Ligand-Controlled Integration of Zn and Tb by Photoactive Terpyridyl

1 hour ago - Synopsis. One water-stable 3D heterometallic metal−organic framework (MOF) were successfully synthesized and structurally characterized...
0 downloads 7 Views 4MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Ligand-Controlled Integration of Zn and Tb by Photoactive Terpyridyl-Functionalized Tricarboxylates as Highly Selective and Sensitive Sensors for Nitrofurans Zhi-Hang Zhou,†,∇ Wen-Wen Dong,†,∇ Ya-Pan Wu,† Jun Zhao,† Dong-Sheng Li,*,† Tao Wu,‡ and Xian-hui Bu*,§ †

College of Materials and Chemical Engineering, Hubei Provincial Collaborative Innovation Center for New Energy Microgrid, Key Laboratory of Inorganic Nonmetallic Crystalline and Energy Conversion Materials, China Three Gorges University, Yichang, 443002, China ‡ The Key Lab of Health Chemistry and Molecular Diagnosis of Suzhou, College of Chemical Engineering and Materials Science Soochow University, Suzhou, Jiangsu 215123, China § Department of Chemistry and Biochemistry, California State University, Long Beach, 1250 Bellflower Boulevard, Long Beach, California 90840, United States S Supporting Information *

ABSTRACT: The integration of terpyridyl and tricarboxylate functionality in a novel ligand allows concerted 3:1 stoichiometric assembly of size-and chargecomplementary Zn2+/Tb3+ ions into a water-stable 3D luminescent framework (CTGU-8) capable of highly selective, sensitive, and recyclable of nitrofurans.



INTRODUCTION Nitrofurans (NFs), which are a group of antimicrobial antibiotics with a common nitrofuran ring but different side chains,1 were widely used as additives in food-producing livestock2 and medicine.3 They have been found in food and drinking water,4 and can therefore accumulate in humans,5 with the potential to cause allergic reactions,6 hearing loss,7 nausea,8,9 acute pancreatitis,10 and other health issues. As a result, the detection of nitrofurans (NFs) has significant environmental and health implications. While methods do exist for the detection of antibiotics, such as high-performance liquid chromatography (HPLC),11 surface-enhanced Raman spectroscopy (SERS),12 gas chromatography−mass spectrometry (GC-MS),13 and liquid chromatography−tandem mass spectrometry (LC-MS/MS),14 there remains a need for fast, simple, and low-cost detection techniques. A promising contender is the luminescent sensing technique, which takes advantage of the change in luminescent spectra induced by interactions between analytes and sensors.15 One key to such applications is the development of advanced sensor materials targeting specific analytes. Lanthanide metal−organic frameworks (Ln-MOFs) are promising multifunctional materials, because of their unique © XXXX American Chemical Society

luminescent behaviors such as relatively long lifetime, high color purity, and intense sharp emission in the near-infrared and visible regions.16 Except for Ln-MOFs, transition-metal− lanthanide (d-f) heterometallic MOFs have been widely explored.17 The d-blocks (such as Zn, Cd, Ag, etc.) can play a significant role in the luminescence properties, because, in addition to emitting their own luminescence, they can facilitate energy transfer between ligands and lanthanide ions by controlling the Ln-ligand spatial relationship. Until now, various MOFs have been widely used in some fields, such as magnetism,17b electrocatalysis,18 photocatalytic degradation,19 the detection of explosives,17e and metal ions.17g,20 However, to the best of our knowledge, few studies involve the use of luminescent MOFs for the detection of antibiotics.21 In particular, few studies are known that involve the use of synergistic effects in heterometaillic MOFs for the detection of antibiotics. Generally, the selection of ligands for luminescent applications places a great consideration in the ability of ligands to serve as chromophores, such as those with large Received: December 19, 2017

A

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry conjugated π-systems.22 Also receiving consideration is another feature of ligands, the presence of both N- and O-donor sites, because this feature helps the integration of d- and f-metal ions due to differing affinities for different metal ions by N and O donors. Here, we are further interested in probing the factors that can only integrate d and f metal ions, as well as chromophores, but also dictate their d/f stoichiometric ratio. Presumably, by employing ligands with specific combination and geometric placement of complementary functional groups, it is possible to not only integrate d- and f-metals to produce novel framework materials, but also provide a certain degree of control over the ratio of heterometals. Given the increasing significance and popularity of heterometallic inorganic systems (luminescent MOFs in particular), we have initiated a project to expand the scope of the inorganic heterometallic system by seeking to probe and establish synthetic and structural factors that can aid in the control of heterometal ratios. Toward this goal, we have chosen a very novel ligand that contains the functionality of two very common low-cost ligands: trimesic acid (H3BTC) and terpyridine (TPY). In addition to the nature of the introduced functional groups, we are also keenly interested in complicating factors such as mutual steric hindrance between functional groups, as well as relative binding strength affected by factors such as chelation. The new ligand in this work provides an excellent platform to study all of these effects. In this work, it is interesting to note that the ratio of d/f ions (Zn2+/Tb3+ = 3) is equal to the ratio of carboxylate/terpyridyl groups, despite steric hindrance of terpyridyl group over the adjacent carboxylate groups, impacting the binding modes of such carboxylate groups to metal sites. Here, we report the synthesis and structural characterization of a new d-f heterometallic MOF, [(CH 3 ) 2 NH 2 ][TbZn3(L)3(HCOO)(H2O)2]·5H2O (CTGU-8) [H3L = 4(2,4,6-tricarboxyl phenyl)- 2,2′:6′,2″- terpyridine]. Strikingly, CTGU-8 shows excellent selective, sensitive, and reversible luminescent probing for nitrofurans (NFs):antibiotics in aqueous solution (NFs include furazolidone (FZD), nitrofurazone (NFZ), and nitrofurantoin (NFT)).



1599(s), 1525(m), 1480(m), 1454(m), 1422(w), 1391(w), 1304(m), 1206(w), 830(m), 800(s), 792(m), 730(m), 708(s). X-ray Crystallography. Single-crystal X-ray diffraction (XRD) data for CTGU-8 were collected on a Rigaku XtaLAB mini diffractometer equipped with a graphite monochromated Cu Kα radiation (λ = 1.54184 Å) by using the ϕ/ω scan technique. A suitable crystal was selected and data was collected on a XtaLAB Pro:Kappa single diffractometer. The crystal was kept at 293(2) K during data collection. Using Olex2,23 the structure was solved with the ShelXT24 structure solution program, using intrinsic phasing and refined with the ShelXL 2014 refinement package using least-squares minimization.25 Positional and anisotropic displacement parameters were refined for all non-hydrogen atoms, unless stated in the text below each structure given. Hydrogen atoms were placed in geometrically calculated positions and refined as part of a riding model. The details of the crystal parameters, data collection, and refinements for CTGU-8 are summarized in Table 1. Selected bond lengths and angles are listed in Table S1 in the Supporting Information.

Table 1. Crystal and Structure Refinement Data for Complex CTGU-8

EXPERIMENTAL SECTION

Materials and Physical Measurements. All the chemicals were received as reagent grade and used without any further purification. Elemental analyses were performed on a PerkinElmer 2400 Series II analyzer. FT-IR spectra were recorded as KBr pellets on a Thermo Electron NEXUS 670 FTIR spectrometer. Powder X-ray diffraction (PXRD) analyses were recorded on a Rigaku Ultima IV diffractometer (Cu Kα radiation, λ = 1.5406 Å). The purity and homogeneity of the bulk products were determined by comparison of the simulated and experimental PXRD patterns. Thermogravimetric (TG) curves were recorded on a Netzsch Model 449C thermal analyzer with a heating rate of 10 °C min−1 under an air atmosphere. UV-spectra analysis were carried out on a Shimadzu Model UV 2550 spectrometer. Photoluminescence analysis was performed on an Edinburgh Model FLS55 luminescence spectrometer at room temperature. Synthesis of Complex [(CH3)2NH2][Tb Zn3(L)3(HCOO)(H2O)2]· 5H2O (CTGU-8). Zn(ClO4)2·6H2O(0.05 mmol, 0.0186g), Tb(NO3)3· 6H2O (0.05 mmol, 0.0226 g), L (0.05 mmol, 0.0112 g) were dissolved in DMF-H2O (v:v = 2:1, 3 mL), and then HNO3(0.8 mL, 3.5M) was added. The mixture was sealed in a Teflon-lined stainless steel vessel (25 mL), heated at 140 °C for 4 days, and then cooled to room temperature over 48 h. Colorless crystals were obtained. Yield: 36% (based on Zn(II)). Elemental analysis (%): calcd for C75H59N10O27TbZn3 (Mr = 1887.35): C 47.73, H 3.15, N 7.42; found: C 47.95, H 3.32, N 7.61. IR (cm−1): 3535(s), 1644(m),

parameter

value

complex chemical formula formula weight temperature (K) crystal system space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z density (g cm−3) Abs coeff. (mm−1) F(000) Rint goodness-of-fit on F2 final R indices [I > 2σ(I)] R indices (all data)

CTGU-8 C75H59N10O27TbZn3 1887.35 293(2) triclinic P1̅ 10.84312(9) 18.23895(14) 20.76839(16) 115.6661(8) 95.3999(7) 94.5121(6) 3652.83(5) 2 1.716 6.586 1900.0 0.0659 1.048 R1 = 0.0451, wR2 = 0.1208 R1 = 0.0483, wR2 = 0.1230

R1 = ∑(|Fo| − |Fc|)/Σ|Fo|. Fo|2)2]}1/2.

a

b

wR2 = {∑[w(|Fo|2− |Fc|2)2]/∑[w(|



RESULTS AND DISCUSSION Description of Crystal Structures. Single-crystal X-ray structure analysis reveals that CTGU-8 crystallizes in the triclinic space group P1,̅ and the asymmetic unit consists of one Tb3+ ion, three Zn2+ ions, three fully deprotonated L3− ligand anions with three different coordination modes, two coordinated water molecules, five lattice water molecules, one coordinated HCOO− anion, and one free [(CH3)2NH2]+ cation in the voids of CTGU-8 from the hydrolysis of DMF (Figure 1). Tb1 center displays a 8-coordination geometry surrounded by four monocarboxylate-O atoms (three from the ligand and one from HCOO−), two chelating carboxylate-O atoms, two water-O atoms. Smaller Zn ions adopt 5coordinated geometry with distorted pentagonal pyramid geometry, consisting of two carboxylate-O and three tpy-N atoms. B

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. Perspective view of the three-dimensional (3D) framework of CTGU-8. Figure 1. X-ray structure of CTGU-8 with thermal ellipsoids (50% probability). Symmetry code: #1, 1 − x, 1 − y, −z; #2, −x, 1 − y, −z; #3, 2 −x, 2 − y, 1 − z; #4, +x, 1 + y, +z; #5, 2 − x, 1 − y, 1 − z; #6, −x, −y, −z.

this net is P63/mmc, which has a helical pattern, as observed here. In addition, the face-to-face π−π stacking interactions between adjacent ligands enhance the stability of the entire structure, which turns out to be beneficial for the application of this material involving an aqueous environment. The centroid− centroid separation distance is 3.766 Å, with an offset of 1.371 Å (see Figure S2 in the Supporting Information). PXRD and TGA Measurements. Considering that the applications for the detection of antibiotics often involves water, the aqueous stability of the sample CTGU-8 was examined by immersing it into water and aqueous solutions of a series of antibiotics for 5 days. The PXRD analysis demonstrates that there is no obvious structure transformation, confirming the strong chemical stability (see Figure S3 in the Supporting Information). The cavity of CTGU-8 are occupied by five free water molecules, one free [(CH3)2NH2]+ cation every unit. The total accessible volume of CTGU-8 is 15.1%, calculated using the PLATON program. The TGA curve of CTGU-8 shows that, from room temperature to 160 °C, the weight loss of 6.2% can be observed, which corresponds to the loss of free and coordinated water molecules (calcd: 6.7%). the decomposition of the framework begins when the temperature rises above 380 °C (see Figure S4 in the Supporting Information). Photoluminescent Properties. MOFs constructed from d10 metal centers, f-metal, and electron-rich π-conjugated ligands are promising candidates for photoactive materials. In this work, the PL spectra of CTGU-8 were investigated in the solid state at room temperature (see Figure S5 in the Supporting Information). Under UV irradiation, CTGU-8 displayed intense green-light emission, upon excitation at 361 nm, and the emission spectrum of CTGU-8 features four emission peaks at 492, 542, 582, and 619 nm, corresponding to the characteristic transitions of 5D4 → 7FJ(J = 6, 5, 4, and 3) from Tb3+ ions. The most intense emission is in the 5D4 → 7F5 transition region at 542 nm and the intensity ratio for I(5D4 → 7 F5)/I(5D4 → 7F6) is ∼3.97, indicating the low-symmetry coordination of TbIII.26 The excitation spectrum shows the lanthanide-centered emission is sensitized by ligand L with π electrons and five-coordinated ZnII units.27 The ligand L, which is a nitrogen-containing ligand and has the regular conjugated π systems (benzene ring), coordinated with the lanthanide ions

The terpyridyl moiety of the L3− anion acts as the tridentatechelating mode with two terminal pyridyl rings in a trans arrangement. The carboxylate groups are all deprotonated with different coordination modes to bridge Zn2+ and Tb3+ ions. Two Zn2+ sites are joined by two different L3− to generate a binuclear Zn2L2 unit (Figure 2a) with nearest distances of

Figure 2. (a) Zn2L2 unit and (b) two-dimensional (2D) net of CTGU8; and (c) Tb2(COO)2 unit in CTGU-8.

5.536, 5.807, and 5.339 Å for Zn1···Zn1, Zn2···Zn2, and Zn3··· Zn3, respectively. Each pair of Tb3+ centers is bridged by two different L3− ligands to form a binuclear Tb2(COO)2 unit (Figure 2c). The Tb2(COO)2 units are linked by two carboxyl groups from two L3− ligands extending into a one-dimensional (1D) double chain running along the a-axis (Figure 2b). Tb3+ ions wrap around the screw axis with the closest Tb1···Tb1 distance of 5.542 Å. The neighboring double chains are connected by Zn2L2 units into 3D framework (Figure 3). Topologically, the binuclear Zn2L2 unit could be regarded as a 4-connected node, and the binuclear Tb2(COO)2 unit serve as a 6-connected node. Thus, the 2-nodal 3D network can be described as a 4,6-connected tcj topology with the point symbol of (4 4 ·6 2 )(4 8 ·6 6 ·8) (see Figure S1 in the Supporting Information). Note that the highest possible symmetry for C

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

The quantitative detection of the target is also highly desirable. The fluorescence quenching efficiency was analyzed with the Stern−Volmer equation: I0 = 1 + K sv[M] I where I0 and I are the luminescence intensities before and after the addition of analytes, Ksv is the quenching effect constant, and [M] is the concentration of the analyte. The SV plots exhibited good linear correlations at low concentrations, but subsequently deviated from linearity and bent upward at higher concentrations (see Figures 6−8, as well as Figures S7, S10, and

directly, making it more efficient in transferring the energy and resulting in stronger emission.28 The excited-state lifetime and quantum yield of CTGU-8 are 1.56 μs and 4.19%, respectively (see Figure S6 in the Supporting Information). Luminescent Sensing. Sample CTGU-8 was fine grinded and dispersed in 0.7 mM aqueous solutions of different antibiotics by utrasonic treatment for 20 min. Five classes of frequently used antibiotics were studied: • NMs (nitroimidazoles, including ronidazole (RDZ), metronidazole (MDZ), dimetridazole (DTZ), ornidazole (ODZ)); • sulfonamides (such as sulfadiazine (SDZ) and sulfamethazine (SMZ)); • chloramphenicols (chloramphenicol (CAP)); • β-lactams (penicillin (PCL)); and • NFs (nitrofurans, including furazolidone (FZD), nitrofurantoin (NFT), and nitrofurazone (NFZ)). As shown in Figure 4, NFs (e.g., furazolidone (FZD), nitrofurazone (NFZ), and nitrofurantoin (NFT)) have an

Figure 6. Fluorescence intensities of CTGU-8 dispersed in different concentrations of FZD in aqueous solution excited at 361 nm. Figure 4. Quenching efficiency of CTGU-8 excited at 361 nm in aqueous solution treated with different antibiotics (0.7 mM).

obvious quenching effect on the emission of CTGU-8 at 542 nm, whereas other antibiotics display much smaller or minor influence on the emission (see Figure 5). The quenching percentages for NFs (FZD, NFZ, and NFT) were up to 95%. These results demonstrate that CTGU-8 shows highly selective sensing behavior toward NFs, compared to other antibiotics.

Figure 7. Fluorescence intensities of CTGU-8 dispersed in different concentrations of NFT in aqueous solution excited at 361 nm.

S13 in the Supporting Information). Such changes in the nonlinear SV plots may be due to an energy-transfer process. The limit of detection (LOD) value is calculated using the equation 3σ LOD = K sv where Ksv is the quenching effect constant and σ is the standard deviation for three repeated fluorescent measurements of blank

Figure 5. Luminescence spectra of CTGU-8 excited at 361 nm in aqueous solution treated with different antibiotics (0.7 mM). D

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(CTGU-8) have advantages over Ln-MOFs for more-effective electron and energy transfer processes, thus leading to better NFs sensing performance. In heterometallic MOFs (CTGU-8), the Zn center bridges organic linker to form a Zn2L2 unit, and the Zn2L2 unit could be seen as a metalloligand and also as antennae with high electron-transferring efficiencies to analytes.30 This type of antenna structure facilitates intracrystal energy migration to the surface of the sample, followed by an efficient electron transfer quenching process at the MOFs/ solution interface.31 Considering the small pore size of CTGU8 and dimension of analytes, we can find that the complex might not encapsulate antibiotics. Small-sized crystals make it easy to disperse in solution, and it is needed for sample to make a good contact with the aqueous solutions of different antibiotics. In addition, fine powder with large surface area facilitates the migration of excited-state energy from interior to surface of crystals. Therefore, it is necessary to get the fine powder by grinding and ultrasonic treatment.32 Recyclability of the sensor plays an important role in detecting antibiotics. The recyclability experiments were conducted to study the potential of CTGU-8 for practical applications. It was observed that the luminescence intensity of CTGU-8 were regained by centrifugation and washed several times with water (see Figures S17−S19 in the Supporting Information). The PXRD pattern shows that the framework was not damaged after antibiotics sensing experiments. These results indicate the potential value of CTGU-8 as a sensing material. Figure S20 in the Supporting Information shows that the quenching efficiency was augmented with the increases in the concentrations. The values of quenching efficiencies of NFs are significant, even in the range of low concentrations, but the quenching efficiencies of other antibiotics attain an insignificant level at lower concentrations and show small changes at higher concentrations. Moreover, we have carried out experiments about the interference of other analytes. The changes of the luminescence intensity of CTGU-8 have been recorded, and the selective quenching effects cannot be disturbed in the presence of selected antibiotics. These studies prove the proper selectivity of CTGU-8 toward particular antibiotics (see Figure S21 in the Supporting Information).

Figure 8. Fluorescence intensities of CTGU-8 dispersed in different concentrations of NFZ in aqueous solution excited at 361 nm.

solutions. CTGU-8 has Ksv values of 1.151 × 106 M−1, 1.83 × 106 M−1, 9.25 × 105 M−1 for FZD, NFZ, and NFT, respectively (see Figures S8 and S9, Figures S11 and S12, and Figures S14 and S15, respectively, in the Supporting Information). The LOD values for these NFs were estimated to be 49, 39, and 52 ppb for FZD, NFZ, NFT, respectively. The LOD, as shown in Tables S2 and S3 in the Supporting Information, is lower than most reported values, especially lower than those using Zr(IV)based MOFs,21a Cd(II)-based MOFs,21b Eu-based MOFs21c for NFs detection, showing that CTGU-8 has superior sensitivity as an NF probe. At the same time, CTGU-8 has high quenching efficiency toward NFs, compared with other antibiotics, indicating excellent selectivity in the detection of NFs. Mechanism of Luminescent Sensing. The quenching mechanism of CTGU-8 toward NFs (FZD, NFZ, and NFT) is explored to better understand the fluorescence quenching effect. The “antenna effect” is considered as one possible reason for the luminescence quenching. UV-vis absorption for different antibiotics was recorded (Figure S16 in the Supporting Information). The absorption spectrum of FZD, NFZ, and NFT has extensive overlap with the excitation spectrum of CTGU-8, and no obvious overlap was observed for other antibiotics. Thus, the competition absorption of the excitation light between antibiotics and CTGU-8, reducing the efficiency of energy transfer from the ligand to the lanthanide ions, may lead to luminescence quenching. Another possible luminescence quenching mechanism is the transfer of electrons. Because the energy level of the conduction band (CB) of MOFs is usually higher than those of the LUMOs of analytes, electron-transfer can happen easily from electron-rich CUGU-8 to the electron-deficient NFs. HOMO and LUMO values of selected antibiotics are listed in Table S4 in the Supporting Information. The estimated LUMO values of NFs are lower than those of other antibiotics, which implies that the electron-transfer process between CUGU-8 and NFs is easier to achieve, compared to other antibiotics. For this reason, the highest quenching efficiency of NFs agrees fairly with lowest LUMO energy values.21a,c,29 To the best of our knowledge, one Ln-MOF was reported for NFs sensing.21c The mechanism of luminescence quenching is also described as electron-transfer and inner-filter effect (IFE), which is similar to that of CTGU-8. The heterometallic MOFs



CONCLUSION In conclusion, a 3D heterometallic Zn−Tb luminescent metal− organic framework has been obtained from a novel ligand with multifunctional, complementary coordination and optical properties. This material has excellent sensitivity, selectivity, and reversibility for the detection of NFs sensor with good stability during the sensing process, especially in the aqueous environment. This work indicates the great potential of heterometallic luminescent metal−organic frameworks as sensing materials for selective detection of the antibiotics in the environment.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b03200. PXRD, TGA, UV-vis spectroscopy, solid-state fluorescence decay curves of complex, IR spectra, additional figures, and PL spectra (PDF) E

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Accession Codes

(12) Xie, Y. F.; Zhu, X. Y.; Sun, Y. Y.; Wang, H. Y.; Qian, H.; Yao, W. R. Rapid detection method for nitrofuran antibiotic residues by surface-enhanced Raman Spectroscopy. Eur. Food Res. Technol. 2012, 235, 555−561. (13) Yang, S.; Zhu, X.; Wang, J.; Jin, X.; Liu, Y.; Qian, F.; Zhang, S.; Chen, J. Combustion of hazardous biological waste derived from the fermentation of antibiotics using TG-FTIR and Py−GC/MS techniques. Bioresour. Technol. 2015, 193, 156−163. (14) 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. (15) (a) Han, M.; Xu, G.; Li, D.; Azofra, L. M.; Zhao, J.; Chen, B.; Sun, C. A Terbium-Organic Framework Material for Highly Sensitive Sensing of Fe3+ in Aqueous and Biological Systems: Experimental Studies and Theoretical Analysis. Chemistryselect 2016, 1, 3555−3561. (b) Zhang, X.; Hu, Q.; Xia, T.; Zhang, J.; Yang, Y.; Cui, Y.; Chen, B.; Qian, G. Turn-on and Ratiometric Luminescent Sensing of Hydrogen Sulfide Based on Metal-Organic Frameworks. ACS Appl. Mater. Interfaces 2016, 8, 32259−32265. (c) Pan, M.; Zhu, Y.; Wu, K.; Chen, L.; Hou, Y.; Yin, S.; Wang, H.; Fan, Y.; Su, C. Epitaxial Growth of Hetero-Ln-MOF Hierarchical Single Crystals for Domain- and Orientation-Controlled Multicolor Luminescence 3D Coding Capability. Angew. Chem., Int. Ed. 2017, 56, 14582−14586. (d) Li, X.; Shi, W.; Wang, X.; Ma, L.; Hou, L.; Wang, Y. Luminescence Modulation, White Light Emission, and Energy Transfer in a Family of Lanthanide Metal-Organic Frameworks Based on a Planar π-Conjugated Ligand. Cryst. Growth Des. 2017, 17, 4217−4224. (16) (a) Li, H.; Wei, Y.; Dong, X.; Zang, S.; Mak, T. C. W. Novel TbMOF Embedded with Viologen Species for Multi-Photofunctionality: Photochromism, Photomodulated Fluorescence, and Luminescent pH Sensing. Chem. Mater. 2015, 27, 1327−1331. (b) Aulsebrook, M. L.; Biswas, S.; Leaver, F. M.; Grace, M. R.; Graham, B.; Barrios, A. M.; Tuck, K. L. A luminogenic lanthanide-based probe for the highly selective detection of nanomolar sulfide levels in aqueous samples. Chem. Commun. 2017, 53, 4911−4914. (c) Dai, Y.; Zhou, H.; Song, X.; Zhang, J.; Hao, C.; Di, L.; Wang, Y.; Ni, J.; Wang, H. Two (5,5)connected isomeric frameworks as highly selective and sensitive photoluminescent probes of nitroaromatics. CrystEngComm 2017, 19, 2786−2794. (d) Li, G.; Liu, G.; Li, Y.; Hou, L.; Wang, Y.; Zhu, Z. Uncommon Pyrazoyl-Carboxyl Bifunctional Ligand-Based Microporous Lanthanide Systems: Sorption and Luminescent Sensing Properties. Inorg. Chem. 2016, 55, 3952−3959. (17) (a) Chen, M.; Chen, M.; Okamura, T.; Lv, M.; Sun, W.; Ueyama, N. A series of silver(i) -lanthanide(iii) heterometallic coordination polymers: syntheses, structures and photoluminescent properties. CrystEngComm 2011, 13, 3801−3810. (b) Wang, K.; Liu, T.; Liu, Y.; Tian, X.; Sun, J.; Zhang, Q. Fluorescent heterometallic MOFs: tunable framework charges and application for explosives detection. CrystEngComm 2016, 18, 8301−8308. (c) Zhang, H.; Fan, R.; Chen, W.; Fan, J.; Dong, Y.; Song, Y.; Du, X.; Wang, P.; Yang, Y. 3D Lanthanide Metal-Organic Frameworks Based on Mono-, Tri-, and Heterometallic Tetranuclear Clusters as Highly Selective and Sensitive Luminescent Sensor for Fe3+ and Cu2+ Ions. Cryst. Growth Des. 2016, 16, 5429−5440. (d) Liu, Q.; Ge, S. Z.; Zhong, J. C.; Sun, Y. Q.; Chen, Y. P. Two novel 2D lanthanide-cadmium heterometal-organic frameworks based on nanosized heart-like Ln6Cd6O12 wheel-clusters exhibiting luminescence sensing to the polarization and concentration of cations. Dalton Trans. 2013, 42, 6314−6317. (e) Zhang, S.; Du, D.; Qin, J.; Li, S.; He, W.; Lan, Y.; Su, Z. 2D Cd(II)-Lanthanide(III) Heterometallic-Organic Frameworks Based on Metalloligands for Tunable Luminescence and Highly Selective, Sensitive, and Recyclable Detection of Nitrobenzene. Inorg. Chem. 2014, 53, 8105−8113. (f) Jia, R.; Li, H.; Chen, P.; Gao, T.; Sun, W.; Li, G.; Yan, P. Synthesis, structure, and tunable white light emission of heteronuclear Zn2Ln2 arrays using a zinc complex as ligand. CrystEngComm 2016, 18, 917− 923. (g) Li, Y.; Wang, D.; Liao, Z.; Kang, Y.; Ding, W.; Zheng, X.; Jin, L. Luminescence tuning of the Dy-Zn metal-organic framework and its application in the detection of Fe(III) ions. J. Mater. Chem. C 2016, 4,

CCDC 1572262 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 data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K.; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (D.-S. Li). *E-mail: [email protected] (X.-h. Bu). ORCID

Dong-Sheng Li: 0000-0003-1283-6334 Tao Wu: 0000-0003-4443-1227 Xian-hui Bu: 0000-0002-2994-4051 Author Contributions ∇

These authors contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (Grants Nos. 21673127, 21373122, 21671119, 51502155, and 51572152).



REFERENCES

(1) EFSA Panel on Contaminants in the Food Chain (CONTAM), European Food Safety Authority (EFSA). Scientific Opinion on nitrofurans and their metabolites in food. EFSA J.2015, 13, 4140− 414310.2903/j.efsa.2015.4140 (2) Vass, M.; Hruska, K.; Franek, M. Nitrofuran antibiotics: A review on the application, prohibition and residual analysis. Vet. Med. (Prague, Czech Repub.) 2018, 53, 469−500. (3) Browne, H. Antibiotics, gut bugs and the young. Nat. Rev. Microbiol. 2016, 14, 336−336. (4) Michael, I.; Rizzo, L.; McArdell, C. S.; Manaia, C. M.; Merlin, C.; Schwartz, T.; Dagot, C.; Fatta-Kassinos, D. Urban wastewater treatment plants as hotspots for the release of antibiotics in the environment: A review. Water Res. 2013, 47, 957−995. (5) Kowalski, P.; Plenis, A.; Oledzka, I. Optimization and validation of capillary electrophoretic method for the analysis of amphenicols in poultry tissues. Acta Pol. Pharm. 2008, 65, 45−50. (6) Abo Salem, E.; Fowler, J. C.; Attari, M.; Cox, C. D.; Perez Verdia, A.; Panikkath, R.; Nugent, K. Antibiotic-induced Cardiac Arrhythmias. Cardiovasc. Ther. 2014, 32, 19−25. (7) Bai, H.; Yuan, H.; Nie, C.; Wang, B.; Lv, F.; Liu, L.; Wang, S. A Supramolecular Antibiotic Switch for Antibacterial Regulation. Angew. Chem., Int. Ed. 2015, 54, 13208−13213. (8) Pilehvar, S.; Gielkens, K.; Trashin, S. A.; Dardenne, F.; Blust, R.; De Wael, K. Electro) Sensing of Phenicol AntibioticsA Review. Crit. Rev. Food Sci. Nutr. 2016, 56, 2416−2429. (9) Nagel, O.; Molina, M. P.; Althaus, R. Microbiological system in microtitre plates for detection and classification of antibiotic residues in milk. Int. Dairy J. 2013, 32, 150−155. (10) Marot, J. C.; Jonckheere, S.; Munyentwali, H.; Belkhir, L.; Vandercam, B.; Yombi, J. C. Tigecycline-induced acute pancreatitis: About two cases and review of the literature. Acta Clin. Belg. 2012, 67, 229−232. (11) Gbylik-Sikorska, M.; Posyniak, A.; Sniegocki, T.; Zmudzki, J. Liquid chromatography−tandem mass spectrometry multiclass method for the determination of antibiotics residues in water samples from water supply systems in food-producing animal farms. Chemosphere 2015, 119, 8−15. F

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry 4211−4217. (h) Zheng, S.; Tan, J.; Cai, S.; Fan, J.; Zhang, W. Construction of six new luminescent Ln(III)-Zn(II) heterometallic coordination polymers based on heterometallic secondary building units. CrystEngComm 2016, 18, 8672−8682. (i) Wen, H.; Dong, P.; Liu, S.; Liao, J.; Liang, F.; Liu, C. 3d-4f heterometallic trinuclear complexes derived from amine-phenol tripodal ligands exhibiting magnetic and luminescent properties. Dalton Trans. 2017, 46, 1153− 1162. (18) Wu, Y.; Zhou, W.; Zhao, J.; Dong, W.; Lan, Y.; Li, D.; Sun, C.; Bu, X. Back Cover: Surfactant-Assisted Phase-Selective Synthesis of New Cobalt MOFs and Their Efficient Electrocatalytic Hydrogen Evolution Reaction. Angew. Chem., Int. Ed. 2017, 56, 13001−13005. (19) Wu, Y.; Wu, X.; Wang, J.; Zhao, J.; Dong, W.; Li, D.; Zhang, Q. Assembly of Two Novel Cd3/(Cd3+Cd5)-Cluster-Based MetalOrganic Frameworks: Structures, Luminescence, and Photocatalytic Degradation of Organic Dyes. Cryst. Growth Des. 2016, 16, 2309− 2316. (20) Wu, Y.; Xu, G.; Dong, W.; Zhao, J.; Li, D.; Zhang, J.; Bu, X. Anionic Lanthanide MOFs as a Platform for Iron-Selective Sensing, Systematic Color Tuning, and Efficient Nanoparticle Catalysis. Inorg. Chem. 2017, 56, 1402−1411. (21) (a) Wang, B.; Lv, X.; Feng, D.; Xie, L.; Zhang, J.; Li, M.; Xie, Y.; Li, J.; Zhou, H. 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. (b) Zhao, D.; Liu, X.; Zhao, Y.; Wang, P.; Liu, Y.; Azam, M.; AlResayes, S. I.; Lu, Y.; Sun, W. Luminescent Cd(II)-organic frameworks with chelating NH2 sites for selective detection of Fe(III) and antibiotics. J. Mater. Chem. A 2017, 5, 15797−15807. (c) Zhang, F.; Yao, H.; Chu, T.; Zhang, G.; Wang, Y.; Yang, Y. A Lanthanide MOF Thin-Film Fixed with Co3O4 Nano-Anchors as a Highly Efficient Luminescent Sensor for Nitrofuran Antibiotics. Chem.Eur. J. 2017, 23, 10293−10300. (d) Hu, Z.; Lustig, W. P.; Zhang, J.; Zheng, C.; Wang, H.; Teat, S. J.; Gong, Q.; Rudd, N. D.; Li, J. Effective Detection of Mycotoxins by a Highly Luminescent Metal-Organic Framework. J. Am. Chem. Soc. 2015, 137, 16209−16215. (e) Chen, M.; Gan, N.; Zhou, Y.; Li, T.; Xu, Q.; Cao, Y.; Chen, Y. An electrochemical aptasensor for multiplex antibiotics detection based on metal ions doped nanoscale MOFs as signal tracers and RecJ(f) exonucleaseassisted targets recycling amplification. Talanta 2016, 161, 867−874. (f) Han, M.; Wen, G.; Dong, W.; Zhou, Z.; Wu, Y.; Zhao, J.; Li, D.; Ma, L.; Bu, X. A heterometallic sodium−europium-cluster-based metal-organic framework as a versatile and water-stable chemosensor for antibiotics and explosives. J. Mater. Chem. C 2017, 5, 8469−8474. (g) Zhang, Q.; Lei, M.; Yan, H.; Wang, J.; Shi, Y. A Water-Stable 3D Luminescent Metal−Organic Framework Based on Heterometallic [EuIII6ZnII] Clusters Showing Highly Sensitive, Selective, and Reversible Detection of Ronidazole. Inorg. Chem. 2017, 56, 7610− 7614. (22) Xu, L.; Xu, G.; Chen, Z. Recent advances in lanthanide luminescence with metal-organic chromophores as sensitizers. Coord. Chem. Rev. 2014, 273-274, 47−62. (23) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339−341. (24) Sheldrick, G. M. SHELXTIntegrated space-group and crystalstructure determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, 71, 3−8. (25) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (26) (a) Richardson, F. S. Terbium(III) and europium(III) ions as luminescent probes and stains for biomolecular systems. Chem. Rev. 1982, 82, 541−552. (b) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent Functional Metal-Organic Frameworks. Chem. Rev. 2012, 112, 1126−1162. (27) (a) Gai, Y.; Jiang, F.; Chen, L.; Wu, M.; Su, K.; Pan, J.; Wan, X.; Hong, M. Europium and Terbium Coordination Polymers Assembled from Hexacarboxylate Ligands: Structures and Luminescent Properties. Cryst. Growth Des. 2014, 14, 1010−1017. (b) Singha, D. K.;

Bhattacharya, S.; Majee, P.; Mondal, S. K.; Kumar, M.; Mahata, P. Optical detection of submicromolar levels of nitro explosives by a submicron sized metal-organic phosphor material. J. Mater. Chem. A 2014, 2, 20908−20915. (28) (a) Chandler, B. D.; Cramb, D. T.; Shimizu, G. K. H. Microporous Metal-Organic Frameworks Formed in a Stepwise Manner from Luminescent Building Blocks. J. Am. Chem. Soc. 2006, 128, 10403−10412. (b) Zhou, Y.; Guo, Y.; Xu, S.; Zhang, L.; Ahmad, W.; Shi, Z. Photoluminescent 3D Lanthanide-Organic Frameworks Based on 2,5-Dioxo-1,4-piperazinylbis(methylphosphonic) Acid Formed via in Situ Cyclodehydration of Glyphosates. Inorg. Chem. 2013, 52, 6338−6345. (29) Acharyya, K.; Mukherjee, P. S. A fluorescent organic cage for picric acid detection. Chem. Commun. 2014, 50, 15788−15791. (30) Xie, Z.; Ma, L.; deKrafft, K. E.; Jin, A.; Lin, W. Porous Phosphorescent Coordination Polymers for Oxygen Sensing. J. Am. Chem. Soc. 2010, 132, 922−923. (31) Kent, C. A.; Liu, D.; Ma, L.; Papanikolas, J. M.; Meyer, T. J.; Lin, W. Light Harvesting in Microscale Metal-Organic Frameworks by Energy Migration and Interfacial Electron Transfer Quenching. J. Am. Chem. Soc. 2011, 133, 12940−12943. (32) Zhang, Y.; Yuan, S.; Day, G.; Wang, X.; Yang, X.; Zhou, H. C. Luminescent sensors based on metal-organic frameworks. Coord. Chem. Rev. 2018, 354, 28−45.

G

DOI: 10.1021/acs.inorgchem.7b03200 Inorg. Chem. XXXX, XXX, XXX−XXX