Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-05-14T16:47:06.332Z Has data issue: false hasContentIssue false

Forty Years of Laser-Induced Breakdown Spectroscopy and Laser and Particle Beams

Published online by Cambridge University Press:  01 January 2024

Vincenzo Palleschi*
Affiliation:
Applied and Laser Spectroscopy Laboratory, Institute of Chemistry of Organometallic Compounds, National Research Council Research Area, Via G. Moruzzi 1–56124, Pisa, Italy
*
Correspondence should be addressed to Vincenzo Palleschi; vincenzo.palleschi@cnr.it
Rights & Permissions [Opens in a new window]

Abstract

The laser-induced breakdown spectroscopy (LIBS) technique is one of the most promising laser-based analytical techniques. Coincidentally, the LIBS acronym was proposed by Radziemski and Loree in two seminal papers published in 1981, almost at the same time in which the Laser and Particle Beams journal started its publication. In this contribution, the evolution of the LIBS technique is discussed following a chronological collection of key papers in LIBS, some of which were in fact published on LPB.

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © 2023 Vincenzo Palleschi.

1. Introduction

Laser-induced breakdown spectroscopy (LIBS) is an atomic emission spectroscopic technique based on the spectral analysis of the plasma induced by a pulsed laser beam in gas, liquid, or solid targets to obtain information about the materials under study.

The principles of the LIBS technique are deeply rooted on the preexisting knowledge on flame and plasma spectroscopy which largely precedes the discovery of the laser. After the introduction of the laser, techniques similar to modern LIBS were proposed [Reference Brech and Cross1]; however, as in the 1962 Brech and Cross paper, at those times, the laser was used essentially for ablating a solid sample, while the excitation of the material was obtained through an electrical spark. The main characteristics of LIBS as it is now practiced are, on the other hand, to use the laser for obtaining at the same time the sampling of the material and its heating for producing the atomic optical emission [Reference Palleschi2].

This characteristic is peculiar of the LIBS technique and brings the exceptional advantages of operating on not treated materials in very short time, which in turn allow the use of the technique for remote in situ analysis in hostile environments. On the other hand, the use of a single tool for sampling and excitation prevents the possibility of independent optimization of the two processes, leading to analytical performances that are usually considered modest with respect to other conventional laboratory spectrochemical techniques.

A typical LIBS experiment involves the use of a pulsed laser, typically Nd : YAG, at the fundamental wavelength of 1064 nm, emitting pulses of a few nanoseconds with energy of several tens of milli-Joules and maximum repetition rates of a few Hertz. The light emitted by the laser-induced plasma is collected and sent to a spectrometer (gated or ungated, narrow- or wideband) where the signal is analysed using a suitable delay after the laser pulse, to reduce the continuum bremsstrahlung emission (see Figure 1).

Figure 1: A typical LIBS setup.

Many alternative experimental configurations have been realized, though. A description of some of them is given in [Reference Tognoni, Palleschi, Corsi and Cristoforetti3].

A typical LIBS spectrum, acquired with a broadband spectrometer, is shown in Figure 2.

Figure 2: A typical broadband LIBS spectrum (steel).

The attribution of the emission lines to the corresponding atomic species is usually done manually, based on the information contained in the NIST database of atomic lines [Reference Kramida, Olsen and Ralchemko4], although methods for automatic identification of the lines were also proposed [Reference Amato, Cristoforetti and Legnaioli5].

In the following, we will discuss the exceptional evolution of the LIBS technique in the last 40 years, also highlighting the important role that the Laser and Particle Beams journal has had in this evolution.

2. 1981–1990: The Early Years

The first papers where the acronym LIBS was originally proposed were published by Radziemski and Loree [Reference Radziemski and Loree6, Reference Loree and Radziemski7] in 1981. The two authors, researchers of the Los Alamos National Laboratory (Los Alamos, New Mexico, USA), outlined the principle of the LIBS technique in two companion papers, the first dealing with time-integrated detection of the plasma emission, the second discussing the analytical advantages of a time-resolved detection. The authors analysed by LIBS sodium and potassium in a coal gasifier product, airborne beryllium and phosphorous, sulphur, fluorine, and chlorine (the latter three elements particularly complex to detect by LIBS) in atmosphere. For the next 10 years, the research on LIBS remained essentially confined in North America; with the arrival of the Radziemski group of David Cremers, the application of the LIBS technique was extended to many other interesting fields, such as the study of aerosols [Reference Radziemski, Loree, Cremers and Hoffman8], the analysis of liquids [Reference Cremers, Radziemski and Loree9], detection of beryllium in air [Reference Radziemski, Cremers and Loree10] and in beryllium-copper alloys [Reference Millard, Dalling and Radziemski11], and detection of uranium in solution [Reference Wachter and Cremers12] and cadmium, lead, and zinc in aerosol [Reference Essien, Radziemski and Sneddon13].

3. 1991–2000: Evolution of LIBS

In 1991, the Pisa group published a paper dealing with the quantitative determination of pollutants in air by LIBS [Reference Palleschi, Harith, Salvetti, Singh and Vaselli14]. The paper was published on Laser and Particle Beams, and it represented the first work on LIBS published by a group outside the USA.

In the following years, other works on LIBS were published in Europe (determination of carbon in steel by the Spanish group of Aragón et al. [Reference Aragón, Aguilera and Campos15, Reference Aguilera, Aragon and Campos16]) and in Canada (quantitative analysis of aluminium alloys [Reference Sabsabi and Cielo17]).

In the 1990–2000 decade, several papers on LIBS were published using the “LIPS” (laser-induced plasma spectroscopy) acronym. This occurred mostly in Europe (see, for example [Reference Haisch and Panne18]), but some groups in Canada [Reference St-Onge, Sabsabi and Cielo19] and USA [Reference Martin and Cheng20] also adopted this terminology, which was considered more general than the original “LIBS.”

The LIPS acronym is now deprecated, after the First International LIBS Conference (LIBS 2000), organized by the Pisa group in Tirrenia, Italy [Reference Palleschi21]. It is nevertheless curious that two of the major contributions to LIBS, which introduced two techniques still widely used nowadays, were in fact referring to LIPS as the name of the technique.

The first key paper was published in 1988 by the Sabsabi group in Canada and reported on an alternative experimental configuration in which the laser energy is delivered on the sample surface in two pulses, suitably delayed [Reference St-Onge, Sabsabi and Cielo22]. The authors reported a considerable intensity enhancement in the spectral signal which was substantially independent on the interpulse delay. The physical explanation of the enhancement in double-pulse configuration was fully explained only several years after by the Palleschi group in Pisa [Reference Corsi, Cristoforetti and Giuffrida23, Reference Wen, Mao and Russo24], in terms of the reduced plasma shielding of the plasma produced by the second laser pulse, due to a reduction of the environmental gas density behind the shock wave produced by the first laser pulse. It is worth noting that the essential role of the first shock wave in double-pulse LIBS was firstly hypothesized by the Russian researcher Sergei Pershin; unfortunately, his research, published on a Russian journal [Reference Pershin25], went generally unnoticed, and the author was credited for his original intuition only recently [Reference Cristoforetti and Palleschi26].

The second important paper was published by the Pisa group in 1999 and proposed a new procedure for standard-less LIBS analysis called calibration-free LIPS (now known as CF-LIBS) [Reference Ciucci, Corsi, Palleschi, Rastelli, Salvetti and Tognoni27]. Interestingly enough, an extended description of the method was published the same year on Laser and Particle Beams [Reference Ciucci, Palleschi, Rastelli, Salvetti, Singh and Tognoni28]. The Ciucci et al. paper is the most quoted research paper (thus excluding books and reviews) in the history of LIBS.

Many applications and improvements of the CF-LIBS method have been proposed since the original papers of 1999. One of the most useful procedures is the compensation of self-absorption effects in the LIBS plasma, which produce a non-linear dependence between the analyte concentration and the LIBS line intensity (see Figure 3).

Figure 3: Schematic representation of the self-absorption effect, showing the balance between spontaneous emission and absorption (stimulated emission is normally negligible in LIBS plasmas).

The effect of self-absorption in laser-induced plasmas was studied in a key paper by the Winefordner group at the University of Florida in Gainesville, USA [Reference Gornushkin, Anzano, King, Smith, Omenetto and Winefordner29], but the implication of this research would not be transferred to CF-LIBS until the beginning of the XXI century.

4. 2001–2021: XXI Century LIBS

The first proposal to use the curve-of-growth approach to compensate for self-absorption effects in LIBS plasmas was published by the Pisa group in 2002 [Reference Bulajic, Corsi and Cristoforetti30]. A simple experimental method for evaluating the self-absorption effect and compensating it using a duplicating optical path mirror was proposed by the Gainesville group of Nicolò Omenetto in 2009 [Reference Moon, Herrera, Omenetto, Smith and Winefordner31]. This method is conceptually very simple, although its realization is rather complex and limited to a laboratory setup. A more versatile method was proposed by the Palleschi group [Reference El Sherbini, El Sherbini and Hegazy32], as a generalization of the theoretical work of Amamou et al. [Reference Amamou, Bois, Ferhat, Redon, Rossetto and Matheron33]. According to this method, the degree of self-absorption of a given emission line can be calculated and, eventually, compensated, by measuring the intensity and full width at half maximum of the emission line and the measured plasma electron number density, once the Stark broadening coefficient of the line is known [Reference Griem34].

The method proposed by the Pisa group offers the possibility of measuring the plasma electron number density from the broadening of the Balmer alpha hydrogen line [Reference Gigosos, González and Cardeñoso35]. The problem was studied in 2013 by Pardini et al. [Reference Pardini, Legnaioli and Lorenzetti36].

Despite the fact that the Pisa method was initially developed for improving the predictions of the calibration-free LIBS technique, its applications have been extended to many situations in which the self-absorption effects are important (see [Reference Rezaei, Cristoforetti, Tognoni, Legnaioli, Palleschi and Safi37] for a detailed discussion). Particularly important in this framework is the criticism to the “branching ratio” method for assessing the self-absorption effect of a spectral line published in 2021 by Urbina Medina et al. [Reference Urbina Medina, Carneiro, Rocha, Farias, Bredice and Palleschi38].

The ability of compensating for self-absorption effects opened new perspectives in the determination of spectroscopic fundamental parameters as transition probabilities [Reference Urbina, Carneiro, Rocha, Farias, Bredice and Palleschi39, Reference Aguilera, Aragón and Manrique40] and Stark broadening coefficients [Reference Poggialini, Campanella and Jafer41, Reference Bredice, Borges and Sobral42].

One of the fundamental hypotheses for the application of the calibration-free LIBS is the fulfilment of the local thermal equilibrium (LTE) condition [Reference Griem43]. An important result, obtained at the end of the first decade of the century, was the extension of the McWhirter criterion for local thermal equilibrium to non-stationary and non-homogeneous LIBS plasmas [Reference Cristoforetti, de Giacomo and Dell’Aglio44]. The implication of that research confirmed the necessity, for the use of CF-LIBS, to use time-resolved detectors. A paper by Grifoni et al. in 2014 [Reference Grifoni, Legnaioli, Lezzerini, Lorenzetti, Pagnotta and Palleschi45] provided a simple tool for extracting time-resolved information from time-integrated spectra, exploiting the differences between two or more spectra taken at different time delays.

The first decade of the century witnessed a great improvement in the performances of the LIBS technique, which accelerated its acceptation as a powerful analytical technique. In 2004, the Gainesville group published a paper, in which LIBS was defined as a possible future superstar among the atomic spectrometric techniques [Reference Winefordner, Gornushkin, Correll, Gibb, Smith and Omenetto46], and in 2010, David Hann and Nicolò Omenetto published an important review on Basic Diagnostics and Plasma-Particle Interactions [Reference Hahn and Omenetto47] which at present time is the most quoted review paper in the LIBS history.

Many applications of LIBS in the analysis of biological materials [Reference Gaudiuso, Melikechi and Abdel-Salam48Reference Rehse, Salimnia and Miziolek55], cultural heritage and archaeology [Reference Scholten, Teule, Zafiropulos and Heeren56Reference Arafat, Na’es and Kantarelou62], industry [Reference Bulajic, Cristoforetti and Corsi63Reference Li, Feng, Oderji, Luo and Ding65], and environment [Reference Karpate, Nayak and Libs66Reference Nunes, De Carvalho, Santos and Krug69] have been reported.

The group of Javier Laserna at Malaga University, Spain, demonstrated the feasibility of performing stand-off LIBS analysis at long distances (>10 meters) using an open-path configuration [Reference Palanco and Laserna70, Reference Fortes and Laserna71], thus extending dramatically the possible applications of LIBS for the analysis of industrial or environmental samples in hostile environment. Among the many exploitations of the LIBS technique, it is worth mentioning the results of a recent European project (LACOMORE—laser-based continuous monitoring and resolution of steel grades in sequence casting machines), aimed at the optimization of the continuous casting process of steel [Reference Lorenzetti, Legnaioli, Grifoni, Pagnotta and Palleschi72, Reference Cabalín, Delgado, Ruiz, Mier and Laserna73]. The project involved the world’s two most active groups in LIBS development and research of Malaga and Pisa. In the framework of that project, an open-path double-pulse LIBS instrument was successfully used for remote analysis (about 6 meters) of steel up to 900 C temperature.

An important evolution of the LIBS technique, in the first decade of the century, was the introduction of ultra-short lasers sources for plasma generation [Reference Le Drogoff, Margot and Chaker74Reference Margetic, Niemax and Hergenröder76]. Femtosecond laser pulses produce neater craters on the sample surface compared to nanosecond lasers, and the resulting spectra are characterized by a lower continuum emission because of the temporal separation between the ablation phenomenon and the creation of the plasma, which inhibits the laser-plasma interaction phenomena. The advantages of femtosecond LIBS were exploited for sub-micrometric in-depth measurements [Reference Margetic, Bolshov, Stockhaus, Niemax and Hergenröder77], analysis of biological tissues [Reference Moon, Han, Choi, Shin, Kim and Jeong78], and environmental applications [Reference Hu, Shi, Yan, Wu and Zeng79]. Combined nanosecond-femtosecond [Reference Scaffidi, Pender and Pearman80, Reference Babushok, DeLucia, Gottfried, Munson and Miziolek81] and femtosecond-femtosecond [Reference Piñon, Fotakis, Nicolas and Anglos82] dual-pulse analysis was also proposed. The possibility of maintaining the laser beam collimation at very long distances, due to the filamentation/self-focusing effects that characterize the propagation of fs-laser beams in atmosphere [Reference Hu, Shi, Yan, Wu and Zeng79, Reference Stelmaszczyk, Rohwetter and Méjean83Reference Hu, Xu, Yuan and Zeng85], has triggered many innovative applications of stand-off LIBS, including the remote analysis of cultural heritage [Reference Tzortzakis, Anglos and Gray86], biological materials [Reference Xu, Méjean and Liu87], geological samples [Reference Abdul Kalam, Balaji Manasa Rao, Jayananda and Venugopal Rao88], and explosives [Reference Shaik, Epuru and Syed89, Reference Shaik and Soma90].

After the modest results of the proposals to use LIBS for Homeland Defense, in the years following the 9/11 tragic events [Reference Fortes and Laserna71, Reference González, Lucena, Tobaria and Laserna91Reference Gottfried, De Lucia, Munson and Miziolek94], LIBS regained some public consideration for its possible application in the field of space exploration [Reference Knight, Scherbarth, Cremers and Ferris95]. Finally, on August 6, 2012, the NASA Curiosity rover landed on Mars, carrying a LIBS instrument [Reference Wiens, Maurice and Barraclough96] which is still operating after more than 10 years of activity, resulting in hundreds of thousands of spectra taken on Martian rocks. The data obtained by the LIBS instrument on Mars certainly helped in better understanding the Martian geology and contributed to the search of former life traces on the planet, opening the way to the use of LIBS in two other missions landed on Mars in 2021 (the NASA SuperCam instrument, mounted on the Perseverance rover [Reference Wiens, Maurice and Robinson97], and the Chinese MarSCoDe mounted on the Zhurong Mars rover [Reference Xu, Liu and Yan98]). However, the success of LIBS on Mars benefitted more, if possible, the development of LIBS research on Earth. In fact, the first LIBS Mars mission spurred the development of compact hand-held LIBS spectrometers, which rapidly arrived on the market of scientific instrumentation for metal analysis [Reference Afgan, Hou and Wang99, Reference Poggialini, Campanella, Legnaioli, Pagnotta, Raneri and Palleschi100], nuclear industry [Reference Garlea, Bennett, Martin, Bridges, Powell and Leckey101], and environmental applications [Reference VerMeulen, Clausen, Mosell, Morgan, Messan and Beal102], to cite some of the most important applications.

The hand-held LIBS instruments represented an impressive advance with respect to the conventional laboratory or mobile LIBS instrumentation [Reference Senesi, Harmon and Hark103], but their compact size unavoidably imposes some compromise in the analytical performances of the instrumentation. The limitations of the hand-held LIBS hardware require the use of sophisticated chemometric techniques for extracting useful information from the spectra [Reference Afgan, Hou and Wang99, Reference Poggialini, Campanella, Legnaioli, Pagnotta, Raneri and Palleschi100, Reference Zhou, Luo, Zheng and Huang104].

The use of advanced chemometric tools (artificial neural network, ANN) was firstly introduced in LIBS in 1998 [Reference Sattmann, Mönch and Krause105]. However, these techniques became widely used only in the second decade of the century. The chemometric techniques can be used for simplification (for example, principal component analysis [Reference Moncayo, Duponchel and Mousavipak106]), classification (self-organizing map (SOM) [Reference Pagnotta, Grifoni, Legnaioli, Lezzerini, Lorenzetti and Palleschi107], support vector machine (SVM) [Reference Dingari, Barman, Myakalwar, Tewari and Kumar Gundawar108], graph clustering (GC) [Reference Grifoni, Legnaioli, Lorenzetti, Pagnotta and Palleschi109], random forest (RF) [Reference Sheng, Zhang and Niu110], ANN [Reference Pokrajac, Vance and Lazarević111], etc.), and quantification of the LIBS spectra (partial least squares (PLS) analysis [Reference Yaroshchyk, Death and Spencer112], ANN [Reference Sirven, Bousquet and Canioni113], etc.). Numerous applications based on machine learning and chemometric analysis of LIBS spectra have been proposed in recent years. The most impressive are probably the applications to animal and human health (early detection of cancer, for example, [Reference Gaudiuso, Melikechi and Abdel-Salam48]), but many other examples can be cited in cultural heritage [Reference Duchene, Detalle, Bruder and Sirven114], energy production [Reference Legnaioli, Campanella, Pagnotta, Poggialini and Palleschi115], space exploration [Reference Sirven, Sallé, Mauchien, Lacour, Maurice and Manhès116], and several other fields. LIBS elemental imaging [Reference Sancey, Motto-Ros and Busser117, Reference Busser, Moncayo, Coll, Sancey and Motto-Ros118], which represents one of the most interesting developments of the technique, can produce millions of spectra. The construction and interpretation of LIBS elemental maps also benefit the chemometric algorithms [Reference Jolivet, Leprince, Moncayo, Sorbier, Lienemann and Motto-Ros119].

The growing complexity of the chemometric algorithms currently used in LIBS has stimulated some research aimed to obtain a better interpretability of the results obtained [Reference Safi, Campanella and Grifoni120, Reference Képeš, Vrábel and Adamovsky121]. An up-to-date description of the most advanced chemometric techniques can be found in [Reference Palleschi122].

Among the emerging new approaches to LIBS analysis, it is worth mentioning two interesting variations of LIBS which were proposed in the second decade of the century.

In 2011, Rick Russo proposed a technique for isotopic analysis by LIBS based on the detection of molecular emission in the spectrum. The isotopic shift of diatomic oxides or fluorides of the elements is typically larger than the ones of the corresponding atoms, thus allowing for an easier separation of the characteristic emission pattern of the different isotopes. The technique was called laser ablation molecular isotopic spectrometry (LAMIS) [Reference Russo, Bol’šhakov, Mao, McKay, Perry and Sorkhabi123], and its effectiveness was demonstrated for the isotopic analysis of several interesting elements [Reference Bol’šhakov, Mao, González and Russo124].

Another very interesting alternative approach to LIBS analysis was suggested by Alessandro De Giacomo and his group in Bari, Italy, in 2013. de Giacomo et al. reported on the use of nanoparticles to enhance the LIBS signal on metal targets [Reference de Giacomo, Gaudiuso, Koral, Dell’Aglio and de Pascale125]. The enhancement observed was comparable to the ones obtained in double-pulse LIBS, but obtained with a simpler, in principle, experimental apparatus. The mechanism of the enhancement in nanoparticle-enhanced LIBS (NELIBS) is very different from the one observed in DP-LIBS (total ablated mass, plasma electron density, and temperature are found very similar between NELIBS and conventional LIBS at the same energy). The mechanism hypothesized by the inventors of the technique involves a much larger atomization of the ablated mass in NELIBS, produced by the intense electric field which is created between one nanoparticle and the other. A model of the phenomenon has been recently published [Reference de Giacomo, Salajkova and Dell’aglio126], which also explains the dependence of the enhancement on the distance between the nanoparticles (which in turn reflects in a strong dependence of the enhancement on the concentration of the deposited nanoparticle).

The full potential of NELIBS has probably not been yet explored, but this method can possibly change the conventional narrative of LIBS as a mediocre analytical technique. The applications of NELIBS should be mainly confined to the laboratory, though, because of the need of treating, although minimally, the sample under analysis. Proposals have been presented for putting together NELIBS and DP-LIBS, in a way to possibly combine the two enhancements [Reference Poggialini, Campanella, Legnaioli, Pagnotta and Palleschi127]. Finally, in the analysis of insulators, NELIBS offers the advantage of enhancing the signal while reducing the surface damages, a feature that makes this approach particularly useful for the study of precious stones [Reference Koral, Dell’Aglio, Gaudiuso, Alrifai, Torelli and de Giacomo128], for example.

The possibility of obtaining readable LIBS spectra from minimal quantity of ablated material was specifically studied by the Malaga group of Javier Laserna, which demonstrated the feasibility of LIBS for analysis of single nanoparticles [Reference Fortes, Fernández-Bravo and Javier Laserna129] and, in more recent papers [Reference Purohit, Fortes and Laserna130Reference Purohit, Fortes and Laserna132], essentially debunked the standard narrative describing LIBS as a low sensitivity technique (in [Reference Purohit, Fortes and Laserna130], a limit of detection of 60 attograms was demonstrated in the LIBS analysis of single copper nanoparticles).

5. Conclusions

The LIBS technique has made incredible progresses during its 40 years of existence. Certainly, the evolution of LIBS has been favoured by the technological evolution of the key instrumental parts (lasers, spectrometers, and detectors); however, the improved knowledge of the basic phenomena involved in the laser-sample, laser-plasma, and plasma-sample interactions has helped in better modelling the complex chemical and physical phenomena involved in the generation of the LIBS spectral signal. In this contribution, we have tried to retrace the history of LIBS to its fundamental papers, several of which were published in this journal. The list is far from being complete. A search on the Scopus® database with keywords (LIBS OR LIPS) reports, after removal of non-pertinent works, more than 12,000 papers published on the topic since the first two in 1981, with a growth rate in 2021 of more than 1,000 papers per year. Nevertheless, we hope to have given an idea of the distance that the technique has travelled from the first laboratory application to the present interplanetary missions.

There are still many steps to do and many obstacles to overcome before the LIBS technique could be considered to have growth to its full potential, but we have no doubts that some of these important steps will be presented and commented on this journal, as it was in the past with several key publications which still represent fundamental milestones in LIBS research.

Data Availability

The relevant data are available from the author upon reasonable request.

Conflicts of Interest

The author declares that there are no conflicts of interest regarding the publication of this paper.

References

Brech, F., Cross, L., “Optical microemission stimulated by a ruby MASER,Applied Spectroscopy, vol. 16, p. 59, 1962.Google Scholar
Palleschi, V., “Laser-induced breakdown spectroscopy: principles of the technique and future trends,ChemTexts, vol. 6, no. 2, pp. 18–16, 2020.CrossRefGoogle Scholar
Tognoni, E., Palleschi, V., Corsi, M., and Cristoforetti, G., “Quantitative micro-analysis by laser-induced breakdown spectroscopy: a review of the experimental approaches,Spectrochimica Acta, Part B: Atomic Spectroscopy, vol. 57, no. 7, pp. 11151130, 2002.CrossRefGoogle Scholar
Kramida, A., Olsen, K., and Ralchemko, Y., “NIST LIBS database,2022, https://physics.nist.gov/PhysRefData/ASD/LIBS/libs-form.html.Google Scholar
Amato, G., Cristoforetti, G., Legnaioli, S. et al., “Progress towards an unassisted element identification from laser induced breakdown spectra with automatic ranking techniques inspired by text retrieval,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 65, no. 8, pp. 664670, 2010.CrossRefGoogle Scholar
Radziemski, L. J. and Loree, T. R., “Laser-induced breakdown spectroscopy: time-resolved spectrochemical applications,Plasma Chemistry and Plasma Processing, vol. 1, no. 3, pp. 281293, 1981.CrossRefGoogle Scholar
Loree, T. R. and Radziemski, L. J., “Laser-induced breakdown spectroscopy: time-integrated applications,Plasma Chemistry and Plasma Processing, vol. 1, no. 3, pp. 271279, 1981.CrossRefGoogle Scholar
Radziemski, L. J., Loree, T. R., Cremers, D. A., and Hoffman, N. M., “Time-resolved laser-induced breakdown Spectrometry of aerosols,Analytical Chemistry, vol. 55, no. 8, pp. 12461252, 1983.CrossRefGoogle Scholar
Cremers, D. A., Radziemski, L. J., and Loree, T. R., “Spectrochemical analysis of liquids using the laser spark,Applied Spectroscopy, vol. 38, no. 5, pp. 721729, 1984.CrossRefGoogle Scholar
Radziemski, L. J., Cremers, D. A., and Loree, T. R., “Detection of beryllium by laser-induced-breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 38, no. 1-2, pp. 349355, 1983.CrossRefGoogle Scholar
Millard, J. A., Dalling, R. H., and Radziemski, L. J., “Time-resolved laser-induced breakdown Spectrometry for the rapid determination of beryllium in beryllium-copper alloys,Applied Spectroscopy, vol. 40, no. 4, pp. 491494, 2016.CrossRefGoogle Scholar
Wachter, J. R. and Cremers, D. A., “Determination of uranium in solution using laser-induced breakdown spectroscopy,Applied Spectroscopy, vol. 41, no. 6, pp. 10421048, 2016.CrossRefGoogle Scholar
Essien, M., Radziemski, L. J., and Sneddon, J., “Detection of cadmium, lead and zinc in aerosols by laser-induced breakdown Spectrometry,J Anal At Spectrom, vol. 3, no. 7, pp. 985988, 1988.CrossRefGoogle Scholar
Palleschi, V., Harith, M. A., Salvetti, A., Singh, D. P., and Vaselli, M., “Time-resolved lies experiment for quantitative determination of pollutant concentrations in air,Laser and Particle Beams, vol. 9, 1991.Google Scholar
Aragón, C., Aguilera, J. A., and Campos, J., “Determination of carbon content in molten steel using laser-induced breakdown spectroscopy,Applied Spectroscopy, vol. 47, no. 5, pp. 606608, 1993.CrossRefGoogle Scholar
Aguilera, J. A., Aragon, C., and Campos, J., “Determination of carbon content in steel using laser-induced breakdown spectroscopy,Applied Spectroscopy, vol. 46, no. 9, pp. 13821387, 1992.CrossRefGoogle Scholar
Sabsabi, M. and Cielo, P., “Quantitative analysis of aluminum alloys by laser-induced breakdown spectroscopy and plasma characterization,Applied Spectroscopy, vol. 49, no. 4, pp. 499507, 1995.CrossRefGoogle Scholar
Haisch, C. and Panne, U., “Laser-induced plasma spectroscopy (LIPS) in action,Spectroscopy Europe, vol. 9, pp. 814, 1997.Google Scholar
St-Onge, L., Sabsabi, M., and Cielo, P., “Quantitative analysis of additives in solid zinc alloys ByLaser-induced plasma Spectrometry,J Anal At Spectrom, vol. 12, no. 9, pp. 9971004, 1997.CrossRefGoogle Scholar
Martin, M. and Cheng, M.-D., “Detection of chromium aerosol using time-resolved laser-induced plasma spectroscopy,Applied Spectroscopy, vol. 54, no. 9, pp. 12791285, 2000.CrossRefGoogle Scholar
Palleschi, V., “The first international conference on laser-induced plasma spectroscopy and applications (LIBS 2000),Spectrochimica Acta Part B Atomic Spectroscopy, vol. 56, 2001.Google Scholar
St-Onge, L., Sabsabi, M., and Cielo, P., “Analysis of solids using laser-induced plasma spectroscopy in double-pulse mode,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 53, no. 3, pp. 407415, 1998.CrossRefGoogle Scholar
Corsi, M., Cristoforetti, G., Giuffrida, M. et al., “Three-dimensional analysis of laser induced plasmas in single and double pulse configuration,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 59, no. 5, pp. 723735, 2004.CrossRefGoogle Scholar
Wen, S.-B., Mao, X., Russo, R. E. et al. “Comment on “three-dimensional analysis of laser induced plasmas in single and double pulse configuration”,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 60, no. 6, pp. 870872, 2005.CrossRefGoogle Scholar
Pershin, S., “Transformation of the optical spectrum of a laser plasma when a surface is irradiated with a double pulse,Soviet Journal of Quantum Electronics, vol. 19, no. 2, pp. 215218, 1989.CrossRefGoogle Scholar
Cristoforetti, G. and Palleschi, V., Double-Pulse Laser Ablation of Solid Targets in Ambient Gas: Mechanisms and Effects, Nova Science Publishers, Inc, Hauppauge, NY, USA, 2011.Google Scholar
Ciucci, A., Corsi, M., Palleschi, V., Rastelli, S., Salvetti, A., and Tognoni, E., “New procedure for quantitative elemental analysis by laser-induced plasma spectroscopy,Applied Spectroscopy, vol. 53, no. 8, pp. 960964, 1999.CrossRefGoogle Scholar
Ciucci, A., Palleschi, V., Rastelli, S., Salvetti, A., Singh, D. P., and Tognoni, E. C. F.-L. I. P. S., “CF-LIPS: a new approach to LIPS spectra analysis,Laser and Particle Beams, vol. 17, no. 4, pp. 793797, 1999.CrossRefGoogle Scholar
Gornushkin, I. B. B., Anzano, J. M. M., King, L. A. A., Smith, B. W. W., Omenetto, N., and Winefordner, J. D. D., “Curve of growth methodology applied to laser-induced plasma emission spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 54, no. 3-4, pp. 491503, 1999.CrossRefGoogle Scholar
Bulajic, D., Corsi, M., Cristoforetti, G. et al., “A procedure for correcting self-absorption in calibration free-laser induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 57, no. 2, pp. 339353, 2002.CrossRefGoogle Scholar
Moon, H.-Y., Herrera, K. K., Omenetto, N., Smith, B. W., and Winefordner, J. D., “On the usefulness of a duplicating mirror to evaluate self-absorption effects in laser induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 64, no. 7, pp. 702713, 2009.CrossRefGoogle Scholar
El Sherbini, A. M., El Sherbini, T. M., Hegazy, H. et al., “Evaluation of self-absorption coefficients of aluminum emission lines in laser-induced breakdown spectroscopy measurements,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 60, no. 12, pp. 15731579, 2005.CrossRefGoogle Scholar
Amamou, H., Bois, A., Ferhat, B., Redon, R., Rossetto, B., and Matheron, P., “Correction of self-absorption spectral line and ratios of transition probabilities for homogeneous and LTE plasma,Journal of Quantitative Spectroscopy and Radiative Transfer, vol. 75, no. 6, pp. 747763, 2002.CrossRefGoogle Scholar
Griem, H. R., “Stark broadening,Advances in Atomic and Molecular Physics, vol. 11, pp. 331359, 1976.CrossRefGoogle Scholar
Gigosos, M. A., González, M. Á., and Cardeñoso, V., “Computer simulated balmer-alpha, beta and gamma Stark line profiles for non-equilibrium plasmas Diagnostics,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 58, no. 8, pp. 14891504, 2003.CrossRefGoogle Scholar
Pardini, L., Legnaioli, S., Lorenzetti, G. et al., “On the determination of plasma electron number density from Stark broadened hydrogen balmer series lines in laser-induced breakdown spectroscopy experiments,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 88, pp. 98103, 2013.CrossRefGoogle Scholar
Rezaei, F., Cristoforetti, G., Tognoni, E., Legnaioli, S., Palleschi, V., and Safi, A., “A review of the current analytical approaches for evaluating, compensating and exploiting self-absorption in laser induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 169, Article ID 105878, 2020.CrossRefGoogle Scholar
Urbina Medina, I. A., Carneiro, D. D., Rocha, S., Farias, E. E., Bredice, F. O., and Palleschi, V., “Branching ratio method for assessing optically thin conditions in laser-induced plasmas,Applied Spectroscopy, vol. 75, pp. 000370282110067000370282110780, 2021.CrossRefGoogle ScholarPubMed
Urbina, I., Carneiro, D., Rocha, S., Farias, E., Bredice, F., and Palleschi, V., “Measurement of atomic transition probabilities with laser-induced breakdown spectroscopy using the 3D Boltzmann plot method,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 154, pp. 9196, 2019.CrossRefGoogle Scholar
Aguilera, J. A., Aragón, C., and Manrique, J., “Method for measurement of transition probabilities by laser-induced breakdown spectroscopy based on CSigma graphs–application to Ca II spectral lines,Journal of Quantitative Spectroscopy and Radiative Transfer, vol. 160, pp. 1018, 2015.CrossRefGoogle Scholar
Poggialini, F., Campanella, B., Jafer, R. et al., “Determination of the Stark broadening coefficients of tantalum emission lines by time-independent extended C-sigma method,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 167, Article ID 105829, 2020.CrossRefGoogle Scholar
Bredice, F., Borges, F. O. O., Sobral, H. et al., “Measurement of Stark broadening of MnI and MnII spectral lines in plasmas used for laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 62, no. 11, pp. 12371245, 2007.CrossRefGoogle Scholar
Griem, H. R., “Validity of local thermal Equilibrium in plasma spectroscopy,Physical Review A, vol. 131, no. 3, pp. 11701176, 1963.CrossRefGoogle Scholar
Cristoforetti, G., de Giacomo, A., Dell’Aglio, M. et al., “Local thermodynamic Equilibrium in laser-induced breakdown spectroscopy: beyond the McWhirter criterion,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 65, no. 1, pp. 8695, 2010.CrossRefGoogle Scholar
Grifoni, E., Legnaioli, S., Lezzerini, M., Lorenzetti, G., Pagnotta, S., and Palleschi, V., “Extracting time-resolved information from time-integrated laser-induced breakdown spectra,Journal of Spectroscopy, vol. 2014, Article ID 849310, 5 pages, 2014.CrossRefGoogle Scholar
Winefordner, J. D., Gornushkin, I. B., Correll, T., Gibb, E., Smith, B. W., and Omenetto, N., “Comparing several atomic spectrometric methods to the super stars: special emphasis on laser induced breakdown Spectrometry, LIBS, a future super star,J Anal At Spectrom, vol. 19, no. 9, pp. 10611083, 2004.CrossRefGoogle Scholar
Hahn, D. W. and 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,Applied Spectroscopy, vol. 64, no. 12, pp. 335A336A, 2010.CrossRefGoogle ScholarPubMed
Gaudiuso, R., Melikechi, N., Abdel-Salam, Z. A. et al., “Laser-induced breakdown spectroscopy for human and animal health: a review,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 152, pp. 123148, 2019.CrossRefGoogle Scholar
Corsi, M., Cristoforetti, G., Hidalgo, M. et al., “Application of laser-induced breakdown spectroscopy technique to hair tissue mineral analysis,Applied Optics, vol. 42, no. 30, p. 6133, 2003.CrossRefGoogle ScholarPubMed
Multari, R. A., Cremers, D. A., and Bostian, M. L., “Use of laser-induced breakdown spectroscopy for the differentiation of pathogens and viruses on substrates,Applied Optics, vol. 51, no. 7, pp. B57B64, 2012.CrossRefGoogle ScholarPubMed
Multari, R. A., Cremers, D. A., Scott, T., and Kendrick, P., “Detection of pesticides and dioxins in tissue fats and rendering oils using laser-induced breakdown spectroscopy (LIBS),Journal of Agricultural and Food Chemistry, vol. 61, no. 10, pp. 23482357, 2013.CrossRefGoogle ScholarPubMed
Multari, R. A., Cremers, D. A., Dupre, J. M., and Gustafson, J. E., “The use of laser-induced breakdown spectroscopy for distinguishing between bacterial pathogen species and strains,Applied Spectroscopy, vol. 64, no. 7, pp. 750759, 2010.CrossRefGoogle ScholarPubMed
Multari, R. A., Cremers, D. A., Dupre, J. A. M., and Gustafson, J. E., “Detection of biological contaminants on foods and food surfaces using laser-induced breakdown spectroscopy (LIBS),Journal of Agricultural and Food Chemistry, vol. 61, no. 36, pp. 86878694, 2013.CrossRefGoogle ScholarPubMed
Rehse, S. J., “Biomedical applications of LIBS,Springer Series in Optical Sciences, vol. 182, pp. 457488, 2014.CrossRefGoogle Scholar
Rehse, S. J., Salimnia, H., and Miziolek, A. W., “Laser-induced breakdown spectroscopy (LIBS): an overview of recent progress and future potential for biomedical applications,Journal of Medical Engineering and Technology, vol. 36, no. 2, pp. 7789, 2012.CrossRefGoogle ScholarPubMed
Scholten, J. H., Teule, J. M., Zafiropulos, V., and Heeren, R. M. A., “Controlled laser cleaning of painted artworks using accurate beam manipulation and on-line LIBS-detection,Journal of Cultural Heritage, vol. 1, pp. S215S220, 2000.CrossRefGoogle Scholar
Klein, S., Hildenhagen, J., Dickmann, K., Stratoudaki, T., and Zafiropulos, V., “LIBS-spectroscopy for monitoring and control of the laser cleaning process of stone and medieval glass,Journal of Cultural Heritage, vol. 1, pp. S287S292, 2000.CrossRefGoogle Scholar
Anglos, D. and Detalle, V., “Cultural heritage applications of LIBS,” in Laser-Induced Breakdown Spectroscopy, pp. 531554, Springer, Berlin, Germany, 2014.CrossRefGoogle Scholar
Simileanu, M., “Libs quantitative analyses of bronze objects for cultural heritage applications,Romanian Reports in Physics, vol. 68, pp. 203209, 2016.Google Scholar
Osorio, L. M., Cabrera, L. V. P., García, M. A. A., Reyes, T. F., and Ravelo, I., “Portable LIBS system for determining the composition of multilayer structures on objects of cultural value,J Phys Conf Ser, vol. 274, Article ID 012093, 2011.CrossRefGoogle Scholar
Cristoforetti, G., Legnaioli, S., Palleschi, V., Pardini, L., Salvetti, A., and Tognoni, E. M., “A new mobile instrument for in situ standard-less LIBS analysis of cultural heritage,Proceedings of the Proceedings of SPIE The International Society for Optical Engineering, vol. 5857, 2005.Google Scholar
Arafat, A., Na’es, M., Kantarelou, V. et al., “Combined in situ micro-XRF, LIBS and SEM-EDS analysis of base metal and corrosion products for islamic copper alloyed artefacts from umm qais museum, Jordan,Journal of Cultural Heritage, vol. 14, no. 3, pp. 261269, 2013.CrossRefGoogle Scholar
Bulajic, D., Cristoforetti, G., Corsi, M. et al., “Diagnostics of high-temperature steel pipes in industrial environment by laser-induced breakdown spectroscopy technique: the LIBSGRAIN project,Spectrochimica Acta, Part B: Atomic Spectroscopy, vol. 57, no. 7, pp. 11811192, 2002.CrossRefGoogle Scholar
Noll, R., Sturm, V., Stepputat, M., Whitehouse, A., Young, J., and Evans, P., “Industrial applications of LIBS,” in Laser-Induced Breakdown Spectroscopy (LIBS), Miziolek, A. W., Palleschi, V., Schechter, I., Eds., pp. 400439, Cambridge University Press, Cambridge, UK, 2006.Google Scholar
Li, C., Feng, C. L., Oderji, H. Y., Luo, G. N., and Ding, H., “Review of LIBS application in nuclear fusion technology,Frontiers in Physiology, vol. 11, no. 6, Article ID 114214, 2016.Google Scholar
Karpate, T., Nayak, R., and Libs, S., A Potential Tool for Industrial/Agricultural Waste Water Analysis, Harvard University, Cambridge, MA, USA, 2016.Google Scholar
Popov, A. M., Colao, F., and Fantoni, R., “Spatial confinement of laser-induced plasma to enhance LIBS sensitivity for trace elements determination in soils,J Anal At Spectrom, vol. 25, no. 6, p. 837, 2010.CrossRefGoogle Scholar
Senesi, G. S., Dell’Aglio, M., Gaudiuso, R. et al., “Heavy metal concentrations in soils as determined by laser-induced breakdown spectroscopy (LIBS), with special emphasis on chromium,Environmental Research, vol. 109, no. 4, pp. 413420, 2009.CrossRefGoogle ScholarPubMed
Nunes, L. C., De Carvalho, G. G. A., Santos, D., and Krug, F. J., “Determination of Cd, Cr and Pb in phosphate fertilizers by laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 97, pp. 4248, 2014.CrossRefGoogle Scholar
Palanco, S. and Laserna, J., “Remote sensing instrument for solid samples based on open-path atomic emission Spectrometry,Review of Scientific Instruments, vol. 75, no. 6, pp. 20682074, 2004.CrossRefGoogle Scholar
Fortes, F. J. and Laserna, J. J., “The development of fieldable laser-induced breakdown spectrometer: No limits on the horizon,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 65, no. 12, pp. 975990, 2010.CrossRefGoogle Scholar
Lorenzetti, G., Legnaioli, S., Grifoni, E., Pagnotta, S., and Palleschi, V., “Laser-based continuous monitoring and resolution of steel grades in sequence casting machines,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 112, pp. 15, 2015.CrossRefGoogle Scholar
Cabalín, L. M., Delgado, T., Ruiz, J., Mier, D., and Laserna, J. J., “Stand-off laser-induced breakdown spectroscopy for steel-grade intermix detection in sequence casting operations. At-line monitoring of temporal evolution versus predicted mathematical model,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 146, pp. 93100, 2018.CrossRefGoogle Scholar
Le Drogoff, B., Margot, J., Chaker, M. et al., “Temporal characterization of femtosecond laser pulses induced plasma for spectrochemical analysis of aluminum alloys,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 56, no. 6, pp. 9871002, 2001.CrossRefGoogle Scholar
Eland, K. L., Stratis, D. N., Gold, D. M., Goode, S. R., and Angel, S. M., “Energy dependence of emission intensity and temperature in a LIBS plasma using femtosecond excitation,Applied Spectroscopy, vol. 55, no. 3, pp. 286291, 2001.CrossRefGoogle Scholar
Margetic, V., Niemax, K., and Hergenröder, R., “A study of non-linear calibration graphs for brass with femtosecond laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 56, no. 6, pp. 10031010, 2001.CrossRefGoogle Scholar
Margetic, V., Bolshov, M., Stockhaus, A., Niemax, K., and Hergenröder, R., “Depth profiling of multi-layer samples using femtosecond laser ablation,J Anal At Spectrom, vol. 16, no. 6, pp. 616621, 2001.CrossRefGoogle Scholar
Moon, Y., Han, J. H., Choi, J., Shin, S., Kim, Y.-C., and Jeong, S., “Mapping of cutaneous melanoma by femtosecond laser-induced breakdown spectroscopy,Journal of Biomedical Optics, vol. 24, no. 03, p. 1, 2018.CrossRefGoogle ScholarPubMed
Hu, M., Shi, S., Yan, M., Wu, E., and Zeng, H., “Femtosecond laser-induced breakdown spectroscopy by multidimensional plasma grating,J Anal At Spectrom, vol. 37, no. 4, pp. 841848, 2022.CrossRefGoogle Scholar
Scaffidi, J., Pender, J., Pearman, W. et al., “Dual-pulse laser-induced breakdown spectroscopy with combinations of femtosecond and nanosecond laser pulses,Applied Optics, vol. 42, no. 30, p. 6099, 2003.CrossRefGoogle ScholarPubMed
Babushok, V. I., DeLucia, F. C., Gottfried, J. L., Munson, C. A., and Miziolek, A. W., “Double pulse laser ablation and plasma: laser induced breakdown spectroscopy signal enhancement,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 61, no. 9, pp. 9991014, 2006.CrossRefGoogle Scholar
Piñon, V., Fotakis, C., Nicolas, G., and Anglos, D., “Double pulse laser-induced breakdown spectroscopy with femtosecond laser pulses,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 63, no. 10, pp. 10061010, 2008.CrossRefGoogle Scholar
Stelmaszczyk, K., Rohwetter, P., Méjean, G. et al., “Long-distance remote laser-induced breakdown spectroscopy using filamentation in air,Applied Physics Letters, vol. 85, no. 18, pp. 39773979, 2004.CrossRefGoogle Scholar
Polynkin, P., Burger, M., Jovanovic, I., Jovanovic, I., and Jovanovic, I., “Filament-induced breakdown spectroscopy with structured beams,Optics Express, vol. 28, no. 24, pp. 3681236821, 2020.Google Scholar
Hu, M., Xu, S., Yuan, S., and Zeng, H., “Breakdown spectroscopy induced by nonlinear interactions of femtosecond laser filaments and multidimensional plasma gratings,Ultrafast Science, vol. 3, 2023.CrossRefGoogle Scholar
Tzortzakis, S., Anglos, D., and Gray, D., “Ultraviolet laser filaments for remote laser-induced breakdown spectroscopy (LIBS) analysis: applications in cultural heritage monitoring,Optics Letters, vol. 31, no. 8, pp. 11391141, 2006.CrossRefGoogle ScholarPubMed
Xu, H. L., Méjean, G., Liu, W. et al., “Remote detection of similar biological materials using femtosecond filament-induced breakdown spectroscopy,Applied Physics B: Photophysics and Laser Chemistry, vol. 87, no. 1, pp. 151156, 2007.CrossRefGoogle Scholar
Abdul Kalam, S., Balaji Manasa Rao, S. V., Jayananda, M., and Venugopal Rao, S., “Standoff femtosecond filament-induced breakdown spectroscopy for classification of geological materials,J Anal At Spectrom, vol. 35, no. 12, pp. 30073020, 2020.CrossRefGoogle Scholar
Shaik, A. K., Epuru, N. R., Syed, H. et al. “Femtosecond laser induced breakdown spectroscopy based standoff detection of explosives and discrimination using principal component analysis,Optics Express, vol. 26, no. 7, pp. 80698083, 2018.CrossRefGoogle ScholarPubMed
Shaik, A. K. and Soma, V. R., “Standoff discrimination and trace detection of explosive molecules using femtosecond filament induced breakdown spectroscopy combined with silver nanoparticles,OSA Continuum, vol. 2, no. 3, pp. 554562, 2019.CrossRefGoogle Scholar
González, R., Lucena, P., Tobaria, L. M., and Laserna, J. J., “Standoff LIBS detection of explosive residues behind a barrier,J Anal At Spectrom, vol. 24, no. 8, p. 1123, 2009.CrossRefGoogle Scholar
Lucena, P., Doña, A., Tobaria, L. M., and Laserna, J. J., “New challenges and insights in the detection and spectral identification of organic explosives by laser induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 66, no. 1, pp. 1220, 2011.CrossRefGoogle Scholar
DeLucia, F. C., Samuels, A. C., Harmon, R. S. et al., “Laser-induced breakdown spectroscopy (LIBS): a promising versatile chemical sensor technology for hazardous material detection,IEEE Sensors Journal, vol. 5, no. 4, pp. 681689, 2005.CrossRefGoogle Scholar
Gottfried, J. L., De Lucia, F. C., Munson, C. A., and Miziolek, A. W., “Standoff detection of chemical and biological threats using laser-induced breakdown spectroscopy,Applied Spectroscopy, vol. 62, no. 4, pp. 353363, 2008.CrossRefGoogle ScholarPubMed
Knight, A. K., Scherbarth, N. L., Cremers, D. A., and Ferris, M. J., “Characterization of laser-induced breakdown spectroscopy (LIBS) for application to space exploration,Applied Spectroscopy, vol. 54, no. 3, pp. 331340, 2000.CrossRefGoogle Scholar
Wiens, R. C., Maurice, S., Barraclough, B. et al., “The ChemCam instrument suite on the Mars science laboratory (MSL) rover: body unit and combined system tests,Space Science Reviews, vol. 170, no. 1-4, pp. 167227, 2012.CrossRefGoogle Scholar
Wiens, R. C., Maurice, S., Robinson, S. H. et al., “The SuperCam instrument suite on the NASA Mars 2020 rover: body unit and combined system tests,Space Science Reviews, vol. 217, pp. 187, 2021.CrossRefGoogle ScholarPubMed
Xu, W., Liu, X., Yan, Z. et al., “The MarSCoDe instrument suite on the Mars rover of China’s tianwen-1 mission,Space Science Reviews, vol. 217, pp. 158, 2021.CrossRefGoogle Scholar
Afgan, M. S., Hou, Z., and Wang, Z., “Quantitative analysis of common elements in steel using a handheld μ-LIBS instrument,J Anal At Spectrom, vol. 32, no. 10, pp. 19051915, 2017.CrossRefGoogle Scholar
Poggialini, F., Campanella, B., Legnaioli, S., Pagnotta, S., Raneri, S., and Palleschi, V., “Improvement of the performances of a commercial hand-held laser-induced breakdown spectroscopy instrument for steel analysis using multiple artificial neural Networks,Review of Scientific Instruments, vol. 91, no. 7, Article ID 073111, 2020.CrossRefGoogle ScholarPubMed
Garlea, E., Bennett, B. N., Martin, M. Z., Bridges, R. L., Powell, G. L., and Leckey, J. H., “Novel use of a hand-held laser induced breakdown spectroscopy instrument to monitor hydride corrosion in uranium,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 159, Article ID 105651, 2019.CrossRefGoogle Scholar
VerMeulen, H., Clausen, J., Mosell, A., Morgan, M., Messan, K., and Beal, S., “Application of laser induced breakdown (LIBS) for environmental, chemical, and biological sensing,Advanced Environmental, Chemical, and Biological Sensing Technologies XV, pp. 6890, Article ID 11007, 2019.Google Scholar
Senesi, G. S., Harmon, R. S., and Hark, R. R., “Field-portable and handheld laser-induced breakdown spectroscopy: historical review, current status and future prospects,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 175, Article ID 106013, 2021.CrossRefGoogle Scholar
Zhou, L., Luo, Y., Zheng, Y. Q., and Huang, C. G., “Field determination of chromium, nickel, manganese, copper and silicon in steel product by hand-held laser induced breakdown spectrometer,Yejin Fenxi/Metallurgical Analysis, vol. 39, pp. 812, 2019.Google Scholar
Sattmann, R., Mönch, I., Krause, H. et al., “Laser-induced breakdown spectroscopy for polymer identification,Applied Spectroscopy, vol. 52, no. 3, pp. 456461, 1998.CrossRefGoogle Scholar
Moncayo, S., Duponchel, L., Mousavipak, N. et al., “Exploration of megapixel hyperspectral LIBS images using principal component analysis,J Anal At Spectrom, vol. 33, no. 2, pp. 210220, 2018.CrossRefGoogle Scholar
Pagnotta, S., Grifoni, E., Legnaioli, S., Lezzerini, M., Lorenzetti, G., and Palleschi, V., “Comparison of brass alloys composition by laser-induced breakdown spectroscopy and self-organizing maps,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 103-104, pp. 7075, 2015.CrossRefGoogle Scholar
Dingari, N. C., Barman, I., Myakalwar, A. K., Tewari, S. P., and Kumar Gundawar, M., “Incorporation of Support vector machines in the LIBS toolbox for sensitive and robust classification amidst unexpected sample and system variability,Analytical Chemistry, vol. 84, no. 6, pp. 26862694, 2012.CrossRefGoogle ScholarPubMed
Grifoni, E., Legnaioli, S., Lorenzetti, G., Pagnotta, S., and Palleschi, V., “Application of Graph theory to unsupervised classification of materials by laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 118, pp. 4044, 2016.CrossRefGoogle Scholar
Sheng, L., Zhang, T., Niu, G. et al., “Classification of iron ores by laser-induced breakdown spectroscopy (LIBS) combined with random forest (RF),J Anal At Spectrom, vol. 30, no. 2, pp. 453458, 2015.CrossRefGoogle Scholar
Pokrajac, D., Vance, T., Lazarević, A. et al., “Performance of multilayer perceptrons for classification of LIBS protein spectra,” in Proceedings of the 10th Symposium on Neural Network Applications in Electrical Engineering, pp. 171174, Belgrade, Serbia, September 2010.Google Scholar
Yaroshchyk, P., Death, D. L., and Spencer, S. J., “Comparison of principal components regression, partial Least Squares regression, multi-block partial Least Squares regression, and serial partial Least Squares regression algorithms for the analysis of Fe in iron ore using LIBS,J. Anal. At. Spectrom, vol. 27, no. 1, pp. 9298, 2012.CrossRefGoogle Scholar
Sirven, J. B., Bousquet, B., Canioni, L. et al., “Qualitative and quantitative investigation of chromium-polluted soils by laser-induced breakdown spectroscopy combined with neural Networks analysis,Analytical and Bioanalytical Chemistry, vol. 385, no. 2, pp. 256262, 2006.CrossRefGoogle ScholarPubMed
Duchene, S., Detalle, V., Bruder, R., and Sirven, J., “Chemometrics and laser induced breakdown spectroscopy (LIBS) analyses for identification of wall paintings pigments,Current Analytical Chemistry, vol. 6, no. 1, pp. 6065, 2010.CrossRefGoogle Scholar
Legnaioli, S., Campanella, B., Pagnotta, S., Poggialini, F., and Palleschi, V., “Determination of ash content of coal by laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 155, pp. 123126, 2019.CrossRefGoogle Scholar
Sirven, J. B., Sallé, B., Mauchien, P., Lacour, J. L., Maurice, S., and Manhès, G., “Feasibility study of rock identification at the surface of Mars by remote laser-induced breakdown spectroscopy and three chemometric methods,J Anal At Spectrom, vol. 22, no. 12, pp. 14711480, 2007.CrossRefGoogle Scholar
Sancey, L., Motto-Ros, V., Busser, B. et al., “Laser Spectrometry for multi-elemental imaging of biological tissues,Scientific Reports, vol. 4, no. 1, p. 6065, 2014.CrossRefGoogle ScholarPubMed
Busser, B., Moncayo, S., Coll, J. L., Sancey, L., and Motto-Ros, V., “Elemental imaging using laser-induced breakdown spectroscopy: a new and promising approach for biological and medical applications,Coordination Chemistry Reviews, vol. 358, pp. 7079, 2018.CrossRefGoogle Scholar
Jolivet, L., Leprince, M., Moncayo, S., Sorbier, L., Lienemann, C.-P., and Motto-Ros, V., “Review of the recent advances and applications of LIBS-based imaging,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 151, pp. 4153, 2019.CrossRefGoogle Scholar
Safi, A., Campanella, B., Grifoni, E. et al., “Multivariate calibration in laser-induced breakdown spectroscopy quantitative analysis: the dangers of a ‘black box’ approach and how to avoid them,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 144, pp. 4654, 2018.CrossRefGoogle Scholar
Képeš, E., Vrábel, J., Adamovsky, O. et al., “Interpreting Support vector machines applied in laser-induced breakdown spectroscopy,Analytica Chimica Acta, vol. 1192, Article ID 339352, 2022.CrossRefGoogle ScholarPubMed
Palleschi, V., Chemometrics and Numerical Methods in LIBS, John Wiley, Hoboken, NY, USA, 2022.CrossRefGoogle Scholar
Russo, R. E., Bol’šhakov, A. A., Mao, X., McKay, C. P., Perry, D. L., and Sorkhabi, O., “Laser ablation molecular isotopic Spectrometry,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 66, no. 2, pp. 99104, 2011.CrossRefGoogle Scholar
Bol’šhakov, A. A., Mao, X., González, J. J., and Russo, R. E., “Laser ablation molecular isotopic Spectrometry (LAMIS): current state of the art,J Anal At Spectrom, vol. 31, no. 1, pp. 119134, 2016.CrossRefGoogle Scholar
de Giacomo, A., Gaudiuso, R., Koral, C., Dell’Aglio, M., and de Pascale, O., “Nanoparticle-enhanced laser-induced breakdown spectroscopy of metallic samples,Analytical Chemistry, vol. 85, no. 21, pp. 1018010187, 2013.CrossRefGoogle ScholarPubMed
de Giacomo, A., Salajkova, Z., and Dell’aglio, M., “A quantum chemistry approach based on the analogy with π-system in polymers for a rapid estimation of the resonance wavelength of nanoparticle systems,Nanomaterials, vol. 9, no. 7, pp. 9291016, 2019.CrossRefGoogle ScholarPubMed
Poggialini, F., Campanella, B., Legnaioli, S., Pagnotta, S., and Palleschi, V., “Investigating double pulse nanoparticle-enhanced laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 167, Article ID 105845, 2020.CrossRefGoogle Scholar
Koral, C., Dell’Aglio, M., Gaudiuso, R., Alrifai, R., Torelli, M., and de Giacomo, A., “Nanoparticle-enhanced laser induced breakdown spectroscopy for the noninvasive analysis of transparent samples and gemstones,Talanta, vol. 182, pp. 253258, 2018.CrossRefGoogle ScholarPubMed
Fortes, F. J., Fernández-Bravo, A., and Javier Laserna, J., “Chemical characterization of single micro- and nano-particles by optical catapulting-optical trapping-laser-induced breakdown spectroscopy,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 100, pp. 7885, 2014.CrossRefGoogle Scholar
Purohit, P., Fortes, F. J., and Laserna, J. J., “Spectral identification in the attogram regime through laser-induced emission of single optically trapped nanoparticles in air,Angewandte Chemie International Edition, vol. 56, no. 45, pp. 1417814182, 2017.CrossRefGoogle ScholarPubMed
Purohit, P., Fortes, F. J., and Laserna, J. J., “Subfemtogram simultaneous elemental detection in multicomponent nanomatrices using laser-induced plasma emission spectroscopy within atmospheric pressure optical traps,Analytical Chemistry, vol. 91, no. 11, pp. 74447449, 2019.CrossRefGoogle ScholarPubMed
Purohit, P., Fortes, F. J., and Laserna, J. J., “Atomization efficiency and photon yield in laser-induced breakdown spectroscopy analysis of single nanoparticles in an optical trap,Spectrochimica Acta Part B: Atomic Spectroscopy, vol. 130, pp. 7581, 2017.CrossRefGoogle Scholar
Figure 0

Figure 1: A typical LIBS setup.

Figure 1

Figure 2: A typical broadband LIBS spectrum (steel).

Figure 2

Figure 3: Schematic representation of the self-absorption effect, showing the balance between spontaneous emission and absorption (stimulated emission is normally negligible in LIBS plasmas).