Next Article in Journal
Comparative Chemical Profiling and Citronellol Enantiomers Distribution of Industrial-Type Rose Oils Produced in China
Next Article in Special Issue
Development of a Ratiometric Fluorescent Cu(II) Indicator Based on Poly(N-isopropylacrylamide) Thermal Phase Transition and an Aminopyridyl Cu(II) Ligand
Previous Article in Journal
Analysis of Single-Step Pretreatments for Lignocellulosic Platform Isolation as the Basis of Biorefinery Design
Previous Article in Special Issue
A Ratiometric Fluorescent Sensor Based on Dye/Tb (III) Functionalized UiO-66 for Highly Sensitive Detection of TDGA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Naphthalimide-Piperazine Derivatives as Multifunctional “On” and “Off” Fluorescent Switches for pH, Hg2+ and Cu2+ Ions

by
Kristina Pršir
1,
Mislav Matić
1,
Marlena Grbić
1,
Gerhard J. Mohr
2,
Svjetlana Krištafor
1 and
Ivana Murković Steinberg
1,*
1
Department of General and Inorganic Chemistry, Faculty of Chemical Engineering and Technology, University of Zagreb, Marulićev trg 19, 10000 Zagreb, Croatia
2
Joanneum Research Forschungsgesellschaft mbH—Materials, Franz-Pichler-Straße 30, A-8160 Weiz, Austria
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(3), 1275; https://doi.org/10.3390/molecules28031275
Submission received: 7 December 2022 / Revised: 13 January 2023 / Accepted: 24 January 2023 / Published: 28 January 2023
(This article belongs to the Special Issue Fluorescence Chemosensors: Design, Synthesis, and Application)

Abstract

:
Novel 1,8-naphthalimide-based fluorescent probes NI-1 and NI-2 were designed and screened for use as chemosensors for detection of heavy metal ions. Two moieties, methylpyridine (NI-1) and hydroxyphenyl (NI-2), were attached via piperazine at the C-4 position of the napthalimide core resulting in a notable effect on their spectroscopic properties. NI-1 and NI-2 are pH sensitive and show an increase in fluorescence intensity at around 525 nm (switch “on”) in the acidic environment, with pKa values at 4.98 and 2.91, respectively. Amongst heavy metal ions only Cu2+ and Hg2+ had a significant effect on the spectroscopic properties. The fluorescence of NI-1 is quenched in the presence of either Cu2+ or Hg2+ which is attributed to the formation of 1:1 metal-ligand complexes with binding constants of 3.6 × 105 and 3.9 × 104, respectively. The NI-1 chemosensor can be used for the quantification of Cu2+ ions in sub-micromolar quantities, with a linear range from 250 nM to 4.0 μM and a detection limit of 1.5 × 10−8 M. The linear range for the determination of Hg2+ is from 2 μM to 10 μM, with a detection limit of 8.8 × 10−8 M. Conversely, NI-2 behaves like a typical photoinduced electron transfer (PET) sensor for Hg2+ ions. Here, the formation of a complex with Hg2+ (binding constant 8.3 × 103) turns the green fluorescence of NI-2 into the “on” state. NI-2 showed remarkable selectivity towards Hg2+ ions, allowing for determination of Hg2+ concentration over a linear range of 1.3 μM to 25 μM and a limit of detection of 4.1 × 10−7 M.

1. Introduction

Due to the well-known impact of heavy metals in the environment and on human health, the development of new, cost-effective, fast and reliable methods for their detection has become urgent [1,2]. Heavy metal detection methods have variously been proposed based on sensing mechanisms that include electrochemical and optical techniques with specifically designed recognition elements, such as molecularly imprinted polymers [3], nanomaterials [4], metal-organic frameworks [5] and N-heterocyclic receptors [6].
Fluorescence-based optical sensors and molecular probes in particular have attracted interest as metal ion sensors due to favorable properties such as high sensitivity and selectivity, the relatively simple detection process and intrinsic low cost. Various fluorescent chemosensors have been synthesized and investigated for heavy metal ion-determination based on photophysical mechanisms such as chelation-induced enhanced fluorescence (CHEF), chelation enhancement quenching (CHEQ), intramolecular charge transfer (ICT), photo-induced electron transfer (PET) and aggregation-induced emission (AIE) [7,8,9,10,11,12,13,14,15,16].
Photo-induced electron transfer sensors are particularly popular in molecular fluorescent “switch” designs. In PET systems, a fluorophore is covalently linked with a receptor through a short aliphatic spacer, while both segments maintain their individual properties. Upon interaction with the target analyte, the PET process is inhibited thereby causing changes in the quantum yield or fluorescence intensity [17,18,19].
The potent toxicity of mercuric ions to living systems has inspired the development of new fluorescent chemosensing molecules and sensing schemes for detection of Hg2+ in environmental, industrial and biological samples [2,20,21,22]. Chemosensors for Cu2+, a metal essential for life but also highly toxic to some organisms, have also been the subject of much research and development effort [22,23,24,25]. It is known that Cu2+ and Hg2+ act as quenching agents to many fluorescent probes. Cu2+ has intrinsic paramagnetic properties, which cause an energy or electron transfer, consequently decreasing the fluorescence intensity of the fluorophore [26]. Being a heavy atom, Hg2+ decreases the quantum yield of fluorescence by an external spin–orbital coupling effect, which induces a non-radiative intersystem crossing [27]. There is a great interest in developing probes for Hg2+ based on intensity enhancement, rather than quenching, since the latter often show lower sensitivity [28].
One notable group of fluorescent probes are the derivatives of 1,8-naphthalimide. Known for their chemical stability, these molecules exhibit excellent thermal, electrochemical, electroluminescent, optical and photophysical properties [29,30,31]. By introducing different functional moieties on the imide and amino nitrogen atom, it is possible to modify the physical and chemical properties of the naphthalimide. Such modifications then open the possibility of using the resulting molecules as chemical and biochemical sensors. In general, the molecular fragment bound to the 4-amino nitrogen forms the receptor part of the molecule that interacts with the analyte, while the fragment attached to the imide nitrogen atom affects the physical properties. 1,8-naphthalimide is a photostable conjugate system consisting of two aromatic rings and two double bonds with oxygen. This allows the π–π* electron transfer due to light absorption in the ultraviolet (UV) region [17]. In addition, naphthalimide derivatives exhibit large Stokes shifts. Thanks to these favorable photophysical properties and the relatively simple synthesis, naphthalimides have been used for both cell imaging and for measuring extracellular and intracellular pH. For example, naphthalimide-based pH-probes incorporating a protonable piperazine ring [32] and a hydroxyl group linked to the naphthalimide moiety at the C-4 position [33] have recently been reported. These demonstrate some of the many practical applications of naphthalimide derivatives as pH sensors in living systems for physiology and cell biology research. Naphthalimide probes and sensors for alkali [34], earth alkali [35] and heavy metal ions [36,37,38,39,40] have also been devised and investigated. Naphthalimides may also be used as components of functional materials [41] and as components of drug delivery systems [42], and are particularly suitable here due to the ease of structural modification by modulation of moieties attached to the amino nitrogen atom [43].
In this paper, we investigate the applicability of two compounds, NI-1 and NI-2, as optical probes for pH and heavy metal ion detection (Figure 1).
The NI-1 and NI-2 compounds each have a typical donor–spacer–acceptor structure (D-π-A), comprising a 1,8-naphthalimide core, which represents the fluorophore, and either a pyridin-2-ylmethyl or hydroxyphenyl moieties attached to the core through piperazine. The moieties act as additional receptor sites for metal ion binding. Due to the presence of the N-substituent containing a 2-hydroxyethylsulfonyl group, the covalent immobilization of these naphthalimide probes to different substrate materials and to nanoparticles (e.g., cellulose, textiles) is also possible using well-established vinyl sulfonyl immobilization chemistry. This molecular feature opens up a wide range of solid state applications for these probes in the future [44].
Structures of studied naphthalimide derivatives NI-1 and NI-2 are shown in Figure 1. The main structural difference between them is a molecular fragment attached via piperazine moiety at 2-hydroxyethylsulfonyl-functionalised naphthalimide.

2. Results and Discussion

2.1. Spectral Characteristics of NI-1 and NI-2 in Methanol

Structures of studied naphthalimide derivatives NI-1 and NI-2 are shown in Figure 1. The main structural difference between them is a molecular fragment attached via piperazine moiety at 2-hydroxyethylsulfonyl-functionalised naphthalimide.
Absorption and emission spectra of compounds are shown in Figure 2a.
Both compounds displayed absorption spectra with the absorption maximum at 410 nm with molar absorption coefficients, ε = 2.41 × 104 M−1 cm–1 for NI-1 and ε = 8.52 × 103 M−1 cm−1 for NI-2, which can be assigned to π–π* transitions. To the naked eye, NI-1 and NI-2 displayed a very light yellow color (Figure 2b). Considering the fluorescence emission spectra, compounds NI-1 and NI-2 exhibited a single emission band with emission maxima at 523 and 525 nm, respectively, but the intensity was strongly dependent on the attached moiety. NI-1-bearing pyridine functionality is strongly green fluorescence emitting fluorophore, compared to hydroxyphenyl-substituted NI-2. Since the lone electron pair on the oxygen atom is a part of the conjugated system, overall electron density of the hydroxyphenyl moiety is increased, which might repulse the lone electron pair on the piperazine nitrogen. That in turn induces the electron transfer from the piperazine nitrogen to the fluorophore, which lowers the emission rate. On the other hand, the lone electron pair on the pyridine nitrogen is not involved in the conjugated system, making the ring overall electron-deficient. Compounds NI-1 and NI-2 exhibited large Stokes shifts in neutral pH (5558 cm–1 and 5629 cm−1) and therefore have great potential for the application as molecular probes, despite the observed low fluorescence quantum yield (Φ(NI-1) = 0.027, Φ(NI-2) = 0.004) [45].

2.2. Effect of pH on Spectral Properties

The spectral properties of NI-1 and NI-2 as a function of pH were investigated using both absorption and fluorescence emission spectroscopy. The potential use of NI-1 and NI-2 as pH-sensing probes is based on the spectral changes caused by the protonation of the piperazine nitrogen atom. Optical pH sensors are characterized by their apparent pKa values in buffer methanol mixtures (2:1), determined by spectrophotometry and spectrofluorimetry. The pH values correspond to the pH values of the aqueous buffers used.

2.2.1. UV–Visible Absorption Spectra

Changes in the pH of solutions caused minor changes in the absorbance spectra of the studied compounds (Figure S1 (Supplementary Materials)). As the pH decreases from 7.0 to 1.6 the absorption band of NI-1 gradually shifts from 410 nm to 388 nm, generally followed by an increase in absorbance of the maximum. The absorbance ratio at peak values (for protonated and neutral forms) has been used for calculating the apparent pKa value (Figure S1a,b). NI-2 showed the same behavior upon the pH changes, with hypsochromic shift of absorption maxima from 410 nm to 389 nm upon the protonation along with the increase in the peak intensity, as well as a gradual hyperchromic shift in the pH range 5.0–8.0 (Figure S1c).

2.2.2. Fluorescence Emission Spectra

The emission spectral responses to pH for NI-1 and NI-2 and the dependence of fluorescence intensity (at 523 nm for NI-1 and 525 nm for NI-2) upon the pH value are shown in Figure 3 and Figure S2. The fluorescence intensity changes of NI-1 and NI-2 upon pH were described well by a sigmoidal function (Boltzmann fitting), Figure 3b.
The effect of pH is more pronounced in the fluorescence spectra than in the absorption spectra of compounds. It is known that the protonation of the nitrogen atom of piperazine stops the PET process from amine donor to the naphthalimide core and therefore, the fluorescence intensity increases gradually with the decrease in pH value [45,46,47]. The fluorescence intensity of both NI-1 and NI-2 were considerably increased upon the protonation, up to 29.6-fold for NI-1 and 162.5-fold for NI-2. Fluorescence turn-ON in acidic solution supports the inhibition of the PET process in the piperazine–naphthalimide derivative upon the protonation of piperazine, as shown in Figure 4.
The pH response curves exhibit excellent adjustment and apparent pKa values were calculated to be 4.98 for NI-1 (R2 = 0.9923) and 2.91 for NI-2 (R2 = 0.9982).
NI-1 and NI-2 have different substituents on their pH-sensitive amino group. The substituents did not affect the pH sensitivities, as both compounds displayed fluorescence enhancement upon the protonation. The different nature of substituent moieties attached to the protonable piperazine moiety (alkyl or aryl) significantly influenced the photophysical properties of derivatives, showing that the suitable substituents play a crucial role for tuning the pKa and pH sensitivity. Moreover, NI-1 and NI-2 dyes were immobilized onto cellulose layers and the pKa values of were determined to be 6.05 and 2.98, respectively [43]. These results indicate that both the nature of substituents and the immobilization technique have a significant impact on spectroscopic and pH sensing properties and could provide a different application of fluorescent naphthalimide dyes [48,49,50].
The properties of NI-1 and NI-2 in acidic (pH 1.6) and neutral solution (pH 7.0), including absorption and emission wavelength (λabs, λemiss), molar absorption coefficient (ε), fluorescence intensity, Stokes shift and calculated apparent pKa values are summarized in Table 1.

2.3. Metal Ion Sensing

The difference in molecular fragments attached via piperazine moiety at 2-hydroxyethylsulfonyl-functionalised naphthalimide, affected not only the emission spectra of NI-1 and NI-2, but also their response upon binding with metal cations. Spectroscopic measurements of metal ion sensing were performed in MeOH/buffer (v/v, 1/2, pH 5.5). For the spectroscopic measurements, compound concentration and experimental conditions were adjusted for these two compounds separately, because of their different behavior in solution and response in the presence of metal cations. A pH value of 5.5 for the metal ion studies was chosen as a compromise to maximize sensitivity and to minimize effects of competing equilibrium reactions, i.e., protonation of NI at lower pH values and the metal ion hydrolysis at higher pH values.

2.3.1. NI-1

  • Visual detection of metal ion sensitivity
To evaluate the metal ion sensitivity, preliminary tests were performed under a UV lamp (365 nm) by adding 30.0 μL metal ion stock solution to the compound in MeOH/buffer (v/v, 1/2, pH = 5.5) (c (NI-1) = 6.56 × 10−6 M) and the results are presented in Figure S3.
NI-1 exhibits bright green fluorescence in MeOH/buffer solution and with the addition of Cu2+, fluorescence quenching can be seen by naked eye. Addition of other metal ions did not cause any observable changes in the fluorescence of NI-1;
  • Fluorescence titration experiments
The experiments were performed at compound concentration 5.00 × 10−7 M in MeOH/buffer (1/2, v/v, pH = 5.5) in the presence of 2, 20 and 95 equiv. of metal ions and the results are summarized in Figure 5. A small addition of metal ions (2 equiv.) did not significantly affect the fluorescence intensity of NI-1, while the presence of 20 equiv. of Cu2+ and Hg2+ resulted in fluorescence quenching of NI-1 intensity for 46.4% and 22.7%, respectively. Furthermore, 95 equiv. of these two cations induced further quenching of fluorescence intensity of NI-1 for 60.0% upon Cu2+ and 39.8% upon Hg2+ addition). Moreover, Al3+, Zn2+, Ba2+, Pb2+ and Cd2+ ions present in high concentrations (95 equiv.) induced NI-1 fluorescence intensity increase in the range of 31.1 to 36.0% for Zn2+, Ba2+, Pb2+ and Cd2+ and 47.9% increase in the presence of Al3+, as summarized in Figure 5.
Encouraged by these findings, we performed titration experiments with Cu2+ and Hg2+ and the results are presented in Figure 6 and Figure 7.
It can be seen from the emission spectra that an increase in the concentration of metal ions leads to a decrease in the fluorescence intensity, and a significant decrease in intensity is achieved at higher concentrations. In the case of Cu2+ ions, a hypochromic shift occurs, with the maximum of fluorescence intensity at 523 nm, while in the case of Hg2+ ions, a bathochromic shift occurs with a decrease in intensity. The wavelength of the fluorescence intensity maximum in the presence of Hg2+ ions shifts from 523 nm to 541 nm. Hypochromic shift is common in naphthalimide derivatives whose fluorescence is quenched in presence of Cu2+ ions [39]. However, bathochromic shift in the presence of Hg2+ ions is not common and the possibility of obtaining two different fluorescent responses can potentially be used to discriminate Cu2+ from Hg2+ using NI-1.
Fluorescence properties of immobilized napthalimide derivatives with (methylpyridine)piperazine receptor in presence of Hg2+ and Cu2+ ions have been reported previously. Covalently immobilized probes on a glass carrier were examined at pH 7 in the presence of Hg2+ ions and have shown an increase in fluorescence with Hg2+ ions and a slight decrease with Cu2+ ions [37]. Additionally, Liu et al. [51] synthesized two naphthalimide derivatives, also with (methylpyridine)piperazine as receptor of NI-1, and immobilized them on (poly)styrene substrate. The resulting optodes differ only in the linker (N-substituent of naphthalimide) through which the compounds were attached to the carrier. In both cases, Hg2+ ions caused a significant increase in fluorescence intensity. It is interesting to notice that, although the receptor part of all four compounds is the same, their response to Hg2+ and Cu2+ ions differs significantly. This implies that other factors, such as immobilization method, material of the substrate, length of the linker part on the N-substituent, test conditions—particularly the pH of the solution, solvents and anion composition, etc., have a significant effect on the mechanism of the fluorescence response, possibly causing steric hindrance and influencing binding stoichiometry as well.
Unlike fluorescence emission spectra, UV-Vis absorption spectra of NI-1 were not significantly affected by presence of Hg2+ and Cu2+ ions (Figure S4), showing an absorption maximum at 410 nm which can be assigned to the naphthalimide core;
  • Reversibility tests
Reversibility tests were performed with the addition of EDTA to the solution containing NI-1 and metal cations, in five cycles, as shown in Figures S5 and S6 for Cu2+ and Hg2+, respectively. In general, upon addition of EDTA, the fluorescence of NI-1 recovers. In the case of Cu2+, it was noticed that the recovered fluorescence intensity of NI-1 increases by approx. 10% upon each addition of EDTA in repeating cycles. This behavior could be attributed to the enhancement of fluorescence of NI-1 caused by other effects, possibly related to the gradually changing composition of the solution with each cycle. In comparison to free NI-1, the solutions containing NI-1, EDTA and excess Hg2+ ion show fluorescence intensity quenching, varying by approximately 8%, while the maximum shifts from 523 nm to 536 nm (Δλmax = 13 nm);
  • Metal-binding studies
The stoichiometric coefficients of Cu2+ and Hg2+ ions upon binding with NI-1 were assumed using Job’s plot as shown in Figure 8 and a 1:1 binding ratio is suggested for both ions, which is in accordance with previously published results for the same substituent [36,37].
The equilibrium association (binding) constant for Cu2+ and Hg2+ ions and NI-1 binding was determined by Benesi–Hildebrand’s plot using the equation:
1 F F 0 = 1 K b F 0 C u 2 + n + 1 F m a x F 0
where F0 is the fluorescence intensity in the absence of metal ions; Fmax—the fluorescence intensity in presence of saturated metal; F—the intensity of the metal-probe complex in the linear range and n—factor of stoichiometric binding (1 for 1:1 ratio). The correlation factor for both ions is R2 > 0.96 with the binding constant being 3.9 × 104 for Hg2+ and 3.6 × 105 for Cu2+ (Figure S7);
  • LOD (limit of detection)
The limit of detection (LOD) of NI-1 for Cu2+ and Hg2+ ions was determined using the IUPAC-recommended equation:
LOD = 3σ/k
where σsp is the standard deviation of blank measurements and k is the slope between the fluorescence intensity and metal concentration in linear range as shown in Figure S8. The standard deviation for four blank measurements is σsp = 0.3295. The detection limit of NI-1 is 1.5 × 10−8 M and 8.8 × 10−8 M for Cu2+ and Hg2+ ion, respectively. These values are comparable to similar fluorescent quenching chemosensors utilizing the naphthalimide structure with aliphatic amino receptors, reported previously [52,53,54,55]. Wu et al. [37] determined the limit of detection of Hg2+ ions by naphthalimide-based optode to be 2.00 × 10−6 M at pH 7.01 in 0.05 M Tris-HCl which is significantly higher than the detection limit of the compound NI-1. Furthermore, Gao et al. determined the detection limit of Cu2+ ions for three naphthalimide derivatives, which have the same receptor moiety as NI-1, in which fluorescence quenching occurs. The determined detection limits reported are in range from 2.62 × 10−9 to 1.17 × 10−8 M [39]. The Stern–Volmer constants for NI-1 were calculated to be 152570 M−1 for Cu2+, and 23909 M−1 for Hg2+, as determined from the respective Stern–Volmer plots (Figure S9). Based on this evidence from the literature and the known behavior of Cu(II) as a strong paramagnetic quencher, we assume that the quenching here is also induced by chelation, rather than being collisional.
Table 2 shows a comparison between different copper sensors reported in the literature and the napthalimide sensor NI-1 presented in this work.

2.3.2. NI-2

  • Visual detection of metal ion sensitivity
NI-2 did not fluoresce in the MeOH/buffer mixture at pH 5.5. Upon addition of metal cations, a drastic increase in the fluorescence intensity was observed only in the presence of Hg2+ ions (Figure 9). These initial experiments suggested that NI-2 could be used as selective fluorimetric naked-eye chemosensor for Hg2+;
  • Titration experiments
To investigate the possibility of an NI-2 compound to detect Hg2+ ions, fluorescence titration was conducted. The fluorescence spectra were recorded at excitation wavelength of 410 nm. As shown in Figure 10, the maximum fluorescence band is centered at 525 nm. The spectrum of the NI-2 compound without Hg2+ exhibits a low intensity peak around 533 nm, typical for the naphthalimide derivatives containing an amine substituent at the C-4 position of the naphthalene imide ring [61]. The addition of Hg2+ to the compound NI-2 results in a remarkable enhancement of fluorescence signal but has nearly no effect on its UV-Vis spectrum (Figure S10). The fluorescence intensity increases steadily in response to the gradual increase in concentration of added Hg2+ ion in MeOH/buffer (v/v, 1/2, pH 5.5) at a compound concentration 6.53 × 10–6 M, while a hypsochromic shift ΔλF = 6 nm was observed. Such a small shift suggests that upon addition of the metal ions the fluorophore ICT state is not affected and that the complexation takes place via a nitrogen atom on the piperazinyl part and hydroxyphenyl group of the substituent simultaneously, rather than the imide moiety. These results support the potential of the NI-2 compound to be utilized as a highly selective and sensitive fluorescent sensor for Hg2+ ions;
  • Selectivity
The selectivity towards Hg2+ ions was investigated by adding 150 µL of Hg2+ (1.00 × 10−3 M) to the NI-2 solution (6.54 × 10−6 M) containing 150 µL of background metal ion (1.00 × 10−3 M) in MeOH/buffer mixture (v/v, 1/2, pH 5.5) (Figure 11).
Compound NI-2 shows a greater selectivity relative to NI-1, which can be attributed to the structural difference of their receptors. Namely, in NI-2 the hydroxyphenyl group is attached directly to piperazine via a nitrogen atom which makes the receptor structure more rigid compared to NI-1 in which pyridine is linked to piperazine by a methylene bond. The stiffer receptor structure in NI-2 prevents it from coordinating various ions as is the case in NI-1;
  • Metal-binding studies
Using Job’s method, analogous to NI-1, the supposed stoichiometric binding coefficient of Hg2+ and NI-2 was determined in MeOH/buffer mixture at pH 5.5. As shown in Figure S11, the dependence of fluorescence intensity at 525 nm on the molar content of NI-2 in solution in the presence of Hg2+ ions might vaguely suggest a 2:1 metal–ligand binding ratio, which, given the structure of ligand, is however unlikely. We could conclude here that Job’s plot failed to reveal the real stoichiometry of the complex which is not unusual given the known limitations of the method [62].
The Benesi–Hildebrand’s method was employed by plotting the F0/(FF0) against 1/[Hg2+], where F is the fluorescence intensity at 525 nm in the presence of the metal ion at the given concentration, and F0 is the fluorescence intensity of the free compound (Figure S12). The association constant was evaluated from the ratio of the intercept and the slope to be 8278 M−1, which is similar to previously reported results [37];
  • LOD (limit of detection)
As one of the most important sensor features for useful applications, the limit of detection (LOD) of NI-2 for Hg2+ ions was determined using the IUPAC-recommended equation.
Figure S13 showed good linearity between the emission at 525 nm and concentrations of Hg2+ in the range from 0 to 25.0 μM, indicating that the compound NI-2 can measure specific concentrations of Hg2+ quantitatively.
In the linear range between 0–25 µM of the calibration curve, the limit of detection can be obtained as LOD = 3σ/k, where σ is standard deviation of blank measurements and k is the slope between the fluorescence intensity and the metal concentration. The standard deviation for four blank measurements is σ = 0,621 and the limit of detection was calculated to be 4.1 × 10−7 M, being comparable with similar chemosensors based on naphthalimides [26,37,63];
  • Reversibility
The reversibility test was performed by the addition of Hg2+ and EDTA to the NI-2 solution. The addition of Hg2+ increased fluorescence intensity of NI-2 as shown in Figure S14, but there was no change in the spectrum after the addition of EDTA to the solution of NI-2-Hg2+.
Table 3 shows a comparison between the different mercury sensors and napthalimide sensors NI-1 and NI-2 presented in this work.
It is important to note that nearly all the literature data in the area of chemosensors research for determination of complex stoichiometry, binding constants and quenching constants are based on the use of “old” graphical methods, such as Job’s plot, Benesi–Hildebrand (BH) or Stern–Volmer (SV) plots. Those methods show significant limitations and are often used non-critically to draw conclusions that cannot be made without more detailed analysis [67,68,69]. In this context, it is important to stress that all constants reported in our work are conditional constants. They were calculated under defined conditions, i.e., certain mixtures of solvents and pH values of buffer solutions. In addition to chelation equilibrium reactions assumed to be happening during metal-ion titrations, there are other possible reactions simultaneously happening in the solution including protolytic equilibrium of the sensor molecules and hydrolysis of metal cations. Calculation of exact binding constants would require more sophisticated calculation tools and different measurements, which for purely analytical applications is usually not necessary. Nevertheless, one has to be aware of the limitations of these graphical methods and use them accordingly.

3. Materials and Methods

3.1. Reagents and Instruments

Indicator dyes NI-1 and NI-2 were from Joanneum Research Forschungsgesellschaft mbH (Austria) [43]. All chemicals and solvents were purchased from commercial suppliers Acros, Sigma Aldrich or Fluka and were used as received without further purification. Methanol for spectroscopic measurements was of spectroscopic (HPLC) grade. A phosphate buffer was used for the near neutral pH range (5.0–8.0) and citric buffer for the acidic pH range (1.2–4.8). Buffer solutions were prepared using chemicals of analytical grade following the procedure [70]. The pH values of buffer solutions were monitored using MA 5740 pH meter.
The 1.0 × 10−3 M stock solutions of metal ions (Ag+, Mg2+, Ba2+, Cu2+, Hg2+, Zn2+, Co2+, Fe2+, Ni2+, Pb2+, Mn2+, Cd2+, Fe3+, Al3+) were prepared by dissolving metal salts (AgNO3, MgCl2, BaCl2, CuCl2, Hg(NO3)2, ZnSO4, Co(CH3CO2)2, FeSO4, NiSO4, Pb(NO3)2, MnCl2, (CH3CO2)2Cd, FeCl3, AlCl3) in 0.01 M HNO3. The UV-Vis spectra were recorded, against the solvent, at 25 °C, on a Varian Cary spectrophotometer in double-beam mode using quartz cuvette (1 cm). Wavelength scan was performed between 200 nm and 800 nm. Baseline was recorded prior to each set of experiments. Fluorescence measurements were carried out on a Varian Cary Eclipse fluorescence spectrophotometer. Emission spectra were recorded from 380 nm to 800 nm with excitation wavelength at 410 nm, at 25 °C using 1 cm path quartz cells. Relative fluorescence quantum yields were determined according to Miller using Equation (3):
Φ x = Φ x × A s D x n x 2 / A x D s n s 2
where Φ is the emission quantum yield; A is the absorbance at the excitation wavelength; D is the area under the corrected emission curve and n is the refractive index of the solvents used. The subscripts s and x refer to the standard and to the unknown, respectively. The standard employed was quinine sulphate with a published fluorescence quantum yield of 0.54 [71].

3.2. UV-Vis Absorbance and Fluorescence Spectra Measurements

Stock solutions of compounds were prepared by dissolving 1.0 mg of compound in 100.0 mL methanol to form a 1.97 × 10−5 M (NI-1) and 1.96 × 10−5 M (NI-2) solutions.

3.2.1. The Effect of pH

In order to study the effect of pH on the spectroscopic properties of new compounds, UV-Vis and fluorescence emission spectra were recorded covering the pH range 1.6–8.0 for NI-1 or 1.2–8.0 for NI-2. Testing solutions of compounds were prepared by diluting 1.0 mL of stock solution with 2.0 mL buffer and the final concentrations of NI-1 and NI-2 for the absorbance and fluorescence measurements were 6.56 × 10−6 M and 6.54 × 10−6 M, respectively.

3.2.2. Metal Ion Sensing

All measurements of fluorescent properties were obtained in MeOH/buffer pH 5.5 solution (1:2) with excitation wavelength at 410 nm. The visual evaluation of NI-1 and NI-2 sensing properties towards metal cations were performed prior to spectroscopic measurements. The experiment was carried out in 1.0 mL MeOH/buffer mixture (1:2) at a compound concentration 6.56 × 10−6 M and 6.54 × 10−6 M, respectively, followed by addition of 30.0 µL (2.00 equiv.) of each metal ion stock solution.
The spectroscopic evaluation of the metal-sensing properties of NI-1 was carried out at a concentration of 5.0 × 10−7 M. The effect of metal cation on emission spectra was tested by adding 3.0, 30.0 and 150.0 µL of stock solution of each metal cation. For titration experiments, in the prepared solution of compound, small aliquots of stock solutions of metal ions were added. The resulting solution of compound and metal ions were mixed and the fluorescence spectra were recorded after 2 min.
For the evaluation of the sensing properties of NI-2 towards different metal cations, spectroscopic measurements were obtained at a compound concentration of 6.54 × 10−6 M, followed by addition of 150.0 µL of metal ion stock solution.

4. Conclusions

In this paper we investigated the photophysical characteristics of two novel 1,8-naphthalimide-based fluorescent compounds, NI-1 and NI-2, and their potential use as optical probes for pH and heavy metal ions. Different molecular moieties attached at the C-4 position in NI-1 and NI-2 show a notable effect on the optical properties of the derivatives. NI-1 bearing a methyl-pyridyl substituent exhibits 6-fold higher quantum yield than hydroxyphenyl substituted NI-2. Both NI-1 and NI-2 showed fluorescence intensity increase with decreasing pH. The corresponding pKa values for NI-1 and NI-2 were determined to be 4.98 and 2.91, respectively.
When it comes to metal sensing, NI-1 and NI-2 show opposite behaviors. The compound NI-1 shows high sensitivity and selectivity toward Cu2+ and Hg2+, via fluorescence quenching (switch “off” probe), while the compound NI-2 undergoes significant fluorescence enhancement exclusively by Hg2+ ions (switch “on” probe).
The fluorescence intensity of NI-1 decreases upon complexation with Cu2+ and the chemosensor can be applied to the quantification of Cu2+ ions within a linear range from 2.5 × 10−7 M to 4.0 × 10−6 M with a detection limit of 1.5 × 10−8 M. The linear range for Hg2+ was determined to be 2.0 × 10−6 M to 1.0 × 10−5 M, with a detection limit of 8.8 × 10−8 M.
Coordination of the receptor in NI-2 with Hg2+ hinders the PET mechanism from the aliphatic amino group to the naphthalimide core, resulting in fluorescence enhancement. Therefore, the NI-2 compound can be utilized as a turn-on PET sensor for Hg2+ ions. The results show a remarkable selectivity towards Hg2+ ions, with a linear detection range from 1.3 × 10−6 to 2.5 × 10−5 M and a limit of detection of 4.1 × 10−7 M.
Due to the presence of a 2-hydroxyethylsulfonyl group, the covalent immobilization of these naphthalimide probes to different substrate materials and to nanoparticles (e.g., cellulose, textiles) is possible, opening up a wide range of solid state applications for these molecules in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28031275/s1, Figure S1: Absorbance spectra of: (a) NI-1 and (c) NI-2 in a buffer/methanol mixture (2:1) at different pH; plot of absorbance ratio at 388 nm and 410 nm vs. pH for: (b) NI-1 and (d) NI-2; Figure S2: (a) Fluorescence spectra of NI-2 in buffer methanol mixture (2:1) at different pH; (b) fluorescence intensity at 525 nm vs. pH; Figure S3: NI-1 in the presence of metal ions (from left: blank, Zn2+, Fe2+, Fe3+, Cu2+, Al3+, Hg2+, Pb2+, Co2+, Ni2+, Ag+, Mn2+, Ba2+, Cd2+, Mg2+, Ca2+) c (metal ion) = 3.0 × 10−5 M; Figure S4: UV-Vis spectra of NI-1 solution upon the addition of 20 equiv. of (a) Hg(II) and (b) Cu(II) ions; Figure S5: Changes in the emission spectra of NI-1 upon gradual addition of Cu2+ and EDTA; Figure S6: Changes in the emission spectra of NI-1 upon gradual addition of Hg2+ and EDTA; Figure S7: Benesi–Hildebrand’s plot in linear range for NI-1 binding with (a) Cu2+ (b) Hg2+; Figure S8: Dependence of fluorescence intensity of NI-1 (5.00 × 10−7 M) in MeOH/buffer pH 5.5 (v/v, 1/2) at 523 nm on the ion concentration in the linear range of dependence for (a) Cu2+ (b) Hg2+ ion; Figure S9: Stern–Volmer plot of the ratio of the fluorescence intensities (F0/F) of NI-1 against the increasing concentration of (a) Cu2+ and (b) Hg2+ ions in linear range; Figure S10: UV-Vis spectra of NI-2 upon the addition of Hg(II) ions; Figure S11: Job’s plot fluorescence intensity at 525 nm on the molar content of NI-2 in presence of Hg2+ ions at a total concentration of 5.00 × 10−6 M; The extrapolated directions of intensity dependence on the molar fraction of NI-2 are presented; Figure S12: Benesi–Hildebrand’s plot in linear range for NI-2 binding with Hg2+; Figure S13: Dependence of fluorescence intensity of NI-2 (6.53 × 10−6 M) in MeOH/buffer 5.5 (v/v, 1/2) at 525 nm on the ion concentration in the linear range of dependence for Hg2+ ion; Figure S14: Reversibility test for NI-2 and Hg(II) binding.

Author Contributions

Conceptualization, I.M.S., G.J.M. and S.K.; methodology, I.M.S. and G.J.M.; validation, K.P. and I.M.S.; formal analysis, K.P., M.M., I.M.S. and S.K.; investigation, M.M., K.P. and M.G.; resources, I.M.S. and G.J.M.; data curation, M.M., K.P. and M.G.; writing—original draft preparation, K.P. and M.M.; writing—review and editing, I.M.S. and S.K., visualization, K.P., M.M. and M.G.; supervision, I.M.S. and S.K.; funding acquisition, I.M.S. and G.J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Zagreb, under research grant numbers 121054 (2020), “Integrated analytical chemical systems: development and application of chemical sensors and biosensors”, and grant number 121083 (2021) “Integrated analytical chemical systems: development and application of wearable sensors”. G. J. M. acknowledges funding by the Austrian Research Promotion Agency FFG within Bridge 1 project GlucoVin (project number 880559).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article and the Supplementary Materials.

Acknowledgments

The authors would also like to acknowledge Matthew Steinberg for English language and style proofreading and correction.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Li, M.; Gou, H.; Al-Ogaidi, I.; Wu, N. Nanostructured Sensors for Detection of Heavy Metals: A Review. ACS Sustain. Chem. Eng. 2013, 1, 713–723. [Google Scholar] [CrossRef]
  2. Bhardwaj, V.; Nurchi, V.M.; Sahoo, S.K. Mercury Toxicity and Detection Using Chromo-Fluorogenic Chemosensors. Pharmaceuticals 2021, 14, 123. [Google Scholar] [CrossRef]
  3. Tchekwagep, P.M.S.; Crapnell, R.D.; Banks, C.E.; Betlem, K.; Rinner, U.; Canfarotta, F.; Lowdon, J.W.; Eersels, K.; van Grinsven, B.; Peeters, M.; et al. A Critical Review on the Use of Molecular Imprinting for Trace Heavy Metal and Micropollutant Detection. Chemosensors 2022, 10, 296. [Google Scholar] [CrossRef]
  4. Ullah, N.; Chohan, M.M.; Khan, I.; Qurashi, A. Nanomaterial-based optical chemical sensors for the detection of heavy metals in water: Recent advances and challenges. TrAC Trends Anal. Chem. 2018, 100, 155–166. [Google Scholar] [CrossRef]
  5. Garg, N.; Deep, A.; Sharma, A.L. Recent Trends and Advances in Porous Metal-Organic Framework Nanostructures for the Electrochemical and Optical Sensing of Heavy Metals in Water. Crit. Rev. Anal. Chem. 2022, 13, 1–25. [Google Scholar]
  6. Nagarajan, R.; Varadaraju, C.; Lee, K.H. Recent advancements in the role of N-Heterocyclic receptors on heavy metal ion sensing. Dyes Pigm. 2021, 191, 109331. [Google Scholar] [CrossRef]
  7. De Acha, N.; Elosua, C.; Corres, J.M.; Arregui, F.J. Fluorescent Sensors for the Detection of Heavy Metal Ions in Aqueous Media. Sensors 2019, 19, 599. [Google Scholar] [CrossRef] [Green Version]
  8. Valeur, B.; Leray, I. Design principles of fluorescent molecular sensors for cation recognition. Coord. Chem. Rev. 2000, 205, 3–40. [Google Scholar] [CrossRef]
  9. Murkovic, I.; Wolfbeis, O.S. Fluorescence-based sensor membrane for mercury(II) detection. Sens. Actuators B: Chem. 1997, 39, 246–251. [Google Scholar] [CrossRef]
  10. Murkovic, I.; Oehme, I.; Mohr, G.J.; Ferber, T.; Wolfbeis, O.S. Optode membrane for continuous measurement of silver ions. Mikrochim. Acta 1995, 121, 249–258. [Google Scholar] [CrossRef]
  11. Hranjec, M.; Horak, E.; Tireli, M.; Pavlović, G.; Karminski-Zamola, G. Synthesis, crystal structure and spectroscopic study of novel benzimidazoles and benzimidazo[1,2-a]quinolines as potential chemosensors for different cations. Dyes Pigm. 2012, 95, 644–656. [Google Scholar] [CrossRef]
  12. de Silva, A.P.; Gunaratne, H.Q.N.; Gunnlaugsson, T.; Huxley, A.J.M.; McCoy, C.P.; Rademacher, J.T.; Rice, T.E. Signaling Recognition Events with Fluorescent Sensors and Switches. Chem. Rev. 1997, 97, 1515–1566. [Google Scholar] [CrossRef]
  13. Quang, D.T.; Kim, J.S. Fluoro- and Chromogenic Chemodosimeters for Heavy Metal Ion Detection in Solution and Biospecimens. Chem. Rev. 2010, 110, 6280–6301. [Google Scholar] [CrossRef]
  14. Lobnik, A.; Turel, M.; Urek, S.K. Optical Chemical Sensors:Design and Applications. In Advances in Chemical Sensors; Wang, W., Ed.; IntechOpen: London, UK, 2012; pp. 3–28. [Google Scholar]
  15. Carter, K.P.; Young, A.M.; Palmer, A.E. Fluorescent Sensors for Measuring Metal Ions in Living Systems. Chem. Rev. 2014, 114, 4564–4601. [Google Scholar] [CrossRef]
  16. Wu, D.; Sedgwick, A.C.; Gunnlaugsson, T.; Akkaya, E.U.; Yoon, J.; James, T.D. Fluorescent chemosensors: The past, present and future. Chem. Soc. Rev. 2017, 46, 7105–7123. [Google Scholar] [CrossRef] [Green Version]
  17. Duke, R.M.; Veale, E.B.; Pfeffer, F.M.; Kruger, P.E.; Gunnlaugsson, T. Colorimetric and fluorescent anion sensors: An overview of recent developments in the use of 1,8-naphthalimide-based chemosensors. Chem. Soc. Rev. 2010, 39, 3936–3953. [Google Scholar] [CrossRef] [Green Version]
  18. Daly, B.; Ling, J.; de Silva, A.P. Current developments in fluorescent PET (photoinduced electron transfer) sensors and switches. Chem. Soc. Rev. 2015, 44, 4203–4211. [Google Scholar] [CrossRef] [Green Version]
  19. Yordanova-Tomova, S.; Cheshmedzhieva, D.; Stoyanov, S.; Dudev, T.; Grabchev, I. Synthesis, Photophysical Characterization, and Sensor Activity of New 1,8-Naphthalimide Derivatives. Sensors 2020, 20, 3892. [Google Scholar] [CrossRef]
  20. Nolan, E.M.; Lippard, S.J. Tools and tactics for the optical detection of mercuric ion. Chem. Rev. 2008, 108, 3443–3480. [Google Scholar] [CrossRef]
  21. Escandar, G.M.; Olivieri, A.C. A Critical Review on the Development of Optical Sensors for the Determination of Heavy Metals in Water Samples. The Case of Mercury(II) Ion. ACS Omega 2022, 7, 39574–39585. [Google Scholar] [CrossRef]
  22. Choudhury, N.; Saha, B.; De, P. Recent progress in polymer-based optical chemosensors for Cu2+ and Hg2+ Ions: A comprehensive review. Eur. Polym. J. 2021, 145, 110233. [Google Scholar] [CrossRef]
  23. Gerdan, Z.; Saylan, Y.; Denizli, A. Recent Advances of Optical Sensors for Copper Ion Detection. Micromachines 2022, 13, 1298. [Google Scholar] [CrossRef] [PubMed]
  24. Tang, L.; Sharma, S.; Pandey, S.S. Synthesis and Characterization of Newly Designed and Highly Solvatochromic Double Squaraine Dye for Sensitive and Selective Recognition towards Cu2+. Molecules 2022, 27, 6578. [Google Scholar] [CrossRef]
  25. Queirós, C.; Almodovar, V.A.S.; Martins, F.; Leite, A.; Tome, A.C.; Silva, A.M.G. Synthesis of Novel Diketopyrrolopyrrole-Rhodamine Conjugates and Their Ability for Sensing Cu2+ and Li+. Molecules 2022, 27, 7219. [Google Scholar] [CrossRef]
  26. Georgiev, N.I.; Dimitrova, M.D.; Mavrova, A.T.; Bojinov, V.B. Synthesis, fluorescence-sensing and molecular logic of two water-soluble 1,8-naphthalimides. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 183, 7–16. [Google Scholar] [CrossRef]
  27. McGlynn, S.P.; Sunseri, R.; Christodouleas, N. External Heavy-Atom Spin-Orbital Coupling Effect. I. The Nature of the Interaction. J. Chem. Phys. 1962, 37, 1818–1824. [Google Scholar] [CrossRef]
  28. Tian, M.; Wang, C.; Ma, Q.; Bai, Y.; Sun, J.; Ding, C. A Highly Selective Fluorescent Probe for Hg2+ Based on a 1,8-Naphthalimide Derivative. ACS Omega 2020, 5, 18176–18184. [Google Scholar] [CrossRef]
  29. Jain, N.; Kaur, N. A comprehensive compendium of literature of 1,8-Naphthalimide based chemosensors from 2017 to 2021. Coord. Chem. Rev. 2022, 459, 214454. [Google Scholar] [CrossRef]
  30. Georgiev, N.I.; Krasteva, P.V.; Bakov, V.V.; Bojinov, V.B. A Highly Water-Soluble and Solid State Emissive 1,8-Naphthalimide as a Fluorescent PET Probe for Determination of pHs, Acid/Base Vapors, and Water Content in Organic Solvents. Molecules 2022, 27, 4229. [Google Scholar] [CrossRef]
  31. Bakov, V.V.; Georgiev, N.I.; Bojinov, V.B. A Novel Fluorescent Probe for Determination of pH and Viscosity Based on a Highly Water-Soluble 1,8-Naphthalimide Rotor. Molecules 2022, 27, 7556. [Google Scholar] [CrossRef]
  32. Li, Z.; Chen, W.; Dong, L.; Song, Y.; Li, R.; Li, Q.; Qu, D.; Zhang, H.; Yang, Q.; Li, Y. A novel ratiometric and reversible fluorescent probe based on naphthalimide for the detection of Al3+ and pH with excellent selectivity. New J. Chem. 2020, 44, 3261–3267. [Google Scholar] [CrossRef]
  33. Zhu, H.; Liu, C.; Su, M.; Rong, X.; Zhang, Y.; Wang, X.; Wang, K.; Li, X.; Yu, Y.; Zhang, X.; et al. Recent advances in 4-hydroxy-1,8-naphthalimide-based small-molecule fluorescent probes. Coord. Chem. Rev. 2021, 448, 214153. [Google Scholar] [CrossRef]
  34. Zhou, X.; Su, F.; Gao, W.; Tian, Y.; Youngbull, C.; Johnson, R.H.; Meldrum, D.R. Triazacryptand-based fluorescent sensors for extracellular and intracellular K+ sensing. Biomaterials 2011, 32, 8574–8583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Panchenko, P.A.; Fedorov, Y.V.; Perevalov, V.P.; Jonusauskas, G.; Fedorova, O.A. Cation-Dependent Fluorescent Properties of Naphthalimide Derivatives with N-Benzocrown Ether Fragment. J. Phys. Chem. A 2010, 114, 4118–4122. [Google Scholar] [CrossRef] [PubMed]
  36. Li, C.Y.; Xu, F.; Li, Y.-F.; Zhou, K.; Zhou, Y. A fluorescent chemosensor for Hg2+ based on naphthalimide derivative by fluorescence enhancement in aqueous solution. Anal. Chim. Acta 2012, 717, 122–126. [Google Scholar] [CrossRef]
  37. Wu, X.-F.; Ma, Q.-J.; Wei, X.-J.; Hou, Y.-M.; Zhu, X. A selective fluorescent sensor for Hg2+ based on covalently immobilized naphthalimide derivative. Sens. Actuators B Chem. 2013, 183, 565–573. [Google Scholar] [CrossRef]
  38. Kim, S.Y.; Hong, J.-I. Naphthalimide-based fluorescent Zn2+ chemosensors showing PET effect according to their linker length in water. Tetrahedron Lett. 2009, 50, 2822–2824. [Google Scholar] [CrossRef]
  39. Gao, Y.-G.; Liu, F.-L.; Patil, S.; Li, D.-J.; Qadir, A.; Lin, X.; Tian, Y.; Li, Y.; Qian, A.-R. 1,8-Naphthalimide-Based Multifunctional Compounds as Cu2+ Probes, Lysosome Staining Agents, and Non-viral Vectors. Front. Chem. 2019, 7, 616. [Google Scholar] [CrossRef] [PubMed]
  40. Aderinto, S.O.; Imhanria, S. Fluorescent and colourimetric 1, 8-naphthalimide-appended chemosensors for the tracking of metal ions: Selected examples from the year 2010 to 2017. Chem. Pap. 2018, 72, 1823–1851. [Google Scholar] [CrossRef]
  41. Gopikrishna, P.; Meher, N.; Iyer, P.K. Functional 1,8-Naphthalimide AIE/AIEEgens: Recent Advances and Prospects. ACS Appl. Mater. Interfaces 2018, 10, 12081–12111. [Google Scholar] [CrossRef] [PubMed]
  42. Singh, N.; Srivastava, R.; Singh, A.; Singh, R.K. Synthesis and Photophysical Studies on Naphthalimide Derived Fluorophores as Markers in Drug Delivery. J. Fluoresc. 2016, 26, 1431–1438. [Google Scholar] [CrossRef] [PubMed]
  43. Mohr, G.J. Synthesis of naphthalimide-based indicator dyes with a 2-hydroxyethylsulfonyl function for covalent immobilisation to cellulose. Sens. Actuators B Chem. 2018, 275, 439–445. [Google Scholar] [CrossRef]
  44. de Silva, A.P. Crossing the divide: Experiences of taking fluorescent PET (photoinduced electron transfer) sensing/switching systems from solution to solid. Dyes Pigm. 2022, 204, 110453. [Google Scholar] [CrossRef]
  45. Gan, J.; Chen, K.; Chang, C.-P.; Wu, W. Luminescent properties and photo-induced electron transfer of naphthalimides with piperazine substituent. Dyes Pigm. 2003, 57, 21–28. [Google Scholar] [CrossRef]
  46. Tian, Y.; Su, F.; Weber, W.; Nandakumar, V.; Shumway, B.R.; Jin, Y.; Zhou, X.; Holl, M.R.; Johnson, R.H.; Meldrum, D.R. A series of naphthalimide derivatives as intra and extracellular pH sensors. Biomaterials 2010, 31, 7411–7422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Lee, M.H.; Park, N.; Yi, C.; Han, J.H.; Hong, J.H.; Kim, K.P.; Kang, D.H.; Sessler, J.L.; Kang, C.; Kim, J.S. Mitochondria-immobilized pH-sensitive off–on fluorescent probe. J. Am. Chem. Soc. 2014, 136, 14136–14142. [Google Scholar] [CrossRef] [PubMed]
  48. Kriltz, A.; Löser, C.; Mohr, G.J.; Trupp, S. Covalent immobilization of a fluorescent pH-sensitive naphthalimide dye in sol-gel films. J. Sol-Gel Sci. Technol. 2012, 63, 23–29. [Google Scholar] [CrossRef]
  49. Fu, Y.; Zhang, J.; Wang, H.; Chen, J.L.; Zhao, P.; Chen, G.R.; He, X.P. Intracellular pH sensing and targeted imaging of lysosome by a galactosyl naphthalimide-piperazine probe. Dyes Pigm. 2016, 133, 372–379. [Google Scholar] [CrossRef]
  50. Lee, M.H. Bis(Naphthalimide-Piperazine)-Based Off-On Fluorescent Probe for Acids. J. Fluoresc. 2016, 26, 807–811. [Google Scholar] [CrossRef]
  51. Liu, Y.; Zhang, J.; Feng, T.; Li, Y. Synthesis, structure–fluorescence relationships and density functional theory studies of novel naphthalimide–piperazine–pyridine-based polystyrene sensors for Hg(II) detection. RSC Adv. 2020, 10, 25281–25289. [Google Scholar] [CrossRef]
  52. Zhang, H.; Qu, Y.; Zhao, K.; Wang, C.; Wu, Y.; Wu, H. A fluorescence “on-off” sensor for the highly selective and sensitive detection of Cu2+ ion. J. Chin. Chem. Soc. 2020, 67, 1062–1069. [Google Scholar] [CrossRef]
  53. Qu, Y.; Wu, Y.; Wang, C.; Zhao, K.; Wu, H. A selective fluorescence probe for copper(II) ion in aqueous solution based on a 1,8-naphthalimide Schiff base derivative. Z. Naturforsch. B 2019, 74, 665–670. [Google Scholar] [CrossRef] [Green Version]
  54. Huang, G.Z.; Li, C.; Han, X.; Aderinto, S.O.; Shen, K.; Mao, S.; Wu, H. Sensitive and selective detection of Cu(II) ion: A new effective 1,8-naphthalimide-based fluorescence "turn off’ sensor. Luminescence 2018, 33, 660–669. [Google Scholar] [CrossRef]
  55. Dong, H.Q.; Wei, T.-B.; Ma, X.-Q.; Yang, Q.-Y.; Zhang, Y.-F.; Sun, Y.-J.; Shi, B.-B.; Yao, H.; Zhang, Y.-M.; Lin, Q. 1,8-Naphthalimide-based fluorescent chemosensors: Recent advances and perspectives. J. Mater. Chem. C 2020, 8, 13501–13529. [Google Scholar] [CrossRef]
  56. Anand, T.; Sivaraman, G.; Chellappa, D. Quinazoline copper(II) ensemble as turn-on fluorescence sensor for cysteine and chemodosimeter for NO. J. Photochem. Photobiol. A Chem. 2014, 281, 47–52. [Google Scholar] [CrossRef]
  57. Arslan, F.N.; Geyik, G.A.; Koran, K.; Ozen, F.; Aydin, D.; Elmas, S.N.K.; Gorgulu, A.O.; Yilmaz, I. Fluorescence “Turn On–Off” Sensing of Copper (II) Ions Utilizing Coumarin–Based Chemosensor: Experimental Study, Theoretical Calculation, Mineral and Drinking Water Analysis. J. Fluoresc. 2020, 30, 317–327. [Google Scholar] [CrossRef]
  58. Kim, G.; Choi, D.; Kim, C. A Benzothiazole-Based Fluorescence Turn-on Sensor for Copper(II). J. Fluoresc. 2021, 31, 1203–1209. [Google Scholar] [CrossRef]
  59. Silpcharu, K.; Soonthonhut, S.; Sukwattanasinitt, M.; Rashatasakhon, P. Fluorescent Sensor for Copper(II) and Cyanide Ions via the Complexation−Decomplexation Mechanism with Di(bissulfonamido)spirobifluorene. ACS Omega 2021, 6, 16696–16703. [Google Scholar] [CrossRef]
  60. Tamil Selvan, G.; Varadaraju, C.; Tamil Selvan, R.; Enoch, I.V.M.V.; Selvakumar, M. On/Off Fluorescent Chemosensor for Selective Detection of Divalent Iron and Copper Ions: Molecular Logic Operation and Protein Binding. ACS Omega 2018, 3, 7985–7992. [Google Scholar] [CrossRef] [Green Version]
  61. Xu, J.-H.; Hou, Y.-M.; Ma, Q.-J.; Wu, X.-F.; Wei, X.-J. A highly selective fluorescent sensor for Fe3+ based on covalently immobilized derivative of naphthalimide. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 112, 116–124. [Google Scholar] [CrossRef]
  62. Brynn Hibbert, D.; Thordarson, P. The death of the Job plot, transparency, open science and online tools, uncertainty estimation methods and other developments in supramolecular chemistry data analysis. Chem. Comm. 2016, 52, 12792–12805. [Google Scholar] [CrossRef] [PubMed]
  63. Li, G.; Gao, G.; Cheng, J.; Chen, X.; Zhao, Y.; Ye, Y. Two new reversible naphthalimide-based fluorescent chemosensors for Hg2+. Luminescence 2016, 31, 992–996. [Google Scholar] [CrossRef] [PubMed]
  64. Lei, F.; Shi, W.; Ma, J.; Yabin Chen, Y.; Kui, F.; Hui, Y.; Xie, Z. A novel thiosemicarbazone Schiff base derivative with aggregation-induced emission enhancement characteristics and its application in Hg2+ detection. Sens. Actuators B Chem. 2016, 237, 563–569. [Google Scholar]
  65. Sie, Y.-W.; Li, C.-L.; Wan, C.-F.; Chen, J.-H.; Hu, C.-H.; Wu, A.-T. 1,10-Phenanthroline based colorimetric and fluorescent sensor for Hg2+ in water: Experimental and DFT study. Inorganica Chim. Acta 2018, 469, 397–401. [Google Scholar] [CrossRef]
  66. Su, Q.; Niu, Q.; Sun, T.; Li, T. A simple fluorescence turn-on chemosensor based on Schiff-base for Hg2+-selective detection. Tetrahedron Lett. 2016, 57, 4297–4301. [Google Scholar] [CrossRef]
  67. Gamov, G.A. Processing of the spectrofluorimetric data using the graphical methods and the maximum likelihood approach. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2021, 249, 119334. [Google Scholar] [CrossRef]
  68. Davidge, J.; Thomas, C.P.; Williams, D.R. Conditional formation constants or chemical speciation data? Chem. Speciat. Bioavailab. 2001, 13, 129–134. [Google Scholar] [CrossRef]
  69. van de Weert, M.; Stella, L. Fluorescence quenching and ligand binding: A critical discussion of a popular methodology. J. Mol. Struct. 2011, 998, 144–150. [Google Scholar] [CrossRef]
  70. Lentner, C. Wissenschaftliche Tabellen Geigy, 8th ed.; Ciba-Geigy: Basel, Switzerland, 1985. [Google Scholar]
  71. Fletcher, A.N. Quinine sulfate as a fluorescence quantum yield standard. Photochem. Photobiol. 1969, 9, 439–444. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of NI-1 and NI-2.
Figure 1. Structures of NI-1 and NI-2.
Molecules 28 01275 g001
Figure 2. (a) UV-Vis absorption and fluorescence emission spectra of NI-1 and NI-2 solutions in methanol (c = 1.0 × 10−5 M, λexc = 410 nm); (b) Images of solutions in daylight (above) and under UV lamp (365 nm) (down).
Figure 2. (a) UV-Vis absorption and fluorescence emission spectra of NI-1 and NI-2 solutions in methanol (c = 1.0 × 10−5 M, λexc = 410 nm); (b) Images of solutions in daylight (above) and under UV lamp (365 nm) (down).
Molecules 28 01275 g002
Figure 3. (a) Fluorescence spectra of NI-1 in buffer methanol mixture (2:1) at different pH; (b) fluorescence intensity at 523 nm vs. pH.
Figure 3. (a) Fluorescence spectra of NI-1 in buffer methanol mixture (2:1) at different pH; (b) fluorescence intensity at 523 nm vs. pH.
Molecules 28 01275 g003
Figure 4. Schematic representation of PET process using NI-1 as an example.
Figure 4. Schematic representation of PET process using NI-1 as an example.
Molecules 28 01275 g004
Figure 5. Change of fluorescence intensity at 523 nm for NI-1 (orange) and in the presence of metal ions in MeOH/buffer (v/v, 1/2, pH = 5.5) solution.
Figure 5. Change of fluorescence intensity at 523 nm for NI-1 (orange) and in the presence of metal ions in MeOH/buffer (v/v, 1/2, pH = 5.5) solution.
Molecules 28 01275 g005
Figure 6. (a) Fluorescence titration spectra of NI-1 (c = 5.00 × 10−7 M) in the presence of changing concentration of Cu2+ ions (c(Cu2+) = 5.00 × 10−8 M–5.00 × 10−5 M, as shown in the graph ranging from 0.1 eq to 100 eq) in MeOH/buffer (v/v, 1/2, pH = 5.5), λexc = 410 nm. (b) Plot of fluorescence quenching at 523 nm of NI-1 as a function of Cu2+ ion concentration.
Figure 6. (a) Fluorescence titration spectra of NI-1 (c = 5.00 × 10−7 M) in the presence of changing concentration of Cu2+ ions (c(Cu2+) = 5.00 × 10−8 M–5.00 × 10−5 M, as shown in the graph ranging from 0.1 eq to 100 eq) in MeOH/buffer (v/v, 1/2, pH = 5.5), λexc = 410 nm. (b) Plot of fluorescence quenching at 523 nm of NI-1 as a function of Cu2+ ion concentration.
Molecules 28 01275 g006
Figure 7. (a) Fluorescence titration spectra of NI-1 (c = 5.00 × 10−7 M) with Hg2+ in concentration range 5.00 × 10−8 M–5.00 × 10−5 M (shown in the graph ranging from 0.1 eq to 100 eq) in MeOH/buffer (v/v, 1/2, pH = 5.5), λexc = 410 nm. (b) Plot of fluorescence quenching at 523 nm of NI-1 as a function of Hg2+ ion concentration.
Figure 7. (a) Fluorescence titration spectra of NI-1 (c = 5.00 × 10−7 M) with Hg2+ in concentration range 5.00 × 10−8 M–5.00 × 10−5 M (shown in the graph ranging from 0.1 eq to 100 eq) in MeOH/buffer (v/v, 1/2, pH = 5.5), λexc = 410 nm. (b) Plot of fluorescence quenching at 523 nm of NI-1 as a function of Hg2+ ion concentration.
Molecules 28 01275 g007
Figure 8. Dependence of fluorescence intensity at 523 nm on the molar content of NI-1 in solution with (a) Cu2+ ions and (b) Hg2+ ions at a total concentration of 5.00 × 10−6 M and addition of buffer pH 5.5. The extrapolated directions of intensity dependence on the molar fraction of NI-1 are presented.
Figure 8. Dependence of fluorescence intensity at 523 nm on the molar content of NI-1 in solution with (a) Cu2+ ions and (b) Hg2+ ions at a total concentration of 5.00 × 10−6 M and addition of buffer pH 5.5. The extrapolated directions of intensity dependence on the molar fraction of NI-1 are presented.
Molecules 28 01275 g008
Figure 9. NI-2 in the presence of metal ions in MeOH/buffer mixture.
Figure 9. NI-2 in the presence of metal ions in MeOH/buffer mixture.
Molecules 28 01275 g009
Figure 10. (a) Fluorescence titration spectra of NI-2 (5.00 × 10−7 M) with Hg2+ in concentration range 1.31 × 10−7 M–2.26 × 10−5 M (shown in the graph ranging from 0.01 eq to 9.58 eq) in MeOH/buffer (1/2, v/v, pH = 5.5), λexc = 410 nm. (b) Plot of fluorescence enhancement at 525 nm of NI-2 as a function of Hg2+ ion concentration.
Figure 10. (a) Fluorescence titration spectra of NI-2 (5.00 × 10−7 M) with Hg2+ in concentration range 1.31 × 10−7 M–2.26 × 10−5 M (shown in the graph ranging from 0.01 eq to 9.58 eq) in MeOH/buffer (1/2, v/v, pH = 5.5), λexc = 410 nm. (b) Plot of fluorescence enhancement at 525 nm of NI-2 as a function of Hg2+ ion concentration.
Molecules 28 01275 g010
Figure 11. Relative fluorescence intensity changes at 525 nm of NI-2 (6.54 × 10−6 M) in MeOH/buffer solution (v/v, 1/2, pH 5.5) after addition of 150 µL stock solution of Hg2+ and other metal ions after 5 min from the addition of Hg2+ ions. The concentrations of Hg2+ ions and competing metal ions are equal (4.55 × 10−5 M).
Figure 11. Relative fluorescence intensity changes at 525 nm of NI-2 (6.54 × 10−6 M) in MeOH/buffer solution (v/v, 1/2, pH 5.5) after addition of 150 µL stock solution of Hg2+ and other metal ions after 5 min from the addition of Hg2+ ions. The concentrations of Hg2+ ions and competing metal ions are equal (4.55 × 10−5 M).
Molecules 28 01275 g011
Table 1. Effect of pH on spectral properties of NI-1 and NI-2 in neutral and protonated/cationic forms at pH 1.6 and pH 7.0, respectively.
Table 1. Effect of pH on spectral properties of NI-1 and NI-2 in neutral and protonated/cationic forms at pH 1.6 and pH 7.0, respectively.
pH (Form)λabs/nmε/M−1 cm−1λemiss/nmRel. Fluo. Int.Stokes Shift/cm−1pKa,abspKa,emiss
NI-11.6 (NI-1H+)3882662.90523836.266535.094.98
7.0 (NI-1)4102020.5853130.25558
NI-21.6 (NI-2H+)388988.53525769.867263.182.91
7.0 (NI-2)410916,58533<4.955629
Table 2. Solvent systems, fluorescence changes, mechanism, stoichiometry (Ligand: Analyte), K and LOD values of sensors on complexation with Cu2+ ions.
Table 2. Solvent systems, fluorescence changes, mechanism, stoichiometry (Ligand: Analyte), K and LOD values of sensors on complexation with Cu2+ ions.
Sensor StructureSolvent SystemsFluorescence Changes and MechanismStoichiometryK/M−1LOD/MRef.
Molecules 28 01275 i001DMSO: H2O
(1:9, v/v) at pH = 7.4 (HEPES)
Turn-off at 434 and 494 nm,
Paramagnetic quenching
1:1--[56]
Molecules 28 01275 i002CH3CN: HEPES (95:5, v/v)Turn-off at 458 nm,
Paramagnetic quenching
1:19.0 × 104 (BH)2.45 × 10−8[57]
Molecules 28 01275 i003CH3CNFluorescence enhancement at 540 nm (turn-on),
CHEF and ICT
1:12.0 × 103 (BH)3.3 × 10−6[58]
Molecules 28 01275 i004CH3CN: HEPES (1:1, v/v)
pH = 7.0
Quenching at 530 nm
(turn-off)
1:1-9.82 × 10−8[59]
Molecules 28 01275 i005H2O: DMF (9.9:0.1, v/v)Turn-off at 410 nm,
MLCT and paramagnetic quenching
1:29.23 × 108 (SV)1 × 10−7[60]
Molecules 28 01275 i006DMF: Tris-HCl buffer (1:1, v/v)Turn-off at 528 nm,
Paramagnetic quenching
2:14 × 10121.92 × 10−7[50]
Molecules 28 01275 i007MeOH: HEPES (1:1, v/v)Turn-off at 525 nm,
Paramagnetic quenching
2:11.1 × 1045.67 × 10−7[53]
Molecules 28 01275 i008DMSO: HEPES (1:1, v/v)Turn-off at 523 nm1:11.14 × 1064.67 × 10−8[54]
Molecules 28 01275 i009Tris-HCl bufferTurn-off at 541 nm1:2n.d.2.62 × 10−9[39]
Molecules 28 01275 i010HEPESTurn-off at 396 nm,
Paramagnetic quenching
1:1
1:1
5.7 × 105
2.5 × 105
1.5 × 10−6
2.5 × 10−6
[26]
Molecules 28 01275 i011MeOH: phosphate buffer (1:2, v/v)Turn-off at 523 nm1:13.6 × 105 (BH)
1.5 × 105 (SV)
1.5 × 10−8This work
K values were determined by using Benesi–Hildebrand (BH) or Stern–Volmer (SV) graphical methods.
Table 3. Solvent systems, fluorescence changes, mechanism, stoichiometry (Ligand: Analyte), K and LOD values of sensors on complexation with Hg2+ ions.
Table 3. Solvent systems, fluorescence changes, mechanism, stoichiometry (Ligand: Analyte), K and LOD values of sensors on complexation with Hg2+ ions.
Sensor StructureSolvent SystemsFluorescence Changes And MechanismStoichiometryK/M−1LOD/MRef.
Molecules 28 01275 i012DMSO: H2O
(1:9, v/v)
Quenching at 510 nm and red shift (50 nm),
AIEE
1:22.51 × 105 (BH)9.07 × 10−7[64]
Molecules 28 01275 i013H2O with 0.5 % DMSOQuenching at 480 and 540 nm1:21.77 × 105 (SV)21.72 ppb[65]
Molecules 28 01275 i014CH3CN: H2O (1:1, v/v)Turn-on at 415 nm and new band at 473 nm,
C=N isomerization and ESIPT inhibition
2:14.484 × 105 (BH)2.27 × 10−8[66]
Molecules 28 01275 i015Tris-HCl bufferTurn-on t at 520 nm, PET inhibition1:13.5 × 103 (BH)2.0 × 10−6[37]
Molecules 28 01275 i016CH3CN: HEPES (1:1, v/v)Enhancement 405 nm, PET inhibition--1.98 × 10−6[49]
Molecules 28 01275 i017Tris-HNO3 bufferEnhancement at 529 nm, PET inhibition1:12.08 × 105 (BH)4.93 × 10−8[36]
Molecules 28 01275 i018Phosphate bufferQuenching at 550 nm1:1-2.1 × 10−6[63]
Molecules 28 01275 i019HEPESQuenching at 396 nm,
PET inhibition
1:11.8 × 105
2.1 × 105 (SV)
1.9 × 10−6
2.3 × 10−6
[26]
Molecules 28 01275 i020MeOH: phosphate buffer (1:2, v/v)Quenching at 523 nm and red shift (18 nm)1:13.9 × 104 (BH)
2.4 × 104 (SV)
8.8 × 10−8This work
Molecules 28 01275 i021MeOH: phosphate buffer (1:2, v/v)Enhancement at 525 nm-8278 (BH)4.1 × 10−7This work
K values were determined by using Benesi–Hildebrand (BH) or Stern–Volmer (SV) graphical method.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pršir, K.; Matić, M.; Grbić, M.; Mohr, G.J.; Krištafor, S.; Steinberg, I.M. Naphthalimide-Piperazine Derivatives as Multifunctional “On” and “Off” Fluorescent Switches for pH, Hg2+ and Cu2+ Ions. Molecules 2023, 28, 1275. https://doi.org/10.3390/molecules28031275

AMA Style

Pršir K, Matić M, Grbić M, Mohr GJ, Krištafor S, Steinberg IM. Naphthalimide-Piperazine Derivatives as Multifunctional “On” and “Off” Fluorescent Switches for pH, Hg2+ and Cu2+ Ions. Molecules. 2023; 28(3):1275. https://doi.org/10.3390/molecules28031275

Chicago/Turabian Style

Pršir, Kristina, Mislav Matić, Marlena Grbić, Gerhard J. Mohr, Svjetlana Krištafor, and Ivana Murković Steinberg. 2023. "Naphthalimide-Piperazine Derivatives as Multifunctional “On” and “Off” Fluorescent Switches for pH, Hg2+ and Cu2+ Ions" Molecules 28, no. 3: 1275. https://doi.org/10.3390/molecules28031275

Article Metrics

Back to TopTop