Open access peer-reviewed chapter

Coumarins as Fluorescent Labels of Biomolecules

Written By

António Pereira, Sérgio Martins and Ana Teresa Caldeira

Submitted: 27 February 2019 Reviewed: 20 March 2019 Published: 07 June 2019

DOI: 10.5772/intechopen.85973

From the Edited Volume

Phytochemicals in Human Health

Edited by Venketeshwer Rao, Dennis Mans and Leticia Rao

Chapter metrics overview

1,497 Chapter Downloads

View Full Metrics

Abstract

Important areas such as environmental sciences, medicine, pharmacy, and cellular biology are dependent on very sensitive analytical techniques. One of the most common methodologies used for their bioanalytical purposes is the fluorescent labelling. The synthesis of new fluorophores and the great development of fluorescent-labelling techniques combined with the enormous technological advances in the field of fluorescence microscopy allowed to deepen the structural knowledge of biomolecules. This new organic fluorophores form covalent bonds with the sample to be analyzed, producing stable bioconjugates that show fluorescence in a wide range of wavelengths, depending on the label used. Coumarin derivatives represent one of the most important chemical classes of organic fluorescent materials being one of the most extensively investigated and commercially significant groups of organic fluorescent materials. In this chapter, it is reviewed the use of fluorescent coumarin derivatives and their application to labelling biomolecules. These fluorescent labels allow researchers to study, and understand, biomolecular assemblies that exhibit complex sensitivity and selectivity. Reactive fluorescent coumarin derivatives are actually widely used in labelling biomolecules as peptides, proteins, oligonucleotides, nucleic acids, and carbohydrates, among other biological molecules.

Keywords

  • coumarins
  • fluorophores
  • labelling
  • biomolecules
  • bioconjugation

1. Introduction

Important areas such as environmental sciences, medicine, medicinal chemistry, and cellular biology are dependent on very sensitive analytical techniques to detect and track biomolecules (amino acids, peptides, proteins, antibodies, oligonucleotides, nucleic acids, carbohydrates, and other biological molecules). Many of these techniques often require labelling with reporters or sensors, such as isotope labels [1], radioactive tracers [2], colorimetric biosensors [3], photoswitchable biomaterials [4], photochromic compounds [5, 6], electrochemical sensors [7], or fluorescent labels [8, 9]. The fluorescent labelling presents numerous advantages, when compared to the other techniques, due to the high sensitivity of the fluorescence technique and also due to its non-destructive nature that allows the use of small sample quantities and their fluorescent labels. The fluorescence process occurs in certain molecules called fluorophores or fluorescent dyes, and a fluorescent probe is nothing more than a fluorophore enabled to detect particular components of complex biomolecular assemblies, including live cells, with complex sensitivity and selectivity [10]. The organic fluorophores may form covalent or non-covalent linkages with the sample to be analyzed, producing the respective bioconjugates (or complexes) that can show fluorescence, from short to very long wavelengths, depending on the label used. The bioconjugation technique depends on two interrelated chemistries: the reactive functionality present on the fluorescent label and the functional groups present on the target biomolecules to be labeled. The knowledge of the basic mechanisms by which the reactive groups couple to target functionalities provides the means to intelligently design the bioconjugation strategy. Choosing the correct fluorescent label that can react with the chemical groups available on target biomolecules forms the basis for successful labelling [11].

In general, the fluorescent label should be small in size and chemically stable, with minimal interference on the structure and biological functions of the unlabeled biomolecules, producing high fluorescence quantum yield bioconjugates.

On the other hand, the labelling reaction should be extremely efficient with high yields, preferably establishing a stable covalent linkage between the fluorescent label and a specific residue in the target biomolecule. The efficiency and selectivity of several fluorescent-labeled biomolecules have been used to study and understand their dynamics, kinetics, and photophysical properties [12, 13, 14, 15, 16, 17, 18].

The amine reactive fluorescent labels are the most frequently used to prepare stable bioconjugates to a great number of biological applications since amino groups are either abundant or easily introduced into biomolecules. In contrast, to study some particular protein structures and functions, thiol-reactive reagents are chosen due to the smaller presence of thiol groups, when compared with lysine, in biomolecules [19]. In this context, cysteine is generally the amino acid chosen to label when it is desired to label selectively a protein in vitro, due to its relatively low abundance and high nucleophilicity compared to other amino acid side chains. Specific and noninterfering dual fluorescent labelling in a peptide or protein molecule allows conformational investigations in terms of intramolecular distances [20].

The expeditious development of the fluorescent-labelling techniques allowed to explore and discover several cellular functions. To study, and understand, the activity of signal transduction by visualizing protein binding or folding, the fluorescence correlation spectroscopy (FCS) and the fluorescence resonance energy transfer (FRET) are widely used [21]. Molecular tags that specifically bind to particular membrane-permeable dyes [22] allow to study protein dynamics and trafficking by fluorescence recovery after photobleaching (FRAP) as well the protein turnover [23, 24].

The great development of fluorescent-labelling techniques combined with the enormous technological advances in the field of fluorescence microscopy allowed to study, in vivo and in vitro systems, the protein distribution as well as their translocation and their interactions [25]. With specific and efficient fluorescent labelling, the proteins can be visualized in real time for the elucidation of their functions in a complex biological network, which also allows the detection of the protein-protein interactions, fundamental to understand intra- and intercellular communications [26].

Coumarins (benzopyranones or 2H-chromen-2-ones), whether natural products or synthetic ones, have also aroused a growing interest of the scientific community in the last decades due to their very significant pharmacological activity [27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37]. The nature and substitution pattern in the coumarins grant them diversified and exceptional optical properties with high fluorescence quantum yields [38]. Coumarins constitute the major class of fluorescent dyes [39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63], used as fluorescent labels and probes for physiological measurement [43, 44, 45, 46, 47], fluorescent whiteners [48], optical brighteners [49, 50], nonlinear optical chromophores [51, 52, 53], emission layers in organic light-emitting diodes (OLED) [54, 55, 56, 57], and more recently, in caging [58, 59, 60, 61], and labelling [62, 63]. Due to strong blue fluorescence of coumarin, it is easy to distinguish its light from green, yellow, and red, an enormous advantage in multicolored fluorescence investigation. Developments from the last decade show that the introduction of appropriated substituents into the coumarin ring contributes to structures with improved photophysical and spectroscopic properties [64, 65, 66]. The synthesis of new fluorophores, with absorption and emission at long wavelengths, is of extreme importance for biological purposes, and the coumarins may play a leading role in this field.

Advertisement

2. Chemical labelling

Of all different fluorescent-labelling techniques, the chemical labelling is actually one of the most used as it allows novel types of experiments in biomolecules using a wider range of reactive fluorescent chromophores available. The covalent attachment of the chemical probes with specific amino acid has the advantage of being an irreversible process when compared to the non-covalent binding [67]. The chemical labelling methodology produces very stable bioconjugates, easy to manipulate with high efficiency, in a great number of available fluorophores that can be coupled covalently to the target biomolecule. Chemical labelling methods produce better results in in vitro studies rather than in vivo [18]. The most used methods in chemical labelling, in the biomolecules’ native functional groups, under mild aqueous conditions, and using fluorescent coumarins, are discussed below.

2.1 Amine reactive fluorescent coumarins

Presently, amine reactive fluorescent coumarins are widely used to label biomolecules, as peptides, proteins, oligonucleotides, and nucleic acids, among others. The fluorescent bioconjugates obtained are very useful in fluorescence in situ hybridization (FISH), receptor labelling immunochemistry, cell tracing, and fluorescent analog cytochemistry studies. Almost all of the techniques used in these tests implicate a robust fluorescent conjugate able to support rigorous incubation, hybridization, and washing steps, which is provided by the stability of the covalent bond between the amine reactive dye and biomolecule. Chemically, the amine labelling reaction proceeds usually through acylation pathway producing stable amide (or thiourea) bonds. The “ideal” reactions are those which require the same conditions as proteins, like functional group tolerance, compatibility, selectivity, water as solvent (or pH ~ 7), room temperature, high reaction rates, low reactant concentration, and nontoxic reagents.

A number of fluorescent amino-reactive coumarins have been developed to label various biomolecules, and the resultant conjugates are widely used in biological applications. Four major classes of amine-reactive fluorescent reagents are currently used to label biomolecules: succinimidyl esters (SE), 4-sulfotetrafluorophenyl (STP) esters, sulfonyl chlorides, and isothiocyanates [68]. Figure 1 represents, in a general schematic diagram, the referred labelling reactions, between an amine group of a biomolecule and a fluorescent amino-reactive coumarin.

Figure 1.

Schematic diagram of amine labelling techniques using succinimidyl esters (A), 4-sulfotetrafluorophenyl esters (B), sulfonyl chlorides (C), and isothiocyanates (D).

2.1.1 Fluorescent coumarin succinimidyl esters

Succinimidyl esters (SE) are proven to be very good reagents for amine modifications. These kinds of reagents are generally stable and show good reactivity and selectivity with aliphatic amines, such as the amine group of lysine side chain. Some of these kinds of reactive dyes are hydrophobic molecules and should be previously dissolved in anhydrous dimethylformamide (DMF) or dimethylsulfoxide (DMSO), but the sulfo-succinimidyl esters are water soluble. The amine labelling reaction with succinimidyl esters has a handicap, due to its great pH dependence. Succinimidyl esters react with non-protonated aliphatic amine groups, and the amine acylation reaction must be carried out at pH > 7.5. In the specific case of protein labelling by succinimidyl esters, the reactions require a pH between 7.5 and 8.5. Buffers used in labelling reactions shall not contain nucleophilic compounds because they may react with the labelling reagent to form unstable intermediates that could destroy the reactive dye. Most conjugations are done at room temperature, but either high or low temperature may be required for a particular labelling reaction. Some of the fluorescent coumarin succinimidyl esters contain a seven-atom aminohexanoyl spacer between the fluorophore and the reactive group, providing better solubility and spatial separation between the fluorophore and the target molecule being labeled. This separation potentially reduces the quenching that typically occurs upon conjugation and makes the dye more available for recognition by secondary detection reagents [68]. The most important fluorescent coumarin succinimidyl esters used for labelling biomolecules are shown in Table 1, as the corresponding values of maximal excitation (Ex) and emission (Em) wavelengths and their physicochemical features and biological applications [19, 68].

Table 1.

Fluorescent coumarin succinimidyl esters used for biomolecule labelling.

2.1.2 Fluorescent coumarin 4-sulfotetrafluorophenyl (STP) esters

Some succinimidyl esters may not be compatible with a specific application due to their insolubility in aqueous solution. To overcome these limitations, the 4-sulfotetrafluorophenyl (STP) ester can be used. These sulfonated esters have higher water solubility than simple succinimidyl esters and sometimes eliminate the need for organic solvents in the conjugation reaction, which is a great advantage to maintain the native characteristics of biomolecules. They are, however, more polar than succinimidyl esters, which makes them less likely to react with buried amines in proteins or to penetrate cell membranes [68, 94]. Table 2 presents the single fluorescent coumarin 4-sulfotetrafluorophenyl (STP) ester used for labelling biomolecules, as the corresponding values of maximal excitation (Ex) and emission (Em) wavelengths and their physicochemical features and biological applications [95, 96].

Table 2.

Fluorescent coumarin 4-sulfotetrafluorophenyl (STP) ester used for biomolecule labelling.

2.1.3 Fluorescent coumarin sulfonyl chlorides

Sulfonyl chlorides (SC) are highly reactive and are unstable in water, especially at high pH required for reaction with aliphatic amines. The labelling reactions with sulfonyl chlorides must be performed, carefully, at very low temperature in a place with local exhaust ventilation. Sulfonyl chlorides present a major reactive handicap as they can also easily react with other reactive groups present in biomolecules as phenols, thiols, aliphatic alcohols, imidazoles, and many others. Fortunately, this kind of reactions rarely occurs in proteins or in aqueous solution, allowing the use of this type of chromophores to label proteins. Sulfonyl chloride dyes are generally hydrophobic molecules and should be dissolved in anhydrous dimethylformamide (DMF), but never in dimethylsulfoxide (DMSO) due to their highly instability in this solvent.

The labelling reactions of amines with SC reagents are strongly pH dependent, and the sulfonylation-based conjugations may require a pH 9.0–10.0 for optimal conjugations, which potentiates the sulfonyl chlorides’ degradation by hydrolysis reactions. In general, sulfonylation-based conjugations have much lower yields than the succinimidyl ester-based conjugations. As in the case of succinimidyl esters, the buffers used in sulfonyl chloride reactions shall not contain nucleophilic compounds, because they may react with the labelling reagent to form unstable intermediates that could destroy the reactive dye [19, 97, 98, 99]. Table 3 shows fluorescent coumarin sulfonyl chlorides used for labelling biomolecules, as the corresponding values of maximal excitation (Ex) and emission (Em) wavelengths and their physicochemical features and biological applications. In addition to the coumarins presented in Table 3, new sulfonyl chloride coumarins have been developed, with high potential as fluorescent probes [100, 101].

Table 3.

Fluorescent coumarin sulfonyl chlorides used for biomolecule labelling.

2.1.4 Fluorescent coumarin isothiocyanates

Isothiocyanates form thioureas upon reaction with amines, but some thiourea products are much less stable than the conjugates that are prepared from the corresponding succinimidyl esters. Most part of isothiocyanate-reactive dyes are hydrophobic molecules and should be dissolved either in anhydrous dimethylformamide (DMF) or in dimethylsulfoxide (DMSO), and their reactions may require a pH 9.0–10.0 for optimal conjugations. As in the previous cases, the buffers used shall not contain nucleophilic compounds. The isothiocyanate conjugations are done at room temperature, but either high or low temperature may be required for a particular labelling reaction [19, 102]. The unique fluorescent coumarin isothiocyanate used for labelling biomolecules is shown in Table 4, but new isothiocyanate coumarins have been synthesized, with high potential as fluorescent probes [103, 104].

Table 4.

Fluorescent coumarin isothiocyanate used for biomolecule labelling.

2.2 Thiol-reactive fluorescent coumarins

Cysteine is, in comparison with lysine, a rare amino acid present in biomolecules, and, for this reason, thiol-reactive reagents are used to label selectively a biomolecule at a defined site, probing their function, interaction, and biological structure. A great number of thiol-reactive dyes have been developed to analyze the proteins’ topography in biological membranes, to measure the distances within (or between) proteins, and to observe and understand the changes in protein conformation using environmental sensitive probes.

Maleimides and iodoacetamides are the principal types of thiol-reactive coumarin dyes reported in the literature. Despite many similarities in their reactivity and selectivity toward thiol-reactive moieties, maleimides have a great advantage in relation to iodoacetamides, due to their high stability, solubility in simple solvent mixtures, and their high reactivity in the neutral pH range. Air oxidation of thiol compounds (to disulfides) is a major competing reaction for the iodoacetamide modifications of thiol compounds [18, 19, 105]. Due to the disinterest on the development of new coumarin iodoacetamides, for the above reasons, only the fluorescent coumarin maleimides will be focused in this section. Figure 2 represents, in a general schematic diagram, the thiol-labelling reaction with fluorescent coumarin maleimides.

Figure 2.

Schematic diagram of thiol-labelling technique using maleimides.

2.2.1 Fluorescent coumarin maleimides

Maleimides readily react with thiol moieties of biomolecules to form thioether conjugates even under neutral conditions. The thioether bond formed is quite stable and is known to be responsible for the light produced, especially in the solution. Maleimides require conjugation conditions less rigorous than those of iodoacetamides and do not react with histidine and methionine under physiological conditions. Most labelling reactions can be done at room temperature at neutral pH. However, either elevated or reduced pH or temperature may be required for a particular labelling reaction [18, 19, 68]. In Table 5, the most important fluorescent coumarin maleimides used for labelling biomolecules are presented, as the corresponding values of maximal excitation (Ex) and emission (Em) wavelengths and their physicochemical features and biological applications.

Table 5.

Fluorescent coumarin maleimides used for biomolecule labelling.

2.3 Tyrosine-reactive fluorescent coumarins

The hydroxyl groups of the amino acids can be labeled with the same reagents used for the lysine residues, but the labelling reaction is carried out in organic solvent, like anhydrous dimethylformamide (DMF) or dimethylsulfoxide (DMSO), which absorbs the formed water molecule avoiding possible hydrolysis reactions. The amino acid hydroxyl groups do not allow highly specific labelling reactions due to the existence of several hydroxyl groups in biomolecules (serine, threonine, and tyrosine) [113].

One of the well-known labelling methods is the reaction with diazonium salts resulting in the formation of azo compounds, as 4-trifluoromethylcoumarin-7-diazonium chloride [114]. Although these aryl diazonium ions are promising for the desired application, their storage and delivery are challenging, and they often require in situ generation. The pH range should be between 8 and 10 for the formation of a phenolate anion [115].

Advertisement

3. Concluding remarks

Reactive fluorescent coumarins have been increasingly attracting special interest as fluorescent labels, with a wide range of applications in bioimaging and biolabelling, due to their extremely attractive and stable scaffold. Coumarins will allow the development of new low-cost fluorescent dyes due to its easy synthesis with high yields, large Stokes shift, pH independence of absorbance and emission, and excellent photostability, which represents a great value for the biological fluorescence imaging techniques.

Advertisement

Acknowledgments

The authors thank the European Regional Development Fund ALENTEJO 2020 through the project “MEDUSA-Microorganisms Monitoring and Mitigation—Developing and Unlocking novel Sustainable Approaches (ALT20-03-0145-FEDER-000015)” for financial support. Sérgio Martins acknowledge Fundação para a Ciência e a Tecnologia (FCT) for the economic support through the doctoral grant SFRH/BD/128807/2017.

Advertisement

Conflict of interest

There are no conflicts of interest to declare.

References

  1. 1. Ong S, Mann M. A practical recipe for stable isotope labeling by amino acids in cell culture (SILAC). Nature Protocols. 2006;1:2650-2660. DOI: 10.1038/nprot.2006.427
  2. 2. Varki A. Radioactive tracer techniques in the sequencing of glycoprotein oligosaccharides. The FASEB Journal. 1991;5(2):226-235. DOI: 10.1096/fasebj.5.2.2004668
  3. 3. Li H, Rothberg L. Colorimetric detection of DNA sequences based on electrostatic interactions with unmodified gold nanoparticles. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:14036-14039. DOI: 10.1073/pnas.0406115101
  4. 4. Kumita J, Smart O, Woolley G. Photo-control of helix content in a short peptide. Proceedings of the National Academy of Sciences of the United States of America. 2000;97:3803-3808. DOI: 10.1073/pnas.97.8.3803
  5. 5. Soh N, Yoshida K, Nakajima H, Nakano K, Imato T, Fukaminato T, et al. A fluorescent photochromic compound for labeling biomolecules. Chemical Communications. 2007;48:5206-5208. DOI: 10.1039/b713663c
  6. 6. Benedetto F, Mele E, Camposeo A, Athanassiou A, Cingolani R, Pisignano D. Photoswitchable organic nanofibers. Advanced Materials. 2008;20:314-318. DOI: 10.1002/adma.200700980
  7. 7. Staveren D, Metzler-Nolte N. Bioorganometallic chemistry of ferrocene. Chemical Reviews. 2004;104:5931-5985. DOI: 10.1021/cr0101510
  8. 8. Giepmans B, Adams S, Ellisman M, Tsien R. The fluorescent toolbox for assessing protein location and function. Science. 2006;312:217-224. DOI: 10.1126/science.1124618
  9. 9. Zimmer M. Green fluorescent protein (GFP): Applications, structure, and related photophysical behavior. Chemical Reviews. 2002;102:759-781. DOI: 10.1021/cr010142r
  10. 10. Hanson G, Hanson B. Fluorescent probes for cellular assays. Combinatorial Chemistry & High Throughput Screening. 2008;11(7):505-513. DOI: 10.2174/138620708785204090
  11. 11. Hermanson T. Bioconjugate Techniques. 3rd ed. London: Elsevier; 2013. DOI: 10.1016/C2009-0-64240-9
  12. 12. Kurien B, Scofield R. A brief review of other notable protein detection methods on blots. Methods in Molecular Biology. 2009;536:557-571. DOI: 10.1007/978-1-59745-542-8_56
  13. 13. Maurel D, Comps-Agrar L, Brock C, Rives M, Bourrier E, Ayoub M, et al. Cell-surface protein–protein interaction analysis with time-resolved FRET and snap-tag technologies: Application toGPCR oligomerization. Nature Methods. 2008;5:561-567. DOI: 10.1038/nmeth
  14. 14. Yin J, Straight P, McLoughlin S, Zhou Z, Lin A, Golan D, et al. Genetically encoded short peptide tag for versatile protein labeling by Sfp phosphopantetheinyl transferase. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(44):15815-15820. DOI: 10.1073/pnas.0507705102
  15. 15. Evans J, Yue D. New turf for CFP/YFP FRET imaging of membrane signaling molecules. Neuron. 2003;38:145-147. DOI: 10.1016/S0896-6273(03)00234-4
  16. 16. Pan D, Qin H, Cooperman B. Synthesis and functional activity of tRNAs labeled with fluorescent hydrazides in the D-loop. RNA. 2009;15(2):346-354. DOI: 10.1261/rna.1257509
  17. 17. Katritzky A, Narindoshvilia T. Fluorescent amino acids: Advances in protein-extrinsic fluorophores. Organic & Biomolecular Chemistry. 2009;7(4):627-634. DOI: 10.1039/b818908k
  18. 18. Sahoo H. Fluorescent labeling techniques in biomolecules: A flashback. RSC Advances. 2012;2:7017-7029. DOI: 10.1039/c2ra20389h
  19. 19. Biomol. Classic Reactive Fluorescent Labeling Dyes & Their Applications [Internet]. 2018. Available from: https://www.biomol.com/resources/biomol-blog/classic-fluorescent-labeling-dyes [Accessed: February 8, 2019]
  20. 20. Sahoo H. Förster resonance energy transfer–A spectroscopic nanoruler: Principle and applications. Journal of Photochemistry and Photobiology C. 2011;12:20-30. DOI: 10.1016/j.jphotochemrev.2011.05.001
  21. 21. Mochizuki N, Yamashita S, Kurokawa K, Ohba Y, Nagai T, Miyawaki A, et al. Spatio-temporal images of growth-factor-induced activation of Ras and Rap1. Nature. 2001;411(6841):1065-1068. DOI: 10.1038/35082594
  22. 22. Gaietta G, Deerinck T, Adams S, Bouwer J, Tour O, Laird D, et al. Multicolor and electron microscopic imaging of connexin trafficking. Science. 2002;296(5567):503-507
  23. 23. Miyashita T. Confocal microscopy for intracellular co-localization of proteins. Methods in Molecular Biology. 2004;261:399-410. DOI: 10.1385/1-59259-762-9:399
  24. 24. Zuleger N, Kelly D, Richardson A, Kerr A, Goldberg M, Goryachev A, et al. System analysis shows distinct mechanisms and common principles of nuclear envelope protein dynamics. The Journal of Cell Biology. 2011;193:109-123. DOI: 10.1083/jcb.201009068
  25. 25. Fernández-Suárez M, Ting A. Fluorescent probes for super-resolution imaging in living cells. Nature Reviews Molecular Cell Biology. 2008;9(12):929-943. DOI: 10.1038/nrm2531
  26. 26. Bustamante C, Cheng W, Mejia Y. Revisiting the central dogma one molecule at a time. Cell. 2011;144(4):480-497. DOI: 10.1016/j.cell.2011.01.033
  27. 27. Riveiro M, De Kimpe N, Moglioni A, Vázquez R, Monczor F, Shayo C, et al. Coumarins: Old compounds with novel promising therapeutic perspectives. Current Medicinal Chemistry. 2010;17(13):1325-1338. DOI: 10.2174/092986710790936284
  28. 28. Hoult J, Payá M. Pharmacological and biochemical actions of simple coumarins: Natural products with therapeutic potential. General Pharmacology. 1996;27(4):713-722. DOI: 10.1016/0306-3623(95)02112-4
  29. 29. Pierson J, Dumètre A, Hutter S, Delmas F, Laget M, Finet J, et al. Synthesis and antiprotozoal activity of 4-arylcoumarins. European Journal of Medicinal Chemistry. 2010;45(3):864-869. DOI: 10.1016/j.ejmech.2009.10.022
  30. 30. Combes S, Barbier P, Douillard S, McLeer-Florin A, Bourgarel-Rey V, Pierson J, et al. Synthesis and biological evaluation of 4-arylcoumarin analogues of combretastatins. Part 2. Journal of Medicinal Chemistry. 2011;54(9):3153-3162. DOI: 10.1021/jm901826e
  31. 31. Musa M, Cooperwood J, Khan M. A review of coumarin derivatives in pharmacotherapy of breast cancer. Current Medicinal Chemistry. 2008;15(26):2664-2679. DOI: 10.2174/092986708786242877
  32. 32. Kostova I. Synthetic and natural coumarins as cytotoxic agents. Current Medicinal Chemistry Anti-Cancer Agents. 2005;5(1):29-46. DOI: 10.2174/1568011053352550
  33. 33. Chin Y, Huang W, Hsu F, Lin Y, Lin M. Synthesis and evaluation of antibacterial activities of 5,7-dihydroxycoumarin derivatives. Archiv Der Pharmazie. 2011;344(6):386-393. DOI: 10.1002/ardp.201000233
  34. 34. Carotti A, Altomare C, Catto M, Gnerre C, Summo L, De Marco A, et al. Lipophilicity plays a major role in modulating the inhibition of monoamine oxidase B by 7-substituted coumarins. Chemistry & Biodiversity. 2006;3(2):134-149. DOI: 10.1002/cbdv.200690017
  35. 35. Thuong P, Hung T, Ngoc T, Ha do T, Min B, Kwack S, et al. Antioxidant activities of coumarins from Korean medicinal plants and their structure-activity relationships. Phytotherapy Research. 2010;24(1):101-106. DOI: 10.1002/ptr.2890
  36. 36. Kabeya L, de Marchi A, Kanashiro A, Lopes N, da Silva C, Pupo M, et al. Inhibition of horseradish peroxidase catalytic activity by new 3-phenylcoumarin derivatives: Synthesis and structure-activity relationships. Bioorganic & Medicinal Chemistry. 2007;15(3):1516-1524. DOI: 10.1016/j.bmc.2006.10.068
  37. 37. Jung J, Park O. Synthetic approaches and biological activities of 4-hydroxycoumarin derivatives. Molecules. 2009;14(11):4790-4803. DOI: 10.3390/molecules14114790
  38. 38. Kuznetsova N, Kaliya O. The photochemistry of coumarins. Russian Chemical Reviews. 1992;61(7):683-696. DOI: 10.1070/RC1992v061n07ABEH000992
  39. 39. Jones G, Jackson W, Choi C, Bergmark W. Solvent effects on emission yield and lifetime for coumarin laser dyes. Requirements for a rotatory decay mechanism. The Journal of Physical Chemistry. 1985;89(2):294-300. DOI: 10.1021/j100248a024
  40. 40. Jagtap A, Satam V, Rajule R, Kanetkar V. The synthesis and characterization of novel coumarin dyes derived from 1,4-diethyl-1,2,3,4-tetrahydro-7-hydroxyquinoxalin-6-carboxaldehyde. Dyes and Pigments. 2009;82:84-89. DOI: 10.1016/j.dyepig.2008.11.007
  41. 41. Wagner B. The use of coumarins as environmentally-sensitive fluorescent probes of heterogeneous inclusion systems. Molecules. 2009;14(1):210-237. DOI: 10.3390/molecules14010210
  42. 42. Katerinopoulos H. The coumarin moiety as chromophore of fluorescent ion indicators in biological systems. Current Pharmaceutical Design. 2004;10(30):3835-3852. DOI: 10.2174/1381612043382666
  43. 43. Kim J, Kim H, Kim S, Lee J, Do J, Kim H, et al. Fluorescent coumarinyldithiane as a selective chemodosimeter for mercury(II) ion in aqueous solution. Tetrahedron Letters. 2009;50(43):5958-5961. DOI: 10.1016/j.tetlet.2009.08.045
  44. 44. Sheng R, Wang P, Gao Y, Wu Y, Liu W, Ma J, et al. Colorimetric test kit for Cu2+ detection. Organic Letters. 2008;10(21):5015-5018. DOI: 10.1021/ol802117p
  45. 45. Kim H, Park J, Choi M, Ahn S, Chang S. Selective chromogenic and fluorogenic signalling of Hg2+ ions using a fluorescein-coumarin conjugate. Dyes and Pigments. 2010;84:54-58. DOI: 10.1016/j.dyepig.2009.06.009
  46. 46. Lin W, Yuan L, Cao X, Tan W, Feng Y. A coumarin-based chromogenic sensor for transition-metal ions showing ion-dependent bathochromic shift. European Journal of Organic Chemistry. 2008;29:4981-4987. DOI: 10.1002/ejoc.200800667
  47. 47. Jung H, Kwon P, Lee J, Kim J, Hong C, Kim J, et al. Coumarin-derived Cu(2+)-selective fluorescence sensor: Synthesis, mechanisms, and applications in living cells. Journal of the American Chemical Society. 2009;131(5):2008-2012. DOI: 10.1021/ja808611d
  48. 48. Wilze K, Johnson A. Handbook of Detergents, Chemistry, Production, and Application of Fluorescent Whitening Agents, Part F. Boca Raton: Taylor & Francis. CRC Press; 2007. DOI: 10.1201/9781420014655.ch28
  49. 49. Dorlars A, Schellhammer C, Schroeder J. Heterocycles as structural units in new optical brighteners. Angewandte Chemie International Edition. 1975;14:665-679. DOI: 10.1002/anie.197506651
  50. 50. Kido J, Iizumi Y. Fabrication of highly efficient organic electroluminescent devices. Applied Physics Letters. 1998;73:2721-2723. DOI: 10.1063/1.122570
  51. 51. Moylan C. Molecular hyperpolarizabilities of coumarin dyes. Physical Chemistry. 1994;98(51):13513-13516. DOI: 10.1021/j100102a014
  52. 52. Painelli A, Terenziani F. Linear and non-linear optical properties of push-pull chromophores: Vibronic and solvation effects beyond perturbation theory. Synthetic Metals. 2001;124(1):171-173. DOI: 10.1016/S0379-6779(01)00431-3
  53. 53. Benight S, Johnson L, Barnes R, Olbricht B, Bale D, Reid P, et al. Reduced dimensionality in organic electro-optic materials: Theory and defined order. The Journal of Physical Chemistry B. 2010;114:11949-11956. DOI: 10.1021/jp1022423
  54. 54. Lee M, Yen C, Yang W, Chen H, Liao C, Tsai C, et al. Efficient green coumarin dopants for organic light-emitting devices. Organic Letters. 2004;6(8):1241-1244. DOI: 10.1021/ol049903d
  55. 55. Swanson S, Wallraff G, Chen J, Zhang W, Bozano L, Carter K, et al. Stable and efficient fluorescent red and green dyes for external and internal conversion of blue OLED emission. Chemistry of Materials. 2003;15(12):2305-2312. DOI: 10.1021/cm021056q
  56. 56. Yu T, Zhang P, Zhao Y, Zhang H, Meng J, Fan D. Synthesis, characterization and high-efficiency blue electroluminescence based on coumarin derivatives of 7-diethylamino-coumarin-3-carboxamide. Organic Electronics. 2009;10:653-660. DOI: 10.1016/j.orgel.2009.02.026
  57. 57. Chang C, Cheng H, Lu Y, Tien K, Lin H, Lin C, et al. Enhancing color gamut of white OLED displays by using microcavity green pixels. Organic Electronics. 2010;11(4):247-254. DOI: 10.1016/j.orgel.2009.11.002
  58. 58. Mayer G, Heckel A. Biologically active molecules with a “light switch”. Angewandte Chemie International Edition. 2006;45(30):4900-4921. DOI: 10.1002/anie.200600387
  59. 59. Geissler D, Antonenko Y, Schmidt R, Keller S, Krylova O, Wiesner B, et al. (Coumarin-4-yl)methyl esters as highly efficient, ultrafast phototriggers for protons and their application to acidifying membrane surfaces. Angewandte Chemie International Edition. 2005;44(8):1195-1198. DOI: 10.1002/anie.200461567
  60. 60. Yu H, Li J, Wu D, Qiu Z, Zhang Y. Chemistry and biological applications of photo-labile organic molecules. Chemical Society Reviews. 2010;39(2):464-473. DOI: 10.1039/b901255a
  61. 61. Pinheiro A, Baptista P, Lima J. Light activation of transcription: Photocaging of nucleotides for control over RNA polymerization. Nucleic Acids Research. 2008;36(14):e90. DOI: 10.1093/nar/gkn415
  62. 62. Goddard J, Reymond J. Recent advances in enzyme assays. Trends in Biotechnology. 2004;22(7):363-370. DOI: 10.1016/j.tibtech.2004.04.005
  63. 63. Heiner S, Detert H, Kuhn A, Kunz H. Hydrophilic photolabelling of glycopeptides from the murine liver–intestine (LI) cadherin recognition domain. Bioorganic & Medicinal Chemistry. 2006;14:6149-6164. DOI: 10.1016/j.bmc.2006.06.014
  64. 64. Mizukami S, Okada S, Kimura S, Kikuchi K. Design and synthesis of coumarin-based Zn(2+) probes for ratiometric fluorescence imaging. Inorganic Chemistry. 2009;48(16):7630-7638. DOI: 10.1021/ic900247r
  65. 65. Hara K, Kurashige M, Dan-oh Y, Kasada C, Shinpo A, Suga S, et al. Design of new coumarin dyes having thiophene moieties for highly efficient organic-dye-sensitized solar cells. New Journal of Chemistry. 2003;27(5):783-785. DOI: 10.1039/b300694h
  66. 66. Camur M, Bulut M, Kandaz M, Guney O. Effects of coumarin substituents on the photophysical properties of newly synthesized phthalocyanine derivatives. Supramolecular Chemistry. 2009;21:624-631. DOI: 10.1080/10610270802613596
  67. 67. Gonçalves M. Fluorescent labeling of biomolecules with organic probes. Chemical Reviews. 2009;109:190-212. DOI: 10.1021/cr0783840
  68. 68. Johnson I, Spence M. Molecular Probes™ Handbook: A Guide to Fluorescent Probes and Labeling Technologies. 11th ed. New York: Life Technologies. Thermo Fisher Scientific; 2010. ISBN 13:978-0982927915
  69. 69. Webb M, Corrie J. Fluorescent coumarin-labeled nucleotides to measure ADP release from actomyosin. Biophysical Journal. 2001;81:1562-1569. DOI: 10.1016/S0006-3495(01)75810-9
  70. 70. Arguello J, Kaplan J. Glutamate 779, an intramembrane carboxyl, is essential for monovalent cation binding by the Na,K-ATPase. The Journal of Biological Chemistry. 1994;269(9):6892-6899
  71. 71. Arguello J, Kaplan J. Evidence for essential carboxyls in the cation-binding domain of the Na,K-ATPase. The Journal of Biological Chemistry. 1991;266(22):14627-14635
  72. 72. Berthelot T, Talbot J, Laïn G, Déleris G, Latxague L. Synthesis of nepsilon-(7-diethylaminocoumarin-3-carboxyl)- and nepsilon-(7-methoxycoumarin-3-carboxyl)-l-fmoc lysine as tools for protease cleavage detection by fluorescence. Journal of Peptide Science. 2005;11:153-160. DOI: 10.1002/psc.608
  73. 73. Oda Y, Kinoshita M, Nakayama K, Ikeda S, Kakehi K. Flow injection analysis of binding reaction between fluorescent lectin and cells. Analytical Biochemistry. 1999;269(2):230-235. DOI: 10.1006/abio.1999.4020
  74. 74. Gabor G, Chadha S, Walt D. Sensitivity enhancement of fluorescent pH indicators using pH-dependent energy transfer. Analytica Chimica Acta. 1995;313:131-137. DOI: 10.1016/0003-2670(95)00248-X
  75. 75. Exley D, Ekeke G. Fluoroimmunoassay of 5a-dihydrotestosterone. Journal of Steroid Biochemistry. 1981;14:1297-1302. DOI: 10.1016/0022-4731(81)90335-6
  76. 76. Tisljar U, Knight C, Barrett A. An alternative quenched fluorescence substrate for Pz-peptidase. Analytical Biochemistry. 1990;186(1):112-115. DOI: 10.1016/0003-2697(90)90582-T
  77. 77. Mita H, Yasueda H, Hayakawa T, Shida T. Quantitation of platelet-activating factor by high performance liquid chromatography with fluorescent detection. Analytical Biochemistry. 1989;180:131-135. DOI: 10.1016/0003-2697(89)90100-0
  78. 78. Khalfan H, Abuknesha R, Rand-Weaver M, Price R, Robinson D. Aminomethyl coumarin acetic acid: A new fluorescent labelling agent for proteins. The Histochemical Journal. 1986;18(9):497-499. DOI: 10.1007/BF01675617
  79. 79. Eldaw A, Khalfan H. Aminomethyl coumarin acetic acid and fluorescein isothiocyanate in detection of leishmanial antibodies: A comparative study. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1988;82(4):561-562. DOI: 10.1016/0035-9203(88)90506-8
  80. 80. Panchuk-Voloshina N, Haugland R, Bishop-Stewart J, Bhalgat M, Millard P, Mao F, et al. Alexa dyes, a series of new fluorescent dyes that yield exceptionally bright, photostable conjugates. The Journal of Histochemistry and Cytochemistry. 1999;47(9):1179-1188. DOI: 10.1177/002215549904700910
  81. 81. Cox W, Singer V. Fluorescent DNA hybridization probe preparation using amine modification and reactive dye coupling. BioTechniques. 2004;36(1):114-122. DOI: 10.2144/04361RR02
  82. 82. Li M, Reddy L, Bennett R, Silva N, Jones L, Thomas D. A fluorescence energy transfer method for analyzing protein oligomeric structure: Application to phospholamban. Biophysical Journal. 1999;76(5):2587-2599. DOI: 10.1016/S0006-3495(99)77411-4
  83. 83. Nguyen T, Joshi N, Francis M. An affinity-based method for the purification of fluorescently-labeled biomolecules. Bioconjugate Chemistry. 2006;17(4):869-872. DOI: 10.1021/bc060130i
  84. 84. Weerachatyanukul W, Xu H, Anupriwan A, Carmona E, Wade M, Hermo L, et al. Acquisition of arylsulfatase A onto the mouse sperm surface during epididymal transit. Biology of Reproduction. 2003;69(4):1183-1192. DOI: 10.1095/biolreprod.102.010231
  85. 85. Lewis B, Rathman S, McMahon R. Detection and quantification of biotinylated proteins using the storm 840 optical scanner. The Journal of Nutritional Biochemistry. 2003;14(4):196-202. DOI: 10.1016/S0955-2863(02)00283-8
  86. 86. Kumari S, Wälchli S, Fallang L, Yang W, Lund-Johansen F, Schumacher T, et al. Alloreactive cytotoxic T cells provide means to decipher the immunopeptidome and reveal a plethora of tumor-associated self-epitopes. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(1):403-408. DOI: 10.1073/pnas.1306549111
  87. 87. Saavedra-Lozano J, Cao Y, Callison J, Sarode R, Sodora D, Edgar J, et al. An anti-CD45RO immunotoxin kills HIV-latently infected cells from individuals on HAART with little effect on CD8 memory. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:2494-2499. DOI: 10.1073/pnas.0308381100
  88. 88. Irish J, Myklebust J, Alizadeh A, Houot R, Sharman J, Czerwinski D, et al. B-cell signaling networks reveal a negative prognostic human lymphoma cell subset that emerges during tumor progression. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(29):12747-12754. DOI: 10.1073/pnas.1002057107
  89. 89. Karachunski P, Ostlie N, Monfardini C, Conti-Fine B. Absence of IFN-gamma or IL-12 has different effects on experimental myasthenia gravis in C57BL/6 mice. Journal of Immunology. 2000;164(10):5236-5244. DOI: 10.4049/jimmunol.164.10.5236
  90. 90. Henegariu O, Bray-Ward P, Ward D. Custom fluorescent-nucleotide synthesis as an alternative method for nucleic acid labeling. Nature Biotechnology. 2000;18(3):345-348. DOI: 10.1038/73815
  91. 91. Zhang P, Beck T, Tan W. Design of a molecular beacon DNA probe with two fluorophores. Angewandte Chemie International Edition. 2001;40(2):402-405. DOI: 10.1002/1521-3773
  92. 92. Lumiprobe. Life science solutions [Internet]. 2019. Available from: https://www.lumiprobe.com/p/coumarin-343-x-nhs-ester [Accessed: February 15, 2019]
  93. 93. Zheng L, Zhao H, Han Y, Qian H, Vukovic L, Mecinović J, et al. Catalytic transport of molecular cargo using diffusive binding along a polymer track. Nature Chemistry. 2019;11:359-366. DOI: 10.1038/s41557-018-0204-7
  94. 94. Gee K, Archer E, Kang H. 4-Sulfotetrafluorophenyl (STP) esters: New water-soluble amine-reactive reagents for labeling biomolecules. Tetrahedron Letters. 1999;40(8):1471-1474. DOI: 10.1016/S0040-4039(98)02695-1
  95. 95. González-Pérez M, Ooi S, Martins S, Ramalho J, Pereira A, Caldeira A. Gaining insight into the photophysical properties of a coumarin STP ester with potential for bioconjugation. New Journal of Chemistry. 2018;42:16635-16645. DOI: 10.1039/c8nj03548b
  96. 96. Yin S, Ramalho J, Pereira A, Martins S, Salvador C, Caldeira A. A simple method for labelling and detection of proteinaceous binders in art using fluorescent coumarin derivatives. The European Physical Journal Plus. 2019;134:71-80. DOI: 10.1140/epjp/i2019-12478-4
  97. 97. Al-Kindy S, Miller J. Coumarin-6-sulphonyl chloride: A novel label in fluorimetry and phosphorimetry Part 1. Synthesis and luminescence properties anal. Chimica Acta. 1989;227:145-153. DOI: 10.1016/S0003-2670(00)82653-7
  98. 98. Al-Kindy S, Suliman F, Al-Hamadi A. Fluorescence enhancement of coumarin-6-sulfonyl chloride amino acid derivatives in cyclodextrin media. Analytical Sciences. 2001;17(4):539-543. DOI: 10.2116/analsci.17.539
  99. 99. Signore G, Nifosì R, Albertazzi L, Bizzarri R. A novel coumarin fluorescent sensor to probe polarity around biomolecules. Journal of Biomedical Nanotechnology. 2009;5(6):722-729. DOI: 10.1166/jbn.2009.1089
  100. 100. Jashari A, Hey-Hawkins E, Mikhova B, Draeger G, Popovsk E. An improved synthesis of 4-chlorocoumarin-3-sulfonyl chloride and its reactions with different bidentate nucleophiles to give pyrido[1′,2′:2,3]- and thiazino[3′,2′:2,3]-1,2,4-thiadiazino[6,5-c]benzopyran-6-one 7,7-dioxides. Molecules. 2007;12(8):2017-2028. DOI: 10.3390/12082017
  101. 101. Al-Kindy S, Al-Sharji N, Al-Harasi A, Suliman F, AL-Lawati H, Schulman S. Synthesis and spectroscopic study of 2,7-diethylamino-2-oxo-2H-chromen-3-yl benzothiazole-6-sulfonyl chlorides and its derivatives. Arabian Journal of Chemistry. 2017;10:S114-S120. DOI: 10.1016/j.arabjc.2012.06.015
  102. 102. Bhusal R, Cho P, Kim S-A, Park H, Kim H. Synthesis of green emitting coumarin bioconjugate for the selective determination of flu antigen. Bulletin of the Korean Chemical Society. 2011;32(5):1461-1462. DOI: 10.5012/bkcs.2011.32.5.1461
  103. 103. Mayekar S, Chaskar A, Mulwad V. Facile synthesis of coumarinyl isothiocyanate from amino coumarin. Synthetic Communications. 2010;1(40):46-51. DOI: 10.1080/00397910902916080
  104. 104. Patil S, Deokar H. Synthesis and structural determination of novel heterocyclic derivatives of coumarin isothiocyanates. Der Chemica Sinica. 2014;5(6):74-78
  105. 105. Sippel T. New fluorochromes for thiols: Maleimide and iodoacetamide derivatives of a 3-phenylcoumarin fluorophore. The Journal of Histochemistry and Cytochemistry. 1981;29(2):314-316. DOI: 10.1177/29.2.7019305
  106. 106. Kuiper J, Pluta R, Huibers W, Fusetti F, Geertsma E, Poolman B. A method for site-specific labeling of multiple protein thiols. Protein Science. 2009;18(5):1033-1041. DOI: 10.1002/pro.113
  107. 107. Zhou Z, Koglin A, Wang Y, McMahon A, Walsh C. An eight residue fragment of an acyl carrier protein suffices for post-translational introduction of fluorescent pantetheinyl arms in protein modification in vitro and in vivo. Journal of the American Chemical Society. 2008;130(30):9925-9930. DOI: 10.1021/ja802657n
  108. 108. Zavala L, Pardo-López L, Cantón P, Gómez I, Soberón M, Bravo A. Domains II and III of Bacillus thuringiensis Cry1Ab toxin remain exposed to the solvent after insertion of part of domain I into the membrane. The Journal of Biological Chemistry. 2011;286(21):19109-19117. DOI: 10.1074/jbc.M110.202994
  109. 109. Ramachandiran V, Grigoriev V, Lan L, Ravkov E, Mertens S, Altman J. A robust method for production of MHC tetramers with small molecule fluorophores. Journal of Immunological Methods. 2007;319(1-2):13-20. DOI: 10.1016/j.jim.2006.08.014
  110. 110. Kunzelmann S, Webb M. A biosensor for fluorescent determination of ADP with high time resolution. The Journal of Biological Chemistry. 2009;284(48):33130-33138. DOI: 10.1074/jbc.M109.047118
  111. 111. Brune M, Corrie J, Webb M. A fluorescent sensor of the phosphorylation state of nucleoside diphosphate kinase and its use to monitor nucleoside diphosphate concentrations in real time. Biochemistry. 2001;40(16):5087-5094. DOI: 10.1021/bi002484h
  112. 112. Anderson S, Williams C, O’donnell M, Bloom L. A function for the psi subunit in loading the Escherichia coli DNA polymerase sliding clamp. The Journal of Biological Chemistry. 2007;282(10):7035-7045. DOI: 10.1074/jbc.M610136200
  113. 113. Boutureira O, Bernardes G. Advances in chemical protein modification. Chemical Reviews. 2015;115(5):2174-2195. DOI: 10.1021/cr500399p
  114. 114. Shadmehr M, Davis G, Mehari B, Jensen S, Jewett J. Coumarin triazabutadienes for fluorescent labeling of proteins. Chembiochem. 2018;19:2550-2552. DOI: 10.1002/cbic.201800599
  115. 115. Gavrilyuk J, Ban H, Nagano M, Hakamata W, Barbas C. Formylbenzene diazonium hexafluorophosphate reagent for tyrosine-selective modification of proteins and the introduction of a bioorthogonal aldehyde. Bioconjugate Chemistry. 2012;23(12):2321-2328. DOI: 10.1021/bc300410p

Written By

António Pereira, Sérgio Martins and Ana Teresa Caldeira

Submitted: 27 February 2019 Reviewed: 20 March 2019 Published: 07 June 2019