DOI:
10.1039/C6RA09535F
(Paper)
RSC Adv., 2016,
6, 59073-59080
A fluorescent sensor for Cu2+ and Fe3+ based on multiple mechanisms†
Received
13th April 2016
, Accepted 13th June 2016
First published on 15th June 2016
Abstract
A new fluorescent sensor for the selective and sensitive sensing of Cu2+ and Fe3+ based on the rational control of multiple mechanisms has been developed. Sensor L is weakly-fluorescent (ϕ = 0.029) due to the C
N isomerization and PET process. Upon binding with Cu2+, the complex L + Cu2+ can give a remarkable fluorescence enhancement (ϕ = 0.182) with a limit of detection down to 1.3 × 10−8 mol L−1 because of both inhibition of the C
N isomerization and PET process. However, upon binding with Fe3+, the original weak fluorescence of L is completely quenched (ϕ = 0.002) due to the paramagnetic quenching property. In addition, the complex L + Cu2+ could be further used as a new platform for the ultrasensitive detection of Fe3+ based on the fluorescence “turn-off” response with a limit of detection down to 2.4 × 10−8 mol L−1 in aqueous solution.
Introduction
The development of highly selective and sensitive fluorescent sensors that can be used to detect bioactive metal ions has gained enormous importance.1–4 Metal ions are important for a variety of fundamental biological processes, but can also cause serious environmental and health problems due to their high toxicity. Copper is a significant environmental pollutant and yet also an essential trace element that acts as a cofactor in all currently known life forms.5–11 The limit of Cu2+ in drinking water has been set to 1.3 ppm (20 μM). Exposure to a high level of Cu2+ can cause a wide variety of symptoms (gastrointestinal disease, Wilson's disease, dyslexia, hypoglycemia, and infant liver damage).12–17 Thus, the detection of Cu2+ from various sources including those in waste water outlets, electroplating waste and other metal processing industries is important. Cu2+ complexation is well known to induce intrinsic fluorescence quenching,18–24 while chemosensors with fluorescence enhancement were more encouraging because of their simplicity in practical applications.25–30 Therefore, fluorescence ‘turn-on’ chemosensors with high selectivity and sensitivity towards Cu2+ are highly desirable.
As an important physiologically relevant metal ion, Fe3+ performs a significant role in a variety of vital cell functions, and both its excess (hyperferremia) and deficiency (hypoferremia) can induce a variety of diseases.31–34 The security limit for Fe3+ was restricted to 2 mg L−1 by the WHO.35,36 In the form of ferritin or hemosiderin iron is stored in liver, spleen and bone marrow of human body. However, excess of Fe3+ leads to increasing incidence of certain cancers and dysfunction of body organs (gastric upset, constipation, nausea, abdominalpain, vomiting and faintness), even causes hyperferremic and hypotransferrinemic disorder when Fe3+ accumulates in the body.37,38 Similarly, the deficiency of Fe3+ causes a disease namely anaemia and functional deficits associated with anaemia (gastrointestinal disturbances and impaired cognitive function, immune function, exercise or work performance, and body temperature regulation).39–41 In this regard, the development of chemosensor that can detect Fe3+ is very important to prevent and solve environmental and health problems induced by Fe3+.
Till now, a lot of effective fluorescent sensors had been reported which showed excellent optical response towards Cu2+ or Fe3+. However, most of them could detect only one metal ion (Cu2+ or Fe3+) in the solution, and this certainly limited their practical application.18,19,27 For these fluorescent sensors, there was only one mechanism playing a leading role (such as charge transfer or paramagnetic quenching) in the detection process of metal ions. In contrast, the sensors which could be used as platforms to detect multiple metal ions based on the rational control of multiple mechanisms exhibited a lot of advantages, such as high efficiency, multiplicity and low cost. Unfortunately, there are few articles carefully describing the role of multiple mechanisms and giving an accurate sensor as an example for the metal ions detection. As a result, it is still in urgent need of designing an ultrasensitive fluorescent sensor for the accurate detection of Cu2+ and Fe3+ without interference from other anions based on the rational control of multiple mechanisms.
In this work, a novel fluorescent sensor (L) based on the fluorophore of benzimidazo[2,1-a]benz[de]isoquinoline-7-one-12-carbohydrazide was prepared. With the introduction of 4-oxo-4H-chromene-3-carbaldehyde as a receptor in the molecule, the fluorophore and receptor were linked by a C
N bond and the PET (photoinduced electron transfer) process was directed from the receptor to fluorophore, giving L potential ability to be used as a fluorescent sensing platform towards multiple metal ions. It was found that the ligand L was actually weak-fluorescent. After binding with Cu2+, a new complex, L + Cu2+, was formed and showed a strong fluorescence enhancement due to the inhibition of both C
N isomerization and PET process. Upon binding with Fe3+, the original weak fluorescence of L completely quenched due to the paramagnetic quenching property. What is more, the lighted complex L + Cu2+ could be further used as a new sensing platform for ultrasensitive detection of Fe3+ based on the fluorescence turn-off response. Base on the rational control of multiple sensing mechanisms, Cu2+ and Fe3+ could be very sensitively and selectively detected.
Experimental section
Measurements
UV-vis spectra were recorded on a Shimadzu 3100 spectrometer. Fluorescence measurements were carried out using an HITACHI Instruments F-4600 fluorescence spectrophotometer. 1H NMR spectra were recorded on a Bruker AV III 400 MHz NMR spectrometer with tetramethysilane (TMS) as an internal standard. Infrared spectra were recorded using a Bruker Vertex 70 FT-IR spectrometer with KBr pellets.
The 13C CP/MAS NMR spectra were recorded on a Bruker AVANCE III 400 WB spectrometer equipped with a 4 mm standard bore CPMAS probehead whose X channel was tuned to 100.62 MHz for 13C and the other channel was tuned to 400.18 MHz for broad band 1H decoupling, using a magnetic field of 9.39 T at 297 K. The dried and finely powdered samples were packed in the ZrO2 rotor closed with Kel-F cap which were spun at 8 kHz rate. The experiments were conducted at a contact time of 2 ms. A total of 1500 scans were recorded with 3 s recycle delay for each sample. All 13C CP MAS chemical shifts are referenced to the resonances of adamantane (C10H16) standard (δCH2 = 38.4).
Sample preparation
All tests described in this paper were carried out at room temperature (25 °C) with distilled water. In the experiments of titration with various metal ions, the sensor was dissolved in HEPES acetonitrile–H2O (9
:
1) buffer solution to afford the test solution (1 × 10−5 M). Stock solutions (1 × 10−5 M) of the metal salts of LiCl, NaCl, KCl, MgCl2, CaCl2, BaCl2, NiCl2, CuCl2, ZnCl2, CdCl2, HgCl2, PbCl2, AgNO3, MnCl2, FeCl3, CoCl2, CrCl3, SrCl3 and AlCl3 in water were prepared.
Theoretical calculation
Density functional theory (DFT) structural optimizations were calculated with the Gaussian 09 program.42 In all cases, the structures were optimized using the B3LYP functional and the mixed basis sets 6-31+G(d) and LANL2DZ. Each structure was subsequently subjected to TD-DFT calculation using the B3LYP functional. For all optimized structures, frequency calculations were carried out to confirm the absence of imaginary frequencies. The molecular orbitals were visualized and plotted with the GaussView 5.0 program.
Calculation of quantum yield
The quantum yield of sensor L was determined according to the following equation:
where ϕ is fluorescence quantum yield; F is integrated area under the corrected emission spectra; A is the absorbance at the excitation wavelength; n is the refractive index of the solution; the subscripts u and s refer to the unknown and the standard, respectively. Rhodamine B in ethanol solution was used as the standard, which has a quantum yield of 0.97.
Synthesis
7-Oxo-N′-((4-oxo-4H-chromen-3-yl)methylene)-7H-benzo[de]benzo[4,5]imidazo[2,1-a]isoquinoline-12-carbohydrazide (L). Compound 1, 2, 3 and 4 was synthesized according to our previous work.43,44 Compound 4 (0.05 g, 0.15 mmol) was suspended in 30 mL absolutely ethyl alcohol and then 4-oxo-4H-chromene-3-carbaldehyde (0.040 g, 0.17 mmol) was added. The mixture was stirred at room temperature for 3 h. The resulting suspension was filtered, and the filter cake was washed with methanol (30 mL × 3) and got the pure compound L. Yield: 0.028 g, 42%. ESI-MS: m/z = 485.1 [M + H]. FTIR (KBr, cm−1): 1718 (C
O), 1230 (C–N). 1H NMR (400 MHz, CDCl3): 13.11 (s, 1H), 9.08–8.71 (m, 5H), 8.52 (d, 1H), 8.46–8.27 (m, 3H), 7.95 (dd, 2H), 7.82–7.76 (m, 1H), 7.69–7.51 (m, 3H). 13C CP/MAS NMR: δ 175.18, 160.35, 159.29, 154.06, 147.26, 138.72–116.94. Melting point: 190–193 °C. Element analysis for C29H16N4O4 (%): C 71.87, H 3.35, N 11.60, O 13.11. Calculated C 71.90, H 3.33, N 11.56, O 13.21.
Results and discussion
Synthesis of L
L was synthesized in moderate yield according to the synthetic route shown in Scheme 1. For the preparation of L, 4-oxo-4H-chromene-3-carbaldehyde was added to the ethanol solution of compound 4 (benzimidazo[2,1-a]benz[de]isoquinoline-7-one-12-carbohydrazide), and the mixture was stirred at 50 °C for 3 hours. The chemical structures of the synthesized compounds were characterized by 1H NMR, 13C CP/MAS NMR spectra, FTIR, element analysis and mass spectrum. All of the data in the spectra were in good accordance with the structure.
 |
| Scheme 1 Synthetic routes of L. Conditions: (i) CH2Cl2, at room temperature overnight; (ii) CHCl3, at room temperature for 1 h; (iii) dioxane, refluxed at 70 °C for 3 h; (iv) CH3CH2OH, at room temperature for 3 h. | |
The optical properties of L towards metal ions
The fluorescence behavior of L was investigated in acetonitrile–H2O (9
:
1) solution (0.01 M HEPES buffer at pH 7.4). As expected, L showed weak fluorescence intensity with a quantum yield (ϕ) of 0.029. As show in Fig. 1, upon addition of various metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ba2+, Ni2+, Cu2+, Zn2+, Cd2+, Hg2+, Pb2+, Ag+, Mn2+, Fe3+, Co2+, Cr3+, Sr3+ and Al3+), only Cu2+ and Fe3+ could induce obvious changes in the fluorescence intensity. After addition of Cu2+, the emission band of L showed remarkable fluorescence enhancement (ϕ = 0.182) centred at 490 nm, indicating a fluorescence turn-on response. What is different, after addition of Fe3+, the original weak fluorescence intensity of L showed a completely quenching response (ϕ = 0.002), displaying an efficient turn-off behavior. However, with addition of various other metal ions, almost negligible changes of fluorescence intensity were induced. As shown in Scheme 2, when 10 equivalents of Cu2+ or Fe3+ were added to the acetonitrile–H2O (9
:
1) HEPES buffer (pH 7.4) solution, L responded with a dramatic color change, from weak-fluorescence to strong-fluorescence for Cu2+, from weak-fluorescence to non-fluorescence for Fe3+, respectively. The apparent color changes could be distinguished by naked eye. Based on the fluorescence experiments, we found that L could be used as a fluorescent sensor for Cu2+ and Fe3+ in acetonitrile–H2O buffer solution.
 |
| Fig. 1 Fluorescence spectra of L (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) solution in the presence of various metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ba2+, Ni2+, Cu2+, Zn2+, Cd2+, Hg2+, Pb2+, Ag+, Mn2+, Fe3+, Co2+, Cr3+, Sr3+ and Al3+). Excitation wavelength is at 370 nm. | |
 |
| Scheme 2 The proposed binding mode of sensor L with Cu2+ and Fe3+. | |
The Cu2+ sensing
To get a further insight into the sensing property of L towards Cu2+, the fluorescence titration experiment was investigated in HEPES buffer solution at an excitation wavelength of 370 nm. As shown in Fig. 2, upon incremental addition of Cu2+, the fluorescence emission maximum of L at 490 nm gradually increased and reached a plateau when the concentration of Cu2+ was 10 equiv. Particularly, as shown in Fig. 3, the fluorescence intensity of L linearly increased as the concentration of Cu2+ changed from 0 to 10 equiv. By linearly fitting the changes of fluorescence as the function of concentration of Cu2+, we obtained the slope as 7.3 × 106 for L. The limit of detection (LOD) was found to be 1.3 × 10−8 mol L−1 based on LOD = 3σ/s where σ is the standard deviation of blank measurements, and s is the slope between fluorescence intensity versus Cu2+ concentration. The Job's method monitored by fluorescence intensities was applied to examine the stoichiometry of the L + Cu2+ complex, indicating a 1
:
1 stoichiometry of L to Cu2+ in the complex (Fig. 4). The titration results indicated that L had an excellent sensitivity towards Cu2+ based on the fluorescence turn-on response, and even was applicable for quantitative detection of Cu2+.
 |
| Fig. 2 Fluorescence spectra of L (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) upon titration with Cu2+ (1 × 10−5 M). Excitation is at 370 nm. | |
 |
| Fig. 3 Change ratio of L (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) upon titration with Cu2+ (1 × 10−5 M). Emission is monitored at 490 nm. | |
 |
| Fig. 4 Job's plot of the L in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) at 25 °C. The total concentration of L and Cu2+ was 0.05 mM. Excitation is at 370 nm, and emission is monitored at 490 nm. | |
As well all known, the selective sensing of sensor towards guest in the presence of other competitive species was another important property. Thus, to study the influence of other metal ions on the formation of L + Cu2+ complex, competitive experiments were performed with 20.0 equiv. of other metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ba2+, Ni2+, Zn2+, Cd2+, Hg2+, Pb2+, Ag+, Mn2+, Fe3+, Co2+, Cr3+, Sr3+ and Al3+) in the presence of 10.0 equiv. of Cu2+. As shown in Fig. 5, it was found that the fluorescence enhancement caused by the mixture of Cu2+ with most of other metal ions was similar to that caused by Cu2+ only. However, in the presence of Cu2+ mixed with Fe3+, a completely quenched fluorescence of L was observed, indicating there was interference from Fe3+ to the detection of Cu2+. The competitive experiments verified that the sensing ability of L towards Cu2+ was not affected by the presence of most other metal ions, and the new formation of lighted complex L + Cu2+ could be further used as a fluorescence quenching sensor for Fe3+.
 |
| Fig. 5 The fluorescence intensity of L (1 × 10−5 M) in the presence of 10 equiv. of Cu2+ and followed by 20 equiv. of various other metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ba2+, Ni2+, Zn2+, Cd2+, Hg2+, Pb2+, Ag+, Mn2+, Fe3+, Co2+, Cr3+, Sr3+ and Al3+) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) at 25 °C. Excitation is at 370 nm, and emission is monitored at 490 nm. | |
The Fe3+ sensing
As shown in Fig. 6, the fluorescence titration experiment of L towards Fe3+ was investigated in HEPES buffer solution at an excitation wavelength of 370 nm. Upon incremental addition of Fe3+, the weak fluorescence intensity at 490 nm gradually decreased. When the concentration of Fe3+ reached 5 equiv., the fluorescence intensity quenched to the end. At this state, the emission spectra of L was almost a flat line parallel to the abscissa (ϕ = 0.002). Particularly, as shown in Fig. 7, upon the concentration of Fe3+ increased from 0 to 3.5 equiv., a good linear relationship was observed between fluorescence intensity and [Fe3+]. By linearly fitting the changes of fluorescence as the function of concentration of Fe3+, the LOD for Fe3+ was found to be 4.3 × 10−7 mol L−1 based on LOD = 3σ/s, where σ is the standard deviation of blank measurements, and s is the slope between fluorescence intensity versus Fe3+ concentration. The titration results indicated that L had an excellent sensitivity towards Fe3+ based on the fluorescence turn-off response.
 |
| Fig. 6 Fluorescence spectra L (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) upon titration with Fe3+ (1 × 10−5 M). Excitation is at 370 nm. | |
 |
| Fig. 7 Change ratio of L (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) upon titration with Fe3+ (1 × 10−5 M). Emission is monitored at 490 nm. | |
As shown in Fig. 8, to study the influence of other metal ions on the fluorescence detection of Fe3+, competitive experiments were performed with 10.0 equiv. of other metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ba2+, Ni2+, Zn2+, Cd2+, Hg2+, Pb2+, Ag+, Mn2+, Fe3+, Co2+, Cr3+, Sr3+ and Al3+) in the presence of 5.0 equiv. of Fe3+. It was found that the fluorescence quenching caused by the mixture of Fe3+ with other metal ions was very similar to that caused by Fe3+ only. The competitive experiments verified that the sensing ability of L towards Fe3+ was not affected by the presence of other metal ions.
 |
| Fig. 8 The fluorescence intensity of L (1 × 10−5 M) in the presence of 5 equiv. of Fe3+ and followed by 10 equiv. of various other metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ba2+, Ni2+, Zn2+, Cu2+, Cd2+, Hg2+, Pb2+, Ag+, Mn2+, Co2+, Cr3+, Sr3+ and Al3+) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) at 25 °C. Excitation is at 370 nm, and emission is monitored at 490 nm. | |
As mentioned above, the formed complex L + Cu2+ could be further used as a new platform for ultrasensitive detection of Fe3+. To investigate the sensing property of L + Cu2+ towards Fe3+ in detail, the fluorescence titration experiment was carried out. As shown in Fig. 9, in acetonitrile–H2O (9
:
1, HEPES 0.01 M, pH = 7.4) solution, the complex L + Cu2+ showed a strong fluorescence intensity centered at 490 nm (ϕ = 0.182). Upon incremental addition of Fe3+, the strong fluorescence intensity gradually decreased and completely quenched when the concentration of Fe3+ reached 15 equiv. As shown in Fig. 10, upon the concentration of Fe3+ increased from 0 to 12 equiv., the fluorescence intensity of L + Cu2+ linearly decreased. By linearly fitting the changes of fluorescence as the function of concentration of Fe3+, we obtained the slope as 2.3 × 106 for L + Cu2+. The limit of detection (LOD) was found to be 2.4 × 10−8 mol L−1 based on LOD = 3σ/s. where σ is the standard deviation of blank measurements, and s is the slope between fluorescence intensity versus Fe3+ concentration. The titration results indicated that the complex L + Cu2+ had an excellent sensitivity towards Fe3+ based on the fluorescence turn-off response.
 |
| Fig. 9 Fluorescence spectra of L + Cu2+ (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) upon titration with Fe3+ (1 × 10−5 M). Excitation is at 370 nm. | |
 |
| Fig. 10 Change ratio of L + Cu2+ (1 × 10−5 M) in acetonitrile–H2O (9 : 1, HEPES 0.01 M, pH = 7.4) upon titration with Fe3+ (1 × 10−5 M). Emission is monitored at 490 nm. | |
The mechanism of binding mode
Based on the results of fluorescence titration experiments, we proposed two plausible binding modes for L + Cu2+ and L + Fe3+ as shown in Scheme 2. The weak fluorescence of L was likely due to multiple quenching mechanisms in the molecule (C
N isomerization and PET process). With the introduction of 4-oxo-4H-chromene-3-carbaldehyde as a receptor, the fluorophore and receptor were linked by a C
N bond. C
N isomerization was always thought as the predominant decay process of the excited states for compounds with an unbridged C
N structure, so that those compounds were often weak-fluorescent.45 Meanwhile, the free 4-oxo-4H-chromene-3-carbaldehyde group have a lone pair of suitable energy for causing sufficient PET effect in the molecule directed from receptor to fluorophore, also resulting in a weak fluorescence of L. After binding with Cu2+, the C
N isomerization was inhabited and the stiffness of the molecular became better. In addition, the lone pair of receptor was occupied and the electron transfer was thermodynamically disallowed, resulting in the suppression of PET communication between receptor and fluorophore. Although Cu2+ was a paramagnetic cation which could always induce fluorescence quenching response, the inhibition of both C
N isomerization and PET processes played a dominant role and override the paramagnetic quenching property in the sensing process of L towards Cu2+, resulting in a strong fluorescence enhancement of the complex L + Cu2+.
What is different, Fe3+ was also paramagnetic with an unfilled d shell and could strongly quench the emission of fluorophore near it through electron and/or energy transfer processes. Upon binding with Fe3+, the weak fluorescence of L was completely quenched which could be ascribed to the ligand–metal charge transfer mechanism (LMCT).46,47 In the complex L + Fe3+, the paramagnetism and unfilled d shell of Fe3+ could prompt the fluorophore to open a nonradiative deactivation channel and facilitate the transfer of electron and/or energy, resulting in the fluorescence quenching response of L.48–50 Although the C
N isomerization was probably also inhibited to some extent during the binding mode, the LMCT could be firstly and quickly occurred and the strength of paramagnetic quenching property of Fe3+ was much stronger that could cover the other mechanisms. Similarly, for the lighted complex L + Cu2+, upon addition of Fe3+, the Fe3+ could instantly displace Cu2+ from L + Cu2+, resulting in the formation of a new complex L + Fe3+. It was suggested that in this case, the paramagnetic quenching property of Fe3+ played the leading role, so L + Fe3+ showed a completely quenched fluorescence. Because of the rapidly response and rational control of multiple mechanisms, the lightened complex L + Cu2+ had a limit of detection for Fe3+ as low as 2.4 × 10−8 mol L−1.
To understand the PET mechanism underlying the experimentally observed photophysical characteristics of L upon complexation with Cu2+, we had investigated the structural, electronic, and optical properties using ab initio density functional theory (DFT) combined with time-dependent density functional theory (TDDFT) calculations as implemented in the Gaussian 09 package. We had adopted the hybrid B3LYP exchange and correlation functional using an effective core potential with the LANL2DZ basis set for transition-metal Cu and the 6-31+G(d) basis set for all the other elements in the calculations. As shown in Fig. 11, in L, the electron densities of the highest occupied molecular orbital (HOMO) were distributed over the whole fluorophore moiety, while those of the lowest unoccupied molecular orbital (LUMO) transferred to carbonyl group and receptor moiety. Upon excitation of L, an electron would be transferred from the 4-oxo-4H-chromene-3-carbaldehyde unit to the fluorophore, resulting in the weak fluorescence, which was corresponded with the PET process. In contrast, after binding with Cu2+, the orbitals were localized on fluorophore for both HOMO and LUMO. Upon excitation, the PET process was inhibited and there was no electron transfer between the receptor and fluorophore, leading to the strong fluorescence of complex L + Cu2+.
 |
| Fig. 11 PET mechanism: HOMO and LUMO of L and L + Cu2+, calculated with DFT/TDDFT at the B3LYP/6-31G(d) level using Gaussian 09. | |
Application in water samples
In order to investigate the practicability of L, the amounts of Cu2+ and Fe3+ in tap water were determined by our proposed fluorescence assay method. As demonstrated in Table 1, Cu2+ added to water samples could be accurately measured with good recovery (93.1–104.3%) and the relative standard deviation (RSD) of five measurements was less than 1.6%, indicating that probe L has potential application for quantitative detection of Cu2+ in real water samples. Furthermore, as shown in Table 2, the concentration of Fe3+ added to the water samples could also be accurately determined (92.0–108.0%), and the data revealed a good agreement between the added and the found concentrations of Fe3+ (RSD < 1.8%), indicating that the complex of L + Cu2+ could be used for quantitative detection of Fe3+ in real water samples.
Table 1 Determination of the Cu2+ concentration in water samples
Sample |
Cu2+ added (mol L−1) |
Cu2+ recovered (mol L−1) |
Recovery (%) |
RSD (%) |
1 |
1.0 × 10−5 |
0.9 × 10−5 |
93.1 |
0.84 |
2 |
3.0 × 10−5 |
3.1 × 10−5 |
104.3 |
0.83 |
3 |
5.0 × 10−5 |
4.8 × 10−5 |
96.8 |
1.55 |
4 |
7.0 × 10−5 |
7.0 × 10−5 |
100.4 |
0.39 |
5 |
10.0 × 10−5 |
10.1 × 10−5 |
101.1 |
0.98 |
Table 2 Determination of the Fe3+ concentration in water samples
Sample |
Fe3+ added (mol L−1) |
Fe3+ recovered (mol L−1) |
Recovery (%) |
RSD (%) |
1 |
1.0 × 10−5 |
0.9 × 10−5 |
92.0 |
0.76 |
2 |
3.0 × 10−5 |
3.2 × 10−5 |
108.0 |
1.22 |
3 |
5.0 × 10−5 |
5.0 × 10−5 |
101.6 |
0.89 |
4 |
9.0 × 10−5 |
8.9 × 10−5 |
99.4 |
1.71 |
Conclusions
In summary, an ideal fluorescent sensor L was designed and synthesized based on benzimidazo[2,1-a]benz[de]isoquinoline-7-one-12-carbohydrazide. In acetonitrile–H2O (9
:
1) solution (0.01 M HEPES buffer at pH 7.4), L showed great selectivity and sensitivity towards Cu2+ with fluorescence turn-on response based on the inhibition of C
N isomerization and PET process. Furthermore, in the same condition, L and the lighted complex L + Cu2+ could be used as new sensors for ultrasensitive detection of Fe3+ based on the paramagnetic quenching process with a limit of detection down to 2.4 × 10−8 mol L−1. The sensing mechanisms were demonstrated by DFT and TDDFT calculations by Gaussian 09 package. Because of the rational control of multiple mechanisms, L could be used as excellent selective and sensitive sensor towards Cu2+ and Fe3+, especially for the dual detection of Fe3+. In addition, L was successfully applied to determine the accurate amounts of Cu2+ and Fe3+ in tap water. Series of similar sensors based on this sensing strategy are preparing in our lab, and our further efforts will be focused on the investigation of optical properties and practical applications of sensors.
Acknowledgements
The authors thank Henan Sanmenxia Aoke Chemical Industry Co., Ltd. (w0920) for financial support and thank Prof. Guangyou Zhang for the experimental assistance.
Notes and references
- B. Valeur and I. Leray, Coord. Chem. Rev., 2000, 205, 3–40 CrossRef CAS.
- S. Dey, A. Roy, G. P. Maiti, S. K. Mandal, P. Banerjee and P. Roy, New J. Chem., 2016, 40, 1365–1376 RSC.
- K. Boonkitpatarakul, J. Wang, N. Niamnont, B. Liu, L. Mcdonald, Y. Pang and M. Sukwattanasinitt, ACS Sens., 2016, 1, 144–150 CrossRef CAS.
- K. P. Carter, A. M. Young and A. E. Palmer, Chem. Rev., 2014, 114, 4564–4601 CrossRef CAS PubMed.
- B. Halliwell and J. M. Gutteridge, Biochem. J., 1984, 219, 1 CrossRef CAS PubMed.
- S. Hu, P. Furst and D. Hamer, New Biol., 1990, 2, 544 CAS.
- Handbook of Metalloproteins, ed. A. Messerschmidt, R. Huber, T. Poulos and K. Weighardt, John Wiley & Sons, Inc., New York, 2001, vol. 2, p. 1149 Search PubMed.
- H. Tapiero, D. M. Townsend and K. D. Tew, Biomed. Pharmacother., 2003, 57, 386 CrossRef CAS.
- L. I. Bruijn, T. M. Miller and D. W. Cleveland, Annu. Rev. Neurosci., 2004, 27, 723 CrossRef CAS PubMed.
- K. J. Barnham, C. L. Masters and A. I. Bush, Nat. Rev. Drug Discovery, 2004, 3, 205 CrossRef CAS PubMed.
- D. R. Brown and H. Kozlowski, Dalton Trans., 2004, 1907 RSC.
- K. C. Ko, J. S. Wu, H. J. Kim, P. S. Kwon, J. W. Kim, R. A. Bartsch, J. Y. Lee and J. S. Kim, Chem. Commun., 2011, 47, 3165 RSC.
- A. K. Jain, V. K. Gupta, L. P. Singh and J. R. Raisoni, Talanta, 2005, 66, 1355 CrossRef CAS PubMed.
- A. Mokhir, A. Kiel, D. P. Herten and R. Kraemer, Inorg. Chem., 2005, 44, 5661 CrossRef CAS PubMed.
- J. Liu and Y. Lu, J. Am. Chem. Soc., 2007, 129, 9838 CrossRef CAS PubMed.
- P. G. Georgopoulos, A. Roy, M. J. Yonone-Lioy, R. E. Opiekun and P. J. Lioy, J. Toxicol. Environ. Health, Part B, 2001, 4, 341–394 CAS.
- R. V. Rathod, S. Bera, M. Singh and D. Mondal, RSC Adv., 2016, 6, 34608–34615 RSC.
- A. Torrado, G. K. Walkup and B. Imperiali, J. Am. Chem. Soc., 1998, 120, 609 CrossRef CAS.
- P. Grandini, F. Mancin, P. Tecilla, P. Scrimin and U. Tonellato, Angew. Chem., Int. Ed., 1999, 38, 3061 CrossRef CAS PubMed.
- G. Klein, D. Kaufmann, S. SchVrch and J. L. Reymond, Chem. Commun., 2001, 561–562 RSC.
- M. Boiocchi, L. Fabbrizzi, M. Licchelli, D. Sacchi, M. Vázquez and C. Zampa, Chem. Commun., 2003, 1812 RSC.
- Y. Zheng, J. Orbulescu, X. Ji, F. M. Andreopoulos, S. M. Pham and R. M. Leblanc, J. Am. Chem. Soc., 2003, 125, 2680 CrossRef CAS PubMed.
- C. Roy, B. Chandra, D. Hromas and S. Mallik, Org. Lett., 2003, 5, 11 CrossRef PubMed.
- S. H. Kim, J. S. Kim, S. M. Park and S. K. Chang, Org. Lett., 2006, 8, 371 CrossRef CAS PubMed.
- Q. Y. Wu and E. V. Anslyn, J. Am. Chem. Soc., 2004, 126, 14682 CrossRef CAS PubMed.
- J. V. Mello and N. S. Finney, J. Am. Chem. Soc., 2005, 127, 10124 CrossRef PubMed.
- E. W. Miller, A. E. Albers, A. Pralle, E. Y. Isaco and C. J. Chang, J. Am. Chem. Soc., 2005, 127, 16652 CrossRef CAS PubMed.
- D. Y. Sasaki, D. R. Shnek, D. W. Pack and F. H. Arnold, Angew. Chem., Int. Ed., 1995, 34, 905 CrossRef CAS.
- R. Kramer, Angew. Chem., Int. Ed., 1998, 37, 772 CrossRef CAS.
- M. Royzen, Z. H. Dai and J. W. Canary, J. Am. Chem. Soc., 2005, 127, 1612 CrossRef CAS PubMed.
- B. D'Autreaux, N. P. Tucker, R. Dixon and S. Spiro, Nature, 2005, 437, 769–772 CrossRef PubMed.
- J. W. Lee and J. D. Helmann, Nature, 2006, 440, 363–367 CrossRef CAS PubMed.
- H. Weizman, O. Ardon, B. Mester, J. Libman, O. Dwir, Y. Hadar, Y. Chen and A. Shanzer, J. Am. Chem. Soc., 1996, 118, 12368–12375 CrossRef CAS.
- J. P. Sumner and R. Kopelman, Analyst, 2005, 130, 528–533 RSC.
- R. M. Roat Malone, Bioinorganic Chemistry-A Short Course, John Wiley and Sons, New Jersey, NJ, 2007 Search PubMed.
- I. Bertini, H. B. Gray, E. I. Stiefel and J. S. Valentine, J. Biol. Inorg. Chem., 2007, 12, 1197–1206 CrossRef PubMed.
- F. Ge, H. Ye, H. Zhang and B. X. Zhao, Dyes Pigm., 2013, 99, 661–665 CrossRef CAS.
- E. D. Weinberg, Science, 1974, 184, 952–956 CrossRef CAS PubMed.
- S. Sidhu, K. Kumari and M. Uppal, J. Hum. Ecol., 2007, 21, 265–267 CrossRef.
- A. R. Nissenson and J. Strobos, Kidney Int., 1999, 55, S18–S21 CrossRef.
- R. K. Schindhelm, M. Schoorl and J. V. Pelt, Diabetes Care, 2012, 35, 797–802 CrossRef PubMed.
- M. J. Frisch, et al., Gaussian 09 (Revision D.01), Gaussian Inc., Pittsburg, PA, 2009 Search PubMed.
- Z. Liu, Y. Qi, C. Guo, Y. Zhao, X. Yang, M. Pei and G. Zhang, RSC Adv., 2014, 4, 56863–56869 RSC.
- Z. Liu, C. Peng, Y. Wang, M. Pei and G. Zhang, Org. Biomol. Chem., 2016, 14, 4260–4266 Search PubMed.
- J. S. Wu, W. M. Liu, X. Q. Zhuang, F. Wang, P. F. Wang, S. L. Tao, X. H. Zhang, S. K. Wu and S. T. Lee, Org. Lett., 2007, 9, 33–36 CrossRef CAS PubMed.
- Y. Ma, W. Luo, P. J. Quinn, Z. Liu and R. C. Hider, J. Med. Chem., 2004, 47, 6349–6362 CrossRef CAS PubMed.
- R. Nudelman, O. Ardon, Y. Hadar, Y. Chen, J. Libman and A. Shanzer, J. Med. Chem., 1998, 41, 1671–1678 CrossRef CAS PubMed.
- L. Yang, W. Yang, D. Xu, Z. Zhang and A. Liu, Dyes Pigm., 2013, 97, 168–174 CrossRef CAS.
- J. H. Xu, Y. M. Hou, Q. J. Ma, X. F. Wu and X. J. Wei, Spectrochim. Acta, Part A, 2013, 112, 116–124 CrossRef CAS PubMed.
- Z. X. Li, L. F. Zhang, W. Y. Zhao, X. Y. Li, Y. K. Guo, M. M. Yu and J. X. Liu, Inorg. Chem. Commun., 2011, 14, 1656–1658 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: NMR, MS spectra and calculation details. See DOI: 10.1039/c6ra09535f |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.