Fluorescent Metal–Organic Framework (MOF) as a Highly Sensitive

Jan 8, 2018 - A Zn(II)-based fluorescent metal–organic framework (MOF) was synthesized and applied as a highly sensitive and quickly responsive chem...
0 downloads 5 Views 1MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Fluorescent Metal−Organic Framework (MOF) as a Highly Sensitive and Quickly Responsive Chemical Sensor for the Detection of Antibiotics in Simulated Wastewater Xian-Dong Zhu,*,‡,§ Kun Zhang,‡ Yu Wang,† Wei-Wei Long,‡ Rong-Jian Sa,§ Tian-Fu Liu,*,§ and Jian Lü*,†,§ †

Fujian Provincial Key Laboratory of Soil Environmental Health and Regulation, College of Resources and Environment, Fujian Agriculture and Forestry University, Fuzhou 350002, P.R. China ‡ School of Biological & Chemical Engineering, Anhui Polytechnic University, Wuhu 241000, P.R. China § State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, P.R. China S Supporting Information *

ABSTRACT: A Zn(II)-based fluorescent metal−organic framework (MOF) was synthesized and applied as a highly sensitive and quickly responsive chemical sensor for antibiotic detection in simulated wastewater. The fluorescent chemical sensor, denoted FCS-1, exhibited enhanced fluorescence derived from its highly ordered, 3D MOF structure as well as excellent water stability in the practical pH range of simulated antibiotic wastewater (pH = 3.0−9.0). Remarkably, FCS-1 was able to effectively detect a series of sulfonamide antibiotics via photoinduced electron transfer that caused detectable fluorescence quenching, with fairly low detection limits. Two influences impacting measurements related to wastewater treatment and water quality monitoring, the presence of heavy-metal ions and the pH of solutions, were studied in terms of fluorescence quenching, which was nearly unaffected in sulfonamide-antibiotic detection. Additionally, the effective detection of sulfonamide antibiotics was rationalized by the theoretical computation of the energy bands of sulfonamide antibiotics, which revealed a good match between the energy bands of FCS-1 and sulfonamide antibiotics, in connection with fluorescence quenching in this system.



antibiotics,13 the existing methods appear to be either too expensive or too time- and labor-consuming.14−17 Therefore, the development of effective methodologies for the fast and sensitive detection of antibiotics in aqueous systems is much more practical. Fluorescence-based chemical sensors (FCS) have attracted enormous research attention because of their fast responses, high sensitivities, and the ease of preparation.12,18,19 In this context, metal−organic frameworks (MOFs) have been recently advanced as viable platforms to detect trace amounts of pollutants present in water,20−23 because of their large specific surface areas, ultrahigh porosities, and tunable structures and functions.24−27 Fluorescent MOFs, in particular, have demonstrated their fast, convenient, and visually capable to detect heavy-metal ions and organic pollutants,28−32 taking advantage of their high color purity and the long lifetimes of their excited states. It is noteworthy that only a few fluorescent MOF materials have shown the ability to detect aromatic organic compounds in aqueous media,18,23 despite a large number of fluorescent MOF materials having been employed as promising chemical sensors. In this context, the detection of antibiotics in

INTRODUCTION Pharmaceutical antibiotics, which are currently being used worldwide in human therapy and the farming industry, and their transformation products are considered to be one of the most noticeable classes of environmental pollutants.1,2 Furthermore, they are often poorly metabolized and absorbed by humans and animals, resulting in the potential to develop and subsequently spread antibiotic-resistant bacteria (ARB) and antibiotic-resistance genes (ARGs) in the long term and have serious impacts on the ecosystem.3 Exposures to residues of antibiotics and their transformed products might cause a variety of adverse effects, including acute and chronic toxicity and microorganism antibiotic resistance.4,5 In view of increasing concerns related to the environment and the public,6 there is an urgent demand for the development of effective technologies to remove antibiotics from pharmaceutical wastewater. Previous research efforts have mainly been directed toward the discovery of viable absorbents7−9 for the elimination of antibiotics from aqueous systems. However, the detection of antibiotics in wastewater, prior to the application of advanced technologies for antibiotic removal,10 and the detection of antibiotic residues after wastewater purification11,12 are important issues related to antibiotic pollution. Although various spectroscopic and analytical techniques are known for the detection of © XXXX American Chemical Society

Received: September 26, 2017

A

DOI: 10.1021/acs.inorgchem.7b02471 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

used as a stock solution (pH ≈ 8.0). N.B. Antibiotics must be handled and disposed of with care! Fluorescence Measurements. The FCS-1 suspension was prepared by adding 20 mg of the FCS-1 powder sample to 20 mL of deionized water, ultrasonicating the solution for 30 min, and placing it overnight to let it form a uniform emulsion. In a typical fluorescencequenching experiment, 2 mL of the above emulsion (FCS-1 content: ca. 7.4 μg, concentration: ca. 3.7 ppm) was added to a quartz cuvette (1 × 1 cm), and the fluorescence-emission spectra with excitations at 321 nm were recorded in situ after each incremental addition of a 1 mM antibiotic solution at room temperature. The fluorescence-quenching efficiencies were calculated by using the formula (1 − I/I0) × 100%, where I0 and I are the fluorescence intensities before and after the addition of an antibiotic, respectively. The fluorescence-quenching efficiencies were further analyzed using the Stern−Volmer (S−V) equation, (I0/I) = Ksv[Q]+1, in which [Q] is the molar concentration of the antibiotic, and Ksv is the quenching constant (M−1). The quenching constant, Ksv, for each individual antibiotic was calculated using the mean values of the fluorescence-quenching efficiency at different antibiotic concentrations. X-ray Crystallographic Study. Single-crystal X-ray data of FCS-1 were collected on an Agilent SuperNova diffractometer. The structure was solved by direct method and developed by difference Fourier techniques, both of which used the SHELXL software package.36 All non-hydrogen atoms were treated anisotropically. The positions of hydrogen atoms attached to carbon atoms were generated geometrically. Idealized positions of the hydrogen atoms of water molecules were located from Fourier-difference maps and refined isotropically. CCDC 1560982 contains the supplementary crystallographic data for this paper, and a summary of the crystallographic data is given in Table S1. Computational Details for the Lowest Unoccupied Molecular Orbitals (LUMOs) and Highest Occupied Molecular Orbitals (HOMOs) of the Antibiotics. All the calculations were carried out with the Gaussian 09 software package37 at the Supercomputing Center of the Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences (Table S2). The geometric structures were optimized with the DFT method at the M062X/6-31G (d,p) level. The harmonic frequency calculation at the same computational level was performed to confirm that the obtained structure was the local minimum.

simulated wastewater by using FCS materials is challenging from both an experimental and a practical point of view. Herein, we report a fluorescent MOF material based on trinuclear zinc clusters with a formula of {[Zn3(μ3-OH)(HL)L(H2O)3]·H2O}n (FCS-1, H3L = 5-(4-carboxy-phenoxymethyl)isophthalic acid). FCS-1 has been prepared and subsequently applied as a highly efficient and quickly responsive chemical sensor for the detection of sulfonamide antibiotics in simulated wastewater, which are some of the most widely used antibiotics.33−35 Interestingly, FCS-1 exhibited viable and effective detection of sulfonamide antibiotics in the aqueous phase based on the fluorescence-quenching effect. Moreover, the detection of sulfonamide antibiotics using FCS-1 maintains the detection efficiency in the wide pH range of 3.0−9.0 and is not influenced by the presence of heavy-metal ions, which is promising for possible applications in wastewater treatment and water-quality monitoring. Furthermore, the detection mechanism has been studied on the basis of computational and experimental results.



EXPERIMENTAL SECTION

Materials and General Methods. All commercially available reagents and starting materials were of reagent-grade quality and used without further purification. The sulfonamide antibiotics were of analytical grade. Elemental analyses (EA, CHN) were carried out on an Elementar Vario EL III analyzer. Infrared (IR) spectra were measured on a PerkinElmer Spectrum One instrument with KBr pellets. Thermogravimetric analysis (TGA) was performed using NETZSCH STA 449C under a nitrogen atmosphere with a heating rate of 10 °C min−1. Diffuse reflectance spectra (DRS) were recorded on a Shimadzu UV−vis spectrophotometer (UV-2550) with BaSO4 as the background. 1 H NMR spectra were recorded on a Bruker DPX-400 spectrometer. Raman spectra were collected with a Thermo Nicolet Nexus 870 spectrometer (Laser excitation line: 325 nm). Powder X-ray diffraction (PXRD) patterns were collected on a Bruker D8 ADVANCE diffractometer with Cu Kα radiation (λ = 1.5418 Å). The fluorescence excitation and emission spectra were recorded on a Hitachi F-4500 spectrophotometer. Scanning electron microscopy (SEM) was performed using a Phenom ProX. Synthesis of 5-(4-Carboxy-phenoxymethyl)-isophthalic Acid (H3L). A mixture of methyl-p-hydroxybenzoate (1.52 g, 10.0 mmol), dimethyl-5-bromomethyl-1,3-benzene-dicarboxylate (2.87 g, 10.0 mmol), and K2CO3 (5 g) in 100 mL of DMF was placed in a three-neck round-bottom flask and refluxed at 100 °C for 24 h. The reaction mixture was extracted with DCM and evaporated to get a crude solid, which was then transferred to a flask with 100 mL of distilled water, 20 mL of ethanol, and 5 g of NaOH to hydrolyze at 80 °C for 24 h. The pH of the above solution was then adjusted to be acidic using concentrated hydrochloric acid (HCl) to obtain the final product, H3L. Yield: ca. 85%. 1H NMR (DMSO-d6): δ 6.98 (s, 2H, phenyl), δ 5.20 (s, 2H, −OCH2Ph−), δ 8.02 (s, 2H, phenyl), δ 8.40 (s, 2H, phenyl), and δ 8.93 (m, H, phenyl). Synthesis of {[Zn 3 (μ 3 -OH)(HL)L(H 2 O) 3 ]·H 2 O} n (FCS-1). A mixture of Zn(NO3)2·6H2O (45 mg, 0.15 mmol), H3L (31 mg, 0.1 mmol), NaHCO3 (25 mg, 0.3 mmol), and 10 mL of distilled water was sealed in a 25 mL Teflon-lined stainless-steel autoclave, heated at 130 °C for 72 h, and then cooled to room temperature within 12 h. Colorless block crystals were collected by filtration, washed with distilled water, and dried in air at ambient temperature. Yield: ca. 80% (based on Zn). Elemental analysis (EA) for C32H28O19Zn3 (Mr = 912.70): C, 42.11%; H, 3.09%. Found: C, 41.89%; H, 3.17%. IR (KBr, cm−1): 3429(w), 1472(s), 1382(s), 1246(s), 1166(m), 1009(w), 855(s), 770(s), 718(m), 681(w). Preparation of Antibiotic Stock Solutions. In a typical procedure, for example, that of sulfamethazine (SMZ), 13.9 mg of SMZ (0.05 mmol) was added to a 14 mL solution of NaOH (0.01 mol·L−1) and dissolved under magnetic stirring. Deionized water was then added to a constant volume of 50 mL, and the resulting solution was



RESULTS AND DISCUSSION

The hydrothermal reaction of Zn(NO3)2·6H2O, H3L (31 mg 0.1 mmol), and NaHCO3 at 130 °C resulted in the formation of the crystalline solid material FCS-1. The phase purity and thermal stability of FCS-1 were identified by IR, TGA, and PXRD studies (Figures S1, S2, and S3 in the ESI). Singlecrystal X-ray structure determination (Table S1 in the ESI) revealed that FCS-1 crystallized in the triclinic space group P1̅. The structure of FCS-1 consists of μ3-OH-bridged trinuclear zinc clusters {Zn3} surrounded by two HL2− and three L3− (Figure 1a and Figure S4a in the ESI). The HL2− connects two {Zn3} clusters, and each L3− holds three {Zn3}, by which a three-dimensional (3D) crystal lattice forms (Figure 1b). The network topology of FCS-1 was interpreted using ToposPro38 as a 3,5-connected (63)(69·8)-hms net (Figure 1b). Moreover, network interpenetration was observed between two identical hms nets related by full interpenetration symmetry elements (FISE, class IIa, Z = 2), resulting in an overall 2-fold-interpenetrated 3D structure (Figure 1c and Figure S4b in the ESI). The fluorescence properties of the organic ligand (H3L) and FCS-1 were examined in the solid state at room temperature. As shown in Figure 2a, the pristine ligand exhibits an emission band at ca. 375 nm (λex = ca. 313 nm), whereas a greatly enhanced luminescence is observed at ca. 382 nm for FCS-1 (λex = ca. 321 nm) under the same testing conditions. The fluorescence of B

DOI: 10.1021/acs.inorgchem.7b02471 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. (a) View of the secondary building units (SBUs) in FCS-1, (b) topological view of the hms net derived from FCS-1, and (c) view of the 2-fold-interpenetrated network.

Figure 2. (a) Excitation and emission spectra of H3L and FCS-1; (b) representative fluorescence-quenching spectrum of FCS-1 upon the addition of sulfonamide antibiotics (SMZ), inset: optical photographs of the quartz cuvette before and after the fluorescence-quenching experiments; and (c) Raman spectra and (d) simulated and experimental PXRD patterns of FCS-1 before and after the fluorescence-quenching experiments.

FCS-1 is mainly attributed to the fluorescent emission of the organic ligand, H3L, with significant enhancement upon the formation of the ordered 3D MOF structure. The water stability and pH tolerance of FCS-1 were further demonstrated by PXRD measurements of the samples treated in aqueous solutions at various pH values. The satisfactory match between the simulated and experimental PXRD patterns confirms the excellent water stability and pH tolerance of FCS-1 (Figure S3 in the ESI). The minor shifts of the characteristics at 12.7 and 27.1° (2θ) were ascribed to local structural changes due to the presence of flexible functional groups in the organic ligand (i.e., the ether and methylene groups). FCS-1 was then applied as a chemical sensor for the detection of various sulfonamide antibiotics in simulated wastewater. As expected, FCS-1 showed visually observable fluorescence-quenching effects toward eight selected sulfonamide antibiotics in aqueous media (Scheme S1), including sulfadiazine (SDZ), sulfamethazine (SMZ), sulfamethoxazole (SMX), sulfachlorpyridazine (SCP), sulfameter (SM), sulfamonomethoxine (SMM), sulfaquinoxaline (SQX), and sulfathiazole (STZ). For example, the maximum fluorescence intensity of FCS-1 reduced gradually upon the addition of the SMZ solution (Figure 2b), indicating the considerable quenching

effect of SMZ. Moreover, recovered FCS-1 retains its structural integrity and phase purity, as indicated by the Raman and PXRD patterns (Figure 2c,d). The fluorescent detection of different sulfonamide antibiotics have been measured and plotted in Figure 3a (see also Figure S5 in the ESI). The sulfonamide antibiotics exhibit significant quenching of FCS-1 in water, with quenching percentages varying from ca. 28% (SMX) to ca. 86% (SQX). From the linear fitting of the plots using the Stern−Volmer (S−V) equation (Figure 3b), the calculated Ksv values for SQX and STZ were found to be 3.7 × 104 and 2.2 × 104 M−1, which indicated the excellent quenching ability of FCS-1 toward SQX and STZ, in comparison with the Ksv values of 8.7 × 103, 5.4 × 103, 5.5 × 103, 5.7× 103, 3.1 × 103, and 1.87 × 103 M−1 for SMZ, SM, SDZ, SCP, SMM, and SMX, respectively. It is also worth noting that surface adsorption of sulfonamide antibiotics onto the FCS-1 host was observed for the recovered materials after the fluorescence-quenching experiments (Figure 4). The estimated decision limit of FCS-1, based on the fluorescence quenching of SQX, is comparable to those from spectroscopic and analytical techniques known for the detection of antibiotics, that is, capillary-electrophoresis tandem mass spectrometry14 C

DOI: 10.1021/acs.inorgchem.7b02471 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Concentration-dependent-fluorescence-quenching scheme (a) and Stern−Volmer plots (b) of various sulfonamide antibiotics.

The fluorescence-quenching effects of sulfonamide antibiotics can be ascribed to the π−π interactions that favor interplay between the antibiotics and FCS-1 via surface adsorption, and fluorescence quenching depends greatly on electron transfer from FCS-1 in an excited state to the sulfonamide antibiotics (Figure 5a). The lowest-unoccupied-molecular-orbital (LUMO) energy levels of the electron-deficient sulfonamide antibiotics should be generally lower than the conduction band of FCS-1, and thus electron transfer from the conduction band of FCS-1 to the LUMO of the sulfonamide antibiotics might occur upon excitation and cause fluorescence quenching.32,40 In order to rationalize the quenching efficiency of various sulfonamide antibiotics, theoretical highest-occupied-molecular-orbital (HOMO) and LUMO energies for the sulfonamide antibiotics were calculated (Figure 5b; see also Table S2 and Figure S7 in the ESI). Furthermore, Mott−Schottky plots of FCS-1 and SMZ, which showed the highest LUMO among the applied antibiotics, were measured at frequencies of 500, 1000, and 1500 Hz (Figure S8c,f in the ESI). The flat-band potential (Vfb) of FCS-1 and SMZ estimated from the intersection was approximately −2.2 V versus Ag/AgCl (−2.0 V vs NHE) and −1.08 V versus Ag/AgCl (−0.88 V vs NHE), respectively. With band gap energies (Eg) of 3.0 eV (FCS-1) and 3.14 eV (SMZ) (Figure S8a,b,d,e in the ESI), the HOMOs of FCS-1 and SMZ were calculated to be 1.0 V versus NHE and 2.26 V versus NHE, respectively (Figure S8c,f, inset, in the ESI). These experimental results further confirmed our hypothesis about the use of the electron-transfer pathway during fluorescence quenching, as illustrated in Figure 5a. The calculated LUMO energies for the sulfonamide antibiotics follow an order of SQX > SCP > SM ≈ STZ ≈ SDZ > SMX ≈ SMM ≈ SMZ, which was almost consistent with the

Figure 4. SEM images and elemental mapping of FCS-1 before and after the fluorescence-quenching experiments (color code: C, cyan; O, blue; N, navy; Zn, pink.).

and liquid-chromatography tandem mass spectrometry17 as well as MOF chemosensors.23 Ion and pH tolerances are key factors that can evaluate the reliability of chemical sensors in wastewater treatment, and thus the fluorescent detection of SMZ, as a representative of sulfonamide antibiotics, at different pH conditions and with the presence of heavy-metal ions interference was studied. The quenching efficiency of SMZ at the practical pH range of simulated antibiotic wastewater (pH = 3.0−9.0) was nearly unchanged (Figure S6a in the ESI). Similarly, the quenching efficiency of SMZ remained in an acceptable range in the presence of various heavy-metal ions (Figure S6b in the ESI). Notably, in the presence of Pb2+, the quenching efficiency at a low SMZ loading was reasonably higher than those with other metal ions, as a result of the fluorescence-quenching effects of Pb2+ known for similar systems.39 These results suggest that FCS-1 is potentially applicable as a chemical sensor for the detection of trace amounts of sulfonamide antibiotics in wastewater.

Figure 5. (a) Schematic representation of electron transfer from the LUMO of FCS-1 to the LUMO of sulfonamide antibiotics (ABs) and (b) theoretical HOMO and LUMO energies for the sulfonamide antibiotics. D

DOI: 10.1021/acs.inorgchem.7b02471 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Notes

experimental results of their quenching efficiencies. Moreover, the experimental quenching efficiency for SCP was considerably lower than expected, whereas those for STZ and SMZ were higher. The discrepancy is probably caused by the differences in the local electron distribution of SCP, STZ, and SMZ, originating from the substituent groups on the aromatic rings of the sulfonamide antibiotics. The electron-withdrawing chlorine atoms in SCP reduced the local electron densities of the diazine groups, thereby affecting the π−π interactions between SCP and FCS-1. Moreover, chlorine atoms are not ideally participating in hydrogen-bonding interactions with FCS-1. Thus, SCP might lack efficient supramolecular π−π and hydrogenbonding interactions with the host material, FCS-1, which would account for its unexpectedly low fluorescence-quenching efficiency. On the other hand, the electron-withdrawing and hydrogen-bond-donating methyl groups in STZ and SMZ greatly favored supramolecular interactions between STZ or SMX and FCS-1, and thus STZ and SMX exhibit enhanced fluorescence-quenching efficiencies.

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors gratefully acknowledge financial support from the National Natural Science Foundation of China (NSFC, grants 91622114 and 21441008) and the State Key Laboratory of Structural Chemistry (grants 20170032 and 20140006). J.L. thanks the New Century Excellent Talents in Fujian Province University and the International Science and Technology Cooperation and Exchange Project of Fujian Agriculture and Forestry University (grant KXGH17010) for funding. X.-D.Z. thanks the Excellent Young Scholar Foundation of Anhui Province (grant gxyqZD2017055) and the Distinguished Young Scholar Foundation of Anhui Polytechnic University (grant 2017JQ01) for financial support.





(1) Martinez, J. L. Environmental pollution by antibiotics and by antibiotic resistance determinants. Environ. Pollut. 2009, 157, 2893− 2902. (2) Hirsch, R.; Ternes, T. A.; Haberer, K.; Kratz, K.-L. Occurrence of antibiotics in the aquatic environment. Sci. Total Environ. 1999, 225, 109−118. (3) Halling-Sørensen, B.; Nors-Nielsen, S.; Lanzky, P. F.; Ingerslev, F.; Holten-Lützhøft, H. C.; Jørgensen, S. E. Occurrence, fate, and effects of pharmaceutical substances in the environment  a review. Chemosphere 1998, 36, 357−393. (4) Boxall, A. B. A.; Kolpin, D. W.; Halling-Sørensen, B.; Tolls, J. Are veterinary medicines causing environmental risks? Environ. Sci. Technol. 2003, 37, 286A−294A. (5) Schmitt, H.; Stoob, K.; Hamscher, G.; Smit, E.; Seinen, W. Tetracyclines and tetracycline resistance in agricultural soils: Microcosm and field studies. Microb. Ecol. 2006, 51, 267−276. (6) Kim, S.; Eichhorn, P.; Jensen, J. N.; Weber, A. S.; Aga, D. S. Removal of antibiotics in wastewater: Effect of hydraulic and solid retention times on the fate of tetracycline in the activated sludge process. Environ. Sci. Technol. 2005, 39, 5816−5823. (7) Van Wieren, E. M.; Seymour, M. D.; Peterson, J. W. Interaction of the fluoroquinolone antibiotic, ofloxacin, with titanium oxide nanoparticles in water: Adsorption and breakdown. Sci. Total Environ. 2012, 441, 1−9. (8) Xu, L.; Dai, J.; Pan, J.; Li, X.; Huo, P.; Yan, Y.; Zou, X.; Zhang, R. Performance of rattle-type magnetic mesoporous silica spheres in the adsorption of single and binary antibiotics. Chem. Eng. J. 2011, 174, 221−230. (9) Nakagawa, K.; Namba, A.; Mukai, S. R.; Tamon, H.; Ariyadejwanich, P.; Tanthapanichakoon, W. Adsorption of phenol and reactive dye from aqueous solution on activated carbons derived from solid wastes. Water Res. 2004, 38, 1791−1798. (10) Ternes, T. A.; Joss, A.; Siegrist, H. Scrutinizing pharmaceuticals and personal care products in wastewater treatment. Environ. Sci. Technol. 2004, 38, 392A−399A. (11) Pena, A.; Lino, C. M.; Alonso, R.; Barceló, D. Determination of tetracycline antibiotic residues in edible swine tissues by liquid chromatography with spectrofluorometric detection and confirmation by mass spectrometry. J. Agric. Food Chem. 2007, 55, 4973−4979. (12) Fernández, F.; Sánchez-Baeza, F.; Marco, M.-P. Nano gold probe enhanced Surface Plasmon Resonance immunosensor for improved detection of antibiotic residues. Biosens. Bioelectron. 2012, 34, 151−158. (13) Nagarkar, S. S.; Joarder, B.; Chaudhari, A. K.; Mukherjee, S.; Ghosh, S. K. Highly selective detection of nitro explosives by a luminescent metal−organic framework. Angew. Chem., Int. Ed. 2013, 52, 2881−2885. (14) Moreno-González, D.; Lara, F. J.; Jurgovská, N.; Gámiz-Gracia, L.; García-Campaña, A. M. Determination of aminoglycosides in

CONCLUSION In summary, a strongly fluorescent metal−organic framework (MOF) material, FCS-1, was prepared and applied as a highly sensitive and quickly responsive chemical sensor for the detection of sulfonamide antibiotics in simulated wastewater. The FCS-1 exhibited a viable and effective detection performance based on fluorescence quenching upon its exposure to sulfonamide antibiotics. Fluorescence quenching was not influenced by the presence of heavy-metal ions and pH changes related to the water treatment process. This work displays a powerful platform to advance fluorescent chemical sensors for their practical applications in water-quality monitoring. Future studies will focus on the further enhancement of the detection sensitivity in this and related systems, by means of tuning the fluorescence and guest-available docks in the MOF materials, so as to increase its detection selectivity for various antibiotics.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b02471. Crystallographic data, DFT calculations of HOMO and LUMO energies, IR spectrum, TGA curve, PXRD patterns, fluorescence spectra, and additional figures (PDF) Accession Codes

CCDC 1560982 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.



REFERENCES

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (X.-D.Z.). *E-mail: tfl[email protected] (T.-F.L.). *E-mail: [email protected] (J.L.). ORCID

Jian Lü: 0000-0002-0015-8380 E

DOI: 10.1021/acs.inorgchem.7b02471 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

quantitatively sensing of HgII and CrVI in aqueous solution. Inorg. Chem. 2015, 54, 7133−7135. (33) Zhang, Q.-Q.; Ying, G.-G.; Pan, C.-G.; Liu, Y.-S.; Zhao, J.-L. 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. (34) Kim, S. C.; Carlson, K. Temporal and spatial trends in theoccurrence of human and veterinary antibiotics in aqueous and river sediment matrices. Environ. Sci. Technol. 2007, 41, 50−57. (35) Kolpin, D. W.; Furlong, E. T.; Meyer, M. T.; Thurman, E. M.; Zaugg, S. D.; Barber, L. B.; Buxton, H. T. Pharmaceuticals, hormones, and other organic wastewater contaminants in US streams, 1999− 2000: A national reconnaissance. Environ. Sci. Technol. 2002, 36, 1202−1211. (36) Sheldrick, G. M. SHELXS97. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112. (37) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A.; Peralta, Jr., J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision A.02; Gaussian, Inc.: Wallingford, CT, 2009. (38) Blatov, V. A.; Shevchenko, A. P.; Proserpio, D. M. Applied topological analysis of crystal structures with the program package ToposPro. Cryst. Growth Des. 2014, 14, 3576−3586. (39) Qiu, Y.; Deng, H.; Mou, J.; Yang, S.; Zeller, M.; Batten, S. R.; Wu, H.; Li, J. In situ tetrazole ligand synthesis leading to a microporous cadmium−organic framework for selective ion sensing. Chem. Commun. 2009, 36, 5415−5417. (40) Malik, A. H.; Iyer, P. K. Conjugated polyelectrolyte based sensitive detection and removal of antibiotics tetracycline from water. ACS Appl. Mater. Interfaces 2017, 9, 4433−4439.

honey by capillary electrophoresis tandem mass spectrometry and extraction with molecularly imprinted polymers. Anal. Chim. Acta 2015, 891, 321−328. (15) 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. (16) Tabrizchi, M.; ILbeigi, V. Detection of explosives by positive corona discharge ion mobility spectrometry. J. Hazard. Mater. 2010, 176, 692−696. (17) Blasco, C.; Corcia, A. D.; Picó, Y. Determination of tetracyclines in multi-specie animal tissues by pressurized liquid extraction and liquid chromatography-tandem mass spectrometry. Food Chem. 2009, 116, 1005−1012. (18) Hou, B.-L.; Tian, D.; Liu, J.; Dong, L.-Z.; Li, S.-L.; Li, D.-S.; Lan, Y.-Q. A water-stable metal−organic framework for highly sensitive and selective sensing of Fe3+ ion. Inorg. Chem. 2016, 55, 10580−10586. (19) Joarder, B.; Desai, A. V.; Samanta, P.; Mukherjee, S.; Ghosh, S. K. Selective and sensitive aqueous-phase detection of 2,4,6Trinitrophenol (TNP) by an amine-functionalized metal−organic framework. Chem. - Eur. J. 2015, 21, 965−969. (20) Fu, R. B.; Hu, S. M.; Wu, X. T. Rapid and sensitive detection of nitroaromatic explosives by using new 3D lanthanide phosphonates. J. Mater. Chem. A 2017, 5, 1952−1956. (21) Bagheri, M.; Masoomi, M. Y.; Morsali, A.; Schoedel, A. Two dimensional host−guest metal−organic framework sensor with high selectivity and sensitivity to picric acid. ACS Appl. Mater. Interfaces 2016, 8, 21472−21479. (22) Zhou, E.-L.; Huang, P.; Qin, C.; Shao, K.-Z.; Su, Z.-M. A stable luminescent anionic porous metal−organic framework for moderate adsorption of CO2 and selective detection of nitro explosives. J. Mater. Chem. A 2015, 3, 7224−7228. (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) Mason, J. A.; Oktawiec, J.; Taylor, M. K.; Hudson, M. R.; Rodriguez, J.; Bachman, J. E.; Gonzalez, M. I.; Cervellino, A.; Guagliardi, A.; Brown, C. M.; Llewellyn, P. L.; Masciocchi, N.; Long, J. R. Methane storage in flexible metal−organic frameworks with intrinsic thermal management. Nature 2015, 527, 357−361. (25) Yang, S.; Lin, X.; Lewis, W.; Suyetin, M.; Bichoutskaia, E.; Parker, J. E.; Tang, C. C.; Allan, D. R.; Rizkallah, P. J.; Hubberstey, P.; Champness, N. R.; Thomas, K. M.; Blake, A. J.; Schröder, M. A partially interpenetrated metal−organic framework for selective hysteretic sorption of carbon dioxide. Nat. Mater. 2012, 11, 710−716. (26) Furukawa, H.; Ko, N.; Go, Y. B.; Aratani, N.; Choi, S. B.; Choi, E.; Yazaydin, A.Ö .; Snurr, R. Q.; O’Keeffe, M.; Kim, J.; Yaghi, O. M. Ultrahigh porosity in metal−organic frameworks. Science 2010, 329, 424−428. (27) Cooper, A. I.; Rosseinsky, M. J. Metal−organic frameworks: Improving pore performance. Nat. Chem. 2009, 1, 26−27. (28) Lan, A.; Li, K.; Wu, H.; Olson, D. H.; Emge, T. J.; Ki, W.; Hong, M.; Li, J. A luminescent microporous metal−organic framework for the fast and reversible detection of high explosives. Angew. Chem., Int. Ed. 2009, 48, 2334−2338. (29) Pramanik, S.; Zheng, C.; Zhang, X.; Emge, T. J.; Li, J. New microporous metal−organic framework demonstrating unique selectivity for detection of high explosives and aromatic compounds. J. Am. Chem. Soc. 2011, 133, 4153−4155. (30) Yang, X.-L.; Chen, X.; Hou, G.-H.; Guan, R.-F.; Shao, R.; Xie, M.-H. A multiresponsive metal−organic framework: direct chemiluminescence, photoluminescence, and dual tunable sensing applications. Adv. Funct. Mater. 2016, 26, 393−398. (31) Zhao, S.-S.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Fluorescent aromatic tag-functionalized MOFs for highly selective sensing of metal ions and small organic molecules. Inorg. Chem. 2016, 55, 2261−2273. (32) Wen, L.; Zheng, X.; Lv, K.; Wang, C.; Xu, X. Two aminodecorated metal−organic frameworks for highly selective and F

DOI: 10.1021/acs.inorgchem.7b02471 Inorg. Chem. XXXX, XXX, XXX−XXX