Next Article in Journal
α-Borophene Nanoribbons: Edge-Dependent Metallic and Magnetic Properties for Low-Dimensional Nanoelectronics
Previous Article in Journal
Hyperpolarized Carbon-13 Metabolic Imaging Differentiates Distinctive Molecular Phenotypes in Diffuse Midline Gliomas
Previous Article in Special Issue
Modern Methods for Detection of Fentanyl and Its Analogues: A Comprehensive Review of Technologies and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Fundamentals of Laser-Induced Breakdown Spectroscopy to Recent Advancements in Cancer Detection and Calcified Tissues Analysis: An Overview (2015–2025)

by
Muhammad Mustafa Dastageer
1,
Khurram Siraj
1,*,
Johannes David Pedarnig
2,*,
Dacheng Zhang
3,
Muhammad Qasim
1,
Muhammad Shahzad Abdul Rahim
1,
Saba Mushtaq
1,
Qaneeta Younas
1 and
Bareera Hussain
1
1
Laser & Optronics Centre, Department of Physics, University of Engineering and Technology (UET), Main Campus, G.T. Road, Lahore 54890, Pakistan
2
Institute of Applied Physics, Johannes Kepler University Linz, A-4040 Linz, Austria
3
School of Optoelectronic Engineering, Xidian University, 2 South Taibai Road, Xi’an 710071, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(21), 4176; https://doi.org/10.3390/molecules30214176 (registering DOI)
Submission received: 21 July 2025 / Revised: 10 October 2025 / Accepted: 14 October 2025 / Published: 24 October 2025
(This article belongs to the Special Issue Review Papers in Analytical Chemistry, 2nd Edition)

Abstract

Laser-induced breakdown spectroscopy (LIBS) is a promising elemental analysis technique that has rapidly evolved in numerous fields, including biomedical research and medical sciences, over the last two decades. In combination with other methods, it has the potential to examine complex biological structures and their species distributions. The present work first develops the basic understanding of LIBS and then reviews its evolution in oncological diagnosis and calcified tissue analysis from medical perspectives over the last 11 years. LIBS can potentially improve early cancer detection and monitor treatment outcomes, ultimately enhancing patient care and diagnosis. It has effectively differentiated between malignant and normal tissues and also classifies cancer stages and types based on disease severity. Its applications for categorising and identifying calcified tissues are attractive for inspecting minerals, while soft tissue is more challenging, given the potential for significant matrix effects. This review article deals with the following aspects of LIBS and its application: (i) the fundamentals of this analytical measurement method, (ii) the matrix effect and its influence on the LIBS analyses of various biological tissues, (iii) the role of signal enhancement methodologies and artificial intelligence models to advance the method for analyses of biological sample materials, and (iv) applications of LIBS in cancer and calcified tissues investigations. This article also addresses challenges and opportunities encountered in these applications and discusses prospects, providing a comprehensive overview of the current state and potential advancement in LIBS technology.

Graphical Abstract

1. Introduction

LIBS has garnered considerable attention from researchers in biomedical fields over recent years, yielding prolific outcomes either alone or in combination with other supporting techniques [1,2,3]. This laser-induced plasma spectroscopy method stands out for its speed, non-destructiveness, real-time analysis, minimal or no sample preparation requirement, multi-element detection, and relatively inexpensive instrumentation [4]. It has been effectively used for environmental monitoring [5,6], agricultural materials inspection [7], archaeological investigation [2,8], geological mining [9,10], industrial applications [4,11], space exploration [12], forensic sciences [13], health sciences [14,15], and biological detections [16,17].
Cancer remains the leading cause of death globally, mainly due to its silent onset and lack of early symptoms [18,19,20]. Despite advances in bioscience, early detection—a critical factor for successful treatment—remains a significant challenge. Cancer diagnosis today relies on conventional biopsies, imaging [computed tomography (CT)/magnetic resonance imaging (MRI)/X-ray/positron emission tomography (PET)/mammography (MMG)], and laboratory tests [21,22,23,24], which are effective but often slow, expensive, and invasive [21,25]. Alternatively, optical analytical techniques [atomic absorption spectroscopy (AAS) [26], laser-induced breakdown spectroscopy (LIBS) [27], Raman spectroscopy (RS) [28], inductively coupled plasma mass spectrometry (ICP-MS) [29], and X-ray fluorescence (XRF) [30] offer rapid, cost-effective, and less invasive diagnostic efficacy. Among others, LIBS is considered an emerging state-of-the-art tool with significant advancements in understanding cancer biology [31,32].
Calcified tissues serve as elemental archives due to hydroxyapatite’s (HA) affinity for toxic metals and metabolic markers [33]. Teeth disorders and metabolic bone diseases alter HA crystallography, changing crystal morphology, orientation, and alignment with collagen, as well as elemental composition, in a manner distinct from that found in normal calcified tissue [34]. Their composition is analysed by two methods: dissolution, which dissolves samples and hinders in vivo research, and spectroscopy, which is semi-destructive and provides precise spatial mapping, enabling more accurate in vivo and in vitro analysis than the former [16].
LIBS enable rapid elemental ana lysis of cancerous and calcified tissues, serving as a powerful tool for pathological diagnosis and physiological monitoring [35,36,37]. Classification of pathologies corresponds to variations in laser-induced plasma characteristics, elemental concentrations, detection of toxic metals, molecular decomposition, de/remineralisation, chemical imbalances, calcification, and carbonisation [34,38]. LIBS achieved significant milestones in cancer diagnostics and the investigation of calcified tissue, as highlighted in Figure 1. However, its clinical adoption remains limited due to challenges, for instance the matrix effect, issues in signal reproducibility, sensitivity issues from tissue heterogeneity, and ablation-induced sample destruction [39,40,41]. The path forward lies in intelligent calibration systems and hybrid analytical approaches that compensate for these technical challenges by providing linearity, sensitivity, and robustness to signals [42]. In 2019, Gaudiuso et al. [43] devoted review sections to LIBS-based analysis of calcified tissues and the application of LIBS for cancer diagnosis using fluid specimens of humans and animals. In 2022, Khan M.N. et al. [40] reviewed advancements in LIBS for diagnosing various cancers using different tissue samples and body fluids and Khan Z. et al. [44] reviewed trace element detection in biomaterials and other materials. The importance of LIBS in calcified tissue analysis was highlighted by Singh et al. [16] a decade ago.
The lack of availability of LIBS-oriented research and its unexplored avenues in cancer diagnosis, calcified tissue analysis, and the effect of cancer on calcified tissues urge us to write this article to attract, convince, and engage researchers and readers toward med-LIBS. The primary goal of writing the present article is to keep readers updated on the latest ongoing research trends (in med-LIBS) by covering literature (from 2015 to 2025) related to LIBS application in cancer diagnosis and calcified tissue examination. This work provides a brief overview of LIBS, including the setbacks associated with the matrix effect, the methods used for signal enhancement (SE), and the artificial intelligence (AI) models applied for spectral analysis. It includes a critical evaluation of LIBS as a diagnostic tool in multiple types of cancers: skin, breast, blood, lungs, stomach, colon, ovarian, gallbladder, oral, cervical, and brain. Furthermore, an overview is presented on published research work on human and animal calcified tissues (teeth, bones, kidney stones, and chicken shells) from LIBS-medical aspects. Finally, before concluding, the article discusses shortcomings, potential solutions, and prospects.

2. Fundamentals of LIBS and Elemental Composition of Human Body

LIBS is a spectro-analytical technique in which a laser pulse ablates a small volume of material from the sample surface, creating a microplasma. As the plasma cools, it emits characteristic wavelengths of light that act like a fingerprint for the elements it contains (Figure 2a).
The spectrometer records these emissions to get the spectrum. These spectra were further analysed to determine the elemental composition of the target [47], and the images formed are shown in Figure 2b. LIBS fundamental principle can be summarised in four steps: (i) laser ablation process, (ii) plasma formation and plasma plume expansion, (iii) plasma emission, and (iv) spectral analysis [54,55,56].
A basis to advance understanding of LIBS can be developed going through these aspects: fundamentals (laser parameters, specimen properties, and plasma chemistry govern the complex interaction processes) [54], instrumentations and methodologies [34,56,57,58,59,60], pulsed laser–tissue ablation [61], plasma formation and plasma diagnostics [56,60,62,63,64,65], spectral analysis [54,55,56], advance data processing (AI-based machine learning ML models) [66,67], and elemental imaging [68,69]. Laser–matter interaction, various forms of ablated material and plasma interaction, and laser–plasma plume interaction (caused by plasma shielding) are underlying mechanisms for understanding LIBS analytical outcomes [62]. Due to the high complexity of interaction phenomena, specific theoretical models exist, and the most commonly used for plasma modelling is local thermodynamic equilibrium (LTE) [65]. There are two primary LIBS formalisms adopted for the elemental quantification of specimens: calibration curve (CC-LIBS) and calibration-free (CF-LIBS). In CC-LIBS, standard reference materials are required to quantify the elemental content of the specimen by plotting a calibration curve. In contrast, the CF-LIBS approach is employed to determine the elemental concentration of diverse samples without standard references or matrix-matched materials, eliminating the need for a calibration curve [70,71,72,73].
In nanosecond (ns) pulsed laser ablation, plasma is formed during laser pulse, and the trailing part of the pulse interacts with plasma, leading to plasma reheating and persistence. The plasma formation time depends on the laser pulse length because the pulse duration determines how long energy is delivered to the target materials, thereby impacting the mechanisms and time scales of plasma formation. For ns laser pulses, it is an order of a few ns to tens of ns, whereas for fs laser pulses, it is an order of 1 ps [74]. Ultra-short laser pulses with durations ranging from picoseconds (1 ps = 10−12 s) to femtoseconds (1 fs = 10−15 s) are commonly used [75]. The fs pulse plasma evolves faster, with a lifetime of several hundred ns, depending on laser fluence and various other factors. In contrast, ns plasma evolves relatively slowly and has a longer lifetime of µs [74,75]. In femtosecond pulsed laser ablation, plasma is formed after the laser pulse, so interaction between plasma and laser can be avoided. Due to the absence of plasma–laser interaction and the very short thermal penetration length for fs pulses the absorption of laser energy is localised to the irradiated area and the heat-affected zone (HAZ) is strongly reduced compared to ns pulses [76,77].
Fs laser pulses offer high ablation efficiency as the ablation thresholds are lower than with ns pulses. Fs laser ablation depth of 6 µm on a thin tissue section of liver metastases (from a colorectal cancer (CRC) patient) was reported, allowing fast in-depth multi-elemental profiling at cellular spatial resolution [78]. A maximum ablation rate of 0.66 mm3/s in porcine femur for laser pulses with wavelength of 515 nm and repetition rate of 250 kHz is reported in [79]. These pulses promoted less selective ablation and reduced dependence on the material matrix, which favours determination of elements in dental tissue (dentine) [80]. Fs LIBS offers reproducible spectra in the pathological tissues (section of liver metastases of CRC patient, breast tissues with tumour, and lymph node with metastasis breast cancer) at lower energies due to less interaction between fs pulses and plasma plumes [81]. Fs-LIBS elemental imaging of melanoma tumour tissue (skin cancer) provides a spatial resolution of 15 µm [82]. It is also used as a real-time feedback control system to ensure the removal of bone tumours, thereby reducing the need of repeat surgeries [83]. Their ability to reduce damage to dental tissues and surrounding nerves from undesirable thermal destruction enables the use of a feedback loop in which the clinician can both diagnose and remove carious lesions [84].
Elemental concentration in the human body can vary for several reasons, primarily related to physiological and quality of living standards [85], including age [86], gender [86,87], diet [88], habitat [89], and environmental exposure [90]. Another source of elemental variations can be disease [91]; among many, our prioritised content focused on various forms of cancer diseases and pathological calcified tissue abnormalities, and researchers identified elemental biomarkers based on LIBS data collection. It is worth noting that the interpretation of LIBS data has a significant impact on determining relevance from a biological perspective [92]. Even for the same physiological or pathological conditions, elemental variation may not be unique. LIBS investigators have not succeeded in defining particular criteria for certain types of disease diagnosis, and novel research still requires confirmation. Additionally, spectro-analytical outcomes highly depend on experimental instruments and conditions, the nature of tissues, plasma dynamics, approximations for spectral analysis, formalisms for quantification, statistical methods, and the use of AI models for characterising biological tissues [85,93].
Major/macro, micro/minor/lesser/minute, and trace elements constitute 98.5%, 0.7%, and 0.8% of total human body weight, respectively, as illustrated below in Figure 3 [94]. Macroelements are crucial for regulating glucose metabolism and detoxifying contaminants, whereas micronutrients are essential for the body in minute quantities for the proper growth and development of organs. Trace elements, though scant, are pivotal: their imbalance triggers the disease, from hyperglycemia to toxic metal poisoning, serving as silent biomarkers for cancer and calcified tissue disorders [68,91,92]. Yet quantifying them remains a challenge that one LIBS could overcome [92] if its limitations, particularly matrix effects, are addressed.

2.1. Matrix Effect

The matrix effect is the influence of a specimen’s chemical composition and physical properties on analytical signals, causing variations in emission line intensities and errors in analyte quantification [95]. The measured intensity of an analyte chemical element may show a dependence on the concentration of another element in the sample material, for instance (cross-sensitivity). Also, the spectrochemical analyte signal may depend on the sample material’s condition, e.g., its hardness, mass density, microstructure, and water content, even if the concentration of the analyte is independent of such conditions. Laser ablation of tissue is initiated by the absorption of laser radiation in the surface of the sample material and, in case of short ns pulses, in the early plasma induced by the trailing edge of the pulse. The penetration depth of absorbed laser pulse energy into the sample and, therefore, the amount of ablated material per pulse, the energy of ablated species in the plasma, and the composition of the plasma plume are dependent on the condition/state of the tissue investigated. The measured intensity of continuous radiation which masks relevant elemental information [62,96] and, most importantly, the intensity of specific analyte emission lines and bands at later stages of plasma expansion are therefore dependent on such conditions. The matrix effect has direct involvement in LIBS signals and cannot be completely eliminated; however, its adverse effects can be reduced to improve LIBS quantifications [97]. This unwanted phenomenon can cause (i) reduction in signal sensitivity and detection limit due to variations in plasma properties [98], (ii) non-linearity in calibration curves due to non-uniform response of the LIBS measurement system to the concentration of analytes across different matrices [99], (iii) non-uniform ablation due to different physical and chemical properties of matrices [100], and (iv) elemental fractionation due to non-stoichiometric ablation [101].
In LIBS (and also for other laser ablation based spectrochemical techniques) the matrix effect in soft and hard tissue is very significant. For cancer diagnostics soft tissues are typically characterised. The effect primarily originates from organic and inorganic content of tissues which have distinct elemental compositions that, in turn, influence their plasma parameters [49,102]. Research [103] has demonstrated that plasma temperature and electron number density are inversely proportional to carbon content in samples, which is associated with ablation rates and the recombination process within the plasma. Laser-induced plasma of organic tissues is typically dense, with several optically thick spectral lines, resulting in a pronounced background that can obscure the subtle analytical signals. These organic tissues contain water which acts as a thermal buffer, reducing ablation efficiency and plasma temperature. Consequently, this leads to signal suppression and poor reproducibility. Protein and lipid increase molecular emission (CN and C2) due to their high carbon content which can interfere with metal emissions. Accurate quantification of metallic trace elements is essential for assessing the physio-pathological states of tissues. Relatively, soft tissues exhibit weaker mineral emission making the detection of trace elements more challenging [104]. Due to the matrix effect, poor sensitivity and a high limit of detection (LOD) are hurdles in Tag-LIBS readout for the accurate quantification of human serum albumin (HSA) [105], multiple tissues, tooth-bone, are encountered in the same specimen. Their examination is challenging [106], and matrix/non-matrix elements of dental tissues (healthy or carious) and filling material (amalgam) behave differently at varying laser fluences and pulse numbers [107]. Approaches to reduce the undesirable matrix effect in LIBS include the following: (i) adoption of suitable sample preparation methods (e.g., with pressed samples plasma temperature fluctuation can be reduced [103]); (ii) optimisation of experimental parameters (e.g., laser beam defocusing and spectrometer delay [99], use of ultraviolet laser wavelength (266 nm) for higher ablation efficiency, improved plasma stability, and sharper spectral lines in calcified matrices [108]); and (iii) selection of data processing methods (e.g., matrix-matched external calibration [109,110], automated matrix recognition for identification of different tissues in a sample using ML algorithm improved elemental quantification [109], image fusion techniques extracting matrix independent information of bones from LIBS data profiling [111]). Additionally, various matrix-matched reference materials are available to enhance the accuracy and precision of quantitative analyses. Several calibration materials and methods are used for the trace elemental quantification of soft tissues (such as brain, liver, and hair) and hard tissues (teeth and bones) via LIBS and related techniques such as LA-ICP-MS [110]. The optical diagnostic approach of front-face fluorescence (FFF), used as a complementary method, can reduce the reliance of LIBS elemental results on the tooth matrix for early detection of pathologies [112].
Despite employing multiple strategies, researchers have not reached a consensus on which method is most suitable to reduce the matrix effect for different types of tissue. An overview on the reported influence of the biological matrix on the LIBS analytical performance of tissues is given in Table 1. The heterogeneous nature of samples contributes to this effect. LIBS analyses of tissues with similar elemental compositions (e.g., fat and nerves) but different structure (matrix) require different approaches for analyte quantification. Results from ex vivo analysis cannot be reliable transferred to in vivo applications, where the tissues are immersed in blood and other body fluids and surrounded by other organs [113]. A universal procedure to correct LIBS spectra and data for this effect is lacking due to complexity and diversity of biological tissues. Few efforts have been made to mitigate the matrix effect on LIBS performance using animal tissues [49,96,109,113,114,115]. Studies on human tissues [116] are lacking, which may be due to ethical issues and the unavailability of human tissues.

2.2. LIBS Signal Enhancement Strategies

LIBS is encountering numerous difficulties (mentioned in Table 1) in examining biological tissues. Several methods have been tried to address these issues. Among many methodologies, nanoparticles enhanced (NE-LIBS) is recommended due to its potential for on-site analysis (in vivo studies) of biological tissue. But in this method, enhancement of emission signals depends on the size, shape, distribution, and concentration of nanoparticles (NPs), which are not easily controlled. Colloidal silver nanoparticles (Ag-NPs) can enhance the LIBS spectral intensity of metals in serum samples deposited on filtration paper by mitigating the coffee ring effect [126]. The reported enhancement factors for Ca, Mg, and K are 1.76, 1.85, and 3.10, respectively [127]. Sprinkling of bio-synthesised Ag-NPs on bovine bones before analysis enhanced the sensitivity of LIBS signal [128]. Ag-NPs on antibodies enhanced the detection of europium (Eu) and ytterbium (Yb) up to 12 times [129]. The formation of zinc oxide nanoparticles in acid (eugenol) for teeth analysis [130] poses health issues for in vivo examination. Its limitation is that it can only improve the sensitivity and LOD of elements, yielding linear calibration curves.
Tag-LIBS readout sensitivity for human serum albumin (HSA) detection is improved by using upconversion nanoparticles (UCNPs), collinear DP-LIBS configuration, and a modified optical collection system. The reported LOD of TAG-LIBS for HSA is 0.29 ng/mL which is comparable to 0.37 ng/mL when measured with a standard enzyme-linked immunosorbent assay (ELISA). However, the performance of Tag-LIBS is still significantly lower than that of the gold standard readout, the upconversion-linked immunosorbent assay (ULISA) (LOD: 0.17 ng/mL). Additionally, for Tag-LIBS the signal-to-background ratio (SBR) is 3.5 times lower than that of ELISA, and the results demand validation on multiple biomedical specimens [105]. NETag-LIBS, a combination of NE-LIBS and Tag-LIBS, has enhanced sensitivity in the detection of biomarkers when an optimum concentrations (0.05–0.1) mg/mL of Ag-NPs is used [129].
Spark discharge (SD)/spark-assisted (SA) LIBS methods have been employed to discriminate between healthy and cancerous gastric tissues [123] and healthy (infiltrate) and brain tumour tissues [131] based on emission signals of Ca and Mg. However, emission enhancement is not significant enough, and signal enhancement factors are not reported, which may be due to experimental uncertainties. In surface-enhanced (SE-LIBS) superhydrophobic substrates, polydimethylsiloxane (PDMS), can elevate the detection of trace elements in the blood serum of an oral cancer patient, improving sensitivity and specificity to 88% and 96%, respectively [132].
Fs LIBS enables one to map the distribution of tooth elements with spatial resolution of 100 µm [80]. Fs-DP-LIBS signals are enhanced up to 5-fold compared to fs-SP-LIBS when exploited to bovine tissues (liver and muscle). However, the enhancement factor is depending on several factors including physicochemical matrix effects, sample preparation, environmental conditions, and instrument specifications [133]. Fs-pulses enable the removal of caries from dental tissues through a precise ablation process, but optimisation of the fs-LIBS experimental setup is the prerequisite for controlling temperature [84].
Very few studies have exploited dual/double pulse (DP-LIBS) for biological specimens. One is for bovine tissue analysis and the reported signals are 5-fold those of SP-LIBS [133]. The other is about HSA where a collinear DP-LIBS configuration is employed [105]. Although the potential of DP-LIBS, electric field-assisted (EF-LIBS), and microwave-assisted (MW-LIBS) has been demonstrated in other scientific domains [134,135,136,137,138,139,140], their specific advantages and limitations in the analysis of medical samples remain unexplored and require systematic investigation.

2.3. Artificial Intelligence in LIBS

LIBS-integrated ML and DL algorithms offer advantages in large and complex spectral data processing, including pattern recognition for rapid classification, spectral feature extraction, and high precision and accuracy by mitigating sample matrix effects and compensating for self-absorption (SA) interference [141]. Generally, model training requires complex data pre-processing to filter signals using various approaches including baseline correction, normalisation, and outlier detection [142]. Nowadays, the spectroscopic output contains thousands of variables and millions of spectra which are impossible to process manually, but AI algorithms can efficiently and effectively accomplish such complex processing [143]. Although these models have several advantages, for data from biological samples they are encountering many challenges (overfitting, underfitting, optimisation and validation losses, inconsistency, absence of physical embodiment, lack of interpretability, and probabilistic outcomes) [144].
Principal component analysis (PCA) is an unsupervised ML algorithm. It is commonly applied for dimensionality reduction of LIBS spectra to enhance model robustness for the diagnosis of oncological diseases [31,48,119,122,132,145] and the analysis of hard tissues [83,96,112,146,147,148,149]. For diagnosis of human blood malignancies (lymphoma and myeloma) the measured spectral data of order 4800 (number of spectra) × 24 (number of spectral lines) was reduced to 4684 × 12 [48], for discrimination of lymphoma the original data set 3240 × 16 was minimised to 3240 × 6 [119], and for classification of tooth tissues the data matrix 5589 × 9 was compressed to 5589 × 2 [149]. After data reduction through PCA, the average sensitivity of the supervised ML algorithms linear discriminant analysis (LDA), quadratic discriminant analysis (QDA), and kernel nearest neighbour (KNN) for serum samples (of blood cancer patients and healthy subjects) was 82%, 92%, and 96%, respectively [48]. LDA accuracy for discrimination of lymphoma was 99.78%. Validation losses and misclassification cases, even for a small dataset of 17 patients, remain a significant concern for the algorithm’s performance [119].
Artificial neural network (ANN)-supervised ML models are frequently used in the classification of LIBS spectral data for cancer [27,81,118,122,150,151] and calcified [147,152] tissues with an accuracy of over 89%. The major concern is that none of these studies [27,81,118,122,147,150,151,152] reported all the required model parameters: architecture (number of hidden layers, number of neurons per layer, hidden activation function, and output activation function), learning (learning rate, batch size, and loss function), and regularisation (early stopping, dropout rate, and L2 regularisation rate). It is necessary to incorporate such information in published research work; without it, these studies cannot be reproduced. Consequently, the results remain specific to the application of a particular lab. In addition, many other ML models (PLS-DA [118,152], logistics regression (LR) [147,153], support vector machine (SVM) [31,154,155,156,157,158], random forest (RF) [81], boosting tree [159], bagged tree [160], bagging voting fusion (BVF) [158], and frameworks (XGBoost [121] and AdaBoost [150])) are used in both of the mentioned medical fields. However, the limitations of ML models and frameworks prevent LIBS technology from being applied outside of laboratories. To overcome such issues, data scientists built a deep learning architecture and introduced reinforcement learning.
Deep neural network (DNN) and convolutional neural network (CNN) are deep learning algorithms designed for tabular/structural and imaging/visual data, respectively [144]. Despite the restriction on using CNN on imaging data, a few researchers apply them to sequential data [121,161,162,163]. They claimed classification accuracy of 97.72% for various cancer cells (cervical, liver, and colorectal) [161], 82% for breast cancer identification from healthy tissues [162], and 92% classification sensitivity (first batch) for breast cancer [163], and 93% for lung cancer staging [121]. Biases and small perturbations intentionally introduced in the input data raised questions about the credibility of this research. In addition, the unavailability of architectural and model parameters in several works, (i) DNN for skin cancer diagnosis [52], (ii) RF-1D ResNet for lung tumour identification [164], and (iii) GRAN for breast cancer [163], does not support their adoption in the medical field, because these frameworks inherit certain limitations (substantial amount of training data, unclear working mechanisms, requirement of high computation power, environmental impact, security vulnerability, and poor generalisation) [144]. Table 2 highlights some shortcomings of integrating AI models for the interpretation of LIBS data from oncology and calcified tissue research, along with possible solutions and suggestions.

3. Implementation of LIBS in Oncology

The first research on cancer via LIBS was performed in 2004, focusing on discriminating cancerous tissues from healthy ones (in dogs) [167]. This foundational research work laid the groundwork for subsequent studies examining LIBS in multiple cancer diagnostics. LIBS cancer screening and pathology identification efficacy provides simpler and more realistic objectives as a complementary approach to traditional histopathological examination [117].
Several methodologies are employed to investigate LIBS as a diagnostic tool in pathological examinations, including (I) classical LIBS approach [168], which involves comparing spectral features, plasma parameters, and elemental composition of cancerous and healthy tissues, and validated against traditional analytical methods. This typical method has been deployed to investigate multiple types of cancers: skin cancer [117], breast cancer [35,47], stomach cancer [123], colon cancer [47,50,120], brain cancer [131], bone invasive oral cancer [153], and gallbladder cancer [39]. (II) LIBS elemental imaging, where spatial resolution is a critical parameter, as it governs the ability to distinguish fine structural and compositional features within a sample. It is crucial for cancer diagnosis, where detailed elemental mapping can reveal subtle tissue changes. The technique involves scanning the sample with micrometre-scale laser pulses in a raster pattern. Each pulse ablates a small area, producing a LIBS spectrum that reflects the elemental composition at that point.
Advanced spectral analysis quantifies elements of interest, typically biomarkers, and generates high-resolution 2D or 3D maps that visually represent their spatial distribution [27,47,169,170]. This methodology is used in the identification of skin cancer [125,171] and malignant pleural mesothelioma (MPM) cancer [29]. (III) AI-assisted LIBS (supervised machine learning methods [168]) focuses on spectral features related to cancer markers and matches spectral peaks to identify elements associated with pathologies. Spectral features (peak intensity, area under the curve, and ratio between elements) are extracted, such as to use them as inputs for ML model. The steps are as follows: (i) training of model (using 70% known data set (training set)) to learn the relationship between spectral feature and cancer classification, (ii) optimisation of model (using 15% known data set (validation set)) by controlling hyperparameters (to improve model accuracy and prevent overfitting) to attain optimum classification performance, and (iii) testing and evaluating the model performance (using 15% unseen data set (test set)) using accuracy, sensitivity, specificity, and receiver operating characteristic-area under curve (ROC-AUC) to measure effectiveness of model in diagnosing cancer. This methodology has been used in several cancer studies, including skin cancer [52,119,145,150,168], breast cancer [81,162], lung cancer [159,160,164], ovarian cancer [27,172], colon cancer [81,161], cervical cancer [31,161], blood cancer [166], thyroid cancer [154], and brain tumours [151,165]. (IV) LIBS elemental imaging and machine learning, in which methods II and III are combined for screening, diagnosing, and monitoring of cancer, has been employed for skin cancer [51,82]. In addition, other strategies are also incorporated with LIBS to enhance cancer diagnostic and treatment efficacy, signal enhancement methodologies (surface-enhanced (SEN-LIBS) [132], and spark-assisted LIBS [123,131]), Tag-LIBS [45,173], NE-LIBS [129], ultrashort LIBS [81], and hybrid technology [35,121,171,174]. Before delving into each cancer type, it is essential to have basic information on LIBS instrumental and experimental parameters used in various oncological investigations, as presented in Table 3.

3.1. Skin Cancer

Skin cancer leads global cancer statistics, with over 1.5 million cases reported in 2022. Melanoma, the most lethal form, shows higher prevalence in white populations and males. Non-melanoma types, such as basal cell carcinoma (BCC), squamous cell carcinoma (SCC), and Merkel cell carcinoma (MCC), further add to substantial diagnostic load [176,177]. Multiple non-invasive optical modalities, including dermoscopy, multiphoton microscopy, confocal microscopy, dermatofluoroscopy (DMF), and optical coherence tomography (OCT), offer real-time imaging for early detection [19,21,23,24,178,179,180]. These techniques face critical limitations, including poor penetration depth, low resolution, and an inability to distinguish between ambiguous lesions. Consequently, they are insufficient as standalone diagnostic tools and continue to rely on histopathological confirmation via biopsy. Optical imaging remains a valuable adjunct, but not a replacement, for more definitive microscopic and spectroscopic analyses [21]. In this context, LIBS combined with optical imaging methods can be an excellent choice for addressing skin diseases clinically, where pathologists can effectively use it to screen, diagnose, and perform therapeutic operations before making additional treatment recommendations [25].
Khan and his co-workers (2020) [118] proposed a stage-wise classification of melanoma skin cancer using ML algorithms. The identification of melanoma in skin tissues through multiple classification models and the declared achievement of classification accuracy (100%). Pyun S.H. et al. (2023) [52] used LIBS technology in real-time skin cancer detection. The DNN model is trained and tested based on the LIBS-identified biomarkers during in vivo analysis. They achieved sensitivity (94.6%) and specificity (88.9%) for diagnosing skin malignancies while preserving tissue integrity at both microscopic and macroscopic levels. Concluded that LIBS is an attractive candidate for skin treatment and has real potential in the field of dermatology. However, the primary concern is that these models have not been evaluated on real patient data (from hospitals) and have not been consulted with practitioners to determine whether the classifiers are specific to such cancers or originate from other sources. These models are predictive and unable to provide causal explanations. To date, LIBS-ML studies have not incorporated demographic, medical history, social activity, genetic, and environmental factors of patients and participants into the training and testing of algorithms. Causal inference models have the potential to incorporate the mentioned factors and play a pivotal role in understanding disease propagation [181].
Kiss et al. (2021) [51] conducted a spectral study, and the analysis of four biogenic elements (Mg, Ca, Na, and K) showed higher concentrations of Mg, Na, and K in tumour-affected tissues. The progression of the tumour was examined by capturing biotic elemental (Mg, Na, and K) imaging. It was correlated with the concentration and spatial distribution of elements within malignant tissues, as illustrated in Figure 4. The distribution and proliferation of calcium are not well understood outside the tumour region. The authors declared the novelty of their work in hyperspectral data processing and k-clustering analysis for discriminating cutaneous tissue from its healthy surroundings. In research work, LIBS-3D imaging involves a complex sample preparation procedure, such as FFPE, which contrasts with the fact that LIBS typically require minimal or no sample preparation. An experimental method is lengthy, as it involves background correction for every single laser shot. The destructive effect on tissue, spatial resolution, and self-organising map (SOM) data interpretation information is not reported for in-depth understanding. Choi et al. (2021) [82] performed ultraviolet (UV-fs-LIBS) imaging of melanoma-affected skin tissue to classify melanoma and dermis. During the experiment, femtosecond laser operated at the UV region (343 nm) with a laser energy of 50 µJ provided a consistent crater pattern with minimal debris on the periphery of the exposed tissues. Spectral analysis revealed the intensity ratio I K / I CN and peak intensity ( I CN ) of the element as a classifier to identify melanoma regions. LIBS imaging was performed, and results were validated with hematoxylin and eosin (H&E) staining images, confirming the identification of cancer with LIBS. They suggested that a UV-fs-laser is more appropriate for imaging, and achieved a spatial resolution of 15 µm.
In the future, it could be a useful alternative or complementary tool for melanoma diagnosis. In this work, as in any other research, data fairness remains a concern, as biases in data (such as selection bias, social bias, and sample bias) can impact the model’s performance. Table 4 compares the LIBS research studies with the considered gold standard optical and spectroscopic methods applied in skin cancer diagnosis.

3.2. Breast Cancer

Breast cancer affects more women than any other cancer and has the highest mortality rate among female cancers. In 2020, an estimated 1,233,465 breast cancer cases were recorded worldwide among females aged 15–59 years, according to WHO statistics [189]. They reported a higher incidence rate in developed countries. However, the availability of advanced medical equipment helped them to fight against such diseases, which resulted in low casualties. Conversely, fewer cases are recorded by governing bodies in developing countries, but their survival rates are significantly lower due to a lack of awareness, inadequate medical resources, and high treatment costs [190,191]. Contemporary methods such as MRI, ultrasound, thermography, biopsy, and tissue sampling are widely used for screening breast cancer [18,22,192]. LIBS was used to study breast cancer in 2012 by quantifying the concentration of trace elements [155]. Idrees et al. (2023) [175] compare the effect of sample selection (whole blood and serum) on the diagnostic accuracy of LIBS-ML data interpretation. Accordingly, the choice of whole blood samples yields a lower error rate and might assist in early cancer screening. The research suggests LIBS cancer diagnosis may depend on the state and quantity of blood samples, but the point needs further confirmation.
Ghasemi et al. (2016, 2017) [35,47] used the classical LIBS approach to discriminate cancerous (breast, larynx, tongue, and colon) tissue from normal ones, but the measured plasma parameters and trace elemental ratios of malignant and benign tissues are very close to each other, and higher and intense for abnormal ones. The minor error in measuring plasma temperature propagates to cause high uncertainty in determining other parameters, electron density, and plasma frequency, which ultimately affect the accuracy of elemental quantification. The approach demands cross-validation, which is an extra effort and time-consuming. Table 5 lists several LIBS breast diagnostic investigation and their comparison with other state-of-the-art technologies that have evolved over the last 11 years.

3.3. Colon Cancer

Colon/colorectal cancer is the third most common cancer in terms of morbidity rate and the second leading cause of cancer-related mortalities worldwide, as statistically shown in Figure 5. Gondal et al. (2020) [50] determined the elemental concentration of heavy toxic metals in both normal and affected colon tissues using two distinct techniques (LIBS and ICP-OES). Researchers analysed cancerous colon tissue and correlated with accumulated heavy metals (Hg, Cr, Pb) using CF-LIBS. They compared the results and found that no toxic metals were detected in healthier tissue. In contrast, the presence of highly toxic metals (Hg, Cr, Pb) in cancerous colon tissues indicates tumour growth. The use of small sample sizes (n = 15), the lack of statistical analysis, and the absence of a follow-up study at a larger scale restrict the generalisation of the results and prevent confirmation of the conclusion.

3.4. Stomach Cancer

Globally, gastric/stomach cancer accounts for the fifth-highest cancer incidence. Its incidence has experienced a significant decline over the last five decades. Unfortunately, it remains a serious health concern as the third most important cause of cancer-related global deaths [197]. Spark discharge-assisted LIBS (SD-LIBS) was employed to investigate the feasibility of differentiating malignant tissues from normal stomach tissues for diagnosing gastric cancer. Neoplastic (malignant) and non-neoplastic (normal) stomach tissues were identified based on their respective distinct atomic emission spectra and measured elemental concentration. It was found that the Ca and Mg content is higher in cancerous tissues than in normal tissues for the same individual. However, only five patients participated in this research activity, which was insufficient to reach a firm conclusion. Findings are neither statistically nor medically validated and have not been confirmed using traditional elemental quantification methodologies [123].

3.5. Lung Cancer

Lung cancer has been a serious public health concern due to its high incidence and mortality rate. Globally, the lung cancer death rate in 2020 was the highest among all types of cancers, as shown in Figure 5. Approximately 1.8 million people are diagnosed with this disease, and 1.6 million die annually; these statistics are relatively higher than other oncology-related threats [20,198]. Lin et al. [159] differentiated between lung tumour tissue and boundary tissue by using LIBS-ML models. After pre-processing the LIBS spectral data of 90 tissues from 45 patients, principal component analysis (PCA) is used for dimensionality reduction, and random forest (RF) is used for feature selection. Limitations of PCA and RF (variance may not be informative, biasing towards high cardinality features, and inability to provide causal interpretation) may cause the loss of important information required for diagnosis. PCA-SVM and RF-BT models are trained and tested with compromised features. RF-BT diagnostic performance is declared to be superior to that of other models. However, the difference among model index indicators is less than 6% for all models. Li et al. (2024) [158] developed a bagging voting fusion (BVF) algorithm and proposed it as a method to overcome the limitations of single models in identifying complex cancers. In BVF, five models (SVM, ANN, KNN, QDA, and RF) were fused at both training and decision levels to process LIBS data to increase the diagnostic accuracy of multiple types of cancers (liver, lungs, and esophageal). The model claimed to achieve an accuracy of ~92% and a recall of ~93% for all serum samples, outperforming the best single model (SVM: accuracy of ~76%, recall of ~78%). To conclude, LIBS-BVF enables rapid (<3-min) and precise detection of a multitude of cancers, representing a transformative approach to clinical cancer diagnostics. In the above studies, dimension reduction approaches (PCA, FS, FE) are used to enhance model performance by sacrificing some of the information that must be given as input to build them. The high accuracy of LIBS-ML algorithms does not guarantee a reliable cancer diagnosis, as some ML models lack causal explanation mechanisms, which complicates the evaluation of their scientific validity and undermines the trust of oncologists [181].

3.6. Cancer Studies in Animal Samples

Han et al. (2016) [145] explored the feasibility of using LIBS to distinguish melanoma lesions from surrounding dermal tissue. They conducted elemental analysis on optimised pellet samples from melanoma-implanted mice. However, the limited sample size (n = 10, 1470 spectra) prevents the drawing of robust statistical conclusions, and the use of homogenised samples raises questions about clinical relevance. Most crucially, the identified biomarkers require validation in actual human melanoma cases. Further, this work lights the path forward but underscores the need for larger and more physiologically human-relevant studies. Zhao et al. (2024) [150] employed adaptive boosting (Ada-Boost) combined with a backpropagation neural network (BPNN) model for the early screening and staging of melanoma. They stated that the screening and staging accuracies of the models 83% and 96%, respectively. Based on Kruskal–Wallis (KW) statistical analysis, Ca and Na are identified as biomarkers for early screening and staging, whereas K and Mg play significant roles in the staging of melanoma. Melikechi et al. (2016) [172] developed models (LDA and RF) for differentiating blood plasma specimens of mice suffering from ovarian carcinoma. The accuracy achieved by the two models (LDA 70–75% and RF 79–81%) for the three age-specific groups, 8-, 12-, and 16-week-old mice, is similar and considered comparable across all age groups.

3.7. Miscellaneous Cancer Studies

Yue et al. (2021) [27] evaluated the effect of data reduction methods (PCA and SelectKBest (SKB) on the performance of a LIBS-NN model for early-stage ovarian cancer screening. Spectral analysis was performed on 176 blood plasma samples from cancer patients, including ovarian cysts and normal cases. Essential electrolytes in blood plasma for preserving homeostasis in the body are metal elements (K, Na, Mg, and Ca), and an imbalance in their concentration indicates a state of abnormality in patients. Cancer detection sensitivity and specificity are expressed at up to 71% and 86%, respectively. Teng et al. (2020) [165] identified glioma (a brain tumour) from its surrounding tissues based on attributed biomarkers (Ca and Mg) to the progression of abnormality. Similarly, Mohammadimatin et al. (2023) [131] detect the same biomarkers for discriminating between two types of lethal brain cancers, glioblastoma multiforme (GBM) and oligodendroglioma (OG), and healthy infiltrated brain tissues by exploiting SD-LIBS. In addition to these types of cancers, several further studies related to oral, prostate, and cervical cancers are tabulated in Table 6 and Table 7. The respective experimental information of these studies can be found in Table 3. A statistical summary of the LIBS-cancer research is visualized in Figure 6 at the end of Section 3.

4. LIBS in Calcified Tissues Analysis

Calcified tissues are biological tissues that contain the mineral hydroxyapatite (HA, Ca 10 ( PO 4 ) 6 OH 2 ). According to the definition, primarily teeth and bones are calcified. Their characteristics, including biocompatibility, slow degradation, structural support, and metabolic functions, allow them to participate in biological processes. Additionally, their chemical compositional understanding offers insight into physiology and pathological conditions [102,200]. LIBS has been exploited for such purpose; since the last decade, a variety of studies have been performed with improvisations such as the following: (i) LIBS has been employed to quantify the relative intensity of various elements, enabling detailed elemental profiling of biological and inorganic materials including studies on teeth [201,202,203,204,205,206,207,208,209,210,211,212,213], bones [79,97,214,215], and gallbladder stones [148]; (ii) investigators have integrated LIBS spectral profiles to compute area to determine the ablation thresholds of dental tissues [216,217,218], providing critical insights into laser–tissue interactions; (iii) LIBS is used to study laser-induced plasma parameters, for understanding the physicochemical properties of dental tissues [203,204,205,219] and bones [97]; (iv) CF-LIBS framework has been applied for the quantitative analysis of bones [49]; (v) CC-LIBS has been utilised to establish calibration curves for teeth [124,130] and kidney stones [220], enhancing the accuracy of elemental quantification; and (vi) LIBS signal enhancement methods NE-LIBS have been integrated to improve signal intensity for teeth [130] and bones [128], increasing the sensitivity and precision of measurements. It has also been employed for spatial elemental mapping of teeth [221], providing detailed distribution patterns of elements within the tissue (vii) LIBS data has been combined with ML algorithms to enhance the analysis of teeth [84,96,112,147], and bones [96,128,146,152], enabling automated classification, pattern recognition, and predictive modelling (viii) The use of ultrafast femtosecond lasers in LIBS has been explored to examine teeth [84] and bones [222], offering higher precision and reduced thermal damage compared to conventional nanosecond LIBS. Table 8 (in Section 4.1) and Table 9 (in Section 4.3) categorise and summarise the LIBS research work on human and animal calcified tissues, respectively. In addition, Figure 7 (at the end of Section 4.3) gives a statistical overview of recent LIBS research on calcified tissues.

4.1. Evolution of LIBS in Dentistry

In 1964, the first research on using lasers as a surgical tool in dentistry was published; at that time, the energetic pulsed ruby laser was used to eradicate caries [223]. These preliminary studies define fundamental concepts of laser dental interaction. However, its practical applications were not permitted due to the thermal damage induced in dental tissues by the available lasers in the early 1960s. Therefore, further research was carried out, leading to a significant breakthrough in this field of study with the development of laser technology in subsequent decades. Particularly, ultra-short laser pulses have proven to be practically valuable by enabling precise control over the ablation of carious tissues, allowing for the ablation of dental fillings without damaging the surrounding healthier tissues, and providing online feedback of endodontic treatment via laser-induced plasma-emitted radiations. The latter aspect is that when LIBS is introduced into dental studies to inspect the degree of thermal, mechanical, and optical damage induced in teeth by lasers. Successive LIBS research has been conducted to exploit plasma-emitted radiations as a diagnostic parameter for online monitoring of caries removal processes. Teeth are the most suitable sample for LIBS analysis, composed of four tissues: three of them, enamel, dentin, and cementum, are hard, while the fourth is soft, known as pulp. The outer layer, enamel, contains ~95 wt% hydroxyapatite (HA), ~4 wt% water, and only ~1 wt% organic substances; the middle layer, dentin, is comprised of ~70 wt% HA, ~10 wt% water and ~20 wt% organic matter; tooth root covering tissue cementum is least mineralised with ~45 wt% hydroxyapatite, ~35 wt% organic matter, and ~20 wt% water; the inner part, pulp, is non-mineralised and primarily composed of blood vessels and nerves [224]. The fundamental building block of teeth is HA crystal; its presence was confirmed by LIBS detection of its constituent elements, as evidenced by intense emission lines of Ca, P, O, and H [204,217].
Caries is the most common reason for dental infections (pericoronitis, periodontitis, pulpitis, osteomyelitis, gingivitis, cellulitis, and cracked tooth syndrome) caused by the quick reproduction of pathogens (bacteria and viruses) on tooth surfaces that lead to the demineralisation of dental tissues [16]. Caries alter the dental elemental composition, and loss of Ca is substituted by other cations detected by the LIBS analytical technique. There is no considerable relation between chemical properties and the type of teeth. However, it is one of the requirements to know the chemical profile of teeth elements to design a novel class of biomaterials for clinical purposes [225]. It is essential to understand the role of elements in human teeth to comprehend the purpose behind their detection. Multiple trace elements are indispensable for human growth, particularly in ensuring normal tooth development. Excess or deficiency of elements can be detrimental to the body, having a direct association with environmental conditions, lifestyle habits (such as diet and drug use), and diseases (diabetes, hypertension) [226]. These elements enter to the human body and are deposited into the tooth, causing demineralisation or decalcification processes that primarily lead to tooth decay due to a deficiency of calcium ions. Trace elements must be within the permissible limit provided by WHO; otherwise, adverse effects on teeth have been reported [227].
LIBS can discriminate between healthy and carious tissues by determining plasma parameters and elemental variations produced by caries-induced demineralisation, which are matched with the constituent concentrations of healthy tissues or standard references. Several such studies exist in the literature and are reported in this article: Khalid et al. (2015) [211] measured plasma parameters and elemental composition of healthy and carious tooth tissues (enamel, dentin, and cementum). They reported that Ca concentration is lower in carious enamel than in healthy ones compared with the artificially fabricated pellet reference (CaCO3). The highest contents of microminerals (Pb, Sr, Zn, and Fe) and low concentrations of Ca were found in enamel and turned into the most carious-affected tissues among other dental tissues. Batool et al. (2021) [204] recently measured the temperature and electron density of an Nd:YAG laser with both its fundamental and second harmonic outputs. They concluded that Nd:YAG laser in the second harmonic mode induced less temperature on the enamel surface and higher electron density than its fundamental mode. In laser dentistry, the second-harmonic Nd:YAG laser is declared as a suitable candidate for cavity preparation, as it produces well-defined ablations while preserving adjacent healthy tissue. The observed effects correlate with distinct electron densities between laser modes, corresponding to their emitted photon energies.
The study by Gazmeh et al. (2015) [36] demonstrated that LIBS combined with PLS-DA could effectively distinguish between sound teeth and those with caries. It was worthwhile to use statistical analysis to classify samples because there was little difference in spectral line intensities (P, Ca, Mg, Z, Sr, C, Na, H, and O) between healthy and carious tissues. Elemental emission line intensities are considered variables to build the model. The author claims high model prediction accuracy in classifying unknown samples (healthy and carious tooth samples. Hadeethi et al. (2016) [216] and Mustafa et al. (2022) [217,218] have measured the dependence of Ca emission intensities in dental tissues (enamel, dentin) on the fluence of the laser pulse using an Nd:YAG laser-based LIBS. The thresholds of laser fluence for calcium in enamel and dentin tissue were found to be 1.41 J/cm2 and 0.38 J/cm2, respectively. In the case of enamel, a higher threshold laser fluence is required for Ca emission in enamel than in dentine due to its calcified nature and rigid structure. To date, the literature lacks sufficient data on the structural and chemical composition changes in teeth associated with ageing. It cannot be advised to children and adults on similar preventive measures and treatment procedures in modern dentistry. Therefore, age-specific morphological and chemical analyses of teeth are strongly recommended to support the development of optimised materials, including anti-caries hygiene products and dental filling materials [228]. Briefly, 28 LIBS-human dental research works, along with 5 studies on human bones and kidney stone examination, are summarised chronologically in Table 8.
Table 8. LIBS analyses of human calcified tissues.
Table 8. LIBS analyses of human calcified tissues.
TissuesRef.Purpose/ApplicationMethodFindingsComments
Teeth[221]Imaging of teeth using long-pulse laser to evaluate LIBS performance/dental anatomyLIBS elemental imagingVisualise variation in elemental distribution from enamel to dentineLimited to major elements, alternative needs to be explored for trace elements
[106]Migration of species from dental restorative materials to tooth matrix/clinical applications (pathological identifications)Measuring line intensitiesOne source of trace elemental accumulation in tooth matrix is dental filling materials.Hg in healthy tissues not detected, LOD higher than concentration (0.10 μg g−1)
[124]Elemental quantification of teeth using appropriate calibration method/pre-clinical dentistryCC-LIBSQuantification of Ca, Mg, C, and Zn, error rate measured for different methodsConsiderable error range for LIBS with respect to standard techniques (ICP-MS, EDX); inaccuracies due to matrix effect
[130]Effect of NPs deposited on teeth on LIBS performance/nano-dentistryCC-LIBSSignal enhancement and improved calibration for Ca with ZnO NP Control of size, shape, and distribution of NPs difficult; reproducibility is of concern
[147]Sex identification/odontologyLIBS-ANNCa, Mg, P, Sr, and H used for classification; accuracy for males 98%, for females 99%Small sample size (n = 10) and limited number of spectra (N = 500) are insufficient for definitive conclusion
[201]Classification of teeth according to age and sex for healthy and carious teeth/pre-clinical dentistryIntensity ratio of spectral linesHigh concentration of Mg and Pb in carious teeth; trace elements: intensity decreases with age, female teeth have higher intensity Small sample size (n = 8); no statistical, cross, and clinical validation performed
[205]Effect of laser wavelength and irradiance on plasma parameters/pre-clinical dentistryBoltzmann and Saha–Boltzmann plot methodsHigh plasma temperature at 1064 nm; plasma frequency and Debye length increase with irradianceLimited clinical relevance; results published earlier confirmed
[219]Monitoring the migration of dental filling material into tooth/laser dentistryIntensity ratio and Stark broadening methodsPlasma temperatures for composite, dentine, and amalgam are different Element concentration not quantified; limited to ex vivo analysis; other relevant materials are to be analysed
[213]Effect of antimicrobial agents on remineralisation of dentine/tissue restorationChanges in intensity vs. wavelength emission profilesBiomimetic remineralisation of carious dentine by activating an antimicrobial agent using fs laserUse of different materials will change results (matrix effect); no statistical cross validation and clinical validation
[229] Variation in elemental composition in ankylotic tissues/orthodonticElemental imagingAnkylotic tissues have higher concentration of Ca and PLimited spatial resolution of 30 µm (Ca, Mg, P); role of Mg is to be explored
[217]Variation of Ca and Mg in dental tissues after laser irradiation/laser dentistryIntegrated peak area vs. laser energy densityThreshold laser fluence for Ca in enamel and dentin determined Threshold laser fluence for Mg not detected (instrumental limitations)
[230]Age and sex identification/orthodontic treatmentsMeasured average intensitiesCa, P, and Fe concentrations decrease with age, higher concentrations in females than malesVariation is linked to orthodontic abnormalities; cross validation and clinical validation should be considered
[112]Diagnosis of dental pathologies/pathological identificationRelative intensityCa/P ratio decreases in presence of plaqueCa/P signal range is close for healthy and pathological regions
[204]Characterisation of dental tissues/clinical dentistry (dental implants and cavity preparation)intensity ratio and Stark broadening methodsPlasma temperature higher and electron density lower for Nd:YAG laser wavelength 1064 nm than for 532 nmOutcomes are to validated by practitioners using other lasers (CO2, diode, Er:YAG) used in laser dentistry
[107]Assess interaction of laser beam with dental tissues and dental material/dental health and safetyLIBS and photoacoustic sensorMore intense plasma on carious than on healthy tissues. Release of toxic elements (Hg, Ag, Cu, Sn) from amalgam hazardousUsed approximation for acoustic wave propagation may not be appropriate for clinical applications
[202]Detect early signs of osteoporosis in periodontal patients/clinical dentistryMean spectral intensityLower (Ca) and higher (K, Mg) content associated with osteoporotic group compared to control groupControl group had periodontal disease, which can have specific effects of osteoporosis (on periodontal tissues)
[231]Examination of dental tissues, enamel (apical and buccal) and dentine/pre-clinical dentistryIntensity ratio of spectral linesApical enamel is the hardest among buccal enamel and dentineNo statistical, cross, and clinical validation; in vitro analysis
[84]Thermal effect of fs laser ablation on teeth for caries removal/laser surgerySimulation method (thermal model)Minimal thermal damage to surrounding nerve tissues; acceptable removal ratesOptimisation of laser fluence below carbonization threshold of each tissue is challenging
[80]Evaluation of diffusion of mercury to dental tissues/tooth restorationFs-LIBSHg penetration depth for deciduous and permanent teeth determinedSpatial resolution limited to 100 µm; penetration depths of other metals in tissues are to be determined
[216]Ablation threshold fluence for enamel and dentine/laser dentistryEmission intensity vs. laser energy density Ablation threshold for enamel and dentine determined Ablation damages tissues inducing variations in surface topography and structural morphology
[232]Elemental variations in teeth associated with cariogenic and periodontal pathologies/oral surgeryLIBS intensity of emission linesHigher concentration of C, O, K, F, Na aggregates in periodontal teeth compared to those with cavities Statistical cross validation and clinical validations to be done; in vitro analysis
[203]Detection of toxic elements in smokers, non-smokers, and teeth/periodontal probingCalibration curve for quantificationHigher concentration of hazardous elements (Pb, Cd, As) in smoker group than in non-smoker group.No statistical analysis performed; correlation of results to smoking habits or chronic periodontitis difficult
[211]Examination of deciduous teeth to measure toxic elements/dental health and safetyComparing spectra with matrix-matched referencesHigh amounts of toxic elements in enamel make it the most affected tissue in the tooth Small sample size (n = 4); no statistical, cross, and clinical validation; in vitro analysis
[210]Improve sensitivity of system for caries identification/optical feedback in laser dentistryMeasuring intensity variationsIntensity ratio Zn/Ca increases in presence of cariesUse of two different lasers (Nd:YAG, Er:YAG) makes treatment difficult
Bone[152]Performance evaluation of ML models in bone classification/pre-orthopaedic surgery LIBS-ML (PLS-DA, LDA, LR, SVM, SIMCA, CART, NN)NN has exceptional performance in terms of sensitivity, robustness, generalisationAccuracy of NN (100%) may be due to overfitting; other models’ accuracy is 42–66%
[83]Discriminate between normal and pathological bone/optical feedback in orthopaedic surgery Fs-LIBS-PCAHigher Mg intensity relative to Ca in pathological bone compared to normal boneUnable to detect bones suffering from severe pathologies
Kidney stones[148]Discrimination of gallbladder stones (mixed vs. pigment GB 2)/diagnosis LIBS-PCA and PASVariation in intensity of Ca, Sr, K, and CN; absence of calcium phosphate in GB 2Different sample preparation methods for LIBS and PAS measurements
[46,220]Measure major and trace elements in gallstones/diagnosis LIBS/WD-XRF/FTIRTrace element (Zn, Pb, Cr, Cd) concentration exceeded safety limitLIBS LOD for trace element 10–19 ppm

4.2. LIBS in Bone and Kidney Stones Analysis

The major chemical compounds in bones are ~69 wt% hydroxyapatite, ~10 wt% water, ~20 wt% collagen, and ~1 wt% proteins [224]. Physiological bone assessment is of great importance in defining therapeutic procedures. To detect irregularities, it is necessary to measure imbalances in chemical compositions and observe morphological alterations in cases of bone abnormalities. Moncayo et al. [152] evaluated LIBS-ML algorithms (soft independent modelling of class analogy (SIMCA), PLS-DA, LDA, classification and regression trees (CART), LR, SVM, and NN) for their classification accuracy on human bone samples. NN performance is relatively satisfactory in discriminating highly similar human bones. Ruby et al. [83] proposed fs-LIBS as a feedback control system for the removal of primary bone tumour, where a higher Mg peak intensity relative to Ca is related to abnormality. But the study is unable to define diagnostic criteria for secondary bone tumour due to their inherent heterogeneity. Overlapping spectra on the PCA plot indicate that fs-LIBS may not be appropriate for late-stage tumour detection. Alternatives need to be explored by refining sample preparation strategy and spatial mapping of LIBS signals. Several articles on LIBS-bone analysis are available in the literature for archaeological, environmental, and forensic applications; however, from a health and medical perspective, their application to human bone specimens is rare.
Zainab et al. [148] performed compositional analysis on two gallbladder stones (mixed (GB 1) and pigment (GB 2)) using two spectroscopic techniques: LIBS for elemental analysis and Photoacoustic spectroscopy (PAS) for molecular information. LIBS-PCA was performed based on the variation in emission intensity of classifiers Ca, Sr, K, and CN for GB 1 and GB 2 to classify them. PAS molecular analysis of GB 1 (calcium carbonate, calcium phosphate, bile acid, bilirubin, and fatty acid) and GB 2 (calcium carbonate, bile acid, bilirubin, and fatty acid) revealed the absence of calcium phosphate in the latter. Gondal et al. [220] quantified carcinogenic metal content in three kidney stones (extracted from three different patients) using SP-LIBS with Nd:YAG laser (266 nm, 8 ns, 20 Hz, 25–50 mJ). Calibration curves are drawn to quantify the concentration of detected metals (Ca, Zn, Cr, Cd, and Pb), and LIBS outcomes are validated with ICP-MS. They concluded that metal concentrations in stones are beyond the safety limits of U.S. Environmental Protection Agency (EPA) and Food and Drug Administration (FDA). Pilot studies [148,220] have yet to demonstrate clinical implications, so they require further extension to broader populations, including diverse demographics, various stone compositions, and different ethnic groups.

4.3. LIBS in Animal Calcified Tissues

Beverage consumption is trending for humanity’s pleasure, relief, and to maintain energetic health. Its effect on teeth was enlightened by Manno et al. (2018, 2020) [206,207]. The impact of coffee immersion on rat teeth (in vitro) was investigated through a comparative study of coffee consumption by different age groups and their respective age-matched control groups. Coffee can induce decalcification of teeth, leading to erosion and exposing the dentin by thinning the enamel layer. In vivo and in vitro, hot and cold beverages (green tea and water) influenced rat teeth. Drinking green tea at room temperature protects teeth against erosion. However, caution is required when consuming hot tea, as it can cause enamel degradation due to the interaction between hot green tea catechins and hydroxyapatite (HA) in the teeth. To reinforce this, atomic force microscopy (AFM) is combined with scanning electron microscopy (SEM) to evaluate surface topographic characteristics and quantify erosion. Refraining from hot beverages reduces direct exposure to teeth, consequently preserving the dental structure, making them less likely to decay, and ensuring their long life [206].
Tariq et al. [233] evaluate the hardness of calcium-rich tissue by measuring the Ca/P ratio of HA (extracted from bovine bone) using laser-induced plasma emission spectral lines. They reported a relative error of <6% between LIBS and EDX findings for Ca/P. Although high accuracy and repeatability in measurements were obtained and verified by varying plasma, further research studies are required to verify the authenticity of using LIBS to determine Ca/P.
Roldan et al. [49] applied CF-LIBS for elemental analysis of rib bone (wild boar) by measuring plasma temperature and electron number density. Spectral analyses were performed across the ultraviolet to near-infrared (UV-NIR) and vacuum ultraviolet (VUV) ranges, as light elements are challenging to detect in UV-NIR due to the low signal-to-noise ratio for emission lines of such elements within a complex matrix. A broad spectral range enabled the detection of more elemental emission lines; even P is detected in the VUV region, further allowing the calculation of the atomic ratio of Ca/P (1.6 ± 0.4) for a gate delay of 500 ns to assess bone hardness. Salam et al. [128] utilised NE-LIBS, biosynthesised Ag-NPs sprinkled on animal feed and bovine bone (ancient and modern) samples to enhance the emission intensity of spectral lines in NE-LIBS spectra. PCA was applied to the data to distinguish between bone and fodder types for assessing the feeding strategy of livestock. Spectrochemical analytical data revealed the presence of numerous common elements in bones and feed that can inform feeding strategies for animals regarding their health throughout their lifetime. Additionally, this study can be beneficial for human health, as it depends on farm animals as their primary food source.
Table 9. LIBS analysis of animal calcified tissues.
Table 9. LIBS analysis of animal calcified tissues.
TissuesRef.Purpose/
Application
Method FindingsComments
Rat dental tissues[206,207]Effect of beverages on rat dental tissues/veterinary dental careLine intensity ratioCoffee induces loss of Ca and P; decalcification of enamelMolecular mechanism of enamel protection by green tea needs to be assessed
Chameleon oral tissues (tooth and bone)[212]Age-related variation in tooth-bone fusion area/physiological conditionElemental imagingIntensity of Ca, P, and Mg correlated with age in junction area of tooth and boneComparative analysis required to support claim that ankylosis is pathological state in mammals but is physiological for chameleon
Pig bones[215]Effect of acidic environment on bones/physiological stateLine intensity ratioIntensity of Ca II/Ca I increases with concentration of sulfuric acid and time; hardness of bone surface is associated to calcium sulphate Clinical relevance not demonstrated
Porcine bones (femora)[115,146]Differentiate hard (bone) and soft tissues (muscle and fat)/laser osteotomyFO-LIBS-MLSensitivity of canonical DFA model for differentiating fat (84.6%) is lower than that of bone (100%)Significant difference in model’s accuracy for soft tissue analysis revealed limitations of LIBS for such tissue examination
Boar bones (rib)[49]Analysis of light elements/Orthopaedic CF-LIBSRatio of Ca/P measured Difficult to fulfil prerequisites for CF-LIBS of biological samples; uncertainties of plasma parameters; comparative validation required
Bovine bones (femur)[128]Discrimination of different bones and different fodders/health and medicalNE-LIBS-PCADeposition of Ag NPs on bones and fodder enhances sensitivity of LIBSAging of bone and metabolism of cattle affect element concentration; complicates correlation of feeding strategy and bone composition
Pig (oral mucosa, peripheral nerve, dental pulp, dentine, enamel, cortical bone, cancellous bone)[96]Differentiation of hard and soft tissues/laser surgeryLIBS-PCA-LDARatio of Ca/C identified as classifier (higher in hard tissues than in soft ones)Sensitivity and specificity of LDA below 70% for enamel-dentine pair classification is serious issue in hard tissues examination (matrix effect lower than in soft tissues)
Bovine (bones and muscles)[222]Effect of laser repetition rate on ablation of dry bone/laser surgeryFs-LIBSHigher repetition rate allows fast cutting of bone (fluence above thresholds)Temperature rise during ablation of bone should not exceed threshold for irreversible damage of nerves
Rat bones (thigh)[97] Role of lead in bone/animal health and medical scienceRelative line intensityCa and Mg signals decrease with increasing Pb concentration Error (18%) of measured plasma parameters high; quantification of elements hindered
Chicken eggshells[234] Discrimination of organic eggs from inorganic/food quality control: health and safetyLIBS-NNClassification accuracy of NN for eggshell across various groups is 100%Research focused on Ca; quantitative analysis of other elements (Mg, Fe, Al, Sr, Zn, Mn) to be performed
Chicken eggshells[235]Measurement of plasma parameters/pre-veterinary Boltzmann plot and Stark broadening Determination of temperature and electron number density and of inverse Bremsstrahlung absorption coefficientLarge uncertainty (5–50%) of used transition probability data limits precision of determined parameters; practical application is lacking
Chicken eggshells[236] Determination of heavy and trace elements in organic and regular chicken shell/food quality control: health and safety CF-LIBS (OLCF and SC)Ba detected only in organic shell; SC-LIBS is more reliable for quantification of heavy metals and trace elementsLOD values for toxic metals (Pb, Cd, Hg, As) remain uncertain
Figure 7. Summary of journal-published articles on LIBS of calcified tissues from 2015 to 2025 (50 original articles). Statistical information about (a) number of publications per year; (b) number of articles on teeth, bones, kidney stones, and others; (c) adoption of methodologies and validation procedures (cross, statistical, clinical); (d) number of studies performed on human and non-human tissues. Types of calcified tissue and related references: Teeth [80,84,106,107,112,124,130,147,201,202,203,204,205,206,207,210,211,212,213,216,217,219,221,229,230,231,232,237], bones [49,83,96,115,128,146,152,199,206,207,212,215,222,233,238], kidney stones [46,148,220], and others [234,235,236,239].
Figure 7. Summary of journal-published articles on LIBS of calcified tissues from 2015 to 2025 (50 original articles). Statistical information about (a) number of publications per year; (b) number of articles on teeth, bones, kidney stones, and others; (c) adoption of methodologies and validation procedures (cross, statistical, clinical); (d) number of studies performed on human and non-human tissues. Types of calcified tissue and related references: Teeth [80,84,106,107,112,124,130,147,201,202,203,204,205,206,207,210,211,212,213,216,217,219,221,229,230,231,232,237], bones [49,83,96,115,128,146,152,199,206,207,212,215,222,233,238], kidney stones [46,148,220], and others [234,235,236,239].
Molecules 30 04176 g007
Abbasi et al. [115] explored LIBS as a feedback system for differentiating bone from soft tissues during laser-osteotomy, aiming to improve surgical precision and safety. The classification of three sample groups (bone, muscle, and fat) based on the intensity ratio of selected peaks was performed using discriminant functional analysis (DFA). The ROC curve analysis depicted an accuracy of 99% for hard–soft tissues (bone–muscle and bone–fat) and a relatively lower accuracy (90%) associated with soft tissues (muscle–fat). When ex vivo results are applied to in vivo analysis, the presence of various biological fluids and dynamic physiological conditions in a living body presents significant challenges that are not addressed in this study.
Ying et al. [239] used LIBS to track the shell growth in sea shells (ezo scallop shell) from biological perspectives. They proposed that Ca intensity is constant on the shell surface and can be used as an internal reference, whereas Sr content increases with the growth of the shell. Applications of LIBS in different animal tissues (teeth, bones, and eggshells) from medical perspectives are tabulated in Table 9.

5. LIBS Hybrid Technology

Biomedical examination requires a comprehensive analysis of specimens, encompassing their elemental, molecular, and structural components. However, it is challenging to obtain such detailed information using standalone techniques due to their limitations. Hence, hybrid models in which two or three methods are integrated into a single device are extensively used to obtain comprehensive information and achieve the best possible outcomes [42]. LIBS can be established as a diagnostic tool that complements other analysis techniques, such as RS and ICP-MS, rather than replacing them [68]. For a better understanding, some structural designs of LIBS-coupled techniques used in various medical applications are illustrated in Figure 8.
Integration of LIBS and Raman technology enables elemental and molecular information about tissue composition. Khan et al. (2022) [174] fused LIBS and Raman spectral data for the classification of melanoma tissues using ML models (ELM, PLS-DA, KNN). The higher intensity of Mg in the LIBS spectra and the spectral shift related to amide III and lipid molecules in the Raman spectra of tumour tissues are identified as biomarkers. Matrix effect should be addressed when tissues were collected from different body parts (lymph node and skin). It highlights the issue of confounding results, as spectral differences can reflect the anatomical site rather than the state of cancer alone. The reported average classification accuracy for ELM is 99.3%, which raises concerns about overfitting due to the small data set (10 samples from 2 patients). Additionally, statistical cross-validation and clinical validation are required for more reliable analysis. Lin et al. (2024) [121] used a bimodal approach (LIBS + Raman) to improve the diagnostic accuracy of the CNN model. For this purpose, LIBS (Ca, Mg, Fe, and Cu) and Raman (phenylalanine, tyrosine, tryptophan, amide III, and protein) identifiers were measured for lung cancer staging. These biomarkers are not cancer-specific and can vary due to obvious reasons (dietary habits, environmental conditions, and tissue heterogeneity), which may potentially confound staging accuracy. The bimodal information is combined using a decision-level Bayesian model, which enhances the classification efficacy up to 99% in the CNN framework. The model may oversimplify the biological interaction, and the results need to be validated against the gold standards (histopathology, immunochemistry, and CT).
Batch effect is caused by operator variability and instrumental variations, consequently reducing the performance of diagnostic models. Shi et al. (2025) [163] fused LIBS and FTIR spectral data from serum samples of breast cancer patients to enhance the detection ability of CNN and GRAN by correcting the batch effect. Classifiers from LIBS (Na, Ca, and Mg) and FTIR (Amide 1, II, and A) were identified. Although technical details of the algorithms are not provided, an in-depth knowledge of data science is required to understand the mechanism behind the improvement. GRAN model accuracy in detection is 89% and validating the model in a clinical setting at a large scale is recommended for generalisation.
Sasazawa et al. (2015) [210] developed an optical-fibre-based LIBS system combined with coaxial Argon gas flow aiming for in vivo analysis of carious enamel, as illustrated in Figure 8a. Specimens are segregated into three groups based on the level of decay: (i) early stages of decay (decay and cavities stayed only on enamel), (ii) advanced stages of decay (cavities reached dentin), and (iii) healthy teeth (without decay or cavities). Two different lasers are used for experimentation: conventional dental Er:YAG laser (2.94 µm) is used for ablation purposes, and Nd:YAG laser (1064 nm) for LIBS analysis. Zn is strongly detected in an early stage of dental decay and considered a biomarker of decayed teeth. The intensity ratio of Zn/Ca increases as caries develops, reaching a high of 0.013 for dentin caries, whereas lowest is 0.0044 for healthy teeth. However, there is doubt about the presence of Zn whether it originates from stains or dietary habits, and the method is unable to detect it in dentine caries. Research lacks histological confirmations, and reliance on a single element for treatment is quite risky. These issues need to be addressed before transitioning from in vitro to in vivo analyses.
Khosroshahi et al. (2020) [107] designed LIBS-integrated photoacoustic (PA) sensors to measure the safety of laser-dental treatments, as shown in Figure 8c. Polyvinylidene fluoride (PVDF) is a common material used in biomedical applications as a PA sensor. It is used to detect the corresponding stress waves generated by heating as a result of photon absorption and propagating through the material. LIBS-PA provides insights into the mechanical effects and the elemental compositional changes that occur during laser interaction. Pressure due to thermal waves increases non-linearly with laser fluence, and a maximum pressure of 8 kPa was measured for healthy teeth. Plasma colour varies with pulse numbers, indicating different temperatures of ejecta. In relation to healthy teeth, the formation of intense plasma in carious teeth is attributed to the presence of C and Sr. In the case of amalgam combustion, the intense emissions of Hg, Ag, Cu, and Sn lead to a high plasma temperature of 15,000 K. Inhaling these heavy metal elements causes serious health problems. The limitations of this research include visual sample inspection, a large variation in plasma temperature (590–15,000 K) without measurement of errors, ex vivo analyses, several assumptions, and a lack of clinical relevance, which may prevent this system from being adopted in dental clinics.

6. Limitations, Possible Solutions, and Recommendations

Spectrochemical analysis, including LIBS, faces several challenges, such as background drift, low signal-to-noise ratio, peak shifting, peak broadening or narrowing, and peak overlapping, which affect the sensitivity, specificity, stability, and reproducibility of analytical signals [142]. Potential causes of these issues include non-optimised laser parameters, variation in ablation efficiency, non-ideal plasma conditions, the transient nature of plasma, matrix effects, fractionation, spectral absorption, heterogeneous sample, complex data analysis, instrumental limitations, and environmental instrumental influences. It is necessary to resolve the above issues to enhance the diagnostic properties of LIBS and fully realise its potential as a stand-alone or complementary technique. Many thousands of laser shots can be beneficial in addressing sample inhomogeneity issues. Pre-sample treatment has improved LIBS signal sensitivity by intensifying plasma emission, particularly in biological studies [241]. In addition, several potential approaches (mentioned in Section 2.1) can be used to minimise matrix effects.
AI-assisted LIBS methodology in diagnosis requires expertise to handle complex data and avoid poor feature selection, which can lead to overfitting. Models’ generalisability is missing due to the choice of a smaller sample size. They often lack interpretability, which makes it challenging to correlate spectral features with biological specimens [52]. Despite encouraging outcomes of ML models being reported, many studies offer limited discussion on model interpretability and potential clinical integration. Thus, future research should prioritise transparent model evaluation, cross-institutional validation, and alignment with clinical diagnostic requirements to ensure robustness and translational value. Reinforcement learning (RL) and imitation learning (IL) algorithms can explain causal effects, provided that the system designs in this context are suitable. The approach has the potential for more accurate disease classification, improved efficiency of treatment procedures, and personalised medical prescriptions [181].
Descriptions of multiple studies suggest that investigators still lack a unified understanding of which abnormalities exhibit variations in element composition or the extent to which these variations occur. It has been more than two decades since LIBS detected the first cancer, and it is still in a research phase that requires confirmations, establishing protocols, and consulting with experts. Most studies are conducted in vitro, which means specimens are extracted from the body, and characterisation is performed in a laboratory under controlled conditions. The issue is whether their findings can be applied to in vivo analysis in real-world scenarios. The studies are recommended where the researchers reproduce their results fairly on real patients. The incorporation of improved spectroscopic instruments, LIBS configurations, multiple LIBS hyphenated techniques, and LIBS-ML has tremendous potential to overcome the shortcomings of LIBS for diagnostic applications. The handheld LIBS technology needs to be upgraded to eliminate the limitations that prevent it from providing the desired features for biological sample diagnosis, including in-depth profiling, in situ detection capability, a micro-destructive nature, minimal or no sample preparation, and the ability to perform in vivo analysis under ambient conditions [242].
Despite advancements in cancer research, certain cancers, including pancreatic, paediatric, gynaecologic, prostrate, neuroendocrine, calcified tissue cancer, rare cancers, penile, anal, salivary gland, and small intestine, have not been extensively studied via LIBS. Similarly, a range of calcified tissues, including calcified tendons and ligaments, dermal bone structures, coral skeletons, exoskeletons, eggshells, and otoliths, can be analysed using LIBS technology to fill existing research gaps and understand the compositional heterogeneity of these tissues. Ideally and potentially, LIBS in vivo surgery is possible for surgeons and dentists, allowing them to ablate tumours or carious tissues while analysing ablation events and categorising spectroscopically, shot by shot, to obtain real-time feedback on the tissue being removed.

7. Conclusions

LIBS is an innovative chemical elemental analysis technique progressing in diverse research fields, including biomedical sciences. It is a well-established technique from a qualitative perspective, but reaching quantitative goals for biological specimens requires considerable effort to overcome the current challenges of self-absorption and matrix effects. Identifying elemental biomarkers associated with cancer pathologies and calcified tissues helps in early disease diagnosis and adaptive therapeutic procedures, ultimately reducing mortality rates. Standalone LIBS cannot achieve exemplary results because pathological identifiers are in minute quantities. Therefore, employing signal-enhanced variants (e.g., DP-LIBS, EF-LIBS, and NE-LIBS) for spectral collection, advanced data processing models for spectral analysis, and ultrashort fs LIBS for high resolution elemental imaging are significant. In DL architectures, integration of reinforcement learning and transfer learning methods can provide more realistic data interpretation. LIBS, in conjunction with medical equipment, can provide advanced diagnostic capabilities for diseases. An analysis of published data reveals that the application of LIBS in cancer research and studies on calcified tissues remains limited, with approximately 45 and 50 original articles published in high-ranking journals, respectively, between 2015 and 2025. Small sample size (n) in cancer studies ranging from n = 2 to n = 35 [31,39,50,81,117,119,120,123,125,131,150,153,164,165,166] and calcified tissues investigation ranging from n = 1 to n = 15 [49,96,97,124,128,130,146,147,201,206,211,220,222] is insufficient for robust statistical analysis to draw a definitive conclusion. To date, no studies have been identified that simultaneously validate LIBS outcomes across all three statistical, medical, and clinical domains. Most research articles lack medical validation of LIBS results (comparison of outcomes with standard diagnostics methods MRI, biopsy, and histopathology) and clinical validation (usage of tools in hospitals on real patient populations). Extensive in vivo research is required for cancer diagnosis and calcified tissue analysis, utilising large samples and incorporating appropriate theoretical modelling, statistical analysis with causal inference, and medical relevance to establish standardisation protocols and guidelines. LIBS cannot be applied in real biomedical applications, particularly in hospitals, until it proposes more reliable analysis for complex biological tissues compared to many mature elemental techniques that are available and considered standard reference techniques (AAS, LA-ICP-MS, and XRF). LIBS-hyphenated techniques can be combined with optical spectroscopic techniques and medical equipment to achieve diagnostic efficiency. It is available in physicochemical laboratories, and analytical results are reliable. There is perseverance in transferring LIBS from lab-based research to real medical diagnostic applications. Ideally, LIBS is envisioned to be in the hands of doctors, nurses, pharmacists, and practitioners within the next couple of years or more; however, in reality, challenges can persist in the field and pose significant obstacles to unleashing its true potential.

Author Contributions

Conceptualization, M.M.D., K.S. and J.D.P.; methodology, M.M.D., K.S. and J.D.P.; software, M.M.D.; validation, K.S., J.D.P. and D.Z.; formal analysis, D.Z., M.Q., M.S.A.R., S.M., Q.Y. and B.H.; investigation, M.M.D.; resources, K.S. and J.D.P.; data curation, M.M.D.; writing—original draft preparation, M.M.D.; writing—review and editing, M.M.D.; visualization, M.M.D.; supervision, K.S. and J.D.P.; project administration, K.S. and J.D.P.; funding acquisition, J.D.P. All authors have read and agreed to the published version of the manuscript.

Funding

Supported by Johannes Kepler University (JKU) via the Johannes Kepler Open Access Publishing Fund and the federal state Upper Austria.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be available on request.

Acknowledgments

During the preparation of this manuscript, the authors used ChatGPT (https://www.chatgpt.com, Open AI, Inc., San Francisco, CA, USA) for fine-tuning and refining the scientific English and DeepSeek (https://www.deepseek.com/, Hangzhou DeepSeek Artificial Intelligence Co., Ltd., Hangzhou, China) for the purpose of better writing quality. The authors have reviewed and edited the output of such tools and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AASAtomic absorption spectroscopy
AdaBoostAdaptive Boosting
AFMAtomic force microscopy
AIArtificial intelligence
ANNArtificial neural network
AUCArea under curve
BPNNBack propagation neural network
BP_AdaBoostBackpropagation neural network with adaptive boosting
BVFBagging voting fusion
BTBoosting tree
BCCBasal cell carcinoma
CARTClassification and regression tree
CC-LIBSCalibration curve LIBS
CF-LIBSCalibration-free LIBS
CLSMConfocal laser scanning microscopy
CNNConvolutional neural network
CRMConfocal Raman microspectroscopy
CTComputed tomography
DCSDual comb spectroscopy
DFADiscriminant function analysis
DLDeep learning
DLMDeep learning model
DMFDermatofluoroscopy
DNNDeep neural network
DP-LIBSDual/double pulse LIBS
EDS/EDXEnergy dispersive X-ray spectroscopy
ED-XRFEnergy dispersive X-ray fluorescence
EF-LIBSElectric field-assisted LIBS
EPAEnvironmental Protection Agency
Er: YAGErbium-doped yttrium aluminium garnet
ELISAEnzyme-linked immunosorbent assay
FDAFood and Drug Administration
FFFFront face fluorescence
FFPEFormalin-fixed, paraffin-embedded
FsFemtosecond pulses
Fs-LIBSFemtosecond LIBS
FS-SVMFeature selection followed by support vector machine
Fs-DP-LIBSFemtosecond double-pulse LIBS
FTIRFourier transform infrared
GISTGastrointestinal stromal tumour
GRANGradient reversal adversarial network
HAHydroxyapatite
HAZHeat-affected zone
H&EHematoxylin and eosin
HSAHuman serum albumin
IBInverse bremsstrahlung
ICP-MSInductively coupled plasma mass spectroscopy
ILIntuition learning
IRInfrared
KNNKernel nearest neighbour
KPCAKernel principal component analysis
KPCA-SVMKernel principal component analysis followed by support vector machine
KWKruskal–Wallis
LA-ICP-MSLaser ablation inductively coupled plasma mass spectroscopy
LA-ICP-TOF–MSLaser ablation mass spectroscopy, inductively coupled time of flight mass spectroscopy
LC-OCTLine-field confocal optical coherence tomography
LDALinear discriminant analysis
LIBSLaser-induced breakdown spectroscopy
LIBS-FTIRLaser-induced breakdown spectroscopy and Fourier transform infrared
LIBS-MLLIBS-integrated machine learning
LIBS-RSLIBS-integrated Raman spectroscopy
LIFLaser-induced fluorescence
LIPLaser-induced plasma
LODLimit of detection
LRLogistic regression
LTELocal thermodynamic equilibrium
MCCMerkel cell carcinoma
MedMedical
MLMachine learning
MMGMammography
MLAMachine learning algorithm
MPMMalignant pleural mesothelioma
MRIMagnetic reasoning imaging
Nd:YAGNeodymium-doped yttrium aluminium garnet laser
NE-LIBSNanoparticles enhanced LIBS
Ns-LIBSNanosecond LIBS
NNNeural network
OCTOptical coherence tomography
OLCFOne-line calibration free
PAPhotoacoustic
PASPhotoacoustic spectroscopy
PCAPrincipal component analysis
PCA-KNNPrincipal component analysis followed by kernel nearest neighbour
PCA-LDAPrincipal component analysis followed by linear discriminant analysis
PDMSPolydimethylsiloxane
PETPositron emission tomography
PIXEParticle-induced X-ray emission technique
PLS-DAPartial least squares discriminant analysis
PSCNNParallel spectral convolutional neural network
Ps-LIBSPicosecond LIBS
QDAQuadratic discriminant analysis
ResNetResidual network
RFRandom forest
RF-1D-ResNetRadio frequency one-dimensional residual network
RLReinforcement learning
ROCReceiver operating characteristics
RSRaman spectroscopy
RS-DLMRaman spectroscopy-deep learning model
RSMRefined spatial module
RSM-LDARefined spatial module-linear discriminant analysis
SASelf-absorption
SA-LIBSSpark assisted LIBS
SBRSignal-to-background ratio
SCSelf-calibrated
SCCSquamous cell carcinoma
SD-LIBSSpark discharge LIBS
SEN-LIBSSurface-enhanced LIBS
SEMScanning electron microscopy
SIMCASoft independent modelling by class analogy
SKBSelectkbest
SMLSupervised machine learning
SNNSpiking neural network
SNRSignal-to-noise ratio
SOMSelf-organising map
SPSample preparation
SP-LIBSSingle pulse LIBS
SVMSupport vector machine
TAType of analysis
ULISAUpconversion-linked immunosorbent assay
Us-LIBSUltrashort LIBS
UV-Fs-LIBSUltraviolet femtosecond LIBS
UV-NIRUltraviolet to near infrared
VUVVacuum ultraviolet
WHOWorld Health Organization
XGBoostExtreme gradient boosting
XRFX-ray fluorescence

References

  1. Nicolodelli, G.; Cabral, J.; Menegatti, C.R.; Marangoni, B.; Senesi, G.S. Recent advances and future trends in LIBS applications to agricultural materials and their food derivatives: An overview of developments in the last decade (2010–2019). Part I. Soils and fertilizers. TrAC Trends Anal. Chem. 2019, 115, 70–82. [Google Scholar] [CrossRef]
  2. Ruan, F.; Zhang, T.; Li, H. Laser-induced breakdown spectroscopy in archeological science: A review of its application and future perspectives. Appl. Spectrosc. Rev. 2019, 54, 573–601. [Google Scholar] [CrossRef]
  3. Cabral, J.S.; Menegatti, C.R.; Nicolodelli, G. Laser-induced breakdown spectroscopy in cementitious materials: A chronological review of cement and concrete from the last 20 years. TrAC Trends Anal. Chem. 2023, 160, 116948. [Google Scholar] [CrossRef]
  4. Pedarnig, J.D.; Trautner, S.; Grünberger, S.; Giannakaris, N.; Eschlböck-Fuchs, S.; Hofstadler, J. Review of element analysis of industrial materials by in-line laser—Induced breakdown spectroscopy (LIBS). Appl. Sci. 2021, 11, 9274. [Google Scholar] [CrossRef]
  5. Ji, H.; Ding, Y.; Zhang, L.; Hu, Y.; Zhong, X. Review of aerosol analysis by laser-induced breakdown spectroscopy. Appl. Spectrosc. Rev. 2021, 56, 193–220. [Google Scholar] [CrossRef]
  6. Palásti, D.J.; Metzinger, A.; Ajtai, T.; Bozóki, Z.; Hopp, B.; Kovács-Széles, É.; Galbács, G. Qualitative discrimination of coal aerosols by using the statistical evaluation of laser-induced breakdown spectroscopy data. Spectrochim. Acta Part B At. Spectrosc. 2019, 153, 34–41. [Google Scholar] [CrossRef]
  7. Wang, L.; Tolok, G.; Fu, Y.; Xu, L.; Li, L.; Gao, H.; Zhou, Y. Application and Research Progress of Laser-Induced Breakdown Spectroscopy in Agricultural Product Inspection. ACS Omega 2024, 9, 24203–24218. [Google Scholar] [CrossRef]
  8. Botto, A.; Campanella, B.; Legnaioli, S.; Lezzerini, M.; Lorenzetti, G.; Pagnotta, S.; Poggialini, F.; Palleschi, V. Applications of laser-induced breakdown spectroscopy in cultural heritage and archaeology: A critical review. J. Anal. At. Spectrom. 2019, 34, 81–103. [Google Scholar] [CrossRef]
  9. Qiao, S.; Ding, Y.; Tian, D.; Yao, L.; Yang, G. A review of laser-induced breakdown spectroscopy for analysis of geological materials. Appl. Spectrosc. Rev. 2015, 50, 1–26. [Google Scholar] [CrossRef]
  10. Rifai, K.; Michaud Paradis, M.-C.; Swierczek, Z.; Doucet, F.; Özcan, L.; Fayad, A.; Li, J.; Vidal, F. Emergences of new technology for ultrafast automated mineral phase identification and quantitative analysis using the CORIOSITY Laser-Induced Breakdown Spectroscopy (LIBS) system. Minerals 2020, 10, 918. [Google Scholar] [CrossRef]
  11. Pedarnig, J. Application of laser-induced breakdown spectroscopy to the analysis of secondary materials in industrial production. In Laser Spectroscopy for Sensing; Elsevier: Amsterdam, The Netherlands, 2014; pp. 496–521. [Google Scholar] [CrossRef]
  12. Rammelkamp, K.; Schröder, S.; Kubitza, S.; Vogt, D.S.; Frohmann, S.; Hansen, P.B.; Böttger, U.; Hanke, F.; Hübers, H.W. Low-level LIBS and Raman data fusion in the context of in situ Mars exploration. J. Raman Spectrosc. 2020, 51, 1682–1701. [Google Scholar] [CrossRef]
  13. Ding, J.; Zhang, T.; Li, H. Recent advances in laser-induced breakdown spectroscopy for explosive analysis. TrAC Trends Anal. Chem. 2023, 166, 117197. [Google Scholar] [CrossRef]
  14. Peng, J.; Song, K.; Zhu, H.; Kong, W.; Liu, F.; Shen, T.; He, Y. Fast detection of tobacco mosaic virus infected tobacco using laser-induced breakdown spectroscopy. Sci. Rep. 2017, 7, 44551. [Google Scholar] [CrossRef]
  15. Velásquez-Ferrín, A.; Babos, D.V.; Marina-Montes, C.; Anzano, J. Rapidly growing trends in laser-induced breakdown spectroscopy for food analysis. Appl. Spectrosc. Rev. 2021, 56, 492–512. [Google Scholar] [CrossRef]
  16. Singh, V.K.; Kumar, V.; Sharma, J. Importance of laser-induced breakdown spectroscopy for hard tissues (bone, teeth) and other calcified tissue materials. Lasers Med. Sci. 2015, 30, 1763–1778. [Google Scholar] [CrossRef] [PubMed]
  17. Rehse, S.J. A review of the use of laser-induced breakdown spectroscopy for bacterial classification, quantification, and identification. Spectrochim. Acta Part B At. Spectrosc. 2019, 154, 50–69. [Google Scholar] [CrossRef]
  18. Houghton, S.C.; Hankinson, S.E. Cancer progress and priorities: Breast cancer. Cancer Epidemiol. Biomark. Prev. 2021, 30, 822–844. [Google Scholar] [CrossRef]
  19. Wolner, Z.J.; Yélamos, O.; Liopyris, K.; Rogers, T.; Marchetti, M.A.; Marghoob, A.A. Enhancing skin cancer diagnosis with dermoscopy. Dermatol. Clin. 2017, 35, 417–437. [Google Scholar] [CrossRef]
  20. Tanoue, L.T.; Tanner, N.T.; Gould, M.K.; Silvestri, G.A. Lung cancer screening. Am. J. Respir. Crit. Care Med. 2015, 191, 19–33. [Google Scholar] [CrossRef]
  21. Meng, X.; Chen, J.; Zhang, Z.; Li, K.; Li, J.; Yu, Z.; Zhang, Y. Non-invasive optical methods for melanoma diagnosis. Photodiagnosis Photodyn. Ther. 2021, 34, 102266. [Google Scholar] [CrossRef]
  22. Mann, R.M.; Kuhl, C.K.; Moy, L. Contrast-enhanced MRI for breast cancer screening. J. Magn. Reson. Imaging 2019, 50, 377–390. [Google Scholar] [CrossRef]
  23. Wan, B.; Ganier, C.; Du-Harpur, X.; Harun, N.; Watt, F.; Patalay, R.; Lynch, M. Applications and future directions for optical coherence tomography in dermatology. Br. J. Dermatol. 2021, 184, 1014–1022. [Google Scholar] [CrossRef] [PubMed]
  24. Ruini, C.; Schuh, S.; Gust, C.; Kendziora, B.; Frommherz, L.; French, L.E.; Hartmann, D.; Welzel, J.; Sattler, E.C. Line-field confocal optical coherence tomography for the in vivo real-time diagnosis of different stages of keratinocyte skin cancer: A preliminary study. J. Eur. Acad. Dermatol. Venereol. 2021, 35, 2388–2397. [Google Scholar] [CrossRef] [PubMed]
  25. Lim, L.; Nichols, B.; Migden, M.R.; Rajaram, N.; Reichenberg, J.S.; Markey, M.K.; Ross, M.I.; Tunnell, J.W. Clinical study of noninvasive in vivo melanoma and nonmelanoma skin cancers using multimodal spectral diagnosis. J. Biomed. Opt. 2014, 19, 117003. [Google Scholar] [CrossRef]
  26. Uddin, A.H.; Khalid, R.S.; Alaama, M.; Abdualkader, A.M.; Kasmuri, A.; Abbas, S. Comparative study of three digestion methods for elemental analysis in traditional medicine products using atomic absorption spectrometry. J. Anal. Sci. Technol. 2016, 7, 6. [Google Scholar] [CrossRef]
  27. Yue, Z.; Sun, C.; Chen, F.; Zhang, Y.; Xu, W.; Shabbir, S.; Zou, L.; Lu, W.; Wang, W.; Xie, Z. Machine learning-based LIBS spectrum analysis of human blood plasma allows ovarian cancer diagnosis. Biomed. Opt. Express 2021, 12, 2559–2574. [Google Scholar] [CrossRef] [PubMed]
  28. Auner, G.W.; Koya, S.K.; Huang, C.; Broadbent, B.; Trexler, M.; Auner, Z.; Elias, A.; Mehne, K.C.; Brusatori, M.A. Applications of Raman spectroscopy in cancer diagnosis. Cancer Metastasis Rev. 2018, 37, 691–717. [Google Scholar] [CrossRef]
  29. Bonta, M.; Gonzalez, J.J.; Quarles, C.D.; Russo, R.E.; Hegedus, B.; Limbeck, A. Elemental mapping of biological samples by the combined use of LIBS and LA-ICP-MS. J. Anal. At. Spectrom. 2016, 31, 252–258. [Google Scholar] [CrossRef]
  30. Li, F.; Ge, L.; Tang, Z.; Chen, Y.; Wang, J. Recent developments on XRF spectra evaluation. Appl. Spectrosc. Rev. 2020, 55, 263–287. [Google Scholar] [CrossRef]
  31. Wang, J.; Li, L.; Yang, P.; Chen, Y.; Zhu, Y.; Tong, M.; Hao, Z.; Li, X. Identification of cervical cancer using laser-induced breakdown spectroscopy coupled with principal component analysis and support vector machine. Lasers Med. Sci. 2018, 33, 1381–1386. [Google Scholar] [CrossRef]
  32. Wang, Q.; Xiangli, W.; Teng, G.; Cui, X.; Wei, K. A brief review of laser-induced breakdown spectroscopy for human and animal soft tissues: Pathological diagnosis and physiological detection. Appl. Spectrosc. Rev. 2021, 56, 221–241. [Google Scholar] [CrossRef]
  33. Asaduzzaman, K.; Khandaker, M.U.; Baharudin, N.A.B.; Amin, Y.B.M.; Farook, M.S.; Bradley, D.; Mahmoud, O. Heavy metals in human teeth dentine: A bio-indicator of metals exposure and environmental pollution. Chemosphere 2017, 176, 221–230. [Google Scholar] [CrossRef]
  34. Galbács, G. A critical review of recent progress in analytical laser-induced breakdown spectroscopy. Anal. Bioanal. Chem. 2015, 407, 7537–7562. [Google Scholar] [CrossRef]
  35. Ghasemi, F.; Parvin, P.; Motlagh, N.S.H.; Amjadi, A.; Abachi, S. Laser induced breakdown spectroscopy and acoustic response techniques to discriminate healthy and cancerous breast tissues. Appl. Opt. 2016, 55, 8227–8235. [Google Scholar] [CrossRef]
  36. Gazmeh, M.; Bahreini, M.; Tavassoli, S.H. Discrimination of healthy and carious teeth using laser-induced breakdown spectroscopy and partial least square discriminant analysis. Appl. Opt. 2015, 54, 123–131. [Google Scholar] [CrossRef]
  37. Cáceres, J.O.; de Los Terreros, J.Y.S. A real-world approach to identifying animal bones and Lower Pleistocene fossils by laser induced breakdown spectroscopy. Talanta 2021, 235, 122780. [Google Scholar] [CrossRef] [PubMed]
  38. Gimenez, Y.; Busser, B.; Trichard, F.; Kulesza, A.; Laurent, J.; Zaun, V.; Lux, F.; Benoit, J.-M.; Panczer, G.; Dugourd, P. 3D imaging of nanoparticle distribution in biological tissue by laser-induced breakdown spectroscopy. Sci. Rep. 2016, 6, 29936. [Google Scholar] [CrossRef]
  39. Eum, C.; Park, J.; Kumar, S.; Jang, E.; Lee, Y.; Nam, S.-H.; Choi, D.; Chung, H. Feasibility of laser-induced breakdown spectroscopy as a direct raw bile analysis tool for screening of gallbladder cancer. J. Anal. At. Spectrom. 2022, 37, 823–832. [Google Scholar] [CrossRef]
  40. Khan, M.N.; Wang, Q.; Idrees, B.S.; Xiangli, W.; Teng, G.; Cui, X.; Zhao, Z.; Wei, K.; Abrar, M. A review on laser-induced breakdown spectroscopy in different cancers diagnosis and classification. Front. Phys. 2022, 10, 10. [Google Scholar] [CrossRef]
  41. Singh, V.K.; Sharma, J.; Pathak, A.K.; Ghany, C.T.; Gondal, M. Laser-induced breakdown spectroscopy (LIBS): A novel technology for identifying microbes causing infectious diseases. Biophys. Rev. 2018, 10, 1221–1239. [Google Scholar] [CrossRef]
  42. V S, D.; George, S.D.; Kartha, V.; Chidangil, S.; V K, U. Hybrid LIBS-Raman-LIF systems for multi-modal spectroscopic applications: A topical review. Appl. Spectrosc. Rev. 2021, 56, 463–491. [Google Scholar] [CrossRef]
  43. Gaudiuso, R.; Melikechi, N.; Abdel-Salam, Z.A.; Harith, M.A.; Palleschi, V.; Motto-Ros, V.; Busser, B. Laser-induced breakdown spectroscopy for human and animal health: A review. Spectrochim. Acta Part B At. Spectrosc. 2019, 152, 123–148. [Google Scholar] [CrossRef]
  44. Khan, Z.; Ullah, M.H.; Rahman, B.; Talukder, A.I.; Wahadoszamen, M.; Abedin, K.; Haider, A. Laser-induced breakdown spectroscopy (LIBS) for trace element detection: A review. J. Spectrosc. 2022, 2022, 3887038. [Google Scholar] [CrossRef]
  45. Markushin, Y.; Sivakumar, P.; Connolly, D.; Melikechi, N. Tag-femtosecond laser-induced breakdown spectroscopy for the sensitive detection of cancer antigen 125 in blood plasma. Anal. Bioanal. Chem. 2015, 407, 1849–1855. [Google Scholar] [CrossRef] [PubMed]
  46. Jaswal, B.B.S.; Kumar, V.; Sharma, J.; Rai, P.K.; Gondal, M.A.; Gondal, B.; Singh, V.K. Analysis of heterogeneous gallstones using laser-induced breakdown spectroscopy (LIBS) and wavelength dispersive X-ray fluorescence (WD-XRF). Lasers Med. Sci. 2016, 31, 573–579. [Google Scholar] [CrossRef] [PubMed]
  47. Ghasemi, F.; Parvin, P.; Reif, J.; Abachi, S.; Mohebbifar, M.; Razzaghi, M. Laser induced breakdown spectroscopy for the diagnosis of several malignant tissue samples. J. Laser Appl. 2017, 29, 042005. [Google Scholar] [CrossRef]
  48. Chen, X.; Li, X.; Yu, X.; Chen, D.; Liu, A. Diagnosis of human malignancies using laser-induced breakdown spectroscopy in combination with chemometric methods. Spectrochim. Acta Part B At. Spectrosc. 2018, 139, 63–69. [Google Scholar] [CrossRef]
  49. Marín-Roldán, A.; Manzoor, S.; Krištof, J.; Veis, P. Enlarged spectral range in Calibration Free-Laser Induced Breakdown Spectroscopy for the qualitative and quantitative analysis of a complex bone matrix. Spectrochim. Acta Part B At. Spectrosc. 2019, 156, 13–19. [Google Scholar] [CrossRef]
  50. Gondal, M.; Aldakheel, R.; Almessiere, M.; Nasr, M.; Almusairii, J.; Gondal, B. Determination of heavy metals in cancerous and healthy colon tissues using laser induced breakdown spectroscopy and its cross-validation with ICP-AES method. J. Pharm. Biomed. Anal. 2020, 183, 113153. [Google Scholar] [CrossRef]
  51. Kiss, K.; Šindelářová, A.; Krbal, L.; Stejskal, V.; Mrázová, K.; Vrábel, J.; Kaška, M.; Modlitbová, P.; Pořízka, P.; Kaiser, J. Imaging margins of skin tumors using laser-induced breakdown spectroscopy and machine learning. J. Anal. At. Spectrom. 2021, 36, 909–916. [Google Scholar] [CrossRef]
  52. Pyun, S.H.; Min, W.; Goo, B.; Seit, S.; Azzi, A.; Wong, D.Y.-S.; Munavalli, G.S.; Huh, C.-H.; Won, C.-H.; Ko, M. Real-time, in vivo skin cancer triage by laser-induced plasma spectroscopy combined with a deep learning–based diagnostic algorithm. J. Am. Acad. Dermatol. 2023, 89, 99–105. [Google Scholar] [CrossRef]
  53. Livingston, K.M.; Zejdlik, K.; Baudelet, M. Reassociation of Skeletal Remains Using Laser-Induced Breakdown Spectroscopy. Anal. Chem. 2024, 96, 9478–9485. [Google Scholar] [CrossRef]
  54. Palleschi, V. Laser-induced breakdown spectroscopy: Principles of the technique and future trends. ChemTexts 2020, 6, 18. [Google Scholar] [CrossRef]
  55. Singh, V.K.; Rai, A.K. Prospects for laser-induced breakdown spectroscopy for biomedical applications: A review. Lasers Med. Sci. 2011, 26, 673–687. [Google Scholar] [CrossRef]
  56. Hahn, D.W.; Omenetto, N. Laser-induced breakdown spectroscopy (LIBS), part I: Review of basic diagnostics and plasma–particle interactions: Still-challenging issues within the analytical plasma community. Appl. Spectrosc. 2010, 64, 335A–366A. [Google Scholar] [CrossRef]
  57. Pasquini, C.; Cortez, J.; Silva, L.; Gonzaga, F.B. Laser induced breakdown spectroscopy. J. Braz. Chem. Soc. 2007, 18, 463–512. [Google Scholar] [CrossRef]
  58. Song, K.; Lee, Y.-I.; Sneddon, J. Recent developments in instrumentation for laser induced breakdown spectroscopy. Appl. Spectrosc. Rev. 2002, 37, 89–117. [Google Scholar] [CrossRef]
  59. Fortes, F.J.; Moros, J.; Lucena, P.; Cabalín, L.M.; Laserna, J.J. Laser-induced breakdown spectroscopy. Anal. Chem. 2013, 85, 640–669. [Google Scholar] [CrossRef] [PubMed]
  60. Hahn, D.W.; Omenetto, N. Laser-induced breakdown spectroscopy (LIBS), part II: Review of instrumental and methodological approaches to material analysis and applications to different fields. Appl. Spectrosc. 2012, 66, 347–419. [Google Scholar] [CrossRef] [PubMed]
  61. Vogel, A.; Venugopalan, V. Mechanisms of pulsed laser ablation of biological tissues. Chem. Rev. 2003, 103, 577–644. [Google Scholar] [CrossRef]
  62. Cremers, D.A.; Radziemski, L.J. Handbook of Laser-Induced Breakdown Spectroscopy; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar] [CrossRef]
  63. Yu, J.; Ma, Q.; Motto-Ros, V.; Lei, W.; Wang, X.; Bai, X. Generation and expansion of laser-induced plasma as a spectroscopic emission source. Front. Phys. 2012, 7, 649–669. [Google Scholar] [CrossRef]
  64. Wang, Z.; Afgan, M.S.; Gu, W.; Song, Y.; Wang, Y.; Hou, Z.; Song, W.; Li, Z. Recent advances in laser-induced breakdown spectroscopy quantification: From fundamental understanding to data processing. TrAC Trends Anal. Chem. 2021, 143, 116385. [Google Scholar] [CrossRef]
  65. Rai, V.; Thakur, S. Physics of plasma in laser-induced breakdown spectroscopy. In Laser-Induced Breakdown Spectroscopy; Singh, J.P., Thakur, S.N., Eds.; Elsevier: Amsterdam, The Netherlands, 2007; pp. 83–111. [Google Scholar] [CrossRef]
  66. Pořízka, P.; Klus, J.; Képeš, E.; Prochazka, D.; Hahn, D.W.; Kaiser, J. On the utilization of principal component analysis in laser-induced breakdown spectroscopy data analysis, a review. Spectrochim. Acta Part B At. Spectrosc. 2018, 148, 65–82. [Google Scholar] [CrossRef]
  67. Zhang, T.; Tang, H.; Li, H. Chemometrics in laser-induced breakdown spectroscopy. J. Chemom. 2018, 32, e2983. [Google Scholar] [CrossRef]
  68. Limbeck, A.; Brunnbauer, L.; Lohninger, H.; Pořízka, P.; Modlitbová, P.; Kaiser, J.; Janovszky, P.; Kéri, A.; Galbács, G. Methodology and applications of elemental mapping by laser induced breakdown spectroscopy. Anal. Chim. Acta 2021, 1147, 72–98. [Google Scholar] [CrossRef] [PubMed]
  69. Jolivet, L.; Leprince, M.; Moncayo, S.; Sorbier, L.; Lienemann, C.-P.; Motto-Ros, V. Review of the recent advances and applications of LIBS-based imaging. Spectrochim. Acta Part B At. Spectrosc. 2019, 151, 41–53. [Google Scholar] [CrossRef]
  70. Messaoud Aberkane, S.; Safi, A.; Botto, A.; Campanella, B.; Legnaioli, S.; Poggialini, F.; Raneri, S.; Rezaei, F.; Palleschi, V. Laser-induced breakdown spectroscopy for determination of spectral fundamental parameters. Appl. Sci. 2020, 10, 4973. [Google Scholar] [CrossRef]
  71. Zhang, N.; Ou, T.; Wang, M.; Lin, Z.; Lv, C.; Qin, Y.; Li, J.; Yang, H.; Zhao, N.; Zhang, Q. A brief review of calibration-free laser-induced breakdown spectroscopy. Front. Phys. 2022, 10, 887171. [Google Scholar] [CrossRef]
  72. Poggialini, F.; Campanella, B.; Cocciaro, B.; Lorenzetti, G.; Palleschi, V.; Legnaioli, S. Catching up on calibration-free LIBS. J. Anal. At. Spectrom. 2023, 38, 1751–1771. [Google Scholar] [CrossRef]
  73. Tognoni, E.; Cristoforetti, G.; Legnaioli, S.; Palleschi, V. Calibration-Free Laser-Induced Breakdown Spectroscopy: State of the art. Spectrochim. Acta Part B At. Spectrosc. 2010, 65, 1–14. [Google Scholar] [CrossRef]
  74. Abubaker Hassan, H. Effects of Different Laser Pulse Regimes (Nanosecond, Picosecond and Femtosecond) on the Ablation of Materials for Production of Nanoparticles in Liquid Solution. In High Energy and Short Pulse Lasers; Richard, V., Ed.; IntechOpen: Rijeka, Croatia, 2016; Volume Chapter 12, pp. 305–325. [Google Scholar] [CrossRef]
  75. Bäuerle, D. Laser Processing and Chemistry; Springer Science & Business Media: London, UK, 2011; p. 874. [Google Scholar] [CrossRef]
  76. Méndez-López, C.; González-Gago, C.; Pisonero, J.; Bordel, N. Spatially resolved spectroscopical characterization of one-and two-component structured femtosecond laser induced plasmas. Spectrochim. Acta Part B At. Spectrosc. 2024, 216, 106949. [Google Scholar] [CrossRef]
  77. Labutin, T.A.; Lednev, V.N.; Ilyin, A.A.; Popov, A.M. Femtosecond laser-induced breakdown spectroscopy. J. Anal. At. Spectrom. 2016, 31, 90–118. [Google Scholar] [CrossRef]
  78. Pisonero, J.; Calon, A.; Linares, J.; Mendez-Vicente, A.; Martinez-Nistal, A.; Bordel, N. Evaluation of femtosecond-LA-ICP-TOFMS for multi-elemental mapping at cellular resolution of human-tissue from cancer patients. Opt. Laser Technol. 2025, 180, 111527. [Google Scholar] [CrossRef]
  79. Al-Bourgol, S.; Machinet, G.; Bakkali, A.; Faucon, M.; Gemini, L. Real-Time Monitoring of Thermal Phenomena during Femtosecond Ablation of Bone Tissue for Process Control. Bioengineering 2024, 11, 309. [Google Scholar] [CrossRef]
  80. Bello, L.T.; da Ana, P.A.; Santos, D., Jr.; Krug, F.J.; Zezell, D.M.; Vieira, N.D., Jr.; Samad, R.E. Mercury amalgam diffusion in human teeth probed using femtosecond LIBS. Appl. Spectrosc. 2017, 71, 659–669. [Google Scholar] [CrossRef]
  81. Sarpe, C.; Ciobotea, E.R.; Morscher, C.B.; Zielinski, B.; Braun, H.; Senftleben, A.; Rüschoff, J.; Baumert, T. Identification of tumor tissue in thin pathological samples via femtosecond laser-induced breakdown spectroscopy and machine learning. Sci. Rep. 2023, 13, 9250. [Google Scholar] [CrossRef] [PubMed]
  82. Choi, J.-H.; Shin, S.; Moon, Y.; Han, J.H.; Hwang, E.; Jeong, S. High spatial resolution imaging of melanoma tissue by femtosecond laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2021, 179, 106090. [Google Scholar] [CrossRef]
  83. Gill, R.K.; Smith, Z.J.; Panchal, R.R.; Bishop, J.W.; Gandour-Edwards, R.; Wachsmann-Hogiu, S. Preliminary fsLIBS study on bone tumors. Biomed. Opt. Express 2015, 6, 4850–4858. [Google Scholar] [CrossRef] [PubMed]
  84. de Menezes, R.F.; Harvey, C.M.; de Martínez Gerbi, M.E.M.; Smith, Z.J.; Smith, D.; Ivaldi, J.C.; Phillips, A.; Chan, J.W.; Wachsmann-Hogiu, S. Fs-laser ablation of teeth is temperature limited and provides information about the ablated components. J. Biophotonics 2017, 10, 1292–1304. [Google Scholar] [CrossRef] [PubMed]
  85. Costas-Rodríguez, M.; Delanghe, J.; Vanhaecke, F. High-precision isotopic analysis of essential mineral elements in biomedicine: Natural isotope ratio variations as potential diagnostic and/or prognostic markers. TrAC Trends Anal. Chem. 2016, 76, 182–193. [Google Scholar] [CrossRef]
  86. Sehrawat, J.S.; Singh, M. Application of Trace Elemental Profile of Known Teeth for Sex and Age Estimation of Ajnala Skeletal Remains: A Forensic Anthropological Cross-Validation Study. Biol. Trace Elem. Res. 2020, 193, 295–310. [Google Scholar] [CrossRef]
  87. Tamburo, E.; Varrica, D.; Dongarrà, G. Gender as a key factor in trace metal and metalloid content of human scalp hair. A multi-site study. Sci. Total Environ. 2016, 573, 996–1002. [Google Scholar] [CrossRef]
  88. Schomburg, L. Dietary selenium and human health. Nutrients 2017, 9, 22. [Google Scholar] [CrossRef]
  89. López-Costas, O.; Lantes-Suarez, O.; Cortizas, A.M. Chemical compositional changes in archaeological human bones due to diagenesis: Type of bone vs soil environment. J. Archaeol. Sci. 2016, 67, 43–51. [Google Scholar] [CrossRef]
  90. Briffa, J.; Sinagra, E.; Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 2020, 6, e04691. [Google Scholar] [CrossRef]
  91. Siddiqui, K.; Bawazeer, N.; Scaria Joy, S. Variation in macro and trace elements in progression of type 2 diabetes. Sci. World J. 2014, 2014, 461591. [Google Scholar] [CrossRef] [PubMed]
  92. Skalny, A.V.; Korobeinikova, T.V.; Aschner, M.; Baranova, O.V.; Barbounis, E.G.; Tsatsakis, A.; Tinkov, A.A. Medical application of Laser-Induced Breakdown Spectroscopy (LIBS) for assessment of trace element and mineral in biosamples: Laboratory and clinical validity of the method. J. Trace Elem. Med. Biol. 2023, 79, 127241. [Google Scholar] [CrossRef] [PubMed]
  93. Clases, D.; Gonzalez de Vega, R. Facets of ICP-MS and their potential in the medical sciences—Part 1: Fundamentals, stand-alone and hyphenated techniques. Anal. Bioanal. Chem. 2022, 414, 7337–7361. [Google Scholar] [CrossRef] [PubMed]
  94. Composition of the Human Body. Available online: https://en.wikipedia.org/wiki/Composition_of_the_human_body (accessed on 6 July 2024).
  95. Takahashi, T.; Thornton, B. Quantitative methods for compensation of matrix effects and self-absorption in Laser Induced Breakdown Spectroscopy signals of solids. Spectrochim. Acta Part B At. Spectrosc. 2017, 138, 31–42. [Google Scholar] [CrossRef]
  96. Rohde, M.; Mehari, F.; Klämpfl, F.; Adler, W.; Neukam, F.W.; Schmidt, M.; Stelzle, F. The differentiation of oral soft-and hard tissues using laser induced breakdown spectroscopy–a prospect for tissue specific laser surgery. J. Biophotonics 2017, 10, 1250–1261. [Google Scholar] [CrossRef]
  97. Shahedi, A.; Eslami, E.; Nourani, M.R. Influence of Lead on the Interpretation of Bone Samples with Laser-Induced Breakdown Spectroscopy. J. Spectrosc. 2016, 2016, 8205479. [Google Scholar] [CrossRef]
  98. Chan, G.C.Y.; Chan, W.-T. Plasma-related matrix effects in inductively coupled plasma—Atomic emission spectrometry by group I and group II matrix-elements. Spectrochim. Acta Part B At. Spectrosc. 2003, 58, 1301–1317. [Google Scholar] [CrossRef]
  99. Wang, W.; Sun, L.; Zhang, P.; Chen, T.; Zheng, L.; Qi, L. Study of matrix effects in laser-induced breakdown spectroscopy by laser defocus and temporal resolution. J. Anal. At. Spectrom. 2021, 36, 1977–1985. [Google Scholar] [CrossRef]
  100. Popov, A.M.; Zaytsev, S.M.; Seliverstova, I.V.; Zakuskin, A.S.; Labutin, T.A. Matrix effects on laser-induced plasma parameters for soils and ores. Spectrochim. Acta Part B At. Spectrosc. 2018, 148, 205–210. [Google Scholar] [CrossRef]
  101. Amais, R.S.; Donati, G.L.; Zezzi Arruda, M.A. ICP-MS and trace element analysis as tools for better understanding medical conditions. TrAC Trends Anal. Chem. 2020, 133, 116094. [Google Scholar] [CrossRef]
  102. Martinez, M.; Bayne, C.; Aiello, D.; Julian, M.; Gaume, R.; Baudelet, M. Multi-elemental matrix-matched calcium hydroxyapatite reference materials for laser ablation: Evaluation on teeth by laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2019, 159, 105650. [Google Scholar] [CrossRef]
  103. Lin, J.; Li, Y.; Ding, K.; Lin, X.; Che, C. Investigating the mechanistic impact of pork soft tissue preparation techniques on the classification precision of laser-induced breakdown spectroscopy. Anal. Methods 2024, 16, 3654–3662. [Google Scholar] [CrossRef] [PubMed]
  104. Małajowicz, J.; Khachatryan, K.; Kozłowska, M. Properties of water activated with low-temperature plasma in the context of microbial activity. Beverages 2022, 8, 63. [Google Scholar] [CrossRef]
  105. Farka, Z.; Vytisková, K.; Makhneva, E.; Zikmundová, E.; Holub, D.; Buday, J.; Prochazka, D.; Novotný, K.; Skládal, P.; Pořízka, P. Comparison of single and double pulse laser-induced breakdown spectroscopy for the detection of biomolecules tagged with photon-upconversion nanoparticles. Anal. Chim. Acta 2024, 1299, 342418. [Google Scholar] [CrossRef]
  106. Tareq, M.S.; Hamad, T.K. Heavy metal migration from dental filling materials to calcified oral tissues: An in vitro analysis using LIBS and ICP-MS techniques. Odontology 2025, 113, 1613–1623. [Google Scholar] [CrossRef]
  107. Khosroshahi, M.E.; Valizadeh, S. Measurement of pulse Nd:YAG laser-induced stress and analysis of dental tissue and amalgam plume using uniaxial polyvinylidene fluoride-based photoacoustic sensor and plasma spectroscopy. Opt. Laser Technol. 2020, 128, 106239. [Google Scholar] [CrossRef]
  108. Kasem, M.; Gonzalez, J.; Russo, R.; Harith, M. LIBS analysis of artificial calcified tissues matrices. Talanta 2013, 108, 53–58. [Google Scholar] [CrossRef] [PubMed]
  109. Janovszky, P.; Kéri, A.; Palásti, D.J.; Brunnbauer, L.; Domoki, F.; Limbeck, A.; Galbács, G. Quantitative elemental mapping of biological tissues by laser-induced breakdown spectroscopy using matrix recognition. Sci. Rep. 2023, 13, 10089. [Google Scholar] [CrossRef] [PubMed]
  110. Martinez, M.; Baudelet, M. Calibration strategies for elemental analysis of biological samples by LA-ICP-MS and LIBS–A review. Anal. Bioanal. Chem. 2020, 412, 27–36. [Google Scholar] [CrossRef]
  111. Youssef, D.; Motto-Ros, V.; Abdel-Harith, M. Machine learning-based novel approach of image analysis in LIBS elemental mapping for discriminating archeological human bone. Spectrochim. Acta Part B At. Spectrosc. 2025, 224, 107108. [Google Scholar] [CrossRef]
  112. Nouir, R.; Cherni, I.; Ghalila, H.; Hamzaoui, S. Early diagnosis of dental pathologies by front face fluorescence (FFF) and laser-induced breakdown spectroscopy (LIBS) with principal component analysis (PCA). Instrum. Sci. Technol. 2022, 50, 465–480. [Google Scholar] [CrossRef]
  113. Mehari, F.; Rohde, M.; Kanawade, R.; Knipfer, C.; Adler, W.; Klämpfl, F.; Stelzle, F.; Schmidt, M. Investigation of the differentiation of ex vivo nerve and fat tissues using laser-induced breakdown spectroscopy (LIBS): Prospects for tissue-specific laser surgery. J. Biophotonics 2016, 9, 1021–1032. [Google Scholar] [CrossRef]
  114. Li, X.; Yang, S.; Fan, R.; Yu, X.; Chen, D. Discrimination of soft tissues using laser-induced breakdown spectroscopy in combination with k nearest neighbors (kNN) and support vector machine (SVM) classifiers. Opt. Laser Technol. 2018, 102, 233–239. [Google Scholar] [CrossRef]
  115. Abbasi, H.; Rauter, G.; Guzman, R.; Cattin, P.C.; Zam, A. Differentiation of femur bone from surrounding soft tissue using laser-induced breakdown spectroscopy as a feedback system for smart laserosteotomy. In Proceedings of the SPIE Photonics Europe 2018, Proceedings Volume 10685. Biophotonics: Photonic Solutions for Better Health Care VI, Strasbourg, France, 23–26 April 2018; pp. 172–179. [Google Scholar] [CrossRef]
  116. Moncayo, S.; Manzoor, S.; Ugidos, T.; Navarro-Villoslada, F.; Caceres, J. Discrimination of human bodies from bones and teeth remains by laser induced breakdown spectroscopy and neural networks. Spectrochim. Acta Part B At. Spectrosc. 2014, 101, 21–25. [Google Scholar] [CrossRef]
  117. Kopřivová, H.; Kiss, K.; Krbal, L.; Stejskal, V.; Buday, J.; Pořízka, P.; Kaška, M.; Ryška, A.; Kaiser, J. Imaging the elemental distribution within human malignant melanomas using Laser-Induced Breakdown Spectroscopy. Anal. Chim. Acta 2024, 1310, 342663. [Google Scholar] [CrossRef]
  118. Khan, M.N.; Wang, Q.; Idrees, B.S.; Teng, G.; Cui, X.; Wei, K. Discrimination of melanoma using laser-induced breakdown spectroscopy conducted on human tissue samples. J. Spectrosc. 2020, 2020, 1–11. [Google Scholar] [CrossRef]
  119. Chen, X.; Li, X.; Yang, S.; Yu, X.; Liu, A. Discrimination of lymphoma using laser-induced breakdown spectroscopy conducted on whole blood samples. Biomed. Opt. Express 2018, 9, 1057–1068. [Google Scholar] [CrossRef]
  120. Emara, E.M.; Song, H.; Imam, H.; Elwekeel, W.M.; Gao, X.; Mohammed, M.M.; Liu, S. Detection of hypokalemia disorder and its relation with hypercalcemia in blood serum using LIBS technique for patients of colorectal cancer grade I and grade II. Lasers Med. Sci. 2022, 37, 1081–1093. [Google Scholar] [CrossRef]
  121. Lin, J.; Li, Y.; Lin, X.; Che, C. Decision-level data fusion based on laser-induced breakdown and Raman spectroscopy: A study of bimodal spectroscopy for diagnosis of lung cancer at different stages. Talanta 2024, 275, 126194. [Google Scholar] [CrossRef] [PubMed]
  122. Ponce, A.; Flores, T.; Ponce, L. Fast Detection of Prostate Malignant Tissue by Multipulsed Laer-Induced Breakdown Spectroscopy (LIBS). Rev. Cuba. De Física 2023, 39, 81. [Google Scholar]
  123. Seifalinezhad, A.; Bahreini, M.; Matin, M.M.H.; Tavassoli, S.H. Feasibility study on discrimination of neo-plastic and non-neoplastic gastric tissues using spark discharge assisted laser induced breakdown spectroscopy. J. Lasers Med. Sci. 2019, 10, 64. [Google Scholar] [CrossRef]
  124. Tareq, M.S.; Hamad, T.K. Quantitative analysis of human teeth by using LIBS technology with different calibration methods: In vitro study. J. Opt. 2024, 54, 2064–2074. [Google Scholar] [CrossRef]
  125. Moon, Y.; Han, J.H.; Choi, J.-h.; Shin, S.; Kim, Y.-C.; Jeong, S. Mapping of cutaneous melanoma by femtosecond laser-induced breakdown spectroscopy. J. Biomed. Opt. 2019, 24, 031011. [Google Scholar] [CrossRef]
  126. Zhang, X.; Li, X.; Chen, X.; Shi, M.; Ren, T. Evaluation of coffee-ring effect of serum and Ag NPs mixture drop stains using laser-induced breakdown spectroscopy mapping. Spectrochim. Acta Part B At. Spectrosc. 2025, 227, 107158. [Google Scholar] [CrossRef]
  127. Zhang, X.; Li, X.; Chen, X.; Shi, M.; Ren, T. Nanoparticle-enhanced laser-induced breakdown spectroscopy for serum element analysis using an Ag NP-coated filter paper substrate. J. Anal. At. Spectrom. 2024, 39, 1332–1342. [Google Scholar] [CrossRef]
  128. Abdel-Salam, Z.; Palleschi, V.; Harith, M. Study of the feeding effect on recent and ancient bovine bones by nanoparticle-enhanced laser-induced breakdown spectroscopy and chemometrics. J. Adv. Res. 2019, 17, 65–72. [Google Scholar] [CrossRef] [PubMed]
  129. Safi, A.; Landis, J.E.; Adler, H.G.; Khadem, H.; Eseller, K.E.; Markushin, Y.; Honarparvaran, S.; De Giacomo, A.; Melikechi, N. Enhancing biomarker detection sensitivity through tag-laser induced breakdown spectroscopy with NELIBS. Talanta 2024, 271, 125723. [Google Scholar] [CrossRef]
  130. Tareq, M.S.; Hamad, T.K. Nanoparticle enhanced laser Induced Breakdown Spectroscopy (NELIBS) analysis of human Teeth by using ZnO and Al2O3 nanoparticles. J. Opt. 2024. [Google Scholar] [CrossRef]
  131. Mohammadimatin, P.; Parvin, P.; Jafargholi, A.; Jahanbakhshi, A.; Ahmadinouri, F.; Tabibkhooei, A.; Heidari, O.; Salarinejad, S. Signal enhancement in spark-assisted laser-induced breakdown spectroscopy for discrimination of glioblastoma and oligodendroglioma lesions. Biomed. Opt. Express 2023, 14, 5795–5816. [Google Scholar] [CrossRef]
  132. Keerthi, K.; George, S.D.; Ongole, R.; Unnikrishnan, V. Fabrication of a Low Cost Superhydrophobic Substrate for Surface Enhanced Laser-Induced Breakdown Spectroscopy and Its Utility through Identification of Electrolyte Variation for Oral Cancer Detection. ACS Biomater. Sci. Eng. 2024, 10, 1153. [Google Scholar] [CrossRef]
  133. Ciobotea, E.R.; Sarpe, C.; Zielinski, B.; Braun, H.; Senftleben, A.; Dutta, S.; Mayer, G.; Florian, C.; Baumert, T. Signal enhancement with double-pulse LIBS on biological samples and better discrimination of tissues through machine learning algorithms. Spectrochim. Acta Part B At. Spectrosc. 2024, 222, 107063. [Google Scholar] [CrossRef]
  134. Grünberger, S.; Ehrentraut, V.; Eschlböck-Fuchs, S.; Hofstadler, J.; Pissenberger, A.; Pedarnig, J.D. Overcoming the matrix effect in the element analysis of steel: Laser ablation-spark discharge-optical emission spectroscopy (LA-SD-OES) and Laser-induced breakdown spectroscopy (LIBS). Anal. Chim. Acta 2023, 1251, 341005. [Google Scholar] [CrossRef]
  135. Grünberger, S.; Eschlböck-Fuchs, S.; Hofstadler, J.; Pissenberger, A.; Duchaczek, H.; Trautner, S.; Pedarnig, J.D. Analysis of minor elements in steel and chemical imaging of micro-patterned polymer by laser ablation-spark discharge-optical emission spectroscopy and laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2020, 169, 105884. [Google Scholar] [CrossRef]
  136. Gschwandtner, P.; Rudinger, F.; Trautner, S.; Ramsauer, C.; Hölzl, G.; Röder, T.; Pedarnig, J.D. Quantification of ZnSO4 in aqueous solution by liquid-solid matrix transfer and double-pulse laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2025, 229, 107215. [Google Scholar] [CrossRef]
  137. Ahamer, C.M.; Pedarnig, J.D. Femtosecond double pulse laser-induced breakdown spectroscopy: Investigation of the intensity enhancement. Spectrochim. Acta Part B At. Spectrosc. 2018, 148, 23–30. [Google Scholar] [CrossRef]
  138. Giannakaris, N.; Haider, A.; Ahamer, C.M.; Grünberger, S.; Trautner, S.; Pedarnig, J.D. Femtosecond single-pulse and orthogonal double-pulse laser-induced breakdown spectroscopy (LIBS): Femtogram mass detection and chemical imaging with micrometer spatial resolution. Appl. Spectrosc. 2022, 76, 926–936. [Google Scholar] [CrossRef]
  139. Giannakaris, N.; Siozos, P.; Piñon, V.; Banerjee, S.P.; Sentis, M.; Pedarnig, J.D.; Anglos, D. UV femtosecond single-pulse and collinear double-pulse laser-induced breakdown spectroscopy (LIBS) for depth-resolved characterization of nano-scaled films. Appl. Surf. Sci. 2023, 640, 158354. [Google Scholar] [CrossRef]
  140. Viljanen, J.; Sun, Z.; Alwahabi, Z.T. Microwave assisted laser-induced breakdown spectroscopy at ambient conditions. Spectrochim. Acta Part B At. Spectrosc. 2016, 118, 29–36. [Google Scholar] [CrossRef]
  141. Chen, T.; Zhang, T.; Li, H. Applications of laser-induced breakdown spectroscopy (LIBS) combined with machine learning in geochemical and environmental resources exploration. TrAC Trends Anal. Chem. 2020, 133, 116113. [Google Scholar] [CrossRef]
  142. Debus, B.; Parastar, H.; Harrington, P.; Kirsanov, D. Deep learning in analytical chemistry. TrAC Trends Anal. Chem. 2021, 145, 116459. [Google Scholar] [CrossRef]
  143. Kościelniak, P. Unified principles of univariate analytical calibration. TrAC Trends Anal. Chem. 2022, 149, 116547. [Google Scholar] [CrossRef]
  144. Mahmud, M.; Kaiser, M.S.; McGinnity, T.M.; Hussain, A. Deep learning in mining biological data. Cogn. Comput. 2021, 13, 1–33. [Google Scholar] [CrossRef]
  145. Han, J.H.; Moon, Y.; Lee, J.J.; Choi, S.; Kim, Y.-C.; Jeong, S. Differentiation of cutaneous melanoma from surrounding skin using laser-induced breakdown spectroscopy. Biomed. Opt. Express 2016, 7, 57–66. [Google Scholar] [CrossRef]
  146. Abbasi, H.; Guzman, R.; Cattin, P.C.; Zam, A. All-fiber-optic LIBS system for tissue differentiation: A prospect for endoscopic smart laser osteotomy. Opt. Lasers Eng. 2022, 148, 106765. [Google Scholar] [CrossRef]
  147. Tarai, A.K.; Junjuri, R.; Dhobley, A.; Gundawar, M.K. Classification of human tooth using laser-induced breakdown spectroscopy combined with machine learning. J. Opt. 2024, 53, 3810–3820. [Google Scholar] [CrossRef]
  148. Gazali, Z.; Kumar, R.; Rai, P.K.; Rai, P.K.; Rai, A.K.; Thakur, S.N. Discrimination of gallbladder stone employing laser-induced breakdown spectroscopy (LIBS) and photoacoustic spectroscopy (PAS). Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 260, 119948. [Google Scholar] [CrossRef]
  149. Pathak, A.; Singh, A.; Kumar, R.; Rai, A. Laser-induced breakdown spectroscopy coupled with PCA study of human tooth. Natl. Acad. Sci. Lett. 2019, 42, 87–90. [Google Scholar] [CrossRef]
  150. Zhao, Z.; Xu, X.; Bao, M.; Zheng, Y.; Luo, T.; Lu, B.; Teng, G.; Wang, Q.; Khan, M.N.; Yong, J. Early screening and staging of melanoma using blood based on laser-induced breakdown spectroscopy. Microchem. J. 2024, 203, 110955. [Google Scholar] [CrossRef]
  151. Teng, G.; Wang, Q.; Yang, H.; Qi, X.; Zhang, H.; Cui, X.; Idrees, B.S.; Xiangli, W.; Wei, K.; Khan, M.N. Pathological identification of brain tumors based on the characteristics of molecular fragments generated by laser ablation combined with a spiking neural network. Biomed. Opt. Express 2020, 11, 4276–4289. [Google Scholar] [CrossRef] [PubMed]
  152. Moncayo, S.; Manzoor, S.; Navarro-Villoslada, F.; Caceres, J. Evaluation of supervised chemometric methods for sample classification by Laser Induced Breakdown Spectroscopy. Chemom. Intell. Lab. Syst. 2015, 146, 354–364. [Google Scholar] [CrossRef]
  153. Winnand, P.; Ooms, M.; Heitzer, M.; Vohl, N.; Lammert, M.; Hölzle, F.; Boernsen, K.O.; Modabber, A. Assessment of the bony resection margin distance in bone-invasive oral cancer using laser-induced breakdown spectroscopy. Clin. Oral Investig. 2024, 28, 474. [Google Scholar] [CrossRef]
  154. Luo, M.; Li, X.; He, Q.; Liu, J.; Lu, H.; Liu, Q.; Yang, X. Elemental analysis and identification of papillary thyroid cancer tissues using laser-induced breakdown spectroscopy. J. Anal. At. Spectrom. 2022, 37, 833–840. [Google Scholar] [CrossRef]
  155. Imam, H.; Mohamed, R.; Eldakrouri, A.A. Primary study of the use of laser-induced plasma spectroscopy for the diagnosis of breast cancer. Opt. Photonics J. 2012, 2, 3. [Google Scholar] [CrossRef]
  156. Idrees, B.S.; Wang, Q.; Khan, M.N.; Teng, G.; Cui, X.; Xiangli, W.; Wei, K. In-vitro study on the identification of gastrointestinal stromal tumor tissues using laser-induced breakdown spectroscopy with chemometric methods. Biomed. Opt. Express 2022, 13, 26–38. [Google Scholar] [CrossRef]
  157. Lin, J.; Li, Y.; Lin, X.; Che, C. Precise and rapid diagnosis of lung cancer: Leveraging laser-induced breakdown spectroscopy with optimized kernel methods in machine learning. J. Anal. At. Spectrom. 2024, 39, 2049–2057. [Google Scholar] [CrossRef]
  158. Li, J.; Pan, X.; Guo, L.; Chen, Y. Cancer diagnosis based on laser-induced breakdown spectroscopy with bagging-voting fusion model. Med. Eng. Phys. 2024, 132, 104207. [Google Scholar] [CrossRef] [PubMed]
  159. Lin, X.; Sun, H.; Gao, X.; Xu, Y.; Wang, Z.; Wang, Y. Discrimination of lung tumor and boundary tissues based on laser-induced breakdown spectroscopy and machine learning. Spectrochim. Acta Part B At. Spectrosc. 2021, 180, 106200. [Google Scholar] [CrossRef]
  160. Li, H.; Sun, H.; Gao, X. Laser-Induced Breakdown Spectroscopy Combined with Machine Learning for the Identification of Lung Cancer Tumors. J. Appl. Spectrosc. 2025, 92, 166–174. [Google Scholar] [CrossRef]
  161. Wang, Y.; Huang, D.; Shu, K.; Xu, Y.; Duan, Y.; Fan, Q.; Lin, Q.; Tuchin, V.V. Optimization of machine learning classification models for tumor cells based on cell elements heterogeneity with laser-induced breakdown spectroscopy. J. Biophotonics 2023, 16, e202300239. [Google Scholar] [CrossRef] [PubMed]
  162. Shi, S.; Pi, L.; Peng, L.; Zhang, D.; Ma, H.; Liu, Y.; Deng, N.; Wang, X.; Guo, L. High-precision identification of breast cancer based on end-to-end parallel spectral convolutional neural network assisted laser-induced breakdown spectroscopy. J. Anal. At. Spectrom. 2025, 40, 478–486. [Google Scholar] [CrossRef]
  163. Shi, S.; Pi, L.; Hou, Z.; He, L.; Wang, X.; Guo, L. Batch Effect Correction for LIBS-FTIR Spectral Fusion in Breast Cancer Serum Detection Based on Gradient Reversal Adversarial Network. Talanta 2025, 295, 128324. [Google Scholar] [CrossRef] [PubMed]
  164. Lin, J.; Li, Y.; Lin, X.; Che, C. Fusion of laser-induced breakdown spectroscopy technology and deep learning: A new method to identify malignant and benign lung tumors with high accuracy. Anal. Bioanal. Chem. 2024, 416, 993–1000. [Google Scholar] [CrossRef]
  165. Teng, G.; Wang, Q.; Zhang, H.; Xiangli, W.; Yang, H.; Qi, X.; Cui, X.; Idrees, B.S.; Wei, K.; Khan, M.N. Discrimination of infiltrative glioma boundary based on laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2020, 165, 105787. [Google Scholar] [CrossRef]
  166. Chu, Y.; Chen, F.; Sheng, Z.; Zhang, D.; Zhang, S.; Wang, W.; Jin, H.; Qi, J.; Guo, L. Blood cancer diagnosis using ensemble learning based on a random subspace method in laser-induced breakdown spectroscopy. Biomed. Opt. Express 2020, 11, 4191–4202. [Google Scholar] [CrossRef]
  167. Kumar, A.; Yueh, F.-Y.; Singh, J.P.; Burgess, S. Characterization of malignant tissue cells by laser-induced breakdown spectroscopy. Appl. Opt. 2004, 43, 5399–5403. [Google Scholar] [CrossRef]
  168. Gaudiuso, R.; Ewusi-Annan, E.; Melikechi, N.; Sun, X.; Liu, B.; Campesato, L.F.; Merghoub, T. Using LIBS to diagnose melanoma in biomedical fluids deposited on solid substrates: Limits of direct spectral analysis and capability of machine learning. Spectrochim. Acta Part B At. Spectrosc. 2018, 146, 106–114. [Google Scholar] [CrossRef]
  169. Wei, H.; Zhao, Z.; Lin, Q.; Duan, Y. Study on the molecular mechanisms against human breast cancer from insight of elemental distribution in tissue based on laser-induced breakdown spectroscopy (LIBS). Biol. Trace Elem. Res. 2021, 199, 1686–1692. [Google Scholar] [CrossRef]
  170. Sancey, L.; Motto-Ros, V.; Busser, B.; Kotb, S.; Benoit, J.-M.; Piednoir, A.; Lux, F.; Tillement, O.; Panczer, G.; Yu, J. Laser spectrometry for multi-elemental imaging of biological tissues. Sci. Rep. 2014, 4, 6065. [Google Scholar] [CrossRef]
  171. Kiss, K.; Kopřivová, H.; Stejskal, V.; Krbal, L.; Buday, J.; Brunnbauer, L.; Képeš, E.; Pořízka, P.; Ryška, A.; Kaška, M. Assessing spatial distribution of bioindicator elements in various cutaneous tumors using correlative imaging with laser-ablation-based analytical methods. Talanta 2024, 279, 126651. [Google Scholar] [CrossRef]
  172. Melikechi, N.; Markushin, Y.; Connolly, D.C.; Lasue, J.; Ewusi-Annan, E.; Makrogiannis, S. Age-specific discrimination of blood plasma samples of healthy and ovarian cancer prone mice using laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2016, 123, 33–41. [Google Scholar] [CrossRef]
  173. Zikmundová, E.; Sklenárová, D.; Kočí, E.; Zatloukalová, T.; Bačová, T.; Makhneva, E.; Holub, D.; Macháčová, E.; Kopřivová, H.; Vytisková, K. Magnetic microbead-based upconversion immunoassay with laser-induced breakdown spectroscopy readout for the detection of prostate-specific antigen. Microchim. Acta 2024, 191, 656. [Google Scholar] [CrossRef]
  174. Khan, M.N.; Wang, Q.; Idrees, B.S.; Teng, G.; Xiangli, W.; Cui, X.; Wei, K. Evaluation of human melanoma and normal formalin paraffin-fixed samples using Raman and LIBS fused data. Lasers Med. Sci. 2022, 37, 2489–2499. [Google Scholar] [CrossRef] [PubMed]
  175. Idrees, B.S.; Teng, G.; Israr, A.; Zaib, H.; Jamil, Y.; Bilal, M.; Bashir, S.; Khan, M.N.; Wang, Q. Comparison of whole blood and serum samples of breast cancer based on laser-induced breakdown spectroscopy with machine learning. Biomed. Opt. Express 2023, 14, 2492–2509. [Google Scholar] [CrossRef] [PubMed]
  176. Orthaber, K.; Pristovnik, M.; Skok, K.; Perić, B.; Maver, U. Skin cancer and its treatment: Novel treatment approaches with emphasis on nanotechnology. J. Nanomater. 2017, 2017, 1–20. [Google Scholar] [CrossRef]
  177. Apalla, Z.; Lallas, A.; Sotiriou, E.; Lazaridou, E.; Ioannides, D. Epidemiological trends in skin cancer. Dermatol. Pract. Concept. 2017, 7, 2. [Google Scholar] [CrossRef] [PubMed]
  178. Mehrabi, J.N.; Baugh, E.G.; Fast, A.; Lentsch, G.; Balu, M.; Lee, B.A.; Kelly, K.M. A clinical perspective on the automated analysis of reflectance confocal microscopy in dermatology. Lasers Surg. Med. 2021, 53, 1011–1019. [Google Scholar] [CrossRef]
  179. Shahriari, N.; Grant-Kels, J.M.; Rabinovitz, H.; Oliviero, M.; Scope, A. Reflectance confocal microscopy: Diagnostic criteria of common benign and malignant neoplasms, dermoscopic and histopathologic correlates of key confocal criteria, and diagnostic algorithms. J. Am. Acad. Dermatol. 2021, 84, 17–31. [Google Scholar] [CrossRef]
  180. Hofmann, M.A.; Keim, U.; Jagoda, A.; Forschner, A.; Fink, C.; Spaenkuch, I.; Tampouri, I.; Eigentler, T.; Weide, B.; Haenssle, H.A. Dermatofluoroscopy diagnostics in different pigmented skin lesions: Strengths and weaknesses. JDDG J. Der Dtsch. Dermatol. Ges. 2020, 18, 682–690. [Google Scholar] [CrossRef]
  181. Barbierato, E.; Gatti, A. The challenges of machine learning: A critical review. Electronics 2024, 13, 416. [Google Scholar] [CrossRef]
  182. Zhao, J.; Lui, H.; Kalia, S.; Lee, T.K.; Zeng, H. Improving skin cancer detection by Raman spectroscopy using convolutional neural networks and data augmentation. Front. Oncol. 2024, 14, 1320220. [Google Scholar] [CrossRef]
  183. Ayadh, M.; Waszczuk, L.; Ogien, J.; Dauce, G.; Augis, L.; Tfaili, S.; Tfayli, A.; Perrot, J.-L.; Dubois, A. AI-assisted identification of nonmelanoma skin cancer structures based on combined line-field confocal optical coherence tomography and confocal Raman microspectroscopy. J. Biomed. Opt. 2025, 30, 076008. [Google Scholar] [CrossRef] [PubMed]
  184. 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] [PubMed]
  185. Weihmann, K.; Schleusener, J.; Eigentler, T.K.; Ghoreschi, F.C.; Moritz, R.K.; Szyc, L.; Dronnik, E.; Meinke, M.C. In vivo study to evaluate an intelligent algorithm for time efficient detection of malignant melanoma using dermatofluoroscopy. Ski. Pharmacol. Physiol. 2024, 37, 97–103. [Google Scholar] [CrossRef] [PubMed]
  186. Peñaranda, F.; Naranjo, V.; Lloyd, G.R.; Kastl, L.; Kemper, B.; Schnekenburger, J.; Nallala, J.; Stone, N. Discrimination of skin cancer cells using Fourier transform infrared spectroscopy. Comput. Biol. Med. 2018, 100, 50–61. [Google Scholar] [CrossRef]
  187. Huttunen, M.J.; Hristu, R.; Dumitru, A.; Floroiu, I.; Costache, M.; Stanciu, S.G. Multiphoton microscopy of the dermoepidermal junction and automated identification of dysplastic tissues with deep learning. Biomed. Opt. Express 2019, 11, 186–199. [Google Scholar] [CrossRef]
  188. Ruini, C.; Schlingmann, S.; Jonke, Ž.; Avci, P.; Padrón-Laso, V.; Neumeier, F.; Koveshazi, I.; Ikeliani, I.U.; Patzer, K.; Kunrad, E. Machine learning based prediction of squamous cell carcinoma in ex vivo confocal laser scanning microscopy. Cancers 2021, 13, 5522. [Google Scholar] [CrossRef] [PubMed]
  189. Cancer Today. Available online: https://gco.iarc.fr/today/online-analysis-table?v=2020&mode=cancer&mode_population=continents&population=900&populations=900 (accessed on 7 March 2024).
  190. Akram, M.; Iqbal, M.; Daniyal, M.; Khan, A.U. Awareness and current knowledge of breast cancer. Biol. Res. 2017, 50, 33. [Google Scholar] [CrossRef] [PubMed]
  191. Lauby-Secretan, B.; Scoccianti, C.; Loomis, D.; Benbrahim-Tallaa, L.; Bouvard, V.; Bianchini, F.; Straif, K. Breast-cancer screening—Viewpoint of the IARC Working Group. N. Engl. J. Med. 2015, 372, 2353–2358. [Google Scholar] [CrossRef]
  192. Sun, Y.-S.; Zhao, Z.; Yang, Z.-N.; Xu, F.; Lu, H.-J.; Zhu, Z.-Y.; Shi, W.; Jiang, J.; Yao, P.-P.; Zhu, H.-P. Risk factors and preventions of breast cancer. Int. J. Biol. Sci. 2017, 13, 1387. [Google Scholar] [CrossRef]
  193. Escudero-Cernuda, S.; Clases, D.; Eiro, N.; González, L.O.; Fraile, M.; Vizoso, F.J.; Fernández-Sánchez, M.L.; Gonzalez de Vega, R. Quantitative distribution of essential elements and non-essential metals in breast cancer tissues by LA-ICP-TOF–MS. Anal. Bioanal. Chem. 2025, 417, 361–371. [Google Scholar] [CrossRef]
  194. Naidu, B.G.; Srikanth, S.; Raju, G.N.; Sarita, P. PIXE analysis of blood serum of breast cancer patients undergoing successive chemotherapy. J. Radioanal. Nucl. Chem. 2020, 323, 1307–1316. [Google Scholar] [CrossRef]
  195. Huthwelker, T.; Borca, C.N.; Altamura, D.; De Caro, L.; Vanna, R.; Corsi, F.; Morasso, C.; Banfi, G.; Arpa, G.; Bunk, O. Microcalcifications in breast cancer tissue studied by X-ray absorption, emission, scattering and diffraction. Appl. Crystallogr. 2025, 58, 233–250. [Google Scholar] [CrossRef]
  196. Eyecioğlu, Ö. Determination of Se, Cr, Mn, Zn, Co, Na, and K in blood samples of breast cancer patients to investigate their variation using ICP-MS and ICP-OES. At. Spectrosc. 2019, 40, 11–16. [Google Scholar] [CrossRef]
  197. Balakrishnan, M.; George, R.; Sharma, A.; Graham, D.Y. Changing trends in stomach cancer throughout the world. Curr. Gastroenterol. Rep. 2017, 19, 36. [Google Scholar] [CrossRef]
  198. Visser, M.; Van Grimbergen, I.; Hölters, J.; Barendregt, W.; Vermeer, L.; Vreuls, W.; Janssen, J. Performance insights of endobronchial ultrasonography (EBUS) and mediastinoscopy for mediastinal lymph node staging in lung cancer. Lung Cancer 2021, 156, 122–128. [Google Scholar] [CrossRef] [PubMed]
  199. Winnand, P.; Boernsen, K.O.; Ooms, M.; Heitzer, M.; Vohl, N.; Lammert, M.; Hölzle, F.; Modabber, A. Assessment of the electrolyte heterogeneity of tissues in mandibular bone-infiltrating head and neck cancer using laser-induced breakdown spectroscopy. Int. J. Mol. Sci. 2024, 25, 2607. [Google Scholar] [CrossRef]
  200. Samek, O.; Beddows, D.; Telle, H.; Kaiser, J.; Liška, M.; Caceres, J.; Urena, A.G. Quantitative laser-induced breakdown spectroscopy analysis of calcified tissue samples. Spectrochim. Acta Part B At. Spectrosc. 2001, 56, 865–875. [Google Scholar] [CrossRef]
  201. Abdullah, K.F.; Kadhem, S.J. Analysis of laser induced breakdown spectra for distinguish between healthy and carious teeth. J. Opt. 2024, 54, 1100–1107. [Google Scholar] [CrossRef]
  202. Ayed, M.S.; Shafiq, S.S.; Diab, H.M.; Alahmari, A.D.; Divakar, D.D. Assessing periapical dental radiographs as a screening parameter for early indications of osteoporosis in postmenopausal periodontal patients and root surface evaluation using spectrochemical analysis. Saudi Med. J. 2018, 39, 719. [Google Scholar] [CrossRef] [PubMed]
  203. Alhasmi, A.M.; Gondal, M.A.; Nasr, M.M.; Shafik, S.; Habibullah, Y.B. Detection of toxic elements using laser-induced breakdown spectroscopy in smokers’ and nonsmokers’ teeth and investigation of periodontal parameters. Appl. Opt. 2015, 54, 7342–7349. [Google Scholar] [CrossRef]
  204. Batool, J.; Amin, N.; Jamil, Y.; Shaikh, N.; Al Islam, S. Rapid elemental analysis of human teeth using laser induced breakdown spectroscopy. Phys. B Condens. Matter 2021, 602, 412495. [Google Scholar] [CrossRef]
  205. Dakhil, S.S.; Hamad, T.K. Spectrometric measurement of human teeth plasma produced by nanosecond Nd:YAG pulsed laser. J. Opt. 2024, 53, 4107–4112. [Google Scholar] [CrossRef]
  206. Manno, S.H.; Manno, F.A.; Tian, L.; Khan, M.S.; Ahmed, I.; Liu, Y.; Li, V.W.; Xu, S.; Xie, F.; Hung, T.F. Spectroscopic and microscopic examination of teeth exposed to green tea at different temperatures. PLoS ONE 2020, 15, e0244542. [Google Scholar] [CrossRef] [PubMed]
  207. Manno, S.H.; Manno, F.A.; Ahmed, I.; Ahmed, R.; Shu, L.; Li, L.; Xu, S.; Xie, F.; Li, V.W.; Ho, J. Spectroscopic examination of enamel staining by coffee indicates dentin erosion by sequestration of elements. Talanta 2018, 189, 550–559. [Google Scholar] [CrossRef]
  208. Ahmed, A.N.; Alwazzan, M.J.; Ismael, M.A. Study effects of pulse laser energy on human primary teeth and extraction caries area by using image processing techniques. NeuroQuantology 2020, 18, 36. [Google Scholar] [CrossRef]
  209. Hamzaoui, S.; Nouir, R.; Jaidene, N. The study of carious teeth by laser-induced breakdown spectroscopy. J. Appl. Spectrosc. 2017, 84, 82–86. [Google Scholar] [CrossRef]
  210. Sasazawa, S.; Kakino, S.; Matsuura, Y. Optical-fiber-based laser-induced breakdown spectroscopy for detection of early caries. J. Biomed. Opt. 2015, 20, 065002. [Google Scholar] [CrossRef]
  211. Khalid, A.; Bashir, S.; Akram, M.; Hayat, A. Laser-induced breakdown spectroscopy analysis of human deciduous teeth samples. Lasers Med. Sci. 2015, 30, 2233–2238. [Google Scholar] [CrossRef]
  212. Dosedělová, H.; Štěpánková, K.; Zikmund, T.; Lesot, H.; Kaiser, J.; Novotný, K.; Štembírek, J.; Knotek, Z.; Zahradníček, O.; Buchtová, M. Age-related changes in the tooth–bone interface area of acrodont dentition in the chameleon. J. Anat. 2016, 229, 356–368. [Google Scholar] [CrossRef] [PubMed]
  213. Kandil, H.; Ahmed, E.; Fouad, N.; Ali Dabbous, O.; Niazy, M.; Mohamed, T. Using Femtosecond Laser Light-Activated Materials: The Biomimetic Dentin Remineralization Was Monitored by Laser-Induced Breakdown Spectroscopy. Medicina 2023, 59, 591. [Google Scholar] [CrossRef]
  214. Tofanelli, M.; Pardini, L.; Borrini, M.; Bartoli, F.; Bacci, A.; D’Ulivo, A.; Pitzalis, E.; Mascherpa, M.C.; Legnaioli, S.; Lorenzetti, G. Spectroscopic analysis of bones for forensic studies. Spectrochim. Acta Part B At. Spectrosc. 2014, 99, 70–75. [Google Scholar] [CrossRef]
  215. Marković, M.; Kuzmanović, M.; Ranković, D.; Bajuk-Bogdanović, D.; Šajić, A.; Dimić, D. From Structure to Strength: Analyzing the Impact of Sulfuric Acid on Pig Bone Demineralization Through FTIR, LIBS, and AAS. Int. J. Mol. Sci. 2024, 25, 12250. [Google Scholar] [CrossRef]
  216. Al-Hadeethi, Y.; Al-Jedani, S.; Razvi, M.; Saeed, A.; Abdel-Daiem, A.; Ansari, M.S.; Babkair, S.S.; Salah, N.A.; Al-Mujtaba, A. Data fitting to study ablated hard dental tissues by nanosecond laser irradiation. PLoS ONE 2016, 11, e0156093. [Google Scholar] [CrossRef]
  217. Mustafa, M.; Latif, A.; Jehangir, M.; Siraj, K. Nd:YAG laser irradiation consequences on calcium and magnesium in human dental tissues. Lasers Dent. Sci. 2022, 6, 107–115. [Google Scholar] [CrossRef]
  218. Mustafa, M.; Jehangir, M.; Latif, A. Laser-Induced Breakdown Spectroscopy and Microscopy Study of Human Dental Tissues. In Electron Microscopy; Mhadhbi, M., Ed.; IntechOpen: Rijeka, Croatia, 2022. [Google Scholar] [CrossRef]
  219. PetroviĆ, J.; MarinkoviĆ, J.; ŽivkoviĆ, S.; RadenkoviĆ, M.; CiganoviĆ, J.; MarkoviĆ, D.; MomČiloviĆ, M. The TEA CO2 Laser as a Promising tool for Elemental Analysis of Restored Teeth and Evaluation of Restoration Constituent Penetration into Dental Tissues. Plasma Chem. Plasma Process. 2023, 43, 1237–1253. [Google Scholar] [CrossRef]
  220. Khalil, A.A.I.; Gondal, M.A.; Shemis, M.; Khan, I.S. Detection of carcinogenic metals in kidney stones using ultraviolet laser-induced breakdown spectroscopy. Appl. Opt. 2015, 54, 2123–2131. [Google Scholar] [CrossRef] [PubMed]
  221. Li, S.; Teng, F.; Ren, L.; Yan, Q.; Ye, W.; Tian, Y.; Li, Y.; Guo, J.; Deguchi, Y.; Zheng, R. Long-pulsed laser-induced breakdown spectroscopy for elemental imaging: An evaluation with human teeth. Spectrochim. Acta Part B At. Spectrosc. 2025, 229, 107201. [Google Scholar] [CrossRef]
  222. Gill, R.K.; Smith, Z.J.; Lee, C.; Wachsmann-Hogiu, S. The effects of laser repetition rate on femtosecond laser ablation of dry bone: A thermal and LIBS study. J. Biophotonics 2016, 9, 171–180. [Google Scholar] [CrossRef]
  223. Goldman, L.; Hornby, P.; Meyer, R.; Goldman, B. Impact of the laser on dental caries. Nature 1964, 203, 417. [Google Scholar] [CrossRef]
  224. Kareem, R.O.; Bulut, N.; Kaygili, O. Hydroxyapatite biomaterials: A comprehensive review of their properties, structures, medical applications, and fabrication methods. J. Chem. Rev. 2024, 6, 1–26. [Google Scholar] [CrossRef]
  225. Sarna-Boś, K.; Boguta, P.; Skic, K.; Wiącek, D.; Maksymiuk, P.; Sobieszczański, J.; Chałas, R. Physicochemical properties and surface characteristics of ground human teeth. Molecules 2022, 27, 5852. [Google Scholar] [CrossRef] [PubMed]
  226. Fernández-Escudero, A.C.; Legaz, I.; Prieto-Bonete, G.; López-Nicolás, M.; Maurandi-López, A.; Pérez-Cárceles, M.D. Aging and trace elements in human coronal tooth dentine. Sci. Rep. 2020, 10, 9964. [Google Scholar] [CrossRef]
  227. Saghiri, M.A.; Vakhnovetsky, J.; Vakhnovetsky, A.; Ghobrial, M.; Nath, D.; Morgano, S.M. Functional role of inorganic trace elements in dentin apatite tissue—Part 1: Mg, Sr, Zn, and Fe. J. Trace Elem. Med. Biol. 2022, 71, 126932. [Google Scholar] [CrossRef]
  228. Kunin, A.A.; Evdokimova, A.Y.; Moiseeva, N.S. Age-related differences of tooth enamel morphochemistry in health and dental caries. EPMA J. 2015, 6, 3. [Google Scholar] [CrossRef]
  229. Pořízka, P.; Konečná, A.; Šindelářová, A.; Šulcová, M.; Modlitbová, P.; Prochazka, D.; Nevoránková, P.; Navrátil, M.; Vrlíková, L.; Buchtová, M. Feasibility of laser-induced breakdown spectroscopy to elucidate elemental changes in human tooth ankylosis. Spectrochim. Acta Part B At. Spectrosc. 2023, 206, 106727. [Google Scholar] [CrossRef]
  230. Unnikrishnan, V.; Mitra, P.; Srikant, N.; Baptist, J.; Keerthi, K.; Ongole, R. Validity of laser induced breakdown spectroscopy (LIBS) in determining age and sex from tooth specimens. Heliyon 2022, 8, e10946. [Google Scholar] [CrossRef]
  231. Suyanto, H.; Trisnawati, N.; Putra, I.; Suprihatin, I. Characterization of human teeth by laser-induced breakdown spectroscopy. J. Phys. Conf. Ser. 2018, 1120, 012018. [Google Scholar] [CrossRef]
  232. Gassama, B.C.; Gueye, M.; Traore, A.; Tamba, B.; Babamou, K.; Tine, S.D.; Wague, A.; Rey, G.; Faye, B. Laser-induced breakdown spectroscopy of pathological mineralized dental tissues in a population of Senegalese individuals: A study of 75 cases. J. Dent. Oro. Surg. 2016, 1, 115. [Google Scholar] [CrossRef]
  233. Tariq, U.; Haider, Z.; Hussain, R.; Tufail, K.; Ali, J. LIBS analysis of hydroxyapatite extracted from bovine bone for Ca/P ratio measurements. AIP Conf. Proc. 2017, 1824, 030027. [Google Scholar] [CrossRef]
  234. Subki, K.; Putri, A.; Qusthalani, A.; Sari, N.; Subianto, M.; Sadami, K.; Ramli, M.; Lahna, K.; Idris, N. Analysis of organic egg authenticity determination using laser-induced breakdown spectroscopy (LIBS) assisted by artificial neural networks (ANN). IOP Conf. Ser. Earth Environ. Sci. 2025, 1510, 012088. [Google Scholar] [CrossRef]
  235. Luo, W.; Zhao, X.; Lv, S.; Zhu, H. Measurements of egg shell plasma parameters using laser-induced breakdown spectroscopy. Pramana 2015, 85, 105–114. [Google Scholar] [CrossRef]
  236. Fahad, M.; Shah, K.H.; Ali, S.; Abrar, M. Spectroscopic diagnostics of eggshells plasma using LIBS with self-absorption correction in conjunction with WDXRF, EDS and FTIR techniques. Eur. Phys. J. D 2024, 78, 153. [Google Scholar] [CrossRef]
  237. Manard, B.T.; Hintz, C.J.; Quarles, C.D., Jr.; Burns, W.; Zirakparvar, N.A.; Dunlap, D.R.; Beiswenger, T.; Cruz-Uribe, A.M.; Petrus, J.A.; Hexel, C.R. Determination of fluorine distribution in shark teeth by laser-induced breakdown spectroscopy. Metallomics 2022, 14, mfac050. [Google Scholar] [CrossRef]
  238. Abbasi, H.; Rauter, G.; Guzman, R.; Cattin, P.C.; Zam, A. Laser-induced breakdown spectroscopy as a potential tool for autocarbonization detection in laserosteotomy. J. Biomed. Opt. 2018, 23, 071206. [Google Scholar] [CrossRef] [PubMed]
  239. Ying, L.; Yanhong, G.; Ying, Z.; Yuandong, L.; Yuan, L. Analytical study of seashell using laser-induced breakdown spectroscopy. Plasma Sci. Technol. 2017, 19, 52–55. [Google Scholar] [CrossRef]
  240. Chirinos, J.R.; Oropeza, D.D.; Gonzalez, J.J.; Hou, H.; Morey, M.; Zorba, V.; Russo, R.E. Simultaneous 3-dimensional elemental imaging with LIBS and LA-ICP-MS. J. Anal. At. Spectrom. 2014, 29, 1292–1298. [Google Scholar] [CrossRef]
  241. Jantzi, S.C.; Motto-Ros, V.; Trichard, F.; Markushin, Y.; Melikechi, N.; De Giacomo, A. Sample treatment and preparation for laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2016, 115, 52–63. [Google Scholar] [CrossRef]
  242. Senesi, G.S.; Harmon, R.S.; Hark, R.R. Field-portable and handheld laser-induced breakdown spectroscopy: Historical review, current status and future prospects. Spectrochim. Acta Part B At. Spectrosc. 2021, 175, 106013. [Google Scholar] [CrossRef]
Figure 1. LIBS timeline highlights the achieved milestones in the field of cancer detection and calcified tissue analysis over the past decade [39,45,46,47,48,49,50,51,52,53].
Figure 1. LIBS timeline highlights the achieved milestones in the field of cancer detection and calcified tissue analysis over the past decade [39,45,46,47,48,49,50,51,52,53].
Molecules 30 04176 g001
Figure 2. (a) schematic diagram of LIBS instrument showing primary components: pulsed laser used to generate plasma of sample placed in moveable sample stage, optical collection system connected to spectrometer by an optical fibre, and computer system for qualitative and quantitative elemental analysis; (b) spectral analysis by homemade elemental imaging software built in LabVIEW environment (adapted from [38] with permission).
Figure 2. (a) schematic diagram of LIBS instrument showing primary components: pulsed laser used to generate plasma of sample placed in moveable sample stage, optical collection system connected to spectrometer by an optical fibre, and computer system for qualitative and quantitative elemental analysis; (b) spectral analysis by homemade elemental imaging software built in LabVIEW environment (adapted from [38] with permission).
Molecules 30 04176 g002
Figure 3. (a) Most abundant element in the human body; (b) categorisation of elements into major, minor, and trace elements based on percentage in the human body (data collection [94]).
Figure 3. (a) Most abundant element in the human body; (b) categorisation of elements into major, minor, and trace elements based on percentage in the human body (data collection [94]).
Molecules 30 04176 g003
Figure 4. Histological [hematoxylin and erosion staining (HES)] images complemented with LIBS elemental images of Mg and Ca for basal cell carcinoma (BCC) tumour sample, adapted from [51] with permission, licensed under CC by 3.0.
Figure 4. Histological [hematoxylin and erosion staining (HES)] images complemented with LIBS elemental images of Mg and Ca for basal cell carcinoma (BCC) tumour sample, adapted from [51] with permission, licensed under CC by 3.0.
Molecules 30 04176 g004
Figure 5. Reported distinct cancer cases globally in 2020: (a) incident rate; (b) casualties [189].
Figure 5. Reported distinct cancer cases globally in 2020: (a) incident rate; (b) casualties [189].
Molecules 30 04176 g005
Figure 6. Distribution of journal-published articles on LIBS in cancer research for the period 2015–2025 (45 original articles). Statistical information about (a) number of articles related to different types of cancer; (b) adoption of methodologies and clinical trials; (c) number of studies performed on human and non-human specimens. Types of cancer and related references: skin [51,52,82,117,118,125,145,150,168,171,174], breast [35,47,127,162,163,169,175], lung [121,157,158,159,160,164], blood [48,119,166], brain [131,151,165], stomach [123,156], colon [50,81,120], cervical [31,161], ovarian [27,172], prostate [122], and others [39,83,132,153,199].
Figure 6. Distribution of journal-published articles on LIBS in cancer research for the period 2015–2025 (45 original articles). Statistical information about (a) number of articles related to different types of cancer; (b) adoption of methodologies and clinical trials; (c) number of studies performed on human and non-human specimens. Types of cancer and related references: skin [51,52,82,117,118,125,145,150,168,171,174], breast [35,47,127,162,163,169,175], lung [121,157,158,159,160,164], blood [48,119,166], brain [131,151,165], stomach [123,156], colon [50,81,120], cervical [31,161], ovarian [27,172], prostate [122], and others [39,83,132,153,199].
Molecules 30 04176 g006
Figure 8. Schematic design of LIBS coupled with (a) optical fibre (adapted from [210] with permission); (b) RS (adapted from [42] with permission); (c) polyvinylidene fluoride (PVDF) based photoacoustic sensor (adapted from [107] with permission); (d) LA-ICP-MS (adapted from [240] with permission).
Figure 8. Schematic design of LIBS coupled with (a) optical fibre (adapted from [210] with permission); (b) RS (adapted from [42] with permission); (c) polyvinylidene fluoride (PVDF) based photoacoustic sensor (adapted from [107] with permission); (d) LA-ICP-MS (adapted from [240] with permission).
Molecules 30 04176 g008
Table 1. Influence of matrix effect (ME), challenges to LIBS analyses of human and animal tissues (hard and soft).
Table 1. Influence of matrix effect (ME), challenges to LIBS analyses of human and animal tissues (hard and soft).
Material/TissuesMatrix Elements/Non-Matrix ElementsChallengesCommentsRef.
Human samplesSkin tissues (cancerous and healthy)C, N, H, O/Ca, Mg, Na, KLimited lateral resolution, unable to map trace elements with precisionRatio of intensities used for standardisation to eliminate ME; spatial resolution in LIBS imaging lower than for LA-ICP-MS [51,117]
Skin tissuesC, N, H, O/Ca, Mg, K, P, Fe, NaSpectral fluctuationExtensive pre-processing methods (standard normal variate, autoscaling, auto centring, normalisation by area) used to enhance model efficiencies [118]
BloodC, N, O, H, Ca, P/Fe, K, Na, MgSpectral fluctuationME of filter paper substrate and blood add up; signals of filter paper not subtracted, normalisation of intensities insufficient[119]
SerumK, Na, Ca, Mg/Zn, CuPoor signal-to-noise ratio (SNR)Self-absorption factor (0.6) to be reduced for reliable analysis[120]
Lung tumourC, N, H, O, P, S/Ca, Mg, Cu, FeDetection of anomalous spectraPre-processing and z-score method used to reduce fluctuation; uncertainty about variation in intensity (ME or abnormalities)[121]
Prostate malignant tissueC, N, H, O/Na, FeLow intensities, high background noiseTrace elemental detection and quantification are lacking[122]
Gastric tissuesC, N, H, O, Ca, Na/MgPoor SNRBiopsies and extensive sample preparation required [123]
TeethCa, P/-Poor quantification precisionCertified reference materials for dental tissues not available; plasma properties of tooth samples and reference materials considerably different [124]
TeethCa, P/Al, Ba, Hg, Pb, SrHigh limit of detection for trace elementsNon-matrix elements from filling materials migrate to matrix, unreliable quantification[106]
Teeth, bonesCa, P/Mg, SrPoor repeatability, LIBS-ML model inefficient for classification LIBS-NN model cannot be generalised; different models used for classification, even for specimens having similar matrices[116]
Animal samplesSoft pork tissues (fat, skin, and muscle)C, N, H, O, K, Ca, Na/-Spectral fluctuation LIBS-SVM classification sensitivity (74%) for similar muscle tissues is limited[114]
Porcine organic tissues (liver, brain, kidney, heart, lung, and skeletal muscle)C, N, H, O/Ca, Mg, Na, K, FePoor quantitative analysisMatrix-matched reference materials required; larger data set of animal tissues required for validation [109]
Porcine tissues (bone, muscle, and fat)C, N, H, O/Ca, K, Na, Zn, Fe, ClBackground noiseLIBS-DFA model sensitivity and specificity for discrimination of soft tissues (muscle and fat) 10% less than that for bones [115]
Pig’s soft tissues (fats and nerves)C, N, H, O, P, S/Ca, Na, KPoor classification accuracyModel performances to be improved; ex vivo analysis may not be applicable for in vivo studies[113]
Rib bones of boarCa, P/Mg, SrUncertainty of measured plasma parameters restricts accurate elemental quantificationOptimisation of experimental parameters (gate delay, gas pressure, spectral range); enhanced CF-LIBS performance for all matrix elements of all tissues uncertain [49]
Mice skin tissues (melanoma lesion and normal) C, N, H, O, P/MgLow average spatial resolution due to instrumental limitations Unable to discriminate between highly similar tissues; conventional histological imaging used for cross-validation[125]
Table 2. Reported pitfalls of AI models for the interpretation of LIBS data in various cancer investigations, along with possible solutions.
Table 2. Reported pitfalls of AI models for the interpretation of LIBS data in various cancer investigations, along with possible solutions.
Type of Cancer/Methods/Refs.PitfallsRemarks
Gastric/LIBS-KNN/[156]KNN: poor scalability, curse of dimensionality, lack of interpretabilityGeneralised additive models (GAM) and XGBoost could be better alternatives
Lungs/LIBS-PCA/[159]PCA: compromises important features having low variations, ineffective for non-linear relationshipsFor non-linear issues, kernel PCA can be employed
Lungs/LIBS-1D-ResNet/[164]1D ResNet: domain-sensitive means model trained on one type of spectrometer; for data from another spectrometer transfer learning methods requiredFine-tuning can be effective in resolving domain sensitivity; regularising techniques and weight loss functions can reduce over-fitting
Lungs/LIBS-KPCA/[157]Kernel has its own critical hyperparameters for tuning bandwidth and their values are not reported; KPCA is sensitive to outliers It can be useful for a non-linear data set when the goal is to get predictive outcomes without understanding the features
Lungs/LIBS-Bagged tree model/[160]Bagged tree model suffers from overfitting when a smaller number of spectra (280) is usedOverfitting can be reduced by eliminating noise from the LIBS spectrum
Cervical/LIBS-SVM/[31]SVM is kernel-dependent, wrong selection leads to poor accuracy; effective for binary classification, unable to deal with multiple stages of cancerUse of cross-validation is recommended for kernel selection because they have different regularisation parameters
Cervical/LIBS-CNN/[161]CNN decision interpretation is not convincing; vulnerable to adversarial attacks; difficult to handle and has poor robustness Adversarial training and multiple CNNs can provide generalisation
Brain/LIBS-SNN/[151]Heuristic algorithms (cuckoo search algorithm used) computationally challenging; new method, not yet applied in other LIBS-medical studies Usage of neuromorphic hardware can increase computational speed; hybrid architectures (ANN-SNN or CNN-SNN) can be more effective
Brain/LIBS-SVM/[165]Identification accuracy of SVM is low (~60%); least reliable in intraoperative tumour classificationsBayesian optimisation method can be applied for tuning penalty and kernel function parameters instead of using particle swarm optimisation (PSO)
Ovarian/LIBS-BPNN/[27]BPNN model detail activation functions, regularisation and training strategy (epochs, batch size, and learning rate) is missingModel description is required for reproducing the
results in other labs
Breast/LIBS-PSCC/[162]Significant drop in performance of CNN when shared pre-processing module is removed (accuracy decreased from 90% to 82%); PSCNN outcomes are instrument-dependent In limited scenarios, shared pre-processing can be effective
Blood/LIBS-RSM-LDA/[166]Randomness excludes critical features; can repeatedly pick correlated features, which reduces diversity among learners and weakens ensemble effectMechanistic correlation improvement in subtype discrimination
Teeth/LIBS-ANN/[147]ANN requires large data set, 500 spectra are insufficient for analysis; it is an extension of SVM based on sigmoid functionsTools (LIME and SHAP) can interpret ANN predictions for medical use
Bones/LIBS-SIMCA/[152]Non-discriminative model which requires pre-processing and PCA before applying SIMCA, limited accuracy of 58% is attainedKernel PCA or non-linear relationship kernel SIMCA can be used for biomedical data
Table 3. Instrumentation and environmental conditions employed for LIBS measurements of different types of cancer. Laser parameters are wavelength, repetition rate, pulse duration, energy per pulse, and fluence. Spectrometer parameters are name/model/available information, spectral resolution, and spectral coverage range. Detector parameters are gate delay and integration time. The environment, type of cancer, and reference are included. Asterisk symbol (*) indicates cancer research was conducted on non-human samples.
Table 3. Instrumentation and environmental conditions employed for LIBS measurements of different types of cancer. Laser parameters are wavelength, repetition rate, pulse duration, energy per pulse, and fluence. Spectrometer parameters are name/model/available information, spectral resolution, and spectral coverage range. Detector parameters are gate delay and integration time. The environment, type of cancer, and reference are included. Asterisk symbol (*) indicates cancer research was conducted on non-human samples.
Ref.Laser (Parameters)SpectrometerSpectral
Resolution
Spectral RangeGate DelayIntegration TimeEnvironmentCancer Type
[150]Nd:YAG (1064 nm, 1 Hz, 10 ns, 30 mJ)Avantes AvaSpec 20480.20–0.30 nm190–1100 nm1.28 µs1.05 msAirSkin
[145] Nd:YAG (532 nm, 5 Hz, 5 ns, 7.49 mJ)Multichannel
Instruments
0.1 nm197–1045 nm0.2 µs 1.05 msArgonSkin
[48] Nd:YAG (1064 nm, 5 Hz, 8 ns, 73 mJ)Avantes AvaSpec ULS2048-40.09–0.22 nm200–850 nm5 µs-AirSkin, Blood
[118] Nd:YAG (1064 nm, 1 Hz, 5 ns, 64 mJ)Avantes AvaSpec 2048-2-USB20.2–0.3 nm190–1100 nm1.28 µs2 msAirSkin
[51] Nd:YAG (532 nm, 20 Hz, 10 ns, 7.49 mJ)Czerny Turner (SR-500i-B2-R)-275–775 nm0.5 µs -ArSkin
[117] Nd:YAG (266 nm, 50 Hz, 8 ns, 8 mJ)Czerny Turner-240–407 nm0.3 µs-ArSkin
[52] Nd:YAG (1064 nm, 4 ns)Single channel0.7 nm270–800 nm-1 msAirSkin
[168] Ti:Sapphire (775 nm, 150 fs, 1.20 mJ & 1.44 mJ) Echelle--50 ns700 µs HeliumSkin *
[125] Ytterbium (1030 nm, 550 fs, 250 µJ)-0.1 nm200–900 nm0.1 µs-ArSkin *
[82] Ytterbium (1030 nm, 343 nm, 550 fs, 30–80 µJ, 7.42 J/cm2)Single spectrometer0.4 nm240–800 nm20 ns-AirSkin *
[175]Nd:YAG (532 nm, 10 Hz, 103 mJ)Avantes Ava Spec 20480.08 nm190–770 nm1 µs2 msAirBreast
[162,163]Nd:YAG (532 nm, 1 Hz, 98.6 mJ)Avantes AvaSpec ULS4096CL-Evo-200–900 nm2 µs-AirBreast
[35,47] Nd:YAG (1064 nm, 1 Hz, 10 ns, 150 mJ)Avantes Ava Spec 20480.4 nm200–1100 nm1.28 µs-AirBreast, Colon, Larynx,
Tongue
[169] Nd:YAG (1064 nm, 10 Hz, 6 mJ)Avantes-182–600 nm2 µs-ArBreast *
[81] Ti:Sapph. (785 nm, 1 KHz, 30 fs, 7 µJ)L.O.T. Oriel Multispec MS1251 nm 23 ns-AirBreast, Liver
[158] Nd:YAG (532 nm, 10 Hz, 8 ns, 175 mJ)Avantes AvaSpec ULS4096CL-EVO-200–950 nm2 µs-AirLungs, Esophageal
[160]Nd:YAG (1064 nm, 10 Hz, 10 ns, 65 mJ)Mechelle Me5000 200–850 nm1 µs1 µsAirLungs
[159] Nd:YAG (1064 nm, 10 Hz, 10 ns, 65 mJ)Mechelle Me5000-200–850 nm1 µs1 msAirLungs
[164] Nd:YAG (1064 nm, 10 Hz, 40 mJ)--240–850 nm6 µs-AirLungs
[121]Nd:YAG (1064 nm, 10 Hz, 5 ns, 50 mJ)Mechelle Me5000-200–900 nm3 µs-AirLungs
[157] Nd:YAG (10 Hz, 10 ns, 40 mJ)--200–900 nm3 µs-AirLungs
[166] Nd:YAG (532 nm, 10 Hz, 8 ns, 30 mJ)Echelle-200–950 nm1 µs-AirBlood
[119] Nd:YAG (1064 nm, 5 Hz, 8 ns, 73 mJ)Avantes AvaSpec ULS2048-40.09–0.22 nm200–850 nm5 µs-AirBlood
[50]Nd:YAG (266 nm, 20 Hz, 8 ns, 50 mJ)SR 500i-A-280–900 nm500 ns-AirColon
[120] Nd:YAG (1064 nm, 10 Hz, 10 ns, 20–30 mJ)Mechelle Me5000-200–975 nm300 ns-AirColon
[123] Nd:YAG (1064 nm, 1 Hz, 6 ns, 30 mJ)Echelle (Kestrel, SE200)-200–800 nm1 µs-AirStomach
[31]Nd:YAG (532 nm, 5 Hz, 8 ns, 30 mJ)Mechelle Me5000-200–900 nm0.9 µs1 sAirCervical
[161]Nd:YAG (1064 nm, 10 Hz, 6 ns, 50 mJ)Echelle (Aryelle 200)-193–840 nm--AirCervical
[27] Nd:YAG (1064 nm, 7 ns, 30 mJ)Mechelle Me5000 230–900 nm0.8 µs-AirOvarian
[45]Ti:Sapphire (775 nm, 150 fs, 1.6 mJ)Mechelle Me5000--50 ns700 µs AirOvarian
[172] Ti:Sapphire (775 nm, 150 fs, 1.54 mJ)Mechelle Me50000.013–0.056 nm220–850 nm50 ns700 µs HeliumOvarian*
[151,165] Nd:YAG (1064 nm, 1 Hz, 5 ns, 40 mJ)Avantes AvaSpec 2048-2-USB20.2–0.3 nm190–1100 nm1.29 µs2 msAirBrain
[131]Nd:YAG (1064 nm, 1 Hz, 10 ns, 50 mJ)Avantes AvaSpec 20480.4 nm200–1100 nm1.2 µs2 msAirBrain
[39] Nd:YAG (1064 nm, 1 kHz, 7 ns, 270 µJ)Five channels-187–887 nm0.5 µs -AirGallbladder
[122]Nd:YAG (1064 nm, 10 Hz, 8 ns, 40 mJ)Czerny-Turner0.3 nm250–800 nm2 µs, 10 1 µs-AirProstrate
[153]Nd:YAG (1064 nm, 10 Hz, 8 ns, 270 µJ)--127–868 nm--AirOral
[132]Nd:YAG (1064 nm, 10 Hz, 6 ns, 8 mJ)Mechelle (Me5000),
Czerny-Turner
0.05 nm, 0.1 nm250–900 nm900 ns-AirOral
Table 4. Comparison of laser-induced breakdown spectroscopy (LIBS) with other spectroscopic and optical methods in skin cancer diagnosis based on purpose of study, sample preparation (SP) method, type of analysis (TA), associated biomarkers, and limitations [48,52,117,118,125,145,171,174,182,183,184,185,186,187,188].
Table 4. Comparison of laser-induced breakdown spectroscopy (LIBS) with other spectroscopic and optical methods in skin cancer diagnosis based on purpose of study, sample preparation (SP) method, type of analysis (TA), associated biomarkers, and limitations [48,52,117,118,125,145,171,174,182,183,184,185,186,187,188].
Laser-Induced Breakdown Spectroscopy (LIBS)Other Optical/Spectroscopical Modalities
ObjectivesSP/TABiomarkersRemarksObjectivesSP/TABiomarkersRemarks
Diagnostic accuracy of LIBS-DNN for skin cancerNo/in vivoHigher intensities of Ca, Na, and Fe for cancerous tissuesUse of limited dataset, further clinical studies required Improvement in skin cancer detection by RS-DLMNo/in vivoHigher line intensities at certain wavenumbers for cancer tissuesBinary classification restricted to exploring staging of disease
Evaluation of LIBS-Raman data fusion method in melanoma diagnosisFFPE/in vitroHigher intensities (Ca, Mg), alterations in Raman bands for cancer tissuesAssumption-based study performed only on two subjectsMorpho-chemical characterisation of skin cancer using LC-OCT and CRMNo/ex vivoHigher intensities of SCC at 821, 1012, 1220, 1446, 1580, 2931 cm−1Fluctuation of intensity for the same tissue, same band shift for different pathologies
* Identification of melanoma lesions from surrounding dermisEmbedded and slicing/in vitroHigher intensities (Ca, Mg) in affected tissuesInjection of melanoma into mice, cannot apply to human studiesMelanoma cell identification from melanocyte cells by RS Incubation and centrifugation/in vitroHigher intensities of Melanoma cells at 645, 947, 1030, 1453, 1582 cm−1 Sample preparation involved; limited number of samples being used
LIBS imaging of cutaneous tumours correlated with LA-ICP-MS imagingFFPE/invitroHigher Ca and Mg content in tumour regionPoor spatial resolution due to ablation; low sensitivity due to matrix effect and self-absorption LA-ICP-MS elemental imaging of cutaneous tissue as a complementary methodTissues on glass slide/in vitroHigh intensities of Ca, Mg, P, and Zn in tumour tissuesSpatial and in-depth resolution limited, ablation heterogeneous and destructive, matrix effects
Spectral analysis using ML algorithms for classification of melanoma stagesFFPE/in vitroP, Ca, Mg, KSemi-destructive nature of LIBS restricts in vivo analysisDMF images of skin lesions processed by AI models to achieve diagnostic accuracyNo/in vivoCancer-induced alteration in signals of molecules (melanin and keratin)Accuracy of diagnostic model not high, time-consuming procedure
Classification of melanoma, lymphoma, and healthy subjects based on LIBS-ML algorithmsSerum drops on dry filter paper/in vitroCa, Na, K, H, O, NReproducibility issues due to shot-to-shot fluctuations, use of filter paper causes uncertaintiesMultiphoton microscopy with DL model for diagnostic information on non-melanoma cancerFFPE/in vitroMPM images exhibit distinct features for both healthy and abnormal surfacesResults not very reliable, cross-validation is required
Diagnosis of subtype of melanoma malignancies using LIBS elemental imagingFFPE/in vitroCa, MgSmall sample size (17) used to classify six subtypes of melanoma CLSM image classification for diagnostic prediction of skin cancerStaining procedure/in vitroAlteration in image processing for two groups of normal and SCC cellsTraining of technicians required due to complex and extensive diagnostic procedure
* Examine the melanoma malignancy using fs-LIBS elemental imagingFrozen sectioning/in vitroCa, MgVariation of line intensities (plasma fluctuation); low spatial resolutionProcessing of FTIR hyperspectral images to classify skin tumour cells Cells grown on crystal surface/Ex vivoMalignant cell lines grow disordered, skin cells are flattenedMisclassification by environmental water vapour interference and artefacts
Symbol (*) indicates cancer study performed on non-human specimens. FFPE: formalin-fixed, paraffin-embedded; RS-DLM: Raman spectroscopy and deep learning models; LC-OCT: line-field confocal optical coherence tomography; RS: Raman spectroscopy; CRM: confocal Raman microscopy; LA-ICP-MS: laser-ablation inductively coupled plasma mass spectroscopy; DMF: dermatofluorscopy; CLSM: confocal laser scanning microscopy; FTIR: Fourier transform infrared spectroscopy.
Table 5. Summary of recent research works on breast cancer diagnosis by LIBS and other techniques considering study objectives, sample preparation (SP) method, type of analysis (TA), outcomes, and challenges.
Table 5. Summary of recent research works on breast cancer diagnosis by LIBS and other techniques considering study objectives, sample preparation (SP) method, type of analysis (TA), outcomes, and challenges.
MethodsObjectivesSP/TAFindingsChallengesRef.
Laser-induced breakdown spectroscopy (LIBS)Identify cancerous tissues (breast, colon, larynx, and tongue) from normal ones.Tissues kept in formalin, cut in slices of ~5  ×  5  ×  2 mm3
/in vitro
Higher plasma temperature for cancerous tissues due to presence of trace elementsErrors in measuring plasma parameters cause uncertainty in element quantification[35,47]
Laser ablation inductively coupled plasma time of flight mass spectroscopy (LA-ICP-TOF–MS)Comparative analysis of LA-ICP-TOF-MS and histological staining for breast cancer investigationParaffin-embedded breast tissues/in vitroBoth methods revealed elevated levels of Cu, Zn, Sr, and Ba in abnormal tissuesCalibration range 0–33 µg g−1, limit of quantification for various metals 11–83 ng g−1.[193]
LIBS and Fourier transform infrared spectroscopy (LIBS-FTIR)LIBS and FTIR data fusion correction of batch effects in serum spectra for accurate classification of breast cancer using GRAN frameworkCentrifugation and storage of the serum sample/in vitroSuperior feature extraction from combined elemental and molecular spectra by CNN model improved diagnosis by mitigating batch effects Different sample compatibility, computational complexity, concern about reproducibility of results, lack of interpretation of an observed effect [163]
Particle-induced X-ray emission (PIXE)Measuring alteration of trace elements in breast cancer patients undergoing chemotherapy and restoration to normal levels after chemotherapy Centrifugation and freezing of serum samples/in vitroElevation of Ca, Cr, Fe, and Cu and alleviation of Ti, Zn, and Se in breast cancer tissues Requires solid sample, light elements not detected, high uncertainty in quantification heterogeneous samples [194]
X-ray fluorescence (XRF), X-ray absorption spectroscopy, wide-angle X-ray scatteringDetermining the degree of microcalcification (MC) and trace elements association with breast cancer malignancies FFPE/in vitroIrregular hydroxyapatite crystal arrangement and distribution of trace elements (Na, S, Cl, Sr and Y) of cancer-affected tissues Difficult to detect light elements, overlapping of spectral peaks, requirement of matrix match standards[195]
Inductively coupled plasma mass spectroscopy (ICP-MS), Inductively coupled plasma optical emission spectroscopy (ICP-OES)Quantification of variation in elemental composition in blood samples of patientsDigestion in microwave oven/in vitroReduction in Se and Cr and elevation in Na content in blood of breast cancer patientsDigestion of sample in acid, unable to detect light elements and to quantify halogens, spectral drift; spectral interferences[196]
Table 6. Overview of LIBS studies on various types of cancer, including sample size and spectral analysis information (number of measured spectra, data pre-processing, measurement of plasma parameters, and adoption of AI models).
Table 6. Overview of LIBS studies on various types of cancer, including sample size and spectral analysis information (number of measured spectra, data pre-processing, measurement of plasma parameters, and adoption of AI models).
Cancer Type and Ref.Sample PreparationSample SizeNumber of SpectraData PreprocessingPlasma ParametersDimensionality ReductionModelModel Accuracy
Blood and skin [48]ModerateModerateKNNHigh
Skin [118] -SmallANN/PLS-DAHigh
Skin * [150] ModerateSmallBP_AdaBoostHigh
Skin [52]LargeLargeDNN-
Skin * [168]--Gradient boostingHigh
Breast [163]LargeSmallCNNLow
Breast [175]LargeModerateSVMLow
Breast [162] LargeModeratePSCNNHigh
Breast/liver [81]SmallLargeRF/ANNHigh
Breast [35]LargeLarge
Breast/colon/larynx/tongue [47]LargeLarge
Lungs/liver/oesophageal [158]LargeModerateSVMModerate
Lungs [121]LargeSmallXGBoostLow
Lungs [159]LargeSmallPCA-Boosting treeHigh
Lungs [164]LargeSmallRF-1D ResNetHigh
Lungs [160]ModerateSmallKNNHigh
Lungs [157]ModerateSmallKPCA-SVMHigh
Blood [166]ModerateSmall--RSM-LDAHigh
Blood [119]ModerateModerateLDA/KNNHigh
Brain [131]---
Brain [165]SmallSmallSVMHigh
Brain [151]SmallSmallSNNModerate
Colon [120]SmallSmall
Colon [50]SmallSmall
Stomach [123]Small-
Stomach [156]SmallSmallSVM/KNN/PLS-DAHigh
* Ovarian [172]LargeLargeRFLow
Ovarian [27]LargeModerateBPNN-
Cervical [31] SmallSmallPCA-SVMHigh
Oral [153]SmallModerateLR-
Sample size: small 1–20, medium 21–50, large > 50. Number of LIBS spectra: small < 3000, medium 3001–15,000, large > 15,000. Model accuracy: low < 80%, moderate 80–90%, high > 90%. Data pre-processing includes baseline correction, normalisation, and standardisation. Plasma parameters are plasma temperature, electron number density, and plasma frequency. Dimensionality reduction by principal component analysis and feature selection methods. Symbol (*) indicates cancer studies performed on non-human samples.
Table 7. Overview of AI models applied on LIBS data from measurements of cancer (diagnosis, screening, and staging). LIBS measurements of different samples and types of cancer. Model classification performance metrics are accuracy, sensitivity, specificity, receiver operating characteristic ROC (area under curve AUC), and cross-validation. Symbol (*) indicates cancer studies performed on non-human samples.
Table 7. Overview of AI models applied on LIBS data from measurements of cancer (diagnosis, screening, and staging). LIBS measurements of different samples and types of cancer. Model classification performance metrics are accuracy, sensitivity, specificity, receiver operating characteristic ROC (area under curve AUC), and cross-validation. Symbol (*) indicates cancer studies performed on non-human samples.
AI Model/SampleAccuracy (%)Sensitivity (%)Specificity (%)ROC Curve (AUC)Cross-Validation Cancer Type, Ref.
PCA-KNN/Serum969795.60.9910-foldsBlood [48]
PCA-KNN/Serum9689.299.40.98610-foldsSkin [48]
DNN/skin tissues-94.688.9 10-foldsSkin [52]
PCA-LDA/pellets for melanoma-99.4100-10-foldsSkin * [145]
PCA-LDA/excised tissues of melanoma-96.799.7-10-foldsSkin * [145]
BP_AdaBoost/serum for early screening86.1---10-foldsSkin * [150]
BP_AdaBoost/serum for staging96.1---10-foldsSkin * [150]
ANN/melanoma FFPE1001001001-Skin [118]
PLS-DA/melanoma FFPE1001001001-Skin [118]
Gradient boosting/Serum on Cu substrate96.3---5-foldsSkin * [168]
CNN/serum (batch 2)59.90.480.710.64-Breast [163]
GRAN/serum (batch 2)89.70.990.800.950-Breast [163]
Narrow NN/whole blood91.797.287.50.9310-foldBreast [175]
Decision fine Tree/serum89.795.283.30.8710-foldBreast [175]
PSCNN/blood plasma9086940.955-foldsBreast [162]
RF, ANN, KNN/breast and liver tissue on quartz glass substrate>94---10-foldsBreast
Liver [81]
BVF/serum samples on silicon substrate92.5392.92---Lungs Liver Esophageal [158]
CNN/lung tissues99.1799.1799.881-Lungs [121]
RF boosting tree/lung tissues98.999.398.60.98210-foldsLungs [159]
RF-1D ResNet/lung tissues91.191.391.30.99-Lungs [164]
Bagged tree/tumour and normal tissues98.998.699.30.98210-foldsLungs [160]
KPCA-SVM/tumour and normal tissues99.0399.7298.890.970-Lungs [157]
RSM-LDA/serum91----Blood [166]
PCA-LDA/blood99.7899.699.8110-foldsBlood [119]
PCA-KNN/blood99.7299.799.7110-foldsBlood [119]
FS-SVM/glioma and infiltrative tissue samples95----Brain [165]
SNN/tumour tissues88.62----Brain [151]
BPNN/blood-71.486.5 -Ovarian [27]
CNN/cervical cancer cells on silicon wafer97.92----Cervical [161]
PCA-SVM/cervical tissues embedded in paraffin wax94.4----Cervical [31]
PCA-NN/prostate tissue microarrays97---5-foldsProstrate [122]
PCA/blood and biological fluids on superhydrophobic (PDMS) substrate-8896--Oral [132]
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

Dastageer, M.M.; Siraj, K.; Pedarnig, J.D.; Zhang, D.; Qasim, M.; Rahim, M.S.A.; Mushtaq, S.; Younas, Q.; Hussain, B. From Fundamentals of Laser-Induced Breakdown Spectroscopy to Recent Advancements in Cancer Detection and Calcified Tissues Analysis: An Overview (2015–2025). Molecules 2025, 30, 4176. https://doi.org/10.3390/molecules30214176

AMA Style

Dastageer MM, Siraj K, Pedarnig JD, Zhang D, Qasim M, Rahim MSA, Mushtaq S, Younas Q, Hussain B. From Fundamentals of Laser-Induced Breakdown Spectroscopy to Recent Advancements in Cancer Detection and Calcified Tissues Analysis: An Overview (2015–2025). Molecules. 2025; 30(21):4176. https://doi.org/10.3390/molecules30214176

Chicago/Turabian Style

Dastageer, Muhammad Mustafa, Khurram Siraj, Johannes David Pedarnig, Dacheng Zhang, Muhammad Qasim, Muhammad Shahzad Abdul Rahim, Saba Mushtaq, Qaneeta Younas, and Bareera Hussain. 2025. "From Fundamentals of Laser-Induced Breakdown Spectroscopy to Recent Advancements in Cancer Detection and Calcified Tissues Analysis: An Overview (2015–2025)" Molecules 30, no. 21: 4176. https://doi.org/10.3390/molecules30214176

APA Style

Dastageer, M. M., Siraj, K., Pedarnig, J. D., Zhang, D., Qasim, M., Rahim, M. S. A., Mushtaq, S., Younas, Q., & Hussain, B. (2025). From Fundamentals of Laser-Induced Breakdown Spectroscopy to Recent Advancements in Cancer Detection and Calcified Tissues Analysis: An Overview (2015–2025). Molecules, 30(21), 4176. https://doi.org/10.3390/molecules30214176

Article Metrics

Back to TopTop