Facile rhodamine-based colorimetric sensors for sequential detections of Cu(II) ions and pyrophosphate (P2O74−) anions

Reguram Arumugaperumala, Venkatesan Srinivasadesikanb, Ming-Chang Linb, Muthaiah Shellaiaha, Tarun Shuklaa and Hong-Cheu Lin*a
aDepartment of Materials Science and Engineering, National Chiao Tung University, Hsinchu 300, Taiwan. E-mail: linhc@mail.nctu.edu.tw
bCenter for Interdisciplinary Molecular Science, Department of Applied Chemistry, National Chiao Tung University, Hsinchu 300, Taiwan

Received 1st October 2016 , Accepted 31st October 2016

First published on 31st October 2016


Abstract

Two rhodamine hydrazine derivatives Rh1 and Rh2 with catechol and ether functionalities have been synthesized and utilized towards sequential colorimetric detections of Cu(II) and pyrophosphate (PPi) ions in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) semi-aqueous medium. Notably, Rh1 and Rh2 are the first example of colorimetric rhodamine-based probes for the sequential detections of Cu2+ ion and PPi anion. Based on the significant colorimetric responses (from colorless to pink) of Rh1 and Rh2 to Cu2+ ions, the detection limits were estimated as low as 1.22 × 10−8 M and 8.0 × 10−7 M, respectively, which signified the utilities of designed probes towards facile detections of Cu2+ ions. Furthermore, upon the successive addition of PPi ion to Rh1–Cu2+ and Rh2–Cu2+ complexes, it has been altered and restored to its origin of Rh1 and Rh2 via re-coordination of PPi to Cu2+ ion, which was confirmed by color changes from pink to colorless. Moreover, computational DFT calculations provided more insights into HOMO–LUMO structures of rhodamine derivatives and their copper complexes. Additionally, the solid state strip-based colorimetric detections of Cu2+ ion were supplemented as a real time application.


1 Introduction

Recently, owing to their environmental and biological significance, new artificial materials have emerged as efficient chemosensors towards the detection and estimation of transition metal ions.1–3 Among them, copper ions were found to be essential and their detection has been paid greater attention over other metal ions.4–7 In this regard, divalent copper ions are more fascinating not only as an environmental pollutant at higher concentrations but also as the third most abundant divalent metal ion in the human body.8–12

Furthermore, copper has been suspected to cause liver and kidney damages for infants in long-term exposure. Similarly, short-term exposures to high levels of copper ion cause gastrointestinal disturbance.13,14 Henceforth, the U.S. Environmental Protection Agency (EPA) has set a limit of copper ions in drinking water as 1.3 ppm (ca. 20 mM).15 Additionally, the average concentration of blood copper in the normal body is fixed as 100–150 mg dL−1 (15.7–23.6 mM).16 However, copper can be toxic over its normal concentration that cause oxidative stresses and disorders associated with neurodegenerative diseases, including Meknes, Wilson diseases, familial amyotrophic lateral sclerosis, Alzheimer's disease and prion diseases. Therefore, the colorimetric visual detection of trace amounts of Cu2+ ions via simplified instrumental set up are become vital.17–21 In this concept, the high sensitivity and selectivity are considered as a fundamental goal for organic and analytical chemists all over the world. In order to achieve the selective detection of Cu2+ against the background of competing analytes, many promising small molecules have been reported as Cu2+ sensor through colorimetric response.22

On the other hand, phosphate-based inorganic and organic molecules play critical roles in both environment and biological systems. Among phosphate anions, pyrophosphate (PPi) is known to be biologically significant one,23 which has been obtained as a product, during the hydrolysis of adenosine triphosphate (ATP) to adenosine monophosphate under cellular condition.24 However, abnormal levels of PPi can lead to several diseases such as chondrocalcinosis or calcium pyrophosphate dihydrate (CPPD) crystal deposition problem and arthritis.25 Recently, several approaches such as electro optical methods, colorimetric and fluorescence assays have been developed for the detection of PPi ions.26–28 Among those techniques, colorimetric chemosensors are most attractive for the detection of PPi due to their selectivity, sensitivity, response time and low-cost.29 Moreover, because of the strong binding affinities between metal ions and PPi, the utilization of metal ion complexes as a colorimetric chemosensor has been found to be the successful strategy for the detection of PPi ions. However, chemosensors based on color changes (by naked eye observation) for detection of PPi are found to be limited.30 Hence, we tend to develop such colorimetric sensors for the detection of PPi ions.

On the consideration of photophysical properties, such as high quantum yield, photo stability, absorption and emission in the visible region, rhodamine fluorophore is seems to be most attractive one.31–34 Further, the sensing behavior of rhodamine towards metal ions is also very interesting. Typically, rhodamine derivatives are non-fluorescent and colorless in their ring-closed spiro lactam structures, whereas the spiro lactams induced ring-opening by specific metal – give rise to strong fluorescence emissions to develop for colorimetric sensors.35–38 Nevertheless, many rhodamine-based florescent and colorimetric chemosensors for detecting metal ions have been studied.39 Colorimetric sensors are more facile due to the detection by naked eye that can be commonly used in low cost, less equipment and easy handling.40–45 There are perceptible color changes in rhodamine derivatives, even at very small amounts of analyte concentrations. Therefore, much effort has been focused on the development of rhodamine-based sensor probes for the detections of heavy metal and anions.46 On the other hand, direct detections of Cu(II) and PPi ions could be accomplished by regulating the acceptors of rhodamine sensor molecules to affect ICT mechanisms, which may evidence via ratiometric responses and ion induce naked-eye color changes.47 Hence, rhodamine-based chemosensory probes for the sequential recognitions of Cu(II) and PPi ions remained highly demands. Yoon and co-workers previously demonstrated rhodamine derivative as a chemosensor for sequential detections of Al3+ and pyrophosphate.48 However, to the best of our knowledge, rhodamine-based selective and sensitive sequential detections of Cu2+ and pyrophosphate via the ratiometric absorption spectral changes have never been explored yet.

Herein, we have successfully developed two reversible colorimetric chemosensor probes Rh1 and Rh2 for sequential detections of Cu2+ and pyrophosphate (PPi) ions. Dimers Rh1 and Rh2 (Fig. 1) contain two symmetrical rhodamine units (on both ends) as the fluorophore and two central cores of dihydroxyphenyl and multiple ethers groups as the chelating units, respectively. The receptor moieties together with the nitrogen and carbonyl of rhodamine serve as ionophores by providing coordinating sites based on spirocyclic ring-opening mechanisms. Additionally, the optical properties of these probes and their possible mechanisms for recognition of Cu2+ ion has been demonstrated by UV-vis titrations. Further, B3LYP/LANL2DZ density functional calculations were carried out to investigate the molecular orbitals and electronic excitations of these two probes. The stabilities of two probes were examined in the absence and presence of Cu2+ ion at wide ranges of pH values. Interestingly, both metal complexes (Rh1–Cu2+ and Rh2–Cu2+) have been utilized in excellent selective detections of PPi via colorimetric changes visualized by naked-eyes.


image file: c6ra24472f-f1.tif
Fig. 1 Chemical structures of Rh1 and Rh2.

2 Experimental

2.1 Materials and instrumentations

All reagents were commercially purchased and were used without further purification. The solvents were dried by distillation over appropriate drying agents. Generally, reactions were monitored by TLC plates and flash chromatography was performed on silica gel. The 1H and 13C NMR spectra were recorded on a 300 MHz spectrometer and samples were dissolved in CDCl3 and DMSO-d6. The chemical shifts were expressed in ppm and coupling constants (J) in Hz. Mass spectra (FAB) were obtained on the respective mass spectrometer. Elemental analysis was carried out by Elemental Varo EL. UV-vis absorption and fluorescence spectra were measured on V-670 spectrophotometer and F-4500 fluorescence spectrophotometer, respectively. The pH [1–12] buffer solutions were freshly prepared.

2.2 Preparation of metal ion solutions for sensor titrations

Rh1 and Rh2 were dissolved in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) at 1 × 10−5 M concentration. The solutions of metal cations Li+, K+, Cs+, Ni2+, Fe3+, Fe2+, Co2+, Zn2+, Cd2+, Pb2+, Ca2+, Cr3+, Cu2+, Ba2+ and Al3+ were dissolved in water medium at 1 × 10−3 M concentration from their chloride salts. Ag+, Hg2+ and Mg2+ were made from AgNO3, Hg(OAc)2 and MgSO4, respectively, in water medium at 1 × 10−3 M concentration. The solutions of all anions were prepared by dissolving the respective TBA salts of F, CN, HPO42−, H2PO4, HSO4, ClO4, Br, OH, SCN, P2O74−, NO3 and OAc in deionized water (1 × 10−3 M). Pentamethyl diethylene triamine (PMDTA) was dissolved in CH3CN at a concentration of 1 × 10−3 M.

2.3 NMR titrations

10 mmol (1 equiv.) of Rh1 and Rh2 in DMSO-d6 were titrated with 20 mmol (2 equiv.) of Cu2+ in D2O.

2.4 Computational methods

Density functional theory (DFT) calculations were employed to elucidate the Cu2+ interactions with Rh1 and Rh2 systems. All computations were carried out using Gaussian 09 software package.49 Geometry optimization of the ground state structures was carried out with DFT at the B3LYP level of theory using LANL2DZ basis set in gas phase.50 The excitation energies (up to 5 states) of the low-lying excited states and oscillator strengths were predicted using the time-dependent density functional theory (TD-DFT) with B3LYP/LANL2DZ level.

2.5 Synthetic procedures

The following compounds (1–3) were prepared according to the literature: 2,3-dihydroxyterephthalaldehyde (1),51 [2,2′-(oxybis(ethane-2,1diyl))bis(oxy)dibenzaldehyde] (2)52 and rhodamine hydrazide (3).53 The detailed synthetic procedures are illustrated in the ESI and Scheme S1.

2.6 Synthesis of compound Rh1

Rhodamine hydrazide 3 (0.6 g, 1.32 mmol) was dissolved in absolute ethanol (25 mL), then 3,6-diformylcatechol (1) (0.1 g, 0.66 mmol) was added and the reaction mixture was refluxed in an oil bath for 12 h. The solvent was evaporated by rotary evaporator and the crude product was purified by silica gel column chromatography using EA[thin space (1/6-em)]:[thin space (1/6-em)]hexane (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) to obtain compound Rh1 as a light pink solid. Yield: 0.32 g (50%). 1H NMR (CDCl3) δ: 10.62 (s, 2H), 9.16 (s, 2H), 7.95–7.93 (m, 2H), 7.51–7.45 (m, 4H), 7.12–7.10 (m, 2H), 6.57 (s, 2H), 6.47 (d, 4H, J = 3.0 Hz), 6.41 (d, 4H, J = 2.7 Hz), 6.21 (dd, 4H, J = 8.7 Hz and 2.4 Hz), 3.30 (q, 16H, J = 6.9 Hz), 1.12 (t, 24H, J = 6.9 Hz); 13C NMR (CDCl3) δ: 164.3, 153.2, 152.5, 151.5, 148.9, 146.9, 133.9, 129.2, 128.3, 127.8, 124.0, 123.2, 120.8, 119.5, 108.0, 105.0, 98.1, 66.4, 44.3, 12.5; IR (KBr, cm−1): 3425, 3079, 1706; MS (FAB+) m/z: calcd for C64H66N8O6: 1042.5; found: 1143.7 (M + H)+; anal. calcd for C64H66N8O6: C, 73.68; H, 6.38; N, 10.74; found: C, 72.66; H, 6.60; N, 10.75.

2.7 Synthesis of compound Rh2

Rhodamine hydrazide 3 (0.5 g, 1.26 mmol) was dissolved in absolute ethanol (25 mL), then 2,2′-((oxybis(ethane-2,1-diyl))bis(oxy))dibenzaldehyde (2) (0.2 g, 0.63 mmol) was added and the reaction mixture was refluxed in an oil bath for 12 h. The solvent was evaporated by rotary evaporator and the crude product was purified by silica gel column chromatography using EA[thin space (1/6-em)]:[thin space (1/6-em)]hexane (6[thin space (1/6-em)]:[thin space (1/6-em)]4, v/v) to obtained compound Rh2 as a light pink solid. Yield: 0.310 g (40%). 1H NMR (CDCl3) δ: 9.03 (s, 2H), 7.97 (dd, 2H, J = 1.8 Hz and 5.4 Hz), 7.85–7.83 (m, 2H), 7.48–7.44 (m, 4H), 7.17–7.09 (m, 4H), 6.84 (t, 2H, J = 7.5 Hz), 6.76 (d, 2H, J = 8.4 Hz), 6.53 (d, 4H, J = 9.0 Hz), 6.43 (d, 4H, J = 2.4), 6.22 (dd, 4H, J = 8.7 Hz and 2.4 Hz), 4.01 (t, 4H, J = 5.1 Hz), 3.82 (t, 4H, J = 5.1 Hz), 3.28 (q, 16H, J = 6.9 Hz), 1.11 (t, 24H, J = 6.9 Hz); 13C NMR (CDCl3) δ: 164.9, 157.3, 153.6, 153.1, 149.8, 143.6, 133.1, 130.8, 129.3, 128.1, 127.4, 126.3, 124.4, 123.7, 121.0, 112.9, 107.9, 106.2, 97.9, 69.7, 66.2, 65.9, 44.30, 12.6; IR (KBr, cm−1): 3460, 3065, 1692; MS (FAB+) m/z: calcd for C74H78N8O7: 1191.6; found: 1192.4 (M + H)+; 1214.5 (M + Na)+; anal. calcd for C74H78N8O7: C, 74.60; H, 6.60; N, 9.40; found C, 74.36; H, 6.97; N, 9.61.

3 Results and discussion

3.1 Colorimetric responses of sensor probes (Rh1 and Rh2) towards metal ions

Selectivity is one of the most important characteristics in analyte-selective chemosensors. Initially, Rh1 and Rh2 (10 μM) in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) displayed no absorption bands corresponding to the rhodamine moiety. As shown in Fig. 2a and b, upon examine the selectivity of Rh1 and Rh2 towards various metal ions (4 equiv.), such as Li+, Ag+, K+, Cs+, Ni2+, Fe3+, Co2+, Zn2+, Cd2+, Pb2+, Ca2+, Cr3+, Cu2+, Ba2+, Al3+, Fe2+, Hg2+ and Mg2+, the UV-vis spectra of Rh1 and Rh2 were only affected by Cu2+, Fe3+ and Cr3+ ions to different extents. Among these three metal ions, upon the addition of Cu2+ both absorption bands of Rh1 and Rh2 had the largest changes with the absorption enhancements as large as 40 and 48 fold, respectively (Fig. S13a and b), where the detailed explanation will be discussed later. Sequentially, the maximum UV-vis absorption intensities of Rh1 and Rh2 at 561 nm were also increased upon the addition of Fe3+ (13 and 20 folds, respectively) or Cr3+ (8 and 10 folds, respectively) ions. Whereas, a significant increase in the absorption intensity at 561 nm may be attributed to the induced spirolactam ring opening upon the addition of metal ions. The photo images exhibit a naked eye colorimetric change of Rh1 towards Cu2+ metal ion, where the most obvious and selective color change from colorless to pink color is observed in Fig. 3. Therefore, our careful investigations suggest that Rh1 and Rh2 could be useful as selective optical chemosensor towards Cu2+ than the other surveyed metal ions.
image file: c6ra24472f-f2.tif
Fig. 2 UV-visible absorption spectra (a) Rh1 and (b) Rh2 (10 μM) upon the addition of several relevant metal ions in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4).

image file: c6ra24472f-f3.tif
Fig. 3 Colorimetric changes (naked eye observations) of probe Rh1 (1 μM) upon the addition of various metal ions (4 equiv.).

Furthermore, to explore the possible applications of Rh1 and Rh2 as practical cation-selective chromogenic chemosensors for Cu2+, two competitive experiments were also carried out by monitoring the changes in UV-vis absorption intensities at 555 nm. To perform these dual metal ion experiments, Rh1 and Rh2 (10 μM) were first mixed with 4 equiv. of various metal ions including Al3+, Cr3+, Fe3+, Hg2+, Pb2+, Ag2+, Zn2+, Mg2+, Ni2+ and Ag+ ions, followed by the addition of 2 equiv. Cu2+ in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) solution. In the presence of miscellaneous competitive cations, the observations of UV-vis spectra for Rh1 and Rh2 did not show any changes in contrast to single Cu2+ ion experiments. However, in contrast to Cu2+ ion, only the cases of Fe3+ and Cr3+ ions were evidenced to affect the absorption peaks to some extents, as illustrated in Fig. 4a and b, where the significant spectral and color changes in dual metal ion systems with Cu2+ were still observed. Therefore, both designed dimeric probes (Rh1 and Rh2) may be promising towards establishing selective colorimetric sensors for Cu2+ ion over the other competing metal ions.


image file: c6ra24472f-f4.tif
Fig. 4 (a) Absorbance changes of Rh1 (10 μM) in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) solution upon the addition of 4 equiv. various metal ions (black bars). The grey bars represent the absorbance changes as 2.4 equiv. Cu2+ ion was added to the mixture of Rh1 (10 μM) 4 equiv. Al3+, Cr3+, Fe3+, Hg2+, Pb2+, Ag2+, Zn2+, Mg2+, Ni2+ and Ag+ ions, respectively. (b) Absorbance changes of Rh2 (10 μM) in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) solution upon the addition of 4 equiv. various metal ions (black bars). The grey bars represent the absorbance changes as 2.2 equiv. Cu2+ ion was added to the mixture of Rh2 (10 μM) 4 equiv. Al3+, Cr3+, Fe3+, Hg2+, Pb2+, Ag2+, Zn2+, Mg2+, Ni2+ and Ag+ ions, respectively.

The spectroscopic properties of Rh1 and Rh2 (10 μM) were investigated in a solvent system of in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4). As shown in Fig. 5a, the free chemosensor Rh1 exhibited an absorption band centered at 323 nm. Upon the addition of Cu2+ ion (0–2.4 equiv.) into Rh1 solution, the absorption intensity at 323 nm gradually decreased with a simultaneous new absorption band appearance at 555 nm due to the possible formation of phenoxide ion, which can be further proven by later discussion. Interestingly, the absorption intensity as well as absorption position at isosbestic point 366 nm were remained constant during the interaction of Cu2+ with Rh1 solution. Moreover, the absorbance values at 323 nm and 555 nm became saturated and leveled off with 2.4 equiv. of Cu2+ ion, and thus demonstrated the formation of a 1[thin space (1/6-em)]:[thin space (1/6-em)]2 complexations between the chemosensor Rh1 and Cu2+ ion. The binding constant Ka and the calculated detection limit (LOD) were found to be 6.18 × 106 M−1 and 1.22 × 10−8 M (Fig. S14a and S15a), respectively.


image file: c6ra24472f-f5.tif
Fig. 5 The absorption spectral changes of (a) Rh1 and (b) Rh2 (10 μM in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4)) solution during the sequential additions of Cu2+.

Similarly, free Rh2 Chemosensor was colorless as shown in Fig. 5b, when Cu2+ was added from 0 to 2 equiv., an absorption band at 556 nm gradually increased with respect to the concentrations of Cu2+ ion, which induced an optical color change from colorless to pink color. Therefore, a selective colorimetric change was observed only in the presence of Cu2+ ion, which was attributed to the spirolactam ring opening mechanism and multiple ether units reported in a previous literature.54 The visible color change of Rh2 towards Cu2+ clearly demonstrates that Rh2 was highly sensitive to Cu2+. Moreover, the binding affinity of Rh2 towards Cu2+ was evaluated from UV-vis titration with an association constant Ka = 2.53 × 105 M−1 and its detection limit was estimated as 8.0 × 10−7 M (see Fig. S14b and S15b; ESI). Furthermore, after careful investigation of the fluorescence properties of dimeric systems Rh1 and Rh2, we have established that they could not induce any fluorescence changes with all metal ions, including Cu2+ ion (see Fig. S16a and b; ESI). The fluorescence quenching process was accompanied by the inherent quenching property of paramagnetic Cu2+ ion. The proposed binding mechanism of dimeric sensor systems was monitored by the UV-vis absorption spectra. In the absence of Cu2+, Rh1 and Rh2 displayed no absorbance above 400 nm, which indicated the closed spirolactam ring at this condition. Whereas, upon the addition of Cu2+ a new absorption band appeared at 555 nm suggested the disclosure of the spirolactam ring in the rhodamine unit. Therefore, we believe that the significant color change was solely due to the alteration of molecular structures between cyclic-spiro and ring-opening conformations of the rhodamine framework. Therefore, these observations suggested that among all metal ions only Cu2+ could induce the spirolactam ring opening of the rhodamine unit.

To further investigate the binding sites of dimeric systems, we did the 1H NMR titrations of probes Rh1 and Rh2 in the presence of Cu2+ metal ion. To perform the experiments, Rh1 and Rh2 were dissolved in DMSO-d6 (1 × 10−3 M) and Cu2+ solution (5 × 10−2 M) was prepared in D2O. After adding Cu2+ (0–2 equiv.) sequentially into Rh1 solution (1 equiv.), the total disappearance of –OH peak at 10.62 ppm (see Fig. S17), was well verified the clear binding of Rh1 with Cu2+ ion through the phenoxide interactions (after the deprotonation of –OH due to binding with Cu2+). In addition, owing to the Schiff-base binding of nitrogen atom with Cu2+ ion, the –CH proton peak was upfield-shifted from 9.07 to 8.59 ppm. In the case of Rh2 (see Fig. 6), during the addition of Cu2+ (0–2 equiv.), the Ha (–CH2) protons of the multiple ether unit (i.e., –O–CH2–CH2–O–) were upfield-shifted from 4.01 to 3.94 ppm. Similarly, the Hb (–CH2) protons (with a lower polarity) at 3.77 ppm were also upfield-shifted in the above experiment. Further, the little upfield-shifts of Schiff-base protons (i.e., –N[double bond, length as m-dash]CH–) were observed because of the complex formation between ‘N’ atoms and Cu2+. Hence, based on the 1H NMR titrations, a possible mechanism is proposed in Scheme 1 to explain the binding mode of probes Rh1 and Rh2 towards Cu2+.


image file: c6ra24472f-f6.tif
Fig. 6 1H NMR spectral changes of Rh2 (10 mM) in DMSO-d6 and titrated with 0–2.0 equiv. of Cu2+ in D2O.

image file: c6ra24472f-s1.tif
Scheme 1 A possible proposed binding mechanism of sensor probes Rh1 and Rh2 towards Cu2+.

Additionally, to ensure the binding site of Rh1 and Rh2 towards Cu2+, the stoichiometries of Cu2+ complexes (i.e., Rh1–Cu2+ and Rh2–Cu2+) were determined from Job's plots (see Fig. S18a and b; ESI) by fixing the final concentrations of Rh1–Cu2+ and Rh2–Cu2+ as 50 μM. The stoichiometries of both metal complexes were established by Job's plots between the mole fractions (Xm) and the maximum absorption changes at 553 nm and 556 nm for Rh1–Cu2+ and Rh2–Cu2+, respectively. After attaining the maximum values, the absorption peaks at 553 and 556 nm were slowly reduced from 0.6 molar fraction of Cu2+. Therefore, these phenomena clearly indicated the formation of 1[thin space (1/6-em)]:[thin space (1/6-em)]2 stoichiometric complexes for both dimeric systems towards Cu2+. To reconfirm the Cu2+ binding mode of the dimeric systems (Rh1 and Rh2), their FT-IR band (see Fig. S19a and b; ESI) of –OH at 3425 cm−1 and the ether linkage band (i.e., C–O–C) at 3460 cm−1 were investigated in the presence of Cu2+ ion. Correspondingly, the IR bands of C[double bond, length as m-dash]O (at 1706 and 1692 cm−1) and –C–H (at 3079 and 3065 cm−1) were also been evaluated for Rh1 and Rh2 dimeric systems. During the complex formation of probes Rh1 and Rh2 to Cu2+ ion, the above characteristic bands were affected incredibly as noted next. The characteristic bands of –OH and ether unit are became very broad along with the disappearance of –CH bands of Rh1 and Rh2 at 3079 and 3065 cm−1, respectively. Further to note, the respective C[double bond, length as m-dash]O stretching bands of Rh1 and Rh2 at 1706 and 1692 cm−1 were also diminished with slight shift. The above FTIR changes were well confirmed the attachment of colorimetric probes Rh1 and Rh2 towards Cu2+ ion.

The reversibilities of sensor complexes Rh1–Cu2+ and Rh2–Cu2+ in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) solutions are important for future applications. The reversible nature of the sensor complexes with Cu2+ ion was further carried out through UV-vis observations by adding penta-methyl ethylene-triamine (PMDTA), which is a well-known metal ion chelator. As depicted in Fig. S20a and b (ESI), upon the addition of PMDTA (2.6 and 2.2 equiv.) to the mixtures of Rh1–Cu2+ and Rh2–Cu2+ (10 μM), the absorption bands were found to be restored immediately. Moreover, by increasing the concentration of PMDTA (0–2.6 equiv. and 0–2.2 equiv., respectively) the reversibilities of Rh1–Cu2+ and Rh2–Cu2+ with PMDTA were clearly established as shown in Fig. 21a and b (ESI). Upon the addition of PMDTA, the absorption peaks of dimeric probes Rh1 and Rh2 were decreased gradually to characterize the reversible sensing capabilities of sensor probes (Rh1 and Rh2). This reversibility allow us to reuse of the sensor probes for effective Cu2+ detection.

3.2 Colorimetric responses of metal complexes (Rh1–Cu2+ and Rh2–Cu2+) towards anions

Next, we used the previously formed metal complexes Rh1–Cu2+ and Rh2–Cu2+ to examine their anion selectivities. The absorbance spectral studies of metal complexes with different anions F, CN, HPO42−, H2PO4, HSO4, ClO4, Br, OH, SCN, P2O74−, NO3 and OAc were carried out in the solution condition of 10 μM in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4). The color changes for metal complexes (Rh1–Cu2+ and Rh2–Cu2+) towards various anions are demonstrated in the photograph images of Fig. 7, where no color changes were observed in all anions, except pyrophosphate (PPi).
image file: c6ra24472f-f7.tif
Fig. 7 Photo-images of color changes for metal complexes (Rh1–Cu2+ and Rh2–Cu2+) towards various anions.

The addition of PPi anion turned the original pink colored solution into a colorless solution. Upon the addition of PPi (1 equiv.), not only these color changes could be easily recognized by naked-eyes, but also the UV-vis spectral changes were apparent as noticed in Fig. 8a and b. On the other hand, metal complexes didn't show any considerable spectral changes in absorption peaks during the addition of other anions (up to 4 equiv.). Therefore, the selectivity of the metal complexes (Rh1–Cu2+ and Rh2–Cu2+) towards PPi anion might be attributed to their release of Rh1 and Rh2 through the strong binding affinity of PPi to copper ion as illustrated in Scheme 2. This result indicates that both metal complexes can serve as potential candidates as “naked-eye” chemosensor for PPi detection in semi-aqueous medium. Meanwhile, we further investigated the binding properties of both metal complexes by UV-vis titration experiments. Fig. 9 showed the colorimetric changes of metal complexes (Rh1–Cu2+ and Rh2–Cu2+) 10 μM in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4). Upon the addition of increasing concentrations (1 equiv.) of PPi, the absorbance at 554 nm was decreased linearly with increasing the concentration of PPi. The binding constants between the metal complexes (Rh1–Cu2+ and Rh2–Cu2+) and PPi were calculated as 1.09 × 105 M−1 and 4.11 × 104 M−1, respectively (see Fig. S22a and b; ESI). Furthermore, based on the results of UV-vis titrations, the detection limits of metal complexes (Rh1–Cu2+ and Rh2–Cu2+) mediated PPi sensors were estimated to be 0.389 × 10−6 M and 1.61 × 10−6 M, respectively, as shown in Fig. S23a and b (ESI). These observations suggest that both metal complexes (Rh1–Cu2+ and Rh2–Cu2+) could be excellent chemosensors for the detection of PPi anion.


image file: c6ra24472f-f8.tif
Fig. 8 (a) UV-visible spectral changes of Rh1–Cu2+ and (b) Rh2–Cu2+ (10 μM) upon the addition of several relevant anions in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4).

image file: c6ra24472f-s2.tif
Scheme 2 Possible proposed sensing mechanism of metal complexes (Rh1–Cu2+ and Rh2–Cu2+) towards pyrophosphate (PPi) anion.

image file: c6ra24472f-f9.tif
Fig. 9 Absorption spectral changes of (a) Rh1–Cu2+ and (b) Rh2–Cu2+ (10 μM) in CH3CN–H2O (v/v = 9[thin space (1/6-em)]:[thin space (1/6-em)]1, 5 mM Tris–HCl, pH 7.4) during the sequential additions of PPi.

For practical applications, we need to evaluate the utilizations of colorimetric sensor probes under wider pH conditions (other than the previous neutral form). As expected, the absorption peaks of un-bonded Rh1 and Rh2 (without Cu2+) were only observable in the ranges of pH = 2.0–5.0, suggesting that the strong protonation and spirolactam ring opening process of sensor probes occurred at certain pH ranges. In order to investigate the applications of Rh1 and Rh2 over a wide pH span of 0–12, the absorption intensities of Rh1 and Rh2 increased by adding Cu2+ under pH = 3.0–8.0 (see Fig. S24a and b; ESI). In this pH zone, Cu2+ induced unique ring opening process in the rhodamine unit was obtained, which clearly demonstrated Rh1 and Rh2 as effective colorimetric probes towards Cu2+ under environmental conditions.

3.3 Theoretical calculations

To evaluate the experimental observations and provide a better insight into probes Rh1 and Rh2 along with their copper complexes, the DFT and TD-DFT calculations were carried out and summarized in Table 1. The model structures have been built for Rh1 and its copper complex as shown in Fig. 10.
Table 1 The predicted values of HOMO, LUMO, HOMO LUMO gap (HLG) and oscillator strength (f) for Rh1, Rh1 + Cu2+, Rh2 and Rh2 + Cu2+ by DFT and TD-DFT calculations
System Tot. Ea HOMO (eV) LUMO (eV) HLGb (eV) Excited states fc Contribution
a Total energy.b HOMO–LUMO gap.c Oscillator strength.
Rh1 −3367.381623 −4.269328 −1.625269 2.64 S1 0.0475 H → L (0.69)
H → L+1 (−0.1)
S2 0.0146 H−1 → L (0.69)
H−1 → L+1 (0.1)
Rh1 + Cu2+ −3759.353208 −7.281195 −5.181813 2.10 S1 0.0001 H → L+2 (0.6)
H → L+3 (−0.33)
H → L+5 (0.16)
S2 0.0477 H → L+2 (0.34)
H → L+3 (0.61)
Rh2 −3830.768067 −4.172246 −1.302659 2.87 S1 0.011 H → L+1 (0.71)
S2 0.009 H+1 → L+2 (0.7)
Rh2 + Cu2+ −4223.285313 −4.172246 −8.467632 1.28 S1 0.0025 H → L+1 (0.7)
S2 0 H+1 → L+1 (0.7)
S3 0.0026 H+1 → L+2 (0.7)



image file: c6ra24472f-f10.tif
Fig. 10 Optimized structures of (a) Rh1 and (b) complex Rh2–Cu2+ at B3LYP/LANL2DZ in the gas phase.

The main absorption bands of transitions were H → L, H → L+2, H → L+1 and H → L+1 observed in Rh1, Rh1–Cu2+, Rh2 and Rh2–Cu2+, respectively. Fig. 11 shows the molecular orbital of Rh1, where the electron density of HOMO was located partially on the catechol unit and further on the oxa-anthracene unit and the LUMO was partially located on the catechol unit and further at the indole units on both sides. As shown in Fig. 12, the electron density of HOMO in complex Rh1–Cu2+ was located on the hydrazine units with part of Rh2–Cu2+ are shown in Fig. S25a and b (ESI). The respective molecular orbitals of Rh2 and Rh2–Cu2+ are shown in Fig. S26 and S27 (ESI). As shown in Fig. S26 (ESI) in Rh2, the electron density of HOMO was located on a spiro unit, and that of LUMO was also located on a spiro unit. The electron density of HOMO has been shifted slightly towards the edge of the spiro unit in Rh2–Cu2+ complex. The LUMO also covered partially one Cu2+ unit with aza-carbonyl and benzene units of the complex. From the molecular orbitals of probes and complexes (Fig. 10, 11, S26 and S27), the electron densities located on the same part of the HOMO and LUMO for the cases of Rh2 and Rh2–Cu2+, but on the different parts of HOMO and LUMO for the cases of Rh1 and Rh1–Cu2+. Based on our calculated molecular orbital diagram in Fig. 11, the conjugated system was broken due to the addition of Cu2+ which led to the stronger ICT mechanism55,56 in the Cu2+ complexes. This could be the possible explanations for the experimental observations of ratiometric responses of the Cu2+ complexes. Furthermore, the HLG value was also reduced in complex Rh1–Cu2+ (2.10 eV) in contrast to that of Rh1 (2.64 eV).


image file: c6ra24472f-f11.tif
Fig. 11 HOMO–LUMO structures of Rh1 at B3LYP/LANL2DZ level in the gas phase.

image file: c6ra24472f-f12.tif
Fig. 12 HOMO–LUMO structures of complex Rh1 + Cu2+ at B3LYP/LANL2DZ level in the gas phase.

3.4 Colorimetric responses on a solid support

In analytical chemistry grade, the sensitivity of a sensing system is crucial. Normally, it remains big challenges to achieve high sensitivities in such colorimetric sensing systems. We systematically investigated the selectivities, sensitivities and reversibilities of Rh1 and Rh2 on TLC plates coated with ready-made silica. Test strips were prepared by immersing silica gel plates into a CH3CN solution of Rh1 (10−4 M) and then were dried by a drier. As shown in Fig. 13, the strips were examined with various metal ions (10−3 M) by dipping into the corresponding solutions. We found that only copper ion induced a remarkable color change from colorless to pink, implying that probe was a highly selective naked-eye sensor probe towards Cu2+ over the other metal ions.
image file: c6ra24472f-f13.tif
Fig. 13 Photographs of sensor tests on Rh1 strips in the presence of various metal ions.

The detection results of test strips containing Rh1 towards Cu2+ with different concentrations are also illustrated in Fig. S28. The higher concentration of copper ion, the more apparent color change from colorless to pink, where the detectable concentration of copper ion could be as low as 4 × 10−6 M. As shown in Fig. 14, further reversibility investigation revealed that the pink color of the test strips containing complex Rh1–Cu2+ faded out in the presence of PMDTA (10−5 M). Therefore, we have developed the notable and reversible naked-eye sensor strips to detect low concentrations of copper ion (4 μM) in this study. Similar results regarding selectivity, sensitivity and reversibility of test strips for Rh2 were also been obtained (see Fig. S29–S31; ESI). Therefore, test strips containing Rh1 and Rh2 can conveniently detect aqueous copper ion, which demonstrate a cost-effective detection method without using any equipments.


image file: c6ra24472f-f14.tif
Fig. 14 Photographs of reversibility tests on Rh1 strips (10−4 M) in the presence of PMDTA (10−5 M). From left to right: Rh1, Rh1–Cu2+ and Rh1–Cu2+ + PMDTA.

4 Conclusions

We have developed reversible colorimetric probes Rh1 and Rh2 for the selective and sensitive sequential detections of Cu2+ and pyrophosphate via the ratiometric absorption spectral changes. Stoichiometry and binding mechanisms for both probes were well characterized and established by the respective spectroscopic techniques and theoretical calculations. Observations revealed a better performance of Rh1 towards the selective and effective colorimetric detection of Cu2+ ion at a lower concentration than that of Rh2. Reversibility may open the avenue towards cost effectiveness and understanding the role of metal ions by switching signals between structural transformations and thus for easy detections of analytes for societal demands. Detections of Cu2+ by the designed probes were performed in solution and on solid supports by simple titration and dipping of metal ion solutions (without utilizing any drastic conditions), which may push the limits of our designed probes to future concerning ecological problems as well as living systems. Importantly, Rh1 and Rh2 are good example of reversible rhodamine-based colorimetric probes for sequential detections of Cu2+ ion and PPi via ratiometric responses.

Acknowledgements

The financial supports of this project are provided by the Ministry of Science and Technology (MOST) in Taiwan through MOST 103-2113-M-009-018-MY3 and MOST 103-2221-E-009-215-MY3. Mr Parthiban Venkatesan is acknowledged for his timely help during the paper writing.

References

  1. Y. Yang, Q. Zhao, W. Feng and F. Li, Chem. Rev., 2012, 113, 192–270 CrossRef PubMed.
  2. Z. Liu, C. Peng, Z. Lu, X. Yang, M. Pei and G. Zhang, Dyes Pigm., 2015, 123, 85–91 CrossRef CAS.
  3. X. Li, M. Yu, F. Yang, X. Liu, L. Wei and Z. Li, New J. Chem., 2013, 37, 2257–2260 RSC.
  4. M. Shellaiah, Y. H. Wu, A. Singh, M. V. Ramakrishnam Raju and H.-C. Lin, J. Mater. Chem. A, 2013, 1, 1310–1318 CAS.
  5. D. Udhayakumari, S. Velmathi, W. C. Chen and S. P. Wu, Sens. Actuators, B, 2014, 204, 375–381 CrossRef CAS.
  6. H. T. Feng, S. Song, Y. C. Chen, C. H. Shen and Y. S. Zheng, J. Mater. Chem. C, 2014, 2, 2353–2359 RSC.
  7. G. Sivaraman, T. Anand and D. Chellappa, RSC Adv., 2013, 3, 17029–17033 RSC.
  8. Y. Zhou, F. Wang, Y. Kim, S. J. Kim and J. Yoon, Org. Lett., 2009, 11, 4442–4445 CrossRef CAS PubMed.
  9. P. Zhang, L. Pei, Y. Chen, W. Xu, Q. Lin, J. Wang, J. Wu, Y. Shen, L. Ji and H. Chao, Chem.–Eur. J., 2013, 19, 15494–15503 CrossRef CAS PubMed.
  10. J. Jeong, B. A. Rao and Y. A. Son, Sens. Actuators, B, 2015, 220, 1254–1265 CrossRef CAS.
  11. E. Gaggelli, H. Kozlowski, D. Valensin and G. Valensin, Chem. Rev., 2006, 106, 1995–2044 CrossRef CAS PubMed.
  12. M. Yang, W. Meng, X. Liu, N. Su, J. Zhou and B. Yang, RSC Adv., 2014, 4, 22288–22293 RSC.
  13. G. Y. Lan, C. C. Huang and H. T. Chang, Chem. Commun., 2010, 46, 1257–1259 RSC.
  14. Y. Chen, Y. Mi, Q. Xie, J. Xiang, H. Fan, X. Luo and S. Xia, Anal. Methods, 2013, 5, 4818–4823 RSC.
  15. H. S. Jung, P. S. Kwon, J. W. Lee, J. I. Kim, C. S. Hong, J. W. Kim, S. Yan, J. Y. Lee, J. H. Lee, T. Joo and J. S. Kim, J. Am. Chem. Soc., 2009, 131, 2008–2012 CrossRef CAS PubMed.
  16. M. Li, H. Ge, R. L. Arrowsmith, V. Mirabello, S. W. Botchway, W. Zhu, S. I. Pascu and T. D. James, Chem. Commun., 2014, 50, 11806–11809 RSC.
  17. D. Udhayakumari, S. Velmathi, Y. M. Sung and S. P. Wu, Sens. Actuators, B, 2014, 198, 285–293 CrossRef CAS.
  18. S. D. S. Parveen, B. S. Kumar, S. R. Kumar, R. I. Khan and K. Pitchumani, Sens. Actuators, B, 2015, 221, 75–80 CrossRef.
  19. V. Tharmaraj and K. Pitchumani, J. Mater. Chem. B, 2013, 1, 1962–1967 RSC.
  20. D. Maity, A. K. Manna, D. Karthigeyan, T. K. Kundu, S. K. Pati and T. Govindaraju, Chem.–Eur. J., 2011, 17, 11152–11161 CrossRef CAS PubMed.
  21. H. Kim, B. A. Rao, J. W. Jeong, S. Mallick, S. M. Kang, C. S. Choi, Y. A. Lee and Y. A. Son, Sens. Actuators, B, 2015, 210, 173–182 CrossRef CAS.
  22. (a) N. Aksuner, E. Henden, I. Yilmaz and A. Cukurovali, Dyes Pigm., 2009, 83, 211–217 CrossRef CAS; (b) K. Ghosh, D. Tarafdar, A. Majumdar, C. G. Daniliuc, A. Samadder and A. R. Khuda-Bukhsh, RSC Adv., 2016, 6, 47802–47812 RSC; (c) G. Li, F. Tao, H. Wang, L. Wang, J. Zhang, P. Ge, L. Liu, Y. Tong and S. Sun, RSC Adv., 2015, 5, 18983–18989 RSC; (d) A. Rai, N. Kumari, R. Nair, K. Singh and L. Mishra, RSC Adv., 2015, 5, 14382–14388 RSC.
  23. Z. Li, W. Zhang, X. Liu, C. Liu, M. Yu and L. Wei, RSC Adv., 2015, 5, 25229–25235 RSC.
  24. D. H Lee, S. Y. Kim and J. I. Hong, Angew. Chem., Int. Ed., 2004, 43, 4777–4780 CrossRef PubMed.
  25. J. H. Wang, J. B. Xiong, X. Zhang, S. Song, Z. H. Zhu and Y. S. Zheng, RSC Adv., 2015, 5, 60096–60100 RSC.
  26. E. Quinlan, S. E. Matthews and T. Gunnlaugsson, J. Org. Chem., 2007, 72, 7497–7503 CrossRef CAS PubMed.
  27. (a) S. Bhowmik, B. N. Ghosh, V. Marjomäki and K. Rissanen, J. Am. Chem. Soc., 2014, 136, 5543–5546 CrossRef CAS PubMed; (b) S. Goswami, S. Paul and A. Manna, RSC Adv., 2013, 3, 10639–10643 RSC.
  28. (a) P. Anzenbacher, M. A. Palacios, K. Jursíková and M. Marquez, Org. Lett., 2005, 7, 5027–5030 CrossRef CAS PubMed; (b) S. Lee, K. K. Y. Yuen, K. A. Jolliffe and J. Yoon, Chem. Soc. Rev., 2015, 44, 1749–1762 RSC; (c) V. E. Zwicker, B. M. Long and K. A. Jolliffe, Org. Biomol. Chem., 2015, 13, 7822–7829 RSC.
  29. L. S. A. Haque, R. L. Bolhofner, B. M. Wong and M. A. Hossain, RSC Adv., 2015, 5, 38733–38741 RSC.
  30. S. Kim, M. S. Eom, S. Yoo and M. S. Han, Tetrahedron Lett., 2015, 56, 5030–5033 CrossRef CAS.
  31. G. Sivaraman and D. Chellappa, J. Mater. Chem. B, 2013, 1, 5768–5772 RSC.
  32. J. W. Jeong, B. A. Rao and Y. A. Son, Sens. Actuators, B, 2015, 208, 75–84 CrossRef CAS.
  33. H. N. Kim, M. H. Lee, H. J. Kim, J. S. Kim and J. Yoon, Chem. Soc. Rev., 2008, 37, 1465–1472 RSC.
  34. Y. Q. Sun, J. Liu, X. Lv, Y. Liu, Y. Zhao and W. Guo, Angew. Chem., Int. Ed., 2012, 51, 7634–7636 CrossRef CAS PubMed.
  35. L. Huang, X. Wang, G. Xie, P. Xi, Z. Li, M. Xu, Y. Wu, D. Bai and Z. Zeng, Dalton Trans., 2010, 39, 7894–7896 RSC.
  36. A. Chatterjee, M. Santra, N. Won, S. Kim, J. K. Kim, S. B. Kim and K. H. Ahn, J. Am. Chem. Soc., 2009, 131, 2040–2041 CrossRef CAS PubMed.
  37. Z. Q. Hu, X. M. Wang, Y. C. Feng, L. Ding and H. Y. Lu, Dyes Pigm., 2011, 88, 257–261 CrossRef CAS.
  38. F.-M. Yang, W. Meng, Q. Ding, N. Su, X. Liu, M. Zhang and B. Yang, New J. Chem., 2015, 39, 4790–4795 RSC.
  39. S. D. S. Parveen, A. Affrose and K. Pitchumani, Sens. Actuators, B, 2015, 221, 521–527 CrossRef.
  40. B. Imene, Z. Cui, X. Zhang, B. Gan, Y. Yin, Y. Tian, H. Deng and H. Li, Sens. Actuators, B, 2014, 199, 161–167 CrossRef CAS.
  41. S. Jayabal, R. Sathiyamurthi and R. Ramaraj, J. Mater. Chem. A, 2014, 2, 8918–8925 CAS.
  42. Q. Li, Y. Guo, J. Xu and S. Shao, Sens. Actuators, B, 2011, 15, 427–431 CrossRef.
  43. H. Li, Z. Cui and C. Han, Sens. Actuators, B, 2009, 143, 87–92 CrossRef.
  44. J. Y. Noh, G. J. Park, Y. J. Na, H. Y. Jo, S. A. Lee and C. Kim, Dalton Trans., 2014, 43, 5652–5656 RSC.
  45. Y. Lv, Y. Guo, J. Xu and S. Shao, J. Fluorine Chem., 2011, 132, 973–977 CrossRef CAS.
  46. G. Sivaraman, T. Anand and D. Chellappa, Analyst, 2012, 137, 5881–5884 RSC.
  47. M. J. Culzoni, A. Munoz de la Pena, A. Machuca, H. C. Goicoechea and R. Babiano, Anal. Methods, 2013, 5, 30–49 RSC.
  48. C. R. Lohani, J. M. Kim, S. Y. Chung, J. Yoon and K. H. Lee, Analyst, 2010, 135, 2079–2084 RSC.
  49. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian Inc, Wallingford CT, 2009.
  50. A. D. Becke, J. Chem. Phys., 1997, 107, 8554–8560 CrossRef CAS.
  51. P. Gou, N. D. Kraut, I. M. Feigel and A. Star, Macromolecules, 2013, 46, 1376–1383 CrossRef CAS.
  52. K. D. Belfield, M. V. Bondar, A. Frazer, A. R. Morales, O. D. Kachkovsky, I. A. Mikhailov, A. M. E. Masunov and O. V. Przhonska, J. Phys. Chem. B, 2010, 114, 9313–9321 CrossRef CAS PubMed.
  53. X. Zhang and Y. Y. Zhu, Sens. Actuators, B, 2014, 202, 609–614 CrossRef CAS.
  54. H. Y. Lee, K. M. K. Swamy, J. Y. Jung, G. Kim and J. Yoon, Sens. Actuators, B, 2013, 182, 530–537 CrossRef CAS.
  55. H. S. Jung, K. C. Ko, G. H. Kim, A. R. Lee, Y. C. Na, C. Kang, J. Y. Lee and J. S. Kim, Org. Lett., 2011, 13, 1498–1501 CrossRef CAS PubMed.
  56. L. Deng, W. Wu, H. Guo, J. Zhao, S. Ji, X. Zhang, X. Yuan and C. Zhang, J. Org. Chem., 2011, 76, 9294–9304 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: NMR and mass spectra of Rh1 and Rh2 and sensor properties mentioned in the text. See DOI: 10.1039/c6ra24472f

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.