Next Article in Journal
Heat Illness Requiring Emergency Care for People Experiencing Homelessness: A Case Study Series
Previous Article in Journal
Human Milk for Vulnerable Infants: Breastfeeding and Milk Sharing Practice among Ghanaian Women
Previous Article in Special Issue
Microalgae Diversity in Interim Wet Storage of Spent Nuclear Fuel in Serpong, Indonesia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Search for an Ionic Liquid with the Best Performance during 210Pb/210Bi Cherenkov Counting in Waters on an LS Counter

1
Faculty of Technical Sciences, University of Novi Sad, Trg Dositeja Obradovića 6, 21000 Novi Sad, Serbia
2
Department of Chemistry, Biochemistry and Environmental Protection, Faculty of Sciences, University of Novi Sad, Trg Dositeja Obradovića 3, 21000 Novi Sad, Serbia
3
Department of Physics, Faculty of Sciences, University of Novi Sad, Trg Dositeja Obradovića 3, 21000 Novi Sad, Serbia
4
Laboratory “Ambients and Radiations (Amb.Ra.)”, Department of Computer Engineering, Electrical Engineering and Applied Mathematics (DIEM), University of Salerno, 84084 Fisciano, Italy
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(24), 16563; https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph192416563
Submission received: 10 October 2022 / Revised: 7 December 2022 / Accepted: 8 December 2022 / Published: 9 December 2022
(This article belongs to the Special Issue Radioecology: The Fate of Radionuclides)

Abstract

:
The research presented in this paper aims to investigate the performance of several newly synthesized ionic liquids during 210Pb/210Bi detection in water on a liquid scintillation spectrometer Quantulus 1220 via Cherenkov counting. These experiments have been triggered by the recent reports that certain ionic liquids can act as wavelength shifters, thus significantly increasing the detection efficiency of Cherenkov radiation. The benefit of ionic liquid’s addition to the analysed samples is reflected in the detection limit’s decrement during 210Pb quantification, which is pertinent considering naturally low levels of 210Pb in aqueous samples. Firstly, it was discovered that ionic liquid, 1-butyl-3-methylimidazolium salicylate, is more efficient than the previously explored 2-hydroxypropylammonium salicylate. Consequently, the impact of a few other ionic liquids on Cherenkov counting efficiency with the same cation group (1-butyl-3-methylimidazolium benzoate, 1-butyl-3-methylimidazolium 3-hydroxybenzoate and 1-butyl-3-methylimidazolium 4-hydroxybenzoate) was also explored to test their potential influence. Molecular simulations have been carried out to reveal which structures of ionic liquids assure wavelength-shifting behavior. The obtained results confirmed that, among the investigated ones, only ionic liquids with the salicylate anion exhibited a wavelength shifting effect.

1. Introduction

The occurrence of 210Pb ( T 1 / 2 = 2.26 y) in aquatic environments originates from 238U decay series, and its determination continues to evoke interest in various scientific studies. At first, its presence in waters poses a significant radiological risk. Consequently, its determination is included in the international legislations for the radiological assessment of drinking water, recommending its maximum acceptable concentration of 0.1 Bq l−1 [1]. Additionally, 210Pb and 210Po distribution in the marine environment has been extensively used to determine the removal rates of particles from the ocean and particle fluxes during transport along the coast [2,3], as well as the particulate organic carbon (POC) export in the upper ocean [4].
Determination of 210Pb content in waters is not a trivial task since its environmental concentrations are very low. At the same time, its decay scheme assumes weak gamma transitions and the emissions of low-energetic beta particles [5]. Its direct detection and quantification are possible by gamma spectrometry. Still, it is time-consuming, and it yields relatively high detection limits, with a large sample volume required for the analysis and the high self-absorption of soft gamma-rays in the sample and the detector [5,6].
There are also two possibilities of indirect 210Pb measurement by detecting its progenies 210Po and 210Bi, if they are in radioactive equilibrium with 210Pb. Determination of 210Po via alpha spectrometry can be performed after two spontaneous depositions of 210Po on a silver foil with sufficiently low detection limits, but it demands an in-growth period of several months for 210Po/210Pb equilibrium to take place, introducing long delays in between sampling separation and counting [7]. 210Bi separation from 210Pb is usually followed by its activity determination via Liquid Scintillation Counting (LSC). It provides faster analysis (the required storage periods are shorter since 210Bi/210Pb equilibrium is reached within 40 days), and the acceptable detection limits (several times higher than alpha spectrometry, but still two orders of magnitude lower than the ones for gamma spectrometry). The drawbacks of LSC methods are the need to correct chemical and/or colour quench effects that occur in the samples and cause erroneous results, and, in the case of simultaneous alpha/beta detection, the precise determination of 210Pb counting efficiency is challenging because the alpha/beta spectra subsequent separation is not 100% efficient [7].
One other possibility is to measure 210Pb on an LS counter if it is in equilibrium with its progeny 210Bi through the Cherenkov counting method. Namely, high-energy beta emitter 210Bi ( T 1 / 2 = 5.012(5) d, E β m a x   = 1162.2(8) keV) emits electrons that have enough energy to produce Cherenkov radiation in water [8]. Cherenkov counting advantage lies in easy sample preparation since 210Bi separation from 210Pb is not necessary (the electrons emitted from 210Pb are not able to generate Cherenkov radiation) but has the drawback of lower counting efficiencies than standard LSC methods, typically from 10% [9] up to 20% [10]. Unlike LSC, the generated Cherenkov spectrum is unaffected by the chemical quenching since no scintillation cocktail is added to the sample [11] and has a lower background count rate which provides acceptable detection limits [10].
Recent reports had demonstrated that the excessive Cherenkov radiation signals from 210Bi were detected when ionic liquid (IL) 2-hydroxypropylammonium salicylate, [HPA][Sal], was added to small amounts into the counting vial because this IL acted as the wavelength shifter [12]. Namely, 1.4 g of [HPA][Sal] increased the efficiency for Cherenkov counting during 210Pb/210Bi detection from ~15% to ~60%. Similar research was reported in [13], confirming that one other IL, 1-butyl-3-methylimidazolium 8-hydroxypyrene-1,3,6-trisulfonate, also manifests wavelength-shifting properties, causing the increase of detection efficiency for 18F and 32P via Cherenkov counting for 124% and 14%, respectively. Another IL, 3-methylpiridinium salicylate, had been reported to exhibit both scintillating and wavelength shifting effects, with a significant influence on 210Pb Cherenkov spectra, 210Pb and 226Ra gross alpha/beta spectra, and even 3H spectra generation [14].
The use of ionic liquids as scintillators is desirable due to their unique properties. Ionic liquids are considered as non-volatile solvents [15], which have good thermal and chemical stability [16]. Ionic liquids can be easily recycled and reused in an analytical process without losing efficiency. Combining many cations and anions makes it possible to tune their physical and chemical properties and obtain ionic liquids as an optimal reagent for an industrial or laboratory process [17].
This paper explores several newly synthesized ILs’ impacts on 210Pb/210Bi Cherenkov counting, offering explanations based on their chemical structure. To better understand the mechanism of wavelength-shifting, the influence of the change of cation and anion structure for five different ionic liquids was investigated. For cations were used 2-hydroxypropylammonium and 1-butyl-3-methylimidazolium, and for an anion salicylate, benzoate, 3-hydroxybenzoate and 4-hydroxybenzoate were used. A well-studied [Bmim]+ ion was chosen as the cation to compare whether the presence of an aromatic cation and additional π-π interactions between the cation and hydroxybenzoate affect the scintillating effect. The results were compared with systems containing an aliphatic [HPA]+ cation.
Regarding the change in the anionic structure, a fine structural variation of the position of the hydroxyl group of the corresponding hydroxy benzoates was investigated. The variation of the positions of -ortho, -meta and -para in the series of hydroxy benzoates in previous works has shown a significant effect on different physicochemical properties such as the micellization of surface-active substances as well as on the toxicity of ionic liquids [18,19]. The research conducted in this paper provides deeper insights and answers to the question of which IL structures assure wavelength-shifting behaviour.

2. Materials and Methods

2.1. Instrumentation and Materials

The detection system used for all measurements was Liquid Scintillation Spectrometer Quantulus 1220TM, convenient for the precise low-level radioactivity measurements that are often necessary for environmental radioanalysis. Quantulus 1220 is equipped with the passive shield made from lead, asymmetrically distributed around the detector assembly that absorbs cosmic radiation. The addition of the copper head of the piston absorbs X-rays generated in the lead by cosmic radiation interactions. In addition to this, the detector possess an active shield that further reduces cosmic and environmental gamma radiation. This asymmetric guard counter contains the liquid scintillation filling surrounded with two extra photomultipliers operating in the summed coincidence. This active guard shield operates in anticoincidence with the sample detector and its two photomultipliers that surround the measurement chamber with the vial [20]. Cherenkov spectra were acquired and analysed by WinQ and EASYView software.
To provide the lowest background counting rate and the highest detection efficiency, low diffusion polyethylene vials (Super PE vial Cat.No. 6008117) were selected since 40K that is present in glass vials induces higher background, whilst the vial’s maximum capacity of 20 mL was kept. The comparison of performance of plastic, glass and low 40K glass vials during Cherenkov counting for 90Sr detection was done in our previous work [21]. The experiments were carried out with the calibration samples prepared with the standard radioactive source produced by the Czech Metrology Institute, Inspectorate for Ionizing Radiation, an aqueous 210Pb solution with a certified activity A (210Pb) = 29.55 Bq ml−1 and combined standard uncertainty 1.0% on a reference date 1 October 2013.
After the counting, R C [s−1] and R 0 [s−1] were obtained as the count rates of calibration (reference standard) and background samples, respectively. Since the reference activity of the calibration sample was A [Bq], the detection efficiency was evaluated from the expression:
ε = R C R 0 A   .

2.2. IL Synthesis

Chemical structures of all investigated ILs are presented in Table 1. The provenance and purity of used compounds are given in Table A1 in Appendix A and were used as received without further purification. NMR spectra of all synthesized ILs are provided in Figure A1, Figure A2, Figure A3, Figure A4 and Figure A5 in Appendix A.

2.2.1. Synthesis of 1-Butyl-3-Methylimidazolium-Based Ionic Liquids

The preparation of 1-butyl-3-methylimidazolium benzoate, [Bmim][Ben], was accomplished by mixing the equimolar amounts of 1-butyl-3-methylimidazolium chloride and sodium benzoate. Both compounds were dissolved in acetone and stirred in a round-bottom flask for 5 h. The resulting white precipitate (sodium chloride) was removed by filtering under vacuum, and the clear, pale yellow liquid was obtained. The remaining solvent was removed using a rotary evaporator. After achieving a constant mass of [Bmim][Ben], the obtained product was additionally dried under the vacuum. A similar procedure was used for all the other ionic liquids synthesized in this work. Only different starting compounds were used to obtain different ionic liquids. Therefore, instead of sodium benzoate, sodium salicylate, sodium 3-hydroxybenzoate and sodium 4-hydroxybenzoate were used to obtain [Bmim][Sal], [Bmim][3HB] and [Bmim][4HB], respectively. The residual chloride in the samples was tested using a spot-test by AgNO3.

2.2.2. Synthesis of Ionic Liquid 2-Hydroxypropylammonium Salicylate

Ionic liquid 2-hydroxypropylammonium salicylate, [HPA][Sal], was prepared in the described way. An equimolar amount of salicylic acid was dissolved in methanol, added dropwise to a 1-amino-2-propanol water solution and cooled in an ice bath. After adding the salicylic acid, the reaction mixture was stirred at room temperature for 2 h. The obtained ionic liquid was dried under vacuum for the next 3 h to remove any traces of methanol and water. The obtained liquid product was stored in a vacuum desiccator over P2O5 for the next 48 h.

2.3. Molecular Simulations

Theoretical predicting and describing of the scintillating effect of the examined ionic liquids by the DFT calculations were applied using Jaguar 9.0 software as a part of Schrödinger Materials Science Suite 2015-4. The B3LYP exchange-correlation functional with empirical correction for dispersion (B3LYP–D3) was used with a 6–31 + G(d,p) basis set. Simulations were performed using Generalized Valence Bond Perfect-Pairing (GVB-PP). This pseudospectral method, extended to electron correlation methods, can predict accurate excitation energies, rotational barriers and bond energies. The Continuum solvation model was used, in particular the Generalized Born model. To ensure the validity of the obtained structures, geometrical optimizations were followed by harmonic frequency analysis, where structures without imaginary wavenumbers were considered for further investigation. The intermolecular non–covalent interactions (NCI) have been examined using the method described in the literature [22].
HOMO and LUMO orbital energies were calculated using monitored surfaces (molecular orbitals, density, potential) and the potential charge of the atomic electrostatics. HOMO energy suggests the region of the molecules which can donate an electron, while LUMO energy signifies the capacity of the molecule to accept the electrons. The difference in HOMO and LUMO energy, also known as HOMO–LUMO gap energy, indicates the electronic excitation energy necessary to compute the stability of the compounds and can describe the scintillating potential of a substance.
The Fukui functions are partial derivatives of the electron and spin density concerning a change in either the electron count or the unpaired spin count. A change in the electron count can result from a reaction with another molecule or by any external charge transfer mechanism. A change in the number of the unpaired electron spins would be induced by electromagnetic radiation to produce an electronically excited state of different spin multiplicity. Hence, the scintillating potential of a compound can be predicted from Fukui functions that describe spin density:
f + = ρ N + δ r ρ N r δ ,
f = ρ N δ r ρ N r δ ,
where N is the number of electrons in the reference state of the molecule, and δ is a fraction of an electron.

3. Results and Discussion

3.1. IL’s Influence on the Detection Efficiency of Cherenkov Counting

ILs presented in Table 1 were added in the increasing mass to the prepared sets of 210Pb calibration samples. All samples contained the same 210Pb activity concentration ( A = 4.95(7) Bq) spiked to distilled water so that the total sample volume was kept at 20 mL. Before ILs’ addition, all samples were stored for 50 days to assure that the secular equilibrium between 210Pb and 210Bi was reached, after which they were counted in several cycles for 100 min. Background samples were used to determine M D A (Minimal Detectable Activity) that decreased with the measurement time and were prepared with 20 mL of distilled water transferred into the vial.
Cherenkov radiation detection was carried out through the counting protocol set up manually on Quantulus 1220, the setup configuration was given in the previous research [21]. The only difference in the counting protocol was coincidence bias selection which was set to low instead of high since it was determined that the electrons emitted from 210Bi generated Cherenkov spectrum with better resolution and higher sensitivity on low coincidence bias [12]. According to the expression (1), the obtained count rates were used for the detection efficiency evaluation.
It was reported earlier that [HPA][Sal] efficiently increases the detection efficiency of Cherenkov counting [12]. This is the reason that the first IL that was synthesized was [Bmim][Sal], which also contained salicylate as an anion like [HPA][Sal]. The obtained detection efficiency is presented in Figure 1a, together with the previously published results for [HPA][Sal] addition. In this way, it was possible to compare these two ILs’ effects on Cherenkov spectra and derive some assumptions on which chemical structure acts more efficiently. It is clear that when the amount of the added IL exceeds 0.5 g, [Bmim][Sal] becomes more efficient than [HPA][Sal], causing a more significant efficiency increment, Figure 1a.
Applying [Bmim][Sal] during 210Pb detection in the water samples via the Cherenkov counting method is a significant discovery since natural 210Pb activity concentrations are extremely low. Therefore, it necessitates precise and sensitive methods for its quantification. According to Currie relation [23], if the detection efficiency is higher, it linearly decreases M D A . Therefore, the increment in the detection efficiency from 16% to >70% in the presence of small amounts (around 0.9 g) of [Bmim][Sal] can reduce the detection threshold by more than four times and offers innovative and unmatched improvement in the existing methods.
The explanation in the increment of the count rates with the addition of both ILs will be elaborated in the following paragraphs.
The threshold for Cherenkov photon production varies with the refractive index of a transverse medium. If this parameter is higher, the required energy for Cherenkov radiation generation is lowered, which would cause an excess in the obtained count rates [24]. However, this explanation could not be maintained since the negligible amounts (approximately up to 1 g) of ILs added to 20 mL of water could not significantly alter the sample’s refractive index.
On the other hand, it was noticed that the addition of ILs did not alter the shape or position of Cherenkov spectra, suggesting that ILs could act as wavelength shifters. A substantial fraction of the emitted Cherenkov photons have energies in the ultraviolet region, which are not detectable by the photomultiplier tubes that operate inside an LS counter [25]. ILs absorb ultraviolet photons and re-emit them at longer wavelengths, thus inducing the wavelength shifting effect. Consequently, the photomultiplier tubes detect more light and the counting efficiency increases.
Another explanation for the efficiency increment could be that ILs induce a scintillating effect. This was investigated when spectra of the samples containing only 210Pb standard and the samples of 210Pb standard + maximal amount of the added ILs were inspected after their counting on the default gross alpha/beta counting protocol on Quantulus 1220. Namely, the gross alpha spectrum was not generated in any case. If ILs acted as a scintillator, 210Po alpha spectrum (210Pb’s progeny) should appear to some extent in the samples that contain ILs.
The conclusion on the two investigated ILs’ behaviour is that their presence does not influence the sample’s refractive index and cannot affect the Cherenkov threshold. These two ILs do not exhibit scintillating properties (if they acted like scintillators, there is no explanation why 210Po alpha spectrum was not generated in their presence). Cherenkov emissions from 210Bi generated excessive signals in the spectra with ILs’ addition due to the spectral wavelength shifting effect.
The fact that [Bmim][Sal] had better performance than [HPA][Sal] during Cherenkov detection induced a hypothesis that 1-butyl-3-methylimidazolium cation structure is responsible for its greater efficiency. This possible explanation had to be tested, thus triggering another research, a synthesis of a few novels ILs with the same cation, 1-butyl-3-methylimidazolium-benzoate, [Bmim][Ben], 1-butyl-3-methylimidazolium 3-hydroxybenzoate [Bmim][3HB] and 1-butyl-3-methylimidazolium 4-hydroxybenzoate [Bmim][4HB]. The sets of samples were prepared in the same manner as the previous calibration samples, and were spiked with the increasing amounts of these ILs. The obtained efficiencies for all four ILs with the same cation structure are displayed in Figure 1b.
Since no other IL except [Bmim][Sal] caused the detection efficiency increment, it should be stated that the cation structure itself does not influence the count rate increment. It is interesting to note that [Bmim][3HB] and [Bmim][4HB] even caused a mild decrement in detection efficiency. They induce colour quench, to which Cherenkov counting is sensitive and demands corrections. Namely, Cherenkov pulse height distribution is being shifted towards lower energy channels in the spectrum. It is generated with lesser intensity in the coloured samples, decreasing the obtained count rate and the counting efficiency [26].

Reproducibility Investigation

Finally, reproducibility was considered a dispersion of the detection efficiencies obtained in replicates of calibration samples prepared with the same 210Pb activity concentration spiked to counting vials, but with [Bmim][Sal] that was synthesized twice. The set with IL that was firstly synthesized was denoted as the first experiment, while the second experiment assumed the addition of IL that was synthesized after the first one. Approximately the same amount of the increasing mass of [Bmim][Sal] was added to two sets of calibration samples. The results of measurements are displayed in Figure 2.
The graph in Figure 2a shows the ratio ε / ε 0 for the different mass addition of [Bmim][Sal] to two sets of samples (the added amounts could not be precisely equal between the sets), where ε is Cherenkov detection efficiency obtained in the presence of a certain mass of [Bmim][Sal] and ε 0 is Cherenkov detection efficiency in spiked 210Pb samples without the addition of [Bmim][Sal], ε 0 ≈ 16%. All results are mutually consistent within the measurement uncertainty for the two samples with similar masses of added [Bmim][Sal]. Furthermore, the correlation function between the obtained detection efficiencies in Figure 2b is rather satisfying, y = 1.11 6 x 0.029 19 , with the correlation coefficient close to 1. This value, 1.11(6) was obtained for the samples that do not contain precisely the same mass of [Bmim][Sal], and it would be even closer to 1 if the added mass matched better between the sets of samples. Therefore, it can be concluded that the method with the addition of ILs during Cherenkov counting provides results with excellent reproducibility.

3.2. Results of Molecular Simulations

The DFT calculations were performed to examine the scintillating activity of synthesized ionic liquids. To obtain a more realistic simulation of the transition state geometry, all simulations were performed using the generalized valence bond settings. Figure 3 shows optimized geometrical structures with the representation of noncovalent interactions.
The charge density around all ionic liquids used in this work is presented in Figure 4. From Figure 4, it can be seen that the blue regions represent positive charge, along with the red areas representing negative charge.
Substances with good scintillating properties are substances that can efficiently produce an excited state. This is usually characteristic of molecules with accessible excited states at lower energies and non-bonded π electrons that can easily be excited. Computational descriptors are good for explaining the excitation potential of molecules by predicting energy for the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO). The values of HOMO and LUMO energies together with energy gaps ( Δ E g a p ) are presented in Table 2. The energy gaps are calculated with the following equation:
Δ E g a p = E LUMO E HOMO .
The rigidity of molecules is significant for wavelength shifters. It is achieved either by condensing aromatic molecules or connecting them to allow efficient delocalization of electrons via unhybridized 2pz atomic orbitals [27]. In the ILs’ case, the structural rigidity of the ion pair can be achieved by increasing the number of non-covalent interactions between cation and anion. The rigidity of the molecule and the efficiency of electron delocalization through the system directly affect the values of the HOMO and LUMO orbitals energies. Our previous research [14] has shown that the energy gap between the HOMO and LUMO orbitals (energy gap) significantly impacts the scintillating and wave-shifting ability of compounds and can be an important parameter for evaluating the efficiency of a compound in an LS counter.
As shown in Table 2, the HOMO-LUMO energy gaps of [Bmim][Sal] ionic liquids are slightly higher than those of other ionic liquids. This data indicates that [Bmim][Sal] demands more energy to become excited and emit a photon. The HOMO and LUMO orbitals were performed and shown in Figure 5 and Figure 6 to understand better and further illustrate differences in the electron structure of all ionic liquids investigated in this research.
The Atomic Fukui indices were calculated to distinguish and quantify differences in HOMO and LUMO orbitals more precisely. From Fukui indices, it can be seen that each index includes two subscripts that can take the values N or S, which represent the electron density and the spin density, respectively. The nature of the partial derivative on which the index is based is indicated with two indices. The first index indicates the property that responds to a change in the property indicated by the second index. According to this, f_NS indicates the change in electron density about an atom when the molecule undergoes a reaction in which its spin multiplicity changes. It is established as a descriptor of how easily a molecule can be excited. If the value of an index is higher, the higher is the change in electron or spin density near the atom of interest. The enumeration of ionic liquids and analysis of HOMO f_NS and LUMO f_NS indices is presented in Figure 7 and Figure 8.
From Figure 8, it can be seen that [HPA][Sal] ionic liquid has higher and positive values of HOMO orbitals than those of the other ionic liquids. The most positive f_NS values indicate the largest changes in the electron density at HOMO orbitals of this ionic liquid when a reaction occurs where the shape of the spin changes. It is more important to pay attention to the values of LUMO f_NS orbitals. By comparing ionic liquids with the same cation and a different anion, it is concluded that [Bmim]+ has better properties as a cation because the whole ring has positive LUMO f_NS values. This means that the ring of [Bmim]+ cation can be excited, which is not the case for [HPA]+ cation. If we compare ionic liquids with the same cation and different anion, it can be seen that [Ben] anion, which does not have a hydroxyl group, is inert in terms of the electron density.
Additionally, from Figure 8, it can be seen that most values of LUMO f_NS for [Bmim][Ben] are near zero. Additionally, from Figure 8, we can see that the values of LUMO f_NS are concentrated around the oxygen atom of the carboxyl group (numeration O19) with a pronounced negative value which indicates that in the case of this ionic liquid excitation will not be good. Significant changes occur with introducing the hydroxyl group into the anion structure. In the case of ionic liquid with [Sal] anion, LUMO f_NS values for the oxygen atom (O19) are slightly positive. From this obtained data, it can be concluded that [Bmim][Sal] ionic liquid is the most active scintillator because all LUMO f_NS values around the cation ring are positive. Results obtained from molecular simulations are in accordance with the previous experimental results.

4. Conclusions

This work explored the performance of several newly synthesized ionic liquids during 210Pb/210Bi Cherenkov counting on an LS counter Quantulus 1220, explaining their behaviour based on ILs’ structure. Among the few ILs with the same cation structure, 1-butyl-3-methylimidazolium [Bmim]+, the only one that significantly increased the detection efficiency was [Bmim][Sal], which contained salicylate as an anion. Moreover, it was shown that [Bmim][Sal] had a more significant impact on the efficiency in comparison with the previously investigated 2-hydroxypropylammonium salicylate, [HPA][Sal]. Other ILs, the one with benzoate anion, [Bmim][Ben], did not increase the obtained count rates, while the structures with 3-hydroxybenzoate and 4-hydroxybenzoate anions, [Bmim][3HB] and [Bmim][4HB], respectively, induced mild colour quench, even reducing the initial detection efficiency. ILs’ behaviour could be explained via analysis of their HOMO f_NS and LUMO f_NS values. The presented research confirmed that salicylates act as wavelength shifters, consequently increasing the detection efficiency of Cherenkov counting. This finding can be very useful during the detection of naturally low levels of 210Pb in waters via the Cherenkov counting method, since the increment in detection efficiency in the presence of small amounts of [Bmim][Sal], about 0.9 g, can reduce detection threshold more than four times. Moreover, ILs could be applied in quantifying other radionuclides besides 210Pb/210Bi via Cherenkov counting. Further research on the synthesis of ionic liquids with a more extensive range of HOMO-LUMO orbitals, i.e., energy gaps, is needed to identify the potential limit value required for the wavelength shifting effect.

Author Contributions

Conceptualization, M.V. and N.T.; methodology, I.S. and M.V.; software, M.V. and T.T.B.; validation, J.N. and N.T.; formal analysis, I.S., M.Z., and T.T.B.; investigation, J.N., T.T.B. and M.Z.; resources, S.G.; writing—original draft preparation, I.S. and M.V.; writing—review and editing, S.M. and S.G.; visualization, J.N., S.M. and M.Z.; supervision, N.T. and S.M.; project administration, N.T. and S.G.; funding acquisition, S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Provincial Secretariat for Higher Education and Scientific Research, Republic of Serbia, grant number 142-451-2373/2021-01 (“Development of environmentally safe method based on ionic liquids for the investigation of radiological content of drinking waters in Vojvodina”), and the Ministry of Education, Science and Technological Development of the Republic of Serbia, grant numbers 451-03-68/2022-14/200125 and 451-03-68/2022-14/200156 (“Innovative scientific and artistic research from the FTS (activity) domain”).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Appendix A

Table A1. Provenance, purity, structure, and molar mass of the used chemicals.
Table A1. Provenance, purity, structure, and molar mass of the used chemicals.
Chemical NameStructureSourceCAS NumberMolar MassMass Fraction Purity
1-amino-2-propanolIjerph 19 16563 i006Sigma Aldrich48-96-675.1193%
1-butyl-3-methylimidazolium chlorideIjerph 19 16563 i007TCI Europe79917-90-1174.67≥98%
Salicylic acidIjerph 19 16563 i008Sigma Aldrich69-72-7138.12≥99%
Sodium benzoateIjerph 19 16563 i009TCI Europe532-32-1144.10>98%
Sodium salicylateIjerph 19 16563 i010Sigma Aldrich54-21-7160.11≥99.5%
Sodium-3-hydroxybenzoateIjerph 19 16563 i011TCI Europe7720-19-6160.11≥99%
Sodium-4-hydroxybenzoateIjerph 19 16563 i012Sigma Aldrich114-63-6160.11≥96%
AcetoneIjerph 19 16563 i013Lachner67-64-158.08≥99%
MethanolIjerph 19 16563 i014Sigma Aldrich57-56-132.04≥99.8%
Figure A1. 1H and 13C NMR spectra for [Bmim][3HB].
Figure A1. 1H and 13C NMR spectra for [Bmim][3HB].
Ijerph 19 16563 g0a1aIjerph 19 16563 g0a1b
Figure A2. 1H and 13C NMR spectra for [Bmim][4HB].
Figure A2. 1H and 13C NMR spectra for [Bmim][4HB].
Ijerph 19 16563 g0a2aIjerph 19 16563 g0a2b
Figure A3. 1H and 13C NMR spectra for [HPA][Sal].
Figure A3. 1H and 13C NMR spectra for [HPA][Sal].
Ijerph 19 16563 g0a3aIjerph 19 16563 g0a3b
Figure A4. 1H and 13C NMR spectra for [Bmim][Sal].
Figure A4. 1H and 13C NMR spectra for [Bmim][Sal].
Ijerph 19 16563 g0a4aIjerph 19 16563 g0a4b
Figure A5. 1H and 13C NMR spectra for [Bmim][Ben].
Figure A5. 1H and 13C NMR spectra for [Bmim][Ben].
Ijerph 19 16563 g0a5aIjerph 19 16563 g0a5b

References

  1. World Health Organization. Guidelines for Drinking-Water Quality, 4th ed.; WHO Press: Genève, Switzerland, 2011; ISBN 9789241548151. [Google Scholar]
  2. Wei, C.-L.; Lin, S.-Y.; Wen, L.-S.; Sheu, D.-D. Geochemical behavior of 210Pb and 210Po in the nearshore waters off western Taiwan. Mar. Pollut. Bull. 2012, 64, 214–220. [Google Scholar] [CrossRef]
  3. Murray, J.W.; Paul, B.; Dunne, J.P.; Chapin, T. 234Th, 210Pb, 210Po and stable Pb in the central equatorial Pacific: Tracers for particle cycling. Deep-Sea Res. Pt. I 2005, 52, 2109–2139. [Google Scholar] [CrossRef]
  4. Kim, T.H.; Kim, G. Important role of colloids in the cycling of 210Po and 210Pb in the ocean: Results from the East/Japan Sea. Geochim. Cosmochim. Acta 2012, 95, 134–142. [Google Scholar] [CrossRef]
  5. Larivière, D.; Reiber, K.M.; Evans, R.D.; Cornett, R.J. Determination of 210Pb at ultra-trace levels in water by ICP-MS. Anal. Chim. Acta 2005, 549, 188–196. [Google Scholar] [CrossRef]
  6. Grahek, Ž.; Rožmarić Mačefat, M.; Lulić, S. Isolation of lead from water samples and determination of 210Pb. Anal. Chim. Acta 2006, 560, 84–93. [Google Scholar] [CrossRef]
  7. Villa, M.; Hurtado, S.; Manjón, G.; García-Tenorio, R. Calibration and measurement of 210Pb using two independent techniques. Radiat. Meas. 2007, 42, 1552–1560. [Google Scholar] [CrossRef]
  8. Arinc, A.; Johansson, L.C.; Gilligan, C.R.D.; Pearce, A.K. Standardization of 210Pb by Čerenkov counting. Appl. Radiat. Isot. 2011, 69, 768–772. [Google Scholar] [CrossRef]
  9. Johansson, L.Y. Determination of Pb-210 and Po-210 in Aqueous Environmental Samples. Doctoral Dissertation, Faculty of Natural Sciences, Gottfried Wilhelm Leibniz Universität Hannover, Hannover, Germany, 28 July 2008. Available online: https://www.nrc.gov/docs/ML1105/ML110560301.pdf (accessed on 25 February 2022).
  10. Al-Marri, M.S.; Hamwi, A.; Mikhlallaty, H. Radiochemical determination of lead-210 in environmental water samples using Cerenkov counting. J. Radioanal. Nucl. Chem. 1997, 219, 73–75. [Google Scholar] [CrossRef]
  11. Manjon, G.; Absi, A.; Gomez, E.; Vaca, F.; Garcia-Leon, M. Efficiency and background in Cerenkov counting affected by color quenching: An optical study. In Proceedings of the LSC 2001, Advances in Liquid Scintillation Spectrometry, Karlsruhe, Germany, 7–11 May 2001. [Google Scholar]
  12. Stojković, I.; Todorović, N.; Nikolov, J.; Tenjović, B.; Gadžurić, S.; Tot, A.; Vraneš, M. 210Pb/210Bi detection in waters by Cherenkov counting—Perspectives and new possibilities. Radiat. Phys. Chem. 2020, 166, 108474. [Google Scholar] [CrossRef]
  13. Mirenda, M.; Rodrigues, D.; Ferreyra, C.; Arenillas, P.; Sarmiento, G.P.; Krimer, N.; Japas, M.L. Ionic liquids as solvents for Čerenkov counting and the effect of a wavelength shifter. Appl. Radiat. Isot. 2018, 134, 275–279. [Google Scholar] [CrossRef]
  14. Vraneš, M.; Stojković, I.; Tot, A.; Todorović, N.; Nikolov, J.; Gadžurić, S. Scintillating and wavelength shifting effect investigation of 3-methylpiridinium salicylate and its application in LSC measurements. Appl. Radiat. Isot. 2021, 172, 109697. [Google Scholar] [CrossRef] [PubMed]
  15. Berthod, A.; Ruiz-Ángel, M.J.; Carda-Broch, S. Recent advances on ionic liquid uses in separation techniques. J. Chromatogr. A 2018, 1559, 2–16. [Google Scholar] [CrossRef] [PubMed]
  16. Baranyai, K.J.; Deacon, G.B.; MacFarlane, D.R.; Pringle, J.M.; Scott, J.L. Thermal degradation of ionic liquids at elevated temperatures. Aust. J. Chem. 2004, 57, 145–147. [Google Scholar] [CrossRef]
  17. Brennecke, J.F.; Rogers, R.D.; Seddon, K.R. Ionic Liquids IV.; ACS Symposium series; American Chemical Society: Washington, WA, USA, 2007; pp. 13–15. ISBN 9780841274457. [Google Scholar]
  18. Vraneš, M.; Tot, A.; Jovanović-Šanta, S.; Karaman, M.; Dožić, S.; Tešanović, K.; Kojić, V.; Gadžurić, S. Toxicity reduction of imidazolium-based ionic liquids by the oxygenation of the alkyl substituent. RSC Adv. 2016, 6, 96289–96295. [Google Scholar] [CrossRef]
  19. Čobanov, I.; Šarac, B.; Medoš, Ž.; Tot, A.; Vraneš, M.; Gadžurić, S.; Bešter-Rogač, M. Thermodynamic and computational study of isomerism effect at micellization of imidazolium based surface-active ionic liquids: Counterion structure. J. Mol. Liq. 2020, 301, 112419. [Google Scholar] [CrossRef]
  20. PerkinElmer Life Sciences. Wallac 1220 Quantulus. In Instrument Manual, Ultra Low Level Liquid Scintillation Spectrometer; PerkinElmer: Turku, Finland, 2002. [Google Scholar]
  21. Todorović, N.; Stojković, I.; Nikolov, J.; Tenjović, B. 90Sr determination in water samples using Čerenkov radiation. J. Environ. Radioact. 2017, 169, 197–202. [Google Scholar] [CrossRef]
  22. Johnson, E.R.; Keinan, S.; Mori-Sanchez, P.; Contreras-Garcia, J.; Cohen, A.J.; Yang, W. Revealing noncovalent interactions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef] [Green Version]
  23. Currie, L.A. Limits for qualitative detection and quantification determination. Application to radiochemistry. Anal. Chem. 1968, 40, 586–593. [Google Scholar] [CrossRef]
  24. L’Annunziata, M.F. Handbook of Radioactivity Analysis, 3rd ed.; Academic Press: Cambridge, MA, USA, 2012; ISBN 978-0-12-384873-4. [Google Scholar]
  25. Wang, Y.; Yang, Y.; Song, L.; Ma, Y.; Luo, M.; Dai, X. Effects of sodium salicylate on the determination of Lead-210/Bismuth-210 by Cerenkov counting. Appl. Radiat. Isot. 2018, 139, 175–180. [Google Scholar] [CrossRef]
  26. Vaca, F.; Manjon, G.; Garcia-Leon, M. Efficiency calibration of a liquid scintillation counter for 90Y Cherenkov counting. Nucl. Instrum. Meth. A 1998, 406, 267–275. [Google Scholar] [CrossRef]
  27. Kuźniak, M.; Szelc, A.M. Wavelength Shifters for Applications in Liquid Argon Detectors. Instruments 2021, 5, 4. [Google Scholar] [CrossRef]
Figure 1. Comparison of IL’s influence on 210Pb/210Bi detection efficiency: (a) [HPA][Sal] and [Bmim][Sal]; (b) Several ILs with [Bmim]+ cations.
Figure 1. Comparison of IL’s influence on 210Pb/210Bi detection efficiency: (a) [HPA][Sal] and [Bmim][Sal]; (b) Several ILs with [Bmim]+ cations.
Ijerph 19 16563 g001
Figure 2. Reproducibility tests for the two independent experiments: (a) Efficiency ratio for similar addition of [Bmim][Sal]; (b) Correlation between the obtained efficiencies.
Figure 2. Reproducibility tests for the two independent experiments: (a) Efficiency ratio for similar addition of [Bmim][Sal]; (b) Correlation between the obtained efficiencies.
Ijerph 19 16563 g002
Figure 3. Optimized structures with a representation of noncovalent interactions for: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Figure 3. Optimized structures with a representation of noncovalent interactions for: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Ijerph 19 16563 g003
Figure 4. The pictorial representation of a charge density for: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Figure 4. The pictorial representation of a charge density for: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Ijerph 19 16563 g004
Figure 5. Representation of HOMO orbitals of: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Figure 5. Representation of HOMO orbitals of: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Ijerph 19 16563 g005
Figure 6. Representation of LUMO orbitals of: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Figure 6. Representation of LUMO orbitals of: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Ijerph 19 16563 g006aIjerph 19 16563 g006b
Figure 7. The structure of ionic liquids and the atom-numbering scheme of: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Figure 7. The structure of ionic liquids and the atom-numbering scheme of: (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Ijerph 19 16563 g007aIjerph 19 16563 g007b
Figure 8. Representation of Fukui indices (black represents HOMO f_NS and red represents LUMO f_NS): (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Figure 8. Representation of Fukui indices (black represents HOMO f_NS and red represents LUMO f_NS): (a) [Bmim][3HB]; (b) [Bmim][4HB]; (c) [HPA][Sal]; (d) [Bmim][Sal]; (e) [Bmim][Ben].
Ijerph 19 16563 g008
Table 1. Chemical structures of synthesized ionic liquids.
Table 1. Chemical structures of synthesized ionic liquids.
Ionic LiquidChemical Structure
1-butyl-3-methylimidazolium 3-hydroxybenzoate [Bmim][3HB]Ijerph 19 16563 i001
1-butyl-3-methylimidazolium 4-hydroxybenzoate [Bmim][4HB]Ijerph 19 16563 i002
2-hydroxypropylammonium salicylate [HPA][Sal]Ijerph 19 16563 i003
1-butyl-3-methylimidazolium salicylate [Bmim][Sal]Ijerph 19 16563 i004
1-butyl-3-methylimidazolium benzoate [Bmim][Ben]Ijerph 19 16563 i005
Table 2. HOMO/LUMO energies and energy gaps for [Bmim][3HB], [Bmim][4HB], [HPA][Sal], [Bmim][Sal], and [Bmim][Ben].
Table 2. HOMO/LUMO energies and energy gaps for [Bmim][3HB], [Bmim][4HB], [HPA][Sal], [Bmim][Sal], and [Bmim][Ben].
Compound E H O M O   [ e V ] E L U M O   [ e V ] Δ E g a p   [ e V ]
[Bmim][3HB]−5.123−0.5204.603
[Bmim][4HB]−4.791−1.1793.612
[HPA][Sal]−6.002−1.0754.927
[Bmim][Sal]−6.024−0.7785.245
[Bmim][Ben]−5.292−0.5374.755
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Stojković, I.; Vraneš, M.; Borović, T.T.; Todorović, N.; Nikolov, J.; Zečević, M.; Gadžurić, S.; Mancini, S. In Search for an Ionic Liquid with the Best Performance during 210Pb/210Bi Cherenkov Counting in Waters on an LS Counter. Int. J. Environ. Res. Public Health 2022, 19, 16563. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph192416563

AMA Style

Stojković I, Vraneš M, Borović TT, Todorović N, Nikolov J, Zečević M, Gadžurić S, Mancini S. In Search for an Ionic Liquid with the Best Performance during 210Pb/210Bi Cherenkov Counting in Waters on an LS Counter. International Journal of Environmental Research and Public Health. 2022; 19(24):16563. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph192416563

Chicago/Turabian Style

Stojković, Ivana, Milan Vraneš, Teona Teodora Borović, Nataša Todorović, Jovana Nikolov, Milka Zečević, Slobodan Gadžurić, and Simona Mancini. 2022. "In Search for an Ionic Liquid with the Best Performance during 210Pb/210Bi Cherenkov Counting in Waters on an LS Counter" International Journal of Environmental Research and Public Health 19, no. 24: 16563. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph192416563

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop