is an adjunct device that is approved for use in Canada but not in the US. The FDA previously required an endpoint sensitivity of 95% for a device that detects melanoma. Vita Imaging, which currently owns the patents to the Aura, expects FDA clearance in 2025. The ESS device DermaSensor is handheld and classifies lesions as “investigate further” or “monitor”. Notably, it was FDA-approved (in January 2024) for non-dermatologists, with the manufacturer targeting primary care physicians to use the device to improve their confidence and sensitivity for referring lesions to specialists. The current marketing model is to rent the devices for monthly fees (USD 399/month for unlimited users). RCM has the highest combination of resolution, sensitivity (88–98%), and specificity (57–98%), which can visualize cytological (nuclear and cytoplasmic) and architectural details of skin, similar to histology. However, it has failed to gain acceptance in the dermatology community. Non-technology-based barriers could include system cost, the low reimbursement rate, the cost to train users in acquiring images and interpreting them, and breaks in clinical workflow (patients typically must travel from dermatologists to separate centers for imaging). Moreover, neither RCM nor any of these other listed methods can stage melanoma (determine melanoma depth), so they are largely in the field of assistive screening technologies.
The technologies in development can be categorized as those solely applicable to melanoma screening and those with potentially wider applications. HSI, quantitative dynamic IR imaging, TPI, and EPR research are focused on developing melanoma detection systems. MPI is focused on biomedical research (mouse models of melanoma), FTIR shows promise for analyzing biopsied tissue, and RTE has been used to detect malignant lymph nodes, but the process is apparently time-consuming. Fiber diffraction research appears to be no longer active. PA microscopy and tomography has the widest range of potential applications for cutaneous melanoma [34], but questions remain whether PA applications can overcome the notably large potential barriers to clinical adoption.
Another barrier to clinical adoption that needs discussion is the difficulty of incorporating AI into medical devices in the US. AI has the potential to improve image resolution and standardize image interpretation, reducing training times and improving the specificity and sensitivity of nearly all imaging-based technologies. However, while there has been a significant surge in FDA approvals recently [367], currently, the FDA does not permit AI algorithms to be adapted after they have been released without further FDA review. More recently, the FDA has released guiding principles for developing predetermined change control plans for AI-enabled device software functions in order to support iterative improvement while maintaining device safety and effectiveness [368]. Their adoption would lead to improvements in a range of facets of melanoma care, from screening through post-treatment monitoring.
More different types of imaging and spectroscopic modalities have been researched for use with melanoma than nearly any other form of cancer, yet screening remains largely limited to dermatologists, often aided by dermoscopy. The ideal melanoma screening system should be highly accurate, low-cost, and sufficiently easy to use so that it can be available in every community, including in low-income and rural communities, since dermatologists can be hard to find in rural and low-income communities in the US and around the world.

7. Conclusions

Non-invasive imaging methods are continually being developed to improve melanoma screening, pathology, staging, detection of metastasis, and treatment monitoring. Novel methods demonstrate their value by showing improved accuracy at these tasks, which is an important starting point. On the other hand, the mere fact that very many different technologies and methodologies have been invented, developed, and promoted without being translated to the clinic highlights the fact that other considerations, beyond accuracy, are important for improving melanoma healthcare. Melanoma healthcare must start with effective screening to ensure that all melanomas are detected before they spread. To adequately detect melanoma, penetration depth should, at a minimum, reach the papillary dermis to differentiate melanoma in situ from invasive melanoma. Also, images with near-cellular-level resolution are desirable to distinguish histologic differences, enabling a definitive diagnosis of a lesion. Both RCM and OCT meet these two criteria. Increasing the opportunity to detect a malignancy at its earliest stage reduces the need for every other form of melanoma healthcare. Most existing and proposed screening imaging modalities are intended for dermatology clinics, which limit their availability. For example, they are not useful for screening people who cannot access dermatology clinics either because they cannot get an appointment or because of the lack of dermatologists in their community. Some screening modalities are targeted for use outside of dermatology clinics, but these must demonstrate high specificity to gain dermatologist (and patient) acceptance to minimize screening bottlenecks. In addition, many modalities, including Raman imaging, multispectral imaging, hyperspectral imaging, EIS, and ESS have very good sensitivity but are not able to provide visual evidence, which can reduce dermatologist acceptance. An ideal screening system would have to demonstrate high specificity and sensitivity, ease of use, and some form of imaging or other quantitative metric to increase dermatologist buy-in, and low capital costs to enable widespread implementation, but no current modality meets all of these criteria. After diagnosis, in order to characterize melanoma for staging, a depth penetration of at least 2 mm is required, recommending the use of HFUS or PAI. Then, for margin delineation for surgical planning, the penetration depth is important as well as a large field of view, making HFUS and PAI a good choice here as well. The detection of lymph node involvement has long been performed by lymphoscintigraphy, but SPECT/CT has recently demonstrated it can detect 50% more sentinel nodes than planar lymphoscintigraphy, reducing melanoma mortality. MRI and PET/CT are both used to find distant metastases, but PET/CT is the modality of choice to predict response to immunotherapies. Emerging technologies, while mainly focused on melanoma screening, also propose improvements for the staging and detection of distant metastases (PAI) and improved pathology (FTIR).
Table 1 summarizes all the modalities discussed in this review for detecting and managing melanoma, including how they are or could be used in different stages of melanoma healthcare, the type of biomarkers they provide, and the main limitations of each modality. A number of the modalities discussed in this review and included in Table 1 are focused on applications other than screening, seeking to replace current processes to reduce mortality and morbidity (detect metastasis sooner, find lymph nodes or circulating tumor cells with higher accuracy, better predict metastasis). These advances are also important, but likely will have to demonstrate not only improved accuracy but cost-effective clinical implementation to change current practices. In Table 1, each technology’s main utility is categorized as “prescreening”, that is, helpful prior to consultation with a dermatologist, “screening”, which happens in the presence of a dermatologist, “staging”, “surgical/treatment guidance”, “decision-making regarding treatment”, and “post-surgical/treatment monitoring”.

Author Contributions

L.H., writing the original draft and manuscript revision, J.W.F., R.M., A.R.-E., D.T., S.O., A.F., S.D., M.J., K.N. and D.M., revising the manuscript, K.A., conceptualization, writing the original draft, and revising the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This project is funded by the Melanoma Research Alliance (No. 624320) and Wayne State University and the Medical Student Summer Research Fund.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Linos, E.; Swetter, S.M.; Cockburn, M.G.; Colditz, G.A.; Clarke, C.A. Increasing Burden of Melanoma in the United States. J. Investig. Dermatol. 2009, 129, 1666–1674. (In English) [Google Scholar] [CrossRef]
  2. Erdmann, F.; Lortet-Tieulent, J.; Schüz, J.; Zeeb, H.; Greinert, R.; Breitbart, E.W.; Bray, F. International trends in the incidence of malignant melanoma 1953–2008—Are recent generations at higher or lower risk? Int. J. Cancer 2013, 132, 385–400. [Google Scholar] [CrossRef]
  3. Kosary, C.L.; Altekruse, S.F.; Ruhl, J.; Lee, R.; Dickie, L. Clinical and prognostic factors for melanoma of the skin using SEER registries: Collaborative stage data collection system, version 1 and version 2. Cancer 2014, 120, 3807–3814. [Google Scholar] [CrossRef]
  4. Guy, G.P., Jr.; Thomas, C.C.; Thompson, T.; Watson, M.; Massetti, G.M.; Richardson, L.C.; Centers for Disease Control and Prevention (CDC). Vital signs: Melanoma incidence and mortality trends and projections—United States, 1982–2030. MMWR Morb. Mortal. Wkly. Rep. 2015, 64, 591–596. (In English) [Google Scholar] [PubMed]
  5. Whiteman, D.C.; Green, A.C.; Olsen, C.M. The Growing Burden of Invasive Melanoma: Projections of Incidence Rates and Numbers of New Cases in Six Susceptible Populations through 2031. J. Investig. Dermatol. 2016, 136, 1161–1171. [Google Scholar] [CrossRef] [PubMed]
  6. Arnold, M.; Singh, D.; Laversanne, M.; Vignat, J.; Vaccarella, S.; Meheus, F.; Cust, A.E.; de Vries, E.; Whiteman, D.C.; Bray, F. Global Burden of Cutaneous Melanoma in 2020 and Projections to 2040. JAMA Dermatol. 2022, 158, 495–503. [Google Scholar] [CrossRef]
  7. Street, W. Cancer Facts & Figures 2017. Cancer Facts and Statistics, American Cancer Society. 2017. Available online: https://www.cancer.org/research/cancer-facts-statistics/all-cancer-facts-figures/cancer-facts-figures-2017.html (accessed on 12 January 2025).
  8. Coory, M.; Baade, P.; Aitken, J.; Smithers, M.; McLeod, G.R.C.; Ring, I. Trends for in situ and invasive melanoma in Queensland, Australia, 1982–2002. Cancer Causes Control 2006, 17, 21–27. (In English) [Google Scholar] [CrossRef] [PubMed]
  9. Matthews, N.H.; Li, W.-Q.; Qureshi, A.A.; Weinstock, M.A.; Cho, E. Epidemiology of Melanoma. In Cutaneous Melanoma: Etiology and Therapy; Ward, W.H., Farma, J.M., Ward, W.H., Farma, J.M., Eds.; Exon Publications: Brisbane, Australia, 2017. [Google Scholar]
  10. American_Cancer_Society. Key Statistics for Melanoma Skin Cancer. Available online: https://www.cancer.org/cancer/types/melanoma-skin-cancer/about/key-statistics.html (accessed on 12 January 2025).
  11. Welch, H.G.; Woloshin, S.; Schwartz, L.M. Skin biopsy rates and incidence of melanoma: Population based ecological study. BMJ 2005, 331, 481. [Google Scholar] [CrossRef]
  12. Bharath, A.; Turner, R. Impact of climate change on skin cancer. J. R. Soc. Med. 2009, 102, 215–218. [Google Scholar] [CrossRef]
  13. Shellenberger, R.; Nabhan, M.; Kakaraparthi, S. Melanoma screening: A plan for improving early detection. Ann. Med. 2016, 48, 142–148. [Google Scholar] [CrossRef]
  14. Noone, A.M.; Howlander, N.; Kraphcho, M.; Miller, D.; Brest, A.; Yu, M.; Ruhl, J.; Tatalovich, Z.; Mariotto, A.; Lewis, D.R.; et al. Cancer Statistics Review, 1975–2015—SEER Statistics; National Cancer Institute: Rockville, MD, USA, 2018. [Google Scholar]
  15. Perlis, C.; Herlyn, M. Recent advances in melanoma biology. Oncologist 2004, 9, 182–187. (In English) [Google Scholar] [CrossRef]
  16. Aneja, S.; Aneja, S.; Bordeaux, J.S. Association of increased dermatologist density with lower melanoma mortality. Arch. Dermatol. 2012, 148, 174–178. [Google Scholar] [CrossRef]
  17. Cortez, J.L.; Vasquez, J.; Wei, M.L. The impact of demographics, socioeconomics, and health care access on melanoma outcomes. J. Am. Acad. Dermatol. 2021, 84, 1677–1683. [Google Scholar] [CrossRef] [PubMed]
  18. Themstrup, L.; Jemec, G.B.E. Chapter 6—Optical Coherence Tomography for Skin Cancer and Actinic Keratosis. In Imaging in Dermatology; Hamblin, M.R., Avci, P., Gupta, G.K., Hamblin, M.R., Avci, P., Gupta, G.K., Eds.; Academic Press: Boston, MA, USA, 2016; pp. 59–67. [Google Scholar]
  19. Garibyan, L.; Fisher, D.E. How sunlight causes melanoma. Curr. Oncol. Rep. 2010, 12, 319–326. [Google Scholar] [CrossRef] [PubMed]
  20. Stern, R.S. Prevalence of a history of skin cancer in 2007: Results of an incidence-based model. Arch. Dermatol. 2010, 146, 279–282. [Google Scholar] [CrossRef]
  21. Abbas, O.; Miller, D.D.; Bhawan, J. Cutaneous malignant melanoma: Update on diagnostic and prognostic biomarkers. Am. J. Dermatopathol. 2014, 36, 363–379. (In English) [Google Scholar] [CrossRef] [PubMed]
  22. Zheng, S.; Yu, H.; Zheng, X.; Wu, U.T.; Ming, W.-k.; Huang, H.; Song, J.; Zhang, X.; Lyu, J.; Deng, L. Analysis and prediction of 5-year survival in patients with cutaneous melanoma: A model-based period analysis. Front. Endocrinol. 2023, 14, 1238086. [Google Scholar] [CrossRef]
  23. Thomas, L.; Tranchand, P.; Berard, F.; Secchi, T.; Colin, C.; Moulin, G. Semiological value of ABCDE criteria in the diagnosis of cutaneous pigmented tumors. Dermatology 1998, 197, 11–17. [Google Scholar] [CrossRef]
  24. Harkemanne, E.; Baeck, M.; Tromme, I. Training general practitioners in melanoma diagnosis: A scoping review of the literature. BMJ Open 2021, 11, e043926. (In English) [Google Scholar] [CrossRef]
  25. Nelson, K.C.; Swetter, S.M.; Saboda, K.; Chen, S.C.; Curiel-Lewandrowski, C. Evaluation of the Number-Needed-to-Biopsy Metric for the Diagnosis of Cutaneous Melanoma: A Systematic Review and Meta-analysis. JAMA Dermatol. 2019, 155, 1167–1174. (In English) [Google Scholar] [CrossRef]
  26. Tes, D.; Aber, A.; Zafar, M.; Horton, L.; Fotouhi, A.; Xu, Q.; Moiin, A.; Thompson, A.D.; Moraes Pinto Blumetti, T.C.; Daveluy, S. Granular cell tumor imaging using optical coherence tomography. Biomed. Eng. Comput. Biol. 2018, 9, 1179597218790250. [Google Scholar] [CrossRef]
  27. Narayanamurthy, V.; Padmapriya, P.; Noorasafrin, A.; Pooja, B.; Hema, K.; Khan, A.a.Y.F.; Nithyakalyani, K.; Samsuri, F. Skin cancer detection using non-invasive techniques. RSC Adv. 2018, 8, 28095–28130. (In English) [Google Scholar] [CrossRef] [PubMed]
  28. Panchal, R.; Horton, L.; Poozesh, P.; Baqersad, J.; Nasiriavanaki, M. Vibration analysis of healthy skin: Toward a noninvasive skin diagnosis methodology. J. Biomed. Opt. 2019, 24, 015001. [Google Scholar] [CrossRef] [PubMed]
  29. State-of-the-Art 3D Imaging Can Spot Skin Cancer Early (Video); Baptist Health South Florida: Coral Gables, FL, USA, 2019.
  30. Freeman, T. Millimetre-Wave Imaging Delivers Non-Invasive Skin Cancer Diagnosis. Physics World. 2019. Available online: https://physicsworld.com/a/millimetre-wave-imaging-delivers-non-invasive-skin-cancer-diagnosis/ (accessed on 12 January 2025).
  31. Jalilian, E.; Xu, Q.; Horton, L.; Fotouhi, A.; Reddy, S.; Manwar, R.; Daveluy, S.; Mehregan, D.; Gelovani, J.; Avanaki, K. Contrast-enhanced optical coherence tomography for melanoma detection: An in vitro study. J. Biophotonics 2020, 13, e201960097. (In English) [Google Scholar] [CrossRef] [PubMed]
  32. Reed, J. Non-Invasive Imaging Techniques for Diagnosis. SkinCancer.net. 2020. Available online: https://skincancer.net/procedures-tests/imaging/non-invasive-techniques (accessed on 12 January 2025).
  33. Shapiro, L.; Basra, M.; Patel, H.; Payne, C.; Brazen, B.; Biglione, A. Utilization of Imaging Modalities in the Diagnosis of Melanoma: A Scoping Review. Cureus 2024, 16, 1–10. [Google Scholar] [CrossRef]
  34. Fakhoury, J.W.; Lara, J.B.; Manwar, R.; Zafar, M.; Xu, Q.; Engel, R.; Tsoukas, M.M.; Daveluy, S.; Mehregan, D.; Avanaki, K. Photoacoustic imaging for cutaneous melanoma assessment: A comprehensive review. J. Biomed. Opt. 2024, 29, S11518. [Google Scholar] [CrossRef]
  35. Atkins, M.B.; Curiel-Lewandrowski, C.; Fisher, D.E.; Swetter, S.M.; Tsao, H.; Aguirre-Ghiso, J.A.; Soengas, M.S.; Weeraratna, A.T.; Flaherty, K.T.; Herlyn, M.; et al. The State of Melanoma: Emergent Challenges and Opportunities. Clin. Cancer Res. 2021, 27, 2678–2697. [Google Scholar] [CrossRef]
  36. Bichakjian, C.K.; Halpern, A.C.; Johnson, T.M.; Foote Hood, A.; Grichnik, J.M.; Swetter, S.M.; Tsao, H.; Barbosa, V.H.; Chuang, T.-Y.; Duvic, M.; et al. Guidelines of care for the management of primary cutaneous melanoma. J. Am. Acad. Dermatol. 2011, 65, 1032–1047. (In English) [Google Scholar] [CrossRef]
  37. Reed, D.; Kudchadkar, R.; Zager, J.S.; Sondak, V.K.; Messina, J.L. Controversies in the Evaluation and Management of Atypical Melanocytic Proliferations in Children, Adolescents, and Young Adults. J. Natl. Compr. Cancer Netw. 2013, 11, 679–686. (In English) [Google Scholar] [CrossRef]
  38. Kauffmann, R.M.; Chen, S.L. Workup and staging of malignant melanoma. Surg. Clin. N. Am. 2014, 94, 963–972. (In English) [Google Scholar] [CrossRef]
  39. Sondak, V.K.; Taylor, J.M.G.; Sabel, M.S.; Wang, Y.; Lowe, L.; Grover, A.C.; Chang, A.E.; Yahanda, A.M.; Moon, J.; Johnson, T.M. Mitotic Rate and Younger Age Are Predictors of Sentinel Lymph Node Positivity: Lessons Learned From the Generation of a Probabilistic Model. Ann. Surg. Oncol. 2004, 11, 247–258. (In English) [Google Scholar] [CrossRef]
  40. Mathew, R.; Messina, J.L. Recent Advances in Pathologic Evaluation and Reporting of Melanoma. Semin. Oncol. 2012, 39, 184–191. (In English) [Google Scholar] [CrossRef]
  41. Gershenwald, J.E.; Scolyer, R.A.; Hess, K.R.; Sondak, V.K.; Long, G.V.; Ross, M.I.; Lazar, A.J.; Faries, M.B.; Kirkwood, J.M.; McArthur, G.A.; et al. Melanoma staging: Evidence-based changes in the American Joint Committee on Cancer eighth edition cancer staging manual. CA Cancer J. Clin. 2017, 67, 472–492. (In English) [Google Scholar] [CrossRef] [PubMed]
  42. Ross, M.I.; Gershenwald, J.E. Sentinel lymph node biopsy for melanoma: A critical update for dermatologists after two decades of experience. Clin. Dermatol. 2013, 31, 298–310. (In English) [Google Scholar] [CrossRef]
  43. Mandalà, M.; Galli, F.; Cattaneo, L.; Merelli, B.; Rulli, E.; Ribero, S.; Quaglino, P.; Giorgi, V.D.; Pigozzo, J.; Sileni, V.C.; et al. Mitotic rate correlates with sentinel lymph node status and outcome in cutaneous melanoma greater than 1 millimeter in thickness: A multi-institutional study of 1524 cases. J. Am. Acad. Dermatol. 2017, 76, 264–273.e262. (In English) [Google Scholar] [CrossRef] [PubMed]
  44. Azzola, M.F.; Shaw, H.M.; Thompson, J.F.; Soong, S.-j.; Scolyer, R.A.; Watson, G.F.; Colman, M.H.; Zhang, Y. Tumor mitotic rate is a more powerful prognostic indicator than ulceration in patients with primary cutaneous melanoma. Cancer 2003, 97, 1488–1498. (In English) [Google Scholar] [CrossRef]
  45. Balch, C.M.; Gershenwald, J.E.; Soong, S.-J.; Thompson, J.F.; Atkins, M.B.; Byrd, D.R.; Buzaid, A.C.; Cochran, A.J.; Coit, D.G.; Ding, S.; et al. Final version of 2009 AJCC melanoma staging and classification. J. Clin. Oncol. 2009, 27, 6199–6206. (In English) [Google Scholar] [CrossRef]
  46. Ward, W.H.; Lambreton, F.; Goel, N.; Yu, J.Q.; Farma, J.M. Clinical Presentation and Staging of Melanoma. In Cutaneous Melanoma: Etiology and Therapy; Ward, W.H., Farma, J.M., Ward, W.H., Farma, J.M., Eds.; Exon Publications: Brisbane, Australia, 2017. [Google Scholar]
  47. Amin, M.B.; Greene, F.L.; Edge, S.B.; Compton, C.C.; Gershenwald, J.E.; Brookland, R.K.; Meyer, L.; Gress, D.M.; Byrd, D.R.; Winchester, D.P. The Eighth Edition AJCC Cancer Staging Manual: Continuing to build a bridge from a population-based to a more “personalized” approach to cancer staging. CA A Cancer J. Clin. 2017, 67, 93–99. (In English) [Google Scholar] [CrossRef] [PubMed]
  48. Torbatian, Z.; Adamson, R.; Bance, M.; Brown, J. A split-aperture transmit beamforming technique with phase coherence grating lobe suppression. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2010, 57, 2588–2595. [Google Scholar] [CrossRef]
  49. Adamson, A.S.; Jarmul, J.A.; Pignone, M.P. Screening for melanoma in men: A cost-effectiveness analysis. J. Gen. Intern. Med. 2020, 35, 1175–1181. [Google Scholar] [CrossRef]
  50. Elder, D.E. Melanoma progression. Pathology 2016, 48, 147–154. [Google Scholar] [CrossRef]
  51. Smoller, B.R. Histologic criteria for diagnosing primary cutaneous malignant melanoma. Mod. Pathol. 2006, 19, S34–S40. [Google Scholar] [CrossRef]
  52. Šitum, M.; Buljan, M.; Kolić, M.; Vučić, M. Melanoma–clinical, dermatoscopical, and histopathological morphological characteristics. Acta Dermatovenerol. Croat. 2014, 22, 1–12. [Google Scholar] [PubMed]
  53. Davis, L.E.; Shalin, S.C.; Tackett, A.J. Current state of melanoma diagnosis and treatment. Cancer Biol. Ther. 2019, 20, 1366–1379. [Google Scholar] [CrossRef]
  54. Elmore, J.G.; Barnhill, R.L.; Elder, D.E.; Longton, G.M.; Pepe, M.S.; Reisch, L.M.; Carney, P.A.; Titus, L.J.; Nelson, H.D.; Onega, T. Pathologists’ diagnosis of invasive melanoma and melanocytic proliferations: Observer accuracy and reproducibility study. bmj 2017, 357, j2813. [Google Scholar] [CrossRef] [PubMed]
  55. Plaza, J.A.; Suster, D.; Perez-Montiel, D. Expression of immunohistochemical markers in primary and metastatic malignant melanoma: A comparative study in 70 patients using a tissue microarray technique. Appl. Immunohistochem. Mol. Morphol. 2007, 15, 421–425. [Google Scholar] [CrossRef]
  56. Nwafor, J.N.; Torere, B.E.; Agu, E.; Kadiku, L.; Ogunyemi, T.; Akinsanya, P.A.; Araromi, O.O.; Akahara, D.E.; Okobi, O.E. The Role of Biomarkers in the Diagnosis and Prognosis of Different Stages of Melanoma. Cureus 2023, 15, e38693. (In English) [Google Scholar] [CrossRef]
  57. Ladstein, R.G.; Bachmann, I.M.; Straume, O.; Akslen, L.A. Ki-67 expression is superior to mitotic count and novel proliferation markers PHH3, MCM4 and mitosin as a prognostic factor in thick cutaneous melanoma. BMC Cancer 2010, 10, 140. [Google Scholar] [CrossRef] [PubMed]
  58. Pizzichetta, M.A.; Stanganelli, I. Total Body Photography and Sequential Digital Dermoscopy for Melanoma Diagnosis. In Technology in Practical Dermatology: Non-Invasive Imaging, Lasers and Ulcer Management; Fimiani, M., Rubegni, P., Cinotti, E., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 121–126. [Google Scholar]
  59. Kränke, T.; Tripolt-Droschl, K.; Röd, L.; Hofmann-Wellenhof, R.; Koppitz, M.; Tripolt, M. New AI-algorithms on smartphones to detect skin cancer in a clinical setting-A validation study. PLoS ONE 2023, 18, e0280670. (In English) [Google Scholar] [CrossRef]
  60. Jahn, A.S.; Navarini, A.A.; Cerminara, S.E.; Kostner, L.; Huber, S.M.; Kunz, M.; Maul, J.-T.; Dummer, R.; Sommer, S.; Neuner, A.D.; et al. Over-Detection of Melanoma-Suspect Lesions by a CE-Certified Smartphone App: Performance in Comparison to Dermatologists, 2D and 3D Convolutional Neural Networks in a Prospective Data Set of 1204 Pigmented Skin Lesions Involving Patients’ Perception. Cancers 2022, 14, 3829. [Google Scholar] [CrossRef]
  61. Ji-Xu, A.; Dinnes, J.; Matin, R. Total body photography for the diagnosis of cutaneous melanoma in adults: A systematic review and meta-analysis. Br. J. Dermatol. 2021, 185, 302–312. [Google Scholar] [CrossRef] [PubMed]
  62. Herman, C. Emerging technologies for the detection of melanoma: Achieving better outcomes. Clin. Cosmet. Investig. Dermatol. 2012, 5, 195. [Google Scholar] [CrossRef]
  63. Berk-Krauss, J.; Polsky, D.; Stein, J.A. Mole Mapping for Management of Pigmented Skin Lesions. Dermatol. Clin. 2017, 35, 439–445. (In English) [Google Scholar] [CrossRef]
  64. Dugonik, B.; Dugonik, A.; Marovt, M.; Golob, M. Image Quality Assessment of Digital Image Capturing Devices for Melanoma Detection. Appl. Sci. 2020, 10, 2876. [Google Scholar] [CrossRef]
  65. Risser, J.; Pressley, Z.; Veledar, E.; Washington, C.; Chen, S.C. The impact of total body photography on biopsy rate in patients from a pigmented lesion clinic. J. Am. Acad. Dermatol. 2007, 57, 428–434. (In English) [Google Scholar] [CrossRef]
  66. Drugge, E.D.; Volpicelli, E.R.; Sarac, R.M.; Strang, S.R.; Elston, D.M.; Drugge, R.J. Micromelanomas identified with time-lapse total body photography and dermoscopy. J. Am. Acad. Dermatol. 2018, 78, 182–183. (In English) [Google Scholar] [CrossRef] [PubMed]
  67. Green, W.H.; Wang, S.Q.; Cognetta, A.B. Total-body cutaneous examination, total-body photography, and dermoscopy in the care of a patient with xeroderma pigmentosum and multiple melanomas. Arch. Dermatol. 2009, 145, 910–915. (In English) [Google Scholar] [CrossRef] [PubMed]
  68. Salerni, G.; Carrera, C.; Lovatto, L.; Martí-Laborda, R.M.; Isern, G.; Palou, J.; Alós, L.; Puig, S.; Malvehy, J. Characterization of 1152 lesions excised over 10 years using total-body photography and digital dermatoscopy in the surveillance of patients at high risk for melanoma. J. Am. Acad. Dermatol. 2012, 67, 836–845. (In English) [Google Scholar] [CrossRef]
  69. Vestergaard, M.E.; Macaskill, P.; Holt, P.E.; Menzies, S.W. Dermoscopy compared with naked eye examination for the diagnosis of primary melanoma: A meta-analysis of studies performed in a clinical setting. Br. J. Dermatol. 2008, 159, 669–676. (In English) [Google Scholar] [CrossRef]
  70. Primiero, C.A.; McInerney-Leo, A.M.; Betz-Stablein, B.; Whiteman, D.C.; Gordon, L.; Caffery, L.; Aitken, J.F.; Eakin, E.; Osborne, S.; Gray, L.; et al. Evaluation of the efficacy of 3D total-body photography with sequential digital dermoscopy in a high-risk melanoma cohort: Protocol for a randomised controlled trial. BMJ Open 2019, 9, e032969. (In English) [Google Scholar] [CrossRef]
  71. Banky, J.P.; Kelly, J.W.; English, D.R.; Yeatman, J.M.; Dowling, J.P. Incidence of new and changed nevi and melanomas detected using baseline images and dermoscopy in patients at high risk for melanoma. Arch. Dermatol. 2005, 141, 998–1006. (In English) [Google Scholar] [CrossRef] [PubMed]
  72. Ferreirinha, A.; Farricha, V.; João, A. Melanoma diagnosis with 3D total-body photography. Actas Dermo-Sifiliográficas 2025. (Accepted; In press). [Google Scholar] [CrossRef]
  73. Olsen, J.; Holmes, J.; Jemec, G.B. Advances in optical coherence tomography in dermatology—A review. J. Biomed. Opt. 2018, 23, 040901. [Google Scholar] [CrossRef]
  74. Longo, C.; Pampena, R.; Moscarella, E.; Chester, J.; Starace, M.; Cinotti, E.; Piraccini, B.M.; Argenziano, G.; Peris, K.; Pellacani, G. Dermoscopy of melanoma according to different body sites: Head and neck, trunk, limbs, nail, mucosal and acral. J. Eur. Acad. Dermatol. Venereol. 2023, 37, 1718–1730. [Google Scholar] [CrossRef] [PubMed]
  75. Roldán-Marín, R.; Puig, S.; Malvehy, J. Dermoscopic criteria and melanocytic lesions. G. Ital. Dermatol. Venereol. 2012, 147, 149–159. [Google Scholar] [PubMed]
  76. Ulrich, M.; Stockfleth, E.; Roewert-Huber, J.; Astner, S. Noninvasive diagnostic tools for nonmelanoma skin cancer. Br. J. Dermatol. 2007, 157 (Suppl. S2), 56–58. (In English) [Google Scholar] [CrossRef]
  77. Farnetani, F.; Scope, A.; Mazzoni, L.; Mandel, V.D.; Manfredini, M.; Magi, S.; Vaschieri, C.; Kaleci, S.; Longo, C.; Ciardo, S.; et al. A comparative dermoscopic and reflectance confocal microscopy study of naevi and melanoma with negative pigment network. J. Eur. Acad. Dermatol. Venereol. 2019, 33, 2273–2282. (In English) [Google Scholar] [CrossRef]
  78. Thomas, L.; Puig, S. Dermoscopy, Digital Dermoscopy and Other Diagnostic Tools in the Early Detection of Melanoma and Follow-up of High-risk Skin Cancer Patients. Acta Derm. Venereol. 2017, 97, 14–21. (In English) [Google Scholar] [CrossRef]
  79. Benvenuto-Andrade, C.; Dusza, S.W.; Agero, A.L.C.; Scope, A.; Rajadhyaksha, M.; Halpern, A.C.; Marghoob, A.A. Differences between polarized light dermoscopy and immersion contact dermoscopy for the evaluation of skin lesions. Arch. Dermatol. 2007, 143, 329–338. [Google Scholar] [CrossRef]
  80. Harrison, K. The Accuracy of Skin Cancer Detection Rates with the Implementation of Dermoscopy Among Dermatology Clinicians: A Scoping Review. J. Clin. Aesthet. Dermatol. 2024, 17 (Suppl. S1), S18–S27. (In English) [Google Scholar]
  81. Ye, Z.; Zhang, D.; Zhao, Y.; Chen, M.; Wang, H.; Seery, S.; Qu, Y.; Xue, P.; Jiang, Y. Deep learning algorithms for melanoma detection using dermoscopic imaging: A systematic review and meta-analysis. Artif. Intell. Med. 2024, 155, 102934. [Google Scholar] [CrossRef]
  82. Jaimes, N.; Chen, L.; Dusza, S.W.; Carrera, C.; Puig, S.; Thomas, L.; Kelly, J.W.; Dang, L.; Zalaudek, I.; Braun, R.P.; et al. Clinical and Dermoscopic Characteristics of Desmoplastic Melanomas. JAMA Dermatol. 2013, 149, 413–421. (In English) [Google Scholar] [CrossRef]
  83. Stolz, W.; Schiffner, R.; Burgdorf, W.H.C. Dermatoscopy for facial pigmented skin lesions. Clin. Dermatol. 2002, 20, 276–278. (In English) [Google Scholar] [CrossRef] [PubMed]
  84. Pagnanelli, G.; Soyer, H.P.; Argenziano, G.; Talamini, R.; Barbati, R.; Bianchi, L.; Campione, E.; Carboni, I.; Carrozzo, A.M.; Chimenti, M.S.; et al. Diagnosis of pigmented skin lesions by dermoscopy: Web-based training improves diagnostic performance of non-experts. Br. J. Dermatol. 2003, 148, 698–702. (In English) [Google Scholar] [CrossRef]
  85. Argenziano, G.; Zalaudek, I.; Corona, R.; Sera, F.; Cicale, L.; Petrillo, G.; Ruocco, E.; Hofmann-Wellenhof, R.; Soyer, H.P. Vascular structures in skin tumors: A dermoscopy study. Arch. Dermatol. 2004, 140, 1485–1489. (In English) [Google Scholar] [CrossRef] [PubMed]
  86. Rubegni, P.; Sbano, P.; Burroni, M.; Cevenini, G.; Bocchi, C.; Severi, F.M.; Risulo, M.; Petraglia, F.; Dell’Eva, G.; Fimiani, M.; et al. Melanocytic skin lesions and pregnancy: Digital dermoscopy analysis. Skin Res. Technol. 2007, 13, 143–147. (In English) [Google Scholar] [CrossRef] [PubMed]
  87. Carli, P.; De Giorgi, V.; Crocetti, E.; Mannone, F.; Massi, D.; Chiarugi, A.; Giannotti, B. Improvement of malignant/benign ratio in excised melanocytic lesions in the ‘dermoscopy era’: A retrospective study 1997–2001. Br. J. Dermatol. 2004, 150, 687–692. (In English) [Google Scholar] [CrossRef]
  88. Cinotti, E.; Tognetti, L.; Campoli, M.; Liso, F.; Cicigoi, A.; Cartocci, A.; Rossi, R.; Rubegni, P.; Perrot, J. Super-high magnification dermoscopy can aid the differential diagnosis between melanoma and atypical naevi. Clin. Exp. Dermatol. 2021, 46, 1216–1222. [Google Scholar] [CrossRef]
  89. Kardynal, A.; Olszewska, M. Modern non-invasive diagnostic techniques in the detection of early cutaneous melanoma. J. Dermatol. Case Rep. 2014, 8, 1–8. (In English) [Google Scholar] [CrossRef]
  90. Del Rosario, F.; Farahi, J.M.; Drendel, J.; Buntinx-Krieg, T.; Caravaglio, J.; Domozych, R.; Chapman, S.; Braunberger, T.; Dellavalle, R.P.; Norris, D.A.; et al. Performance of a computer-aided digital dermoscopic image analyzer for melanoma detection in 1,076 pigmented skin lesion biopsies. J. Am. Acad. Dermatol. 2018, 78, 927–934.e926. (In English) [Google Scholar] [CrossRef]
  91. Marchetti, M.A.; Cowen, E.A.; Kurtansky, N.R.; Weber, J.; Dauscher, M.; DeFazio, J.; Deng, L.; Dusza, S.W.; Haliasos, H.; Halpern, A.C.; et al. Prospective validation of dermoscopy-based open-source artificial intelligence for melanoma diagnosis (PROVE-AI study). npj Digit. Med. 2023, 6, 127. [Google Scholar] [CrossRef] [PubMed]
  92. Combalia, M.; Codella, N.; Rotemberg, V.; Carrera, C.; Dusza, S.; Gutman, D.; Helba, B.; Kittler, H.; Kurtansky, N.R.; Liopyris, K.; et al. Validation of artificial intelligence prediction models for skin cancer diagnosis using dermoscopy images: The 2019 International Skin Imaging Collaboration Grand Challenge. Lancet Digit. Health 2022, 4, e330–e339. [Google Scholar] [CrossRef]
  93. Regio Pereira, A.; Corral-Forteza, M.; Collgros, H.; El Sharouni, M.A.; Ferguson, P.M.; Scolyer, R.A.; Guitera, P. Dermoscopic features and screening strategies for the detection of small-diameter melanomas. Clin. Exp. Dermatol. 2022, 47, 932–941. (In English) [Google Scholar] [CrossRef]
  94. Kittler, H.; Guitera, P.; Riedl, E.; Avramidis, M.; Teban, L.; Fiebiger, M.; Weger, R.A.; Dawid, M.; Menzies, S. Identification of clinically featureless incipient melanoma using sequential dermoscopy imaging. Arch. Dermatol. 2006, 142, 1113–1119. (In English) [Google Scholar] [CrossRef] [PubMed]
  95. Altamura, D.; Avramidis, M.; Menzies, S.W. Assessment of the optimal interval for and sensitivity of short-term sequential digital dermoscopy monitoring for the diagnosis of melanoma. Arch. Dermatol. 2008, 144, 502–506. (In English) [Google Scholar] [CrossRef]
  96. Haenssle, H.A.; Korpas, B.; Hansen-Hagge, C.; Buhl, T.; Kaune, K.M.; Johnsen, S.; Rosenberger, A.; Schön, M.P.; Emmert, S. Selection of patients for long-term surveillance with digital dermoscopy by assessment of melanoma risk factors. Arch. Dermatol. 2010, 146, 257–264. (In English) [Google Scholar] [CrossRef] [PubMed]
  97. Tromme, I.; Devleesschauwer, B.; Beutels, P.; Richez, P.; Praet, N.; Sacré, L.; Marot, L.; Van Eeckhout, P.; Theate, I.; Baurain, J.-F.; et al. Selective Use of Sequential Digital Dermoscopy Imaging Allows a Cost Reduction in the Melanoma Detection Process: A Belgian Study of Patients with a Single or a Small Number of Atypical Nevi. PLoS ONE 2014, 9, e109339. [Google Scholar] [CrossRef] [PubMed]
  98. Fink, C.; Haenssle, H.A. Non-invasive tools for the diagnosis of cutaneous melanoma. Skin Res. Technol. 2017, 23, 261–271. [Google Scholar] [CrossRef]
  99. Welzel, J.; Schuh, S. Noninvasive diagnosis in dermatology. JDDG J. Der. Dtsch. Dermatol. Ges. 2017, 15, 999–1016. (In English) [Google Scholar] [CrossRef]
  100. Malvehy, J.; Hauschild, A.; Curiel-Lewandrowski, C.; Mohr, P.; Hofmann-Wellenhof, R.; Motley, R.; Berking, C.; Grossman, D.; Paoli, J.; Loquai, C.; et al. Clinical performance of the Nevisense system in cutaneous melanoma detection: An international, multicentre, prospective and blinded clinical trial on efficacy and safety. Br. J. Dermatol. 2014, 171, 1099–1107. (In English) [Google Scholar] [CrossRef]
  101. Braun, R.P.; Mangana, J.; Goldinger, S.; French, L.; Dummer, R.; Marghoob, A.A. Electrical Impedance Spectroscopy in Skin Cancer Diagnosis. Dermatol. Clin. 2017, 35, 489–493. [Google Scholar] [CrossRef] [PubMed]
  102. Mohr, P.; Birgersson, U.; Berking, C.; Henderson, C.; Trefzer, U.; Kemeny, L.; Sunderkötter, C.; Dirschka, T.; Motley, R.; Frohm-Nilsson, M. Electrical impedance spectroscopy as a potential adjunct diagnostic tool for cutaneous melanoma. Skin Res. Technol. 2013, 19, 75–83. [Google Scholar] [CrossRef] [PubMed]
  103. Anushree, U.; Shetty, S.; Kumar, R.; Bharati, S. Adjunctive Diagnostic Methods for Skin Cancer Detection: A Review of Electrical Impedance-Based Techniques. Bioelectromagnetics 2022, 43, 193–210. [Google Scholar] [CrossRef] [PubMed]
  104. Chavez-Bourgeois, M.; Ribero, S.; Barreiro, A.; Espinoza, N.; Carrera, C.; Garcia, A.; Alos, L.; Puig, S.; Malvehy, J. Reflectance Confocal Microscopy and Electrical Impedance Spectroscopy in the Early Detection of Melanoma in Changing Lesions during Long-term Follow-up of Very High-risk Patients. Acta Derm. Venereol. 2022, 102, adv00751. (In English) [Google Scholar] [CrossRef]
  105. Scibase. Press Release. Available online: https://scibase.com/us/scibase-receives-increased-us-medicare-fee-schedules-for-the-nevisense-melanoma-test/ (accessed on 12 January 2025).
  106. Smith, L.; MacNeil, S. State of the art in non-invasive imaging of cutaneous melanoma. Skin Res. Technol. 2011, 17, 257–269. (In English) [Google Scholar] [CrossRef]
  107. Ferris, L.K.; Harris, R.J. New Diagnostic Aides for Melanoma. Dermatol. Clin. 2012, 30, 535–545. [Google Scholar] [CrossRef]
  108. Rey-Barroso, L.; Burgos-Fernández, F.J.; Delpueyo, X.; Ares, M.; Royo, S.; Malvehy, J.; Puig, S.; Vilaseca, M. Visible and Extended Near-Infrared Multispectral Imaging for Skin Cancer Diagnosis. Sensors 2018, 18, 1441. (In English) [Google Scholar] [CrossRef]
  109. Braga, J.C.T.; Macedo, M.P.; Pinto, C.; Duprat, J.; Begnami, M.D.; Pellacani, G.; Rezze, G.G. Learning Reflectance Confocal Microscopy of Melanocytic Skin Lesions through Histopathologic Transversal Sections. PLoS ONE 2013, 8, e81205. [Google Scholar] [CrossRef]
  110. Marghoob, A.A.; Swindle, L.D.; Moricz, C.Z.M.; Negron, F.A.S.; Slue, B.; Halpern, A.C.; Kopf, A.W. Instruments and new technologies for the in vivo diagnosis of melanoma. J. Am. Acad. Dermatol. 2003, 49, 777–797. [Google Scholar] [CrossRef]
  111. Gareau, D.S.; Merlino, G.; Corless, C.; Kulesz-Martin, M.; Jacques, S.L. Noninvasive imaging of melanoma with reflectance mode confocal scanning laser microscopy in a murine model. J. Investig. Dermatol. 2007, 127, 2184–2190. (In English) [Google Scholar] [CrossRef]
  112. González, S.; Gilaberte-Calzada, Y. In vivo reflectance-mode confocal microscopy in clinical dermatology and cosmetology. Int. J. Cosmet. Sci. 2008, 30, 1–17. (In English) [Google Scholar] [CrossRef] [PubMed]
  113. Uribe, P.; Collgros, H.; Scolyer, R.A.; Menzies, S.W.; Guitera, P. In Vivo Reflectance Confocal Microscopy for the Diagnosis of Melanoma and Melanotic Macules of the Lip. JAMA Dermatol. 2017, 153, 882–891. (In English) [Google Scholar] [CrossRef]
  114. Calzavara-Pinton, P.; Longo, C.; Venturini, M.; Sala, R.; Pellacani, G. Reflectance confocal microscopy for in vivo skin imaging. Photochem. Photobiol. 2008, 84, 1421–1430. [Google Scholar] [CrossRef] [PubMed]
  115. Navarrete-Dechent, C.; Cordova, M.; Postow, M.A.; Pulitzer, M.; Lezcano, C.; Halpern, A.C.; Rossi, A.M. Evaluation of the Response of Unresectable Primary Cutaneous Melanoma to Immunotherapy Visualized With Reflectance Confocal Microscopy: A Report of 2 Cases. JAMA Dermatol. 2019, 155, 347–352. (In English) [Google Scholar] [CrossRef]
  116. Ahlgrimm-Siess, V.; Laimer, M.; Rabinovitz, H.S.; Oliviero, M.; Hofmann-Wellenhof, R.; Marghoob, A.A.; Scope, A. Confocal Microscopy in Skin Cancer. Curr. Dermatol. Rep. 2018, 7, 105–118. [Google Scholar] [CrossRef]
  117. Soenen, A.; Vourc’h, M.; Khammari, A.; Nguyen, J.-M.; Bossard, C.; Musquer, M.D.; Vergier, B.; Dréno, B. Change in lentigo maligna score assessed by in vivo reflectance confocal microscopy after 1 month of imiquimod treatment for lentigo maligna management. J. Am. Acad. Dermatol. 2022, 86, 1042–1048. [Google Scholar] [CrossRef]
  118. Song, S.; Xu, J.; Wang, R.K. Long-range and wide field of view optical coherence tomography for in vivo 3D imaging of large volume object based on akinetic programmable swept source. Biomed. Opt. Express 2016, 7, 4734–4748. [Google Scholar] [CrossRef]
  119. Pezzini, C.; Kaleci, S.; Chester, J.; Farnetani, F.; Longo, C.; Pellacani, G. Reflectance confocal microscopy diagnostic accuracy for malignant melanoma in different clinical settings: Systematic review and meta-analysis. J. Eur. Acad. Dermatol. Venereol. 2020, 34, 2268–2279. [Google Scholar] [CrossRef]
  120. Pellacani, G.; Guitera, P.; Longo, C.; Avramidis, M.; Seidenari, S.; Menzies, S. The impact of in vivo reflectance confocal microscopy for the diagnostic accuracy of melanoma and equivocal melanocytic lesions. J. Investig. Dermatol. 2007, 127, 2759–2765. [Google Scholar] [CrossRef]
  121. Ge, X.; Chen, S.; Chen, S.; Liu, L. High Resolution Optical Coherence Tomography. J. Light. Technol. 2021, 39, 3824–3835. [Google Scholar] [CrossRef]
  122. Gambichler, T.; Jaedicke, V.; Terras, S. Optical coherence tomography in dermatology: Technical and clinical aspects. Arch. Dermatol. Res. 2011, 303, 457–473. (In English) [Google Scholar] [CrossRef] [PubMed]
  123. Avanaki, M.R.N.; Podoleanu, A.G.; Schofield, J.B.; Jones, C.; Sira, M.; Liu, Y.; Hojjat, A. Quantitative evaluation of scattering in optical coherence tomography skin images using the extended Huygens Fresnel theorem. Appl. Opt. 2013, 52, 1574–1580. (In English) [Google Scholar] [CrossRef] [PubMed]
  124. Xu, Q.; Adabi, S.; Clayton, A.; Daveluy, S.; Mehregan, D.; Nasiriavanaki, M. Swept-Source Optical Coherence Tomography–Supervised Biopsy. Dermatol. Surg. 2018, 44, 768. (In English) [Google Scholar] [CrossRef] [PubMed]
  125. Adabi, S.; Turani, Z.; Fatemizadeh, E.; Clayton, A.; Nasiriavanaki, M. Optical coherence tomography technology and quality improvement methods for optical coherence tomography images of skin: A short review. Biomed. Eng. Comput. Biol. 2017, 8, 1179597217713475. [Google Scholar] [CrossRef]
  126. Avanaki, M.R.N.; Podoleanu, A. En-face time-domain optical coherence tomography with dynamic focus for high-resolution imaging. J. Biomed. Opt. 2017, 22, 056009. [Google Scholar] [CrossRef]
  127. Faiza, M.; Adabi, S.; Daoud, B.; Avanaki, M.R.N. High-resolution wavelet-fractal compressed optical coherence tomography images. Appl. Opt. 2017, 56, 1119–1123. (In English) [Google Scholar] [CrossRef]
  128. Adabi, S.; Fotouhi, A.; Xu, Q.; Daveluy, S.; Mehregan, D.; Podoleanu, A.; Nasiriavanaki, M. An overview of methods to mitigate artifacts in optical coherence tomography imaging of the skin. Skin Res. Technol. 2018, 24, 265–273. [Google Scholar] [CrossRef]
  129. Hojjatoleslami, A.; Avanaki, M.R. OCT skin image enhancement through attenuation compensation. Appl. Opt. 2012, 51, 4927–4935. [Google Scholar] [CrossRef]
  130. Nasiri-Avanaki, M.; Aber, A.; Hojjatoleslami, S.; Sira, M.; Schofield, J.B.; Jones, C.; Podoleanu, A.G. Dynamic focus optical coherence tomography: Feasibility for improved basal cell carcinoma investigation. In Imaging, Manipulation, and Analysis of Biomolecules, Cells, and Tissues X; International Society for Optics and Photonics: San Francisco, CA, USA, 2012; Volume 8225, p. 82252J. [Google Scholar]
  131. Avanaki, M.R.; Hojjatoleslami, A.; Sira, M.; Schofield, J.B.; Jones, C.; Podoleanu, A.G. Investigation of basal cell carcinoma using dynamic focus optical coherence tomography. Appl. Opt. 2013, 52, 2116–2124. [Google Scholar] [CrossRef]
  132. Herman, C.; Cetingul, M.P. Quantitative visualization and detection of skin cancer using dynamic thermal imaging. J. Vis. Exp. JoVE 2011, 51, 2679. (In English) [Google Scholar] [CrossRef]
  133. Turani, Z.; Fatemizadeh, E.; Blumetti, T.; Daveluy, S.; Moraes, A.F.; Chen, W.; Mehregan, D.; Andersen, P.E.; Nasiriavanaki, M. Optical Radiomic Signatures Derived from Optical Coherence Tomography Images Improve Identification of Melanoma. Cancer Res. 2019, 79, 2021–2030. [Google Scholar] [CrossRef] [PubMed]
  134. Gambichler, T.; Regeniter, P.; Bechara, F.G.; Orlikov, A.; Vasa, R.; Moussa, G.; Stücker, M.; Altmeyer, P.; Hoffmann, K.J. Characterization of benign and malignant melanocytic skin lesions using optical coherence tomography in vivo. J. Am. Acad. Dermatol. 2007, 57, 629–637. [Google Scholar] [CrossRef] [PubMed]
  135. Lai, P.-Y.; Shih, T.-Y.; Chang, Y.-H.; Chang, C.-H.; Kuo, W.-C. Deep Learning With Optical Coherence Tomography for Melanoma Identification and Risk Prediction. J. Biophotonics 2024, 18, e202400277. [Google Scholar] [CrossRef] [PubMed]
  136. Schuh, S.; Holmes, J.; Ulrich, M.; Themstrup, L.; Jemec, G.B.E.; Carvalho, N.D.; Pellacani, G.; Welzel, J. Imaging Blood Vessel Morphology in Skin: Dynamic Optical Coherence Tomography as a Novel Potential Diagnostic Tool in Dermatology. Dermatol. Ther. 2017, 7, 187–202. (In English) [Google Scholar] [CrossRef]
  137. Gambichler, T.; Schmid-Wendtner, M.H.; Plura, I.; Kampilafkos, P.; Stücker, M.; Berking, C.; Maier, T. A multicentre pilot study investigating high-definition optical coherence tomography in the differentiation of cutaneous melanoma and melanocytic naevi. J. Eur. Acad. Dermatol. Venereol. 2015, 29, 537–541. (In English) [Google Scholar] [CrossRef]
  138. Hinz, T.; Ehler, L.-K.; Voth, H.; Fortmeier, I.; Hoeller, T.; Hornung, T.; Schmid-Wendtner, M.-H. Assessment of Tumor Thickness in Melanocytic Skin Lesions: Comparison of Optical Coherence Tomography, 20-MHz Ultrasound and Histopathology. Dermatology 2011, 223, 161–168. (In English) [Google Scholar] [CrossRef]
  139. Latriglia, F.; Ogien, J.; Tavernier, C.; Fischman, S.; Suppa, M.; Perrot, J.L.; Dubois, A. Line-Field Confocal Optical Coherence Tomography (LC-OCT) for Skin Imaging in Dermatology. Life 2023, 13, 2268. (In English) [Google Scholar] [CrossRef]
  140. Chauvel-Picard, J.; Bérot, V.; Tognetti, L.; Orte Cano, C.; Fontaine, M.; Lenoir, C.; Pérez-Anker, J.; Puig, S.; Dubois, A.; Forestier, S.; et al. Line-field confocal optical coherence tomography as a tool for three-dimensional in vivo quantification of healthy epidermis: A pilot study. J. Biophotonics 2022, 15, e202100236. [Google Scholar] [CrossRef]
  141. Ogien, J.; Chauvel-Picard, J.; Bérot, V.; Tognetti, L.; Cano, C.O.; Fontaine, M.; Lenoir, C.; Pérez-Anker, J.; Puig, S.; Dubois, A. Three-dimensional microscopic quantification of in vivo healthy epidermis based on line-field confocal optical coherence tomography (LC-OCT) assisted by artificial intelligence. In Proceedings of the Photonics in Dermatology and Plastic Surgery, San Francisco, CA, USA, 22–23 January 2022; SPIE: San Francisco, CA, USA, 2022; Volume 11934, pp. 47–52. [Google Scholar]
  142. Ruini, C.; Schuh, S.; Sattler, E.; Welzel, J. Line-field confocal optical coherence tomography—Practical applications in dermatology and comparison with established imaging methods. Skin Res. Technol. 2021, 27, 340–352. [Google Scholar] [CrossRef]
  143. Perez-Anker, J.; Soglia, S.; Lenoir, C.; Albero, R.; Alos, L.; García, A.; Alejo, B.; Cinotti, E.; Orte Cano, C.; Habougit, C.; et al. Criteria for melanocytic lesions in LC-OCT. J. Eur. Acad. Dermatol. Venereol. 2024, 38, 2005–2016. [Google Scholar] [CrossRef]
  144. Schuh, S.; Ruini, C.; Perwein, M.K.E.; Daxenberger, F.; Gust, C.; Sattler, E.C.; Welzel, J. Line-field confocal optical coherence tomography: A new tool for the differentiation between nevi and melanomas? Cancers 2022, 14, 1140. [Google Scholar] [CrossRef]
  145. Schmitt, J.; Yadlowsky, M.; Bonner, R. Subsurface imaging of living skin with optical coherence microscopy. Dermatology 1995, 191, 93–98. [Google Scholar] [CrossRef]
  146. Aguirre, A.D.; Fujimoto, J.G. Optical coherence microscopy. In Optical Coherence Tomography: Technology and Applications; Springer: Berlin/Heidelberg, Germany, 2008; pp. 505–542. [Google Scholar]
  147. Wan, Q.; Applegate, B.E. Multiphoton coherence domain molecular imaging with pump-probe optical coherence microscopy. Opt. Lett. 2010, 35, 532–534. [Google Scholar] [CrossRef]
  148. Levine, A.; Wang, K.; Markowitz, O. Optical coherence tomography in the diagnosis of skin cancer. Dermatol. Clin. 2017, 35, 465–488. [Google Scholar] [CrossRef] [PubMed]
  149. Suppa, M.; Palmisano, G.; Tognetti, L.; Lenoir, C.; Cappilli, S.; Fontaine, M.; Cano, C.O.; Diet, G.; Perez-Anker, J.; Schuh, S. Line-field confocal optical coherence tomography in melanocytic and non-melanocytic skin tumors. Ital. J. Dermatol. Venereol. 2023, 158, 180–189. [Google Scholar] [CrossRef]
  150. Ni, G.; Ge, X.; Liu, L.; Zhang, J.; Wang, X.; Liu, J.; Liu, L.; Liu, Y. Towards indicating human skin state in vivo using geometry-dependent spectroscopic contrast imaging. IEEE Photonics Technol. Lett. 2020, 32, 697–700. [Google Scholar] [CrossRef]
  151. Ge, X.; Chen, S.; Lin, K.; Ni, G.; Bo, E.; Wang, L.; Liu, L. Deblurring, artifact-free optical coherence tomography with deconvolution-random phase modulation. Opto-Electron. Sci. 2024, 3, 230020. [Google Scholar] [CrossRef]
  152. Chen, S.; Ge, X.; Liu, X.; Ding, Q.; Wang, N.; Wang, X.; Chen, S.; Liang, H.; Deng, Y.; Xiong, Q. Understanding optical reflectance contrast for real-time characterization of epithelial precursor lesions. Bioeng. Transl. Med. 2019, 4, e10137. [Google Scholar] [CrossRef]
  153. Rey-Barroso, L.; Peña-Gutiérrez, S.; Yáñez, C.; Burgos-Fernández, F.J.; Vilaseca, M.; Royo, S. Optical technologies for the improvement of skin cancer diagnosis: A review. Sensors 2021, 21, 252. [Google Scholar] [CrossRef]
  154. Ge, X.; Tang, H.; Wang, X.; Liu, X.; Chen, S.; Wang, N.; Ni, G.; Yu, X.; Chen, S.; Liang, H. Geometry-dependent spectroscopic contrast in deep tissues. Iscience 2019, 19, 965–975. [Google Scholar] [CrossRef]
  155. Sakalauskienė, K.; Valiukevičienė, S.; Raišutis, R.; Linkevičiūtė, G. The significance of spectrophotometric image analysis for diagnosis of the melanocytic skin tumours in association with their thickness. Skin Res. Technol. 2018, 24, 692–698. [Google Scholar] [CrossRef]
  156. Tehrani, H.; Walls, J.; Cotton, S.; Sassoon, E.; Hall, P. Spectrophotometric intracutaneous analysis in the diagnosis of basal cell carcinoma: A pilot study. Int. J. Dermatol. 2007, 46, 371–375. (In English) [Google Scholar] [CrossRef] [PubMed]
  157. Dasgeb, B.; Kainerstorfer, J.; Mehregan, D.; Vreede, A.V.; Gandjbakhche, A. An introduction to primary skin imaging. Int. J. Dermatol. 2013, 52, 1319–1330. (In English) [Google Scholar] [CrossRef]
  158. Moncrieff, M.; Cotton, S.; Claridge, E.; Hall, P. Spectrophotometric intracutaneous analysis: A new technique for imaging pigmented skin lesions. Br. J. Dermatol. 2002, 146, 448–457. [Google Scholar] [CrossRef] [PubMed]
  159. Matts, P.J.; Dykes, P.J.; Marks, R. The distribution of melanin in skin determined in vivo. Br. J. Dermatol. 2007, 156, 620–628. (In English) [Google Scholar] [CrossRef] [PubMed]
  160. Goessinger, E.V.; Dittrich, P.-G.; Nöcker, P.; Notni, G.; Weber, S.; Cerminara, S.; Mühleisen, B.; Navarini, A.A.; Maul, L.V. Classification of melanocytic lesions using direct illumination multispectral imaging. Sci. Rep. 2024, 14, 19036. [Google Scholar] [CrossRef]
  161. Bekina, A.; Diebele, I.; Rubins, U.; Zaharans, J.; Derjabo, A.; Spigulis, J. Multispectral assessment of skin malformations using a modified video-microscope. Latv. J. Phys. Tech. Sci. 2012, 49, 4–8. Available online: https://sciendo.com/article/10.2478/v10047-012-0024-2 (accessed on 12 January 2025). (In English).
  162. Monheit, G.; Cognetta, A.B.; Ferris, L.; Rabinovitz, H.; Gross, K.; Martini, M.; Grichnik, J.M.; Mihm, M.; Prieto, V.G.; Googe, P.; et al. The performance of MelaFind: A prospective multicenter study. Arch. Dermatol. 2011, 147, 188–194. (In English) [Google Scholar] [CrossRef]
  163. Fink, C.; Jaeger, C.; Jaeger, K.; Haenssle, H.A. Diagnostic performance of the MelaFind device in a real-life clinical setting. JDDG J. Der Dtsch. Dermatol. Ges. 2017, 15, 414–419. [Google Scholar] [CrossRef]
  164. Aloupogianni, E.; Ishikawa, M.; Kobayashi, N.; Obi, T. Hyperspectral and multispectral image processing for gross-level tumor detection in skin lesions: A systematic review. J. Biomed. Opt. 2022, 27, 060901. (In English) [Google Scholar] [CrossRef]
  165. GlobeNewswire News Room (Ed.) STRATA Skin Sciences Announces Licensing Agreement for Certain MelaFind Assets. STRATA Skin Sciences, Inc.: Horsham, PA, USA, 2018.
  166. Voit, C.; Van Akkooi, A.C.; Schäfer-Hesterberg, G.; Schoengen, A.; Kowalczyk, K.; Roewert, J.C.; Sterry, W.; Eggermont, A.M. Ultrasound morphology criteria predict metastatic disease of the sentinel nodes in patients with melanoma. J. Clin. Oncol. 2010, 28, 847–852. [Google Scholar] [CrossRef]
  167. Belfiore, M.P.; Reginelli, A.; Russo, A.; Russo, G.M.; Rocco, M.P.; Moscarella, E.; Ferrante, M.; Sica, A.; Grassi, R.; Cappabianca, S. Usefulness of high-frequency ultrasonography in the diagnosis of melanoma: Mini review. Front. Oncol. 2021, 11, 673026. (In English) [Google Scholar] [CrossRef] [PubMed]
  168. Wortsman, X. Ultrasound in Dermatology: Why, How, and When? Semin. Ultrasound CT MRI 2013, 34, 177–195. [Google Scholar] [CrossRef] [PubMed]
  169. Jasaitiene, D.; Valiukeviciene, S.; Linkeviciute, G.; Raisutis, R.; Jasiuniene, E.; Kazys, R. Principles of high-frequency ultrasonography for investigation of skin pathology. J. Eur. Acad. Dermatol. Venereol. 2011, 25, 375–382. (In English) [Google Scholar] [CrossRef]
  170. Polańska, A.; Dańczak-Pazdrowska, A.; Jałowska, M.; Żaba, R.; Adamski, Z. Current applications of high-frequency ultrasonography in dermatology. Postep. Dermatol. Alergol. 2017, 34, 535–542. (In English) [Google Scholar] [CrossRef] [PubMed]
  171. Wortsman, X.C.; Holm, E.A.; Wulf, H.C.; Jemec, G.B.E. Real-time spatial compound ultrasound imaging of skin. Skin Res. Technol. 2004, 10, 23–31. (In English) [Google Scholar] [CrossRef]
  172. Wortsman, X. Chapter 25—Ultrasound Imaging in Dermatology. In Imaging in Dermatology; University of Chile: Santiago, Chile, 2016; pp. 357–374. [Google Scholar]
  173. Mlosek, R.K.; Malinowska, S. Ultrasound image of the skin, apparatus and imaging basics. J. Ultrason. 2013, 13, 212–221. [Google Scholar] [CrossRef]
  174. Catalano, O. Ultrasound of Cutaneous Melanoma: Primary Tumor Assessment and Locoregional Staging. In Textbook of Dermatologic Ultrasound; Wortsman, X., Ed.; Springer International Publishing: Cham, Switzerland, 2022; pp. 213–249. [Google Scholar]
  175. Kato, M.; Mabuchi, T.; Yamaoka, H.; Ikoma, N.; Tamiya, S.; Ozawa, A.; Taguchi, M.; Kuramochi, A.; Tsuchida, T. Diagnostic usefulness of findings in Doppler sonography for amelanotic melanoma. J. Dermatol. 2013, 40, 700–705. (In English) [Google Scholar] [CrossRef]
  176. Bessoud, B.; Lassau, N.; Koscielny, S.; Longvert, C.; Avril, M.-F.; Duvillard, P.; Rouffiac, V.; Leclère, J.; Roche, A. High-frequency sonography and color Doppler in the management of pigmented skin lesions. Ultrasound Med. Biol. 2003, 29, 875–879. (In English) [Google Scholar] [CrossRef]
  177. Wortsman, X. Common Applications of Dermatologic Sonography. J. Ultrasound Med. 2012, 31, 97–111. (In English) [Google Scholar] [CrossRef]
  178. Reginelli, A.; Belfiore, M.P.; Russo, A.; Turriziani, F.; Moscarella, E.; Troiani, T.; Brancaccio, G.; Ronchi, A.; Giunta, E.; Sica, A. A preliminary study for quantitative assessment with HFUS (High-Frequency Ultrasound) of nodular skin melanoma breslow thickness in adults before surgery: Interdisciplinary team experience. Curr. Radiopharm. 2020, 13, 48–55. [Google Scholar]
  179. Oranges, T.; Janowska, A.; Scatena, C.; Faita, F.; Lascio, N.D.; Izzetti, R.; Fidanzi, C.; Romanelli, M.; Dini, V. Ultra-High Frequency Ultrasound in Melanoma Management: A New Combined Ultrasonographic–Histopathological Approach. J. Ultrasound Med. 2023, 42, 99–108. [Google Scholar] [CrossRef] [PubMed]
  180. Crisan, M.; Crisan, D.; Sannino, G.; Lupsor, M.; Badea, R.; Amzica, F. Ultrasonographic staging of cutaneous malignant tumors: An ultrasonographic depth index. Arch. Dermatol. Res. 2013, 305, 305–313. [Google Scholar] [CrossRef] [PubMed]
  181. Kunte, C.; Schuh, T.; Eberle, J.Y.; Baumert, J.; Konz, B.; Volkenandt, M.; Ruzicka, T.; Schmid-Wendtner, M.-H. The Use of High-Resolution Ultrasonography for Preoperative Detection of Metastases in Sentinel Lymph Nodes of Patients with Cutaneous Melanoma. Dermatol. Surg. 2009, 35, 1757–1765. (In English) [Google Scholar] [CrossRef]
  182. Xing, Y.; Bronstein, Y.; Ross, M.I.; Askew, R.L.; Lee, J.E.; Gershenwald, J.E.; Royal, R.; Cormier, J.N. Contemporary diagnostic imaging modalities for the staging and surveillance of melanoma patients: A meta-analysis. J. Natl. Cancer Inst. 2011, 103, 129–142. (In English) [Google Scholar] [CrossRef]
  183. Ophuis, C.O.; Verhoef, C.; Grünhagen, D.J.; Siegel, P.; Schoengen, A.; Röwert-Huber, J.; Eggermont, A.M.; Voit, C.A.; van Akkooi, A. Long-term results of ultrasound guided fine needle aspiration cytology in conjunction with sentinel node biopsy support step-wise approach in melanoma. Eur. J. Surg. Oncol. (EJSO) 2017, 43, 1509–1516. [Google Scholar] [CrossRef]
  184. Schmid-Wendtner, M.-H.; Dill-Müller, D. Ultrasound Technology in Dermatology. Semin. Cutan. Med. Surg. 2008, 1, 44–51. [Google Scholar] [CrossRef] [PubMed]
  185. Badea, R.; Crisan, M.; Platon, M.L.; Fodor, L. Diagnosis and characterization of cutaneous tumors using combined ultrasonographic procedures (conventional and high resolution ultrasonography). Med. Ultrason. 2010, 12, 317–322. (In English) [Google Scholar]
  186. Reginelli, A.; Russo, A.; Berritto, D.; Patane, V.; Cantisani, C.; Grassi, R. Ultra-High-Frequency Ultrasound: A Modern Diagnostic Technique for Studying Melanoma. Ultraschall Med. 2023, 44, 360–378. (In English) [Google Scholar] [CrossRef]
  187. Wortsman, X.; Wortsman, J. Clinical usefulness of variable-frequency ultrasound in localized lesions of the skin. J. Am. Acad. Dermatol. 2010, 62, 247–256. [Google Scholar] [CrossRef]
  188. Kleinerman, R.; Whang, T.B.; Bard, R.L.; Marmur, E.S. Ultrasound in dermatology: Principles and applications. J. Am. Acad. Dermatol. 2012, 67, 478–487. (In English) [Google Scholar] [CrossRef]
  189. Rajeswari, M.R.; Jain, A.; Sharma, A.; Singh, D.; Jagannathan, N.R.; Sharma, U.; Degaonkar, M.N. Evaluation of Skin Tumors by Magnetic Resonance Imaging. Lab. Investig. 2003, 83, 1279–1283. (In English) [Google Scholar] [CrossRef] [PubMed]
  190. Mirrashed, F.; Sharp, J.C. In vivo morphological characterisation of skin by MRI micro-imaging methods. Skin Res. Technol. 2004, 10, 149–160. (In English) [Google Scholar] [CrossRef] [PubMed]
  191. Foster, P.J.; Dunn, E.A.; Karl, K.E.; Snir, J.A.; Nycz, C.M.; Harvey, A.J.; Pettis, R.J. Cellular Magnetic Resonance Imaging: In Vivo Imaging of Melanoma Cells in Lymph Nodes of Mice. Neoplasia 2008, 10, 207–216. [Google Scholar] [CrossRef]
  192. Hatzoglou, V.; Tisnado, J.; Mehta, A.; Peck, K.K.; Daras, M.; Omuro, A.M.; Beal, K.; Holodny, A.I. Dynamic contrast-enhanced MRI perfusion for differentiating between melanoma and lung cancer brain metastases. Cancer Med. 2017, 6, 761–767. (In English) [Google Scholar] [CrossRef] [PubMed]
  193. Reinert, C.P.; Liang, C.; Weissinger, M.; Vogel, J.; Forschner, A.; Nikolaou, K.; la Fougère, C.; Seith, F. Whole-Body Magnetic Resonance Imaging (MRI) for Staging Melanoma Patients in Direct Comparison to Computed Tomography (CT): Results from a Prospective Positron Emission Tomography (PET)/CT and PET/MRI Study. Diagnostics 2023, 13, 1963. [Google Scholar] [CrossRef]
  194. Lanka, B.; Turner, M.; Orton, C.; Carrington, B.M. Cross-sectional imaging in non-melanoma skin cancer of the head and neck. Clin. Radiol. 2005, 60, 869–877. (In English) [Google Scholar] [CrossRef]
  195. Arnaiz, J.; Gallardo, E.; Piedra, T.; Sanz-Jimenez-Rico, J.R.; Trillo Bohajar, E.; Alonso Pena, D. Giant basal cell carcinoma on the lower leg: MRI findings. J. Plast. Reconstr. Aesthet. Surg. 2007, 60, 1167–1168. (In English) [Google Scholar] [CrossRef]
  196. Hong, H.; Sun, J.; Cai, W. Anatomical and molecular imaging of skin cancer. Clin. Cosmet. Investig. Dermatol. 2008, 1, 1–17. (In English) [Google Scholar]
  197. Jouvet, J.C.; Thomas, L.; Thomson, V.; Yanes, M.; Journe, C.; Morelec, I.; Bracoud, L.; Durupt, F.; Giammarile, F.; Berthezene, Y. Whole-body MRI with diffusion-weighted sequences compared with 18 FDG PET-CT, CT and superficial lymph node ultrasonography in the staging of advanced cutaneous melanoma: A prospective study. J. Eur. Acad. Dermatol. Venereol. 2014, 28, 176–185. (In English) [Google Scholar] [CrossRef]
  198. Jansen, Y.J.L.; Willekens, I.; Seremet, T.; Awada, G.; Schwarze, J.K.; De Mey, J.; Brussaard, C.; Neyns, B. Whole-Body MRI for the Detection of Recurrence in Melanoma Patients at High Risk of Relapse. Cancers 2021, 13, 442. [Google Scholar] [CrossRef]
  199. McManus, D. Metastatic Melanoma; Radiopaedia: Victoria, Australia, 2023. [Google Scholar]
  200. Gniadecka, M.; Philipsen, P.A.; Wessel, S.; Gniadecki, R.; Wulf, H.C.; Sigurdsson, S.; Nielsen, O.F.; Christensen, D.H.; Hercogova, J.; Rossen, K. Melanoma diagnosis by Raman spectroscopy and neural networks: Structure alterations in proteins and lipids in intact cancer tissue. J. Investig. Dermatol. 2004, 122, 443–449. [Google Scholar] [CrossRef] [PubMed]
  201. Colthup, N. Introduction to Infrared and Raman Spectroscopy; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  202. Li, Q.; Liu, C.; Wang, X.; Fan, S. Measuring the thermal conductivity of individual carbon nanotubes by the Raman shift method. Nanotechnology 2009, 20, 145702. (In English) [Google Scholar] [CrossRef] [PubMed]
  203. Haka, A.S.; Volynskaya, Z.I.; Gardecki, J.A.; Nazemi, J.; Shenk, R.; Wang, N.; Dasari, R.R.; Fitzmaurice, M.; Feld, M.S. Diagnosing breast cancer using Raman spectroscopy: Prospective analysis. J. Biomed. Opt. 2009, 14, 054023. [Google Scholar] [CrossRef] [PubMed]
  204. Qiu, X.; Wu, X.; Fang, X.; Fu, Q.; Wang, P.; Wang, X.; Li, S.; Li, Y. Raman spectroscopy combined with deep learning for rapid detection of melanoma at the single cell level. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 286, 122029. [Google Scholar] [CrossRef]
  205. Araújo, D.C.; Veloso, A.A.; de Oliveira Filho, R.S.; Giraud, M.-N.; Raniero, L.J.; Ferreira, L.M.; Bitar, R.A. Finding reduced Raman spectroscopy fingerprint of skin samples for melanoma diagnosis through machine learning. Artif. Intell. Med. 2021, 120, 102161. [Google Scholar] [CrossRef]
  206. Zhang, Y.; Moy, A.J.; Feng, X.; Nguyen, H.T.M.; Sebastian, K.R.; Reichenberg, J.S.; Wilke, C.O.; Markey, M.K.; Tunnell, J.W. Assessment of Raman Spectroscopy for Reducing Unnecessary Biopsies for Melanoma Screening. Molecules 2020, 25, 2852. [Google Scholar] [CrossRef]
  207. Lui, H.; Zhao, J.; McLean, D.; Zeng, H. Real-time Raman spectroscopy for in vivo skin cancer diagnosis. Cancer Res. 2012, 72, 2491–2500. (In English) [Google Scholar] [CrossRef]
  208. Franzen, L.; Windbergs, M. Applications of Raman spectroscopy in skin research—From skin physiology and diagnosis up to risk assessment and dermal drug delivery. Adv. Drug Deliv. Rev. 2015, 89, 91–104. [Google Scholar] [CrossRef]
  209. Santos, I.P.; van Doorn, R.; Caspers, P.J.; Bakker Schut, T.C.; Barroso, E.M.; Nijsten, T.E.C.; Noordhoek Hegt, V.; Koljenović, S.; Puppels, G.J. Improving clinical diagnosis of early-stage cutaneous melanoma based on Raman spectroscopy. Br. J. Cancer 2018, 119, 1339–1346. [Google Scholar] [CrossRef]
  210. Wang, H.; Osseiran, S.; Igras, V.; Nichols, A.J.; Roider, E.M.; Pruessner, J.; Tsao, H.; Fisher, D.E.; Evans, C.L. In vivo coherent Raman imaging of the melanomagenesis-associated pigment pheomelanin. Sci. Rep. 2016, 6, 37986. [Google Scholar] [CrossRef]
  211. Canpolat, M.; Akman-Karakas, A.; Gökhan-Ocak, G.A.; Bassorgun, I.C.; Çiftçioglu, A.M.; Alpsoy, E. Diagnosis and demarcation of skin malignancy using elastic light single-scattering spectroscopy: A pilot study. Dermatol. Surg. 2012, 38, 215–223. [Google Scholar] [CrossRef] [PubMed]
  212. Canpolat, M.; Gökhan, A.G.; Çiftçioğlu, M.A.; Erin, N. Differentiation of melanoma from non-cancerous tissue in an animal model using elastic light single-scattering spectroscopy. Technol. Cancer Res. Treat. 2008, 7, 235–240. [Google Scholar] [CrossRef]
  213. Hartman, R.; Uwe Paasch, M.; Kelly Tepedino, M.; Max Fung, M.; McNiff, J.; Grant-Kels, J. Validation of a Handheld Elastic-Scattering Spectroscopy Device on Lesions Concerning for Melanoma. Available online: https://www.dermasensor.com/wp-content/uploads/80-0013.v1-DA3-v2.pdf (accessed on 12 January 2025).
  214. Hartman, R.I.; Trepanowski, N.; Chang, M.S.; Tepedino, K.; Gianacas, C.; McNiff, J.M.; Fung, M.; Braghiroli, N.F.; Grant-Kels, J.M. Multicenter prospective blinded melanoma detection study with a handheld elastic scattering spectroscopy device. JAAD Int. 2024, 15, 24–31. [Google Scholar] [CrossRef]
  215. Shurrab, K.; Kochaji, N.; Bachir, W. Elastic scattering spectroscopy for monitoring skin cancer transformation and therapy in the near infrared window. Lasers Med. Sci. 2020, 35, 701–708. [Google Scholar] [CrossRef] [PubMed]
  216. Bourgeois, A.C.; Pasciak, A.S.; Bradley, Y.C. The Role of Positron Emission Tomography/Computed Tomography in Cutaneous Melanoma. Imaging Dermatol. 2016, 455–466. (In English) [Google Scholar] [CrossRef]
  217. Holtkamp, L.H.; Chakera, A.H.; Fung, S.; Stretch, J.R.; Saw, R.P.; Lee, K.; Ch’ng, S.; Gonzalez, M.; Thompson, J.F.; Emmett, L. Staging 18F-FDG PET/CT influences the treatment plan in melanoma patients with satellite or in-transit metastases. Melanoma Res. 2020, 30, 358–363. [Google Scholar] [CrossRef] [PubMed]
  218. Sachpekidis, C.; Hassel, J.C.; Kopp-Schneider, A.; Haberkorn, U.; Dimitrakopoulou-Strauss, A. Quantitative Dynamic 18F-FDG PET/CT in Survival Prediction of Metastatic Melanoma under PD-1 Inhibitors. Cancers 2021, 13, 1019. [Google Scholar] [CrossRef]
  219. Ayati, N.; Sadeghi, R.; Kiamanesh, Z.; Lee, S.T.; Zakavi, S.R.; Scott, A.M. The value of 18 F-FDG PET/CT for predicting or monitoring immunotherapy response in patients with metastatic melanoma: A systematic review and meta-analysis. Eur. J. Nucl. Med. Mol. Imaging 2021, 48, 428–448. [Google Scholar] [CrossRef]
  220. Annovazzi, A.; Vari, S.; Giannarelli, D.; Pasqualoni, R.; Sciuto, R.; Carpano, S.; Cognetti, F.; Ferraresi, V. Comparison of 18F-FDG PET/CT Criteria for the Prediction of Therapy Response and Clinical Outcome in Patients with Metastatic Melanoma Treated With Ipilimumab and PD-1 Inhibitors. Clin. Nucl. Med. 2020, 45, 187–194. [Google Scholar] [CrossRef]
  221. Moncrieff, M.; Pywell, S.; Snelling, A.; Gray, M.; Newman, D.; Beadsmoore, C.; Pawaroo, D.; Heaton, M. Effectiveness of SPECT/CT Imaging for Sentinel Node Biopsy Staging of Primary Cutaneous Melanoma and Patient Outcomes. Ann. Surg. Oncol. 2022, 29, 767–775. [Google Scholar] [CrossRef]
  222. Quartuccio, N.; Siracusa, M.; Pappalardo, M.; Arnone, A.; Arnone, G. Sentinel node identification in melanoma: Current clinical impact, new emerging spect radiotracers and technological advancements. An update of the last decade. Curr. Radiopharm. 2020, 13, 32–41. [Google Scholar] [CrossRef]
  223. Rietbergen, D.; Arias-Bouda, L.P.; van der Hage, J.; Olmos, R.V. Does 99mTc-tilmanocept, as next generation radiotracer, meet with the requirements for improved sentinel node imaging? Rev. Española Med. Nucl. E Imagen Mol. 2021, 40, 275–280. [Google Scholar] [CrossRef]
  224. Stathaki, M.I.; Kapsoritakis, N.; Michelakis, D.; Anagnostopoulou, E.; Bourogianni, O.; Tsaroucha, A.; Papadaki, E.; de Bree, E.; Koukouraki, S. The impact of sentinel lymph node mapping with hybrid single photon emission computed tomography/computed tomography in patients with melanoma. Comparison to planar radioisotopic lymphoscintigraphy. Melanoma Res. 2023, 33, 239–246. [Google Scholar] [CrossRef] [PubMed]
  225. Quartuccio, N.; Garau, L.M.; Arnone, A.; Pappalardo, M.; Rubello, D.; Arnone, G.; Manca, G. Comparison of 99mTc-labeled colloid SPECT/CT and planar lymphoscintigraphy in sentinel lymph node detection in patients with melanoma: A meta-analysis. J. Clin. Med. 2020, 9, 1680. [Google Scholar] [CrossRef] [PubMed]
  226. Uren, R.F.; Nieweg, O.E.; Thompson, J.F. Lymphoscintigraphy in Patients with Melanoma. In Cutaneous Melanoma; Balch, C., Gershenwald, J., Thompson, J., Atkins, M., Kirkwood, J., Kefford, R., Sober, A., Halpern, A., Garbe, C., Scolyer, R., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 1–33. [Google Scholar]
  227. Uren, R.; Howman-Giles, R.; Thompson, J.; Shaw, H.; Quinn, M.; O’Brien, C.; McCarthy, W. Lymphoscintigraphy to identify sentinel lymph nodes in patients with melanoma. Melanoma Res. 1994, 4, 395–399. [Google Scholar] [CrossRef]
  228. Moncrieff, M.D.; Thompson, J.F. Evaluating and embracing modern imaging technology to guide sentinel node biopsy for melanoma. Ann. Surg. Oncol. 2022, 29, 5350–5352. [Google Scholar] [CrossRef]
  229. March, J.; Hand, M.; Truong, A.; Grossman, D. Practical application of new technologies for melanoma diagnosis: Part II. Mol. approaches. J. Am. Acad. Dermatol. 2015, 72, 943–958. (In English) [Google Scholar] [CrossRef]
  230. Thomsen, I.M.N.; Heerfordt, I.M.; Karmisholt, K.E.; Mogensen, M. Detection of cutaneous malignant melanoma by tape stripping of pigmented skin lesions—A systematic review. Skin Res. Technol. 2023, 29, e13286. [Google Scholar] [CrossRef] [PubMed]
  231. Ferris, L.K.; Gerami, P.; Skelsey, M.K.; Peck, G.; Hren, C.; Gorman, C.; Frumento, T.; Siegel, D.M. Real-world performance and utility of a noninvasive gene expression assay to evaluate melanoma risk in pigmented lesions. Melanoma Res. 2018, 28, 478. (In English) [Google Scholar] [CrossRef]
  232. Lee, N.; Scope, A.; Rabinovitz, H. Assessing Skin Cancer Using Epidermal Genetic Information Retrieved by Adhesive Patch Skin Surface Sampling. Dermatol. Clin. 2017, 35, 521–524. (In English) [Google Scholar] [CrossRef]
  233. Gerami, P.; Yao, Z.; Polsky, D.; Jansen, B.; Busam, K.; Ho, J.; Martini, M.; Ferris, L.K. Development and validation of a noninvasive 2-gene molecular assay for cutaneous melanoma. J. Am. Acad. Dermatol. 2017, 76, 114–120.e112. [Google Scholar] [CrossRef] [PubMed]
  234. Gerami, P.; Cook, R.W.; Russell, M.C.; Wilkinson, J.; Amaria, R.N.; Gonzalez, R.; Lyle, S.; Jackson, G.L.; Greisinger, A.J.; Johnson, C.E.; et al. Gene expression profiling for molecular staging of cutaneous melanoma in patients undergoing sentinel lymph node biopsy. J. Am. Acad. Dermatol. 2015, 72, 780–785.e783. (In English) [Google Scholar] [CrossRef]
  235. Ferris, L.K.; Farberg, A.S.; Middlebrook, B.; Johnson, C.E.; Lassen, N.; Oelschlager, K.M.; Maetzold, D.J.; Cook, R.W.; Rigel, D.S.; Gerami, P. Identification of high-risk cutaneous melanoma tumors is improved when combining the online American Joint Committee on Cancer Individualized Melanoma Patient Outcome Prediction Tool with a 31-gene expression profile–based classification. J. Am. Acad. Dermatol. 2017, 76, 818–825.e813. [Google Scholar] [CrossRef] [PubMed]
  236. Scaini, M.C.; Catoni, C.; Poggiana, C.; Pigozzo, J.; Piccin, L.; Leone, K.; Scarabello, I.; Facchinetti, A.; Menin, C.; Elefanti, L.; et al. A multiparameter liquid biopsy approach allows to track melanoma dynamics and identify early treatment resistance. npj Precis. Oncol. 2024, 8, 78. [Google Scholar] [CrossRef]
  237. Gerami, P.; Alsobrook, J.P.; Palmer, T.J.; Robin, H.S. Development of a novel noninvasive adhesive patch test for the evaluation of pigmented lesions of the skin. J. Am. Acad. Dermatol. 2014, 71, 237–244. (In English) [Google Scholar] [CrossRef]
  238. Kashani-Sabet, M.; Leachman, S.A.; Stein, J.A.; Arbiser, J.L.; Berry, E.G.; Celebi, J.T.; Curiel-Lewandrowski, C.; Ferris, L.K.; Grant-Kels, J.M.; Grossman, D. Early detection and prognostic assessment of cutaneous melanoma: Consensus on optimal practice and the role of gene expression profile testing. JAMA Dermatol. 2023, 159, 545–553. [Google Scholar] [CrossRef]
  239. Gastman, B.R.; Gerami, P.; Kurley, S.J.; Cook, R.W.; Leachman, S.; Vetto, J.T. Identification of patients at risk of metastasis using a prognostic 31-gene expression profile in subpopulations of melanoma patients with favorable outcomes by standard criteria. J. Am. Acad. Dermatol. 2019, 80, 149–157.e144. [Google Scholar] [CrossRef] [PubMed]
  240. Long, G.V.; Eroglu, Z.; Infante, J.; Patel, S.; Daud, A.; Johnson, D.B.; Gonzalez, R.; Kefford, R.; Hamid, O.; Schuchter, L. Long-term outcomes in patients with BRAF V600–mutant metastatic melanoma who received dabrafenib combined with trametinib. J. Clin. Oncol. 2018, 36, 667–673. [Google Scholar] [CrossRef]
  241. Hodi, F.S.; Chiarion-Sileni, V.; Gonzalez, R.; Grob, J.-J.; Rutkowski, P.; Cowey, C.L.; Lao, C.D.; Schadendorf, D.; Wagstaff, J.; Dummer, R. Nivolumab plus ipilimumab or nivolumab alone versus ipilimumab alone in advanced melanoma (CheckMate 067): 4-year outcomes of a multicentre, randomised, phase 3 trial. Lancet Oncol. 2018, 19, 1480–1492. [Google Scholar] [CrossRef]
  242. Garbe, C.; Amaral, T.; Peris, K.; Hauschild, A.; Arenberger, P.; Basset-Seguin, N.; Bastholt, L.; Bataille, V.; del Marmol, V.; Dréno, B.; et al. European consensus-based interdisciplinary guideline for melanoma. Part 1: Diagnostics: Update 2022. Eur. J. Cancer 2022, 170, 236–255. [Google Scholar] [CrossRef]
  243. The_Skincare_Network. Electrical Impedance Spectroscopy. Available online: https://www.skincarenetwork.co.uk/skin-cancer-treatment/electrical-impedance-spectroscopy/ (accessed on 12 January 2025).
  244. Murali, R.; Thompson, J.F.; Uren, R.F.; Scolyer, R.A. Fine-needle biopsy of metastatic melanoma: Clinical use and new applications. Lancet Oncol. 2010, 11, 391–400. [Google Scholar] [CrossRef] [PubMed]
  245. Cha, J.; Kim, S.; Wang, J.; Yun, M.; Cho, A. Evaluation of (18)F-FDG PET/CT Parameters for Detection of Lymph Node Metastasis in Cutaneous Melanoma. Nucl. Med. Mol. Imaging 2018, 52, 39–45. (In English) [Google Scholar] [CrossRef]
  246. Veeramani, N.; Jayaraman, P. A promising AI based super resolution image reconstruction technique for early diagnosis of skin cancer. Sci. Rep. 2025, 15, 5084. [Google Scholar] [CrossRef]
  247. Waqar, S.; George, S.; Jean-Baptiste, W.; Yusuf Ali, A.; Inyang, B.; Koshy, F.S.; George, K.; Poudel, P.; Chalasani, R.; Goonathilake, M.R.; et al. Recognizing Histopathological Simulators of Melanoma to Avoid Misdiagnosis. Cureus 2022, 14, e26127. (In English) [Google Scholar] [CrossRef]
  248. Rudie, J.D.; Gleason, T.; Barkovich, M.J.; Wilson, D.M.; Shankaranarayanan, A.; Zhang, T.; Wang, L.; Gong, E.; Zaharchuk, G.; Villanueva-Meyer, J.E. Clinical assessment of deep learning–based super-resolution for 3D volumetric brain MRI. Radiol. Artif. Intell. 2022, 4, e210059. [Google Scholar] [CrossRef] [PubMed]
  249. Najjar, R. Redefining Radiology: A Review of Artificial Intelligence Integration in Medical Imaging. Diagnostics 2023, 13, 2760. (In English) [Google Scholar] [CrossRef] [PubMed]
  250. Popescu, D.; El-Khatib, M.; El-Khatib, H.; Ichim, L. New trends in melanoma detection using neural networks: A systematic review. Sensors 2022, 22, 496. [Google Scholar] [CrossRef]
  251. Gajera, H.K.; Nayak, D.R.; Zaveri, M.A. A comprehensive analysis of dermoscopy images for melanoma detection via deep CNN features. Biomed. Signal Process. Control 2023, 79, 104186. [Google Scholar] [CrossRef]
  252. Cirrincione, G.; Cannata, S.; Cicceri, G.; Prinzi, F.; Currieri, T.; Lovino, M.; Militello, C.; Pasero, E.; Vitabile, S. Transformer-based approach to melanoma detection. Sensors 2023, 23, 5677. [Google Scholar] [CrossRef]
  253. Hameed, M.; Zameer, A.; Raja, M.A.Z. A Comprehensive Systematic Review: Advancements in Skin Cancer Classification and Segmentation Using the ISIC Dataset. CMES-Comput. Model. Eng. Sci. 2024, 140, 1–35. [Google Scholar] [CrossRef]
  254. Magalhaes, C.; Mendes, J.; Vardasca, R. Systematic Review of Deep Learning Techniques in Skin Cancer Detection. BioMedInformatics 2024, 4, 2251–2270. [Google Scholar] [CrossRef]
  255. Dildar, M.; Akram, S.; Irfan, M.; Khan, H.U.; Ramzan, M.; Mahmood, A.R.; Alsaiari, S.A.; Saeed, A.H.M.; Alraddadi, M.O.; Mahnashi, M.H. Skin Cancer Detection: A Review Using Deep Learning Techniques. Int. J. Environ. Res. Public Health 2021, 18, 5479. [Google Scholar] [CrossRef]
  256. Grant, S.R.; Andrew, T.W.; Alvarez, E.V.; Huss, W.J.; Paragh, G. Diagnostic and Prognostic Deep Learning Applications for Histological Assessment of Cutaneous Melanoma. Cancers 2022, 14, 6231. [Google Scholar] [CrossRef] [PubMed]
  257. Liu, Y.; Primiero, C.A.; Kulkarni, V.; Soyer, H.P.; Betz-Stablein, B. Artificial intelligence for the classification of pigmented skin lesions in populations with skin of color: A systematic review. Dermatology 2023, 239, 499–513. [Google Scholar] [CrossRef] [PubMed]
  258. Durot, C.; Mulé, S.; Soyer, P.; Marchal, A.; Grange, F.; Hoeffel, C. Metastatic melanoma: Pretreatment contrast-enhanced CT texture parameters as predictive biomarkers of survival in patients treated with pembrolizumab. Eur. Radiol. 2019, 29, 3183–3191. [Google Scholar] [CrossRef]
  259. Guerrisi, A.; Russillo, M.; Loi, E.; Ganeshan, B.; Ungania, S.; Desiderio, F.; Bruzzaniti, V.; Falcone, I.; Renna, D.; Ferraresi, V. Exploring CT texture parameters as predictive and response imaging biomarkers of survival in patients with metastatic melanoma treated with PD-1 inhibitor nivolumab: A pilot study using a delta-radiomics approach. Front. Oncol. 2021, 11, 704607. [Google Scholar] [CrossRef]
  260. Gill, A.B.; Rundo, L.; Wan, J.C.M.; Lau, D.; Zawaideh, J.P.; Woitek, R.; Zaccagna, F.; Beer, L.; Gale, D.; Sala, E.; et al. Correlating Radiomic Features of Heterogeneity on CT with Circulating Tumor DNA in Metastatic Melanoma. Cancers 2020, 12, 3493. [Google Scholar] [CrossRef]
  261. Tang, P.; Yan, X.; Nan, Y.; Hu, X.; Menze, B.H.; Krammer, S.; Lasser, T. Joint-individual fusion structure with fusion attention module for multi-modal skin cancer classification. Pattern Recognit. 2024, 154, 110604. [Google Scholar] [CrossRef]
  262. Chen, Q.; Li, M.; Chen, C.; Zhou, P.; Lv, X.; Chen, C. MDFNet: Application of multimodal fusion method based on skin image and clinical data to skin cancer classification. J. Cancer Res. Clin. Oncol. 2023, 149, 3287–3299. [Google Scholar] [CrossRef]
  263. Rey-Barroso, L.; Vilaseca, M.; Royo, S.; Díaz-Doutón, F.; Lihacova, I.; Bondarenko, A.; Burgos-Fernández, F.J. Training State-of-the-Art Deep Learning Algorithms with Visible and Extended Near-Infrared Multispectral Images of Skin Lesions for the Improvement of Skin Cancer Diagnosis. Diagnostics 2025, 15, 355. [Google Scholar] [CrossRef]
  264. Huang, H.-Y.; Hsiao, Y.-P.; Karmakar, R.; Mukundan, A.; Chaudhary, P.; Hsieh, S.-C.; Wang, H.-C. A Review of Recent Advances in Computer-Aided Detection Methods Using Hyperspectral Imaging Engineering to Detect Skin Cancer. Cancers 2023, 15, 5634. [Google Scholar] [CrossRef] [PubMed]
  265. Bsharat, S.M.; Abouelnour, S.; Ahmed, R.; Elkhatib, M.; Gaber, S.; Shehieb, W.; Arshad, K.; Assaleh, K. RGB-to-hyperspectral conversion for accessible melanoma detection: A CNN-based approach. J. Intell. Syst. 2024, 33, 20230271. [Google Scholar] [CrossRef]
  266. Zhou, Y.; Wang, L.V. Chapter 24—Photoacoustic Tomography in the Diagnosis of Melanoma. In Imaging in Dermatology; Hamblin, M.R., Avci, P., Gupta, G.K., Hamblin, M.R., Avci, P., Gupta, G.K., Eds.; Academic Press: Boston, MA, USA, 2016; pp. 341–356. [Google Scholar]
  267. Ku, G.; Wang, L.V. Deeply penetrating photoacoustic tomography in biological tissues enhanced with an optical contrast agent. Opt. Lett. 2005, 30, 507–509. (In English) [Google Scholar] [CrossRef]
  268. Song, K.H.; Stein, E.W.; Margenthaler, J.A.; Wang, L.V. Noninvasive photoacoustic identification of sentinel lymph nodes containing methylene blue in vivo in a rat model. J. Biomed. Opt. 2008, 13, 054033. [Google Scholar] [CrossRef]
  269. Yao, J.; Maslov, K.I.; Hu, S.; Wang, L.V. Evans blue dye-enhanced capillary-resolution photoacoustic microscopy in vivo. J. Biomed. Opt. 2009, 14, 054049. [Google Scholar] [CrossRef]
  270. Kim, C.; Favazza, C.; Wang, L.V. In Vivo Photoacoustic Tomography of Chemicals: High-Resolution Functional and Molecular Optical Imaging at New Depths. Chem. Rev. 2010, 110, 2756–2782. [Google Scholar] [CrossRef] [PubMed]
  271. Kim, C.; Cho, E.C.; Chen, J.; Song, K.H.; Au, L.; Favazza, C.; Zhang, Q.; Cobley, C.M.; Gao, F.; Xia, Y. In vivo molecular photoacoustic tomography of melanomas targeted by bioconjugated gold nanocages. ACS Nano 2010, 4, 4559–4564. [Google Scholar] [CrossRef]
  272. Cai, X.; Li, W.; Kim, C.-H.; Yuan, Y.; Wang, L.V.; Xia, Y. In Vivo Quantitative Evaluation of the Transport Kinetics of Gold Nanocages in a Lymphatic System by Noninvasive Photoacoustic Tomography. ACS Nano 2011, 5, 9658–9667. [Google Scholar] [CrossRef]
  273. Wang, X.; Pang, Y.; Ku, G.; Xie, X.; Stoica, G.; Wang, L.V. Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain. Nat. Biotechnol. 2003, 21, 803–806. (In English) [Google Scholar] [CrossRef]
  274. Zhang, H.F.; Maslov, K.; Stoica, G.; Wang, L.V. Functional photoacoustic microscopy for high-resolution and noninvasive in vivo imaging. Nat. Biotechnol. 2006, 24, 848. [Google Scholar] [CrossRef]
  275. Oh, J.-T.; Li, M.-L.; Zhang, H.F.; Maslov, K.; Stoica, G.; Wang, L.V. Three-dimensional imaging of skin melanoma in vivo by dual-wavelength photoacoustic microscopy. J. Biomed. Opt. 2006, 11, 034032. [Google Scholar] [CrossRef] [PubMed]
  276. Xu, M.; Wang, L.V. Photoacoustic imaging in biomedicine. Rev. Sci. Instrum. 2006, 77, 041101. [Google Scholar] [CrossRef]
  277. Weight, R.M.; Viator, J.A.; Dale, P.S.; Caldwell, C.W.; Lisle, A.E. Photoacoustic detection of metastatic melanoma cells in the human circulatory system. Opt. Lett. 2006, 31, 2998–3000. [Google Scholar] [CrossRef]
  278. Zharov, V.P.; Galanzha, E.I.; Shashkov, E.V.; Khlebtsov, N.G.; Tuchin, V.V. In vivo photoacoustic flow cytometry for monitoring of circulating single cancer cells and contrast agents. Opt. Lett. 2006, 31, 3623–3625. [Google Scholar] [CrossRef]
  279. Holan, S.H.; Viator, J.A. Automated wavelet denoising of photoacoustic signals for circulating melanoma cell detection and burn image reconstruction. Phys. Med. Biol. 2008, 53, N227. [Google Scholar] [CrossRef] [PubMed]
  280. Pramanik, M.; Song, K.H.; Swierczewska, M.; Green, D.; Sitharaman, B.; Wang, L.V. In vivo carbon nanotube-enhanced non-invasive photoacoustic mapping of the sentinel lymph node. Phys. Med. Biol. 2009, 54, 3291. (In English) [Google Scholar] [CrossRef]
  281. Song, K.H.; Kim, C.; Cobley, C.M.; Xia, Y.; Wang, L.V. Near-Infrared Gold Nanocages as a New Class of Tracers for Photoacoustic Sentinel Lymph Node Mapping on a Rat Model. Nano Lett. 2009, 9, 183–188. [Google Scholar] [CrossRef]
  282. Kim, C.; Erpelding, T.N.; Maslov, K.I.; Jankovic, L.; Akers, W.J.; Song, L.; Achilefu, S.; Margenthaler, J.A.; Pashley, M.D.; Wang, L.V.J.J.o.b.o. Handheld array-based photoacoustic probe for guiding needle biopsy of sentinel lymph nodes. J. Biomed. Opt. 2010, 15, 046010. [Google Scholar] [CrossRef] [PubMed]
  283. Pan, D.; Cai, X.; Yalaz, C.; Senpan, A.; Omanakuttan, K.; Wickline, S.A.; Wang, L.V.; Lanza, G.M. Photoacoustic Sentinel Lymph Node Imaging with Self-Assembled Copper Neodecanoate Nanoparticles. ACS Nano 2012, 6, 1260–1267. [Google Scholar] [CrossRef]
  284. Zhou, Y.; Tripathi, S.V.; Rosman, I.; Ma, J.; Hai, P.; Linette, G.P.; Council, M.L.; Fields, R.C.; Wang, L.V.; Cornelius, L.A. Noninvasive determination of melanoma depth using a handheld photoacoustic probe. J. Investig. Dermatol. 2017, 137, 1370–1372. [Google Scholar] [CrossRef]
  285. Stoffels, I.; Jansen, P.; Petri, M.; Goerdt, L.; Brinker, T.J.; Griewank, K.G.; Poeppel, T.D.; Schadendorf, D.; Klode, J. Assessment of nonradioactive multispectral optoacoustic tomographic imaging with conventional lymphoscintigraphic imaging for sentinel lymph node biopsy in melanoma. JAMA Netw. Open 2019, 2, e199020. [Google Scholar] [CrossRef]
  286. Vonk, J.; Kukačka, J.; Steinkamp, P.; De Wit, J.; Voskuil, F.; Hooghiemstra, W.; Bader, M.; Jüstel, D.; Ntziachristos, V.; Van Dam, G. Multispectral optoacoustic tomography for in vivo detection of lymph node metastases in oral cancer patients using an EGFR-targeted contrast agent and intrinsic tissue contrast: A proof-of-concept study. Photoacoustics 2022, 26, 100362. [Google Scholar] [CrossRef] [PubMed]
  287. Grootendorst, D.J.; Jose, J.; Wouters, M.W.; van Boven, H.; Van der Hage, J.; Van Leeuwen, T.G.; Steenbergen, W.; Manohar, S.; Ruers, T.J. First experiences of photoacoustic imaging for detection of melanoma metastases in resected human lymph nodes. Lasers Surg. Med. 2012, 44, 541–549. [Google Scholar] [CrossRef] [PubMed]
  288. Langhout, G.C.; Grootendorst, D.J.; Nieweg, O.E.; Wouters, M.W.; van der Hage, J.A.; Jose, J.; van Boven, H.; Steenbergen, W.; Manohar, S.; Ruers, T.J. Detection of melanoma metastases in resected human lymph nodes by noninvasive multispectral photoacoustic imaging. Int. J. Biomed. Imaging 2014, 2014, 163652. [Google Scholar] [CrossRef]
  289. Martell, M.T.; Haven, N.J.M.; Cikaluk, B.D.; Restall, B.S.; McAlister, E.A.; Mittal, R.; Adam, B.A.; Giannakopoulos, N.; Peiris, L.; Silverman, S.; et al. Deep learning-enabled realistic virtual histology with ultraviolet photoacoustic remote sensing microscopy. Nat. Commun. 2023, 14, 5967. [Google Scholar] [CrossRef]
  290. Lassau, N.; Lamuraglia, M.; Koscielny, S.; Spatz, A.; Roche, A.; Leclere, J.; Avril, M.-F. Prognostic value of angiogenesis evaluated with high-frequency and colour Doppler sonography for preoperative assessment of primary cutaneous melanomas: Correlation with recurrence after a 5 year follow-up period. Cancer Imaging 2006, 6, 24. [Google Scholar] [CrossRef] [PubMed]
  291. Galanzha, E.I.; Menyaev, Y.A.; Yadem, A.C.; Sarimollaoglu, M.; Juratli, M.A.; Nedosekin, D.A.; Foster, S.R.; Jamshidi-Parsian, A.; Siegel, E.R.; Makhoul, I. In vivo liquid biopsy using Cytophone platform for photoacoustic detection of circulating tumor cells in patients with melanoma. Sci. Transl. Med. 2019, 11, eaat5857. [Google Scholar] [CrossRef]
  292. Hai, P.; Qu, Y.; Li, Y.; Zhu, L.; Shmuylovich, L.; Cornelius, L.A.; Wang, L.V. Label-free high-throughput photoacoustic tomography of suspected circulating melanoma tumor cells in patients in vivo. J. Biomed. Opt. 2020, 25, 036002. [Google Scholar] [CrossRef]
  293. Viator, J.A.; Hazur, M.; Sajewski, A.; Tarhini, A.; Sanders, M.E.; Edgar, R.H. Photoacoustic detection of circulating melanoma cells in late stage patients. J. Innov. Opt. Health Sci. 2020, 13, 2050023. [Google Scholar] [CrossRef]
  294. Gutierrez-Juarez, G.; Gupta, S.; Weight, R.M.; Polo-Parada, L.; Papagiorgio, C.; Bunch, J.; Viator, J. Optical photoacoustic detection of circulating melanoma cells in vitro. Int. J. Thermophys. 2010, 31, 784–792. [Google Scholar] [CrossRef]
  295. Benjamin, R.E.; James, T.; Krystal, L.; Hager, G.; Kevan, B.; Deepak, D.; Muba, T.; John, M.; Parsin Haji, R. Exploration of photoacoustic remote sensing virtual histology in skin malignancies. In Proceedings of the Photonics in Dermatology and Plastic Surgery, San Francisco, CA, USA, 28–29 January 2023; SPIE: Bellingham, WA, USA, 2023; Volume PC12352, p. PC1235201. [Google Scholar]
  296. Ecclestone, B.; Bell, K.; Boktor, M.; Pekar, V.; Taher, M.; Dinakaran, D.; Mackey, J.; Haji Reza, P. Label-free histological imaging of skin cancers with photoacoustic remote sensing virtual histology. In Proceedings of the Photonics in Dermatology and Plastic Surgery, San Francisco, CA, USA, 22–23 January 2022; SPIE: Bellingham, WA, USA, 2022. [Google Scholar]
  297. Kratkiewicz, K.; Manwar, R.; Rajabi-Estarabadi, A.; Fakhoury, J.; Meiliute, J.; Daveluy, S.; Mehregan, D.; Avanaki, K.M. Photoacoustic/Ultrasound/Optical Coherence Tomography Evaluation of Melanoma Lesion and Healthy Skin in a Swine Model. Sensors 2019, 19, 2815. [Google Scholar] [CrossRef]
  298. Abbasi, S.; Le, M.; Sonier, B.; Dinakaran, D.; Bigras, G.; Bell, K.; Mackey, J.R.; Reza, P.H. All-optical reflection-mode microscopic histology of unstained human tissues. Sci. Rep. 2019, 9, 13392. [Google Scholar] [CrossRef] [PubMed]
  299. Breathnach, A.; Concannon, E.; Dorairaj, J.J.; Shaharan, S.; McGrath, J.; Jose, J.; Kelly, J.L.; Leahy, M.J. Preoperative measurement of cutaneous melanoma and nevi thickness with photoacoustic imaging. J. Med. Imaging 2018, 5, 015004. [Google Scholar] [CrossRef]
  300. Zhou, Y.; Li, G.; Zhu, L.; Li, C.; Cornelius, L.A.; Wang, L.V. Handheld photoacoustic probe to detect both melanoma depth and volume at high speed in vivo. J. Biophotonics 2015, 8, 961–967. [Google Scholar] [CrossRef] [PubMed]
  301. Haji Reza, P.; Shi, W.; Bell, K.; Paproski, R.J.; Zemp, R.J. Non-interferometric photoacoustic remote sensing microscopy. Light. Sci. Appl. 2017, 6, e16278. [Google Scholar] [CrossRef]
  302. Bell, K.; Abbasi, S.; Dinakaran, D.; Taher, M.; Bigras, G.; van Landeghem, F.K.; Mackey, J.R.; Haji Reza, P. Reflection-mode virtual histology using photoacoustic remote sensing microscopy. Sci. Rep. 2020, 10, 19121. [Google Scholar] [CrossRef]
  303. Reza, P.H.; Bell, K.; Shi, W.; Shapiro, J.; Zemp, R.J. Deep non-contact photoacoustic initial pressure imaging. Optica 2018, 5, 814–820. [Google Scholar] [CrossRef]
  304. Hosking, A.M.; Coakley, B.J.; Chang, D.; Talebi-Liasi, F.; Lish, S.; Lee, S.W.; Zong, A.M.; Moore, I.; Browning, J.; Jacques, S.L. Hyperspectral imaging in automated digital dermoscopy screening for melanoma. Lasers Surg. Med. 2019, 51, 214–222. [Google Scholar] [CrossRef] [PubMed]
  305. Vasefi, F.; MacKinnon, N.; Farkas, D.L. Chapter 16—Hyperspectral and Multispectral Imaging in Dermatology. In Imaging in Dermatology; Hamblin, M.R., Avci, P., Gupta, G.K., Hamblin, M.R., Avci, P., Gupta, G.K., Eds.; Academic Press: Boston, MA, USA, 2016; pp. 187–201. [Google Scholar]
  306. Vasefi, F.; MacKinnon, N.; Farkas, D.L. Toward in vivo diagnosis of skin cancer using multimode imaging dermoscopy: (II) molecular mapping of highly pigmented lesions. In Proceedings of the Imaging, Manipulation, and Analysis of Biomolecules, Cells, and Tissues XII, San Francisco, CA, USA, 3–6 February 2014; Volume 8947, p. 89470J. [Google Scholar] [CrossRef]
  307. Vasefi, F.; MacKinnon, N.; Saager, R.; Kelly, K.M.; Maly, T.; Booth, N.; Durkin, A.J.; Farkas, D.L. Separating melanin from hemodynamics in nevi using multimode hyperspectral dermoscopy and spatial frequency domain spectroscopy. J. Biomed. Opt. 2016, 21, 114001. (In English) [Google Scholar] [CrossRef]
  308. Paoli, J.; Pölönen, I.; Salmivuori, M.; Räsänen, J.; Zaar, O.; Polesie, S.; Koskenmies, S.; Pitkänen, S.; Övermark, M.; Isoherranen, K.; et al. Hyperspectral Imaging for Non-invasive Diagnostics of Melanocytic Lesions. Acta Derm. Venereol. 2022, 102, adv00815. (In English) [Google Scholar] [CrossRef]
  309. Tian, C.; Xu, Y.; Zhang, Y.; Zhang, Z.; An, H.; Liu, Y.; Chen, Y.; Zhao, H.; Zhang, Z.; Zhao, Q. Combining hyperspectral imaging techniques with deep learning to aid in early pathological diagnosis of melanoma. Photodiagnosis Photodyn. Ther. 2023, 43, 103708. [Google Scholar] [CrossRef]
  310. Ng, P.C.; Chi, Z.; Verdie, Y.; Lu, J.; Plataniotis, K.N. Hyper-Skin: A Hyperspectral Dataset for Reconstructing Facial Skin-Spectra from RGB Images. Adv. Neural Inf. Process. Syst. 2024, 36, 24158–24170. [Google Scholar]
  311. Lin, T.-L.; Lu, C.-T.; Karmakar, R.; Nampalley, K.; Mukundan, A.; Hsiao, Y.-P.; Hsieh, S.-C.; Wang, H.-C. Assessing the efficacy of the spectrum-aided vision enhancer (SAVE) to detect acral lentiginous melanoma, melanoma in situ, nodular melanoma, and superficial spreading melanoma. Diagnostics 2024, 14, 1672. [Google Scholar] [CrossRef] [PubMed]
  312. MacKinnon, N.; Vasefi, F.; Booth, N.; Farkas, D.L. Melanoma detection using smartphone and multimode hyperspectral imaging. In Proceedings of the Imaging, Manipulation, and Analysis of Biomolecules, Cells, and Tissues IX, San Francisco, CA, USA, 15–17 February 2016; pp. 143–150. [Google Scholar]
  313. Anbar, M. Clinical thermal imaging today. IEEE Eng. Med. Biol. Mag. 1998, 17, 25–33. [Google Scholar] [CrossRef]
  314. Jones, B.F. A reappraisal of the use of infrared thermal image analysis in medicine. IEEE Trans. Med. Imaging 1998, 17, 1019–1027. [Google Scholar] [CrossRef]
  315. Çetingül, M.P.; Herman, C. A heat transfer model of skin tissue for the detection of lesions: Sensitivity analysis. Phys. Med. Biol. 2010, 55, 5933. [Google Scholar] [CrossRef] [PubMed]
  316. Akhter, N.; Manza, R.; Shaikh, S.; Gawali, B.; Yannawar, P.; Shaikh, S. Diagnosis of Melanoma Using Thermography: A Review. In Proceedings of the International Conference on Applications of Machine Intelligence and Data Analytics (ICAMIDA 2022), Aurangabad, India, 22–24 December 2022; p. 466. [Google Scholar]
  317. Shada, A.L.; Dengel, L.T.; Petroni, G.R.; Smolkin, M.E.; Acton, S.; Slingluff, C.L. Infrared thermography of cutaneous melanoma metastases. J. Surg. Res. 2013, 182, e9–e14. (In English) [Google Scholar] [CrossRef]
  318. Godoy, S.E.; Hayat, M.M.; Ramirez, D.A.; Myers, S.A.; Padilla, R.S.; Krishna, S. Detection theory for accurate and non-invasive skin cancer diagnosis using dynamic thermal imaging. Biomed. Opt. Express 2017, 8, 2301–2323. (In English) [Google Scholar] [CrossRef]
  319. Díaz, S.; Krohmer, T.; Moreira, A.; Godoy, S.E.; Figueroa, M. An instrument for accurate and non-invasive screening of skin cancer based on multimodal imaging. IEEE Access 2019, 7, 176646–176657. [Google Scholar] [CrossRef]
  320. 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. (In English) [Google Scholar] [CrossRef]
  321. Woodward, R.M.; Cole, B.E.; Wallace, V.P.; Pye, R.J.; Arnone, D.D.; Linfield, E.H.; Pepper, M. Terahertz pulse imaging in reflection geometry of human skin cancer and skin tissue. Phys. Med. Biol. 2002, 47, 3853–3863. (In English) [Google Scholar] [CrossRef] [PubMed]
  322. Vilagosh, Z.; Lajevardipour, A.; Wood, A.W. Computational phantom study of frozen melanoma imaging at 0.45 terahertz. Bioelectromagnetics 2019, 40, 118–127. (In English) [Google Scholar] [CrossRef] [PubMed]
  323. Sim, Y.C.; Ahn, K.-M.; Park, J.Y.; Park, C.-S.; Son, J.-H. Temperature-dependent terahertz imaging of excised oral malignant melanoma. IEEE J. Biomed. Health Info. 2013, 17, 779–784. (In English) [Google Scholar] [CrossRef]
  324. Wallace, V.P.; Fitzgerald, A.J.; Shankar, S.; Flanagan, N.; Pye, R.; Cluff, J.; Arnone, D.D. Terahertz pulsed imaging of basal cell carcinoma ex vivo and in vivo. Br. J. Dermatol. 2004, 151, 424–432. (In English) [Google Scholar] [CrossRef]
  325. Zaytsev, K.I.; Kudrin, K.G.; Karasik, V.E.; Reshetov, I.V.; Yurchenko, S.O. In vivo terahertz spectroscopy of pigmentary skin nevi: Pilot study of non-invasive early diagnosis of dysplasia. Appl. Phys. Lett. 2015, 106, 053702. [Google Scholar] [CrossRef]
  326. Li, D.; Yang, Z.; Fu, A.; Chen, T.; Chen, L.; Tang, M.; Zhang, H.; Mu, N.; Wang, S.; Liang, G. Detecting melanoma with a terahertz spectroscopy imaging technique. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 234, 118229. [Google Scholar] [CrossRef]
  327. Qi, X.; Bertling, K.; Torniainen, J.; Kong, F.; Gillespie, T.; Primiero, C.; Stark, M.S.; Dean, P.; Indjin, D.; Li, L.H.; et al. Terahertz in vivo imaging of human skin: Toward detection of abnormal skin pathologies. APL Bioeng. 2024, 8, 016117. [Google Scholar] [CrossRef]
  328. Zipfel, W.R.; Williams, R.M.; Christie, R.; Nikitin, A.Y.; Hyman, B.T.; Webb, W.W. Live tissue intrinsic emission microscopy using multiphoton-excited native fluorescence and second harmonic generation. Proc. Natl. Acad. Sci. USA 2003, 100, 7075–7080. (In English) [Google Scholar] [CrossRef]
  329. Dimitrow, E.; Ziemer, M.; Koehler, M.J.; Norgauer, J.; König, K.; Elsner, P.; Kaatz, M. Sensitivity and specificity of multiphoton laser tomography for in vivo and ex vivo diagnosis of malignant melanoma. J. Investig. Dermatol. 2009, 129, 1752–1758. (In English) [Google Scholar] [CrossRef]
  330. Seidenari, S.; Arginelli, F.; Dunsby, C.; French, P.M.W.; König, K.; Magnoni, C.; Talbot, C.; Ponti, G. Multiphoton Laser Tomography and Fluorescence Lifetime Imaging of Melanoma: Morphologic Features and Quantitative Data for Sensitive and Specific Non-Invasive Diagnostics. PLoS ONE 2013, 8, e70682. (In English) [Google Scholar] [CrossRef]
  331. MPTflex-CARS; Multiphoton Tomography: Berlin, Germany, 2016.
  332. Chernyavskiy, O.; Vannucci, L.; Bianchini, P.; Difato, F.; Saieh, M.; Kubínová, L. Imaging of mouse experimental melanoma in vivo and ex vivo by combination of confocal and nonlinear microscopy. Microsc. Res. Tech. 2009, 72, 411–423. (In English) [Google Scholar] [CrossRef] [PubMed]
  333. Elagin, V.; Gubarkova, E.; Garanina, O.; Davydova, D.; Orlinskaya, N.; Matveev, L.; Klemenova, I.; Shlivko, I.; Shirmanova, M.; Zagaynova, E. In vivo multimodal optical imaging of dermoscopic equivocal melanocytic skin lesions. Sci. Rep. 2021, 11, 1405. [Google Scholar] [CrossRef] [PubMed]
  334. Johnson, T.; Imaging Technique Detects Skin Cancer Without Invasive Biopsy. Tech. Briefs 2019. Available online: https://www.techbriefs.com/component/content/article/33583-imaging-technique-detects-skin-cancer-without-invasive-biopsy (accessed on 12 January 2025).
  335. Heaton, A.R.; Burkard, N.J.; Sondel, P.M.; Skala, M.C. Quantifying in vivo collagen reorganization during immunotherapy in murine melanoma with second harmonic generation imaging. Biophotonics Discov. 2024, 1, 015004. [Google Scholar] [CrossRef]
  336. Zheng, Z.; Liu, H.; Zhai, S.; Zhang, H.; Shan, G.; Kwok, R.T.; Ma, C.; Sung, H.H.; Williams, I.D.; Lam, J.W. Highly efficient singlet oxygen generation, two-photon photodynamic therapy and melanoma ablation by rationally designed mitochondria-specific near-infrared AIEgens. Chem. Sci. 2020, 11, 2494–2503. [Google Scholar] [CrossRef]
  337. James, V.J. Fiber diffraction of skin and nails provides an accurate diagnosis of malignancies. Int. J. Cancer 2009, 125, 133–138. (In English) [Google Scholar] [CrossRef] [PubMed]
  338. James, V.J.; Kirby, N. The Connection between the Presence of Melanoma and Changes in Fibre Diffraction Patterns. Cancers 2010, 2, 1155–1165. (In English) [Google Scholar] [CrossRef]
  339. James, V.J.; O’Malley Ford, J.M. Fibre Diffraction Analysis of Skin Offers a Very Early and Extremely Accurate Diagnostic Test for Prostate Cancer. J. Cancer Res. 2014, 2014, 179608. [Google Scholar] [CrossRef]
  340. Baker, M.J.; Trevisan, J.; Bassan, P.; Bhargava, R.; Butler, H.J.; Dorling, K.M.; Fielden, P.R.; Fogarty, S.W.; Fullwood, N.J.; Heys, K.A.; et al. Using Fourier transform IR spectroscopy to analyze biological materials. Nat. Protoc. 2014, 9, 1771–1791. (In English) [Google Scholar] [CrossRef]
  341. Ghimire, H.; Venkataramani, M.; Bian, Z.; Liu, Y.; Perera, A.G.U. ATR-FTIR spectral discrimination between normal and tumorous mouse models of lymphoma and melanoma from serum samples. Sci. Rep. 2017, 7, 16993. (In English) [Google Scholar] [CrossRef]
  342. Ollesch, J.; Heinze, M.; Heise, H.M.; Behrens, T.; Brüning, T.; Gerwert, K. It’s in your blood: Spectral biomarker candidates for urinary bladder cancer from automated FTIR spectroscopy. J. Biophotonics 2014, 7, 210–221. (In English) [Google Scholar] [CrossRef]
  343. Tosi, G.; Conti, C.; Giorgini, E.; Ferraris, P.; Garavaglia, M.G.; Sabbatini, S.; Staibano, S.; Rubini, C. FTIR microspectroscopy of melanocytic skin lesions: A preliminary study. Analyst 2010, 135, 3213–3219. [Google Scholar] [CrossRef]
  344. Wald, N.; Le Corre, Y.; Martin, L.; Mathieu, V.; Goormaghtigh, E. Infrared spectra of primary melanomas can predict response to chemotherapy: The example of dacarbazine. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2016, 1862, 174–181. [Google Scholar] [CrossRef]
  345. Brézillon, S.; Untereiner, V.; Mohamed, H.T.; Ahallal, E.; Proult, I.; Nizet, P.; Boulagnon-Rombi, C.; Sockalingum, G.D. Label-free infrared spectral histology of skin tissue part II: Impact of a lumican-derived peptide on melanoma growth. Front. Cell Dev. Biol. 2020, 8, 377. [Google Scholar] [CrossRef] [PubMed]
  346. Abuh, S.O.; Barbora, A.; Minnes, R. Metastasis diagnosis using attenuated total reflection-Fourier transform infra-red (ATR-FTIR) spectroscopy. PLoS ONE 2024, 19, e0304071. [Google Scholar] [CrossRef] [PubMed]
  347. Hinz, T.; Wenzel, J.; Schmid-Wendtner, M.-H. Real-time tissue elastography: A helpful tool in the diagnosis of cutaneous melanoma? J. Am. Acad. Dermatol. 2011, 65, 424–426. (In English) [Google Scholar] [CrossRef] [PubMed]
  348. Botar-Jid, C.M.; Cosgarea, R.; Bolboacă, S.D.; Şenilă, S.C.; Lenghel, L.M.; Rogojan, L.; Dudea, S.M. Assessment of Cutaneous Melanoma by Use of Very- High-Frequency Ultrasound and Real-Time Elastography. Am. J. Roentgenol. 2016, 206, 699–704. (In English) [Google Scholar] [CrossRef]
  349. Ogata, D.; Uematsu, T.; Yoshikawa, S.; Kiyohara, Y. Accuracy of real-time ultrasound elastography in the differential diagnosis of lymph nodes in cutaneous malignant melanoma (CMM): A pilot study. Int. J. Clin. Oncol. 2014, 19, 716–721. (In English) [Google Scholar] [CrossRef]
  350. Ying, L.; Hou, Y.; Zheng, H.-M.; Lin, X.; Xie, Z.-L.; Hu, Y.-P. Real-time elastography for the differentiation of benign and malignant superficial lymph nodes: A meta-analysis. Eur. J. Radiol. 2012, 81, 2576–2584. (In English) [Google Scholar] [CrossRef]
  351. Dasgeb, B.; Morris, M.A.; Mehregan, D.; Siegel, E.L. Quantified ultrasound elastography in the assessment of cutaneous carcinoma. Br. J. Radiol. 2015, 88, 20150344. [Google Scholar] [CrossRef]
  352. Blois, M.S.; Zahlan, A.B.; Maling, J.E. Electron spin resonance studies on melanin. Biophys. J. 1964, 4, 471–490. (In English) [Google Scholar] [CrossRef]
  353. Katsuda, H.; Kobayashi, T.; Saito, H.; Matsunaga, T.; Ikeya, M. Electron spin resonance imaging of mouse B16 melanoma. Chem. Pharm. Bull. 1990, 38, 2838–2840. [Google Scholar] [CrossRef]
  354. Wehbi, M.; Harkemanne, E.; Mignion, L.; Joudiou, N.; Tromme, I.; Baurain, J.F.; Gallez, B. Towards Characterization of Skin Melanoma in the Clinic by Electron Paramagnetic Resonance (EPR) Spectroscopy and Imaging of Melanin. Mol. Imaging Biol. 2023, 26, 382–390. (In English) [Google Scholar] [CrossRef]
  355. Godechal, Q.; Ghanem, G.E.; Cook, M.G.; Gallez, B. Electron Paramagnetic Resonance Spectrometry and Imaging in Melanomas: Comparison between Pigmented and Nonpigmented Human Malignant Melanomas. Mol. Imaging 2013, 12, 7290.2012.00037. [Google Scholar] [CrossRef]
  356. Mignion, L.; Desmet, C.M.; Harkemanne, E.; Tromme, I.; Joudiou, N.; Wehbi, M.; Baurain, J.-F.; Gallez, B. Noninvasive detection of the endogenous free radical melanin in human skin melanomas using electron paramagnetic resonance (EPR). Free Radic. Biol. Med. 2022, 190, 226–233. [Google Scholar] [CrossRef] [PubMed]
  357. Ren, X.; Lin, K.; Hsieh, C.-M.; Liu, L.; Ge, X.; Liu, Q. Optical coherence tomography-guided confocal Raman microspectroscopy for rapid measurements in tissues. Biomed. Opt. Express 2021, 13, 344–357. [Google Scholar] [CrossRef] [PubMed]
  358. Mazurenka, M.; Behrendt, L.; Meinhardt-Wollweber, M.; Morgner, U.; Roth, B. Development of a combined OCT-Raman probe for the prospective in vivo clinical melanoma skin cancer screening. Rev. Sci. Instrum. 2017, 88, 105103. [Google Scholar] [CrossRef]
  359. Varkentin, A.; Mazurenka, M.; Blumenröther, E.; Behrendt, L.; Emmert, S.; Morgner, U.; Meinhardt-Wollweber, M.; Rahlves, M.; Roth, B. Trimodal system for in vivo skin cancer screening with combined optical coherence tomography-Raman and colocalized optoacoustic measurements. J. Biophotonics 2018, 11, e201700288. [Google Scholar] [CrossRef]
  360. Kukk, A.F.; Wu, D.; Gaffal, E.; Panzer, R.; Emmert, S.; Roth, B. Multimodal imaging system with ultrasound, photoacoustics, and optical coherence tomography for optical biopsy of melanoma. In Proceedings of the Multimodal Biomedical Imaging XVIII, San Francisco, CA, USA, 28 January 2023; pp. 16–22. [Google Scholar]
  361. Roth, B.; Kukk, A.F.; Wu, D.; Panzer, R.; Emmert, S. Four-modal device comprising optical coherence tomography, photoacoustic tomography, ultrasound, and Raman spectroscopy developed for in vivo skin lesion assessment. Biomed. Opt. Express 2025. [CrossRef]
  362. Wang, C.; Guo, L.; Wang, G.; Ye, T.; Wang, B.; Xiao, J.; Liu, X.J.A.O. In-vivo imaging of melanoma with simultaneous dual-wavelength acoustic-resolution-based photoacoustic/ultrasound microscopy. Appl. Opt. 2021, 60, 3772–3778. [Google Scholar] [CrossRef]
  363. Wang, Y.; Xu, D.; Yang, S.; Xing, D. Toward in vivo biopsy of melanoma based on photoacoustic and ultrasound dual imaging with an integrated detector. Biomed. Opt. Express 2016, 7, 279–286. [Google Scholar] [CrossRef]
  364. Kukk, A.F.; Scheling, F.; Panzer, R.; Emmert, S.; Roth, B. Non-invasive 3D imaging of human melanocytic lesions by combined ultrasound and photoacoustic tomography: A pilot study. Sci. Rep. 2024, 14, 2768. [Google Scholar] [CrossRef] [PubMed]
  365. Kukk, A.F.; Scheling, F.; Panzer, R.; Emmert, S.; Roth, B. Combined ultrasound and photoacoustic C-mode imaging system for skin lesion assessment. Sci. Rep. 2023, 13, 17947. [Google Scholar] [CrossRef] [PubMed]
  366. Wongvibulsin, S.; Yan, M.J.; Pahalyants, V.; Murphy, W.; Daneshjou, R.; Rotemberg, V. Current state of dermatology mobile applications with artificial intelligence features: A scoping review. JAMA Dermatol. 2024, 160, 646–650. [Google Scholar] [CrossRef] [PubMed]
  367. Joshi, G.; Jain, A.; Araveeti, S.R.; Adhikari, S.; Garg, H.; Bhandari, M. FDA-Approved Artificial Intelligence and Machine Learning (AI/ML)-Enabled Medical Devices: An Updated Landscape. Electronics 2024, 13, 498. [Google Scholar] [CrossRef]
  368. U.S. Food & Drug Administration. Marketing Submission Recommendations for a Predetermined Change Control Plan for Artificial Intelligence-Enabled Device Software Functions. 2024. Available online: https://www.fda.gov/regulatory-information/search-fda-guidance-documents/marketing-submission-recommendations-predetermined-change-control-plan-artificial-intelligence (accessed on 12 January 2025).
Figure 1. Overview of technologies included in this review.
Figure 1. Overview of technologies included in this review.
Biosensors 15 00297 g001
Figure 2. Flowchart describing process for inclusion of articles in this review.
Figure 2. Flowchart describing process for inclusion of articles in this review.
Biosensors 15 00297 g002
Table 1. Imaging modalities and how they are or could be used in melanoma healthcare.
Table 1. Imaging modalities and how they are or could be used in melanoma healthcare.
Imaging Modalities Currently Used in Medical Practice
TechnologyBiomarkerMain ApplicationsMain Limitation(s)
Smartphone/digital photographyABCDE criteria.PrescreeningVariable lighting, automated edge enhancement, user focusing decreases accuracy.
Total body photographyABCDE criteria.PrescreeningDetects changes over time, hence misses skin regions in genital, acral, scalp, within body folds. Large image files.
DermoscopyIrregular pigment network, asymmetrical structures, abrupt peripheral streaks, uneven color distribution within a lesion, and vascular features.Prescreening/screeningSuperficial assessment. Poor sensitivity for small melanomas, haphazard monitoring over time.
Electrical Impedance SpectroscopyMelanomas show lower impedance at some frequencies due to higher water content and disrupted extracellular matrix and disrupted cell membranes. Melanomas have more pronounced change in impedance with change in frequency.Prescreening/screeningLow specificity for melanoma, can be fooled by inflammation, ulceration, scar tissue. Requires trained physician to assess results.
Reflectance Confocal Microscopy (RCM)Contrast is based on reflectance of different skin components. Melanin has a high refractive index and appears bright. RCM also detects irregularly shaped cells and disorganized arrangements of cells and structures.ScreeningSmall field of view leads to lengthy imaging sessions for large lesions. Shallow imaging depth.
Optical coherence tomography (OCT)Atypical cell structures and disorganized tissue architecture including honeycombed patterns, pagetoid spread, absence of dermal nests, and atypical melanocytes in the dermis.
Also, optical features extracted and
analyzed via machine learning.
D-OCT: changes in microvascular structures in the skin including increased vascular density (angiogenesis), chaotic vessel architecture, irregular and dilated blood vessels, and irregular blood flow.
ScreeningHD-OCT, OCM and LC-OCT have cellular resolution however do not have sufficient imaging depth for staging. Specificity and sensitivity have not been fully studied.
D-OCT: focuses on blood flow changes caused by melanoma.
Multispectral ImagingDifferential absorption and reflectance of light at multiple wavelengths, capturing variations in melanin, hemoglobin, and oxygenation levels that distinguish malignant melanomas from benign lesions.Prescreening/screeningLow specificity, cannot be used for rare melanomas.
Ultrasound (3.5–14 MHz)Berlin morphology criteria for finding sentinel node metastases: peripheral perfusion, loss of central echoes and balloon shapes.Staging, surgical/treatment guidanceMay not provide reliable preoperative nodal staging.
High-Frequency Ultrasound (>15 MHz)Hypoechoic lesions infiltrating the dermis from the epidermis. Anechoic content in suspicious lesions. Margins are usually sharp. Doppler shows increased and anarchic vascularization.StagingAllows deep penetration (1.5–8 mm) beneficial for estimating tumor thickness. Does not rely on melanin, thus useful for amelanotic melanoma. Poor sensitivity.
Raman SpectroscopyDistinct spectral signature resulting from altered molecular compositions, such as elevated nucleic acids, proteins, lipids and melanin, reflecting the biochemical changes characteristic of malignant melanocytes compared to normal skin tissue.ScreeningModerate specificity for in vivo screening.
Elastic scattering spectroscopyVariation in light scattering patterns caused by differences in cellular and subcellular structures, such as nuclear size and density, which distinguish malignant melanoma from benign skin lesions.ScreeningLimited depth detection, high sensitivity but modest specificity.
Magnetic Resonance ImagingLesions are hyperintense by T1 due to reduced T1 relaxation time associated with melanin. T2 signal is reduced. T1 with contrast shows a heterogeneous peripheral rim enhancement.Staging, surgical/treatment guidanceLow sensitivity, has long scanning times, and requires exogenous contrast agents.
Positron emission tomography/computed tomography (PET/CT)18F-FDG (glucose analog that accumulates in cells with high metabolic activity).Staging, surgical treatment/guidance, post-surgical/treatment monitoringIonizing radiation, requiring tracers, expensive.
Single Photon Emission Computed Tomography (SPECT/CT)99mTc-labeled colloids/99mTc-Tilmanocept.Staging, surgical/treatment guidanceIonizing, requiring tracers, limited availability and cost, low spatial resolution.
Lymphoscintigraphy99mTc-labeled colloids.Surgical guidanceThis nuclear medicine imaging technique, low spatial resolution, limited sensitivity for early-stage metastases, and lack of real-time imaging.
Imaging Modalities Currently in Development
TechnologyBiomarkersProposed Main UtilityStrengths and shortcomings
Photoacoustic Imaging (PAI)Melanin is a strong endogenous chromophore, multispectral imaging enables visualization of melanoma-related angiogenesis, and other changes.
Elevated optical absorption contrast of melanin at specific wavelengths, combined with increased vascularization and oxygenation heterogeneity that reflect the tumor’s metabolic and structural abnormalities.
Screening, staging, surgical/treatment guidance, post-surgical/treatment monitoringGood depth penetration. Functional imaging (melanoma and hemoglobin). Versatile technique that can be implemented with either good depth penetration and/or high spatial resolution.
Hyperspectral ImagingUnique spectral reflectance and absorption patterns across visible and near-infrared wavelengths, driven by variations in melanin concentration, hemoglobin content, and tissue morphology.ScreeningPortable imaging device. Can distinguish hemoglobin from melanin. Moderate specificity.
Quantitative Dynamic Infrared ImagingAbnormal thermal signature and delayed heat dissipation patterns caused by increased metabolic activity and vascularization.ScreeningCan image large areas of skin rapidly. Low specificity, difficulty detecting small melanomas.
Terahertz Pulse ImagingAltered terahertz refractive index and absorption coefficient, reflecting differences in water content, cellular density, and tissue composition between malignant melanomas and benign skin lesions.Screening/possibly stagingCan define rough tumor margins but very limited depth detection. Poor sensitivity.
Multiphoton Imaging—Two Photon Excitation (2PE) Microscopy and Second Harmonic Generation (SHG) microscopyAltered autofluorescence and reduced second-harmonic generation signals, changes in melanin distribution, collagen organization, and metabolic activity in malignant melanomas compared to normal or benign skin tissues. Reduced SHG signal intensity reflects structural abnormalities in the extracellular matrix associated with malignant melanoma progression.ScreeningLimited penetration (a few hundred µm) in depth penetration at subcellular (0.5 µm lateral and 1–2 µm axial) resolution.
Fiber DiffractionAltered diffraction patterns of collagen and other fibrous proteins, indicating changes in molecular packing and structural organization associated with the tumor microenvironment in malignant melanoma.ScreeningChanges in fiber organization may occur in early-stage tumors. Poor sensitivity to cellular-level changes. Requires ordered molecular structures:
fiber diffraction is best suited for analyzing highly ordered, periodic structures like collagen fibrils or keratin networks.
Fourier Transform Infrared SpectroscopyDistinct absorption peaks corresponding to altered lipid, protein, and nucleic acid compositions, reflecting biochemical changes in malignant melanoma cells compared to normal or benign tissues.Pathological stagingRapid detection of melanoma in tissue samples through spectral changes. Not for in vivo use.
Real-time ElastographyIncreased stiffness and altered elastic properties of malignant tissues, reflecting changes in extracellular matrix composition and tumor-induced mechanical heterogeneity.Possibly screening, stagingReal-time imaging, excellent imaging depth up to 10 mm. Technique is time-consuming and labor-intensive. Poor specificity.
Electron paramagnetic resonance spectroscopyThe elevated levels of melanin-associated free radicals and altered paramagnetic properties, reflecting oxidative stress and metabolic changes characteristic of malignant melanoma.Screening, stagingHigh quality 3-dimensional images, excellent penetration depth of 7 mm or more. Struggles with resolution and small melanomas.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Horton, L.; Fakhoury, J.W.; Manwar, R.; Rajabi-Estarabadi, A.; Turk, D.; O’Leary, S.; Fotouhi, A.; Daveluy, S.; Jain, M.; Nouri, K.; et al. Review of Non-Invasive Imaging Technologies for Cutaneous Melanoma. Biosensors 2025, 15, 297. https://doi.org/10.3390/bios15050297

AMA Style

Horton L, Fakhoury JW, Manwar R, Rajabi-Estarabadi A, Turk D, O’Leary S, Fotouhi A, Daveluy S, Jain M, Nouri K, et al. Review of Non-Invasive Imaging Technologies for Cutaneous Melanoma. Biosensors. 2025; 15(5):297. https://doi.org/10.3390/bios15050297

Chicago/Turabian Style

Horton, Luke, Joseph W. Fakhoury, Rayyan Manwar, Ali Rajabi-Estarabadi, Dilara Turk, Sean O’Leary, Audrey Fotouhi, Steven Daveluy, Manu Jain, Keyvan Nouri, and et al. 2025. "Review of Non-Invasive Imaging Technologies for Cutaneous Melanoma" Biosensors 15, no. 5: 297. https://doi.org/10.3390/bios15050297

APA Style

Horton, L., Fakhoury, J. W., Manwar, R., Rajabi-Estarabadi, A., Turk, D., O’Leary, S., Fotouhi, A., Daveluy, S., Jain, M., Nouri, K., Mehregan, D., & Avanaki, K. (2025). Review of Non-Invasive Imaging Technologies for Cutaneous Melanoma. Biosensors, 15(5), 297. https://doi.org/10.3390/bios15050297

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop