Abhijeet
Rana
and
Shyam
Biswas
*
Department of Chemistry, Indian Institute of Technology Guwahati, Guwahati, 781039 Assam, India. E-mail: sbiswas@iitg.ac.in
First published on 29th March 2023
Sulfur is a soft Lewis base and thiocarbonyl has moderate electrophilicity. Our probe's properties are adjusted in such a way that it could simultaneously detect a Lewis acid (Hg2+) and a strong nucleophile (NH2–NH2). Considering the above fact, a thioureido-functionalized robust MOF material was prepared, which was utilized for the selective fluorometric detection of environmentally significant toxic pollutants (Hg2+ and NH2–NH2) in an aqueous medium. The probe detected mercury by quenching the fluorescence emission intensity in a static pathway by soft-soft interaction with the S-atom of the probe. In contrast, the detection of hydrazine was furnished by a reaction-based pathway by the attack of hydrazine to the moderate electrophilic (thiocarbonyl) part of the probe, which resulted in an enhancement in the fluorescence emission intensity of the probe. A MOF-coated cotton composite was developed for the naked-eye detection of Hg2+ and NH2–NH2 for real-life applicability. The MOF was highly sensitive towards detecting Hg2+ and NH2–NH2 with very low detection times, i.e., 10 s and 50 s, respectively. The probe's sensitivity also remained unaltered under a significantly lower concentration of the targeted analytes, i.e., the detection limits for NH2–NH2 and Hg2+ were 1.9 nM and 4.0 nM, respectively. Our probe's response time and LOD are much lower than the other previously reported probes for Hg2+ and NH2–NH2 to date. A 92% fluorescence quenching and 28-fold fluorescence enhancement of the MOF were observed after the interaction of the probe with Hg2+ and NH2–NH2, respectively. The probe has excellent selectivity over the competitive analytes of Hg2+ and NH2–NH2. The MOF could also sense hydrazine in various environmental water specimens. Systematic mechanistic investigations were conducted to know the phenomena behind the fluorescence quenching and enhancement processes.
Hydrazine is used as a catalyst,6 propellant,7,8 blowing agent,9etc. The accumulation of hydrazine into the environment occurs though the discharge of industrial waste. Overexposure to hydrazine can cause detrimental effects on human health. The threshold concentration of hydrazine in drinking water has been set as 10 ppb.10 Therefore, the detection of hydrazine below the safe limit is highly required.
Looking into the detrimental effects of mercury and hydrazine in drinking water, many environmental scientists working in the field of sensing hazardous materials have put forward their contribution by several methods. The selective sensing of hazardous materials has been reported by enormous research groups with the help of ion-exchange chromatography, electrochemical,11 spectrometry,12 voltammetry,13 and fluorescence-based methods38.14–16 Among the abovementioned techniques, the fluorescence-based method has several advantages over other techniques due to the easy handling and simple observation process.17 There are many class of materials utilized for the sensing of hazardous materials including quantum dots,18,19 graphene oxide-based materials,20 carbon nanotubes,21 and organic probes.22 Metal–organic frameworks (MOFs) are porous with very high surface areas and their active sites for sensing could be easily tuned. Therefore, MOF-based materials are advantageous over other classes of materials.
The abovementioned toxic properties of hydrazine and mercury and the easy handling fluorescence method inspired us to develop a fluorescence-based selective probe to detect mercury and hydrazine. The soft nature of mercury pushed us to design a soft center-based probe with sulfur-containing functionality. In contrast, the nucleophilic nature of hydrazine having alpha-effect inspired us to design a probe with moderate electrophilic center. The moderate electrophilic nature guides the probe to selectively react with hydrazine even in the presence of other nucleophilic congeners. The above idea was incorporated by designing a thioureido-functionalized robust Hf-MOF named as Hf-UiO-66-NHCSNHCH3. The soft sulfur atom and moderate electrophilic thiocarbonyl group was the active centers for the selective interaction of Hg2+ and NH2–NH2 with the probe. A 28-fold increment in the fluorescence intensity of Hf-UiO-66-NHCSNHCH3 was observed after the addition of hydrazine solution and a fluorescence quenching efficiency of 92% was observed after adding an aqueous Hg2+ solution. The probe can detect Hg2+ and NH2–NH2 below the permissible limit in drinking water. A thorough study of the mechanistic pathway was explored via various analytical techniques to present the work systematically and be applicable in future endeavours. The selectivity and applicability for real field purposes, the ultralow detection limit, and the relevance of dual sensing purpose make our material a unique sensor of mercury and hydrazine.
The FT-IR spectra of Hf-UiO-66-NH2 MOF possesses two peaks (at 1575 and 1385 cm−1) due to the asymmetric and symmetric stretching frequency of the carboxylic group of Hf-UiO-66-NH2. These two carboxyl frequencies are present in Hf-UiO-66-NHCSNHCH3 MOF, which proves that there is no structural detachment of the carboxylate from the metal center during post-synthetic modification. The FT-IR spectra showed additional peaks at 1280, 1085, and 798 cm−1 due to the –CS bond of Hf-UiO-66-NHCSNHCH3 MOF,24 which confirmed the successful post-synthetic modification (Fig. S7†).
The FE-SEM images (Fig. 1d) of Hf-UiO-66-NHCSNHCH3 and Hf-UiO-66-NH2 demonstrate similar particles, which confirmed no change in the crystallinity of the material after post-synthetic modification. The successful post-synthetic modification was again supported by the EDX elemental analysis. The TEM-EDX elemental analysis of Hf-UiO-66-NH2 confirmed the presence of Hf (5.1%) (Fig. S8†). At the same time, the TEM-EDX elemental analysis of Hf-UiO-66-NHCSNHCH3 confirmed the presence of Hf (4.6%) and S (4.5%) elements (Fig. S10†). The TEM-EDX elemental mapping was also carried out for Hf-UiO-66-NH2 and Hf-UiO-66-NHCSNHCH3, which displayed the homogenous distribution of all the desired elements (Fig. S9 and S11†). The presence of the additional sulfur atom in Hf-UiO-66-NHCSNHCH3 confirmed the incorporation of the isothiocyanate group into the free amine of Hf-UiO-66-NH2. The homogenous distribution of the desired elements in the elemental mapping of Hf-UiO-66-NHCSNHCH3 suggests that the methyl isothiocyanate molecule was not physically adsorbed on the surface of Hf-UiO-66-NH2 but it was bonded with the –NH2 group of Hf-UiO-66-NH2.
Further, the PXRD analysis of Hf-UiO-66-NHCSNHCH3 and 1′@NH2 was performed to confirm their phase purity and crystallinity. Both materials showed similar PXRD patterns such as the simulated one (Fig. 1). The PXRD experiment demonstrated that the above-synthesized materials Hf-UiO-66-NHCSNHCH3 and Hf-UiO-66-NH2 belong to the UiO-66 topology. The Pawley fit and indexing data (Fig. S12 and Table S1†) of the slow scan PXRD data of Hf-UiO-66-NHCSNHCH3 again supported this observation. The Pawley fit data showed that the PXRD pattern of Hf-UiO-66-NHCSNHCH3 exactly fits with the simulated pattern with negligible errors (Rw = 1.0% and Rwp = 1.6%) (Fig. S12†).
The chemical robustness of the material was also examined by stirring the materials in DCM, DMF, H2O, pH 2, and pH 12 solutions. The materials were filtered and the PXRD patterns of the recovered materials were measured individually. The PXRD pattern of the recovered materials precisely matched with the PXRD pattern of Hf-UiO-66-NHCSNHCH3 (Fig. S13†). It concludes that the crystallinity of the material remained unchanged even after 24 h of stirring.37 Therefore, the probe Hf-UiO-66-NHCSNHCH3 is stable enough for the application of the sensing purpose in a variety of solvent media.
The thermogravimetric (TG) experiments for both Hf-UiO-66-NHCSNHCH3 and Hf-UiO-66-NH2 were performed to know the material's thermal stability. Initially, there was a weight loss of 2.9% in the TG-curve of Hf-UiO-66-NH2 due to the removal of 1.3 molecules of water per unit formula of Hf-UiO-66-NH2 at 130 °C (Fig. S14†). The second weight loss was attributed to the loss of the framework structure at 400 °C. The absence of breakpoint due to DMF in the TG trace confirmed the proper activation of Hf-UiO-66-NH2. The material Hf-UiO-66-NHCSNHCH3 also displayed a similar TG-curve with an initial weight loss of 3.7% due to the loss 1.5 molecules of water per unit formula and the second weight loss due to linker dislocation from the framework structure starting from 350 °C to 400 °C. The framework destruction of Hf-UiO-66-NHCSNHCH3 occurred at a slightly lower temperature compared to Hf-UiO-66-NH2 due to the linker defect that arose at the time of post-synthetic modification.
N2 sorption analysis of Hf-UiO-66-NHCSNHCH3 was carried out at −196 °C using liquid nitrogen. The determined surface area for Hf-UiO-66-NHCSNHCH3 was 498 m2 g−1. We noticed a decreased surface area of Hf-UiO-66-NHCSNHCH3 compared to the surface area (784 m2 g−1) of the previously reported Hf-UiO-66-NH2 MOF (Fig. S15†).25 The reduced surface area is attributed to the post-synthetic modification. Because of post-synthetic transformation, the extra functionality of methyl isothiocyanate occupied the pores of Hf-UiO-66-NH2, which caused a decrease in the surface area of Hf-UiO-66-NHCSNHCH3. The decreased surface area of Hf-UiO-66-NHCSNHCH3 again supports the successful incorporation of the methyl isothiocyanate moiety via the free amine of Hf-UiO-66-NH2 (Fig. S16†).
A volume-dependent fluorescence detection experiment was carried out by adding 50 μL 10 mM aqueous Hg2+ solution to 3000 μL MOF suspension in DMF in each step. A sudden quenching in the fluorescence emission intensity was observed after every incremental step and, finally, a saturation of the quenching process was noticed after the addition of 300 μL aqueous Hg2+ solution, as shown in Fig. 2a. A similar volume-dependent sensing experiment was carried out by adding 50 μL 10 mM aqueous hydrazine solution to 3000 μL aqueous MOF suspension in each step. An immediate turn-on in the fluorescence emission intensity was noticed after each incremental addition and, finally, a saturation in fluorescence intensity was detected after the addition of 300 μL 10 mM aqueous hydrazine solution, as displayed in Fig. 2b. The detection of targeted analytes should be reproducible and repeatable, which is a good characteristic of the ideal sensor material. Therefore, we carried out batch experiments multiple times on the same day and different days. The experimental results presented in Tables S2 and S3† confirmed that the sensing processes are reproducible and repeatable for providing the same results toward both the analytes with high precision and accuracy.
Fig. 2 Change in the fluorescence emission intensity of the probe Hf-UiO-66-NHCSNHCH3 after the incremental addition of 300 μL aqueous 10 mM Hg2+ (a) and NH2–NH2 (b) solution. |
The time-dependent experiment was performed to determine the sensitivity of detection. To find the detection time, 300 μL 10 mM aqueous Hg2+ solution was added to a 3000 μL MOF suspension in DMF, and the fluorescence emission intensity was recorded after every 10 s intervals. We found that after 10 s, there was a saturation in the fluorescence intensity and no further appreciable quenching occurred up to 1 min (Fig. 3a and b). A similar procedure was adopted to find the probe's response time for hydrazine detection. In this case, we found that the turn-on nature of the fluorescence intensity of Hf-UiO-66-NHCSNHCH3 became saturated after 50 s (Fig. 3c and d). Therefore, the probe Hf-UiO-66-NHCSNHCH3 is highly sensitive with a very low detection time as compared to other MOF-based mercury and hydrazine sensors displayed in Tables S5 and S6.† The fluorescence increment and quenching efficiency after adding 300 μL hydrazine and Hg2+ solution was calculated using the formulas I/I0 and (I − I0/I) × 100, respectively (I0 is the fluorescence intensity of the probe before the addition of the target analyte and I is the fluorescence intensity of the probe after the addition of the target analyte). We found a 28-fold increment in the fluorescence intensity of the probe after the addition of 300 μL 10 mM hydrazine solution. After the addition of 300 μL 10 mM Hg2+ solution, the fluorescence intensity of the probe quenched up to 92%. The appreciable change in the original fluorescence intensity of the probe by both the targeted analytes made our material a different and more efficient sensor.
The selectivity of the probe over the other competitive analytes was thoroughly investigated. To a MOF suspension of 3000 μL in DMF, 300 μL 10 mM of aqueous competitive analyte solutions (Ag+, K+, Li+, Cd2+, Zn2+, Cu2+, Mg 2+, Mn2+, Pb2+, Ni2+, Na+, Pt2+, Pd2+, Al3+, Cr3+, and Co3+) were added in individual experiments, and the fluorescence intensity was recorded (Fig. 4a). We found a quenching efficiency of 92% in the case of Hg2+, whereas for other analytes, the quenching efficiency only remained below 25% (Fig. S18–S33† and 4a). A similar experiment was performed using an aqueous MOF suspension and competitive analytes (ala, gly, leu, tyr, pro, urea, thiourea, phe, Br−, F−, CH3COO−, HSO4−, NCS−, HCO3−, S2O32−, and NO2−) of hydrazine (Fig. S34–S49†). We found an immediate 28-fold increment in the fluorescence intensity when hydrazine was added to an aqueous MOF suspension. However, a minor enhancement in fluorescence intensity was noticed for other competitive analytes of hydrazine. The above experiments confirmed that the probe Hf-UiO-66-NHCSNHCH3 is highly selective over other competitive analytes for the selective detection of Hg2+ and hydrazine, as shown in Fig. 4b. An ideal probe should not only detect the target analyte selectively, but the detection process should also be repeatable and with a minimum allowed standard error. Therefore, the selectivity experiment was carried out three times, and the error in the 2D-bar plot is presented as the standard deviation in Fig. 4a and b. The minor standard deviations confirmed that our probe is reproducible in providing the results repeatedly with high precesion and accuracy.
Fig. 4 Comparative selectivity bar plots of the probe Hf-UiO-66-NHCSNHCH3 toward Hg2+ (a) and NH2–NH2 (b) with their respective congeners (plots are shown with standard deviations of 3 measurements). |
We evaluated the selectivity of our probe to hydrazine and Hg2+ in a complex medium in the presence of other competitive analytes. A three-step procedure was carried out for the selectivity experiments. In the first step, the fluorescence emission intensity of the MOF suspension was recorded. In the second step, 300 μL 10 mM aqueous solution of a competitive analyte was added and the fluorescence spectrum was recorded. In the last step, 300 μL 10 mM aqueous Hg2+ solution was added and the fluorescence emission intensity was recorded. The same experiment was repeated for all the competitive analytes of Hg2+. Similar experimental procedures were adopted for all the competitive analytes of hydrazine as well. The obtained results showed that probe Hf-UiO-66-NHCSNHCH3 can detect both Hg2+ and hydrazine selectively in the presence of respective competitors. As displayed in Fig. 5a and b, there is no such competitor analyte to question the selectivity of the probe Hf-UiO-66-NHCSNHCH3 for detecting both the targeted analytes (Hg2+ and NH2–NH2).
The low value of limit of detection (LOD) is one of the essential properties of an ideal sensor. Therefore, we performed the fluorescence sensing experiment of Hf-UiO-66-NHCSNHCH3 of our probe toward detecting mercury and hydrazine in the concentration of analytes with as low concentration as possible. After a systematic investigation, we calculated the LOD values using the formula 3σ/k. Here, σ is the standard deviation of the blank fluorescence intensities of the MOF suspension only and k is the slope of the curve (linear fit curve between fluorescence intensity and concentration) (Fig. S50 to S51†). The LOD values of the probe Hf-UiO-66-NHCSNHCH3 were 1.9 ± 0.25 nM and 4.0 ± 0.37 nM for hydrazine and Hg2+ detection, respectively. The obtained LOD values are lower in comparison with any MOF-based sensors of mercury and hydrazine to date (Tables S5 and S6†). The selectivity, sensitivity, and lower LOD value of our probe Hf-UiO-66-NHCSNHCH3 toward detecting mercury and hydrazine make it an ideal sensor.
To understand the quenching process of the probe by Hg2+, we plotted the Stern–Volmer plot. The Stern–Volmer plot in Fig. S66† displays that at lower concentration, the plot is linear, but at higher concentration, it became steeper. The above observation concluded that the quenching process might be due to the static or dynamic pathway.28 Therefore, we carried out the time-resolved fluorescence lifetime experiment. Again, the Stern–Volmer constant (Ksv) obtained from the slope of the plot between I0/I versus the concentration of mercury confirmed that the quenching process must be due to strong interaction between the probe and mercury as the Ksv value is much higher (7.49 × 105 M−1) (Table S1†). There was almost no change in the probe's lifetime before and after adding Hg2+. The minor lifetime change confirmed that the static pathway of fluorescence quenching occurred due to the formation of the ground state complex between the probe and Hg2+ (Fig. S52 and Table S4†). The 3D-Stern–Volmer plot, also presented in Fig. 6, shows that the material is highly selective toward mercury over other analytes in a wide concentration range.
The recyclability of the probe for the sensing of mercury was examined by washing it with DMF and water after each cycle of the sensing process. We performed the recyclability experiment up to seven cycles, which showed that the probe is equally efficient in detecting mercury up to seven cycles, as shown in Fig. S53.†
The sensing of hydrazine in different water specimens was studied in a systematic way. An MOF suspension was prepared in each water specimen (river water, seawater, lake water, tap water, and distilled water). A series of aqueous hydrazine solutions of 3.33 mM, 6.66 mM, and 10 mM concentrations were prepared and utilized for sensing experiments. The fluorescence experiment was carried out in a general manner by recording the fluorescence emission intensity before and after the addition of different concentrations of hydrazine. The obtained results shown in Fig. 7a proved that our probe has efficiency in detecting hydrazine even from a complex water specimen system. These results indicate that the probe can work for real-field application purposes.
Fig. 7 Detection of hydrazine in environmental water specimens at different concentrations (a) and detection of hydrazine in various pH solutions (b). |
Again, the detection of hydrazine in different pH solutions was examined. The obtained results in Fig. 7b indicates that our probe could detect hydrazine in a wide pH range (pH 4 to pH 12). At pH 2, the presence of an acidic medium immediately protonated the added hydrazine and hindered its attack on the thiocarbonyl group. Therefore, a negligible turn-on in the fluorescence emission intensity was noticed after the addition of hydrazine to the MOF suspension.
A naked eye detection method was adopted by adding hydrazine and mercury to the cuvettes containing MOF suspension under a fluorescence lamp. A turn-on blue fluorescence light was observed after the addition of hydrazine, whereas after the addition of mercury, a turn-off in fluorescence was observed under the fluorescence lamp. A cotton composite-based device was also designed by coating MOF powder onto the surface of a cotton cloth. For preparing the MOF-cotton composite, at first, cotton pieces were washed with ethanol and acetone and dried. The clean cotton pieces were immersed in MOF suspension in ethanol. The cotton pieces were stirred in MOF suspension slowly for 24 h. Then, the pieces were removed and dried in an oven. The dried MOF-cotton composite was utilized further for real-field sensing purposes. Cotton possesses free hydroxy groups on its surface, and the Hf-atom of Hf-UiO-66-NHCSNHCH3 is oxophilic in nature. Therefore, the MOF material was bound to the hydroxy groups on the surface of cotton to produce a useful cotton-composite material, which has great utilization for real-field sensing application purposes. After adding mercury and hydrazine to the MOF-coated cotton composite, similar turn-off and turn-on phenomena in fluorescence intensity were observed (Fig. 8). The simple cotton composite-based naked-eye sensing device made of our material applicable in real-world sensing applications.
Fig. 8 Naked eye detection of Hg2+ and NH2–NH2 by MOF-coated cotton composite under a fluorescence lamp. |
To know the exact active center of the reaction, we carried out sensing experiments with Hf-UiO-66-NH2 MOF. The obtained result showed that after adding hydrazine solution to Hf-UiO-66-NH2 suspension, there was little change in the fluorescence emission intensity. The above result indicated that the reaction between the thiocarbonyl group of the probe and hydrazine is the reason behind the fluorescence turn-on behavior. Again, we repeated the fluorescence sensing experiment utilizing a Hf-BDC-NHCON(CH3)2 MOF in place of our probe Hf-UiO-66-NHCSNHCH3. In this case, a similar fluorescence turn-on behavior such as probe Hf-UiO-66-NHCSNHCH3 was observed (Fig. 9). The probe Hf-BDC-NHCON(CH3)2 is not selective for other competitive analytes such as NCS−, HSO4−, and S2O32−. Hf-BDC-NHCON(CH3)2 showed a similar turn-on behaviour similar to NH2–NH2. The higher electrophilicity of Hf-BDC-NHCON(CH3)2 is the cause of the absence of selectivity, whereas with probe Hf-UiO-66-NHCSNHCH3, the moderate electrophilicity of the thiocarbonyl group helps to detect hydrazine without any selectivity issue. The α-effect of hydrazine causes the nucleophilic attack on the thiocarbonyl group. It breaks the conjugation from the aromatic part of the linker, resulting in an enhancement in the fluorescence emission intensity. Further proof of reaction-based nucleophilic attack on the thiocarbonyl functionality of the linker was supported by the 1H NMR and 13C NMR spectra. The aromatic peaks in the 1H NMR spectrum of hydrazine-treated Hf-UiO-66-NHCSNHCH3 displayed an up-field chemical shift due to an increase in the electron density on the benzene ring (Fig. S55†). In the 13C NMR spectrum, the disappearance of the peak at 175 ppm and the appearance of the 149 ppm peak is a strong evidence of the nucleophilic attack by NH2–NH2 (Fig. S56†). The mass spectrum of the probe after the addition of hydrazine showed prominent peaks at 237 and 251 m/z ratios (Fig. S57†), which is due to the formation of the new imine complex (mass 251) by the attack of hydrazine on the thiocarbonyl group. Later, the imine complex dissociated by the attack of water present in the system to give a carbonyl compound (mass 237), as shown below in Scheme 2. Again, the UV-Vis spectra of probe Hf-UiO-66-NHCSNHCH3 before and after the addition of NH2–NH2 was examined. We noticed a considerable change in the absorbance spectrum after adding NH2–NH2. It also supports that the enhancement in the fluorescence emission intensity of the probe Hf-UiO-66-NHCSNHCH3 after adding NH2–NH2 might be due to a reaction-based pathway (Fig. S58†).
The EDX elemental analysis of the recovered probe after hydrazine sensing confirmed the absence of sulfur atoms, indicating a nucleophilic attack by hydrazine on the thiocarbonyl group of the probe, thereby causing loss of sulfur in the form of hydrogen sulfide (H2S) (Fig. S59†). To confirm the removal of sulfur in the form of H2S, we used a previously reported azide-based H2S sensor Nap-but for H2S detection.29 The H2S sensor was spiked with our MOF suspension, and the fluorescence spectrum was recorded. The same was also utilized by Bhabak et al. to detect the H2S from their reaction medium.30 After the addition of hydrazine to the above mixture, we noticed a prominent enhancement in the fluorescence intensity, which confirmed the release of H2S from our probe due to the attack of hydrazine on the thiocarbonyl part of our probe (Fig. S60†). When we added hydrazine to a mixture of our probe and lead acetate in water, we noticed immediate black precipitation due to the formation of lead sulfide (Fig. S61†).31 The source of sulfur must be due to the detachment of sulfur from our probe. As we know, lead acetate is very reactive toward H2S to form lead sulfate. The color of lead sulfate is black, and we noticed a black sediment in the mixed suspension after adding hydrazine. All the above experimental data strongly supported the nucleophilic attack of hydrazine to the thiocarbonyl center of our probe. It caused an enhancement in the fluorescence emission intensity after adding hydrazine to the probe.
The fluorescence lifetime of the probe remained almost the same before and after adding mercury solution to the MOF suspension, as mentioned above. Therefore, the quenching process was not due to a dynamic pathway which kicked out the possibility of fluorescence resonance energy transfer mechanism. The above result confirmed that the fluorescence quenching process occurred through a static quenching pathway, i.e., due to the ground state complexation between the probe and mercury, which allows the electron transfer from the probe to the vacant orbital of mercury. As reported by many previous works, the sulfur atom of the thiocarbonyl group is the center for the interaction with mercury, and in our case, the sulfur of the probe may be the center for soft–soft interaction with mercury.32 Therefore, we performed X-ray photoelectron spectroscopy to find the exact active center of complexation. We executed the XPS analysis of Hf-UiO-66-NHCSNHCH3 before and after interaction with mercury. The binding energies of the 2S orbital of sulfur were 161.96 and 163.09 eV before the interaction with mercury. They became 162.46 and 163.58 eV after the interaction with mercury, providing a direct support for the interaction of mercury with sulfur. The mercury–sulfur interaction allows the transfer of electrons from the probe to mercury and thereby causes quenching in the fluorescence intensity of the probe (Fig. S66†). The shift in the binding energy of all the elements after interaction with Hg2+ also supported that there must be some interaction between the probe and Hg2+ (Fig. S67–S70†). The presence of the XPS characteristic peak of mercury is also direct proof of the interaction between the sulfur atom of the probe and Hg2+ (Fig. S71†).33 The solid state UV-Vis spectrum of our material before and after the addition of Hg2+ showed a 24 nm red shift, which is a strong evidence of ground state complexation between our probe and Hg2+ (Fig. S72†).
Further confirmation of molecular interaction was obtained from a systematic temperature-dependant fluorescence experiment of our probe in the presence of different concentrations of mercury in varying temperature range (298–343 K). The Ksv value was obtained at different temperatures from the linear fit plot of I/I0versus the concentration of mercury (Fig. S73†). There was a decrease in the Ksv value with an increase in temperature, as shown in Table 1. The reduction in the Ksv value with increase in temperature rise supports the soft-soft interaction between mercury and the sulfur atom of our probe.
T (K) | K sv × 105 (L mol−1) | K a × 105 (L mol−1) | ΔG (kJ mol−1) | ΔH (kJ mol−1) | ΔS (J mol−1 K−1) |
---|---|---|---|---|---|
298 | 7.49 | 1.29 | −28.89 | −37.24 | −28.02 |
313 | 6.23 | 0.49 | −28.47 | ||
328 | 5.99 | 0.26 | −28.05 | ||
343 | 5.51 | 0.19 | −27.63 |
The bimolecular binding constant (Ka) at different temperatures was obtained from the intercept of the plot of log[(I0 − I)/I] versus log[Q] using the modified Stern–Volmer equation (Fig. S74†): log[(I0 − I)/I] = logKa + nlog[Q]. The decrease in the magnitude of the binding constant again supported the fact that the quenching process was due to the soft-soft interaction between mercury and the sulfur atom of our probe (Table 1). By increasing the temperature of the system, the randomness of the system increases and the interaction becomes weak, because of which the value of the binding constant decreased with the increase in the temperature.34,35
The thermodynamic parameters, i.e., enthalpy change (ΔH) and entropy change (ΔS), were obtained from the slope and intercept of the logKaversus 1/T plot, respectively (Fig. S75†). As we know, the difference in enthalpy and entropy could help in predicting the interaction process. The positive value of ΔH and ΔS suggests hydrophobic interaction, whereas the negative value of ΔH and positive value of ΔS suggests electrostatic interaction. In our case, both the thermodynamic parameters are negative, indicating molecular interaction.
lnKa = −ΔH/RT + ΔS/R |
Further, we calculated the free energy change (ΔG) at different temperatures using the obtained value of ΔH and ΔS and the following thermodynamic formula. The obtained values of ΔG at various temperatures were negative, which confirmed that the interaction process was spontaneous and exothermic.36 The decrease in the magnitude of ΔG with an increase in temperature also agrees with the Ka values obtained at various temperatures.
ΔG = ΔH TΔS |
Further, we carried out control fluorescence experiments using Hf-UiO-66-NH2 and Hf-UiO-66-NHCON(CH3)2 MOF materials in place of our probe. A negligible quenching in fluorescence emission intensity was observed after adding mercury solution to Hf-UiO-66-NH2 and Hf-UiO-66-NHCON(CH3)2 MOF suspensions (Fig. 9). The experiments again confirmed that the interaction between the sulfur atom of the probe and mercury is the reason behind the fluorescence quenching process.
Footnote |
† Electronic supplementary information (ESI) available: E-SEM images, EDX, NMR, IR, TRPL, XPS and fluorescence spectra, N2 sorption isotherm, XRPD patterns, digital images, TG curves, comparison tables. See DOI: https://doi.org/10.1039/d3qi00206c |
This journal is © the Partner Organisations 2023 |