The Effect of Amine-Functional Group on Heavy Metal Ion ‎Detection of a Cu-Based Metal-Organic Framework

Document Type : Research Paper

Authors

1 Department of Chemistry, Iran University of Science and Technology

2 Department of Chemistry, Iran University of Science and Technology, Tehran 16846-13114, ‎Iran‎

Abstract

Owing to their enriched host-guest chemistry and high porosity, metal-organic frameworks ‎‎(MOFs) applied extensively in the detection of a large variety of hazardous chemicals, ‎especially metal ions, using different instrumental methods, such as photoluminescence. In this ‎study, two pillar-layered MOFs, [Cu2(BDC)2(DABCO)] and its amine-functionalized ‎isostructure, [Cu2(NH2-BDC)2(DABCO)], were successfully prepared to highlight the amino ‎group role in the metal ions sensing via exploiting photoluminescence-based method. These ‎materials could recognize copper(II) cations efficiently by changing the luminescent response ‎in less than 20 minutes. With increasing the concentration of Cu(II) solution, first, the ‎luminescent response changes occurred intensely, then equilibrium was reached and no more ‎changes were observed. In the presence of other metal cations, the sensors almost recognized ‎the analyte selectively. At the end of the experimental tests, the reasonable mechanism of ‎sensing was proposed using FTIR spectroscopy and PXRD patterns. They confirmed that the ‎introduction of analyte does not collapse the structure, so just the chemical interaction between ‎the walls of pores and Cu(II) ions are responsible for the purposed application.‎

Keywords


INTRODUCTION

Copper is one of the heavy metals on the earth that plays vital roles in various fields such as biological, environmental, and chemical systems [1]. Although copper is an essential nutrient for all living organisms in trace concentrations, the accumulation of this element in biological systems can, due to its toxicity, lead to cancer, Parkinson’s, neurodegenerative, Alzheimer’s, and Huntington’s diseases, as well as genetic disorders like Menkes and Wilson’s illnesses [2]. On the other hand, copper is widely used in different areas including building construction, renewable energy, electrical equipment, industrial machinery, and so on. Consequently, it is readily released in the environment as a result of industrial processes. It is for this reason that finding convenient, rapid, highly sensitive, and inexpensive copper ion recognition techniques has become a challenging issue in environmental protection [3].

There are various analytical sensing techniques, such as voltammetry, spectrophotometry, and atomic absorption spectroscopy among which fluorometry-based techniques are in the center of attention because of their special features such as facilitated detection and manipulation, and high sensitivity. Accordingly, fluorescent chemosensors are known as promising candidates for the recognition of metal ions or other small molecules based on their good reliability, high sensitivity, and real-time detection [4-7] .

Metal-organic frameworks (MOFs), constructed from metal ions or clusters and organic linkers through coordination bonds, have drawn immense attention and enthusiasm in the class of porous solids [8]. Diverse structures and topologies, ultrahigh porosity, huge surface area, as well as the availability of in-pore functionality and outer surface modification are some of the advantages of MOFs [9, 10] resulting in enormous applications including heterogeneous catalysis [11, 12], sensing [13, 14], proton conduction [15], biomedical imaging [16], selective adsorption [17, 18], gas storage [19], drug delivery [20], and so forth.

Recently, some studies were conducted to detect Cu2+ in wastewater. For instance, Chen et al. reported a lanthanide-based MOF, [Eu(pdc)1.5(DMF)].(DMF)0.5(H2O)0.5 using pyridine-3,5-dicarboxylate possessing pyridyl sites with KSV value of 89.4 M-1 [21]. In another study, Wang and coworkers combined sulfonated poly(arylene ether nitrile) (SPEN) and MOF-5 to obtain a mixed matrix membrane for Cu2+ sensing with KSV value of 110 M-1 [3]. In 2018, synthesis of a novel MOF formulated as [Zn(BDC)0.5(Tz)].DMF.CH3OH was presented for Cu2+ recognition leading to a relatively high sensitivity with KSV being equal to 9.2×103 M-1 [1]. Therefore, finding a stronger platform for sensitive recognition of Cu2+ in the wastewater is intensely necessary.

One of the best ways of achieving an effective MOF-based sensor is to select π-conjugated linkers along with Lewis basic sites to provide luminescence features and the efficient binding sites. Thus, pendant amine and amide group, pyridyl nitrogen, hydroxyl group, and others can play the role of active binding sites for acquiring better performance [22]. Herein, we choose [Cu2(BDC)2(DABCO)], (Cu(BDC); H2BDC = 1,4-benzene dicarboxylic acid, DABCO = 1,4-diazabicyclo[2.2.2]octane) and its NH2-modified isostracture, [Cu2(NH2-BDC)2(DABCO)] (Cu(NH2BDC); NH2BDC = amino-1,4-benzene dicarboxylic acid), to investigate their sensing behavior toward Cu2+ cations according to fluorescence-based methods. Then we check their selectivity and sensitivity during the recognition process. At the end, we will conclude that the amino group in a luminescent MOF can develop its efficiency in the metallic cation sensing application based on its electron-donating features.

EXPERIMENTAL SECTION

Chemicals, reagents, and apparatus

All reagents and precursors were purchased from Sigma-Aldrich, Merck, and other companies. Cu(NO3)3·3H2O and 1,4-benzenedicarboxylic acid (H2BDC), 1,4-diazabicyclo[2.2.2]octane (DABCO) were used to synthesize Cu(BDC). To obtain the NH2-modified MOF, amino-1,4-benzene dicarboxylic acid (NH2BDC) was used instead of H2BDC. N,N-Dimethylformamide (DMF) was exploited as the solvent to purify both MOFs. Solutions of Cd2+, Zn2+, As3+, Al3+, Cu2+, Pb2+, Co2+, Ni2+, and Fe3+ were prepared from CdCl2·2.5H2O, Zn(NO3)3·6H2O, NaAsO2, Al(NO3)3·9H2O, Cu(NO3)2·3H2O, Pb(NO3)2, Co(NO3)2·6H2O, NiCl2·6H2O, and Fe(NO3)3·9H2O, respectively. The infrared spectra were recorded on a Nicolet Fourier transform IR, Nicolet 100 spectrometer in the range of 500–4000 cm-1 via utilizing the KBr tablet. X-ray powder diffraction (XRD) measurements were performed using a Philips X’pert diffractometer with monochromated Cu-kα radiation (λ = 1.54056A). The simulated XRD powder pattern based on single-crystal data was prepared using Mercury software. The fluorescence experiments were carried out at room temperature on an F4500 fluorescence spectrometer (Hitachi, Japan) with a photomultiplier voltage of 700 V, a scan speed of 2400 nm.min-1, an excitation slit width of 10 nm, an emission slit width of 20 nm, and a 380 nm optical filter. The fluorescent emission spectra were recorded in the wavelength range of 300-700 nm upon excitation at 360 nm for Cu(BDC) and 350 nm for Cu(NH2-BDC) (Fig. 4).

Preparation of MOFs Cu(BDC) and Cu(NH2-BDC)

Cu(NO3)2·3H2O (0.241 g, 1 mmol), 1,4-benzene dicarboxylic acid (H2BDC; 0.166 g, 1 mmol), and of 1,4-diazabicyclo[2.2.2]octane (DABCO; 0.56 g, 0.5 mmol), were dissolved in 15 mL DMF. The resulting solution was transferred to a 100 mL Teflon-lined autoclave to heat at 120 °C for 2 days, and then it was gradually cooled to room temperature. The solid product Cu(BDC) obtained was washed in DMF to remove the impurities, then dried at 80 °C for at least 12 h. Cu(NH2-BDC) was synthesized following the synthesis procedure for Cu(BDC) using NH2-BDC instead of the unmodified one.

Luminescent experiments

Cu(BDC) suspensions for fluorescence experiments were prepared by dispersing 1 mg of Cu(BDC) powder in 10 mL of ethanol under ultrasonication (80 W) for 10 min. To a 1 cm×1 cm quartz cell, Cu(BDC) suspension (4 mL) and 20 µL of Cu2+ solution (10-3 M) were sequentially added. The mixtures were then used for fluorescence measurements. The fluorescence data were collected after 2 min. All steps were also repeated for Cu(NH2-BDC) suspensions.

RESULTS AND DISCUSSION

Characterization of Cu(BDC) and Cu(NH2-BDC)

As the nature of the organic linker impacts the luminescence behavior of MOF toward the target analytes, we decided to assess the effect of the amine group on the sensing process. Accordingly, 1,4-benzene dicarboxylic acid (H2BDC) and its modified derivation, NH2BDC, were used to prepare [Cu2(BDC)2(DABCO)] and [Cu2(NH2-BDC)2(DABCO)], respectively [23]. In both of them, the 2D square grids of Cu2(BDC)2 and Cu2(NH2-BDC)2 dimers are pillared by DABCO molecules to form a 3D framework (Fig. 1). The PXRD patterns of as-synthesized Cu(BDC) and Cu(NH2-BDC) exhibit good agreement with the reported pattern (Fig. 2). FT-IR spectra of Cu(BDC) is shown in Fig. 2b. The peak at 3450 cm-1 corresponds to O-H vibrations of uncoordinated water molecules and the bands at 2950–2850 cm-1 correspond to weak C-H bending vibration (aliphatic). The peaks with significant intensity at around 1600 cm-1 and 1400 cm-1 are attributed to νas(C-O) and νs(C-O) vibration of carboxylate groups. As shown in the paddlewheel structure (Fig. 1), the carboxylate ion is coordinated with two Cu nodes as a bridging bidentate linker in a syn-syn configuration leading to a wavenumber in the 1600−1630 cm-1 region. Cu(NH2-BDC) also exhibits a broadband near the 3000 cm-1 corresponding to NH2 groups.

Photoluminescence feature of the MOFs

Since the selection of organic linkers with active sites plays a key role in obtaining a series of suitable MOFs for molecular sensing, the fluorescent properties of Cu(BDC), as well as free ligands of H2BDC and DABCO were measured at room temperature to check the ligands’ role in the sensing performance. As can be seen in Fig. 5a, the emission spectrum of 440 nm for Cu(BDC) is similar to that of the free H2BDC linker demonstrating the π-π* transfer from the H2BDC ligand. The emission of the MOF is weaker than H2BDC which can be attributed to the ligand-to-metal charge transfer (LMCT) effect. The highest occupied molecular orbital (HOMO) is likely to be the π bonding orbital from the aromatic rings, and the lowest unoccupied molecular orbital (LUMO) is related mainly to the Cu–O (carboxylate) π*-antibonding orbital, which is often localized on the metal centers [7].

On the other hand, the presence of free amine in the pores of Cu(NH2-BDC) is its most significant structural feature which highlights its molecular sensing behavior. When Cu(NH2-BDC) was dispersed in methanol solution, it shows an emission peak at 440 nm with a weaker luminescent intensity compared to its NH2-functionalized linker, as shown in Fig. 5b. This is due to the ligand-to-metal charge transfer (LMCT) effect upon making coordination bond between NH2-BDC with Cu-O clusters to form the framework [6].

Detection of metal ions

In order to check the metal ions detection, 1 mg powder of Cu(BDC) and Cu(NH2-BDC) were separately immersed in 10 mL ethanol solutions containing M(NO3)x (10-3 M; M = Cd2+, Mn2+, Ni2+, Co2+, Cu2+, Fe3+, Al3+, Zn2+, Pb2+). Before luminescence measurements, the suspensions were placed in the ultrasonic bath for 5 min to ensure uniform dispersion. The collected data illustrated that both MOFs detected Cu2+ most sensitively as can be seen in Fig 6. For this reason, we follow our experiments using Cu(II) ion solutions.

To further study, the fluorescence titrations were carried out by changing the concentration of Cu2+ to examine luminescence behaviors. The finely dispersed samples of both MOFs were exposed to different concentrations of Cu2+ ions. As can be seen in Fig. 7a and b, when the Cu(II) concentration increased from 0 to 140 μM, the emission intensity of both compounds gradually changed with a good linear relationship between the emission intensity and the concentration of Cu(II) in the range of 0-100 μM of Cu(II) (Fig. 7c,d). It can be concluded that the change in luminescence intensity depends on the concentration of the analyte.

The change in fluorescence follows the Stern–Völmer (SV) equation: I0/I=1+KSV[M], where I0 and I refer to the luminescence intensity for both MOFs in the absence and presence of metal cations, respectively; [M] is the metal concentration; and KSV is the Stern-Völmer constant. It was also concluded that the SV plot of both MOFs towards Cu(II) is nearly linear at a low concentration range, but subsequently deviates from linearity and bends upwards at higher concentrations (Fig. 7b). KSV value of the amine-modified framework is about 90 times more than that of the unmodified one, suggesting more sensitivity toward Cu(II) due to the high affinity of active amine sites to bind with the analyte. Moreover, the calculated detection limit values of Cu(BDC) and Cu(NH2-BDC) were 76.6×10-6 and 57.51×10-6, respectively, confirming that our proposed MOFs could act as the sensitive sensors toward the target analyte.

Moreover, the time-response characteristic of the Cu(BDC) and Cu(NH2-BDC) sensor toward Cu(II) were tested at λex=360 and 350 nm, respectively. As demonstrated in Fig. 8a, by increasing time, the PL intensity increased for the Cu(BDC)@Cu2+, reaching equilibrium during 20 min. On the other hand, Cu(NH2-BDC)@Cu2+ quenches the luminescence emission faster than the non-functionalized MOF (Fig. 8b). It can happen because of forming faster interactions between amine sites of the latter framework with the analyte, compared to the non-modified framework.

As the other heavy metal ions commonly exist in the real sample of Cu(II), we checked the potential of both MOFs to detect the target cation in the presence of 250 μM of coexisting cations. As can be seen in Fig. 9a, in the presence of other competitive cations, Cu(BDC) also enhances the luminescence intensity; therefore, we can claim that it could selectively recognize Cu(II). In Fig. 9b, the selectivity of Cu(NH2-BDC) is assessed. It is very encouraging that Cu(NH2-BDC) exhibits higher anti-interference compared with other coexisting metal ions except for Fe(III), which further underlines the high selectivity of Cu(NH2-BDC). But in the case of Fe(III) presence, because of the high affinity of N atoms of organic linkers of Cu(NH2-BDC) toward binding with Fe(III), it will interfere in the Cu(II) detection process.

Based on FT-IR and PXRD patterns of the frameworks before and after the sensing process, we attempted to provide a plausible explanation for our observations. Generally, the fluorescence changes caused by MOFs can arise from the following reasons: the collapse or change of the structure, the ionic exchange, and the competitive photon absorption between the analyte and frameworks [24]. The PXRD measurements illustrated that no observable changes occurred in the sensing process, ruling out the possibility of the collapse of the Cu(BDC) framework (Fig. 3a). Therefore, the luminescence intensity changes were not affected by the framework collapse. Furthermore, all FT-IR spectra are similar, confirming the retention of the structure in the recognition procedure (Fig 3b).

For the amino-functionalized MOF, a different phenomenon has occurred. As can be seen in the PXRD patterns (Fig 4a), some peaks disappeared during the sensing process. Moreover, FT-IR spectra demonstrate that the amine group was affected by the analyte, perhaps because of Lewis acid-base interactions between Cu(II) cations and these active sites (Fig. 4b). When copper cations diffuse into the pores of MOF, the amine electrons are transferred to the receptor. As a result, the luminescence is quenched.

Real sample tests

For evaluation of the proposed amine-based sensor for the measurement of Cu2+ in real environmental samples (industrial wastewater of Tehran city), the framework was applied to obtained samples. The resultant data was collected in Table 1. As can be seen, the recovery was satisfactory (>90%), confirming that Cu(NH2-BDC) was applicable for the detection of copper in natural water samples.

CONCLUSION

Free Lewis basic sites in the pores of MOFs play pivotal roles in the recognition process of small Lewis acidic molecules like metal ions; therefore, modification of organic linkers with these groups could enhance sensing potential. Here we utilized the amine-functionalized dicarboxylate ligand in the construction of luminescent MOF as a stable fluorescent sensor for the recognition of Cu(II). All experiments were also performed by unmodified parent MOF to highlight the effect of NH2 presence in detection potential. The amine-modified luminescent sensor exhibits high selectivity for sensing Cu(II) in ethanol solution with significant sensitivity (Ksv ~ 89075 M−1), a detection limit of 57.51×10-6 M, and quick response speed (< 5 min). The present results could propose a facile route to design and synthesize functionalized metal-organic frameworks with fluorescent sensing performance. This sensitivity toward Cu2+ may be due to the relatively high Kf of [Cu(NH3)4]2+ (1.1×1013) compared to the other assessed heavy metals.

ACKNOWLEDGMENTS

The support of this investigation by Iran University of Science and Technology, Iran National Science Foundation: INSF and Iran’s National Elites Foundation is gratefully acknowledged.

CONFLICTS OF INTEREST

The authors announce that there are no conflicts of interest.

 

  1.  

    1. Zhou E-L, Qin C, Tian D, Wang X-L, Yang B-X, Huang L, et al. A difunctional metal–organic framework with Lewis basic sites demonstrating turn-off sensing of Cu2+ and sensitization of Ln3+. Journal of Materials Chemistry C. 2018;6(29):7874-9.
    2. Ye J, Zhao L, Bogale RF, Gao Y, Wang X, Qian X, et al. Highly Selective Detection of 2,4,6-Trinitrophenol and Cu2+Ions Based on a Fluorescent Cadmium-Pamoate Metal-Organic Framework. Chemistry - A European Journal. 2014;21(5):2029-37.
    3. Wang L, Zheng P, Zhang W, Xu M, Jia K, Liu X. Detection of Cu2+ metals by luminescent sensor based on sulfonated poly(arylene ether nitrile)/ metal-organic frameworks. Materials Today Communications. 2018;16:258-63.
    4. Dong Y, Zhang H, Lei F, Liang M, Qian X, Shen P, et al. Benzimidazole-functionalized Zr-UiO-66 nanocrystals for luminescent sensing of Fe 3+ in water. Journal of Solid State Chemistry. 2017;245:160-3.
    5. Fan K, Bao S-S, Nie W-X, Liao C-H, Zheng L-M. Iridium(III)-Based Metal–Organic Frameworks as Multiresponsive Luminescent Sensors for Fe3+, Cr2O72–, and ATP2– in Aqueous Media. Inorganic Chemistry. 2018;57(3):1079-89.
    6. Wang M, Guo L, Cao D. Metal-organic framework as luminescence turn-on sensor for selective detection of metal ions: Absorbance caused enhancement mechanism. Sensors and Actuators B: Chemical. 2018;256:839-45.
    7. Farahani YD, Safarifard V. Highly selective detection of Fe3+, Cd2+ and CH2Cl2 based on a fluorescent Zn-MOF with azine-decorated pores. Journal of Solid State Chemistry. 2019;275:131-40.
    8. Ke F, Jiang J, Li Y, Liang J, Wan X, Ko S. Highly selective removal of Hg 2+ and Pb 2+ by thiol-functionalized Fe 3 O 4 @metal-organic framework core-shell magnetic microspheres. Applied Surface Science. 2017;413:266-74.
    9. Zhao X-L, Sun W-Y. The organic ligands with mixed N-/O-donors used in construction of functional metal–organic frameworks. CrystEngComm. 2014;16(16):3247.
    10. Esrafili L, Safarifard V, Tahmasebi E, Esrafili MD, Morsali A. Functional group effect of isoreticular metal–organic frameworks on heavy metal ion adsorption. New Journal of Chemistry. 2018;42(11):8864-73.
    11. Gharib M, Esrafili L, Morsali A, Retailleau P. Solvent-assisted ligand exchange (SALE) for the enhancement of epoxide ring-opening reaction catalysis based on three amide-functionalized metal–organic frameworks. Dalton Transactions. 2019;48(24):8803-14.
    12. Hu M-L, Safarifard V, Doustkhah E, Rostamnia S, Morsali A, Nouruzi N, et al. Taking organic reactions over metal-organic frameworks as heterogeneous catalysis. Microporous and Mesoporous Materials. 2018;256:111-27.
    13. Asha KS, Bhattacharjee R, Mandal S. Complete Transmetalation in a Metal-Organic Framework by Metal Ion Metathesis in a Single Crystal for Selective Sensing of Phosphate Ions in Aqueous Media. Angewandte Chemie International Edition. 2016;55(38):11528-32.
    14. Yao C, Xu Y, Xia Z. A carbon dot-encapsulated UiO-type metal organic framework as a multifunctional fluorescent sensor for temperature, metal ion and pH detection. Journal of Materials Chemistry C. 2018;6(16):4396-9.
    15. Kim S, Joarder B, Hurd JA, Zhang J, Dawson KW, Gelfand BS, et al. Achieving Superprotonic Conduction in Metal–Organic Frameworks through Iterative Design Advances. Journal of the American Chemical Society. 2018;140(3):1077-82.
    16. Lan G, Ni K, Veroneau SS, Song Y, Lin W. Nanoscale Metal–Organic Layers for Radiotherapy–Radiodynamic Therapy. Journal of the American Chemical Society. 2018;140(49):16971-5.
    17. Huo J-B, Xu L, Chen X, Zhang Y, Yang J-CE, Yuan B, et al. Direct epitaxial synthesis of magnetic Fe3O4@UiO-66 composite for efficient removal of arsenate from water. Microporous and Mesoporous Materials. 2019;276:68-75.
    18. Yu C, Shao Z, Liu L, Hou H. Efficient and Selective Removal of Copper(II) from Aqueous Solution by a Highly Stable Hydrogen-Bonded Metal–Organic Framework. Crystal Growth & Design. 2018;18(5):3082-8.
    19. Jin W-G, Chen W, Xu P-H, Lin X-W, Huang X-C, Chen G-H, et al. An Exceptionally Water Stable Metal-Organic Framework with Amide-Functionalized Cages: Selective CO2/CH4Uptake and Removal of Antibiotics and Dyes from Water. Chemistry - A European Journal. 2017;23(53):13058-66.
    20. Lei B, Wang M, Jiang Z, Qi W, Su R, He Z. Constructing Redox-Responsive Metal–Organic Framework Nanocarriers for Anticancer Drug Delivery. ACS Applied Materials & Interfaces. 2018;10(19):16698-706.
    21. Chen B, Wang L, Xiao Y, Fronczek FR, Xue M, Cui Y, et al. A Luminescent Metal-Organic Framework with Lewis Basic Pyridyl Sites for the Sensing of Metal Ions. Angewandte Chemie International Edition. 2009;48(3):500-3.
    22. Xiang Z, Fang C, Leng S, Cao D. An amino group functionalized metal–organic framework as a luminescent probe for highly selective sensing of Fe3+ ions. Journal of Materials Chemistry A. 2014;2(21):7662.
    23. Tan K, Nijem N, Canepa P, Gong Q, Li J, Thonhauser T, et al. Stability and Hydrolyzation of Metal Organic Frameworks with Paddle-Wheel SBUs upon Hydration. Chemistry of Materials. 2012;24(16):3153-67.
    24. Besheli ME, Rahimi R, Farahani YD, Safarifard V. A porous Ni-based metal-organic framework as a selective luminescent probe to Fe3+ metal ion and MeOH. Inorganica Chimica Acta. 2019;495:118956.