High selectivity up-converted fluorescence turn-on probe for Zn2+ based on PAMAM hydroxy-naphthalene Schiff-bases (C[double bond, length as m-dash]N) half-organic quantum dots

Yan Ji and Ying Qian *
School of Chemistry and Chemical Engineering, Southeast University, Nanjing, 211189, China. E-mail: yingqian@seu.edu.cn; jiyan98@163.com; Tel: +86 13851438813

Received 28th February 2014 , Accepted 12th May 2014

First published on 12th May 2014


Abstract

The dendrimer PNS-G0 with Schiff-base imines (C[double bond, length as m-dash]N) realizes highly selective and common fluorescence (λex = 400 nm) or up-converted fluorescence (λex = 800 nm) turn-on effect to qualitatively and quantitatively detect zinc ions (Zn2+). The PNS-G0 + Zn2+ complex (C[double bond, length as m-dash]N_Zn_O) fluorescence emitting parts were firstly proposed as half-organic quantum dots (HOQDs). The HOQDs probe PNS-G0 for Zn2+ by the coordination method based on imine isomerization (inhibition) mechanism, which has potential applications in biological imaging, analytical chemistry, and optical physics areas.


1. Introduction

Zinc is an important element in biological tissues, for many enzymes and proteins include the zinc ion (Zn2+).1,2 Therefore, the qualitative and quantitative analysis for Zn2+ in tissues and the detection of Zn2+ positions are important in the study of biological processes. Fluorescent probe molecules are simple, sensitive, fast, and efficient tools for detecting Zn2+ ions.3 The two-photon absorption induced up-converted fluorescence probe for Zn2+ has been a topic of interest in biology, chemistry, and even physics areas.4

Several probe mechanisms have been reviewed,5 such as photoinduced electron transfer (PET),6 excimer/exciplex formation,7 intramolecular charge transfer (ICT),8 metal–ligand charge transfer (MLCT),9 fluorescence resonance energy transfer (FRET),10 twisted intramolecular charge transfer (TICT),11 electronic energy transfer (EET),12 aggregation-induced emission (AIE),13 and excited-state intramolecular proton transfer (ESIPT).14 Recently, the Schiff-base imine (C[double bond, length as m-dash]N) isomerization phenomena were investigated and have been used in the design and synthesis of a new type of probe.15 The Schiff-base imine (C[double bond, length as m-dash]N) exhibits no or weak fluorescence for the isomerization phenomena. If the imine isomerization is inhibited by complexing with metal ions through coordination bonds, then the imine exhibits fluorescence emission.5 These C[double bond, length as m-dash]N isomerization-based probes can realize a fluorescence turn-on effect in detecting various metal ions (e.g. Mg2+, Cu2+, Cd2+).16

Based on this strategy, a probe for Zn2+ ions was developed by synthesizing the dendrimer PNS-G0 (Scheme 1) with imine (C[double bond, length as m-dash]N) groups from poly-(amido-amine) generation zero (PAMAM G0) and 2-hydroxy-1-naphthaldehyde through the Schiff base reaction. PNS-G0 has a PAMAM-G0 center core with many amines (tertiary amine, amide, and imine). Furthermore, PAMAM has intramolecular holes between flexible chains, which can complex with metal ions and package nanoparticles or small molecules, and even realize drug delivery. The PAMAM can emit fluorescence after storage or oxidation in air,17 which has elicited interest in related areas such as fluorescence probe design. The end group of PNS-G0 is 2-hydroxy-1-naphthaldehyde. The hydroxyl was reserved for forming the coordination bond. The aldehyde can react with the amine of PAMAM by Schiff base condensation reaction, which can form the imine (C[double bond, length as m-dash]N) group for use in the C[double bond, length as m-dash]N isomerization mechanism.


image file: c4ra01758g-s1.tif
Scheme 1 The structures of PNS-G0 and its complex with Zn2+.

The semiconductor nanoparticles known as quantum dots (QDs),18 are characterized by large Stokes shifts, broad absorption bands, and narrow, size-dependent emission bands without a significant red tail. The multiple color QDs can be excited using a single laser excitation wavelength.19 The size-dependent emission of QDs is the result of a quantum confinement effect (QCE). The QCE occurs when the size of the exciton (exciton Bohr radius, B) exceeds the physical size of the semiconductor nanocrystal (D).20 Semiconductor nanocrystals exhibit particularly interesting properties when the exciton is strongly confined (D < 2B);21 they outperform fluorescent dyes in terms of brightness and photo stability.22 Inorganic QDs composited with organic molecules form a new type of QDs, as half-organic quantum dots (HOQDs) were firstly named. HOQDs have been involved in recent research, such as cysteine thiol QDs23 and PAMAM thiol modified QDs,24 as well as other half-organic part QDs.25 In this paper, the imine_Zn_hydroxyl (C[double bond, length as m-dash]N_Zn_O) complex HOQDs are discussed.

2. Results and discussions

2.1. Selectivity of probing Zn2+

Fig. 1 shows the photographs of PNS-G0 and its complex with several metal ions in DMSO solutions. From Fig. 1, the PNS-G0 gave weak emission under UV light. PNS-G0 gave strong blue fluorescence after complexing with Zn2+. PNS-G0 complexed with other metal ions (Ca2+, Mg2+, Pb2+, and Cd2+) gave weak fluorescence. Fig. 1 shows the fluorescence selectivity of PNS-G0 for the Zn2+ probe as distinguished by the naked eye.
image file: c4ra01758g-f1.tif
Fig. 1 Photographs of the probe PNS-G0 and its complex with different metal ions (Zn2+, Ca2+, Mg2+, Pb2+, Cd2+) in DMSO (5 × 10−5 mol L−1). The upper photograph was in room light; the lower photograph was in UV light (365 nm).

Fig. 2 gives the UV-vis absorption spectra of PNS-G0 and the PNS-G0 complex with Zn2+ and other metal ions. The main absorbance peaks were at 250 nm, 300 nm, and 400 nm. Compared with the UV-vis absorption spectra of the PNS-G0 and PNS-G0 complex, the Zn2+ ion complex has a higher UV-vis absorbance peak at 250 nm and a lower UV-vis absorbance peak at 300 nm than those of other metal ion complexes. The Zn2+ ion complex UV-vis absorbance peaks have 380 nm peaks and show a blue shift compared with other metal ion complexes. The UV-vis absorption spectra show different spectra for the PNS-G0 + Zn2+ complex compared with other metal ions. It can be seen that the Zn2+ complex with PNS-G0 changes the n–π* and π–π* transition absorbance, which shows the unique absorbance characteristic of the Zn2+ + PNS-G0 complex, different from the UV-vis absorption spectra of PNS-G0 and the other metal ion complexes.


image file: c4ra01758g-f2.tif
Fig. 2 The UV-vis absorption spectra of PNS-G0 (5 × 10−5 mol L−1) and PNS-G0 complex with different metal ions (Zn2+, Cd2+, Mg2+, Ca2+) (1 × 10−4 mol L−1).

The fluorescence emission spectra in Fig. 3 show that the PNS-G0 gave ten times fluorescence enhancement after complexing with Zn2+, while other metal ion complexes gave very weak fluorescence emission. Employing the sulfuric acid quinine fluorescence reference method test, the fluorescence quantum yield Φ of PNS-G0 and its metal ion complexes gave ΦPNS-G0 = 0.06; ΦPNS-G0+Ca2+ = 0.01; ΦPNS-G0+Mg2+ = 0.04; ΦPNS-G0+Pb2+ = 0.01; ΦPNS-G0+Cd2+ = 0.02; ΦPNS-G0+Zn2+ = 0.73, which show that the PNS-G0 complex with Zn2+ gave the highest Φ, about ten times than that of others. Fig. 4 shows the PNS-G0 complexes in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1, and the Zn2+ complex has higher emission intensity than other metal ion complexes. The introduction of water to the solvent can enhance biocompatibility, and PNS-G0 shows good fluorescence selectivity in this mixed solvent. It can be seen from Fig. 1–4 that the PNS-G0 probe realized fluorescence enhanced and fluorescence turn-on effect on detecting Zn2+ and showed high selectivity by comparing with other metal ions.


image file: c4ra01758g-f3.tif
Fig. 3 The fluorescence emission of the PNS-G0 probe (5 × 10−5 mol L−1) and its complexes with different metal ions (Zn2+, Ca2+, Mg2+, Pb2+, Cd2+, Mn2+, Fe3+, Fe2+, Ni2+, Cu2+, Co2+, Cr3+, Ag+, H+, NH4+, Na+, K+) (1 × 10−4 mol L−1) in DMSO (excited under 400 nm wavelength light).

image file: c4ra01758g-f4.tif
Fig. 4 The fluorescence emission of probe PNS-G0 (1 × 10−6 mol L−1) and its complexes with different metal ions (Zn2+, Ca2+, Mg2+, Pb2+, Cd2+) (1 × 10−4 mol L−1) in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 (excited under 400 nm wavelength light).

2.2 Quantitative probing of sZn2+

The UV-vis absorption spectra of PNS-G0 and its complex with different concentrations of Zn2+ in DMSO solvent are shown in Fig. 5. The UV-vis absorbance peaks become higher at 250 nm and lower at 300 nm along with increasing Zn2+ concentration. It was noticed that the UV-vis absorbance peaks at about 400 nm gave a blue shift along with the increasing Zn2+ concentration, which shows the Zn2+ influence on the π–π* conjugated imine-naphthalene absorbance after complexing with PNS-G0.
image file: c4ra01758g-f5.tif
Fig. 5 The UV-vis absorption spectra of PNS-G0 (1 × 10−4 mol L−1) and PNS-G0 complex with different concentrations of [Zn2+] (mol L−1).

Fig. 6 shows the fluorescence .emission spectra of PNS-G0 + Zn2+ complex with different concentrations of Zn2+ in DMSO. The PNS-G0 was kept at 5 × 10−5 mol L−1 and added into Zn2+ DMSO solution with concentrations from 1 × 10−8 mol L−1 to 1 × 10−3 mol L−1. The fluorescence increased along with the increasing [Zn2+] concentration.


image file: c4ra01758g-f6.tif
Fig. 6 The fluorescence emission spectra of PNS-G0 (5 × 10−5 mol L−1) and PNS-G0 complex with different concentrations [Zn2+] in DMSO.

PNS-G0 can qualitatively detect Zn2+ and also can quantitatively detect the Zn2+, as shown in Fig. 7. The inset graph in Fig. 7 shows the good linear relationship of the PNS-G0 probing the [Zn2+] (range from 1 × 10−8 mol L−1 to 5 × 10−6 mol L−1). From the data in Fig. 5, the approximate binding constant of the single branch of PNS-G0 was K1 ≈ 2 × 103 M−1 by the equation26 ΔA/L = (([Dt]K1Δε[M])/(1 + K1[M])). The fluorescence spectra in Fig. 8 shows the PNS-G0 qualitatively detected the [Zn2+] in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1. The 1/(FFmin) vs. 1/[Zn2+]26c have good linear relationship. It can be seen from Fig. 5–8 that the probe PNS-G0 realized qualitative detection of the Zn2+ by the fluorescence method. The use of organic solvent and the nearly two times enhanced selective signal in water–DMSO mixed solvents show the dyes are inconvenient for biological imaging and need further modification.


image file: c4ra01758g-f7.tif
Fig. 7 The fluorescence emission spectra of PNS-G0 (5 × 10−5 mol L−1) added with Zn2+ (1 × 10−8 mol L−1 ∼5 × 10−6 mol L−1) in DMSO excited under 400 nm wavelength light inset: 1/(FFmin) vs. 1/[Zn2+], F is the fluorescent emission spectra integrated area; Fmin is the minimum [Zn2+]-added sample's fluorescence spectra integrated area.

image file: c4ra01758g-f8.tif
Fig. 8 The fluorescence emission spectra (top) of PNS-G0 (1 × 10−6 mol L−1) and PNS-G0 complex with different concentrations of [Zn2+] in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 (excited under 400 nm wavelength light). 1/(FFmin) vs. 1/[Zn2+] (bottom): F is fluorescent emission spectra integrated area; Fmin is the minimum [Zn2+]-added sample's fluorescence spectra integrated area.

2.3 Up-converted fluorescence spectra

The up-converted fluorescence (anti-Stokes emissions) can realize long wavelength excitation and short wavelength emission.27 After the advent of lasers, the two- and even three-photon absorption-induced frequency up-converted fluorescence in organic dye materials could be observed by using pulsed laser excitation.28 The long excitation wavelength of up-converted fluorescence can reduce optical damage, achieve deep penetration, dark field imaging, high resolution three-dimensional display, and other advantages of probes used in biological tissues.29

Fig. 9 shows the emission peaks at 535 nm as a result of excitation under 800 nm wavelength light, which is attributed to the two-photon mechanism, up-converted fluorescence. The up-converted fluorescence emission spectra of PNS-G0 and complex PNS-G0 + Zn2+ realize turn-on effect in DMSO (insert in Fig. 9). The up-converted fluorescence properties of PNS-G0 excited under 800 nm also can qualitatively and quantitatively detect the Zn2+, as shown in Fig. 9. The introduction of water mixed with 1% DMSO as solvent extends the application of PNS-G0 in the imaging of cells or tissue.


image file: c4ra01758g-f9.tif
Fig. 9 Up-converted fluorescence emission spectra of PNS-G0 (1 × 10−6 mol L−1) and complex with different concentrations of Zn2+ in solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 (top figure) and up-converted fluorescence spectra of PNS-G0 compared complex with Zn2+ in DMSO (bottom figure). Inset: 1/(FFmin) vs. 1/[Zn2+] figure, F is fluorescent intensity, (R2 = 0.97). The excitation wavelength is 800 nm. Fmin is the minimum [Zn2+]-added sample's fluorescence intensity.

Fig. 10 shows the up-converted fluorescence emission spectra of the PNS-G0 probe and its complexes with different metal ions (Zn2+, Ca2+, Mg2+, Pb2+, Cd2+) in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 excited under 800 nm wavelength light. Fig. 10 shows that the PNS-G0 + Zn2+ complex has higher fluorescence emission peaks than that of PNS-G0 and other metal ion complexes excited at 800 nm wavelength light. It can be seen from Fig. 10 that the probe PNS-G0 realized nearly two-times selective detection of Zn2+ compared with other metal ions by up-conversion of the fluorescence spectra.


image file: c4ra01758g-f10.tif
Fig. 10 The up-converted fluorescence emission spectra of probe PNS-G0 (1 × 10−6 mol L−1) and its complex with different metal ions (Zn2+, Ca2+, Mg2+, Pb2+, Cd2+) (1 × 10−4 mol L−1) in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 (excited under 800 nm wavelength light).

The spectra in Fig. 9 and 10 show the up-converted fluorescence under common linear light source. The experiments were taken by FluoroMax-4 spectrofluorometer using the excitation source xenon, continuous output, ozone-free lamp (150 W) at room temperature (∼20 °C), which replaced the laser instruments. The emission phenomena of linear light source-excited organic molecules, giving up-converted fluorescence, were firstly tested. The FluoroMax-4 spectrofluorometer is the common fluorescence testing instrument. The light source was linear, which is different from the laser source. The linear source-excited, up-converted fluorescence organic molecules included PNS-G0 + Zn2+ dendrimers, which might be of interest in chemistry, physics, and biology areas.

Up-converted fluorescence has three mechanisms:30 the first is second-harmonic generation, the second is two-photon absorption, and the third is the photon avalanche31 process. The UV-vis absorption spectra in Fig. 2 and 5 show no linear absorbance at 800 nm, indicating the existence of nonlinear absorbance that produced the fluorescence emission under 800 nm excitation wavelength. Fig. 9 shows that the emission peaks at about 535 nm are attributed to two-photon absorption induced up-converted fluorescence mechanism.

Fig. 9, top image, shows the PNS-G0 complex with Zn2+ emission fluorescence at about 535 nm in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 and emission at 450 nm (bottom image) in DMSO solvent (excited under 800 nm wavelength light), which represents about an 85 nm shift. The most probable explanation is the influence of solvent. The solvent effect parameters, including polarity, dielectric constant (ε0), refractive index (n), and orientation polarizability (Δf), influence the PNS-G0 complex fluorescence emission. The different dipole–dipole actions of the solvent provide different environments for the PNS-G0, including changed excited-states and dipole moments, to give different fluorescence emission peaks. The solvent also transfers energy to the PNS-G0 under the external light excitation, which might influence the fluorescence emission positions. The influence of different solvent effects of water and DMSO on the PNS-G0 + Zn2+ complex at 800 nm excitation wavelength may cause the 85 nm shifts of up-converted fluorescence emission peaks shown in Fig. 9 (top image).

2.4 Imine (C[double bond, length as m-dash]N) isomerization mechanism

Consider the mechanism illustrated in Fig. 11 showing the Schiff base imine (C[double bond, length as m-dash]N) isomerization with the formation of the single/double bond resonance structures, which release the energy of excited molecular states through non-radiative processes at the microscopic scale, and give no or weak fluorescence at the macroscopic scale. However, if the C[double bond, length as m-dash]N isomerization is inhibited by complexing with metal ions through coordination bonds, then the imine C[double bond, length as m-dash]N can enhance fluorescence emission ratios. PNS-G0 with imine C[double bond, length as m-dash]N bonds were used in this mechanism to probe for Zn2+. The probe PNS-G0 gave weak fluorescence before complexing with Zn2+. When detecting the Zn2+, the imine (C[double bond, length as m-dash]N), hydroxyl (HO–), and amines of PAMAM form coordination bonds with the ions. Then the C[double bond, length as m-dash]N_Zn_O parts with naphthalene give strong fluorescence emission. The imine (C[double bond, length as m-dash]N) is locked by coordination bonds with Zn2+, and then the naphthalene (as donor) and imine (C[double bond, length as m-dash]N) (as acceptor) form a D–A (push–pull type) dipole emission unit. This dipole unit is a coplanar conjugated group and can enhance nonlinear optical properties, such as two-photon, up-converted fluorescence emission. These interactions cause the up-converted fluorescence emission of PNS-G0 and the turn-on effect when probing Zn2+ ions.
image file: c4ra01758g-f11.tif
Fig. 11 Schiff-base imine (C[double bond, length as m-dash]N) isomerization inhibition mechanism.

2.5 Half-organic quantum dots (HOQDs)

PNS-G0 matched Zn2+ gave strong fluorescence emission, which was different from other metal ions complexed with PNS-G0. The fluorescence spectra in Fig. 3, 4, and 10 prove that PNS-G0 has good selectivity for Zn2+. These show the PNS-G0 complexes with Zn2+ to form half-inorganic (Zn ions) + half-organic (hydroxyl-naphthalene imine) structures, which are unique fluorescence emission groups that are different from other metal ions complexed with PNS-G0.

Fig. 4, 8, and 9 show that the sharp emission peaks in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 are different from the spectra in DMSO solvent. Fig. 4 and 8 show peaks ranging from 455 nm to 475 nm, about 20 nm wide. The PNS-G0 gave enhanced emission at these narrow ranges. For example, Fig. 4 shows one sharp peak from 12[thin space (1/6-em)]500 to 24[thin space (1/6-em)]000 a.u., an increase of 11[thin space (1/6-em)]500 a.u. The excitation wavelength also influences the sharp shape of emission peaks. At the excitation wavelength of 400 nm (Fig. 4 and 8), the complex gave sharp peaks with range about 20 nm (455 nm to 475 nm); at the excitation wavelength of 800 nm (Fig. 9), the complex gave sharp peaks with a range of about 10 nm (530 nm to 540 nm).

The PNS-G0 + Zn2+ complex (Fig. 12) has a ZnO unit. ZnO is a traditional, typical quantum dot.32 The sharp type of emission peaks also appeared in the photoluminescence spectrum of the ZnO–PAMAM–G3 nanocomposite dispersed in water after excitation at 350 nm, at room temperature.33 These sharp peaks may be associated with PAMAM, Zn element, water solvent, the excitation wavelength, and even the nano states of molecules or particles. The narrow emission peaks of this PNS-G0 + Zn2+ complex are similar with the narrow emission band characteristics of fluorescence quantum dots (QDs). PNS-G0 + Zn2+ has sharp peaks in the fluorescence emission spectra—with narrow emission peaks—and the PNS-G0 complex with Zn2+ has inorganic metal ions and organic hydroxyl + imine ligands. So this PNS-G0 + Zn2+ complex has organic conjugated parts and ZnO QDs parts, which are firstly called half-organic quantum dots (HOQDs).


image file: c4ra01758g-f12.tif
Fig. 12 The half-organic quantum dots (HOQDs) of C[double bond, length as m-dash]N_Zn_O.

The HOQDs gave different sharp peaks shapes and different emission positions under different excitation wavelengths or solvents. Solvent effect and the environment of complex molecules distributed in solution co-act on the formation of sharp peaks. The complex gave sharp emission peaks in mixed solvent H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 and gave smooth emission peaks in DMSO solvent, which shows that the environment and states of the complex influence the sharp shape of emission peaks. The solvent environment, including polarity, dielectric constant (ε0), refractive index (n), and orientation polarizability (Δf), influence the peaks' shapes. The states of the complex might be the other influential element. The PNS-G0 was well dissolved in DMSO, but it was difficult to dissolve in water. Then the H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 mixed solvent system was used to dissolve PNS-G0 to test the fluorescence properties in the simulated biological environment. The PNS-G0 with Zn2+ formed the ZnO unit complex in this mixed solvent. The ZnO connected with C[double bond, length as m-dash]N and naphthalene, giving these HOQDs the narrow emission bands characteristic; then the peaks' shapes were sharp in the mixed H2O–DMSO = 99[thin space (1/6-em)]:[thin space (1/6-em)]1 solvent.

PNS-G0 has primary amine, amide, imine, and hydroxyl groups that can all take part in the complex, so the complex molar ratios with Zn2+ were difficult to show clearly. Fig. 13 gives the predicted structures, which were used for quantum chemical calculations. These kinds of HOQDs (Fig. 13: P-02 and P-03) all have inorganic metal ions and organic parts, which gave fluorescence emission, similar to traditional QDs. The HOQDs have organic parts that can give these QDs more organic characters, such as solubility, reactivity, biocompatibility, and more chemically modifiable properties. The PNS-G0 + Zn2+ complex has Zn2+_imine fluorescence emission groups that can be used as probe labels. The PAMAM cores also can emit fluorescence under some situations, and can package nanoparticles or drugs for delivery. These give (C[double bond, length as m-dash]N_Zn_O) HOQDs a diverse range of applications.


image file: c4ra01758g-f13.tif
Fig. 13 The end parts of PNS-G0 and PNS-G0 + Zn2+ complex P-01 shows the end parts of PNS-G0; P-02 shows the predicted end parts of PNS-G0 + Zn2+ (ligand number 2); P-03 shows the predicted end parts of PNS-G0 + Zn2+ (ligand number 4).

The HOQDS have organic parts and inorganic parts. PNS-G0 + Zn2+ HOQDs have Zn ions as the inorganic part, and imine conjugated with hydroxyl–naphthalene cycle as the organic part. The QDs have the main characteristic of quantum confinement effect (QCE). HOQDs have quantum confinement effect determined from the Bohr radius compared with the size of in/organic quantum parts. Fig. 13 shows the end parts of PNS-G0 (P-01) and PNS-G0 + Zn2+ (P-02 and P-03).

Fig. 14–16 show the molecular orbits of P-01, P-02, and P-03. The calculation used the time-dependent density functional theory (TD_DFT) b3lyp/6-31g methods of the Gaussian 09 software package.34Tables 1–3 list the absorption and emission spectra of P-01, P-02, and P-03 in gas phase vacuum calculated by the TD_DFT b3lyp/6-31g method. S0 → S1 are the absorption states; S1 → S0 are the emission states.


image file: c4ra01758g-f14.tif
Fig. 14 The molecular orbit clouds and contour figures (HOMO−1, HOMO, LUMO and LUMO+1) of P-01 (end parts of PNS-G0).

image file: c4ra01758g-f15.tif
Fig. 15 The molecular orbit clouds and contour figures (HOMO−1, HOMO, LUMO and LUMO+1) of P-02 (end parts of PNS-G0 + Zn2+, ligand number 2).

image file: c4ra01758g-f16.tif
Fig. 16 The molecular orbit clouds and contour figures (HOMO−1, HOMO, LUMO and LUMO+1) of P-03 (end parts of PNS-G0 + Zn2+, ligand number 4).
Table 1 The absorption and emission spectra of P-01 in gas phase vacuum calculated by TD_DFT b3lyp/6-31g method
Electron transitiona Energy (eV) Calculated wavelength (nm) Main transition configuration Oscillator strength f
a S0 → S1 are the absorption states of P-01; S1 → S0 is the emission state of P-01.
S0 → S1 2.857 433.93 HOMO → LUMO: −0.62210 0.2138
HOMO−1 → LUMO: 0.11833
HOMO−3 → LUMO: 0.31203
S0 → S1 2.940 421.68 HOMO → LUMO:0.30241 0.0794
HOMO−1 → LUMO: 0.43256
HOMO−2 → LUMO: 0.33605
HOMO−3 → LUMO: 0.32663
S0 → S1 3.124 396.87 HOMO → LUMO+2: 0.17661 0.0395
HOMO−1 → LUMO: −0.28034
HOMO−2 → LUMO: 0.58927
HOMO−3 → LUMO: −0.17738
S1 → S0 2.857 433.93 LUMO → HOMO: 0.12852 0.2138


Table 2 The absorption and emission spectra of P-02 in gas phase vacuum calculated by TD_DFT b3lyp/6-31g methoda
Electron transitiona Energy (eV) Calculated wavelength (nm) Main transition configuration Oscillator strength f
a S0 → S1 are the absorption states of P-02; S1 → S0 is the emission state of P-02.
S0 → S1 0.220 5624.95 HOMO → LUMO: −1.03080 0.0049
S0 → S1 1.967 630.12 HOMO → LUMO+1: 0.93489 0.0036
HOMO−1 → LUMO: 0.18387
HOMO−1 → HOMO: −0.26499
S0 → S1 2.113 586.72 HOMO−1 → HOMO: 0.61623 0.0013
HOMO−4 → HOMO: −0.12701
HOMO → LUMO+1: 0.31794
HOMO−1→LUMO: −0.67219
HOMO−4 → LUMO: 0.13763
S1 → S0 0.220 5624.95 LUMO → HOMO: 0.26503 0.0049


Table 3 The absorption and emission spectra of P-03 in gas phase vacuum calculated by TD_DFT b3lyp/6-31g methoda
Electron transitiona Energy (eV) Calculated wavelength (nm) Main transition configuration Oscillator strength f
a S0 → S1 are the absorption states of P-03; S1 → S0 is the emission state of P-03.
S0 → S1 1.321 938.47 HOMO−2 → LUMO: −0.14309 0.0167
HOMO → LUMO: 0.69075
S0 → S1 2.297 539.66 HOMO−1 → LUMO: 0.70362 0.0460
S0 → S1 2.318 534.72 HOMO → LUMO+1: 0.70463 0.0030
S1 → S0 1.321 938.47 LUMO → HOMO: −0.13198 0.0167


These HOQDs have two Bohr radius directions: one is the Bohr radius of Zn ions expended to free space; the other is the imine naphthalene hydroxyl organic conjugated part. The former is the same as the traditional QD Bohr radius. The latter is a π-conjugated organic cycle, which expends the electronic cloud by real atom parts that can be seen in the real atom part Bohr radius. The free space expended Bohr radius and real organic conjugated part Bohr radius form the characteristic HOQDs properties. So the HOQDs might become a new type of QDs to be given more attention.

The P-01 HOMO, LUMO contour figures show that the electronic clouds were distributed at the close side of molecules. The P-02 and P-03 HOMO, LUMO contour figures show the electronic clouds at the Zn2+ parts were distributed far away from the Zn atom, and gave a Bohr radius value of about 10 Bohr (B). The C[double bond, length as m-dash]N_Zn_O six-member cycle form the main part of the HOQDs, which gave size values of about ≈6 Bohr (D). Thus HOQDs obey the D < 2B rule, which shows the quantum confinement effect.

3. Experimental section

PNS-G0 was synthesized from PAMAM-G0 (Fig. 17) and 2-hydroxy-1-naphthaldehyde by Schiff base reaction. PNS-G0 was reacted with the different metal ions (Zn2+, Ca2+, Mg2+, Pb2+, Cd2+, Mn2+, Fe3+, Fe2+, Ni2+, Cu2+, Co2+, Cr3+, Ag+, H+, NH4+, Na+, K+) in DMSO and water + DMSO (99[thin space (1/6-em)]:[thin space (1/6-em)]1) mixed solvent to test the selectivity by UV-vis absorption and fluorescence spectra. PNS-G0 was tested with the different concentrations of Zn2+ ions for sensitivity and quantitative detection. The (up-converted) fluorescence was tested for the turn-on effects of probing for Zn2+.
image file: c4ra01758g-f17.tif
Fig. 17 The reaction of PAMAM-G0 molecules.

3.1 The synthesis of polyamides-amine (PAMAM-G0) dendrimers35

Methyl acrylate (MA) and ethylenediamine (EDA) were used as substrates. The Michael-addition of amine groups in EDA to MA under 50 °C in methanol solution affords the dendritic product of −0.5 generation (G) with ester groups terminated. The amidation of the terminal ester groups of the −0.5 G dendrimer from dissolving in methanol solution by excessive EDA under 50 °C affords the 0 G dendrimer with terminal amine groups. Distillation of exceeded EDA under reduced pressure gives the purified 0 G dendrimer. The PAMAM dendrimers are shown as a yellow, ropy liquid.

(PAMAM-G0) IR (KBr) cm−1: 3325 (NH), 2961, 1647 (CO), 1563, 1465, 1307, 995, 687. MS (m/z): (calculate for: 518.6 [M + 2H]2+); found: 518.3 [M + 2H]2+.

3.2 The synthesis of PNS-G0

The synthesis of 3,3′,3′′,3′′′-(ethane-1,2-diylbis(azanetriyl))tetrakis(N-(2-(((2-hydroxyl naphthalen-1-yl)methylene)amino)ethyl)propan amide) (PNS-G0) is described as follows.

PAMAM(G0), 0.52 g (1 mmol); 2-hydroxy-1-naphthaldehyde, 0.9 g (5 mmol, excessive); and anhydrous sodium sulfate, 0.5 g; were added into 30 mL methanol, refluxed for 2 h and cooled to room temperature. The solution was filtered, and the filtrate was washed by methanol three times, followed by vacuum drying. The result was a yellow brown powder of about 1 g, with a yield of 85%.

IR (KBr) cm−1: 3273 and 3064 (NH), 2936, 2818, 1629 (C[double bond, length as m-dash]N), 1549, 1495, 1447, 1354, 1210, 1184, 1139, 1104, 998, 835, 739, 505. 1H NMR (300 MHz, DMSO-d6, δ): δ 8.95 (4H, m, H–C[double bond, length as m-dash]N), 8.14 (4H, s, Ar–H), 8.00 (8H, d, J = 6.0 Hz, Ar–H), 7.71 (4H, d, N–H), 7.60 (4H, m, J = 6.0 Hz, Ar–H), 7.38 (4H, t, J = 6.0 Hz, Ar–H), 7.15 (4H, t, J = 6.0 Hz, Ar–H), 6.71 (4H, m, HO–), 3.08–4.01 (24H, m, C–H), 2.17–2.27 (12H, m, C–H). 13C NMR (300 MHz, DMSO-d6, δ): δ 176.9, 171.4, 160.2, 159.5, 137.0, 136.8, 134.1, 128.8, 127.8, 125.2, 125.1, 122.3, 122.1, 118.6, 118.4, 105.8, 50.3, 49.0, 40.3, 40.0, 39.7, 39.5, 39.2, 38.9, 38.6; HRMS (m/z): (calculate for: 1134.3480 [M + H]+); found: 1134.348 [M + H]+.

3.3 Methods

PNS-G0 solutions were prepared in H2O–DMSO = (99[thin space (1/6-em)]:[thin space (1/6-em)]1) at 1 × 10−6 mol L−1 or in DMSO at 5 × 10−5 mol L−1. The different metal ion solutions were added to the PNS-G0 solutions to keep a 1 × 10−4 mol L−1 ion concentration and form the complex solutions, which were used for quantitative testing. The [Zn2+] gradient concentration solutions were prepared and added into PNS-G0 solutions for qualitative detection. For the optical test: the solutions were placed in a quartz color dish, and the fluorescence emission spectra were tested using a linear light source spectrofluorometer. The up-converted fluorescence emission experiments were taken by FluoroMax-4 spectrofluorometer using a xenon excitation source, continuous output, ozone-free lamp (150 W) at room temperature (∼20 °C). The excitation wavelength was set at 800 nm, and test ranges were set from 200 nm to 850 nm, using a 3 nm slit.

4. Conclusions

PNS-G0 can realize up-converted fluorescence turn-on effect for high selectively qualitative and quantitative probing of Zn2+. This dendrimer probe is based on the Schiff base imine (C[double bond, length as m-dash]N) isomerization inhibition mechanism. The use of organic solvent and the nearly two times enhanced selective signal in water–DMSO mixed solvents show the dyes need to be further modified and improved. The C[double bond, length as m-dash]N_Zn_O fluorescence emission parts were called half-organic quantum dots (HOQDs), which show similar and new properties compared to traditional QDs. The PNS-G0 and its Zn2+ complex have potential applications in biological cell imaging, metal ions analytical chemistry, and up-converted optical physics areas. Furthermore, the HOQDs can be used to explore a wide variety of new applications.

Acknowledgements

The National Natural Science Foundation of China (no. 61178057) and the Scientific Research Foundation of Graduate School of Southeast University (no. YBPY1209) were greatly appreciated for financial support.

Notes and references

  1. B. L. Vallee and K. H. Falchuk, Physiol. Rev., 1993, 73, 79 CAS.
  2. C. J. Frederickson, J. Y. Koh and A. I. Bush, Nat. Rev. Neurosci., 2005, 6, 449 CrossRef CAS PubMed.
  3. (a) P. Saluja, V. K. Bhardwaj, T. Pandiyan, S. Kaur, N. Kaur and N. Singh, RSC Adv., 2014, 4, 9784 RSC; (b) S. Y. Jiao, L. L. Peng, K. Li, Y. M. Xie, M. Z. Ao, X. Wang and X. Q. Yu, Analyst, 2013, 138, 5762 RSC; (c) H. Nouri, C. Cadiou, L. M. L. Daku, A. Hauser, S. Chevreux, I. D. Olivier, F. Lachaud, R. Ternane, M. T. Ayadi, F. Chuburu and G. Lemercier, Dalton Trans., 2013, 42, 12157 RSC; (d) Y. Mikata, A. Ugai, K. Yasuda, S. Itami, S. Tamotsu, H. Konno and S. Iwatsuki, Chem. Biodiversity, 2012, 9(9), 2064 CrossRef CAS PubMed; (e) S. Comby, S. A. Tuck, L. K. Truman, O. Kotova and T. Gunnlaugsson, Inorg. Chem., 2012, 51(19), 10158 CrossRef CAS PubMed.
  4. (a) H. M. Kim, M. S. Seo, M. J. An, J. H. Hong, Y. S. Tian, J. H. Choi, O. Kwon, K. J. Lee and B. R. Cho, Angew. Chem., Int. Ed., 2008, 47, 5167 CrossRef CAS PubMed; (b) C. J. Chang, J. Jaworski, E. M. Nolan, M. Sheng and S. J. Lippard, Proc. Natl. Acad. Sci. U. S. A., 2004, 101, 1129 CrossRef CAS PubMed; (c) H. M. Kim and B. R. Cho, Acc. Chem. Res., 2009, 42, 863 CrossRef CAS PubMed.
  5. J. S. Wu, W. M. Liu, J. C. Ge, H. Y. Zhang and P. F. Wang, Chem. Soc. Rev., 2011, 40, 3483 RSC.
  6. (a) R. M. Manez and F. Sancenon, Chem. Rev., 2003, 103, 4419 CrossRef PubMed; (b) B. Valeur and I. Leray, Coord. Chem. Rev., 2000, 205, 3 CrossRef CAS; (c) Z. C. Xu, J. Yoon and D. R. Spring, Chem. Soc. Rev., 2010, 39, 1996 RSC; (d) J. S. Kim and D. T. Quang, Chem. Rev., 2007, 107, 3780 CrossRef CAS PubMed.
  7. C. Lodeiro and F. Pina, Coord. Chem. Rev., 2009, 253, 1353 CrossRef CAS PubMed.
  8. A. P. de Silva, H. Q. N. Gunaratne, T. Gunnlaugsson, A. J. M. Huxley, C. P. McCoy, J. T. Rademacher and T. E. Rice, Chem. Rev., 1997, 97, 1515 CrossRef CAS PubMed.
  9. Q. Zhao, F. Y. Li and C. H. Huang, Chem. Soc. Rev., 2010, 39, 3007 RSC.
  10. (a) K. E. Sapsford, L. Berti and I. L. Medintz, Angew. Chem., Int. Ed., 2006, 45, 4562 CrossRef CAS PubMed; (b) H. J. Carlson and R. E. Campbell, Curr. Opin. Biotechnol., 2009, 20, 19 CrossRef CAS PubMed.
  11. W. Rettig and R. Lapouyade, Fluorescence probes based on twisted intramolecular charge transfer (TICT) states and other adiabatic photoreactions, in Topics in Fluorescence Spectroscopy, Probe Design and Chemical Sensing, ed. J. R. Lakowicz, Plenum Press, New York, 1994, vol. 4, p. 109 Search PubMed.
  12. J. F. Callan, A. P. de Silva and D. C. Magri, Tetrahedron, 2005, 61, 8551 CrossRef CAS PubMed.
  13. (a) Y. N. Hong, J. W. Y. Lam and B. Z. Tang, Chem. Commun., 2009, 4332 RSC; (b) M. Wang, G. X. Zhang, D. Q. Zhang, D. B. Zhu and B. Z. Tang, J. Mater. Chem., 2010, 20, 1858 RSC.
  14. M. Y. Berezin and S. Achilefu, Chem. Rev., 2010, 110, 2641 CrossRef CAS PubMed.
  15. 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 CrossRef CAS PubMed.
  16. (a) W. Liu, L. Xu, R. Sheng, P. Wang, H. Li and S. Wu, Org. Lett., 2007, 9, 3829 CrossRef CAS PubMed; (b) D. Ray and P. K. Bharadwaj, Inorg. Chem., 2008, 47, 2252 CrossRef CAS PubMed; (c) V. Chandrasekhar, P. Bag and M. D. Pandey, Tetrahedron, 2009, 65, 9876 CrossRef CAS PubMed; (d) H. S. Jung, K. C. Ko, J. H. Lee, S. H. Kim, S. Bhuniya, J. Y. Lee, Y. Kim, S. J. Kim and J. S. Kim, Inorg. Chem., 2010, 49, 8552 CrossRef CAS PubMed; (e) M. Suresh, A. K. Mandal, S. Saha, E. Suresh, A. Mandoli, R. D. Liddo, P. P. Parnigotto and A. Das, Org. Lett., 2010, 12, 5406 CrossRef CAS PubMed; (f) Z. Li, M. Yu, L. Zhang, M. Yu, J. Liu, L. Wei and H. Zhang, Chem. Commun., 2010, 46, 7169 RSC.
  17. (a) K. E. Sapsford, L. Berti and I. L. Medintz, Angew. Chem., Int. Ed., 2006, 45(28), 4562 CrossRef CAS PubMed; (b) W. I. Lee, Y. Bae and A. J. Bard, J. Am. Chem. Soc., 2004, 126, 8358–8359 CrossRef CAS PubMed; (c) D. J. Wang and T. Imae, J. Am. Chem. Soc., 2004, 126, 13204–13205 CrossRef CAS PubMed; (d) D. C. Wu, Y. Liu, C. B. He and S. H. Goh, Macromolecules, 2005, 38(24), 9906–9909 CrossRef CAS; (e) Y. F. Fan, Y. G. Fan and Y. N. Wang, J. Appl. Polym. Sci., 2007, 106(3), 1640–1647 CrossRef CAS; (f) P. L. Wang, X. Wang, K. Meng, S. Hong, X. Liu, H. Cheng and C. C. Han, J. Polym. Sci., Part A: Polym. Chem., 2008, 46(10), 3424–3428 CrossRef CAS.
  18. M. J. Murcia, D. L. Shaw, E. C. Long and C. A. Naumann, Opt. Commun., 2008, 281, 1771 CrossRef CAS PubMed.
  19. (a) C. B. Murray, D. J. Norris and M. G. Bawendi, J. Am. Chem. Soc., 1993, 115, 8706 CrossRef CAS; (b) A. P. Alivisatos, Science, 1996, 271, 933 CAS; (c) C. Donega, S. G. Hickey, S. F. Wuister, D. Vanmaekelbergh and A. Meijerink, J. Phys. Chem. B, 2003, 107, 489 CrossRef; (d) Z. A. Peng and X. Peng, J. Am. Chem. Soc., 2001, 123, 183 CrossRef CAS; (e) R. Cohen, L. Kronik, A. Shanzer, D. Cahen, A. Liu, Y. Rosenwaks, J. K. Lorenz and A. B. Ellis, J. Am. Chem. Soc., 1999, 121, 10545 CrossRef CAS.
  20. W. L. Wilson, P. F. Szajowski and L. E. Brus, Science, 1993, 262, 1242 CAS.
  21. Springer Handbook of Condensed Matter and Materials Data, ed. W. Martienssen and H. Warlimont, Springer Verlag, 2005, vol. 18, p. 1120 Search PubMed.
  22. W. C. W. Chan, D. J. Maxwell, X. Gao, R. E. Bailey, M. Han and S. Nie, Curr. Opin. Biotechnol., 2002, 13, 40 CrossRef CAS.
  23. W. H. Liu, H. S. Choi, J. P. Zimmer, E. Tanaka, J. V. Frangioni and M. Bawendi, J. Am. Chem. Soc., 2007, 129, 14530 CrossRef CAS PubMed.
  24. R. C. Somers, R. M. Lanning, P. T. Snee, A. B. Greytak, R. K. Jain, M. G. Bawendi and D. G. Nocera, Chem. Sci., 2012, 3, 2980 RSC.
  25. (a) A. A. Hajaj, A. Moquin, K. D. Neibert, G. M. Soliman, F. M. Winnik and D. Maysinger, ACS Nano., 2011, 5(6), 4909 CrossRef PubMed; (b) D. A. Geraldo, E. F. Duran-Lara, D. Aguayo, R. E. Cachau, J. Tapia, R. Esparza, M. J. Yacaman, F. D. Gonzalez-Nilo and L. S. Santos, Anal. Bioanal. Chem., 2011, 400, 483 CrossRef CAS PubMed.
  26. (a) K. A. Connors, Binding Constants: The Measurement of Molecular Complex Stability, John Wiley and Sons, New York, 1987 Search PubMed; (b) S. J. K. Pond, O. Tsutsumi, M. Rumi, O. Kwon, E. Zojer, J. L. Brédas, S. R. Marder and J. W. Perry, J. Am. Chem. Soc, 2004, 126, 9291–9306 CrossRef CAS PubMed; (c) H. M. Kim, C. Jung, B. R. Kim, S. Y. Jung, J. H. Hong, Y. G. Ko, K. J. Lee and B. R. Cho, Angew. Chem., Int. Ed., 2007, 46, 3460–3463 CrossRef CAS PubMed.
  27. F. Auzel, Chem. Rev., 2004, 104, 139 CrossRef CAS PubMed.
  28. (a) W. L. Peticolas, J. P. Goldsborough and K. E. J. Rieckhoff, Phys. Rev. Lett., 1963, 10, 43 CrossRef CAS; (b) W. L. Peticolas and K. E. J. Rieckhoff, Chem. Phys., 1963, 39, 1347 CAS; (c) P. M. Rentzepis, C. J. Mitschele and A. C. Saxman, Appl. Phys. Lett., 1970, 17, 122 CrossRef CAS PubMed; (d) A. C. Seldon, Nat. Phys. Sci., 1971, 229, 210 CrossRef; (e) J. A. Giordmaine, P. M. Rentzepis, S. L. Shapiro and K. W. Wecht, Appl. Phys. Lett., 1967, 11, 216 CrossRef CAS PubMed.
  29. (a) G. S. He, J. D. Bhawalkar, C. F. Zhao, C. K. Park and P. N. Prasad, Opt. Lett., 1995, 20, 2393 CrossRef CAS; (b) G. S. He, C. F. Zhao, J. D. Bhawalkar and P. N. Prasad, Appl. Phys. Lett., 1995, 67, 3703 CrossRef CAS PubMed; (c) W. Denk, J. H. Strickler and W. W. Webb, Science, 1990, 248, 73 CAS; (d) G. S. He, L. S. Tan, Q. D. Zheng and P. N. Prasad, Chem. Rev., 2008, 108, 1245 CrossRef CAS PubMed.
  30. F. Auzel, Chem. Rev., 2004, 104, 139 CrossRef CAS PubMed.
  31. W. E. Case and M. E. Koch, J. Lumin., 1990, 45, 351 CrossRef CAS.
  32. (a) W. F. Hsieh, H. C. Hsu, W. J. Liao, H. M. Cheng, K. F. Lin, W. T. Hsu and C. J. Pan, Adv. Seri. Appl. Phys., 2011, 6, 253–267 CAS; (b) T. Yatsui and M. Ohtsu, Reza Kenkyu, 2011, 39(3), 184–187 CAS; (c) C. C. Yang, Key Eng. Mater., 2010, 444, 133–162 CrossRef CAS; (d) X. Sun, C. G. Zhou, Mi. Xie, H. T. Sun, T. Hu, F. Y. Lu, S. M. Scott, S. M. George and J. Lian, J. Mater. Chem. A, 2014, 2(20), 7319–7326 RSC.
  33. R. O. Moussodia, L. Balan, C. Merlin, C. Mustind and R. Schneider, J. Mater. Chem., 2010, 20, 1147 RSC.
  34. 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, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision A.02, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
  35. D. A. Tomalia, H. Baker, J. Dewald, M. Hall, G. Kallos, S. Martin, J. Roeck, J. Ryder and P. Smith, Polym. J., 1985, 17, 117 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: 1H NMR, 13C NMR spectra, HRMS, and IR of PNS-G0 listed in Fig. S1–S4. MS and IR of PAMAM-G0 listed in Fig. S5, S6. See DOI: 10.1039/c4ra01758g

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.