MASARYK UNIVERSITY FACULTY OF SCIENCE DEPARTMENT OF BIOCHEMISTRY Advanced Immunochemical Biosensors and Assays: From LabelFree to Single-Molecule Detection Habilitation Thesis Zdeněk Farka Brno 2022 Abstract The ability of rapid detection of low analyte concentrations, in particular of biomarkers, microorganisms and their products, or pharmaceuticals, is of fundamental importance in many fields, including clinical diagnostics, food control, and environmental screening. Immunochemical biosensors and assays combine the excellent selectivity provided by antibodies with highly sensitive detection based on various readout techniques. This habitation thesis presents a commented summary of 22 scientific papers focused on advanced immunoanalytical techniques, to which I have contributed as a corresponding author, first author, or co-author. After introducing the field of immunosensing, the thesis starts with labelfree biosensors and continues through catalytic and luminescent labels to the detection by laserinduced breakdown spectroscopy. Numerous assays were developed for a wide range of analytes, starting from small molecules (pharmaceuticals, mycotoxins), through proteins (disease biomarkers), to bacteria (Salmonella, honeybee pathogens). The research was focused not only on testing new methodologies but also on the practical applicability of the sensors, as represented by a large focus on the analysis of representative real samples. Acknowledgments First of all, I would like to express my sincere thanks to all my students, colleagues, and collaborators for any contribution to our research summarized in this thesis. I want to thank all of my past supervisors and mentors who contributed to my scientific education. In particular, I thank Petr Skládal, who allowed me to join his research group in the first year of my bachelor studies, which started my scientific career. Throughout my studies, he shared his knowledge and provided all the instrumentation, allowing me to become proficient in the field of biosensing. I thank Hans-Heiner Gorris, who always willingly hosted me on my numerous stays at the University of Regensburg and who introduced me to the application of photon-upconversion nanoparticles. I also thank Niko Hildebrandt for hosting me at the University of Rouen and allowing me to gain experience in the detection of nucleic acids. I wish to thank all my students and colleagues from the Masaryk University. Special thanks go to Matěj Pastucha, to whom I could always rely on. I am grateful to David Kovář, Veronika Poláchová, Eliška Macháčová, Radka Obořilová, Dorota Sklenárová, Ekaterina Makhneva, Pavlína Botíková, Květa Mertová, Zuzana Mikušová, Tomáš Juřík, Karel Lacina, Veronika Křešťáková, Karel Novotný, Libuše Trnková, Lenka Zajíčková, Hana Šimečková, Ivana Mašlaňová, and Roman Pantůček for their inspiring thoughts, great effort during our common work, and for creative and friendly environment that they created. I am grateful for the perfect operation of core-facilities within CEITEC Masaryk University, especially Nanobiotechnology (Jan Přibyl, Šimon Klimovič), Cryo-Electron Microscopy and Tomography (Vít Vykoukal, Miroslav Peterek), and Proteomics (Zbyněk Zdráhal, Ondrej Šedo, Hana Konečná, and Kamil Mikulášek). I would also like to thank the colleagues and collaborators from other institutes, especially the Institute of Analytical Chemistry of the AS CR (Antonín Hlaváček, František Foret), Brno University of Technology (Pavel Pořízka, Pavlína Modlitbová), University of Regensburg (Julian C. Brandmeier, Simone Rink), University of Turku (Riikka Petlomaa, Tero Soukka), Institute of Macromolecular Chemistry of the AS CR (Uliana Kostiv, Daniel Horák), University of Rouen (Mariia Dekaliuk), Leibniz Institute for Plasma Science and Technology (Katja Fricke), Institute of Physics of the AS CR (Ivana Víšová, Hana Lísalová), and Adam Mickiewicz University in Poznań (Natalia Jurga). A great deal of thanks goes to Matthias J. Mickert, my colleague, friend, and an industrial partner. Together, we have done an incredible amount of scientific work, connected with an even more incredible amount of fun, which I will never forget. Finally, I am immensely grateful to my parents for their limitless support. Table of Contents 1 Commentary to Habilitation Thesis...................................................................................6 2 Introduction (Papers I and II)...........................................................................................11 3 Label-Free Biosensing .....................................................................................................14 3.1 Electrochemical Impedance Spectroscopy Biosensing of Salmonella (Paper III)....14 3.2 Quartz Crystal Microbalance Biosensor for Aerosolized Bacteria (Paper IV) .........16 3.3 Plasma-Polymerized Surfaces for SPR Biosensing (Papers V–VII).........................19 4 Catalytic Labels ...............................................................................................................23 4.1 Amperometric Detection of M. plutonius (Paper VIII).............................................23 4.2 Enzymatic Precipitation-Enhanced SPR Detection of Salmonella (Paper IX) .........24 4.3 Nanozyme-Linked Immunosorbent Assay (Paper X)...............................................27 5 Photon-Upconversion Nanoparticles ...............................................................................30 5.1 Competitive Upconversion-Linked Immunosorbent Assays ....................................30 5.1.1 Competitive ULISA for Diclofenac (Papers XI and XII)..................................30 5.1.2 Competitive ULISA for Zearalenone (Paper XIII)............................................32 5.2 Sandwich Upconversion-Linked Immunosorbent Assays ........................................34 5.2.1 Detection of M. plutonius with BSA-Modified UCNPs (Paper XIV) ...............34 5.2.2 Preparation of PEG-Modified UCNPs and Analysis of HSA (Paper XV) ........35 5.2.3 Detection of P. larvae with PEG-Modified UCNPs (Paper XVI).....................37 5.3 Single-Molecule Upconversion-Linked Immunosorbent Assays .............................38 5.3.1 Digital ULISA for PSA (Papers XVII and XVIII) ............................................38 5.3.2 Digital ULISA for Cardiac Troponin (Paper XIX)............................................41 5.4 UCNP-Based Immunocytochemistry (Paper XX) ....................................................43 6 Laser-Induced Breakdown Spectroscopy (Papers XXI and XXII)..................................45 7 Conclusions and Outlook.................................................................................................47 References................................................................................................................................48 List of Abbreviations ...............................................................................................................60 Appendix..................................................................................................................................62 List of Publications ..............................................................................................................62 Paper I..................................................................................................................................66 Paper II...............................................................................................................................137 Paper III .............................................................................................................................166 Paper IV .............................................................................................................................174 Paper V...............................................................................................................................182 Paper VI .............................................................................................................................192 Paper VII............................................................................................................................202 Paper VIII...........................................................................................................................216 Paper IX .............................................................................................................................225 Paper X...............................................................................................................................233 Paper XI .............................................................................................................................241 Paper XII............................................................................................................................249 Paper XIII...........................................................................................................................257 Paper XIV ..........................................................................................................................267 Paper XV............................................................................................................................277 Paper XVI ..........................................................................................................................290 Paper XVII.........................................................................................................................302 Paper XVIII........................................................................................................................309 Paper XIX ..........................................................................................................................317 Paper XX............................................................................................................................327 Paper XXI ..........................................................................................................................339 Paper XXII.........................................................................................................................350 6 1 Commentary to Habilitation Thesis This habitation thesis presents a commented summary of 22 scientific papers published between 2014 and 2021, to which I have contributed as a corresponding author, first author, or co-author. All these publications are focused on immunochemical biosensors and assays; however, they are based on different sensing schemes and the detection of various analytes. After introducing the field of immunosensing, the thesis starts with label-free biosensors and continues through catalytic and luminescent labels to the detection by laser-induced breakdown spectroscopy. The research was focused not only on testing new methodologies but also on the practical applicability of the sensors, as represented by a large focus on the analysis of representative real samples. The label-free sensors especially provide rapid and straightforward analysis, making them suitable for in-field detection, especially of larger analytes. The thesis discusses the development and performance of biosensor based on electrochemical impedance spectroscopy for Salmonella and quartz crystal microbalance biosensor for aerosolized biological warfare agents. We have also focused on biosensor surface modifications by plasma-polymerized films and their application in surface plasmon resonance biosensing. The catalytic labels are beneficial due to their ability of signal enhancement. Apart from the conventional use of horseradish peroxidase in a sandwich immunoassay for European foulbrood diagnosis, the thesis demonstrates advanced approaches based on enzymaticallycatalyzed precipitation for signal enhancement in surface plasmon resonance and the catalytic Prussian blue nanoparticles as a promising alternative to enzymes. The luminescence detection was done with photon-upconversion nanoparticles, which overcome the optical background interference by the ability to be excited in the near IR region, followed by the emission in the visible range. The methods of their surface modification and conjugation with biomolecules were thoroughly studied. The conjugates were used for immunochemical detection of a wide range of analytes from small molecules, through proteins, to bacteria, demonstrating even the capabilities of single-molecule detection. Finally, laser-induced breakdown spectroscopy was introduced as a novel way of signal readout, which is not dependent on the catalytic or luminescent properties of the labels. This approach was used in the microtiter plate-based immunoassay but also as the readout method in immunocytochemical imaging. Roman numerals will be used to address the individual publications in the following text. Full articles have been reproduced in the appendix with permissions from the respective copyright holders. Asterisk denotes corresponding author. I. Farka, Z.; Juřík, T.; Kovář, D.; Trnková, L.; Skládal, P., Nanoparticle-Based Immunochemical Biosensors and Assays: Recent Advances and Challenges. Chem. Rev. 2017, 117 (15), 9973–10042. Contribution: Literature research, manuscript writing (Supervision 10%, Manuscript 30%, Research direction 30%) 7 II. Farka, Z.; Mickert, M. J.; Pastucha, M.; Mikušová, Z.; Skládal, P.; Gorris, H. H., Advances in Optical Single-Molecule Detection: En Route to Super-Sensitive Bioaffinity Assays. Angew. Chem. Int. Ed. 2020, 59 (27), 10746–10773. (Z.F. and M.J.M. contributed equally) Contribution: Outline of review, literature research, manuscript writing (Supervision 50%, Manuscript 30%, Research direction 40%) III. Farka, Z.; Juřík, T.; Pastucha, M.; Kovář, D.; Lacina, K.; Skládal, P., Rapid immunosensing of Salmonella Typhimurium using electrochemical impedance spectroscopy: the effect of sample treatment. Electroanalysis 2016, 28 (8), 1803–1809. (Z.F. and T.J. contributed equally) Contribution: Design of experiments, development and optimization of EIS immunosensor, characterization of sensing surface by AFM, data evaluation, manuscript writing (Experimental work 30%, Supervision 30%, Manuscript 50%, Research direction 40%) IV. Kovář, D.; Farka, Z.; Skládal, P., Detection of aerosolized biological agents using the piezoelectric immunosensor. Anal. Chem. 2014, 86 (17), 8680–8686. (D.K. and Z.F. contributed equally) Contribution: Development and optimization of QCM immunosensor, data evaluation, manuscript writing (Experimental work 50%, Supervision 10%, Manuscript 50%, Research direction 30%) V. Makhneva, E.; Farka, Z.; Skládal, P.; Zajíčková, L., Cyclopropylamine plasma polymer surfaces for label-free SPR and QCM immunosensing of Salmonella. Sens. Actuators B Chem. 2018, 276, 447–455. Contribution: Development and optimization of SPR and QCM immunosensors, characterization of sensing surface by AFM, data evaluation, participation in manuscript writing (Experimental work 30%, Supervision 10%, Manuscript 30%, Research direction 20%) VI. Makhneva, E.; Farka, Z.*; Pastucha, M.; Obrusník, A.; Horáčková, V.; Skládal, P.; Zajíčková, L., Maleic anhydride and acetylene plasma copolymer surfaces for SPR immunosensing. Anal. Bioanal. Chem. 2019, 411 (29), 7689–7697. Contribution: Design of experiments, development and optimization of SPR immunosensor, characterization of sensing surface by AFM, data evaluation, manuscript writing (Experimental work 20%, Supervision 50%, Manuscript 50%, Research direction 40%) VII. Makhneva, E.; Barillas, L.; Farka, Z.; Pastucha, M.; Skládal, P.; Weltmann, K. D.; Fricke, K., Functional Plasma Polymerized Surfaces for Biosensing. ACS Appl. Mater. Interfaces 2020, 20 (14), 17100–17112. Contribution: Development and optimization of SPR immunosensor, data evaluation, participation in manuscript writing (Experimental work 20%, Supervision 10%, Manuscript 20%, Research direction 20%) VIII. Mikušová, Z.; Farka, Z.*; Pastucha, M.; Poláchová, V.; Obořilová, R.; Skládal, P., Amperometric Immunosensor for Rapid Detection of Honeybee Pathogen Melissococcus plutonius. Electroanalysis 2019, 31 (10), 1969–1976. 8 Contribution: Design of experiments, preparation of immunization antigen and antibody, optimization of electrochemical immunosensor, data evaluation, manuscript writing (Experimental work 30%, Supervision 80%, Manuscript 50%, Research direction 80%) IX. Farka, Z.; Juřík, T.; Pastucha, M.; Skládal, P. Enzymatic Precipitation Enhanced Surface Plasmon Resonance Immunosensor for the Detection of Salmonella in Powdered Milk. Anal. Chem. 2016, 88 (23), 11830–11836. Contribution: Design of experiments, development and optimization of precipitation-enhanced SPR assay, characterization of precipitation reaction by AFM, data evaluation, manuscript writing (Experimental work 40%, Supervision 50%, Manuscript 50%, Research direction 50%) X. Farka, Z.*; Čunderlová, V.; Horáčková, V.; Pastucha, M.; Mikušová, Z.; Hlaváček, A.; Skládal, P., Prussian Blue Nanoparticles as a Catalytic Label in a Sandwich Nanozyme-Linked Immunosorbent Assay. Anal. Chem. 2018, 90 (3), 2348–2354. (Z.F. and V.Č. contributed equally) Contribution: Design of experiments, bioconjugation and characterization of PBNPs, development and optimization of sandwich assay, data evaluation, manuscript writing (Experimental work 40%, Supervision 60%, Manuscript 60%, Research direction 60%) XI. Hlaváček, A.; Farka, Z.; Hübner, M.; Horňáková, V.; Němeček, D.; Skládal, P.; Knopp, D.; Gorris, H. H., Competitive Upconversion-Linked Immunosorbent Assay for the Sensitive Detection of Diclofenac. Anal. Chem. 2016, 88 (11), 6011–6017. Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of competitive immunoassay, data evaluation, participation in manuscript writing (Experimental work 30%, Supervision 10%, Manuscript 30%, Research direction 20%) XII. Hlaváček, A.; Peterek, M.; Farka, Z.; Mickert, M. J.; Prechtl, L.; Knopp D.; Gorris, H. H., Rapid single-step upconversion-linked immunosorbent assay for diclofenac. Microchim. Acta 2017, 184 (10), 4159–4165. Contribution: Development and optimization of competitive immunoassay, data evaluation, participation in manuscript writing (Experimental work 20%, Supervision 10%, Manuscript 20%, Research direction 10%) XIII. Peltomaa, R.; Farka, Z.; Mickert, M. J.; Brandmeier, J. C.; Pastucha, M.; Hlaváček, A.; Martínez-Orts, M.; Canales, Á.; Skládal, P.; Benito-Peña, E.; Moreno-Bondi, M. C.; Gorris, H. H., Competitive upconversion-linked immunoassay using peptide mimetics for the detection of the mycotoxin zearalenone. Biosens. Bioelectron. 2020, 170, 112683. Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of competitive immunoassay, data evaluation, participation in manuscript writing (Experimental work 30%, Supervision 30%, Manuscript 30%, Research direction 30%) XIV. Poláchová, V.; Pastucha, M.; Mikušová, Z.; Mickert, M. J.; Hlaváček, A.; Gorris, H. H.; Skládal, P.; Farka, Z.*, Click-conjugated photon-upconversion nanoparticles in an immunoassay for honeybee pathogen Melissococcus plutonius. Nanoscale 2019, 11 (17), 8343–8351. 9 Contribution: Design of experiments, preparation of immunization antigen and antibody, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing (Experimental work 20%, Supervision 80%, Manuscript 50%, Research direction 80%) XV. Kostiv, U.; Farka, Z.; Mickert, M. J.; Gorris, H. H.; Velychkivska, N.; Pop-Georgievski, O.; Pastucha, M.; Odstrčilíková, E.; Skládal, P.; Horák, D., Versatile bioconjugation strategies of PEG-modified upconversion nanoparticles for bioanalytical applications. Biomacromolecules 2020, 21 (11), 4502–4513. (U.K. and Z.F. contributed equally) Contribution: Design of experiments, bioconjugation of UCNPs, development and optimization of sandwich immunoassay, data evaluation, participation in manuscript writing (Experimental work 30%, Supervision 40%, Manuscript 40%, Research direction 40%) XVI. Pastucha, M.; Odstrčilíková, E.; Hlaváček, A.; Brandmeier, J. C.; Vykoukal, V.; Weisová, J.; Gorris, H. H.; Skládal, P.; Farka Z.*, Upconversion-linked Immunoassay for the Diagnosis of Honeybee Disease American Foulbrood. IEEE J. Sel. Top. Quantum Electron. 2021, 27 (5), 6900311. Contribution: Design of experiments, preparation of immunization antigen and antibody, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing (Experimental work 20%, Supervision 80%, Manuscript 50%, Research direction 80%) XVII. Farka, Z.; Mickert, M. J.; Hlaváček, A.; Skládal P.; Gorris, H. H., Single Molecule Upconversion-Linked Immunosorbent Assay with Extended Dynamic Range for the Sensitive Detection of Diagnostic Biomarkers. Anal. Chem. 2017, 89 (21), 11825–11830. (Z.F. and M.J.M. contributed equally) Contribution: Design of experiments, optimization of single-particle microscope setup, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing (Experimental work 40%, Supervision 20%, Manuscript 40%, Research direction 20%) XVIII. Mickert, M. J.; Farka, Z.; Kostiv, U.; Hlaváček, A.; Horák, D.; Skládal, P.; Gorris, H. H., Measurement of Sub-femtomolar Concentrations of Prostate-Specific Antigen through Single-Molecule Counting with an Upconversion-Linked Immunosorbent Assay. Anal. Chem. 2019, 91 (15), 9435–9441. (M.J.M and Z.F. contributed equally) Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing (Experimental work 30%, Supervision 20%, Manuscript 30%, Research direction 30%) XIX. Brandmeier, J. C.; Raiko, K.; Farka, Z.*; Peltomaa, R.; Mickert, M. J.; Hlaváček, A.; Skládal, P.; Soukka, T.; Gorris, H. H., Effect of Particle Size and Surface Chemistry of PhotonUpconversion Nanoparticles on Analog and Digital Immunoassays for Cardiac Troponin. Adv. Healthc. Mater. 2021, 10 (18), 2100506. Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing 10 (Experimental work 20%, Supervision 30%, Manuscript 30%, Research direction 30%) XX. Farka, Z.*; Mickert, M. J.; Mikušová, Z.; Hlaváček, A.; Bouchalová, P.; Xu, W.; Bouchal, P.; Skládal, P.; Gorris, H. H., Surface design of photon-upconversion nanoparticles for highcontrast immunocytochemistry. Nanoscale 2020, 12 (15), 8303–8313. (Z.F. and M.J.M. contributed equally) Contribution: Design of experiments, optimization of microscope setup, bioconjugation and characterization of UCNPs, development and optimization of ICC assay, data evaluation, manuscript writing (Experimental work 40%, Supervision 40%, Manuscript 50%, Research direction 40%) XXI. Modlitbová, P.; Farka, Z.; Pastucha, M.; Pořízka, P.; Novotný, K.; Skládal, P.; Kaiser, J., Laser-induced breakdown spectroscopy as a novel readout method for nanoparticle-based immunoassays. Microchim. Acta 2019, 186, 629. Contribution: Design of experiments, development and optimization of sandwich immunoassay, data evaluation, participation in manuscript writing (Experimental work 40%, Supervision 30%, Manuscript 30%, Research direction 30%) XXII. Pořízka, P.; Vytisková, K.; Obořilová, R.; Pastucha, M.; Gábriš, I.; Brandmeier, J. C.; Modlitbová, P.; Gorris, H. H.; Novotný, K.; Skládal, P.; Kaiser, J.; Farka, Z., Laser-Induced Breakdown Spectroscopy as a Readout Method for Immunocytochemistry with Upconversion Nanoparticles. Microchim. Acta 2021, 188, 147. Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of ICC assay, data evaluation, manuscript writing (Experimental work 20%, Supervision 40%, Manuscript 50%, Research direction 50%) 11 2 Introduction (Papers I and II) The capability to rapidly detect small analyte concentrations, particularly of low-abundance biomarkers, is critical for diagnosing diseases in their early stages. The majority of bioaffinity methods are employing antibodies;1, 2 however, also aptamers3 and molecularly imprinted polymers (MIPs)4 can be used for specific capture of the target analyte. Antibodies with high affinity can be prepared against generally any analyte molecule. The limit of affinity, represented by a binding constant, is approximately 1010 M−1 ,5 which is worse than 1014 M−1 in the case of (strept)avidin-biotin interaction.6 Due to the relatively large size of antibodies, single binding site antibodies (camelids) are also attracting attention recently.7 Aptamers are beneficial due to their easier large-scale production, and MIPs excel in chemical stability. MIPs are particularly suitable for detecting small molecules with a rigid structure. However, they are less suitable for the detection of bigger, more flexible analytes, as proteins. Two approaches can be used for the detection of binding events: (i) Label-free assays exploit the possibility to generate a signal directly upon analyte binding to the detection element. (ii) The so-called sandwich format is based on binding a second affinity reagent bearing a label that provides signal generation. Both approaches can be carried out in a competitive (or inhibition) mode, based on competition of immunoreagents for a limited amount of binding sites, resulting in lower signals for higher analyte concentrations. The first immunoassays were based on radioactive labels;8 however, these were soon replaced by enzymes, which allow higher safety. Furthermore, a single enzyme molecule can generate a high number of measurable product molecules (signal amplification step). The enzyme-linked immunosorbent assay (ELISA) is nowadays considered as a method of choice for quantitative analysis of various analytes, from clinical diagnosis, through food control, up to environmental protection. Throughout the past 60 years, the progress in immunoassays was primarily focused on enhancing sensitivity, specificity, and reproducibility. Even though ELISA can detect picomolar analyte concentrations, even higher sensitivities are necessary. Only a few toxin molecules can be harmful,3 individual infectious viruses or bacteria can initiate a disease,9 and trace cancer biomarker quantities indicate the onset of a malignant transformation.10 Furthermore, developing immunoassays with higher sensitivity is critical to allow discovering new biomarkers, which cannot be analyzed using the current methodology.11, 12 The conventional ELISA is carried out the laboratory conditions and is based on relatively long incubation times and several washing steps. Therefore, the recent development in the field aims also at faster analysis, with higher throughput and smaller sample consumption. Such assays allow on-line analysis, e.g., at the bedside for clinical tests,13 or in the field for environmental or military applications. Such methods are often referred to as pointof-care (PoC) tests.14 It is preferred to use samples that require minimal invasiveness during their collection, such as urine or saliva. Furthermore, assays without washing steps are desired. The most famous PoC test based on antibodies is the home pregnancy test, the representative of lateral flow immunoassays (LFIAs), which were developed in the 1980s.15 The user-friendly operation, along with the possibility to provide reliable results, is necessary to allow the PoC 12 test to be used in predictive, preventive, personalized, and participatory medicine, commonly termed P4 medicine.16 The group of label-based bioaffinity assays can be further divided according to the used detection label (Figure 1). (i) Enzymes represent the most widespread approach, whereas (ii) fluorescent molecular labels are generally easier, without the requirement of the product generation step. However, the fluorescence immunoassays (FIAs) are typically limited by the fluorescence of the background. Furthermore, fluorescence readout was adapted in homogeneous assays based on fluorescence polarization and methods based on signal amplification (e.g., based on Immuno PCR). Significant progress regarding the limitation of background fluorescence was achieved by the development of time-resolved (TR) approaches that exploit lanthanide-based labels with long lifetimes (μs) compared to small fluorophores (ns).17 The time-gated approach is based on the luminescence excitation, which is not directly followed by the signal acquisition, but the measurement is delayed by a few μs, so the autofluorescence signal decays, and only specific lanthanide signals are measured. The dissociation-enhanced lanthanide fluorescent immunoassay (DELFIA) is currently the most widespread commercially available TR approach.18 Figure 1: History of the development of label-based immunoassays with optical readout. Radioisotopes were replaced by enzymes, fluorescent molecules, and nanoparticles. By choosing a suitable readout method, all these labels can be used for the measurement at the single‐molecule level. Reprinted from Paper II under the permission of Creative Commons Attribution-NonCommercial 4.0 International License. 13 As an alternative label type (iii), various kinds of nanoparticles (NPs) are gaining increasing popularity (Figure 1).2, 19 Gold nanoparticles (Au NPs) are widely used for the readout of LFIAs. Because of their plasmonic properties, Au NPs exhibit strong absorption and scattering of light, which makes them easily visible by the eye, and the color-based readout is possible without the need for sophisticated instrumentation. Apart from the plasmonic NPs, many other types of NPs and nanocomposites are being used for optical detection. Quantum dots (QDs) represent an alternative to organic fluorophores due to their better photostability and higher brightness, which is an essential aspect in immunoassay readout. Photonupconversion nanoparticles (UCNPs) are another kind of luminescent labels, which allow excitation by near-infrared light, followed by the emission of light with a shorter wavelength. This anti-Stokes emission avoids autofluorescence and light scattering, leading to detection without optical background interference.20 Nanocontainers (e.g., liposomes) can be packed by many fluorescent molecules to generate strong signals. Compared to the enzyme-based labels, which produce the fluorophores in situ from the non-fluorescent substrate, the fluorophores encapsulated in nanocontainers can be released on demand, limiting the self-quenching inside the confined environment.21 Furthermore, various mixed detection schemes, e.g., electrochemiluminescence, can be employed to generate strong signals without background. Overall, the various detection schemes present different advantages and disadvantages in terms of sensitivity, analysis time, miniaturization potential, etc. Therefore, a suitable method has to be chosen not only concerning the target analyte but also for the intended application and user base. 14 3 Label-Free Biosensing 3.1 Electrochemical Impedance Spectroscopy Biosensing of Salmonella (Paper III) Electrochemical immunosensors are receiving increasing focus because they can combine highly sensitive measurements with portability and low cost. Electrochemical impedance spectroscopy (EIS) is a technique, which allows the measurement of small changes in the interface between the electrode and solution. EIS provides a fast response in combination with high sensitivity and potential for real-time measurement and miniaturization.22 When used in biosensors, EIS provides insight into the individual immobilized layers and coating on the electrode in general. The EIS measurement is based on applying a low-amplitude sinusoidal potential (or current) through the electrochemical cell with the electrolyte solution, typically ferro/ferricyanide. The output current (or potential) is then measured over a range of frequencies by a potentiostat, allowing the calculation of the impedance parameters. Compared to the other electrochemical techniques, including cyclic voltammetry, the applied potential is smaller, preventing the undesired influence on biomolecular layers and binding processes.23 The biosensors based on EIS typically employ antibodies immobilized in the electrode, directly capturing the target analyte. The accumulated mass hinders the electron transfer; this is evaluated as the increase of impedance. This allows operation in label-free mode, providing robust and straightforward analysis. The label-free EIS can be used for rapid analysis of pathogens within small sample amounts. Our research focused on Salmonella enterica serovar Typhimurium, a non-typhoidal strain, which is one of the leading causes of gastrointestinal diseases. Salmonella is a gramnegative bacterium, which can cause diarrhea, fever, and abdominal spasm within 12 to 72 h after infection. In the worst scenario, Salmonella can enter blood, bones, brain, or nervous system, which can cause even lethal infections.24 The infection is typically caused by consuming contaminated food.25 Salmonella can be present in raw animal food products, including meat, eggs, and unpasteurized dairy products. There are globally 94 million cases of gastroenteritis and 155,000 deaths attributed to Salmonella each year.26 According to the Centers for Disease Control and Prevention (CDC), there are 1.2 million illnesses and 450 deaths per year in the United States caused by non-typhoidal Salmonella strains.26 This highlights the danger of Salmonella to human health and the importance of developing devices that can allow rapid and sensitive Salmonella detection. The standard approaches allowing the detection of Salmonella include traditional cultivation-based methods, ELISA, and polymerase chain reaction (PCR). The cultivationbased approaches are considered a gold standard for Salmonella detection because of the high sensitivity and selectivity. However, the long analysis times (5–7 days) with labor-intensive procedures do not allow using cultivation for rapid screening purposes.27 The ELISA can provide sensitive results generally within 24 h.28 Usually, a time-consuming pre-enrichment step is necessary to increase the bacteria count in the samples.29 PCR overcomes the sensitivity and analysis times on conventional methods; however, it requires expensive instrumentation 15 and trained personnel to carry out the analysis.30 Furthermore, the abovementioned methods are typically limited only to laboratory conditions and do not allow PoC operation. We have developed an EIS approach for the detection of Salmonella Typhimurium, based on a simple, easy to fabricate, and low-cost immunosensor. The screen-printed electrodes (SPEs) were modified by a self-assembled monolayer of cysteamine, followed by binding of glutaraldehyde and specific antibody (Figure 2). The increase of impedance after incubation with the sample revealed the presence of bacteria. Different sample treatment methods (viable bacteria and combinations of heat-treatment and sonication) were tested to find the optimal way of sample preparation regarding the specificity of the chosen antibody. The achieved results have shown that the antibody did not exhibit the necessary affinity towards native Salmonella. After the heattreatment (80 ºC, 40 min), the affinity of the antibody to the microbe increased significantly. This allowed reaching a limit of detection (LOD) of 7×104 CFU/mL with a wide linear range up to 108 CFU/mL. The treatment by heat does not present significant technical difficulty for the real sample analyses. Furthermore, it can even be beneficial to work with the killed or weakened bacteria due to the reduced level of pathogenicity. To improve the sensitivity further, the Salmonella cells were disrupted by sonication. The sonicated heat-treated sample has shown a higher level of specific binding than whole cells, resulting in an LOD of 1×103 CFU/mL; the linear range was up to 108 CFU/mL. The total analysis time (including the incubation of the sensor with the sample) was 20 min. The treatment by sonication is also not a technical problem for practical analysis. Even though additional instrumentation is required, sonication is beneficial because it is less timeconsuming than heat-treatment. The presence of bacteria on the sensor surface was confirmed by atomic force microscopy (AFM; Figure 3). It was shown that the sensor captured a large number of cell fragments, with the size and structure corresponding to heat-treated and sonicated bacteria adsorbed on the glass. In the case of the cross-reactivity with E. coli K-12, only negligible increases of impedance were observed, confirming the excellent selectivity of the method. Because both Salmonella and E. coli are relatively similar gram-negative bacteria, a low level of crossreactivity can also be expected in the case of more phylogenetically distant bacteria.31 Figure 2: Scheme of antibody immobilization (blocking by BSA not shown) and the design of the SPE electrode. Reprinted from Paper III with permission. Copyright 2016 Wiley-VCH. 16 The practical applicability of the sensor was demonstrated by the analysis of real samples of milk spiked with Salmonella. The electrode had to be washed thoroughly after the incubation with complex samples; insufficient washing was connected with the increase of non-specific signals. The detection capabilities in the case of complex samples decreased slightly compared to the detection in the buffer, resulting in the LOD of 9×103 CFU/mL and linear range up to 107 CFU/mL. These results are comparable to the infection dose of Salmonella32 and highlight the potential of the developed method. Figure 3: (A) EIS response of the SPE immunosensor to different concentrations of heat-treated and sonicated Salmonella; (B) AFM scan of the electrode after binding (error signal); (C) negative control (no antibody on the surface); (D) AFM scan of the blank electrode. Reprinted from Paper III with permission. Copyright 2016 Wiley-VCH. 3.2 Quartz Crystal Microbalance Biosensor for Aerosolized Bacteria (Paper IV) In 1959, Sauerbrey described the dependence between the resonant frequency of quartz sensor and the mass accumulated on the sensor surface, which led to the development of the microgravimetric biosensing approach – quartz crystal microbalance (QCM). QCM allows detecting binding events based on the measurement of changes of frequency of the quartz 17 resonator.33 The great advantage of QCM is the ability to provide the results in real-time, finding applications in monitoring of surface modifications, adsorption/desorption effects, and binding interactions. When the analyte is captured on the surface, it increases the loaded mass, which can be directly estimated from the decrease of the resonance frequency. We employed the QCM biosensor for the detection of bioaerosols. Airborne microorganisms (bacteria, viruses, fungi, etc.) are an integral part of the environment. Bacteria can be spread to the air by natural as well as anthropogenic sources; the misuse of pathogens can result in targeted biological attacks.34, 35 Modern history has shown that the potential of abuse of biological warfare agents is very high.36 The contamination of outdoor and indoor air by bacteria can infect a large number of people within a short timeframe.37 The detection of bioaerosols is difficult due to the low levels of target bacteria, combined with a potentially complex sample matrix containing pollen grains, mold, fungi, dust, and ubiquitous microbial organisms.38 Further interferences might be caused by industrial products, which are emitted into the atmosphere in large quantities. The low visibility and lack of odor present additional challenges for sampling and analysis. In contrast to chemical agents, also the delayed effect of biological agents has to be considered, leading to a potentially large number of casualties before protective steps are taken.36 The bioaerosols can contribute to the spread of epidemics and pandemics in places with high population densities. In such cases, rapid detection and identification of the pathogen are critical to allow early treatment. Therefore, the specific detection of bacteria in the air is of particular importance also during peacetime. The bioaerosol analysis consists of two critical parts: collection of the bioaerosol and detection of the bacteria. The systems for the collection of bioaerosols (samplers) have been extensively developed in recent years;39 their function can be based on various physical principles.40 The two main approaches for the detection are based on the analysis of general biological compounds in the air or specific detection of target bacteria. The non-specific detection is typically based on optical methods, exploiting the fluorescence of the molecules present in biological systems (ATP, NADH, tryptophan, etc.).41, 42 These systems, however, cannot differentiate between pathogenic and non-pathogenic bacteria, and the presence of atmospheric pollutants with fluorescence properties can interfere with the analysis.43 Furthermore, this principle is only useful for living bacteria, and it does to allow detection of spores, as they exhibit only low ATP levels.44 The specific detection of bacteria typically exploits culture-based techniques, PCR,45, 46 Raman spectroscopy, or mass spectrometry.47 The culture-based methods are time-consuming and usually require collection of samples for subsequent laboratory analysis. The PCR is highly sensitive; however, various interfering substances can be present in the complex samples, which hinders the analysis.48 Therefore, PCR is often combined with other methods to prevent false-positive results.49 In the case of MS, the widespread use for in-field detection is limited mainly by the requirement of vacuum and extensive instrumentation in general. We have employed the immunosensor technology to overcome these disadvantages, aiming at quick-response, real-time, and on-site detection. First, a bioaerosol chamber was 18 constructed for safe and controlled dissemination of biological agents and applied for experiments with model bacterial aerosols of E. coli (Figure 4A). The samples were disseminated using a piezoelectric humidifier, distribution of bioaerosol inside the chamber was achieved using three 12-cm fans. The disseminated bacteria were collected and preconcentrated using the wetted-wall cyclone SASS 2300; the analysis was done using the on-line linked QCM immunosensor. The measurement was fully automated; the flow system was used for the on-line delivery of the samples from the cyclone to the QCM, allowing to perform one detection cycle within 16 min. The achieved LOD of E. coli in the bioaerosol was 104 CFU/L of air, based on the amounts of the disseminated microbe (Figure 4B). The whole experiment, including sample collection, detection, sensor surface regeneration, and bioaerosol chamber ventilation, took 40 min. The reference experiments based on cultivation showed that the disseminated amount of E. coli was reduced, probably because of the surface adsorption, desiccation, and mechanical stress caused by the cyclone. The great benefit of the developed detection system is the possibility of entirely remote operation; the users do not come into contact with potentially dangerous microorganisms during the experiments. Furthermore, the system is fully portable (desk-top size) and requires only power and ethernet connections. The achieved results proved the suitability of the developed QCM immunosensor combined with cyclone sampling to detect aerosolized bacteria. In the future, multiplexed detection based on monoclonal antibodies specific to different pathogens can allow comprehensive screening of the air quality in a reasonably short time. The system based on single-step analysis provides the necessary simplicity, robustness, and reliability, making it a suitable option for in-field applications. Figure 4: (A) The constructed bioaerosol chamber. (B) Calibration curve for the QCM detection of E. coli in bioaerosol. The inset shows the binding curves of samples captured by the cyclone. Adapted from Paper IV with permission. Copyright 2014 American Chemical Society. 19 3.3 Plasma-Polymerized Surfaces for SPR Biosensing (Papers V–VII) Surface plasmon resonance (SPR) biosensors are based on the oscillations of electrons in conduction bands of metal films (typically gold) upon the excitation by light. This effect strongly depends on the dielectric constant of the environment50 and can be exploited in immunosensing because the biological interactions lead to the changes in oscillation frequency. The measurement can be based on changes of intensity, angle, refractive index, or phase of the reflected light.51 The SPR biosensors can be divided into two main groups: (i) propagating SPRs (PSPRs; also simply referred to as SPRs) and localized SPRs (LSPRs).52 The excitation of PSPR is typically done on continuous metal thin films using a prism or grating. The resonance then spreads along the metal/dielectric surface to a distance up to hundreds of micrometers.53 In the case of the LSPR, the plasmon resonance is not propagating and is excited on nanostructured metal surfaces. The properties can be adjusted by the size, shape, or composition of the nanostructures or nanoparticles.54 The SPR experiments are typically done in a direct,55 sandwich,56 or inhibition format.57 The direct assays are useful for bigger analytes that provide a sufficient response upon binding (Figure 5). The sandwich assay is based on a two-step procedure. The antibody first binds the analyte as in the direct format, followed by the capture of secondary antibodies (potentially labeled with enzymes or nanoparticles) to enhance the signal.58 The inhibition assay is based on mixing the analyte with respective antibodies, followed by the injection to the flow cell containing a chip with a known amount of immobilized analyte. The binding of free antibodies from the solution is evaluated to determine the analyte concentration. To efficiently immobilize the capture molecules to the sensor surface, coatings bearing chemical groups that allow the formation of covalent bonds are typically used.59, 60 Apart from the amount of the functional groups, sufficient layer stability under various conditions is necessary. Typically, the biomolecules are immobilized via primary amines or carboxyl groups; therefore, there should be either carboxyls or primary amines available on the surface to provide immobilization via the common carbodiimide / Nhydroxysuccinimide zero-length coupling reactions providing amide-based bonds. Numerous methodologies were investigated for surface modification with the desired chemical groups. Typically, wet-chemical procedures are being used, including binding of thiols on the surfaces of noble metals.61 However, these methods are time-consuming and often require the use of aggressive chemicals. Alternatively, plasma polymerization can be used to prepare thin Figure 5: Scheme of SPR immunoassay in direct format. Adapted from Paper VII with permission. Copyright 2020 American Chemical Society. 20 polymer films on various surfaces in a fast and eco-friendly procedure without the requirement of washing steps or reagent additions. Physical plasma is (partially) ionized gas consisting of photons, electrons, positive and negative ions, atoms, free radicals, and excited or non-excited molecules. Generation of plasma can be done by applying energy at low-, atmospheric-, or high-pressure conditions. Non-thermal plasmas are gaining increasing attention in many applications due to the possibility to provide enhanced gas-phase chemistry with high concentrations of chemically active species without the need for increased gas temperatures. The prepared plasma-polymerized (PP) coatings typically exhibit high branching and crosslinking, excellent adhesion to practically any surfaces, and high stability. The amine-rich PP coatings are the most commonly used and studied to allow biomolecule immobilization and cell binding.62, 63 The use of amine-rich PP films was demonstrated in biosensing, cell proliferation, and other biological applications.64, 65 Compared with the conventional layers, e.g., carboxymethylated dextran (CMD), faster and more costeffective layer preparation can be achieved by employing plasma processes. PP films with carboxyl groups are usually prepared by plasma polymerization of various acrylates.66 Alternative approaches can be based on gas mixtures of CO2 and ethylene67 or deposition from maleic anhydride (MA). The MA-based coating provided a highly reactive surface; however, the polymerization required fine-tuning as the layer was initially not sufficiently stable.68 We have explored different ways of surface modification by PP films for biosensing applications. First, we explored amine-based PP films composed of cyclopropylamine (CPA). The pulsed plasma polymerization of CPA can be used to prepare reactive nitrogen-containing films in a fast and eco-friendly way. As the layer stability is one of the critical aspects, we have first investigated the behavior of CPA PP films in aqueous media. The immersion in the buffer for 18 h before the glutaraldehyde activation turned out to be a critical step in maintaining longterm layer stability. The FT-IR and ellipsometry showed that the number of amine groups decreased, which was connected by the decreases of thickness by up to 17% after the immersion in the buffer. The results were explained by the hydrolysis of enamines or imines in the CPA PP; the chemical changes without thickness losses were caused by the hydrolysis of nitriles. The activation by glutaraldehyde led to the growth of a 5–7 nm thick film of glutaraldehyde and its oligomers. This surface was used to immobilize antibodies against human serum albumin (HSA) or Salmonella for the biosensing experiments. Furthermore, regeneration with 10 mM NaOH allowed multiple measurements with a single sensor. Since the commercial SPR sensors are most commonly based on carboxyl-containing layers,69 we have also examined the field of carboxyl-rich PP films. In our preliminary study, we have demonstrated the potential of PP films in SPR biosensing.70 We compared two kinds of PP films prepared under different plasma conditions; the first was based on polymerization from gas mixtures of maleic anhydride, acetylene, and argon (MA/C2H2/Ar; Figure 6A), the other on the mixture of CO2, ethylene, and argon (CO2/C2H4/Ar). The capacity of both surfaces to bind anti-HSA antibody was demonstrated, and the specific binding of HSA showed the biosensing potential. The layer based on CO2/C2H4 exhibited lower stability and smaller binding capacity, which resulted in the drift of the baseline and small response upon HSA injection. In contrast, the film based on MA/C2H2 allowed efficient immobilization of the 21 antibody and provided a stable baseline signal, which resulted in a higher response to HSA compared to the immunosensor based on CO2/C2H4. Based on these results and with the aim to develop a robust biosensing layer, we continued studying the preparation and properties of films based on MA/C2H2. We carried out a systematic comparison of MA/C2H2 PP films with sensors based on a mixed self-assembled monolayer of mercaptoundecanoic acid with mercaptohexanol (MUA/MCH) and with carboxymethylated dextrans (2D and 3D CMD) in term of the performance in the detection of HSA (Figure 6B). Compared with MUA/MCH and 2D CMD layers, the MA/C2H2 PP films showed better performance, demonstrated by about two times higher signals. On the other hand, the sensor based on 3D CMD still exhibited higher performance, especially providing a wider working range. This, however, can be explained by the significantly smaller surface area of the planar PP film compared to the 3D CMD. Figure 6: (A) Scheme of sensor surface modification by MA/C2H2 PP film with subsequent immobilization of the antibody via EDC/NHS chemistry. (B) SPR binding curves of HSA on antibodymodified MA/C2H2 PP film. Adapted from Paper VI with permission. Copyright 2019 Springer. In our next contribution, we used a high frequency-driven atmospheric-pressure plasma jet (APPJ; Figure 7) to prepare PP coatings based on 1,2,4-trivinylcyclohexane (TVC), tetrahydrofurfuryl methacrylate (THFMA), and a mixture of thereof. THFMA was selected because of the presence of the vinylidene group that can form polymers chains and tetrahydrofuran (THF) group, which served as a protective group against the breakdown of the THFMA. TVC contains three vinyl groups, which makes it effective as a monomer, but TVC also served as the source of carbon functionalities to adjust the carbon content and stability of the resulting PP films. Under plasma-induced polymerization conditions, the THF or cyclohexane ring-opening can happen, which results in further polymer chain cross-linking. Both the TVC and the THFMA are non-toxic, which fits the eco-friendly procedure of plasma polymerization. 22 The behavior of the films in an aqueous environment was studied. The highest stability was observed in the case of ppTVC, which contained the lowest amount of oxygen. In the case of ppTHFMA and ppTHFMA-co-TVC, more significant thickness losses occurred during the initial storage in water, however, with no impact on the chemical composition. After the stabilization for 24 h in liquid, all films have shown a high level of stability. The initial losses of film thickness can be explained by the removal of loosely bound oligomers from the film surface. AFM has demonstrated that the thickness losses were connected with the formation of characteristic morphological features. These led to an increase in the surface area and can be beneficial for the immobilization of a higher amount of antibody, resulting in higher sensor sensitivity. After the immobilization of the specific antibody, the PP-modified chips were used for the SPR detection of HSA. We have demonstrated that not only the number of functional groups affect the sensitivity of the measurement, but also the layer morphology is an essential factor. The ppTHFMA-co-TVC layer with the highest surface roughness provided the largest binding capacity. The sensors exhibited an excellent level of stability; the regeneration allowed to perform up to 9 measurements with a single sensor. The achieved LOD of 50 ng/mL is comparable with the performance of the 3D CMD chip, which confirms that the PP films are a promising alternative to the conventional surface modification techniques. Figure 7: Scheme of the atmospheric-pressure plasma jet (APPJ; left) and photograph of the operating APPJ (right). Reprinted from Paper VII with Permission. Copyright 2020 American Chemical Society. 23 4 Catalytic Labels 4.1 Amperometric Detection of M. plutonius (Paper VIII) Amperometric biosensors are devices based on the measurement of the current, corresponding to the analyte concentration, as a function of electrode potential or time. The amperometric methods can be used to determine the redox potential of the analyte, its electrochemical activity (adsorption, interaction with modified layers, electrocatalysis), but they are also sensitive to the changes of the electrode surface. In the case of amperometric catalytic biosensors, a suitable enzyme (or multi-enzyme system) is typically immobilized on the electrode, which catalyzes the transformation of the analyte; the concentration is determined from the measured current. In the case of amperometric immunosensors, the most common approach is based on labeling the analyte with a tracer (antibody conjugated with suitable enzyme); the analyte concentration is determined from the current measured upon the addition of the substrate solution. The most common potential-controlled (potentiostatic) measurement techniques include (i) chronoamperometry based on the measurement of a current at a fixed potential in time; (ii) single-potential amperometry based on the measurement of direct current as a function of the potential difference between two electrodes, and (iii) multiple-potential amperometry based on sweeping the potential in time and recording the corresponding current in the whole potential window.2 We have developed a chronoamperometric biosensor for the detection of Melissococcus plutonius, the causative agent of honeybee disease European foulbrood (EFB). EFB can typically be found in honeybee larvae up to five days of age, which get infected by the ingestion of food contaminated with M. plutonius. Upon the infection, the larvae color changes from white to yellow or brown, and larvae usually die displaced in the brood cells instead of normal coiled position.71 Because the infection can affect a large percentage of the brood, it can severely weaken the colony or even cause its collapse. It is crucial to prevent the uncontrolled spreading of the EFB to limit the economic and environmental consequences of the honeybee colony losses. Therefore, there is a high demand for methods that can detect M. plutonius in the stages of EFB infections, ideally in the PoC format. The typical detection approach is based on the microscopic evaluation of smears stained by carbol fuchsin. However, the sensitivity of this approach is not high enough to reveal the EFB in its early stages. Cultivation-based methods traditionally exhibit very high sensitivity at the cost of high time demands. However, in the case of EFB, the analysis is complicated by the low cultivation recoveries of M. plutonius and overgrowing by secondary invaders.71 The sensitivity, time-requirements, and throughput of the conventional methods can be overcome by using molecular detection methods based on either DNA or antibodies. Nowadays, realtime PCR is considered the gold standard for laboratory confirmation of M. plutonius.71 Even though antibody-based methods are widely used to detect various pathogens,72 they are not yet commonly used in the EFB diagnosis. There are no antibodies against M. plutonius commercially available; the need to prepare the antibodies in-house is clearly one of the factors limiting the faster growth of antibody-based methods for EFB. There is only a single report on the ELISA for the laboratory detection of M. plutonius.73 Furthermore, an LFIA assay for EFB 24 was recently reported.74 However, it allows only qualitative disease confirmation, suggesting room for more sensitive approaches. To start working in the field of EFB diagnosis, we have first prepared a rabbit polyclonal anti-M. plutonius antibody, and tested it in an ELISA assay.75 To develop an amperometric immunosensor based on a sandwich assay (Figure 8A), the antibody was immobilized to the gold working electrode via a self-assembled monolayer of cysteamine and glutaraldehyde. As the tracer, antibody conjugated with horseradish peroxidase (HRP) was used to provide electrochemical readout based on reducing the enzymatically oxidized 3,3´,5,5´-tetramethylbenzidine (TMB). Due to the use of the sandwich format, the effect of the complex sample matrix was suppressed compared to the label-free procedure based on EIS. Therefore, the amperometric approach is preferred for the analysis of complex samples of bees and larvae. The specific capture of M. plutonius on the sensor was verified by AFM. Even though the electrode exhibited substantial roughness, the bacteria were clearly visible. The achieved LOD was 6.6×104 CFU/mL for the pure bacterial sample in the buffer, and the sensor provided a working range up to 109 CFU/mL (Figure 8B). In the case of the real sample analysis, LODs of 2.4×105 and 7.0×105 CFU/mL were achieved for homogenized bees and larvae, respectively. The experiments with P. alvei as a negative control confirmed the high selectivity of the assay. The achieved sensitivity, together with a short analysis time of 2 h, confirm the suitability of the developed sensor in PoC diagnosis of EFB. Figure 8: (A) Scheme of an amperometric immunoassay for the detection of M. plutonius. (B) Amperometric response traces after the addition of TMB in the M. plutonius detection. Adapted from Paper VIII with permission. Copyright 2019 Wiley-VCH. 4.2 Enzymatic Precipitation-Enhanced SPR Detection of Salmonella (Paper IX) In recent years, various kinds of nanoparticles are being employed to enhance the performance of SPR immunosensors. For example, the application of gold nanoparticles in a sandwich format allowed to significantly increase the refractive index in the Salmonella detection.76 Magnetic nanoparticles can provide signal amplification due to the increased refractive index and the immunomagnetic preconcentration.77 However, the nanoparticle-based labels often suffer from a higher level of non-specific interactions compared to smaller molecules. To 25 overcome this limitation, a method of amplification based on enzyme-catalyzed precipitation of solid product on the sensor surface was developed.78, 79 The approach found application mainly in electrochemical sensing, with reports on the detection of prostate-specific antigen (PSA),80, 81 carcinoembryonic antigen (CEA),82 as well as E. coli.83 We introduced a method for the detection of Salmonella using SPR immunosensor enhanced by biocatalyzed precipitation. Our strategy aimed to develop a highly sensitive, robust, and straightforward assay while maintaining a reasonably short analysis time. The assay was based on the formation of sandwich immunocomplex of capture antibody, Salmonella, and HRP-conjugated detection antibody (Ab2-HRP). The HRP then catalyzed the conversion of 4-chloro-1-naphthol (4-CN) to insoluble benzo-4-chlorocyclohexadienone, which served as the signal enhancement step (Figure 9A). Figure 9: (A) Scheme of biocatalyzed precipitation-enhanced SPR detection of Salmonella. (B) SPR sensorgrams of the final step of biocatalyzed precipitation enhancement for different Salmonella concentrations. (C) Calibration curve. The inset shows the measuring chip after the precipitation reaction (antibody-modified channel – left, reference – right). Adapted from Paper IX with permission. Copyright 2016 American Chemical Society. 26 At the concentration of 107 CFU/mL, the signal change after precipitation enhancement was 40× higher than the signal with the bare bacterium. A closer look at the SPR signal changes in binding and reference channels revealed that even though some response was observed for the non-specific binding of Salmonella, there was practically no signal change upon injection of Ab2-HRP conjugate. This led to the increase of the ratio between the binding and reference channel from 3.5 (after Salmonella binding) to 8.6 (after biocatalyzed precipitation). Even though the injection of Salmonella in a very low concentration of 100 CFU/mL led to signal change comparable to the level of noise, the following precipitation reaction allowed to increase the signal to clearly distinguishable levels, allowing to reliably determine even very low Salmonella concentrations (Figure 9B). With the increasing concentration of bacteria, the measured signal increased exponentially. This suggests that several Ab2-HRP conjugates can be bound on a single Salmonella cells, leading to the precipitation of a large number of 4-CN molecules. This is a significant advantage to nanoparticle-based signal amplification, which leads only to linear enhancement of the signal.84 The obtained dependence of log(ΔR) on log(c) was linear from 102 to 106 CFU/mL, and the LOD was evaluated to be 100 CFU/mL (Figure 9C). The total assay time was 60 min, which is substantially shorter than the conventional methods for the detection of bacteria, including cultivation (~ days),85 ELISA (~ 10 h),86 and PCR (~ hours).87, 88 The analysis time is also shorter than in other reports on the amplification of SPR response with nanoparticles while achieving similar or better LOD.76, 84 Furthermore, the real-time operation of SPR can reveal higher bacteria concentrations in a short time upon sample injection (~ 10 min), allowing a rapid reaction even before the signal amplification is finished. After the measurement, the SPR chip was removed from the system and analyzed by the AFM. It was visible already by the naked eye that there was more precipitate formed in the measuring channel compared to the reference. AFM revealed the presence of bacteria and a large number of precipitate particles (~ 22,000 particles on the area of 20×20 µm2 ). Even though the reference channel also contained some precipitate particles, the number was significantly lower (~ 3,000 particles on 20×20 µm2 ). The 6.9-fold difference in the number of precipitate particles found by AFM corresponds to the 8.6-fold difference in SPR response for the same concentration. A closer look at the individual Salmonella before and after the precipitation revealed the presence of precipitate particles, leading to a three times increase of the height upon the precipitation (Figure 10). To demonstrate the applicability of the developed method for analyzing real samples, Salmonella was detected in powdered milk. Although immunosensors used specific antibodies to ensure selective detection, components of complex samples can still exhibit non-specific binding towards the sensor surface. However, as the Ab2-HRP conjugate used in our method is specific towards Salmonella, the potential non-specific binding is not transferred to the signal amplification step, contributing to the high selectivity of the method. In powdered milk, the achieved LOD was 103 CFU/mL, which is deterioration by one order of magnitude compared to the analysis in the buffer. This was probably caused by the non-specific adsorption of milk components, which can block some of the antibodies in the sensor surface or conceal some 27 epitopes on Salmonella. The ID50 (the number of bacteria that have to be ingested to result in 50% infection probability) of Salmonella is considered to be > 104 CFU.89 Furthermore, it was shown that ingestion of low Salmonella levels below 102 CFU/g does not pose a risk to human health.90 Therefore, the performance of the developed SPR immunosensor is suitable for the practical analysis of Salmonella in contaminated food samples. Figure 10: 3D representation of the AFM scan of (A) native and (B) precipitate-covered individual Salmonella cells. (C) Cross-sections of the bacteria evaluated as perpendicular lines in the center. Reprinted from Paper IX with permission. Copyright 2016 American Chemical Society. 4.3 Nanozyme-Linked Immunosorbent Assay (Paper X) The typical catalytical labels used in immunoassays are represented by enzymes, especially HRP. However, the enzymes suffer from several disadvantages, including the high cost of their production, limited stability, and activity reduction upon conjugation with immunoreagents. The properties of conventional enzymes can be overcome by using catalytic nanomaterials – nanozymes.91, 92 Compared to the biomolecules, the inorganic nanomaterials provide very high thermal and chemical stability.93 In particular, nanozymes with high peroxidase-like activity are preferred for immunoassays due to the compatible assay procedure with conventional ELISA. Nanozyme production can be done using aqueous solutions and benign precursors, making the procedure eco-friendly.94 The peroxidase-like activity was first discovered for magnetite nanoparticles (Fe3O4), followed by many other nanomaterials, including CeO2, CuO, Co3O4, and MnO2 nanoparticles, graphene oxide nanoplates, or Prussian blue nanoparticles (PBNPs).95, 96 We have introduced a method for the conjugation of PBNPs with antibodies and applied the conjugates in a nanozyme-linked immunosorbent assay (NLISA). The conjugation was based on the modification of the PBNP surface by reductively denatured bovine serum albumin (BSA), followed by the oxidation of antibodies by sodium periodate and binding them to the amino group of BSA (Figure 11). We have developed two sandwich NLISA assays, first for the detection of HSA in urine and the other for the detection of Salmonella in powdered milk. Because the oxidation of TMB to the blue product was utilized in the assay, the readout could be done using a standard colorimetric reader without special requirements on instrumentation. For the analysis of HSA in urine, the possible trace amounts of HSA present in the urine of healthy donors were first removed using centrifugal ultrafiltration on a 10-kDa membrane, followed by spiking known HSA concentrations. Even though the urine contains various ions, 28 which can cause undesired oxidation of TMB, the heterogeneous format with several washing steps allowed to overcome this limitation. As a result, only small differences were observed between HSA analysis in the buffer and urine. Microalbuminuria, which happens due to diabetic nephropathy, is connected with HSA concentrations from 20 to 200 μg/mL in the 24 h specimens.97 The optimized NLISA provided an LOD of 1.2 ng/mL of HSA in urine and a working range up to 1 μg/mL (Figure 12). This confirms that the NLISA is suitable for the practical diagnosis of microalbuminuria. Comparing the performance of NLISA with ELISA based on the same immunoreagents has shown only a small difference in LODs (1.2 ng/mL for NLISA and 3.7 ng/mL for ELISA). The comparable results suggest that the primary limiting step of the assay is not the detection step but rather the antibody affinity. Nevertheless, PBNPs offer several practical advantages, including higher stability, simple and cheap synthesis, and the possibility of catalyzing higher concentrations of TMB. As the shelf-life of PBNPs is practically unlimited (several years at 4 °C), antibodies are becoming the main limiting element. Therefore, the overall stability of the detection label could be, in the future, improved by replacing antibodies with MIPs or aptamers. Figure 12: Scheme of sandwich NLISA and calibration curve for HSA detection in spiked urine. Reprinted from Paper X with permission. Copyright 2018 American Chemical Society. Figure 11: Scheme of PBNP-Ab conjugate synthesis. The PBNPs are modified by denatured BSA to introduce amino groups on their surface, and the oxidized antibody is conjugated via the aldehyde groups. Reprinted from Paper X with permission. Copyright 2018 American Chemical Society. 29 The universal applicability of PBNPs as immunoassay labels was also demonstrated by detecting Salmonella in powdered milk. The incubation with milk served as an additional blocking step and had a minimum effect on the LOD. The optimized assay provided LOD of 6×103 CFU/mL with a working range up to 106 CFU/mL. This performance is better than the published ELISA assays that provide LODs between 104 and 106 CFU/mL.28, 98 The comparison with ELISA based on the same immunoreagents (LOD 3×103 CFU/mL) confirmed that the choice of antibodies affects the assay performance more significantly than the label. Overall, PBNPs proved to be a suitable alternative to conventionally used HRP; the detection of peroxidase-like activity is compatible with standard instrumentation and methodologies. Furthermore, the nanoparticle-based labels provide higher stability than biomolecules with the possibilities for cheap and large-scale production. 30 5 Photon-Upconversion Nanoparticles Photon-upconversion nanoparticles (UCNPs) are lanthanide-doped nanocrystals, which exhibit anti-Stokes emission. The energy transfer upconversion belongs to the non-linear optical processes and is based on the absorption of two or more photons, resulting in the emission of a single photon with higher energy (shorter wavelength).99 Compared with other anti-Stokes processes, including two-photon excitation and second harmonic generation, the excitation of upconversion can be done at lower energy densities. Even though the upconversion process was discovered already in the 1960s,100 it was only used in the form of bulk crystalline or glass materials.101 The composition of inorganic upconversion phosphor is based on a crystalline host matrix (typically NaYF4) with a dopant included at a low concentration (typically Yb3+ and Er3+ or Tm3+ ). The dopant ensures luminescence, while the crystal structure of the host lattice provides a matrix to bring the dopant ions into the optimal position.102 Due to the remarkable progress in nanotechnology, methods for the synthesis of upconversion nanomaterials were discovered, leading to UCNPs with high luminescence efficiency. Compared to conventional luminescence labels, such as organic fluorophores or QDs, UCNPs can be detected without autofluorescence background, they provide large antiStokes shifts allowing easy separation of the excitation and detection channels, exhibit excellent photostability, and the emission wavelength is tunable to enable multiplexed detection.102 The synthesis of UCNPs is typically done in hydrophobic solvents, such as oleic acid and octadecene. Therefore, their surface has to be modified for biological applications.103 One of the most widespread surface modification techniques is silanization. It is based on the hydrolysis and condensation of siloxane precursors, typically tetraethyl orthosilicate104 with other derivatives of silane, to provide functional groups for further bioconjugations.105 Other modification techniques are based on exchanging the hydrophobic surface ligands by ligands with hydrophilic properties. Ligands bearing phosphonate or carboxylate groups can coordinate to the lanthanide ions on the UCNP surface; the functional groups on the other end of the ligand are then used for the bioconjugation.106 Throughout our extensive work in the UCNP field, we have explored different ways of UCNPs surface modifications and employed the conjugates to develop immunoassays for a wide range of analytes, from small molecules to bacteria. 5.1 Competitive Upconversion-Linked Immunosorbent Assays 5.1.1 Competitive ULISA for Diclofenac (Papers XI and XII) Diclofenac (DCF) is a widely used non-steroidal anti-inflammatory drug. The widespread use of DCF for cattle treatment in the Indian subcontinent has led to significant vulture population losses in the 1990s because DCF caused renal failure in vultures feeding on contaminated carcasses.107 DCF is one of the most frequently analyzed pharmaceuticals in the water-cycle in Europe because it cannot be easily degraded in water treatment plants. It was detected in the amounts of low μg/L wastewater effluents and amounts of ng/L in surface waters,108 groundwater, and drinking water.109 The sensitive detection of DCF is typically done by LCTOF-MS or high-resolution mass spectrometers.110 However, these methods require expensive 31 instrumentation with trained personnel, and the analysis is lengthy. On the other hand, ELISA assays are highly suitable for analyzing a large number of samples, even in smaller, lessequipped laboratories.111 In our pioneering work on the immunoassay based on UCNPs – the upconversion linked immunosorbent assay (ULISA) – we have synthesized conjugates of UCNPs with detection anti-mouse antibody and applied them in an indirect competitive assay for DCF (Figure 13A). We have synthesized oleic acid-capped UCNPs and coated them with a silica shell bearing carboxyl groups on the surface. This modification was used to improve the water dispersibility and conjugate antibodies using EDC/sulfo-NHS chemistry.112 The coating antigen in a competitive immunoassay has to be, on hand, in low-enough concentration to allow efficient competition for the binding sites of the detection antibodies, but, on the other hand, its concentration still has to provide strong-enough signals. To achieve the optimal assay performance, we have prepared and characterized two different coating conjugates by modifying bovine serum albumin (BSA) with DCF. The MALDI-TOF mass spectrometry analysis revealed that the conjugates carried either 5.7 or 10 DCF molecules per BSA. Even though the conjugate with the higher degree of derivatization provided about twice as high signals, larger signal fluctuations and hook effect were observed. On the other hand, the conjugate with 5.7 DCF molecules per BSA provided better signal stability, slightly lower IC50 value (1.2 ng/mL compared to 1.5 ng/mL), and lower LOD. The optimized ULISA assay provided an LOD of 0.05 ng/mL, which was five times higher than the LOD of a conventional ELISA (0.01 ng/mL; Figure 13B). However, the ULISA allowed for a faster and easier signal generation. However, it was most notably a first step in the further development of ULISA assays that were eventually going to reach a single-molecule sensitivity. Figure 13: (A) Scheme of indirect competitive ULISA for the detection of DCF. A microtiter plate is coated with a BSA-DCF conjugate, dilution series of DCF are prepared in the microtiter plate followed by the addition of anti-DCF mouse antibody, and the attachment of anti-DCF antibody is detected by a conjugate of secondary antibody with UCNP. (B) Normalized calibration curves of ULISA and ELISA. Adapted from Paper XI with permission. Copyright 2016 American Chemical Society. In the follow-up work, we focused on improving the LOD and reducing the analysis time by designing a single-step ULISA assay (Figure 14A). For the synthesis of the DCF tracer, we have prepared a conjugate to DCF with bovine γ-globulin (BGG) and conjugated it on the surface of UCNPs with a carboxylated silica shell. The DCF-BGG conjugate was used 32 because it provided structural flexibility between the DCF and the UCNPs, and it prevented non-specific binding of the tracer to the microtiter plate. Because the UCNP-DCF tracer was used to directly compete with the analyte DCF for the binding on the immobilized anti-DCF antibody, the assay was done in a single step. Furthermore, we have optimized a method for the lyophilization of the tracer, which did not negatively affect its performance even after prolonged storage at room temperature. The single-step analysis and the possibility of storing the tracer in a dry state without the necessity of cooling makes the ULISA an excellent option for environmental analysis in low-resource settings.113 The optimized assay provided an LOD of 0.02 ng/mL with a signal-to-background (S/B) ratio of 82 (Figure 14B). The high value of S/B was enabled by (i) the high brightness of the used UCNPs with a diameter of 90 nm, (ii) the low level of non-specific interaction due to the coating with BGG, and (iii) the presence of multiple DCF molecules per single UCNP ensures efficient competition even when the molar concentration of tracer is significantly lower than the concentration of DCF. Finally, we have demonstrated the practical potential of the method on the successful analysis of real samples of drinking and river water. Figure 14: (A) Scheme of single-step competitive ULISA for the detection of DCF and (B) calibration curve. Adapted from Paper XII with permission. Copyright 2017 Springer. 5.1.2 Competitive ULISA for Zearalenone (Paper XIII) We have also developed an assay for the analysis of mycotoxin zearalenone, based on the competition with epitope mimicking peptide. Microbial toxins are produced by many pathogenic microorganisms (bacteria and fungi) and act as their virulence agents. They are represented by a heterogeneous group of compounds that interfere with biochemical processes, including the function of membranes, transport of ions, release of transmitters, and synthesis of macromolecules. Exposure to the toxins either in food or in the environment can cause significant health problems; the individual symptoms vary significantly between the different toxins.114 Unlike in the case of viable bacteria, toxins are typically not affected by the heat processing of the product. Zearalenone (ZEA) is a non-steroidal estrogenic mycotoxin produced by several fungi species of the Fusarium genus.115 Even though the acute toxicity of ZEA is relatively low, it is chronically toxic and has been connected with reproduction disorders of farm animals, mainly pigs.116 ZEA exhibits estrogenic, genotoxic, haematotoxic, and anabolic effects.117 Along with 33 other mycotoxins, ZEA is often found in agricultural products, including maize, wheat, barley, rice, and oats.118 Epitope mimicking peptides, also referred to as mimotopes, are used as an alternative to conventional analyte-conjugates in competitive immunosensing. Such peptides mimic the epitope of the analyte and allow competition with the native analyte for binding to the antibody. Even though antibodies with high affinity are required in all immunoassay, competing peptides with lower affinity can be beneficial for competitive assays. They shift the equilibrium towards analyte binding, making a smaller amount of analyte produce the same response level.119 Based on the previously identified amino acid sequence,120 we have synthesized the peptide mimetic of ZEA, introduced biotin on its C-terminus, and used it in an ULISA assay for ZEA detection (Figure 15). The specific anti-ZEA antibody was bound on the surface of the microtiter plate, allowing competition between analyte ZEA and biotinylated peptide mimetic. The detection was carried out using the conjugate of UCNPs with streptavidin. The optimized assay provided an LOD of 20 pg/mL with a working range up to 0.5 ng/mL, representing a 200-fold improvement of LOD and 3-fold improvement of working range compared to the previously reported bioluminescence immunoassay with the same peptide mimetic.120 Figure 15: Scheme of competitive ULISA for ZEA. In the first step, a microtiter plate is coated with an anti-ZEA antibody, and ZEA in the sample competes with the biotinylated peptide mimetic for a limited amount of antibody binding sites. In the second step, the conjugates of UCNPs with streptavidin bind to the biotinylated peptide. Reprinted from Paper XIII with permission. Copyright 2020 Elsevier. 34 To confirm the potential of ULISA for the analysis of complex samples, analysis of ZEA-free maize samples (as confirmed by UPLC-MS/MS) spiked with ZEA was done. The recoveries between 77% and 105% demonstrate the suitable accuracy for quantitative real sample analysis. Furthermore, in comparison with UPLC-MS/MS, ULISA is based only on a simple extraction in methanol and does not require extensive sample pre-treatment. The achieved performance fulfills the requirements given by the European legislation, which confirms the suitability of ULISA as a tool for simple analysis of food samples contaminated by mycotoxins. 5.2 Sandwich Upconversion-Linked Immunosorbent Assays 5.2.1 Detection of M. plutonius with BSA-Modified UCNPs (Paper XIV) UCNPs are also highly useful as labels in sandwich immunoassays. We have introduced a method to conjugate UCNPs with streptavidin based on a copper-free click reaction and used this conjugate to detect M. plutonius, the causative agent of European foulbrood (Figure 16). The conjugation was based on strain-promoted cycloaddition between bicyclo[6.1.0]nonyne (BCN) groups bound on the UCNP surface via BSA and azide-modified streptavidin.121 Apart from serving as an intermediate to bind the BCN, BSA also contributes to the reduction of nonspecific binding of the UCNP conjugates. Figure 16: Scheme of the conjugation of UCNP with streptavidin based on functionalization of UCNP surface with alkyne-modified BSA and click reaction with azide-modified streptavidin. The conjugates are then employed as a label in a sandwich ULISA for the detection of M. plutonius. Reprinted from paper XIV with permission. Copyright 2019 Royal Society of Chemistry. The bioconjugation reaction was followed using agarose gel electrophoresis based on fluorescence labeling of BSA-BCN and streptavidin-azide. The overlap of the fluorescence signals, together with the mobility shift, confirmed the successful progress of the conjugation reaction. Furthermore, the presence of BSA and streptavidin on the UCNP surface was confirmed by mass spectrometry. After the digestion of proteins on the UCNP conjugates by trypsin, the UCNP cores were removed by centrifugation, and the samples were analyzed by 35 LC-MS/MS. Both BSA and streptavidin were successfully identified; according to the integrated signal intensities, BSA was detected as the most abundant protein in the sample. The amount of streptavidin was approximately one order of magnitude lower than the amount of BSA. First, we have optimized ULISA on the detection of M. plutonius in the buffer. The assay provided an LOD of 340 CFU/mL and a wide working range up to 109 CFU/mL. This LOD is 400 times better in comparison with ELISA (Figure 17). Since the same antibodies were used in both assays, the most significant impact on enhancement can be accounted to the high label performance, providing a highly sensitive readout of anti-Stokes emission and the low non-specific binding. To demonstrate the practical applicability of the assay, real samples of bees, larvae, and bottom hive debris were analyzed, representing the typical matrices where M. plutonius has to be detected during the infection by EFB. The achieved LODs were 540 CFU/mL for bee extract, 8.5×103 CFU/mL for larvae extract, and 570 CFU/mL for bottom hive debris. The level of M. plutonius in infected apiaries with clinical symptoms is typically around 105 CFU/mL,122 demonstrating the suitability of the developed ULISA for the practical EFB diagnosis in the early stages of the infection. 5.2.2 Preparation of PEG-Modified UCNPs and Analysis of HSA (Paper XV) In our next work, we have introduced a different strategy for UCNP surface modification based on coating the particles with a PEG linker and applying the conjugates in a sandwich immunoassay for the detection of albuminuria marker HSA. For the surface modifications, we have chosen heterobifunctional PEG with neridronate, and alkyne or maleimide functional endgroups based on these considerations; (i) PEG can sterically stabilize the particles and resist the non-specific interactions with surfaces and biomolecules;123 (ii) neridronate shows strong coordination towards lanthanide ions of UCNPs;124 and (iii) the alkyne or maleimide groups can be used for the subsequent conjugation of biomolecules.125 The first conjugation approach was based on attaching azide-modified antibody or streptavidin to the alkyne groups via click reaction (Figure 18). Alternatively, the disulfide bonds in the antibody were reduced by TCEP, and the generated thiol moieties were bound to the maleimide groups. Figure 17: Comparison of sandwich ULISA and ELISA assays for the detection of M. plutonius. The normalized signals were calculated by dividing the data by yMAX value from the logistic fit. Reprinted from Paper XIV with permission. Copyright 2019 Royal Society of Chemistry. 36 Figure 18: Scheme of the preparation of PEG-based conjugates of UCNPs with streptavidin. The oleic acid on the surface of as-synthesized UCNPs was removed by a ligand exchange reaction with nitrosonium tetrafluoroborate to prepare water-dispersible nanoparticles. The particles were then coated with an alkyne-PEG-neridronate linker, and streptavidin-azide was coupled via copper-catalyzed click reaction. The prepared bioconjugates were used in the sandwich ULISA assay for HSA detection. To allow efficient use in immunoassay, the nanoparticle-based labels must provide not only a high level of modification with biorecognition molecule to allow specific binding, but the conjugates must also show high uniformity with a small number of aggregates to reduce the signal fluctuations. First, we have tested the UCNPs modified by antibody via alkyne-azide click reaction. Two different ratios between the antibody and NHS-PEG-N3 were tested. Even though higher signals were observed in the case of conjugate based on antibody with a higher number of azide molecules, there was no positive effect on the LOD. This can be explained by the higher number of aggregated conjugates, resulting in higher signal fluctuation. The optimized ULISA based on these particles provided an LOD of 3.5 ng/mL and a working range up to 1 μg/mL. In contrast, the antibody conjugates based on maleimide coupling provided higher signals and slightly lower background, resulting in the improvement of the LOD to 0.24 ng/mL and an unchanged working range up to 1 μg/mL. The conjugates with streptavidin reached an even lower LOD of 0.17 ng/mL (Figure 19). This can be explained by the more efficient binding of the streptavidin-coated UCNPs to the biotinylated antibody, caused mainly by the flexibility of the additional antibody present in the immunocomplex compared to the binding directly to the antigen. The comparison with ELISA (LOD 0.56 ng/mL) and fluorescence immunoassay (LOD of 0.59 ng/mL) based on the same immunoreagents demonstrated that the use of UCNPs is advantageous in term of assay performance compared to the conventional labels. Figure 19: ULISA for the detection of HSA with UCNPPEG-SA label. Adapted from Paper XV with permission. Copyright 2020 American Chemical Society. 37 The optimized assay based on streptavidin-conjugated UCNPs was then used for the analysis of HSA in spiked urine. A slightly higher baseline was observed for the urine compared to the buffer, which was probably caused by the presence of some HSA levels, even in the samples from healthy donors. However, this did not affect the ability to specifically detect the HSA, demonstrating the potential of the method for practical applications. 5.2.3 Detection of P. larvae with PEG-Modified UCNPs (Paper XVI) Afterward, we have employed the PEG-based UCNP conjugates to detect spore-forming bacterium Paenibacillus larvae, the causative agent of American foulbrood (AFB). AFB represents the most dangerous honeybee brood disease and causes significant economic losses throughout the world.126 The honeybee larvae are infected by ingesting the fee contaminated by spores. The spores then germinate and colonize the midgut of the larvae, which is followed by spreading the bacteria over the midgut epithelium and the body cavity of the larva.127 The dead larvae are found as a glue-like mass sticking to the side of the honeycomb cell, which is used as the typical sign for the AFB diagnosis. Afterward, the bacteria sporulate, and the spores are spread around the hive by the adult honeybees, which results in the infection of more larvae by the ingestion of contaminated food reserves. The spores can be found not only in the diseased larvae and the resulting dry scales but also in adult worker bees, honey, bottom hive debris, beehive surfaces, and beekeeping equipment.128 Because the spores are highly resilient, the discovery of infection is usually connected with burning down the honeybee colonies, as well as the contaminated equipment.129 Therefore, sensitive diagnostic approaches are required to allow early diagnosis, preventing the infection from spreading further.130 The traditional diagnosis of AFB is based on observing the clinical signs within the hive. Microscopic evaluation of stained smears from diseased larvae can be used for fast detection; however, the sensitivity of this approach is not high enough to diagnose the infection in its early stages.131 On the other hand, culture-based methods provide excellent sensitivity, but the cultivation takes several days, making this approach not suitable for screening purposes.132 Currently, PCR is considered the gold standard for AFB diagnosis, as it combines high sensitivity with fast analysis.132, 133 However, PCR-inhibitors and other contaminants in the honeybee material can complicate the analysis of real samples.134 The development of immunochemical methods for AFB diagnosis is generally hampered by the lack of commercially available antibodies against P. larvae. Even though there was a single report on the ELISA for the AFB diagnosis,135 its sensitivity did not allow detecting sub-clinical P. larvae levels. We have prepared a rabbit polyclonal anti-P. larvae antibody and used it in ULISA assay to allow early diagnosis of AFB. Cell wall fraction of P. larvae was used for the immunization of two rabbits. However, only one serum was used for further experiments because the other showed high cross-reactivity with M. plutonius. Even though affinity purification is generally used to suppress the cross-reactivity of generated antibodies, it was not possible here because of the complex nature of the used antigen.136 Therefore, all IgGs present in the antiserum were isolated by the purification on protein G affinity column. First, an ELISA assay was used for testing the antibody specificity. To allow performing a sandwich assay, the antibody was conjugated with biotin, and the conjugate of streptavidin 38 with HRP (SA-HRP) was used as a tracer. The assay provided an LOD of 6.5×104 CFU/mL and a S/B ratio of 34. This performance is comparable with the LOD of 1×105 CFU/mL published by Olsen et al.;135 however, it is not sufficient to analyze sub-clinical levels of the bacterium. Therefore, the SA-HRP conjugate was replaced by the streptavidin-coated UCNPs (Figure 20A) in the ULISA assay (Figure 20B). The S/B value of 128 was achieved in the optimized assay, representing a 4-fold improvement compared to the ELISA. In the case of negative controls of P. alvei, M. plutonius, and B. laterosporus, only small signal changes were observed compared to the target bacterium P. larvae. The assay provided an LOD of 2.9×103 CFU/mL (Figure 20C), which is 22 times better than the ELISA with the same antibody, clearly demonstrating the advantage of the labels based on UCNPs. Finally, the successful analysis of real samples of bees, larvae, and bottom hive debris showed the potential of ULISA in the practical diagnosis of AFB. Figure 20: (A) Structure of UCNP-PEG-SA conjugate. (B) Scheme of sandwich ULISA for the detection of P. larvae. (C) Cross-reactivity testing of ULISA with P. larvae as a specific target and P. alvei, M. plutonius, and B. laterosporus as negative controls. Adapted from Paper XVI with permission. Copyright 2021 IEEE. 5.3 Single-Molecule Upconversion-Linked Immunosorbent Assays 5.3.1 Digital ULISA for PSA (Papers XVII and XVIII) Due to the very low optical background, it is even possible to detect individual UCNPs. We have developed an approach allowing visualization of single UCNPs under a conventional epiluminescence microscope and used it in a single-molecule assay of cancer biomarker prostate-specific antigen (PSA). Prostate cancer is globally the fifth leading cause of death from cancer and the most often diagnosed cancer type among men.137 PSA is secreted by the epithelial cells of the prostate in a typical concentration in healthy men below 4 ng/mL; the increase above this level can be used in the prostate cancer diagnosis.138 When the carcinoma is removed by the radical prostatectomy, the PSA levels decrease significantly.139 However, 39 repeated monitoring of PSA is still necessary because the increase from levels below 0.1 ng/mL to consistently above 0.2 ng/mL is connected with the biochemical recurrence,140 which occurs in up to 40% of cases after the surgery.141 This highlights the need for sensitive assays to detect the recurrence of cancer as early as possible. In the first step, we have modified an epifluorescence microscope Nikon Eclipse Ti-E for the imaging of UCNPs (Figure 21). The microscope was equipped with a 980-nm excitation laser, 100× heat resistant objective, and suitable filter sets to detect upconversion luminescence of Er- and Tmdoped UCNPs. For optimizing the setup, carboxylated UCNPs with sizes from 37 to 90 nm were immobilized on a glass slide modified by cationized BSA. The excitation by 980-laser resulted in a very low background that – when there were no UCNPs present on the surface – depended only on the signal noise of the camera. All tested UCNP types were visible as individual diffraction-limited spots with a diameter of ~ 400 nm; the number of detected UCNPs was directly proportional to their concentration. Microtiter plates with a thin foil (190 µm) at the bottom of each well were used in an immunoassay because of the short working distance of the objective with the high numerical aperture. The assay was based on the sandwich immunocomplex of the capture antibody, PSA, and the conjugate of silica-coated UCNPs with detection antibody. First, the upconversion luminescence was read out by a conventional microtiter plate reader (analog ULISA), followed by counting the individual immunocomplexes under the microscope (digital ULISA). There was a small number of upconversion spots visible, in the case of the blank with no PSA in the samples, which corresponds to the non-specific binding of UCNPs to the surface of the microtiter plate and defines the LOD similarly as in the conventional immunoassays. However, compared to the analog readout, digital detection offers several advantages: Because the signal of an individual label can be reliably distinguished from the background noise, the background fluctuations do not influence the measurement. Therefore, the LOD is limited only by the affinity and non-specific binding of the immunoreagents used in the assay. Furthermore, in contrast to the intensity-based readout, where a few big aggregates can strongly affect the overall intensity, the digital approach counts the aggregates as single binding events, reducing their effect on the measured signal.142 This enhances the robustness of the measurement and indirectly allows for achieving lower LODs. Figure 21: Scheme of upconversion microscope. The inset shows individual UCNPs as diffractionlimited spots. Adapted from Paper XVII with permission. Copyright 2017 American Chemical Society. 40 Figure 22: Upconversion microscopy images of microtiter plate after binding of serial dilutions of PSA in 25% serum. The calibration curves yield LOD of 1.2 pg/mL in the digital and 20.3 pg/mL in the analog mode. Adapted from Paper XVII with permission. Copyright 2017 American Chemical Society. The analog readout of the immunoassay for PSA in 25% serum provided an LOD of 20.3 pg/mL with a working range from 100 pg/mL to 10 ng/mL. This is comparable with the commercial ELISA assays for the PSA.143 The digital readout of the same microtiter plate (Figure 22) allowed to lower the LOD by more than one order of magnitude down to 1.2 pg/mL, with a working range from 10 pg/mL to 1 ng/mL. It was not possible to analyze higher PSA concentrations because the point-spread functions of the UCNPs started to overlap and did not allow reliable counting. It is also possible to combine both readout options, extending the overall working range to three orders of magnitude from 10 pg/mL to 10 ng/mL. The PSA detection in the buffer provided practically identical results, confirming that the 25% serum has a negligible matrix effect. In our follow-up work, we have further improved the single-molecule assay scheme by replacing the conjugates of silica-coated UCNPs with antibody by PEG-coated UCNPs conjugated with streptavidin (Figure 23). Such conjugate was expected to provide better performance because (i) the PEG provides resistance of UCNPs against aggregation and ensures high dispersibility in water; (ii) the steric hindrance of the PEG reduces the level of non-specific interactions;144, 145 and (iii) the subsequent addition of biotinylated detection antibody and streptavidin-coated UCNPs allows using a relatively high detection antibody concentration to efficiently label all PSA molecules while being able to reduce the UCNP concentration due to the high affinity between streptavidin and biotin,146 which leads to a lower amount of non-specifically adsorbed UCNPs. Figure 23: Upconversion microscopy image of a microtiter plate after the specific capture of PSA and a scheme of individual sandwich immunocomplex. Reprinted from Paper XVIII with permission. Copyright 2019 American Chemical Society. 41 Compared to our previous study,147 the new assay design lowered the LOD by 50 times. The comparison of an assay based on Er-doped (LOD 23 fg/mL, 800 aM) and Tm-doped UCNPs LOD 24 fg/mL, 840 aM; Figure 24A) demonstrated that the advantages of the digital readout are not dependent on the label type. The three times lower value of IC50 further confirms that the two-step label design provides improved binding kinetics. For the real sample analysis, random samples of human serum were collected in the hospital and analyzed by electrochemiluminescence immunoassay as a reference method. The dilution linearity experiments have shown that human serum has a low matrix effect on the assay, as represented by the recovery rate fluctuations below 20%. The results from electrochemiluminescence assay and ULISAs (both analog and digital) were in great agreement, confirming the potential of UCNPs to be used as a label in assays for the diagnosis of prostate cancer (Figure 24B). Figure 24: (A) Calibration curves of the ULISA in the digital (red) and analog (black) mode. The logarithmic scale of the y-axis highlights the signals in the low PSA concentration range. (B) Correlation between the PSA concentrations in human serum samples determined by the digital (red) or analog (black) ULISA and an electroluminescence immunoassay. Reprinted from Paper XVIII with permission. Copyright 2019 American Chemical Society. 5.3.2 Digital ULISA for Cardiac Troponin (Paper XIX) The digital ULISA can be adapted for the detection of other biomarkers by simply exchanging the immunoreagents. Thus, to demonstrate the universal nature of the approach, we focused on detecting cardiac troponin, a biomarker of acute myocardial infarction (AMI). Heart diseases represent the leading cause of death worldwide. There is typically only a short time available since the symptoms start before the treatment is necessary, creating demand for rapid and reliable diagnostic assays.148 Cardiac troponin is one of the recommended biomarkers for the diagnosis of AMI. Because it is located only in myocardial tissue in healthy individuals, its elevated concentration in blood can indicate the onset of AMI.149 Cardiac troponin is a heterotrimeric complex, which consists of three distinct subunits – cTnI, cTnT, and TnC.150 The subunits cTnI and cTnT are present only in the myocardium (heart muscle), and during the AMI, they are released into the bloodstream.151, 152 The diagnosis of AMI can be made based on measuring the changes in the cTnI levels.153 42 However, cTnI is a challenging analyte for the detection by immunochemical methods as its recognition by antibodies can be affected by several factors.154, 155 The N- and C-terminal parts of cTnI are susceptible to proteolytic degradation,156 favoring the use of antibodies that target epitopes in the central region.157 In addition, cTnI is typically present in blood in the form of a binary cTnI-TnC complex,158 thus, the antibodies should recognize free as well as complex form cTnI. Furthermore, the epitopes can be phosphorylated or blocked by autoantibodies or heterophile antibodies, hindering immunochemical recognition.159 For this reason, assays for cTnI are often based on the combination of two capture or two detection antibodies.154 In our work, we studied the impact of size and surface modification of UCNP labels on analog and digital ULISA for the detection of cTnI. The size of UCNP-based detection labels is one of the critical factors affecting assay performance. On the one hand, the particles should be as small as possible to (i) provide stable dispersions, (ii) minimize the level of non-specific binding, and (iii) limit the influence of the UCNPs and the immunochemical interaction. On the other hand, larger size leads to higher brightness of the UCNPs, making them more easily detectable. Furthermore, we tested two ways of surface modification, based on (i) alkyne-PEGneridronate linker and streptavidin and (ii) poly(acrylic acid) (PAA) and antibody; the corresponding assay schemes are shown in Figure 25. Figure 25: ULISA configurations for the detection of cTnI with labels based on (A) UCNP-PEG-SA and (B) UCNP-PAA-Ab. Adapted from Paper XIX under the permission of Creative Commons Attribution-NonCommercial 4.0 International License. We found out that the size and surface modification of UCNPs affect the assay performance more than the difference between analog and digital readout modes. Varying the UCNPs size affected especially the assays in human plasma; the increasing size resulted in a higher level of non-specific binding; however, the smaller UCNPs exhibited a slightly higher degree of aggregation. The highest sensitivity was achieved when using PAA-based UCNPs with a diameter of 48 nm. Surprisingly, the LODs in human plasma provided by analog (8.6 pg/mL) and digital (9.8 pg/mL) readout were very similar. This contradicts with the results 43 achieved when detecting PSA where digital readout showed significantly higher sensitivity than the analog one. This discrepancy might be caused by the different affinity of the antibodyantigen pairs for PSA and cTnI. The analog readout provided a 10-fold lower LOD for PSA compared to cTnI, and the difference increased to 200-fold in the digital mode. These results thus suggest that a large enough antibody affinity is required to allow the digital readout to further increase the sensitivity of the ULISA. 5.4 UCNP-Based Immunocytochemistry (Paper XX) We have also demonstrated that UCNPs can be advantageously used to label breast cancer biomarkers in immunocytochemistry (ICC) and immunohistochemistry (IHC). With around 2.1 million new cases reported every year, breast cancer is the second most common type of cancer worldwide.160 Even though mammographic screening and advances in adjuvant systemic therapy help fighting the disease, its incidence continues growing.161 Human epidermal growth factor receptors (HER or ErbB) belong to membrane receptors, which play essential roles in biological processes, including apoptosis and migration, differentiation, and proliferation of cells. The overexpression of the HER2 receptor on cancer cells happens in 10– 30% of all patients with breast cancer. Due to the association with an increased rate of cell proliferation, which results in rapid cancer growth and poor prognosis, HER2 is often used as a biomarker in cancer diagnostics.162 IHC allows detecting and localizing antigens within histological tissues, which can be used to identify cancerous cells.163 The optimization of protocols and testing of new staining and labeling methods can be done in ICC, which targets cultivated cells prepared similarly as the tissue samples. The most common counterstaining approach is based on the combination of hematoxylin and eosin (H&E).164 However, to allow specific detection of cancer biomarkers, conjugates of antibodies with enzymes,165 fluorophores,166 or nanoparticles167 have to be used. Typically, HRP is employed to oxidize 3,3’-diaminobenzidine to a brown precipitation product, which is evaluated by light microscopy.168 The evaluation of the tissue sections is typically done by a timeconsuming visual inspection by the trained pathologists. To increase the throughput, the current research focuses on the automation of imaging and evaluation aided by artificial intelligence in so-called digital pathology.169 However, the automation requires labels providing high specific signals and low non-specific binding.170 Due to their high brightness and low non-specific adsorption, we have explored the capabilities of BSA-based75 and PEG-based143 conjugates of UCNPs with streptavidin for labeling of HER2 biomarker on cancer cells (Figure 26). The conjugates based on PEG provided a higher S/B ratio, Figure 26: Scheme of the ICC assay. The primary antibody binds to the HER2 receptor on the cell surface, followed by a biotinylated secondary antibody, and detection UCNP-streptavidin conjugate. Adapted from Paper XX with permission. Copyright 2020 Royal Society of Chemistry. 44 probably due to the non-specific adsorption of BSA to the cell surface. The efficiency of labeling was strongly dependent on the blocking conditions. The S/B ratios were calculated as the ratio of signals between HER2-positive BT-474 cells incubated with and without primary antibody, as evaluated by upconversion scanning. Even though the assay buffer based on BSA and BGG allowed to reach an acceptable S/B ratio of 23, the use of commercial SuperBlock solution reduced the non-specific binding more efficiently while even slightly increasing the specific signals (Figure 27). This resulted in the improvement of the S/B ratio by more than one order of magnitude to 319. This finding agrees with the previous results, suggesting that the presence of serum proteins is not optimal for achieving low backgrounds in ICC. The comparison of HER2-positive BT-474 cells with HER2-negative MDA-MB-231 cells under the same experimental conditions also confirmed a high level of specific binding, producing 40 times higher signals in the case of BT-474. We have also shown that upconversion-based labeling is compatible with the H&E counterstaining, suggesting good applicability in IHC, where the H&E is the typical counterstaining method. The performance of UCNP-based labels was compared with conventional fluorescence labeling using a conjugate of streptavidin with carboxyfluorescein (SA-FAM). The fluorescence labeling resulted in S/B value of only 6.1, which is connected with relatively high background signals due to the cellular autofluorescence and cross-talk between the detection channels (fluorescein and DAPI). Due to the 50-fold wider dynamic range, the upconversion labeling allows a much finer distinction of HER2 expression within different cell lines. Furthermore, the high S/B ratio can enable the application of UCNP labeling for IHC with automated data evaluation in digital pathology. Figure 27: Labeling of HER2-positive FFPE BT-474 cells using UCNP-PEG-SA conjugates: (A) DAPI, (B) upconversion, (C) overlay. Negative control (without primary antibody): (D) DAPI, (E) upconversion, (F) overlay. Adapted from Paper XX with permission. Copyright 2020 Royal Society of Chemistry. 45 6 Laser-Induced Breakdown Spectroscopy (Papers XXI and XXII) The direct nanoparticle-based immunoassay readout (i.e., not employing catalytic transformation of the substrate) is typically based on luminescence detection, which combines high sensitivity and simple instrumentation. However, the requirement of luminescence properties limits the range of potential labels. Therefore, there is a demand for alternative readout techniques that would allow universal detection independent of the luminescent or catalytic properties of the label. Laser-induced breakdown spectroscopy (LIBS) is an optical emission technique complementary to the conventional methods in bioimaging.171 LIBS combines high sensitivity, rapid analysis, and the possibility to detect halogens and light elements. However, its main advantage is the possibility of multi-elemental imaging on a large scale (few cm) and with a high resolution (units of μm).172 LIBS can be used to detect different kinds of nanoparticles on various matrices, from the analysis of QDs on a filter paper173 to UCNPs in model organisms.174 Furthermore, LIBS can be used for surface mapping, providing information about the 2D or even 3D element distribution within the sample.175 We have developed a method for the detection of Ag NPs and Au NPs by LIBS from the bottom of the conventional 96-well microtiter plate (Figure 28). The optimized setup was then applied for the readout of a sandwich immunoassay to detect HSA based on streptavidinconjugated Ag NPs. The performance of the LIBS-based assay was compared with a conventional fluorescence readout based on the conjugate of detection antibody with fluorescein isothiocyanate (FITC). Even though slightly higher sensitivity was observed in the case of fluorescence, the great advantage of LIBS was the wider dynamic range. Furthermore, LIBS allows detecting labels without luminescence properties and presents the possibility of multi-elemental analysis without the necessity to consider spectral overlaps of the conventional luminescence labels. In the following work, we pioneered the application of LIBS for the readout of nanoparticle-labeled ICC sections. The cell pellets were labeled with UCNPs according to our previous report,176 and the characteristic signal of the Y II 437.49 nm emission line was used to construct the 2D map of the sample surface with a resolution of 100 μm. The results demonstrated the ability of LIBS to map the yttrium distribution and showed a clear difference between HER2-positive and HER2-negative cells. The results from LIBS were then compared with upconversion optical microscopy and upconversion luminescence Figure 28: Scheme of LIBS immunoassay with label based on conjugate of Ag NPs with streptavidin. Reprinted from Paper XXI with permission. Copyright 2019 Springer. 46 scanning. The main advantage of microscopy compared to scanning-based approaches is the high resolution, which allows studying the target distribution within cellular structures. However, it is not possible to use conventional optical microscopy to quantitatively determine the amount of label (and therefore indirectly also of the target antigen) within the whole cell pellet. The S/B ratio of LIBS was 5, whereas the upconversion scanning of the identical pellets provided S/B of 159 (Figure 29). Because there was the same amount of UCNPs, the worse S/B of LIBS was probably given by the lower measurement sensitivity. Despite the successful results of the preliminary work, further improvements of LIBS are necessary to meet the practical requirement of IHC, especially in terms of sensitivity and lateral resolution. In the future, LIBS can significantly improve the multiplexing capabilities in IHC due to the possibility of using multiple nanoparticle labels without having to deal with spectral overlaps, as in the case of optical readout. Figure 29: Comparison of (A) LIBS and (B) upconversion scanning of BT-474 and MDA-MB-231 cells with HER2 biomarker labeled with UCNP-PEG-SA conjugate. (C) The average intensities evaluated by the two methods. Error bars correspond to standard deviations of intensities within the cell pellet. Reprinted from Paper XXII. Copyright 2021 Springer. 47 7 Conclusions and Outlook This thesis has summarized the recent progress in the rapidly developing field of immunochemical biosensors and assays. This research topic represents an interdisciplinary effort to combine various kinds of physicochemical transducers with appropriate assay strategies to achieve highly sensitive and specific detection. Furthermore, the use of nanoparticles, in particular with catalytic or luminescent properties, can further enhance the performance of such assays. The label-free biosensing was represented by EIS biosensor for Salmonella in milk, QCM biosensor for aerosolized biological warfare agents, and application of plasmapolymerized layers in SPR biosensing of HSA and Salmonella. Overall, the main advantage of these approaches is the rapid analysis and simple procedure. However, the sensitivity can be limited by the lack of the signal-amplification step. On the other hand, the use of catalytic labels is typically connected with higher sensitivity and reduced effect of the complex sample matrix, but the analysis requires the substrate conversion step and, therefore, longer time. Enzyme-based labels were employed in an amperometric biosensor for the diagnosis of EFB and in precipitation-based SPR assay for Salmonella. As an alternative to enzymes, we have also used catalytic PBNPs in an NLISA for HSA and Salmonella. Even though the sensitivity was similar to the conventional ELISA, PBNPs provide several practical advantages, particularly higher stability and the possibility of easy synthesis from cheap precursors. UCNPs allow highly sensitive detection due to the anti-Stokes luminescence and lack of optical background. We have thoroughly studied the different ways of UCNP surface modification and conjugation with biorecognition molecules. The conjugates were then employed in ULISA assays for a wide range of analytes, from small molecules (DCF, ZEA), through proteins (HSA, PSA, troponin), to bacteria (M. plutonius and P. larvae). UCNPs also proved useful for labeling HER2 biomarker on the surface of breast cancer cells in ICC imaging. Finally, we have explored the possibilities of LIBS as an alternative way of signal readout not dependent on the catalytic or luminescent properties of the labels. We have employed it in microtiter plate-based immunoassay for HSA and in ICC detection of HER2 biomarker. Even though reaching ever lower LODs is one of the main challenges from the scientific point of view, it is equally important to address the simplicity and robustness of the assay procedure. An ideal assay should provide results within a few minutes and be based either on a fully automated system or a cheap disposable sensor that requires minimum manipulation. The ongoing advances in transducer technology and labels, especially based on nanoparticles, promise that the performance of immunochemical biosensors and assays will keep improving in the future. This will be beneficial for the biochemical and biological research field, and for the large community of immunoassay users in clinical diagnosis, detection of biological agents, food safety, and environmental protection. 48 References 1. Tighe, P. J.; Ryder, R. R.; Todd, I.; Fairclough, L. C., ELISA in the multiplex era: Potentials and pitfalls. Proteomics Clin. Appl. 2015, 9 (3–4), 406–422. 2. Farka, Z.; Juřík, T.; Kovář, D.; Trnková, L.; Skládal, P., Nanoparticle-Based Immunochemical Biosensors and Assays: Recent Advances and Challenges. Chem. Rev. 2017, 117 (15), 9973–10042. 3. Peltomaa, R.; Benito-Pena, E.; Moreno-Bondi, M. C., Bioinspired recognition elements for mycotoxin sensors. Anal. Bioanal. Chem. 2018, 410 (3), 747–771. 4. Surugiu, I.; Danielsson, B.; Ye, L.; Mosbach, K.; Haupt, K., Chemiluminescence imaging ELISA using an imprinted polymer as the recognition element instead of an antibody. Anal. Chem. 2001, 73 (3), 487–491. 5. Foote, J.; Eisen, H. N., Kinetic and affinity limits on antibodies produced during immune responses. Proc. Natl. Acad. Sci. U.S.A. 1995, 92 (5), 1254–1256. 6. Green, N. M., Avidin and streptavidin. In Methods Enzymol., Wilchek, M.; Bayer, E. A., Eds. Academic Press: 1990; Vol. 184, pp 51–67. 7. Harmsen, M. M.; De Haard, H. J., Properties, production, and applications of camelid single-domain antibody fragments. Appl. Microbiol. Biotechnol. 2007, 77 (1), 13–22. 8. Yalow, R. S.; Berson, S. A., Assay of Plasma Insulin in Human Subjects by Immunological Methods. Nature 1959, 184 (4699), 1648–1649. 9. Kunze, A.; Pei, L.; Elsasser, D.; Niessner, R.; Seidel, M., High performance concentration method for viruses in drinking water. J. Virol. Methods 2015, 222, 132– 137. 10. Anderson, N. L.; Anderson, N. G., The human plasma proteome – History, character, and diagnostic prospects. Mol. Cell. Proteomics 2002, 1 (11), 845–867. 11. Hanash, S. M.; Pitteri, S. J.; Faca, V. M., Mining the plasma proteome for cancer biomarkers. Nature 2008, 452 (7187), 571–579. 12. Walt, D. R., Clinical testing should be individualized, not based on populations. J. Clin. Invest. 2019, 129 (9), 3472–3473. 13. Rusling, J. F.; Kumar, C. V.; Gutkind, J. S.; Patel, V., Measurement of biomarker proteins for point-of-care early detection and monitoring of cancer. Analyst 2010, 135 (10), 2496– 2511. 14. St John, A.; Price, C. P., Existing and Emerging Technologies for Point-of-Care Testing. Clin. Biochem. Rev. 2014, 35 (3), 155–167. 15. Litman, D. J.; Hanlon, T. M.; Ullman, E. F., Enzyme Channeling Immunoassay – A New Homogeneous Enzyme-Immunoassay Technique. Anal. Biochem. 1980, 106 (1), 223– 229. 16. Hood, L., A Personal Journey of Discovery: Developing Technology and Changing Biology. Annu. Rev. Anal. Chem. 2008, 1, 1–43. 17. Siitari, H.; Hemmila, I.; Soini, E.; Lovgren, T.; Koistinen, V., Detection of hepatitis B surface antigen using time-resolved fluoroimmunoassay. Nature 1983, 301 (5897), 258– 260. 18. Rui, M.; Hampe, C. S.; Wang, C.; Ling, Z. D.; Gorus, F. K.; Lernmark, A.; Pipeleers, D. G.; De Pauw, P. E. M., Species and epitope specificity of two 65 kDa glutamate decarboxylase time-resolved fluorometric immunoassays. J. Immunol. Methods 2007, 319 (1–2), 133–143. 49 19. Pastucha, M.; Farka, Z.; Lacina, K.; Mikušová, Z.; Skládal, P., Magnetic nanoparticles for smart electrochemical immunoassays: a review on recent developments. Microchim. Acta 2019, 186 (5), 26. 20. Gorris, H. H.; Wolfbeis, O. S., Photon-Upconverting Nanoparticles for Optical Encoding and Multiplexing of Cells, Biomolecules, and Microspheres. Angew. Chem. Int. Ed. 2013, 52 (13), 3584–3600. 21. Hofmann, C.; Duerkop, A.; Baeumner, A. J., Nanocontainers for Analytical Applications. Angew. Chem. Int. Ed. 2019, 58 (37), 12840–12860. 22. Mejri, M. B.; Baccar, H.; Baldrich, E.; Del Campo, F. J.; Helali, S.; Ktari, T.; Simonian, A.; Aouni, M.; Abdelghani, A., Impedance biosensing using phages for bacteria detection: Generation of dual signals as the clue for in-chip assay confirmation. Biosens. Bioelectron. 2010, 26 (4), 1261–1267. 23. Daniels, J. S.; Pourmand, N., Label-free impedance biosensors: Opportunities and challenges. Electroanalysis 2007, 19 (12), 1239–1257. 24. Scharff, R. L., Economic Burden from Health Losses Due to Foodborne Illness in the United States. J. Food Prot. 2012, 75 (1), 123–131. 25. Desin, T. S.; Koster, W.; Potter, A. A., Salmonella vaccines in poultry: past, present and future. Expert Rev. Vaccines 2013, 12 (1), 87–96. 26. Majowicz, S. E.; Musto, J.; Scallan, E.; Angulo, F. J.; Kirk, M.; O'Brien, S. J.; Jones, T. F.; Fazil, A.; Hoekstra, R. M., The Global Burden of Nontyphoidal Salmonella Gastroenteritis. Clin. Infect. Dis. 2010, 50 (6), 882–889. 27. Alocilja, E. C.; Radke, S. M., Market analysis of biosensors for food safety. Biosens. Bioelectron. 2003, 18 (5–6), 841–846. 28. Roda, A.; Mirasoli, M.; Roda, B.; Bonvicini, F.; Colliva, C.; Reschiglian, P., Recent developments in rapid multiplexed bioanalytical methods for foodborne pathogenic bacteria detection. Microchim. Acta 2012, 178 (1–2), 7–28. 29. de Boer, E.; Beumer, R. R., Methodology for detection and typing of foodborne microorganisms. Int. J. Food Microbiol. 1999, 50 (1–2), 119–130. 30. Velusamy, V.; Arshak, K.; Korostynska, O.; Oliwa, K.; Adley, C., An overview of foodborne pathogen detection: In the perspective of biosensors. Biotechnol. Adv. 2010, 28 (2), 232–254. 31. Aydintug, M. K.; Inzana, T. J.; Letonja, T.; Davis, W. C.; Corbeil, L. B., Cross-Reactivity of Monoclonal Antibodies to Escherichia coli J5 with Heterologous Gram-Negative Bacteria and Extracted Lipopolysaccharides. J. Infect. Dis. 1989, 160 (5), 846–857. 32. Kothary, M. H.; Babu, U. S., Infective dose of foodborne pathogens in volunteers: A review. J. Food Saf. 2001, 21 (1), 49–73. 33. Sauerbrey, G., Verwendung von Schwingquarzen zur Wägung dünner Schichten und zur Mikrowägung. Z. Phys. 1959, 155 (2), 206–222. 34. Sternbach, G., The history of anthrax. J. Emerg. Med. 2003, 24 (4), 463–467. 35. Török, T. J.; Tauxe, R. V.; Wise, R. P.; Livengood, J. R.; Sokolow, R.; Mauvais, S.; Birkness, K. A.; Skeels, M. R.; Horan, J. M.; Foster, L. R., A Large Community Outbreak of Salmonellosis Caused by Intentional Contamination of Restaurant Salad Bars. JAMA 1997, 278 (5), 389–395. 36. Danzig, R.; Berkowsky, P. B., Why should we be concerned about biological warfare? JAMA 1997, 278 (5), 431–432. 50 37. Khardori, N.; Kanchanapoom, T., Overview of biological terrorism: Potential agents and preparedness. Clin. Microbiol. Newsl. 2005, 27 (1), 1–8. 38. Polymenakou, P. N., Atmosphere: A Source of Pathogenic or Beneficial Microbes? Atmosphere 2012, 3 (1), 87–102. 39. Mandal, S.; Koirala, D.; Selvam, S.; Ghimire, C.; Mao, H. B., A Molecular Tuning Fork in Single-Molecule Mechanochemical Sensing. Angew. Chem. Int. Ed. 2015, 54 (26), 7607–7611. 40. Su, W. C.; Tolchinsky, A. D.; Chen, B. T.; Sigaev, V. I.; Cheng, Y. S., Evaluation of physical sampling efficiency for cyclone-based personal bioaerosol samplers in moving air environments. J. Environ. Monit. 2012, 14 (9), 2430–2437. 41. Joshi, D.; Kumar, D.; Maini, A. K.; Sharma, R. C., Detection of biological warfare agents using ultra violet-laser induced fluorescence LIDAR. Spectrochim. Acta Pt. A Mol. Biomol. Spectrosc. 2013, 112, 446–456. 42. Park, C. W.; Park, J. W.; Lee, S. H.; Hwang, J., Real-time monitoring of bioaerosols via cell-lysis by air ion and ATP bioluminescence detection. Biosens. Bioelectron. 2014, 52, 379–383. 43. Pohlker, C.; Huffman, J. A.; Poschl, U., Autofluorescence of atmospheric bioaerosols – fluorescent biomolecules and potential interferences. Atmos. Meas. Tech. 2012, 5 (1), 37–71. 44. Venkateswaran, K.; Hattori, N.; La Duc, M. T.; Kern, R., ATP as a biomarker of viable microorganisms in clean-room facilities. J. Microbiol. Methods 2003, 52 (3), 367–377. 45. Hospodsky, D.; Yamamoto, N.; Peccia, J., Accuracy, Precision, and Method Detection Limits of Quantitative PCR for Airborne Bacteria and Fungi. Appl. Environ. Microbiol. 2010, 76 (21), 7004–7012. 46. Usachev, E. V.; Agranovski, I. E., Internally controlled PCR system for detection of airborne microorganisms. J. Environ. Monit. 2012, 14 (6), 1631–1637. 47. Alam, S. I.; Kumar, B.; Kamboj, D. V., Multiplex Detection of Protein Toxins Using MALDI-TOF-TOF Tandem Mass Spectrometry: Application in Unambiguous Toxin Detection from Bioaerosol. Anal. Chem. 2012, 84 (23), 10500–10507. 48. Green, H. C.; Field, K. G., Sensitive detection of sample interference in environmental qPCR. Water Res. 2012, 46 (10), 3251–3260. 49. Maher, N.; Dillon, H. K.; Vermund, S. H.; Unnasch, T. R., Magnetic bead capture eliminates PCR inhibitors in samples collected from the airborne environment, permitting detection of Pneumacystis carinii DNA. Appl. Environ. Microbiol. 2001, 67 (1), 449–452. 50. Wijaya, E.; Lenaerts, C.; Maricot, S.; Hastanin, J.; Habraken, S.; Vilcot, J. P.; Boukherroub, R.; Szunerits, S., Surface plasmon resonance-based biosensors: From the development of different SPR structures to novel surface functionalization strategies. Curr. Opin. Solid State Mat. Sci. 2011, 15 (5), 208–224. 51. Guo, X. W., Surface plasmon resonance based biosensor technique: A review. J. Biophotonics 2012, 5 (7), 483–501. 52. Stewart, M. E.; Anderton, C. R.; Thompson, L. B.; Maria, J.; Gray, S. K.; Rogers, J. A.; Nuzzo, R. G., Nanostructured plasmonic sensors. Chem. Rev. 2008, 108 (2), 494–521. 53. Zeng, S. W.; Yu, X.; Law, W. C.; Zhang, Y. T.; Hu, R.; Dinh, X. Q.; Ho, H. P.; Yong, K. T., Size dependence of Au NP-enhanced surface plasmon resonance based on differential phase measurement. Sens. Actuators B Chem. 2013, 176, 1128–1133. 51 54. Zeng, S. W.; Yong, K. T.; Roy, I.; Dinh, X. Q.; Yu, X.; Luan, F., A Review on Functionalized Gold Nanoparticles for Biosensing Applications. Plasmonics 2011, 6 (3), 491–506. 55. Nakamura, C.; Hasegawa, M.; Nakamura, N.; Miyake, J., Rapid and specific detection of herbicides using a self-assembled photosynthetic reaction center from purple bacterium on an SPR chip. Biosens. Bioelectron. 2003, 18 (5–6), 599–603. 56. Wei, J. Y.; Mu, Y.; Song, D. Q.; Fang, X. X.; Liu, X.; Bu, L. S.; Zhang, H. Q.; Zhang, G. Z.; Ding, J. H.; Wang, W. Z.; Jin, Q. H.; Luo, G. M., A novel sandwich immunosensing method for measuring cardiac troponin I in sera. Anal. Biochem. 2003, 321 (2), 209–216. 57. Gobi, K. V.; Tanaka, H.; Shoyama, Y.; Miura, N., Continuous flow immunosensor for highly selective and real-time detection of sub-ppb levels of 2-hydroxybiphenyl by using surface plasmon resonance imaging. Biosens. Bioelectron. 2004, 20 (2), 350–357. 58. Farka, Z.; Juřík, T.; Pastucha, M.; Skládal, P., Enzymatic Precipitation Enhanced Surface Plasmon Resonance Immunosensor for the Detection of Salmonella in Powdered Milk. Anal. Chem. 2016, 88 (23), 11830–11836. 59. Howell, S.; Kenmore, M.; Kirkland, M.; Badley, R. A., High-density immobilization of an antibody fragment to a carboxymethylated dextran-linked biosensor surface. J. Mol. Recognit. 1998, 11 (1–6), 200–203. 60. Wadu-Mesthrige, K.; Amro, N. A.; Liu, G. Y., Immobilization of proteins on selfassembled monolayers. Scanning 2000, 22 (6), 380–388. 61. Vericat, C.; Vela, M. E.; Benitez, G.; Carro, P.; Salvarezza, R. C., Self-assembled monolayers of thiols and dithiols on gold: new challenges for a well-known system. Chem. Soc. Rev. 2010, 39 (5), 1805–1834. 62. Chen, Q.; Forch, R.; Knoll, W., Characterization of pulsed plasma polymerization allylamine as an adhesion layer for DNA adsorption/hybridization. Chem. Mater. 2004, 16 (4), 614–620. 63. Jung, D.; Yeo, S.; Kim, J.; Kim, B.; Jin, B.; Ryu, D. Y., Formation of amine groups by plasma enhanced chemical vapor deposition and its application to DNA array technology. Surf. Coat. Technol. 2006, 200 (9), 2886–2891. 64. Yang, Z. L.; Wang, J.; Luo, R. F.; Maitz, M. F.; Jing, F. J.; Sun, H.; Huang, N., The covalent immobilization of heparin to pulsed-plasma polymeric allylamine films on 316L stainless steel and the resulting effects on hemocompatibility. Biomaterials 2010, 31 (8), 2072–2083. 65. Ren, T. B.; Weigel, T.; Groth, T.; Lendlein, A., Microwave plasma surface modification of silicone elastomer with allylamine for improvement of biocompatibility. J. Biomed. Mater. Res. Part A 2008, 86A (1), 209–219. 66. Nisol, B.; Watson, S.; Lerouge, S.; Wertheimer, M. R., Energetics of Reactions in a Dielectric Barrier Discharge with Argon Carrier Gas: III Esters. Plasma Process. Polym. 2016, 13 (9), 900–907. 67. Rupper, P.; Vandenbossche, M.; Bernard, L.; Hegemann, D.; Heuberger, M., Composition and Stability of Plasma Polymer Films Exhibiting Vertical Chemical Gradients. Langmuir 2017, 33 (9), 2340–2352. 68. Mishra, G.; McArthur, S. L., Plasma Polymerization of Maleic Anhydride: Just What Are the Right Deposition Conditions? Langmuir 2010, 26 (12), 9645–9658. 52 69. Rich, R. L.; Myszka, D. G., Advances in surface plasmon resonance biosensor analysis. Curr. Opin. Biotechnol. 2000, 11 (1), 54–61. 70. Makhneva, E.; Obrusník, A.; Farka, Z.; Skládal, P.; Vandenbossche, M.; Hegemann, D.; Zajíčková, L., Carboxyl-rich plasma polymer surfaces in surface plasmon resonance immunosensing. Jpn. J. Appl. Phys. 2018, 57 (1), 5. 71. Forsgren, E.; Budge, G. E.; Charriere, J. D.; Hornitzky, M. A. Z., Standard methods for European foulbrood research. J. Apic. Res. 2013, 52 (1), 14. 72. Ahmed, A.; Rushworth, J. V.; Hirst, N. A.; Millner, P. A., Biosensors for Whole-Cell Bacterial Detection. Clin. Microbiol. Rev. 2014, 27 (3), 631–646. 73. Pinnock, D. E.; Featherstone, N. E., Detection and quantification of Melissococcus pluton infection in honeybee colonies by means of enzyme-linked immunosorbent assay. J. Apic. Res. 1984, 23 (3), 168–170. 74. Tomkies, V.; Flint, J.; Johnson, G.; Waite, R.; Wilkins, S.; Danks, C.; Watkins, M.; Cuthbertson, A. G. S.; Carpana, E.; Marris, G.; Budge, G.; Brown, M. A., Development and validation of a novel field test kit for European foulbrood. Apidologie 2009, 40 (1), 63–72. 75. Poláchová, V.; Pastucha, M.; Mikušová, Z.; Mickert, M. J.; Hlaváček, A.; Gorris, H. H.; Skládal, P.; Farka, Z., Click-conjugated photon-upconversion nanoparticles in an immunoassay for honeybee pathogen Melissococcus plutonius. Nanoscale 2019, 11 (17), 8343–8351. 76. Vaisocherová-Lísalová, H.; Víšová, I.; Ermini, M. L.; Špringer, T.; Song, X. C.; Mrázek, J.; Lamačová, J.; Scott Lynn Jr, N.; Šedivák, P.; Homola, J., Low-fouling surface plasmon resonance biosensor for multi-step detection of foodborne bacterial pathogens in complex food samples. Biosens. Bioelectron. 2016, 80, 84–90. 77. Liu, X.; Hu, Y.; Zheng, S.; Liu, Y.; He, Z.; Luo, F., Surface plasmon resonance immunosensor for fast, highly sensitive, and in situ detection of the magnetic nanoparticles-enriched Salmonella enteritidis. Sens. Actuators B Chem. 2016, 230, 191– 198. 78. Juřík, T.; Skládal, P., Detection of hydrogen peroxide and glucose by enzyme product precipitation on sensor surface. Chem. Pap. 2015, 69 (1), 167–175. 79. Li, J.; Wang, J. J.; Guo, X.; Zheng, Q.; Peng, J.; Tang, H.; Yao, S. Z., Carbon Nanotubes Labeled with Aptamer and Horseradish Peroxidase as a Probe for Highly Sensitive Protein Biosensing by Postelectropolymerization of Insoluble Precipitates on Electrodes. Anal. Chem. 2015, 87 (15), 7610–7617. 80. Akter, R.; Rahman, M. A.; Rhee, C. K., Amplified Electrochemical Detection of a Cancer Biomarker by Enhanced Precipitation Using Horseradish Peroxidase Attached on Carbon Nanotubes. Anal. Chem. 2012, 84 (15), 6407–6415. 81. Hou, L.; Tang, Y.; Xu, M. D.; Gao, Z. Q.; Tang, D. P., Tyramine-Based Enzymatic Conjugate Repeats for Ultrasensitive Immunoassay Accompanying Tyramine Signal Amplification with Enzymatic Biocatalytic Precipitation. Anal. Chem. 2014, 86 (16), 8352–8358. 82. Hou, L.; Cui, Y. L.; Xu, M. D.; Gao, Z. Q.; Huang, J. X.; Tang, D. P., Graphene oxidelabeled sandwich-type impedimetric immunoassay with sensitive enhancement based on enzymatic 4-chloro-1-naphthol oxidation. Biosens. Bioelectron. 2013, 47, 149–156. 53 83. Ruan, C. M.; Yang, L. J.; Li, Y. B., Immunobiosensor chips for detection of Escherichia coli O157:H7 using electrochemical impedance spectroscopy. Anal. Chem. 2002, 74 (18), 4814–4820. 84. Torun, O.; Boyaci, I. H.; Temur, E.; Tamer, U., Comparison of sensing strategies in SPR biosensor for rapid and sensitive enumeration of bacteria. Biosens. Bioelectron. 2012, 37 (1), 53–60. 85. Fratamico, P. M., Comparison of culture, polymerase chain reaction (PCR), TaqMan Salmonella, and Transia Card Salmonella assays for detection of Salmonella spp. in naturally-contaminated ground chicken, ground turkey, and ground beef. Mol. Cell. Probes 2003, 17 (5), 215–221. 86. Cudjoe, K. S.; Hagtvedt, T.; Dainty, R., Immunomagnetic separation of Salmonella from foods and their detection using immunomagnetic particle (IMP)-ELISA. Int. J. Food Microbiol. 1995, 27 (1), 11–25. 87. Li, Y.; Mustapha, A., Simultaneous detection of Escherichia coli O157:H7, Salmonella, and Shigella in apple cider and produce by a multiplex PCR. J. Food Prot. 2004, 67 (1), 27–33. 88. Rodriguez-Lazaro, D.; D'Agostino, M.; Herrewegh, A.; Pla, M.; Cook, N.; Ikonomopoulos, J., Real-time PCR-based methods for detection of Mycobacterium avium subsp. paratuberculosis in water and milk. Int. J. Food Microbiol. 2005, 101 (1), 93–104. 89. Teunis, P. F. M.; Kasuga, F.; Fazil, A.; Ogden, L. D.; Rotariu, O.; Strachan, N. J. C., Dose-response modeling of Salmonella using outbreak data. Int. J. Food Microbiol. 2010, 144 (2), 243–249. 90. Scott, V. N.; Powell, M.; Cabrera, J.; Carullo, M. E.; Martinez, I.; Lohachoompol, V., Development of microbiological criteria to assess the acceptability of a food lot – An example for milk powder. Food Control 2015, 58, 12–16. 91. Kuah, E.; Toh, S.; Yee, J.; Ma, Q.; Gao, Z. Q., Enzyme Mimics: Advances and Applications. Chem. Eur. J. 2016, 22 (25), 8404–8430. 92. Ragg, R.; Tahir, M. N.; Tremel, W., Solids Go Bio: Inorganic Nanoparticles as Enzyme Mimics. Eur. J. Inorg. Chem. 2016, 2016 (13–14), 1906–1915. 93. Gao, L. Z.; Zhuang, J.; Nie, L.; Zhang, J. B.; Zhang, Y.; Gu, N.; Wang, T. H.; Feng, J.; Yang, D. L.; Perrett, S.; Yan, X., Intrinsic peroxidase-like activity of ferromagnetic nanoparticles. Nat. Nanotechnol. 2007, 2 (9), 577–583. 94. Shokouhimehr, M.; Soehnlen, E. S.; Khitrin, A.; Basu, S.; Huang, S. P. D., Biocompatible Prussian blue nanoparticles: Preparation, stability, cytotoxicity, and potential use as an MRI contrast agent. Inorg. Chem. Commun. 2010, 13 (1), 58–61. 95. Wei, H.; Wang, E. K., Nanomaterials with enzyme-like characteristics (nanozymes): next-generation artificial enzymes. Chem. Soc. Rev. 2013, 42 (14), 6060–6093. 96. Wu, J. J. X.; Wang, X. Y.; Wang, Q.; Lou, Z. P.; Li, S. R.; Zhu, Y. Y.; Qin, L.; Wei, H., Nanomaterials with enzyme-like characteristics (nanozymes): next-generation artificial enzymes (II). Chem. Soc. Rev. 2019, 48 (4), 1004–1076. 97. Doumas, B. T.; Peters, T., Serum and urine albumin: A progress report on their measurement and clinical significance. Clin. Chim. Acta 1997, 258 (1), 3–20. 98. Wang, W. B.; Liu, L. Q.; Song, S. S.; Tang, L. J.; Kuang, H.; Xu, C. L., A Highly Sensitive ELISA and Immunochromatographic Strip for the Detection of Salmonella typhimurium in Milk Samples. Sensors 2015, 15 (3), 5281–5292. 54 99. Haase, M.; Schafer, H., Upconverting Nanoparticles. Angew. Chem. Int. Ed. 2011, 50 (26), 5808–5829. 100. Auzel, F., Upconversion and Anti-Stokes Processes with f and d Ions in Solids. Chem. Rev. 2004, 104 (1), 139–173. 101. Chamarro, M. A.; Cases, R., Energy upconversion in (Yb, Ho) and (Yb, Tm) doped fluorohafnate glasses. J. Lumin. 1988, 42 (5), 267–274. 102. Li, X. M.; Zhang, F.; Zhao, D. Y., Lab on upconversion nanoparticles: optical properties and applications engineering via designed nanostructure. Chem. Soc. Rev. 2015, 44 (6), 1346–1378. 103. Wang, M.; Abbineni, G.; Clevenger, A.; Mao, C. B.; Xu, S. K., Upconversion nanoparticles: synthesis, surface modification and biological applications. Nanomed. Nanotechnol. Biol. Med. 2011, 7 (6), 710–729. 104. Li, Z. Q.; Zhang, Y., Monodisperse silica-coated polyvinylpyrrolidone/NaYF4 nanocrystals with multicolor upconversion fluorescence emission. Angew. Chem. Int. Ed. 2006, 45 (46), 7732–7735. 105. Liu, Z. E.; Wang, J.; Li, Y.; Hu, X. X.; Yin, J. W.; Peng, Y. Q.; Li, Z. H.; Li, Y. W.; Li, B. M.; Yuan, Q., Near-Infrared Light Manipulated Chemoselective Reductions Enabled by an Upconversional Supersandwich Nanostructure. ACS Appl. Mater. Interfaces 2015, 7 (34), 19416–19423. 106. Sedlmeier, A.; Gorris, H. H., Surface modification and characterization of photonupconverting nanoparticles for bioanalytical applications. Chem. Soc. Rev. 2015, 44 (6), 1526–1560. 107. Oaks, J. L.; Gilbert, M.; Virani, M. Z.; Watson, R. T.; Meteyer, C. U.; Rideout, B. A.; Shivaprasad, H. L.; Ahmed, S.; Chaudhry, M. J. I.; Arshad, M.; Mahmood, S.; Ali, A.; Khan, A. A., Diclofenac residues as the cause of vulture population decline in Pakistan. Nature 2004, 427 (6975), 630–633. 108. Koutsouba, V.; Heberer, T.; Fuhrmann, B.; Schmidt-Baumler, K.; Tsipi, D.; Hiskia, A., Determination of polar pharmaceuticals in sewage water of Greece by gas chromatography-mass spectrometry. Chemosphere 2003, 51 (2), 69–75. 109. Heberer, T., Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment: a review of recent research data. Toxicol. Lett. 2002, 131 (1–2), 5–17. 110. Petrovic, M.; Hernando, M. D.; Diaz-Cruz, M. S.; Barcelo, D., Liquid chromatographytandem mass spectrometry for the analysis of pharmaceutical residues in environmental samples: a review. J. Chromatogr. A 2005, 1067 (1–2), 1–14. 111. Deng, A. P.; Himmelsbach, M.; Zhu, Q. Z.; Frey, S.; Sengl, M.; Buchberger, W.; Niessner, R.; Knopp, D., Residue analysis of the pharmaceutical diclofenac in different water types using ELISA and GC-MS. Environ. Sci. Technol. 2003, 37 (15), 3422–3429. 112. Algar, W. R.; Prasuhn, D. E.; Stewart, M. H.; Jennings, T. L.; Blanco-Canosa, J. B.; Dawson, P. E.; Medintz, I. L., The Controlled Display of Biomolecules on Nanoparticles: A Challenge Suited to Bioorthogonal Chemistry. Bioconjugate Chem. 2011, 22 (5), 825– 858. 113. Petrie, B.; Barden, R.; Kasprzyk-Hordern, B., A review on emerging contaminants in wastewaters and the environment: Current knowledge, understudied areas and recommendations for future monitoring. Water Res. 2015, 72, 3–27. 114. Scognarniglio, V.; Arduini, F.; Palleschi, G.; Rea, G., Biosensing technology for sustainable food safety. Trends Anal. Chem. 2014, 62, 1–10. 55 115. Bennett, J. W.; Klich, M., Mycotoxins. Clin. Microbiol. Rev. 2003, 16 (3), 497–516. 116. Zinedine, A.; Soriano, J. M.; Molto, J. C.; Manes, J., Review on the toxicity, occurrence, metabolism, detoxification, regulations and intake of zearalenone: An oestrogenic mycotoxin. Food Chem. Toxicol. 2007, 45 (1), 1–18. 117. Maaroufi, K.; Chekir, L.; Creppy, E. E.; Ellouz, F.; Bacha, H., Zearalenone induces modifications of haematological and biochemical parameters in rats. Toxicon 1996, 34 (5), 535–540. 118. Schothorst, R. C.; van Egmond, H. P., Report from SCOOP task 3.2.10 "Collection of occurrence data of Fusarium toxins in food and assessment of dietary intake by the population of EU member states" – Subtask: Trichothecenes. Toxicol. Lett. 2004, 153 (1), 133–143. 119. Xiong, Y.; Leng, Y. K.; Li, X. M.; Huang, X. L.; Xiong, Y. H., Emerging strategies to enhance the sensitivity of competitive ELISA for detection of chemical contaminants in food samples. Trends Anal. Chem. 2020, 126, 19. 120. Peltomaa, R.; Fikacek, S.; Benito-Pena, E.; Barderas, R.; Head, T.; Deo, S.; Daunert, S.; Moreno-Bondi, M. C., Bioluminescent detection of zearalenone using recombinant peptidomimetic Gaussia luciferase fusion protein. Microchim. Acta 2020, 187 (10), 11. 121. Becer, C. R.; Hoogenboom, R.; Schubert, U. S., Click Chemistry beyond MetalCatalyzed Cycloaddition. Angew. Chem. Int. Ed. 2009, 48 (27), 4900–4908. 122. Roetschi, A.; Berthoud, H.; Kuhn, R.; Imdorf, A., Infection rate based on quantitative real-time PCR of Melissococcus plutonius, the causal agent of European foulbrood, in honeybee colonies before and after apiary sanitation. Apidologie 2008, 39 (3), 362–371. 123. Schottler, S.; Becker, G.; Winzen, S.; Steinbach, T.; Mohr, K.; Landfester, K.; Mailander, V.; Wurm, F. R., Protein adsorption is required for stealth effect of poly(ethylene glycol)and poly(phosphoester)-coated nanocarriers. Nat. Nanotechnol. 2016, 11 (4), 372–377. 124. Nsubuga, A.; Sgarzi, M.; Zarschler, K.; Kubeil, M.; Hubner, R.; Steudtner, R.; Graham, B.; Joshi, T.; Stephan, H., Facile preparation of multifunctionalisable 'stealth' upconverting nanoparticles for biomedical applications. Dalton Trans. 2018, 47 (26), 8595–8604. 125. Gorris, H. H.; Resch-Genger, U., Perspectives and challenges of photon-upconversion nanoparticles – Part II: bioanalytical applications. Anal. Bioanal. Chem. 2017, 409 (25), 5875–5890. 126. Genersch, E., American Foulbrood in honeybees and its causative agent, Paenibacillus larvae. J. Invertebr. Pathol. 2010, 103, S10–S19. 127. Ebeling, J.; Knispel, H.; Hertlein, G.; Funfhaus, A.; Genersch, E., Biology of Paenibacillus larvae, a deadly pathogen of honey bee larvae. Appl. Microbiol. Biotechnol. 2016, 100 (17), 7387–7395. 128. Dobbelaere, W.; De Graaf, D. C.; Reybroeck, W.; Desmedt, E.; Peeters, J. E.; Jacobs, F. J., Disinfection of wooden structures contaminated with Paenibacillus larvae subsp. larvae spores. J. Appl. Microbiol. 2001, 91 (2), 212–216. 129. Stephan, J. G.; de Miranda, J. R.; Forsgren, E., American foulbrood in a honeybee colony: spore-symptom relationship and feedbacks. BMC Ecology 2020, 20 (1), 14. 130. Locke, B.; Low, M.; Forsgren, E., An integrated management strategy to prevent outbreaks and eliminate infection pressure of American foulbrood disease in a commercial beekeeping operation. Prev. Vet. Med. 2019, 167, 48–52. 56 131. Hornitzky, M. A. Z.; Wilson, S. C., A system for the diagnosis of the major bacterial brood diseases of honeybees. J. Apic. Res. 1989, 28 (4), 191–195. 132. De Graaf, D. C.; Alippi, A. M.; Brown, M.; Evans, J. D.; Feldlaufer, M.; Gregorc, A.; Hornitzky, M.; Pernal, S. F.; Schuch, D. M. T.; Titera, D.; Tomkies, V.; Ritter, W., Diagnosis of American foulbrood in honey bees: a synthesis and proposed analytical protocols. Lett. Appl. Microbiol. 2006, 43 (6), 583–590. 133. Sopko, B.; Zitek, J.; Nesvorna, M.; Markovic, M.; Kamler, M.; Titera, D.; Erban, T.; Hubert, J., Detection and quantification of Melissococcus plutonius in honey bee workers exposed to European foulbrood in Czechia through conventional PCR, qPCR, and barcode sequencing. J. Apic. Res. 2020, 59 (4), 503–514. 134. Forsgren, E.; Laugen, A. T., Prognostic value of using bee and hive debris samples for the detection of American foulbrood disease in honey bee colonies. Apidologie 2014, 45 (1), 10–20. 135. Olsen, P. E.; Grant, G. A.; Nelson, D. L.; Rice, W. A., Detection of American foulbrood disease of the honeybee, using a monoclonal antibody specific to Bacillus larvae in an enzyme-linked immunosorbent assay. Can. J. Microbiol. 1990, 36 (10), 732–735. 136. Huse, K.; Böhme, H.-J.; Scholz, G. H., Purification of antibodies by affinity chromatography. J. Biochem. Biophys. Methods 2002, 51 (3), 217–231. 137. Torre, L. A.; Bray, F.; Siegel, R. L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A., Global Cancer Statistics, 2012. CA Cancer J. Clin. 2015, 65 (2), 87–108. 138. Center, M. M.; Jemal, A.; Lortet-Tieulent, J.; Ward, E.; Ferlay, J.; Brawley, O.; Bray, F., International Variation in Prostate Cancer Incidence and Mortality Rates. Eur. Urol. 2012, 61 (6), 1079–1092. 139. Giljohann, D. A.; Mirkin, C. A., Drivers of biodiagnostic development. Nature 2009, 462 (7272), 461–464. 140. Rissin, D. M.; Kan, C. W.; Campbell, T. G.; Howes, S. C.; Fournier, D. R.; Song, L.; Piech, T.; Patel, P. P.; Chang, L.; Rivnak, A. J.; Ferrell, E. P.; Randall, J. D.; Provuncher, G. K.; Walt, D. R.; Duffy, D. C., Single-molecule enzyme-linked immunosorbent assay detects serum proteins at subfemtomolar concentrations. Nat. Biotechnol. 2010, 28 (6), 595–599. 141. Thaxton, C. S.; Elghanian, R.; Thomas, A. D.; Stoeva, S. I.; Lee, J. S.; Smith, N. D.; Schaeffer, A. J.; Klocker, H.; Horninger, W.; Bartsch, G.; Mirkin, C. A., Nanoparticlebased bio-barcode assay redefines "undetectable'' PSA and biochemical recurrence after radical prostatectomy. Proc. Natl. Acad. Sci. U. S. A. 2009, 106 (44), 18437–18442. 142. Liebherr, R. B.; Hutterer, A.; Mickert, M. J.; Vogl, F. C.; Beutner, A.; Lechner, A.; Hummel, H.; Gorris, H. H., Three-in-one enzyme assay based on single molecule detection in femtoliter arrays. Anal. Bioanal. Chem. 2015, 407 (24), 7443–7452. 143. Mickert, M. J.; Farka, Z.; Kostiv, U.; Hlaváček, A.; Horák, D.; Skládal, P.; Gorris, H. H., Measurement of Sub-femtomolar Concentrations of Prostate-Specific Antigen through Single-Molecule Counting with an Upconversion-Linked Immunosorbent Assay. Anal. Chem. 2019, 91 (15), 9435–9441. 144. Salvati, A.; Pitek, A. S.; Monopoli, M. P.; Prapainop, K.; Bombelli, F. B.; Hristov, D. R.; Kelly, P. M.; Aberg, C.; Mahon, E.; Dawson, K. A., Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface. Nat. Nanotechnol. 2013, 8 (2), 137–143. 57 145. Shi, Y.; Shi, B. Y.; Dass, A. V. E.; Lu, Y. Q.; Sayyadi, N.; Kautto, L.; Willows, R. D.; Chung, R.; Piper, J.; Nevalainen, H.; Walsh, B.; Jin, D. Y.; Packer, N. H., Stable Upconversion Nanohybrid Particles for Specific Prostate Cancer Cell Immunodetection. Sci. Rep. 2016, 6, 11. 146. Weber, P. C.; Ohlendorf, D. H.; Wendoloski, J. J.; Salemme, F. R., Structural origins of high-affinity biotin binding to streptavidin. Science 1989, 243 (4887), 85–88. 147. Farka, Z.; Mickert, M. J.; Hlaváček, A.; Skládal, P.; Gorris, H. H., Single Molecule Upconversion-Linked Immunosorbent Assay with Extended Dynamic Range for the Sensitive Detection of Diagnostic Biomarkers. Anal. Chem. 2017, 89 (21), 11825–11830. 148. Anderson, J. L.; Morrow, D. A., Acute Myocardial Infarction. New Engl. J. Med. 2017, 376 (21), 2053–2064. 149. Thygesen, K.; Alpert, J. S.; Jaffe, A. S.; Simoons, M. L.; Chaitman, B. R.; White, H. D., Third Universal Definition of Myocardial Infarction. Circulation 2012, 126 (16), 2020– 2035. 150. Takeda, S.; Yamashita, A.; Maeda, K.; Maeda, Y., Structure of the core domain of human cardiac troponin in the Ca2+ -saturated form. Nature 2003, 424 (6944), 35–41. 151. Jaffe, A. S.; Ravkilde, J.; Roberts, R.; Naslund, U.; Apple, F. S.; Galvani, M.; Katus, H., It's time for a change to a troponin standard. Circulation 2000, 102 (11), 1216–1220. 152. Thygesen, K.; Mair, J.; Katus, H.; Plebani, M.; Venge, P.; Collinson, P.; Lindahl, B.; Giannitsis, E.; Hasin, Y.; Galvani, M.; Tubaro, M.; Alpert, J. S.; Biasucci, L. M.; Koenig, W.; Mueller, C.; Huber, K.; Hamm, C.; Jaffe, A. S., Recommendations for the use of cardiac troponin measurement in acute cardiac care. Eur. Heart J. 2010, 31 (18), 2197– 2204. 153. Westermann, D.; Neumann, J. T.; Sorensen, N. A.; Blankenberg, S., High-sensitivity assays for troponin in patients with cardiac disease. Nat. Rev. Cardiol. 2017, 14 (8), 472– 483. 154. Apple, F. S.; Collinson, P. O., Analytical Characteristics of High-Sensitivity Cardiac Troponin Assays. Clin. Chem. 2012, 58 (1), 54–61. 155. Savukoski, T.; Twarda, A.; Hellberg, S.; Ristiniemi, N.; Wittfooth, S.; Sinisalo, J.; Pettersson, K., Epitope Specificity and IgG Subclass Distribution of Autoantibodies to Cardiac Troponin. Clin. Chem. 2013, 59 (3), 512–518. 156. Katrukha, A. G.; Bereznikova, A. V.; Filatov, V. L.; Esakova, T. V.; Kolosova, O. V.; Pettersson, K.; Lovgren, T.; Bulargina, T. V.; Trifonov, I. R.; Gratsiansky, N. A.; Pulkki, K.; Voipio-Pulkki, L. M.; Gusev, N. B., Degradation of cardiac troponin I: implication for reliable immunodetection. Clin. Chem. 1998, 44 (12), 2433–2440. 157. Panteghini, M., Assay-related issues in the measurement of cardiac troponins. Clin. Chim. Acta 2009, 402 (1–2), 88–93. 158. Apple, F. S.; Sandoval, Y.; Jaffe, A. S.; Ordonez-Llanos, J., Cardiac Troponin Assays: Guide to Understanding Analytical Characteristics and Their Impact on Clinical Care. Clin. Chem. 2017, 63 (1), 73–81. 159. Herman, D. S.; Kavsak, P. A.; Greene, D. N., Variability and Error in Cardiac Troponin Testing An ACLPS Critical Review. Am. J. Clin. Pathol. 2017, 148 (4), 281–295. 160. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R. L.; Torre, L. A.; Jemal, A., Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68 (6), 394–424. 58 161. Gluz, O.; Liedtke, C.; Gottschalk, N.; Pusztai, L.; Nitz, U.; Harbeck, N., Triple-negative breast cancer-current status and future directions. Ann. Oncol. 2009, 20 (12), 1913–1927. 162. Swain, S. M.; Kim, S. B.; Cortes, J.; Ro, J.; Semiglazov, V.; Campone, M.; Ciruelos, E.; Ferrero, J. M.; Schneeweiss, A.; Knott, A.; Clark, E.; Ross, G.; Benyunes, M. C.; Baselga, J., Pertuzumab, trastuzumab, and docetaxel for HER2-positive metastatic breast cancer (CLEOPATRA study): overall survival results from a randomised, double-blind, placebo-controlled, phase 3 study. Lancet Oncol. 2013, 14 (6), 461–471. 163. Ross, J. S.; Slodkowska, E. A.; Symmans, W. F.; Pusztai, L.; Ravdin, P. M.; Hortobagyi, G. N., The HER-2 Receptor and Breast Cancer: Ten Years of Targeted Anti-HER-2 Therapy and Personalized Medicine. Oncologist 2009, 14 (4), 320–368. 164. Titford, M., Progress in the development of microscopical techniques for diagnostic pathology. J. Histotechnol. 2009, 32 (1), 9–19. 165. Nakane, P. K.; Pierce, G. B., Enzyme-Labeled Antibodies: Preparation and Application for the Localization of Antigens. J. Histochem. Cytochem. 1966, 14 (12), 929–931. 166. Susaki, E. A.; Ueda, H. R., Whole-body and Whole-Organ Clearing and Imaging Techniques with Single-Cell Resolution: Toward Organism-Level Systems Biology in Mammals. Cell Chem. Biol. 2016, 23 (1), 137–157. 167. Nayak, T. R.; Andreou, C.; Oseledchyk, A.; Marcus, W. D.; Wong, H. C.; Massague, J.; Kircher, M. F., Tissue factor-specific ultra-bright SERRS nanostars for Raman detection of pulmonary micrometastases. Nanoscale 2017, 9 (3), 1110–1119. 168. Fan, L.; Tian, Y. Y.; Yin, R.; Lou, D. D.; Zhang, X. Z.; Wang, M.; Ma, M.; Luo, S. H.; Li, S. Y.; Gu, N.; Zhang, Y., Enzyme catalysis enhanced dark-field imaging as a novel immunohistochemical method. Nanoscale 2016, 8 (16), 8553–8558. 169. Bera, K.; Schalper, K. A.; Rimm, D. L.; Velcheti, V.; Madabhushi, A., Artificial intelligence in digital pathology – new tools for diagnosis and precision oncology. Nat. Rev. Clin. Oncol. 2019, 16 (11), 703–715. 170. Griffin, J.; Treanor, D., Digital pathology in clinical use: where are we now and what is holding us back? Histopathology 2017, 70 (1), 134–145. 171. Modlitbova, P.; Porizka, P.; Kaiser, J., Laser-induced breakdown spectroscopy as a promising tool in the elemental bioimaging of plant tissues. Trends Anal. Chem. 2020, 122, 10. 172. El Haddad, J.; Canioni, L.; Bousquet, B., Good practices in LIBS analysis: Review and advices. Spectrochim. Acta Part B Atom. Spectr. 2014, 101, 171–182. 173. Škarková, P.; Novotný, K.; Lubal, P.; Jebavá, A.; Pořízka, P.; Klus, J.; Farka, Z.; Hrdlička, A.; Kaiser, J., 2D distribution mapping of quantum dots injected onto filtration paper by laser-induced breakdown spectroscopy. Spectrochim. Acta Part B Atom. Spectr. 2017, 131, 107–114. 174. Modlitbová, P.; Hlaváček, A.; Švestková, T.; Pořízka, P.; Šimoníková, L.; Novotný, K.; Kaiser, J., The effects of photon-upconversion nanoparticles on the growth of radish and duckweed: Bioaccumulation, imaging, and spectroscopic studies. Chemosphere 2019, 225, 723–734. 175. Gimenez, Y.; Busser, B.; Trichard, F.; Kulesza, A.; Laurent, J. M.; Zaun, V.; Lux, F.; Benoit, J. M.; Panczer, G.; Dugourd, P.; Tillement, O.; Pelascini, F.; Sancey, L.; MottoRos, V., 3D Imaging of Nanoparticle Distribution in Biological Tissue by Laser-Induced Breakdown Spectroscopy. Sci. Rep. 2016, 6, 29936. 59 176. Farka, Z.; Mickert, M. J.; Mikusova, Z.; Hlavacek, A.; Bouchalova, P.; Xu, W. S.; Bouchal, P.; Skladal, P.; Gorris, H. H., Surface design of photon-upconversion nanoparticles for high-contrast immunocytochemistry. Nanoscale 2020, 12 (15), 8303– 8313. 60 List of Abbreviations 4-CN 4-chloro-1-naphthol Ab antibody AFB American foulbrood AFM atomic force microscopy AMI acute myocardial infarction BSA bovine serum albumin CFU colony-forming unit CMD carboxymethylated dextran CPA cyclopropylamine DCF diclofenac EDC 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide EFB European foulbrood EIS electrochemical impedance spectroscopy ELISA enzyme-linked immunosorbent assay HRP horseradish peroxidase HSA human serum albumin ICC immunocytochemistry IHC immunohistochemistry LFIA lateral flow immunoassay LIBS laser-induced breakdown spectroscopy LOD limit of detection MIP molecularly imprinted polymer NHS N-hydroxysuccinimide NLISA nanozyme-linked immunosorbent assay PAA poly(acrylic acid) PBNP Prussian blue nanoparticle PCR polymerase chain reaction PoC point-of-care PP plasma polymerization 61 PSA prostate-specific antigen QCM quartz crystal microbalance QD quantum dot SPE screen-printed electrode SPR surface plasmon resonance THFMA tetrahydrofurfuryl methacrylate TVC 1,2,4-trivinylcyclohexane UCNP photon-upconversion nanoparticle ULISA upconversion-linked immunosorbent assay ZEA zearalenone 62 Appendix List of Publications 1. Hlaváček, A.; Farka, Z.*; Mickert, M. J.; Kostiv, U.; Brandmeier, J. C.; Horák, D.; Skládal, P.; Foret, F.; Gorris, H. H., Bioconjugates of photon-upconversion nanoparticles for cancer biomarker detection and imaging. Nat. Protoc. Accepted. 2. Obořilová, R.; Šimečková, H.; Pastucha, M.; Klimovič, Š.; Víšová, I.; Přibyl, J.; Vaisocherová-Lísalová, H.; Pantůček, R.; Skládal, P.; Mašlaňová, I.; Farka Z.*, Atomic force microscopy and surface plasmon resonance for real-time single-cell monitoring of bacteriophage-mediated lysis of bacteria. Nanoscale 2021, 13 (31), 13538–13549. 3. Brandmeier, J. C.; Raiko, K.; Farka, Z.*; Peltomaa, R.; Mickert, M. J.; Hlaváček, A.; Skládal, P.; Soukka, T.; Gorris, H. H., Effect of Particle Size and Surface Chemistry of Photon-Upconversion Nanoparticles on Analog and Digital Immunoassays for Cardiac Troponin. Adv. Healthc. Mater. 2021, 10 (18), 2100506. 4. Pastucha, M.; Odstrčilíková, E.; Hlaváček, A.; Brandmeier, J. C.; Vykoukal, V.; Weisová, J.; Gorris, H. H.; Skládal, P.; Farka Z.*, Upconversion-linked Immunoassay for the Diagnosis of Honeybee Disease American Foulbrood. IEEE J. Sel. Top. Quantum Electron. 2021, 27 (5), 6900311. 5. Pořízka, P.; Vytisková, K.; Obořilová, R.; Pastucha, M.; Gábriš, I.; Brandmeier, J. C.; Modlitbová, P.; Gorris, H. H.; Novotný, K.; Skládal, P.; Kaiser, J.; Farka, Z., LaserInduced Breakdown Spectroscopy as a Readout Method for Immunocytochemistry with Upconversion Nanoparticles. Microchim. Acta 2021, 188, 147. 6. Petruš, O.; Macko, J.; Oriňaková, R.; Oriňak, A.; Múdra, E.; Kupková, M.; Pastucha, M.; Farka, Z.; Socha, V., Detection of organic dyes by surface-enhanced Raman spectroscopy using plasmonic NiAg nanocavity films. Spectrochim. Acta A 2020, 249, 119322. 7. Trnková, L.; Třísková, I.; Čechal, J.; Farka, Z., Polymer pencil leads as a porous nanocomposite graphite material for electrochemical applications: The impact of chemical and thermal treatments. Electrochem. Commun. 2021, 126, 107018. 8. Farka Z.*, Nanoparticle-Based Biosensing: En Route to Ultimate Sensitivity (Meet the Board). Anal. Sens. 2021, 1 (4), 136–137. 9. Farka, Z.; Mickert, M. J.; Pastucha, M.; Mikušová, Z.; Skládal, P.; Gorris, H. H., Advances in Optical Single-Molecule Detection: En Route to Super-Sensitive Bioaffinity Assays. Angew. Chem. Int. Ed. 2020, 59 (27), 10746–10773. (Z.F. and M.J.M. contributed equally) 10. Farka, Z.*; Mickert, M. J.; Mikušová, Z.; Hlaváček, A.; Bouchalová, P.; Xu, W.; Bouchal, P.; Skládal, P.; Gorris, H. H., Surface design of photon-upconversion nanoparticles for high-contrast immunocytochemistry. Nanoscale 2020, 12 (15), 8303– 8313. (Z.F. and M.J.M. contributed equally) 11. Peltomaa, R.; Farka, Z.; Mickert, M. J.; Brandmeier, J. C.; Pastucha, M.; Hlaváček, A.; Martínez-Orts, M.; Canales, Á.; Skládal, P.; Benito-Peña, E.; Moreno-Bondi, M. C.; Gorris, H. H., Competitive upconversion-linked immunoassay using peptide mimetics for the detection of the mycotoxin zearalenone. Biosens. Bioelectron. 2020, 170, 112683. 12. Makhneva, E.; Barillas, L.; Farka, Z.; Pastucha, M.; Skládal, P.; Weltmann, K. D.; Fricke, K., Functional Plasma Polymerized Surfaces for Biosensing. ACS Appl. Mater. Interfaces 2020, 20 (14), 17100–17112. 63 13. Kostiv, U.; Farka, Z.; Mickert, M. J.; Gorris, H. H.; Velychkivska, N.; Pop-Georgievski, O.; Pastucha, M.; Odstrčilíková, E.; Skládal, P.; Horák, D., Versatile bioconjugation strategies of PEG-modified upconversion nanoparticles for bioanalytical applications. Biomacromolecules 2020, 21 (11), 4502–4513. (U.K. and Z.F. contributed equally) 14. Víšová, I.; Smolková, B.; Uzhytchak, M.; Vrabcová, M.; Chafai, D. E.; Houska, M.; Pastucha, M.; Skládal, P.; Farka, Z.*; Dejneka, A.; Vaisocherová-Lísalová, H., Functionalizable Antifouling Coatings as Tunable Platforms for the Stress-Driven Manipulation of Living Cell Machinery. Biomolecules 2020, 10 (8), 1146. 15. Šišoláková, I.; Hovancová, J.; Oriňaková, R.; Oriňak, A.; Trnková, L.; Třísková, I.; Farka, Z.; Pastucha, M.; Radoňák, J., Electrochemical determination of insulin on CuNPs/chitosan-MWCNTs and CoNPs/chitosan-MWCNTs modified screen printed carbon electrodes. J. Electroanal. Chem. 2020, 860, 113881. 16. Poláchová, V.; Pastucha, M.; Mikušová, Z.; Mickert, M. J.; Hlaváček, A.; Gorris, H. H.; Skládal, P.; Farka, Z.*, Click-conjugated photon-upconversion nanoparticles in an immunoassay for honeybee pathogen Melissococcus plutonius. Nanoscale 2019, 11 (17), 8343–8351. 17. Mickert, M. J.; Farka, Z.; Kostiv, U.; Hlaváček, A.; Horák, D.; Skládal, P.; Gorris, H. H., Measurement of Sub-femtomolar Concentrations of Prostate-Specific Antigen through Single-Molecule Counting with an Upconversion-Linked Immunosorbent Assay. Anal. Chem. 2019, 91 (15), 9435–9441. (M.J.M and Z.F. contributed equally) 18. Pastucha, M.; Farka, Z.; Lacina, K.; Mikušová, Z.; Skládal, P., Magnetic nanoparticles for smart electrochemical immunoassays: a review on recent developments. Microchim. Acta 2019, 186, 312. 19. Modlitbová, P.; Farka, Z.; Pastucha, M.; Pořízka, P.; Novotný, K.; Skládal, P.; Kaiser, J., Laser-induced breakdown spectroscopy as a novel readout method for nanoparticlebased immunoassays. Microchim. Acta 2019, 186, 629. 20. Mikušová, Z.; Farka, Z.*; Pastucha, M.; Poláchová, V.; Obořilová, R.; Skládal, P., Amperometric Immunosensor for Rapid Detection of Honeybee Pathogen Melissococcus plutonius. Electroanalysis 2019, 31 (10), 1969–1976. 21. Makhneva, E.; Farka, Z.*; Pastucha, M.; Obrusník, A.; Horáčková, V.; Skládal, P.; Zajíčková, L., Maleic anhydride and acetylene plasma copolymer surfaces for SPR immunosensing. Anal. Bioanal. Chem. 2019, 411 (29), 7689–7697. 22. Farka, Z.*; Čunderlová, V.; Horáčková, V.; Pastucha, M.; Mikušová, Z.; Hlaváček, A.; Skládal, P., Prussian Blue Nanoparticles as a Catalytic Label in a Sandwich NanozymeLinked Immunosorbent Assay. Anal. Chem. 2018, 90 (3), 2348–2354. (Z.F. and V.Č. contributed equally) 23. Makhneva, E.; Farka, Z.; Skládal, P.; Zajíčková, L., Cyclopropylamine plasma polymer surfaces for label-free SPR and QCM immunosensing of Salmonella. Sens. Actuators B Chem. 2018, 276, 447–455. 24. Hegemann, D.; Indutnyi, I.; Zajíčková, L.; Makhneva, E.; Farka, Z.; Ushenin, Y.; Vandenbossche, M., Stable, nanometer‐thick oxygen‐containing plasma polymer films suited for enhanced biosensing. Plasma Process. Polym. 2018, 15 (11), 1800090. 25. Modlitbová, P.; Pořízka, P.; Novotný, K.; Drbohlavová, J.; Chamradová, I.; Farka, Z.; Zlámalová-Gargošová, H.; Romih, T.; Kaiser, J., Short-term assessment of cadmium toxicity and uptake from different types of Cd-based Quantum Dots in the model plant Allium cepa L. Ecotoxicol. Environ. Saf. 2018, 153, 23–31. 64 26. Makhneva, E.; Obrusník, A.; Farka, Z.; Skládal, P.; Vandenbossche, M.; Hegemann, D.; Zajíčková, L., Carboxyl-rich plasma polymer surfaces in surface plasmon resonance immunosensing. Jpn. J. Appl. Phys. 2018, 57 (01AG06), 1–5. 27. Modlitbová, P.; Klepárník, K.; Farka, Z.; Pořízka, P.; Skládal, P.; Novotný, K.; Kaiser, J., Time-Dependent Growth of Silica Shells on CdTe Quantum Dots. Nanomaterials 2018, 8 (6), 439. 28. Farka, Z.; Juřík, T.; Kovář, D.; Trnková, L.; Skládal, P., Nanoparticle-Based Immunochemical Biosensors and Assays: Recent Advances and Challenges. Chem. Rev. 2017, 117 (15), 9973–10042. 29. Farka, Z.; Mickert, M. J.; Hlaváček, A.; Skládal P.; Gorris, H. H., Single Molecule Upconversion-Linked Immunosorbent Assay with Extended Dynamic Range for the Sensitive Detection of Diagnostic Biomarkers. Anal. Chem. 2017, 89 (21), 11825–11830. (Z.F. and M.J.M. contributed equally) 30. Hlaváček, A.; Peterek, M.; Farka, Z.; Mickert, M. J.; Prechtl, L.; Knopp D.; Gorris, H. H., Rapid single-step upconversion-linked immunosorbent assay for diclofenac. Microchim. Acta 2017, 184 (10), 4159–4165. 31. Škarková, P.; Novotný, K.; Lubal, P.; Jebavá, A.; Pořízka, P.; Klus, J.; Farka, Z.; Hrdlička, A.; Kaiser, J., 2D distribution mapping of quantum dots injected onto filtration paper by laser-induced breakdown spectroscopy. Spectrochim. Acta B. At. Spectrosc. 2017, 131, 107–114. 32. Kubesa, O.; Horáčková, V.; Moravec, Z.; Farka, Z.; Skládal, P., Graphene and graphene oxide for biosensing. Monatsh. Chem. 2017, 148 (11), 1937–1944. 33. Kovář, D.; Malá, A.; Mlčochová, J.; Kalina, M.; Fohlerová, Z.; Hlaváček, A.; Farka, Z.; Skládal, P.; Starčuk, Z.; Jiřík, R.; Slabý, O.; Hubálek, J., Preparation and Characterisation of Highly Stable Iron Oxide Nanoparticles for Magnetic Resonance Imaging. J. Nanomater. 2017, 2017 (7859289), 1–8. 34. Trnková, L.; Farka, Z., Advanced nano- and biomaterials in biophysical chemistry (Editorial). Monatsh. Chem. 2017, 148 (11), 1899–1900. 35. Farka, Z.; Juřík, T.; Pastucha, M.; Skládal, P. Enzymatic Precipitation Enhanced Surface Plasmon Resonance Immunosensor for the Detection of Salmonella in Powdered Milk. Anal. Chem. 2016, 88 (23), 11830–11836. 36. Farka, Z.; Juřík, T.; Pastucha, M.; Kovář, D.; Lacina, K.; Skládal, P., Rapid immunosensing of Salmonella Typhimurium using electrochemical impedance spectroscopy: the effect of sample treatment. Electroanalysis 2016, 28 (8), 1803–1809. (Z.F. and T.J. contributed equally) 37. Hlaváček, A.; Farka, Z.; Hübner, M.; Horňáková, V.; Němeček, D.; Skládal, P.; Knopp, D.; Gorris, H. H., Competitive Upconversion-Linked Immunosorbent Assay for the Sensitive Detection of Diclofenac. Anal. Chem. 2016, 88 (11), 6011–6017. 38. Juřík, T.; Podešva, P.; Farka, Z.; Kovář, D.; Skládal, P.; Foret, F., Nanostructured gold deposited in gelatin template applied for electrochemical assay of glucose in serum. Electrochim. Acta 2016, 188, 277–285. 39. Trnková, L.; Farka, Z., Physical and electrochemical aspects of bio- and nanomaterials (Editorial). Monatsh. Chem. 2016, 147 (5), 845. 40. Farka, Z.; Kovář, D.; Skládal, P., Rapid detection of microorganisms based on active and passive modes of QCM. Sensors 2015, 15 (1), 79–92. (Z.F. and D.K. contributed equally) 65 41. Kovář, D.; Farka, Z.; Skládal, P., Detection of aerosolized biological agents using the piezoelectric immunosensor. Anal. Chem. 2014, 86 (17), 8680–8686. (D.K. and Z.F. contributed equally) 42. Farka, Z.; Kovář, D.; Přibyl, J.; Skládal, P., Piezoelectric and surface plasmon resonance biosensors for Bacillus atrophaeus spores. Int. J. Electrochem. Sci. 2013, 8 (1), 100–112. 43. Farka, Z.; Kovář, D.; Skládal, P., Piezoelectric biosensor coupled to cyclone air sampler for detection of microorganisms. Chem. Listy 2013, 107 (S3), s302–s304. * corresponding author 66 Paper I Nanoparticle-Based Immunochemical Biosensors and Assays: Recent Advances and Challenges Farka, Z.; Juřík, T.; Kovář, D.; Trnková, L.; Skládal, P. Chem. Rev. 2017, 117 (15), 9973–10042 DOI: 10.1021/acs.chemrev.7b00037 Contribution: Literature research, manuscript writing Copyright 2017 American Chemical Society. Reprinted with permission. 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 Paper II Advances in Optical Single-Molecule Detection: En Route to Super-Sensitive Bioaffinity Assays Farka, Z.; Mickert, M. J.; Pastucha, M.; Mikušová, Z.; Skládal, P.; Gorris, H. H. (Z.F. and M.J.M. contributed equally) Angew. Chem. Int. Ed. 2020, 59 (27), 10746–10773 DOI: 10.1002/anie.201913924 Contribution: Outline of review, literature research, manuscript writing Copyright 2019 the authors. Reprinted under the permission of Creative Commons Attribution-NonCommercial 4.0 International License. 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 Paper III Rapid immunosensing of Salmonella Typhimurium using electrochemical impedance spectroscopy: the effect of sample treatment Farka, Z.; Juřík, T.; Pastucha, M.; Kovář, D.; Lacina, K.; Skládal, P. (Z.F. and T.J. contributed equally) Electroanalysis 2016, 28 (8) 1803–1809 DOI: 10.1002/elan.201600093 Contribution: Design of experiments, development and optimization of EIS immunosensor, characterization of sensing surface by AFM, data evaluation, manuscript writing Copyright 2016 Wiley-VCH. Reprinted with permission. 167 168 169 170 171 172 173 174 Paper IV Detection of aerosolized biological agents using the piezoelectric immunosensor Kovář, D.; Farka, Z.; Skládal, P. (D.K. and Z.F. contributed equally) Anal. Chem. 2014, 86 (17), 8680–8686 DOI: 10.1021/ac501623m Contribution: Development and optimization of QCM immunosensor, data evaluation, manuscript writing Copyright 2014 American Chemical Society. Reprinted with permission. 175 176 177 178 179 180 181 182 Paper V Cyclopropylamine plasma polymer surfaces for label-free SPR and QCM immunosensing of Salmonella Makhneva, E.; Farka, Z.; Skládal, P.; Zajíčková, L. Sens. Actuators B Chem. 2018, 276, 447–455 DOI: 10.1016/j.snb.2018.08.055 Contribution: Development and optimization of SPR and QCM immunosensors, characterization of sensing surface by AFM, data evaluation, participation in manuscript writing Copyright 2018 Elsevier. Reprinted with permission. 183 184 185 186 187 188 189 190 191 192 Paper VI Maleic anhydride and acetylene plasma copolymer surfaces for SPR immunosensing Makhneva, E.; Farka, Z.*; Pastucha, M.; Obrusník, A.; Horáčková, V.; Skládal, P.; Zajíčková, L. Anal. Bioanal. Chem. 2019, 411 (29), 7689–7697 DOI: 10.1007/s00216-019-01979-9 Contribution: Design of experiments, development and optimization of SPR immunosensor, characterization of sensing surface by AFM, data evaluation, manuscript writing Copyright 2019 Springer. Reprinted with permission. 193 194 195 196 197 198 199 200 201 202 Paper VII Functional Plasma Polymerized Surfaces for Biosensing Makhneva, E.; Barillas, L.; Farka, Z.; Pastucha, M.; Skládal, P.; Weltmann, K. D.; Fricke, K. ACS Appl. Mater. Interfaces 2020, 20 (14), 17100–17112 DOI: 10.1021/acsami.0c01443 Contribution: Development and optimization of SPR immunosensor, data evaluation, participation in manuscript writing Copyright 2020 American Chemical Society. Reprinted with permission. 203 204 205 206 207 208 209 210 211 212 213 214 215 216 Paper VIII Amperometric Immunosensor for Rapid Detection of Honeybee Pathogen Melissococcus plutonius Mikušová, Z.; Farka, Z.*; Pastucha, M.; Poláchová, V.; Obořilová, R.; Skládal, P. Electroanalysis 2019, 31 (10), 1969–1976 DOI: 10.1002/elan.201900252 Contribution: Design of experiments, preparation of immunization antigen and antibody, optimization of electrochemical immunosensor, data evaluation, manuscript writing Copyright 2019 Wiley-VCH. Reprinted with permission. 217 218 219 220 221 222 223 224 225 Paper IX Enzymatic Precipitation Enhanced Surface Plasmon Resonance Immunosensor for the Detection of Salmonella in Powdered Milk Farka, Z.; Juřík, T.; Pastucha, M.; Skládal, P. Anal. Chem. 2016, 88 (23), 11830–11836 DOI: 10.1021/acs.analchem.6b03511 Contribution: Design of experiments, development and optimization of precipitation-enhanced SPR assay, characterization of precipitation reaction by AFM, data evaluation, manuscript writing Copyright 2016 American Chemical Society. Reprinted with permission. 226 227 228 229 230 231 232 233 Paper X Prussian Blue Nanoparticles as a Catalytic Label in a Sandwich Nanozyme-Linked Immunosorbent Assay Farka, Z.*; Čunderlová, V.; Horáčková, V.; Pastucha, M.; Mikušová, Z.; Hlaváček, A.; Skládal, P. (Z.F. and V.Č. contributed equally) Anal. Chem. 2018, 90 (3), 2348–2354 DOI: 10.1021/acs.analchem.7b04883 Contribution: Design of experiments, bioconjugation and characterization of PBNPs, development and optimization of sandwich assay, data evaluation, manuscript writing Copyright 2018 American Chemical Society. Reprinted with permission. 234 235 236 237 238 239 240 241 Paper XI Competitive Upconversion-Linked Immunosorbent Assay for the Sensitive Detection of Diclofenac Hlaváček, A.; Farka, Z.; Hübner, M.; Horňáková, V.; Němeček, D.; Skládal, P.; Knopp, D.; Gorris, H. H. Anal. Chem. 2016, 88 (11), 6011–6017 DOI: 10.1021/acs.analchem.6b01083 Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of competitive immunoassay, data evaluation, participation in manuscript writing Copyright 2016 American Chemical Society. Reprinted with permission. 242 243 244 245 246 247 248 249 Paper XII Rapid single-step upconversion-linked immunosorbent assay for diclofenac Hlaváček, A.; Peterek, M.; Farka, Z.; Mickert, M. J.; Prechtl, L.; Knopp D.; Gorris, H. H. Microchim. Acta 2017, 184 (10), 4159–4165 DOI: 10.1007/s00604-017-2456-0 Contribution: Development and optimization of competitive immunoassay, data evaluation, participation in manuscript writing Copyright 2017 Springer. Reprinted with permission. 250 251 252 253 254 255 256 257 Paper XIII Competitive upconversion-linked immunoassay using peptide mimetics for the detection of the mycotoxin zearalenone Peltomaa, R.; Farka, Z.; Mickert, M. J.; Brandmeier, J. C.; Pastucha, M.; Hlaváček, A.; Martínez-Orts, M.; Canales, Á.; Skládal, P.; Benito-Peña, E.; Moreno-Bondi, M. C.; Gorris, H. H. Biosens. Bioelectron. 2020, 170, 112683 DOI: 10.1016/j.bios.2020.112683 Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of competitive immunoassay, data evaluation, participation in manuscript writing Copyright 2020 Elsevier. Reprinted with permission. 258 259 260 261 262 263 264 265 266 267 Paper XIV Click-conjugated photon-upconversion nanoparticles in an immunoassay for honeybee pathogen Melissococcus plutonius Poláchová, V.; Pastucha, M.; Mikušová, Z.; Mickert, M. J.; Hlaváček, A.; Gorris, H. H.; Skládal, P.; Farka, Z.* Nanoscale 2019, 11 (17), 8343–8351 DOI: 10.1039/C9NR01246J Contribution: Design of experiments, preparation of immunization antigen and antibody, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing Copyright 2019 Royal Society of Chemistry. Reprinted with permission. 268 269 270 271 272 273 274 275 276 277 Paper XV Versatile bioconjugation strategies of PEG-modified upconversion nanoparticles for bioanalytical applications Kostiv, U.; Farka, Z.; Mickert, M. J.; Gorris, H. H.; Velychkivska, N.; PopGeorgievski, O.; Pastucha, M.; Odstrčilíková, E.; Skládal, P.; Horák, D. (U.K. and Z.F. contributed equally) Biomacromolecules 2020, 21 (11), 4502–4513 DOI: 10.1021/acs.biomac.0c00459 Contribution: Design of experiments, bioconjugation of UCNPs, development and optimization of sandwich immunoassay, data evaluation, participation in manuscript writing Copyright 2020 American Chemical Society. Reprinted with permission. 278 279 280 281 282 283 284 285 286 287 288 289 290 Paper XVI Upconversion-linked Immunoassay for the Diagnosis of Honeybee Disease American Foulbrood Pastucha, M.; Odstrčilíková, E.; Hlaváček, A.; Brandmeier, J. C.; Vykoukal, V.; Weisová, J.; Gorris, H. H.; Skládal, P.; Farka Z.* IEEE J. Sel. Top. Quantum Electron. 2021, 27 (5), 6900311. DOI: 10.1109/JSTQE.2021.3049689 Contribution: Design of experiments, preparation of immunization antigen and antibody, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing Copyright 2021 IEEE. Reprinted with permission. 291 292 293 294 295 296 297 298 299 300 301 302 Paper XVII Single Molecule Upconversion-Linked Immunosorbent Assay with Extended Dynamic Range for the Sensitive Detection of Diagnostic Biomarkers Farka, Z.; Mickert, M. J.; Hlaváček, A.; Skládal P.; Gorris, H. H. (Z.F. and M.J.M. contributed equally) Anal. Chem. 2017, 89 (21), 11825–11830 DOI: 10.1021/acs.analchem.7b03542 Contribution: Design of experiments, optimization of single-particle microscope setup, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing Copyright 2017 American Chemical Society. Reprinted with permission. 303 304 305 306 307 308 309 Paper XVIII Measurement of Sub-femtomolar Concentrations of ProstateSpecific Antigen through Single-Molecule Counting with an Upconversion-Linked Immunosorbent Assay Mickert, M. J.; Farka, Z.; Kostiv, U.; Hlaváček, A.; Horák, D.; Skládal, P.; Gorris, H. H. (M.J.M and Z.F. contributed equally) Anal. Chem. 2019, 91 (15), 9435–9441 DOI: 10.1021/acs.analchem.9b02872 Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing Copyright 2017 American Chemical Society. Reprinted with permission. 310 311 312 313 314 315 316 317 Paper XIX Effect of Particle Size and Surface Chemistry of PhotonUpconversion Nanoparticles on Analog and Digital Immunoassays for Cardiac Troponin Brandmeier, J. C.; Raiko, K.; Farka, Z.*; Peltomaa, R.; Mickert, M. J.; Hlaváček, A.; Skládal, P.; Soukka, T.; Gorris, H. H. Adv. Healthc. Mater. 2021, 10 (18), 2100506 DOI: 10.1002/adhm.202100506 Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of sandwich immunoassay, data evaluation, manuscript writing Copyright 2021 the authors. Reprinted under the permission of Creative Commons Attribution-NonCommercial 4.0 International License. 318 319 320 321 322 323 324 325 326 327 Paper XX Surface design of photon-upconversion nanoparticles for highcontrast immunocytochemistry Farka, Z.*; Mickert, M. J.; Mikušová, Z.; Hlaváček, A.; Bouchalová, P.; Xu, W.; Bouchal, P.; Skládal, P.; Gorris, H. H. (Z.F. and M.J.M. contributed equally) Nanoscale 2020, 12 (15), 8303–8313 DOI: 10.1039/C9NR10568A Contribution: Design of experiments, optimization of microscope setup, bioconjugation and characterization of UCNPs, development and optimization of ICC assay, data evaluation, manuscript writing Copyright 2020 Royal Society of Chemistry. Reprinted with permission. 328 329 330 331 332 333 334 335 336 337 338 339 Paper XXI Laser-induced breakdown spectroscopy as a novel readout method for nanoparticle-based immunoassays Modlitbová, P.; Farka, Z.; Pastucha, M.; Pořízka, P.; Novotný, K.; Skládal, P.; Kaiser, J. Microchim. Acta 2019, 186, 629 DOI: 10.1007/s00604-019-3742-9 Contribution: Design of experiments, development and optimization of sandwich immunoassay, data evaluation, participation in manuscript writing Copyright 2019 Springer. Reprinted with permission. 340 341 342 343 344 345 346 347 348 349 350 Paper XXII Laser-Induced Breakdown Spectroscopy as a Readout Method for Immunocytochemistry with Upconversion Nanoparticles Pořízka, P.; Vytisková, K.; Obořilová, R.; Pastucha, M.; Gábriš, I.; Brandmeier, J. C.; Modlitbová, P.; Gorris, H. H.; Novotný, K.; Skládal, P.; Kaiser, J.; Farka, Z. Microchim. Acta 2021, 188, 147. DOI: 10.1007/s00604-021-04816-y Contribution: Design of experiments, bioconjugation and characterization of UCNPs, development and optimization of ICC assay, data evaluation, manuscript writing Copyright 2021 Springer. Reprinted with permission. 351 352 353 354 355 356 357 358 359 360