Lanthanide Organic Framework as a Reversible Luminescent Sensor

Mar 14, 2019 - ... lanthanide organic framework, {Eu(BTB)DMF}n (Eu-MOF; DMF = N,N-dimethylformamide), with two one-dimensional channels was obtained, ...
0 downloads 0 Views 4MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Lanthanide Organic Framework as a Reversible Luminescent Sensor for Sulfamethazine Antibiotics Kui Ren,† Shun-Hua Wu, Xiao-Feng Guo,† and Hong Wang* Key Laboratory of Analytical Chemistry for Biology and Medicine (Ministry of Education), College of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, China

Inorg. Chem. Downloaded from pubs.acs.org by EAST CAROLINA UNIV on 03/14/19. For personal use only.

S Supporting Information *

ABSTRACT: A water-stable two-dimensional lanthanide organic framework, {Eu(BTB)DMF}n (Eu-MOF; DMF = N,N-dimethylformamide), with two one-dimensional channels was obtained, and its structure was characterized. With changes in the amount of LiOH·H2O, different sizes of {Eu(BTB)DMF}n were synthesized. The prepared EuMOF powder is easy to disperse in water and exhibits typical Eu red emission. The fluorescence properties showed that Eu-MOF can detect sulfamethzine (SMZ) with high sensitivity and selectivity. Finally, the as-synthesized Eu-MOF was successfully used for the detection of SMZ in surface water by a standard addition method.

1. INTRODUCTION

However, investigations on the detection of antibiotics are not many.21−23 In this work, we chose a classical MOF ligand, 1,3,5-tris(4carboxyphenyl)benzene (H3BTB), to prepare Ln-MOF. H3BTB has a large molecular size to build a large porous framework and also has a conjugated electron-donor system to sensitize the rare-earth metal. Then, the Eu-BTB Ln-MOF {Eu(BTB)DMF}n (Eu-MOF; DMF = N,N-dimethylformamide) was obtained with two different sizes of 20−80 and 2 μm. Powder X-ray diffraction (PXRD) patterns of the EuMOF crystal (20−80 μm) and powder (2 μm) are consistent with the simulated data of the Eu-MOF crystal. The assynthesized Eu-MOF powder exhibits high water stability in a wide pH range from 3 to 9 and is easy to disperse uniformly. The luminescence experiment showed that Eu-MOF can recognize sulfamethazine (SMZ) with high sensitivity and selectivity among various antibiotics.

Antibiotics are commonly used in medicine, agriculture, and livestock farming and have been regarded as an important species of organic contaminants in water. The abuse of antibiotics has led to massive accumulation of antibiotics in food and animals because most antibiotics do not easily degrade naturally.1−5 Most antibiotics enter the human body via wastewater or food polluted by wastewater. Thus, monitoring these specific contaminants in water is necessary. Until now, several instrumental methods, such as liquid chromatography−tandem mass spectrometry,6 capillary electrophoresis,7 mass spectrometry,8 Raman spectroscopy,9 and ion-mobility spectrometry,10 have been reported for the detection of antibiotics. However, these analytical methods have some disadvantages like inconvenience, expensiveness, and the need for complex equipment and well-trained personnel. Therefore, the evolution of feasible, convenient, reliable, and inexpensive detection methods for antibiotics is a major concern. Metal−organic frameworks (MOFs) are new functional materials composed of metal ions and organic ligands to form an extended crystal skeleton. MOFs are considered to be superior platforms for detection/sensing applications because of their unusual optical and electronic properties.11 Lanthanide organic frameworks (Ln-MOFs) are MOFs using lanthanide ions as vertexes, which have unique optical properties arising from 4f electrons through the antenna effect, including large Stokes shifts, stable emission bands, and long luminescence lifetimes.12,13 With these unique properties, Ln-MOFs are thought to be ideal luminescent sensors. Many MOF-based luminescent sensors have been successfully synthesized and are mainly used for the detection of nitrobenzene, acetone, Zn2+, Cu2+, Fe3+, Al3+, and other small molecules/cations.14−20 © XXXX American Chemical Society

2. EXPERIMENTAL SECTION 2.1. Reagents. All antibiotics were purchased from commercial suppliers and stored in a cool dark place (SMZ is photosensitive). H3BTB and Eu(NO3)3·6H2O were purchased from Energy Chemical (Shanghai, China). Other chemical reagents were purchased from various commercial suppliers and used without further purification. 2.2. Materials and Physical Measurements. PXRD was carried out on a Smart Lab 9 kW Advance X-ray diffractometer and Rigaku Miniflex600 in the 5−50° 2θ range using Cu Kα radiation. Scanning electron microscopy (SEM) images were recorded on a Zeiss Merlin Compact microscope. All of the samples were mounted on carbon tape and coated with gold prior to measurement. Fourier transform infrared (FTIR) spectroscopy was performed in the range 4000−400 cm−1 on a Thermo IS10 spectrophotometer using KBr pellets. Thermogravimetric analysis (TGA) was measured on a Setaram Received: November 24, 2018

A

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. Crystal structure of Eu-MOF. (a) Coordinated environments of Eu3+ and BTB3−. (b) 2D framework of Eu-MOF along the a direction. (c) Illustration of the bilayer structure. (d) 2D framework of Eu-MOF along the b direction. Eu3+ is connected to a 1D chain by a carboxyl bridge. SETSYS 16 at a heating rate of 10 °C min−1 from 25 to 800 °C under a nitrogen atmosphere. Fluorescence was measured with a RF-6000 fluorescence spectrophotometer at room temperature. A fluorescence lifetime study was carried out on an Edinburgh Instruments FLS920 instrument. Inductively coupled plasma analysis (ICP) atomic emission spectrometry was obtained by Leeman Laboratories Prodigy7. The nitrogen adsorption isotherms were measured on ASAP 2020 and ASAP 2460 surface area size analyzers at 77 K. 2.3. Synthesis for {Eu(BTB)DMF}n. Eu-MOFs were synthesized by the solvothermal method according to the literature.24−26 Eu(NO3)3·6H2O (0.2 mmol, 40.1 mg) in deionized H2O (3 mL) and H3BTB (0.2 mmol, 87.6 mg) in DMF (4 mL) were mixed to form a light-yellow solution. The reaction was carried out in a 25 mL Teflon-lined stainless steel autoclave at 150 °C for 72 h under autogenous pressure and then cooled to room temperature. After washing with DMF and ethanol three times, respectively, needleshaped colorless Eu-MOF crystals were obtained (yield: 62%). EuMOF powder was synthesized by the same procedures, except for the addition of LiOH·H2O (0.2 mmol, 8.4 mg), based on the previous report,26 and a colorless Eu-MOF powder with strong red fluorescence was obtained (yield: 82%).

anion, one coordinated DMF, and one crystallographically independent Eu3+ ion. An eight-coordinated Eu3+ ion exhibits a distorted square-antiprismatic geometry, which is completed by seven carboxylic oxygen atoms from the BTB3− ligand and one terminal DMF molecule (Figure 1a). The Eu−O (oxygen atom of the carboxylate group in BTB3−) bond lengths are in the range of 2.3118−2.5663 Å, and adjacent Eu3+ ions are connected by carboxylate groups into the europium chain with a Eu···Eu length of 4.0958 Å (Figure 1c). The structural of EuMOF is that the binuclear europium atom acts as a 6connected node and the ligand acts as a 4-connected node, and this two-dimensional (2D) skeleton can be abstracted into a 3,6-connected topological net. The 2D framework has two types of one-dimensional (1D) channels with diameters of about 5 and 4 Å along their directions (Figure S1). All of these parameters are in close agreement with those previously reported for Ln-MOFs.24−27 3.1.2. TGA of Eu-MOFs. TGA was performed to evaluate the stability of Eu-MOFs. The thermal behavior of Eu-MOF displayed a weight loss of 23% before 190 °C, which corresponded to the loss of DMF and noncoordinated water molecules in the channel. The framework was still stable up to 520 °C, which demonstrated the excellent thermal stability of Eu-MOF (Figure S2). 3.1.3. FTIR Spectra of Eu-MOFs. FTIR spectra of Eu-MOF and the ligand are listed in Figure S3. Compared to H3BTB, the IR bands at 2665 and 2549 cm−1 (OCOOH−H) disappeared and that at 1692 cm−1 (OCOOHO) was red shifted to 1662 cm−1 because of coordination.

3. RESULTS AND DISCUSSION 3.1. Characterization of Eu-MOFs. 3.1.1. Single-Crystal X-ray Diffraction of Eu-MOFs. To verify the structure of EuMOF, single-crystal X-ray diffraction analysis was performed on the Eu-MOF crystal, and an ORTEP view of Eu-MOF is given in Figure 1. Detailed crystal data, data collection, and refinement details are given in Table S1. Single-crystal X-ray diffraction indicates that Eu-MOF crystallizes in the P21/c space group. The asymmetric unit consists of one BTB3− B

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. Characterizations of Eu-MOF. (a) SEM images for Eu-MOF crystal (b) SEM images for Eu-MOF powder; (c) PXRD patterns of synthesized Eu-MOFs and simulated from Eu-MOF crystal data; (d) Fluorescence spectrum of Eu-MOF powder at room temperature.

3.1.4. PXRD of Eu-MOFs. As shown in Figure 2c, the PXRD patterns of the Eu-MOF crystal and Eu-MOF powder are well consistent with the simulated one from the Eu-MOF crystal data (CCDC 1816956). The PXRD pattern substantiates the good phase purity of the as-synthesized Eu-MOFs. Slight differences in the PXRD pattern between Eu-MOF synthesized with a DMF−H2O mixed solvent and simulated from the single-crystal model most likely result from the water in the pores of the experimental sample, which is not accounted for in the simulated pattern.28 We optimized the reaction conditions after numerous experiments and found that, if pure DMF was used to synthesize Eu-MOF, the PXRD pattern will match better than that synthesized with a DMF−H2O mixed solvent. However, it is impossible to remove H2O in Eu(NO3)3·6H2O. 3.1.5. SEM Images of Eu-MOFs. SEM images clearly show the Eu-MOF crystal with a size of 20−80 μm (Figure 2a) and the Eu-MOF powder with a size of 2 μm (Figure 2b). 3.2. Fluorescent Response. The fluorescence spectrum of Eu-MOF at room temperature is given in Figure 2d. The excitation spectrum of Eu-MOF exhibits a broad band with a maximum at 310 nm (Figure S4). Under excitation at 287 nm, Eu-MOF displays four well-resolved emission peaks at 594, 614, 651, and 699 nm, which are attributed to 5D0 → 7FJ (J = 1−4) transitions of the Eu3+ ion, respectively. The peak at 574 nm is a double-frequency scattering peak, whose wavelength is 2 times the excitation wavelength (287 nm). Such a phenomenon is very common in fluorescence spectra.29 The strongest 5D0 → 7F2 transition with a bright-red luminescence is attributed to magnetic-dipole-induced transitions. This indicates the occurrence of efficient energy transfer from the sensitizer (H3BTB) to doped Eu3+. In addition, it is worth

noting that the broad emission of the ligand cannot be observed in the emission spectrum of Eu-MOF, which indicates the occurrence of “antenna effects”. The excitation spectrum collected at the wavelength of 614 nm displays a broad band at 250−320 nm. We also studied the luminescence decay curves of a pure ligand and Eu-MOF. Eu-MOF has a long lifetime (τ = 725.5438 μs) compared to a free ligand (τ = 16.2751 ns; Figure S5). In order to study the colloidal stability of the Eu-MOF solution, we monitored the intensity of the 5D0 → 7F2 transition (614 nm) of Eu-MOF standing still from 0 to 60 min by 5 min steps, which revealed that the Eu-MOF powder has a certain antiprecipitation ability (Figure S6). To examine the sensing ability of Eu-MOF, several commonly used antibiotics were investigated including SMZ, nitrofurazone (NFZ), nitrofurantoin (NFT), metronidazole (MDZ), ronidazole (RDZ), dimetridazole (DTZ), sulfadiazine (SDZ), ornidazole (ODZ), chloramphenicol (CAP), thiamphenicol (THI), and penicillin (PCL). The quenching percentages of Eu-MOF after the addition of a certain amount of various antibiotics are depicted in Figure 3. SMZ exhibits a drastic quenching efficiency on fluorescence intensity. The order of the quenching efficiency is SMZ > NFZ > CAP > SDZ > NFT > MDZ > DTZ > RDZ > ODZ > PCL > THI. The higher quenching efficiency for SMZ indicates that Eu-MOF has good application prospects in detecting SMZ in water. To further prove that the fluorescence quenching was induced by SMZ, a concentration-dependent study was carried out with fluorescence of Eu-MOF in the presence of SMZ. It was observed that the fluorescence intensity of Eu-MOF progressively decreases upon the incremental addition of SMZ (Figure 4a). A parallel control experiment of the C

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

the quenching efficiencies were always within an acceptable range from pH 3.0 to 9.0 (Figure S8). The nitrogen adsorption isotherms showed Eu-MOFs held at high Brunauer−Emmett− Teller (BET) surface areas (1103.25 and 988.78 m2 g−1) after exposure to pH 3.0 and 9.0 water solutions for 3 days, which showed no significant difference from the as-synthesized EuMOF (945.12 m2 g−1; Figure S9). PXRD indicated that the frameworks were intact from pH 2.0 to 10.0 (Figure S10). Additionally, the PXRD patterns of Eu-MOF after 72 h of immersion in various solvents revealed that Eu-MOF had excellent solvent stability (Figure S11). 3.3. Interference. SMZ was added to the Eu-MOF suspensions, which contained other coexisting antibiotics, and the luminescence intensities of the resulting mixtures were further examined (Figure 3). Most antibiotics have a certain degree of fluorescence quenching without SMZ. After the addition of SMZ into the mixture of the sensor and other antibiotics, serious quenching could be observed. To gain more insight into whether SMZ can be selectively probed in the presence of different concentrations of SDZ or CAP, we carried out some competition experiments by the sequential addition of SDZ or CAP, followed by SMZ into Eu-MOF, and recorded the corresponding emission spectra.31 The initial addition of different SDZ or CAP showed a slight quenching intensity, but after the addition of a SMZ solution, an effective quenching was observed, which proved that Eu-MOF had special selectivity for SMZ (Figures S12 and S13). All of the results verified the outstanding antiinterference ability of EuMOF. 3.4. Mechanism. To demonstrate the source of the high selectivity toward SMZ, the quenching mechanism was investigated. First, PXRD was performed, and the results showed that Eu-MOF still retained its crystal state after fluorescence detection of SMZ (Figure S14), indicating that the framework remained intact after the addition of SMZ, so quenching was not caused by collapse of the framework. Second, we studied the quenching process by measuring KSV. Usually, KSV increases with rising temperature in dynamic quenching, and the reverse occurs in static quenching. We calculated KSV of Eu-MOF at different temperatures. As shown in Table S2, KSV was observed to increase with increasing temperature. UV absorption spectra also indicated that no complex was formed between SMZ and Eu-MOF (Figure S15). So, the possible quenching mechanism was dynamic. Then we speculated that fluorescence quenching may be

Figure 3. Fluorescence intensity of Eu-MOF immersed in 0.1 mM of different antibiotics and a mixture of competing antibiotics (0.1 mM) with SMZ (0.1 mM).

fluorescence response of a free ligand (H3BTB) in the presence of Eu ions upon the addition of SMZ was also carried out. As shown in Figure S7, compared with Eu-MOF, the characteristic peak of the 5D0 → 7F2 transition (614 nm) of a ligand mixed with Eu ions is very weak, and there is a strong ligand peak at about 360 nm, which indicates the disappearances of the “antenna effects”. Both peaks have no obvious fluorescence intensity change upon the addition of SMZ even when the concentration of SMZ is as high as 300 μM, which demonstrated the specificity of Eu-MOF as a sensing probe. Usually, the quenching efficiency could be quantitatively illustrated by the Stern−Volmer (S−V) equation: I0/I = KSV[C] + 1. In the above equation, I0 and I are the fluorescence intensities in the absence and presence of the quenching agent, respectively, [C] is the quenching agent (SMZ) concentration, and KSV is the quenching constant.30 As shown in Figure 4b, a good linear relationship (R2 = 0.996) between the analytical signal of I 0 /I and the SMZ concentration was obtained from 0 to 80 μM. The KSV value is 4.598 × 104 M−1, revealing a strong quenching effect to EuMOF luminescence. The limit of detection for SMZ is estimated to be 0.6554 μM based on a signal-to-noise ratio (S/ N) of 3. pH tolerance and solvent stability are key factors in evaluating the reliability of chemical sensors. The effects of the pH on Eu-MOF and Eu-MOF with SMZ were studied, and

Figure 4. (a) Fluorescence response (λex = 287 nm; slit widths = 3 nm/3 nm) of Eu-MOF (0.16 mg L−1) upon the addition of SMZ [(0−1) × 10−4 mol L−1]. Spectra were acquired in an ethanol−H2O (6:4, v/v) borate buffer (50 mM; pH 7) solution. (b) Stern−Volmer plot (I0/I) of Eu-MOF immersed in different concentrations of SMZ. D

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry caused by electron transfer from the conduction band (CB) of Eu-MOF to the lowest unoccupied molecular orbital (LUMO) of the electron-deficient analyte. Such electron transfer was a well-established quenching mechanism, in which the energy of the Eu-MOF CB should be higher than that of the LUMO of the analyte. We calculated the highest occupied molecular orbital (HOMO) and LUMO energies of related molecules. As shown in Figure S16, THI and PCL have higher LUMO energies than the free ligand, suggesting that there should be nearly no electron transfer from the ligand to THI and PCL. Also, from Figure 3, we can see that THI and PCL have the smallest impact on the fluorescence intensity of Eu-MOF. However, other antibiotics have well-matched LUMO energies. Thus, electron transfer should not be the only mechanism for selectivity. An inner-filter effect (IFE) is another theory for fluorescence quenching, which was caused by the overlap of the absorption spectrum of the analyte and the excitation/emission spectrum of the fluorophore.32 We recorded the UV/vis absorption spectra of Eu-MOF and various antibiotics. As shown in Figure S17, the wide absorption band from 250 to 290 nm of SMZ overlaps some range of absorption bands of Eu-MOF, and the absorption of SMZ in such a range is much stronger than those of other antibiotics. This implies that UV/vis absorption of SMZ from 250 to 290 nm may prevent excitation of Eu-MOF and induces a decrease or quenching of the fluorescence in the excitation from 250 to 290 nm (Figure 5). When the excitation

Figure 6. Reusability of the MOF implemented with 0.1 mM of a SMZ solution.

processes retained a high nitrogen uptake (about 340 cm3 g−1) and a high BET surface area (1036.66 m2 g−1; Figure S9). The excellent reusability of Eu-MOF could avoid sensor consumption and environmental pollution. 3.6. Detection of SMZ in Lake Water. In order to evaluate the practicality of the sensor on real samples, the standard addition and recovery were studied in a real water sample of East Lake (Wuhan, China). As shown in Table 1, the recoveries of SMZ were between 94% and 105.5%, indicating that the proposed method is practical for the detection of SMZ in real water samples. Table 1. Standard Addition and Recovery Test of SMZ in the Application of Lake Water

Figure 5. Possible mechanism of Eu-MOF sensing for SMZ.

wavelength changed to 320 nm (beyond the absorption of SMZ), the fluorescence intensity of Eu-MOF with the presence of SMZ hardly changed (Figure S18b), which proved the presence of IFE. 3.5. Reusability. Considering the cost of luminescent sensors, reusability is a decisive aspect for practical application. According to the above discussion, Eu-MOF was stable in water, which provides a possibility for recycling. Compared to the soluble luminescent sensors, an insoluble solid sensor has an intrinsic advantage for regeneration. The reusability of EuMOF was verified by the fluorescence intensity of Eu-MOF in a 0.1 mM SMZ aqueous solution and the fluorescence intensity after ultrasonic washing with ethanol three times. As shown in Figure 6, after five switching processes, the Eu-MOF sensor could retain the same sensing property without obvious changes. PXRD of the sample after five switching processes matched well with that of the as-synthesized Eu-MOF, which suggested that the Eu-MOF sensor retained the same framework structure (Figure S14). The ICP results of EuMOF and Eu-MOF after five switching processes (Table S3) showed that, before and after five switching processes, EuMOF had similar Eu concentrations, which verified the reusability of En-MOF as a reversible luminescent sensor for SMZ.20 In addition, the nitrogen adsorption isotherm experiment showed that the sample after five switching

sample

added(μM/L)

found(μM/L)

RSD (%; n = 3)

recovery (%)

1 2 3

10 20 30

9.4 21.1 28.9

7.9 2.1 4.0

94 105.5 96.3

4. CONCLUSION In summary, we have successfully synthesized a water-stable 2D luminescence framework, {Eu(BTB)DMF}n (Eu-MOF), which was well characterized by single-crystal X-ray diffraction, PXRD, SEM, TGA, and FTIR. Then, with the addition of LiOH·H2O in the synthesis procedure, a homogeneous EuMOF powder with the same structure was obtained. The proposed Eu-MOF powder exhibits high sensitivity for reversible sensing of SMZ in aqueous solution mainly based on IFE quenching. The proposed Eu-MOF has vast potential in monitoring SMZ in groundwater with a limit of detection of 0.6554 μM.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b03284. Additional text with further experimental details and additional figures (PDF) E

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Accession Codes

(12) Cui, Y.; Xu, H.; Yue, Y.; Guo, Z.; Yu, J.; Chen, Z.; Gao, J.; Yang, Y.; Qian, G.; Chen, B. A luminescent mixed-lanthanide metal-organic framework thermometer. J. Am. Chem. Soc. 2012, 134, 3979−3982. (13) Hu, Z.; Deibert, B. J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815−5840. (14) Chen, B.; Wang, L.; Zapata, F.; Qian, G.; Lobkovsky, E. B. A luminescent microporous metal-organic framework for the recognition and sensing of anions. J. Am. Chem. Soc. 2008, 130, 6718− 6719. (15) Shi, P. F.; Hu, H. C.; Zhang, Z. Y.; Xiong, G.; Zhao, B. Heterometal-organic frameworks as highly sensitive and highly selective luminescent probes to detect I(−) ions in aqueous solutions. Chem. Commun. 2015, 51, 3985−3988. (16) Xie, W.; Zhang, S. R.; Du, D. Y.; Qin, J. S.; Bao, S. J.; Li, J.; Su, Z. M.; He, W. W.; Fu, Q.; Lan, Y. Q. Stable luminescent metal-organic frameworks as dual-functional materials to encapsulate ln(3+) ions for white-light emission and to detect nitroaromatic explosives. Inorg. Chem. 2015, 54, 3290−3296. (17) Wang, Y.; Zhang, F.; Fang, Z.; Yu, M.; Yang, Y.; Wong, K. L. Tb(III) postsynthetic functional coordination polymer coatings on ZnO micronanoarrays and their application in small molecule sensing. J. Mater. Chem. C 2016, 4, 8466−8472. (18) Xu, H.; Cao, C. S.; Zhao, B. A water-stable lanthanide-organic framework as a recyclable luminescent probe for detecting pollutant phosphorus anions. Chem. Commun. 2015, 51, 10280−10283. (19) Zhao, B.; Kang, X. M.; Cheng, R. R.; Xu, H.; Wang, W. M. A Sensitive Luminescent Probe of Acetylacetone based on Zn-MOF with Six-Fold Interpenetration. Chem. - Eur. J. 2017, 23, 13289− 13293. (20) Xu, H.; Fang, M.; Cao, C. S.; Qiao, W. Z.; Zhao, B. Unique (3,4,10)-Connected Lanthanide-Organic Framework as a Recyclable Chemical Sensor for Detecting Al(3.). Inorg. Chem. 2016, 55, 4790− 4794. (21) Hou, S. L.; Dong, J.; Jiang, X. L.; Jiao, Z. H.; Wang, C. M.; Zhao, B. Interpenetration-Dependent Luminescent Probe in IndiumOrganic Frameworks for Selectively Detecting Nitrofurazone in Water. Anal. Chem. 2018, 90, 1516−1519. (22) Qin, Z.-S.; Dong, W.-W.; Zhao, J.; Wu, Y.-P.; Zhang, Q.; Li, D.S. A water-stable Tb(iii)-based metal−organic gel (MOG) for detection of antibiotics and explosives. Inorg. Chem. Front. 2018, 5, 120−126. (23) Wang, B.; Lv, X. L.; Feng, D.; Xie, L. H.; Zhang, J.; Li, M.; Xie, Y.; Li, J. R.; Zhou, H. C. Highly Stable Zr(IV)-Based Metal−Organic Frameworks for the Detection and Removal of Antibiotics and Organic Explosives in Water. J. Am. Chem. Soc. 2016, 138, 6204− 6216. (24) Zhai, B.; Xu, H.; Li, Z. Y.; Cao, C. S.; Zhao, B. A water-stable metal-organic framework: serving as a chemical sensor of PO 4 3− and a catalyst for CO 2 conversion. Sci. China: Chem. 2017, 60, 1328−1333. (25) Qin, J. H.; Ma, B.; Liu, X. F.; Lu, H. L.; Dong, X. Y.; Zang, S. Q.; Hou, H. Ionic liquid directed syntheses of water-stable Eu- and Tb-organic-frameworks for aqueous-phase detection of nitroaromatic explosives. Dalton. Trans. 2015, 44, 14594−14603. (26) Devic, T.; Serre, C.; Audebrand, N.; Marrot, J. A.; Férey, G. MIL-103, A 3-D Lanthanide-Based Metal Organic Framework with Large One-Dimensional Tunnels and A High Surface Area. J. Am. Chem. Soc. 2005, 127, 12788−12789. (27) Xu, H.; Hu, H. C.; Cao, C. S.; Zhao, B. Lanthanide organic framework as a regenerable luminescent probe for Fe(3+). Inorg. Chem. 2015, 54, 4585−4587. (28) Catarineu, N. R.; Schoedel, A.; Urban, P.; Morla, M. B.; Trickett, C. A.; Yaghi, O. M. Two Principles of Reticular Chemistry Uncovered in a Metal−Organic Framework of Heterotritopic Linkers and Infinite Secondary Building Units. J. Am. Chem. Soc. 2016, 138, 10826−10829. (29) Lakowicz, J. R. Principles of Fluorescence Spectroscopy; Springer, 2006; pp 37−56.

CCDC 1816956 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (H.W.). Tel.: +86-2787218924. Fax: +86-27-68754067. ORCID

Kui Ren: 0000-0002-5842-6624 Author Contributions †

These authors contributed equally to this manuscript.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Nature Science Foundation of China, Beijing, China (Grants 20835004, 31670370, and 21777126).



REFERENCES

(1) Baran, W.; Adamek, E.; Ziemianska, J.; Sobczak, A. Effects of the presence of sulfonamides in the environment and their influence on human health. J. Hazard. Mater. 2011, 196, 1−15. (2) Yin, F.; Dong, H.; Zhang, W.; Zhu, Z.; Shang, B. Antibiotic degradation and microbial community structures during acidification and methanogenesis of swine manure containing chlortetracycline or oxytetracycline. Bioresour. Technol. 2018, 250, 247−255. (3) Zhang, Q.; Ying, G.; Pan, C.; Liu, Y.; Zhao, J. Comprehensive Evaluation of Antibiotics Emission and Fate in the River Basins of China: Source Analysis, Multimedia Modeling, and Linkage to Bacterial Resistance. Environ. Sci. Technol. 2015, 49, 6772−6782. (4) Benito-Pena, E.; Urraca, J. B.; Moreno-Bondi, M. C. Quantitative determination of penicillin V and amoxicillin in feed samples by pressurised liquid extraction and liquid chromatography with ultraviolet detection. J. Pharm. Biomed. Anal. 2009, 49, 289−294. (5) Moreno-González, D.; Lara, F. J.; Jurgovská, N.; Gámiz-Gracia, L.; García-Campaña, A. M. Determination of aminoglycosides in honey by capillary electrophoresis tandem mass spectrometry and extraction with molecularly imprinted polymers. Anal. Chim. Acta 2015, 891, 321−328. (6) Toussaint, B.; Chedin, M.; Vincent, U.; Bordin, G.; Rodriguez, A. R. Determination of (fluoro)quinolone antibiotic residues in pig kidney using liquid chromatography−tandem mass spectrometry: I. Laboratory-validated method. J. Chrom. A 2005, 1088, 40−48. (7) Altria, K. D.; Chanter, Y. L. Validation of a capillary electrophoresis method for the determination of a quinolone antibiotic and its related impurities. J. Chrom. A 1993, 652, 459−463. (8) Zhang, X. H.; Deng, Y.; Zhao, M. Z.; Zhou, Y. L.; Zhang, X. X. Highly-sensitive detection of eight typical fluoroquinolone antibiotics by capillary electrophoresis-mass spectroscopy coupled with immunoaffinity extraction. RSC Adv. 2018, 8, 4063−4071. (9) Moros, J.; Laserna, J. J. New Raman-laser-induced breakdown spectroscopy identity of explosives using parametric data fusion on an integrated sensing platform. Anal. Chem. 2011, 83, 6275−6285. (10) Tabrizchi, M.; Ilbeigi, V. Detection of explosives by positive corona discharge ion mobility spectrometry. J. Hazard. Mater. 2010, 176, 692−696. (11) Lustig, W. P.; Mukherjee, S.; Rudd, N. D.; Desai, A. V.; Li, J.; Ghosh, S. K. Metal-organic frameworks: functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242−3285. F

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (30) Thomas, S. W. I.; Joly, G. D.; Swager, T. M. Chemical Sensors Based on Amplifying Fluorescent Conjugated Polymers. Chem. Rev. 2007, 107, 1339−1386. (31) Sanda, S.; Parshamoni, S.; Biswas, S.; Konar, S. Highly selective detection of palladium and picric acid by a luminescent MOF: a dual functional fluorescent sensor. Chem. Commun. 2015, 51, 6576−6579. (32) Li, P.; Hong, Y.; Feng, H.; Li, S. F. Y. An efficient “off−on” carbon nanoparticle-based fluorescent sensor for recognition of chromium(VI) and ascorbic acid based on the inner filter effect. J. Mater. Chem. B 2017, 5, 2979−2988.

G

DOI: 10.1021/acs.inorgchem.8b03284 Inorg. Chem. XXXX, XXX, XXX−XXX