Skip to content
BY 4.0 license Open Access Published by De Gruyter March 17, 2020

Isatin as a simple, highly selective and sensitive colorimetric sensor for fluoride anion

  • Azeem Haider , Mukhtiar Ahmed , Muhammad Faisal and Muhammad Moazzam Naseer EMAIL logo

Abstract

Herein, we report the fluoride anion sensing properties of a commercially available and inexpensive organic compound, isatin, which is found to be a highly selective and sensitive sensor. In naked-eye experiments, by addition of fluoride anions, isatin shows a dramatic color change from pale yellow to violet at room temperature, while the addition of other anions, i.e. Cl,Br,I,ClO4,H2PO4andPF6,did not induce any colour change. Additionally, recognition and titration studies have also been done through UV/Vis spectroscopy. Isatin displayed a new absorption band at 533 nm after the addition of fluoride anions, which is presumably due to acid-base interaction between isatin and fluoride anions, while other anions did not trigger noticeable spectral changes. The detection limit was observed to be 0.367 ppm. DFT calculations were also performed to further explain the behavior of receptor 1 towards the Fˉ anion. Owing to high sensitivity and selectivity, isatin can be useful in the detection of biologically or environmentally important fluoride anions at very low concentration.

Introduction

In our everyday life, anions are valuable because they regulate and/or are responsible for numerous environmental and biological processes [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16]. For instance, anions are involved in the production of electrical signals, controlling osmotic pressure, activating signal transduction pathways, maintaining cell volume, etc [17]. Therefore, the changes in anion flux across cell membranes is increasingly being identified as the foremost cause of various disorders including osteoporosis [18], Pendred’s syndrome [19], Dent’s disease [20], Bartter’s syndrome [21] and cystic fibrosis [22]. Further, anions are important for many industrial processes and are often found as harmful pollutants [23]. Also, anions are present in almost 70% of all active sites of enzyme, which play a significant part in genetic information storage [24]. Moreover, anions play crucial roles in the fields of catalysis and medicine, although pollutant anions have been mainly linked to carcinogenesis (metabolites of benzoate and acetate) [25], eutrophication of rivers (from the application of phosphate-based fertilizers) [26] and contamination of drinking water (by fluoride pollution) [27].

Among anions, fluoride is the most important, owing to its involvement in numerous environmental and biological processes [28, 29, 30, 31]. Fluoride is known to have a hard Lewis base nature, the smallest ionic radius, and the highest charge density of any anion [32]. Naturally, fluoride ions occur in the sea, in soils and rocks, and in groundwater supplies. Numerous studies have shown that fluoride in drinking water or toothpaste can help with dental health [33, 34]. Also, fluoride can help increase bone mass and hence has the potential for treating osteoporosis [35, 36]. Because of these benefits, water fluoridation had become widely used by 1960 [37]. However, later studies have shown that overexposure to fluoride can be detrimental, especially to the teeth [38], brain [39], kidney [40] and bone [41], resulting in a disease known as fluorosis. This diversity of function of the fluoride anion, both the benefits and perniciousness, makes its detection of considerable interest. In this context, organic colorimetric chemosensors are of particular importance due to their simplicity [29, 42,43, 44, 45, 46]. Color changes visible to the naked eye are preliminarily used as signals for detection of anions without any equipment being required. However, the synthesis of such chemosensors usually requires expensive reagents, and long or tedious synthetic processes, making them less applicable. Therefore, the development of a cheap, non-corrosive, environment friendly, safe, selective, and sensitive chemosensor remains attractive [42, 43].

Isatin (1H-indole-2,3-dione) 1 (Figure 1) is a commercially available and inexpensive organic compound that was first discovered by Erdman and Laurent in 1840 as an oxidation product of indigo [47]. Almost 140 years after its discovery, it was identified as a natural product present in plants, animals, fungi, symbiotic bacteria and marine molluscs [47]. Isatin 1 is also considered important in many physiological pathways [48, 49, 50]. Apart from its applications in medicinal chemistry [47], it has been extensively employed as a signaling unit in various receptors [51, 52] and some of its derivatives have also been investigated as chemosensors for anion detection [53, 54]. Careful analysis of its features reveals that isatin 1 has an acidic NH group and intense color which are prime requirements for a chemosensor [29, 42,43, 44, 45, 46]. The interesting structure combined with our recent interest in isatin-derived compounds [55, 56, 57, 58, 59] and in developing chemosensors [6, 60] motivated us to investigate the behavior of commercially available isatin as an anion sensor. Fascinatingly, as discussed below, isatin is found to be a highly selective and sensitive chemosensor for fluoride anions in the presence of various other anions, as its acidity allows it to interact only with fluoride anions. To the best of our knowledge, this is one of the cheapest chemosensors reported so far for the selective recognition/ sensing of the fluoride anion.

Figure 1 Structure of isatin.
Figure 1

Structure of isatin.

Results and discussion

As isatin 1 is a colored compound,its potential in anion recognition/sensing was first monitored by naked eye experiments. This method of sensing is highly valued due to its low cost and easy detection without the help of any instrument or equipment [29, 42,43, 44, 45, 46]. Interestingly, the addition of one equivalent of Fˉ anion to the solution of isatin 1 resulted in a dramatic color change from pale yellow to violet, clearly visible to the naked eye. Most importantly, the color change was observed even at a low concentration of fluoride anion (10mM). In contrast, the addition of other anions resulted in no visible changes in color (Figure 2).

Figure 2 Color changes of the chemosensor 1 (0.01 μM) in MeCN with the addition of 1 equivalent tetrabutylammonium salts (from left to right: 1, F−,Cl−,Br−,l−,ClO4−,H2PO4− and PF6−).$1,\,{{\mathrm{F}}^{-}},\mathrm{C}{{\mathrm{l}}^{-}},\mathrm{B}{{\mathrm{r}}^{-}},{{\mathrm{l}}^{-}},\mathrm{ClO}_{4}^{-},{{\mathrm{H}}_{2}}\mathrm{PO}_{4}^{-}\,\mathrm{and}\,\mathrm{PF}_{6}^{-}).$
Figure 2

Color changes of the chemosensor 1 (0.01 μM) in MeCN with the addition of 1 equivalent tetrabutylammonium salts (from left to right: 1,F,Cl,Br,l,ClO4,H2PO4andPF6).

After obtaining the preliminary information of selectivity for fluoride anion with naked eye experiments (Figure 2), the chemosensor 1 was further evaluated by UV-Vis spectroscopy. Absorption spectra were recorded after the addition of one equivalent of different anions to the solution (0.01 μM) of isatin 1 in MeCN to observe the selectivity pattern. All the solutions were well shaken to ensure the homogeneity of the solutions before recording the absorption spectrum. Figure 3 shows the changes in the UV-Vis spectra of 1 upon the addition of F,Cl,Br,I,ClO4,H2PO4andPF6(as their tetra-n-butylammonium salts). As shown in the Figure 2, 1, which absorbs at 410 nm in MeCN solvent, demonstrated a new and bathochromic shifted absorption band at 533 nm after the addition of fluoride anion. A bathochromic shift of more than 100 nm may be attributed to the hydrogen bonding/acid-base interaction between 1 and the fluoride anion. In other words, the interaction of fluoride anion with 1 caused hydrogen-bond or negative charge (in case of complete deprotonation)-induced electron delocalization. Under similar conditions, no changes in the absorption spectra could be observed upon addition of other anions, i.e. Clˉ, Br,I,ClO4,H2PO4andPF6.These results clearly indicate the ability of 1 for selective fluoride ion recognition in the presence of various other anions (Figure 3).

Figure 3 Absorption spectral changes of chemosensor 1 (0.01 μM) upon addition of 1 μM solutions of different anions.
Figure 3

Absorption spectral changes of chemosensor 1 (0.01 μM) upon addition of 1 μM solutions of different anions.

After ascertaining that compound 1 can recognize fluoride anions preferentially over other anions, titration studies were performed to check the quantitative behavior of this chemosensor towards different concentration of fluoride anion (Figure 4). In these studies absorption spectra were recorded for each concentration of fluoride anion. It was observed that by gradually increasing the concentration of fluoride anion, the intensity of the absorption band of 1 at 410 nm decreased and the intensity of the absorption band at 533 nm increased with a sharp isobestic point development at 470 nm, indicating the existence of only two species in equilibrium throughout the titration process (Figure 5). From this, it may be deduced that the intermolecular interaction between 1 and the fluoride anion results in complete deprotonation rather than hydrogen bonding, owing to its greater influence on the electron density of the isatin nucleus.

Figure 4 UV-Vis titrations of 1 (0.01 μM) with different concentrations of TBAF in MeCN.
Figure 4

UV-Vis titrations of 1 (0.01 μM) with different concentrations of TBAF in MeCN.

Figure 5 Proposed mechanism of detection of 1.
Figure 5

Proposed mechanism of detection of 1.

The binding interactions of 1 with the analyte was further examined via Job’s plot analysis. As shown in Figure 6, the gradual increase in the intensity of the absorption band of 1 with the increase in fluoride anion concentration showed a linear relationship up to one equivalent of fluoride anion, signifying a 1:1 binding ratio between 1 and the Fˉ ion. Moreover, according to the UV-Vis titration experiment the chemosensor 1 showed a detection limit [61] of 0.367 ppm with R2 = 0.9654.

Figure 6 Linear regression graph between added concentration of TBAF (0-2x10-8 M) and of relative absorbance (at 410 and 533 nm) of chemosensor 1.
Figure 6

Linear regression graph between added concentration of TBAF (0-2x10-8 M) and of relative absorbance (at 410 and 533 nm) of chemosensor 1.

Computational studies

The observed selectivity and sensitivity of chemosensor 1 towards the Fˉ anion can be explained on the basis of density functional theory (DFT) calculations. The optimized structures of 1 and its anion in the ground state were obtained from DFT/B3LYP/6-31G using the Gaussian 09 software package. Figure 7 shows the optimized structures, graphical representations (isodensity surface plots) of the LUMO (lowest unoccupied molecular orbital) and HOMO (highest occupied molecular orbital) of 1 and its anion. Figures 7 also displays the band gaps and energy level values. The HOMO energy level (-5.16 eV) of the isatin anion is higher than that of the chemosensor 1 (-5.54 eV) and the LUMO energy level (-2.79 eV) of the isatin anion is lower than that of the chemosensor 1 (-2.50 eV); hence, the band gap between the LUMO and HOMO of the isatin anion (2.37 eV) is smaller than that of the sensor 1 (3.04 eV), which is in good agreement with the bathochromic shift in absorption (λmax = 533 nm) detected upon treatment of chemosensor 1 with Fˉ anions. It is noted that the relation between the energy band gap and absorption spectra was estimated from the Planck-Einstein relation [62].

Figure 7 Frontier molecular orbitals of chemosensor 1 and its anion.
Figure 7

Frontier molecular orbitals of chemosensor 1 and its anion.

In the case of 1, both the HOMO and LUMO were localized throughout the whole molecule. However, in the case of the anion of 1, spatial separation of FMOs is noticeable, i.e., both HOMO and LUMO were chiefly localized on the electron-richer pyrrolidine ring. This orbital distribution revealed that the deprotonation of 1 should impact the LUMO and HOMO evenly and therefore constitute the basis of an optical response of 1 to addition of Fˉ (Figure 7).

Conclusions

In summary, a commercially available colored organic compound, isatin, was investigated for selective fluoride anion sensing using UV/Vis spectroscopy and naked-eye experiments. The results reveals that isatin is a novel colorimetric chemosensor for selective detection of fluoride ions in acetonitrile. Isatin shows changes in both its color and its UV–Vis absorption spectra upon the addition of fluoride, while the other anions, such as Clˉ, Brˉ, Iˉ, ClO4ˉ, H2PO4ˉ and PF6ˉ, did not produce any noticeable change. The selectivity of the isatin receptor towards the fluoride anion over the other anions is attributed to the small size and greater charge density, hence Lewis basicity, of the fluoride anion, which led to the complete deprotonation of the receptors. However, the other anions/Lewis bases were not strong enough to deprotonate the receptors. The results further demonstrated that the isatin receptor respond in microseconds and in linear fashion over a wide range of fluoride concentrations and a very low concentration of fluoride anion could be detected, i.e., the limit of detection of isatin is 0.367 ppm. Owing to its solubility in water as well as in organic solvents, isatin can be useful in the detection of biologically or environmentally important fluoride anions in both aqueous and organic media. The high selectivity and sensitivity of isatin towards fluoride anions makes this molecule highly attractive and this study will provide an ideal platform to explore this molecule further in the field of anion sensing/recognition.

Experimental

Naked-eye detection studies

For naked eye experiments, 1 (0.01 µM) 3 ml was placed into a glass vial and 1 equivalent of each anion was added to the glass vial.

Recognition studies

The UV-Vis spectra were recorded using a BioMate 3S UV-Visible Spectrophotometer (Thermo Scientific). All absorbance assays were performed in 100 μL quartz cuvettes. The recognition studies were performed at 25oC. 2 µM solutions of 1 in acetonitrile were prepared and 2.5 mL of each was taken in volumetric flasks. The binding behavior of 1 was studied by adding 2.5 mL solutions (4 µM) of different anions (TBAF, TBACl, TBABr, TBAI, TBAPF6, TBAH2PO4 and TBAClO4) into the solutions of 1 and the absorption spectrum of each solution was recorded.

Titration studies

For titration studies, 0.2, 0.4, 0.6, 0.8, 1.0, 1.2, 1.4 and 2.0 equivalents of Fˉ were individually added to volumetric flasks containing 1 and absorption spectra were recorded for set after each aliquot addition of Fˉ.

Selectivity studies

Possible interference due to different anions was studied by adding one equivalent of different anions to 1 µM solution of 1. Absorption spectra were recorded and no interference was found.

Sensitivity studies

The sensitivity of 1 towards fluoride anion was checked by adding a solution of Fˉ (µM) to a solution of 1 (0.01 µM) and absorption spectra were recorded as a function of time at short intervals.


Tel +92-51 9064-2129

Acknowledgments

We are grateful to The World Academy of Sciences (TWAS) for financial support (Project No. 13-419 RG/PHA/AS_CUNESCO FR: 3240279216) and Quaid-i-Azam University for financial support under the URF program.

References

[1] Molina P, Zapata F. Caballero, A. Anion Recognition Strategies Based on Combined Noncovalent Interactions. Chem Rev. 2017;117:9907–72.10.1021/acs.chemrev.6b00814Search in Google Scholar PubMed

[2] Zhao Y, Cotelle Y, Liu L, López-Andarias J, Bornhof AB, Akamatsu M, et al. The Emergence of Anion−π Catalysis. Acc Chem Res. 2018;51:2255–63.10.1021/acs.accounts.8b00223Search in Google Scholar PubMed

[3] He Q, Vargas-Zúñiga GI, Seung Kim H, Kim SK, Sessler JL. Macrocycles as Ion Pair Receptors. Chem Rev. 2019;119:9753–835.10.1021/acs.chemrev.8b00734Search in Google Scholar PubMed

[4] Gale PA, Howe EN, Wu X. Anion Receptor Chemistry. Chem. 2016;1:351–422.10.1016/j.chempr.2016.08.004Search in Google Scholar

[5] Langton MJ, Serpell CJ, Beer PD. Anion Recognition in Water: Recent Advances from a Supramolecular and Macromolecular Perspective. Angew Chem Int Ed. 2016;55:1974–87.10.1002/anie.201506589Search in Google Scholar PubMed PubMed Central

[6] Naseer MM, Jurkschat K. Organotin-based receptors for anions and ion pairs. Chem Commun (Camb). 2017;53:8122–35.10.1039/C7CC02667FSearch in Google Scholar PubMed

[7] Busschaert N, Caltagirone C, Van Rossom W, Gale PA. Applications of Supramolecular Anion Recognition. Chem Rev. 2015;115:8038–155.10.1021/acs.chemrev.5b00099Search in Google Scholar PubMed

[8] Gale PA, Caltagirone C. Anion sensing by small molecules and molecular ensembles. Chem Soc Rev. 2015;44:4212–27.10.1039/C4CS00179FSearch in Google Scholar PubMed

[9] Giese M, Albrecht M, Rissanen K. Anion-pi Interactions with Fluoroarenes. Chem Rev. 2015;115:8867–95.10.1021/acs.chemrev.5b00156Search in Google Scholar PubMed

[10] Gale PA, Perez-Tomas R, Quesada R. Anion Transporters and Biological Systems. Acc Chem Res. 2013;46:2801–13.10.1021/ar400019pSearch in Google Scholar PubMed

[11] Valkenier H, Davis AP. Making a Match for Valinomycin: Steroidal Scaffolds in the Design of Electroneutral, Electrogenic Anion Carriers. Acc Chem Res. 2013;46:2898–909.10.1021/ar4000345Search in Google Scholar

[12] Gale PA. From Anion Receptors to Transporters. Acc Chem Res. 2011;44:216–26.10.1021/ar100134pSearch in Google Scholar

[13] Custelcean R. Anions in crystal engineering. Chem Soc Rev. 2010;39:3675–85.10.1039/b926221kSearch in Google Scholar

[14] Carroll CN, Naleway JJ, Haley MM, Johnson DW. Arylethynyl receptors for neutral molecules and anions: emerging applications in cellular imaging. Chem Soc Rev. 2010;39:3875–88.10.1039/b926231hSearch in Google Scholar

[15] Bianchi A, Bowman-James K, García-España E. Supramolecular Chemistry of Anions. New York: Wiley-VCH; 1997.Search in Google Scholar

[16] Sessler JL, Gale PA, Cho W (Stoddart JF, editor). S Anion Receptor Chemistry. Cambridge, U.K.: Royal Society of Chemistry; 2006.10.1039/9781847552471Search in Google Scholar

[17] Alberts B, Bray D, Lewis J, Raff M, Roberts K, Watson JD. Molecular Biology of the Cell. 3rd ed. New York: Garland Science; 1994.Search in Google Scholar

[18] Kornak U, Kasper D, Bosl MR, Kaiser E, Schweizer M, Schulz A, et al. Loss of the ClC-7 Chloride Channel Leads to Osteopetrosis in Mice and Man. Cell. 2001;104:205–15.10.1016/S0092-8674(01)00206-9Search in Google Scholar

[19] Scott DA, Wang R, Kreman TM, Sheffield VC, Karniski LP. The Pendred syndrome gene encodes a chloride-iodide transport protein. Nat Genet. 1999;21:440–3.10.1038/7783Search in Google Scholar PubMed

[20] Devuyst O, Christie PT, Courtoy PJ, Beauwens R, Thakker RV. Intra-renal and subcellular distribution of the human chloride channel, CLC-5, reveals a pathophysiological basis for Dent’s disease. Hum Mol Genet. 1999;8:247–57.10.1093/hmg/8.2.247Search in Google Scholar PubMed

[21] Simon DB, Bindra RS, Mansfield TA, Nelson-Williams C, Mendonca E, Stone RP, et al. Mutations in the chloride channel gene, CLCNKB, cause Bartter’s syndrome type III. Nat Genet. 1997;17:171–8.10.1038/ng1097-171Search in Google Scholar PubMed

[22] Anderson MP, Gregory RJ, Thompson S, Souza DW, Paul S, Mulligan RC, et al. J Demonstration that CFTR is a chloride channel by alteration of its anion selectivity. Science. 1991;253:202–5.10.1126/science.1712984Search in Google Scholar PubMed

[23] Sessler JL, Davis JM. Sapphyrins: Versatile Anion Binding Agents. Acc Chem Res. 2001;34:989–97.10.1021/ar980117gSearch in Google Scholar PubMed

[24] Thale PB, Borase PN, Shankarling GS. “turn on” fluorescent and chromogenic chemosensor for fluoride anion: experimental and DFT studies. Inorg Chem Front. 2016;3:977–84.10.1039/C6QI00033ASearch in Google Scholar

[25] Heras JD, Aldámiz-Echevarría L, Martínez-Chantar ML, Delgado TC. An update on the use of benzoate, phenylacetate and phenylbutyrate ammonia scavengers for interrogating and modifying liver nitrogen metabolism and its implications in urea cycle disorders and liver disease. Expert Opin Drug Metab Toxicol. 2017;13:439–48.10.1080/17425255.2017.1262843Search in Google Scholar PubMed PubMed Central

[26] Elser JJ, Bracken ME, Cleland EE, Gruner DS, Hillebrand H, Ngai JT, et al. Global analysis of nitrogen and phosphorus limitation of primary producers in freshwater, marine and terrestrial ecosystems. Ecol Lett. 2007;10:1135–42.10.1111/j.1461-0248.2007.01113.xSearch in Google Scholar PubMed

[27] Bose P, Ravikumar I, Ghosh P. Anion binding in the C3v-symmetric cavity of a protonated tripodal amine receptor: potentiometric and single crystal X-ray studies. Inorg Chem. 2011;50:10693–702.10.1021/ic2011502Search in Google Scholar PubMed

[28] Wade CR, Broomsgrove AE, Aldridge S, Gabbaï FP. Fluoride Ion Complexation and Sensing Using Organoboron Compounds. Chem Rev. 2010;110:3958–84.10.1021/cr900401aSearch in Google Scholar PubMed

[29] Zhou Y, Zhang JF, Yoon J. Fluorescence and Colorimetric Chemosensors for Fluoride-Ion Detection. Chem Rev. 2014;114(10):5511–71.10.1021/cr400352mSearch in Google Scholar PubMed

[30] Jagtap S, Yenkie MK, Labhsetwar N, Rayalu S. Fluoride in Drinking Water and Defluoridation of Water. Chem Rev. 2012;112:2454–66.10.1021/cr2002855Search in Google Scholar PubMed

[31] Wu X, Wang H, Yang S, Tian H, Liu Y, Sun B. Highly Sensitive Ratiometric Fluorescent Paper Sensors for the Detection of Fluoride Ions. ACS Omega. 2019;4:4918–26.10.1021/acsomega.9b00283Search in Google Scholar PubMed PubMed Central

[32] Zhou X, Lai R, Li H, Stains CI. The 8-silyloxyquinoline scaffold as a versatile platform for the sensitive detection of aqueous fluoride. Anal Chem. 2015;87:4081–6.10.1021/acs.analchem.5b00430Search in Google Scholar PubMed

[33] Campus G, Cagetti MG, Spano N, Denurra S, Cocco F, Bossù M, et al. Laboratory enamel fluoride uptake from fluoride products. Am J Dent. 2012;25:13–6.Search in Google Scholar

[34] Aoun A, Darwiche F, Al Hayek S, Doumit J. The Fluoride Debate: The Pros and Cons of Fluoridation. Prev Nutr Food Sci. 2018;23:171–80.10.3746/pnf.2018.23.3.171Search in Google Scholar PubMed PubMed Central

[35] Pitt P, Berry H. Fluoride treatment in osteoporosis. Postgrad Med J. 1991;67:323–6.10.1136/pgmj.67.786.323Search in Google Scholar PubMed PubMed Central

[36] Demos L, Kazda H, Cicuttini FM, Sinclair MI, Fairley CK. Water fluoridation, osteoporosis, fractures—recent developments. Aust Dent J. 2001;46:80–7.10.1111/j.1834-7819.2001.tb00561.xSearch in Google Scholar PubMed

[37] Lennon MA. One in a million: the first community trial of water fluoridation. Bull World Health Organ. 2006;84:759–60.10.2471/BLT.05.028209Search in Google Scholar

[38] Rwenyonyi CM, Bjorvatn K, Birkeland JM, Haugejordan O. Altitude as a risk indicator of dental fluorosis in children residing in areas with 0.5 and 2.5 mg fluoride per litre in drinking water. Caries Res. 1999;3:3267–74.10.1159/000016528Search in Google Scholar PubMed

[39] Ba Y, Zhu JY, Yang YJ, Yu B, Huang H, Wang G, et al. Serum calciotropic hormone levels, and dental fluorisis in children exposed to different concentrations of fluoride and iodine in drinking water. Chin Med J (Engl). 2010;123:675–9.Search in Google Scholar

[40] Shashi A, Singh JP, Thapar SP. Toxic effects of fluoride on rabbit kidney. Fluoride. 2002;35:38–50.Search in Google Scholar

[41] Czarnowski W, Krechniak J, Urbańska B, Stolarska K, Taraszewska-Czarnowska M, Muraszko-Klaudel A. The impact of water-borne fluoride on bone density. Fluoride. 1999;32:91–5.Search in Google Scholar

[42] Assadollahnejad N, Kargar M, Darabi HR, Abouali N, Jamshidi S, Sharifi A, et al. A new ratiometric, colorimetric and “turn-on” fluorescent chemosensor for the detection of cyanide ions based on a phenol–bisthiazolopyridine hybrid. New J Chem. 2019;43:13001–9.10.1039/C9NJ02863CSearch in Google Scholar

[43] Li Z, Dai Y, Lu Z, Pei Y, Song Y, Zhang L, et al. A Photoswitchable Triple Chemosensor for Cyanide Anion Based on Dicyanovinyl‐Functionalized Dithienylethene. Eur J Org Chem. 2019;22:3614–21.10.1002/ejoc.201900369Search in Google Scholar

[44] Goswami S, Chakrabarty R. Highly Selective Colorimetric Fluorescent Sensor for Pb2+. Eur J Org Chem. 2010;20:3791–5.10.1002/ejoc.201000142Search in Google Scholar

[45] Cho EJ, Ryu BJ, Lee YJ, Nam KC. Visible Colorimetric Fluoride Ion Sensors. Org Lett. 2005;7:2607–9.10.1021/ol0507470Search in Google Scholar PubMed

[46] Chen J, Shu W, Wang, E. A fluorescent and colorimetric probe based on isatin-appended rhodamine for the detection of Hg2+. Chem Res Chin Univ. 2016;32:742–5.10.1007/s40242-016-6001-1Search in Google Scholar

[47] Vine KL, Matesic L, Locke JM, Scropeta D. Recent highlights in the development of isatin-based anticancer agents. Adv. Anticancer Agents Med. Chem. 2013;2:254–312.10.2174/9781608054961113020008Search in Google Scholar

[48] Pandeya SN, Smitha S, Jyoti M, Sridhar SK. Biological activities of isatin and its derivatives. Acta Pharm. 2005;55:27–46.Search in Google Scholar

[49] Medvedev A, Igosheva N, Crumeyrolle-Arias M, Glover V. Isatin: role in stress and anxiety. Stress. 2005;8:175–83.10.1080/10253890500342321Search in Google Scholar PubMed

[50] Medvedev A, Buneeva O, Glover V. Biological targets for isatin and its analogues: implications for therapy. Biol. Targets Ther. 2007;1:151–62.Search in Google Scholar

[51] Shakir M, Abbasi A. Solvent dependant isatin-based Schiff base sensor as fluorescent switch for detection of Cu2+ and S2− in human blood serum. Inorg Chim Acta. 2017; 465:14–25.10.1016/j.ica.2017.04.057Search in Google Scholar

[52] Dhara A, Guchhait N, Kar SK. A novel Cr3+ fluorescence turn-on probe based on rhodamine and isatin framework. J Fluoresc. 2015;25:1921–9.10.1007/s10895-015-1684-0Search in Google Scholar PubMed

[53] Horváth M, Cigáň M, Filo J, Jakusová K, Gáplovský M, Šándrik R. Gáplovský A. Isatin pentafluorophenylhydrazones: interesting conformational change during anion sensing. RSC Advances. 2016;6:109742–50.10.1039/C6RA22396FSearch in Google Scholar

[54] Liu K, Zhao X, Huo J, Neufeld K, Dong M, Zhu X. Anion-induced optical outputs of an isatin-based colorimetric HNI sensor for construction of sequential molecular logic gates and a set-reset memorized device. Sens Actuators B Chem. 2016;226:465–70.10.1016/j.snb.2015.12.019Search in Google Scholar

[55] Hussain M, Bauzá A, Frontera A, Lo KM, Naseer MM. Structure guided or structure guiding? Mixed carbon/hydrogen bonding in bis-Schiff base of N-allyl isatin. CrystEngComm. 2018;20:150–4.10.1039/C7CE01697BSearch in Google Scholar

[56] Pervez H, Khan N, Iqbal J, Zaib S, Yaqub M, Naseer MM. Synthesis and in vitro bio-activity evaluation of N4-benzyl substituted 5-chloroisatin 3-thiosemicarbazones as urease and glycation inhibitors. Acta Chim Slov. 2018;65:108–18.10.17344/acsi.2017.3649Search in Google Scholar

[57] Pervez H, Khan N, Iqbal J, Zaib S, Yaqub M, Tahir MN, et al. Synthesis, crystal structure, molecular docking studies and bio-evaluation of some N4-benzyl substituted isatin-3-thiosemicarbazones as antiurease and antiglycating agents. Heterocycl Commun. 2018;24:51–8.10.1515/hc-2017-0148Search in Google Scholar

[58] Pervez H, Ahmad M, Hadda TB, Toupet L, Naseer MM. Synthesis and fluorine-mediated interactions in methanol-encapsulated solid state self-assembly of an isatin-thiazoline hybrid. J Mol Struct. 2015;1098:124–9.10.1016/j.molstruc.2015.06.013Search in Google Scholar

[59] Ahmad M, Pervez H, Ben Hadda T, Toupet L, Naseer MM. Synthesis and solid state self-assembly of an isatin-thiazoline hybrid driven by three self-complementary dimeric motifs. Tetrahedron Lett. 2014;55:5400–3.10.1016/j.tetlet.2014.08.021Search in Google Scholar

[60] Ahmed M, Hameed S, Ihsan A, Naseer MM. Fluorescent thiazol-substituted pyrazoline nanoparticles for sensitive and highly selective sensing of explosive 2,4,6-trinitrophenol in aqueous medium. Sens Actuators B Chem. 2017;248:57–62.10.1016/j.snb.2017.03.125Search in Google Scholar

[61] The limit of detection (LOD) was calculated from the linear regression graph shown in Figure 3.5 according to the IUPAC definition. The equation is given below: LOD = 3.3σ/k Where K is the slope of the graph and σ is the standard deviation of the y-intercept of the regression line.Search in Google Scholar

[62] Dharma J, Pisal A, Shelton CT. Simple method of measuring the band gap energy value of TiO2 in the powder form using a UV/Vis/NIR spectrometer. Application Note Shelton (CT): PerkinElmer; 2017.Search in Google Scholar

Received: 2019-10-12
Accepted: 2020-01-24
Published Online: 2020-03-17

© 2020 Azeem Haider et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 5.5.2024 from https://www.degruyter.com/document/doi/10.1515/hc-2020-0003/html
Scroll to top button