High Performance for Luminescence Sensing ... - ACS Publications

Mar 4, 2019 - bridge Crystallographic Data Centre, 12 Union Road,. Cambridge CB2 1EZ, UK; fax: ... narcissus 'Sir Winston Churchill'. J. Nat. Prod. 20...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

A Pillar-Layered Zn-LMOF with Uncoordinated Carboxylic Acid Sites: High Performance for Luminescence Sensing Fe3+ and TNP Chengyang Yu, Xiaodong Sun, Lifei Zou, Guanghua Li, Lirong Zhang,* and Yunling Liu* State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, P. R. China

Inorg. Chem. Downloaded from pubs.acs.org by EAST CAROLINA UNIV on 03/04/19. For personal use only.

S Supporting Information *

ABSTRACT: By using the mixed-linker strategy, a new pillarlayered luminescence Zn-LMOF (JLU-MOF71) ([Zn2 Na 2 (TPHC)(4,4-Bipy)(DMF)]·8H2O) (TPHC = [1,1′:2′,1″-terphenyl]-3,3″,4,4′,4″,5′-hexacarboxylic acid, 4,4-bipy = 4,4-bipyridine, DMF = N,N-dimethylformamide) was successfully synthesized and structurally characterized. JLU-MOF71 is constructed by the 4,4bipy pillars and 2D layers which consist of Zn2+ and TPHC ligands, and displays a rare fsh topology. Benefiting from the uncoordinated carboxylate sites in the framework, JLU-MOF71 not only can sensitively detect trace amounts of individual Fe3+ and 2,4,6trinitrophenol (TNP) through luminescence quenching but also exhibits high selectivity when other competing analytes exist. Besides, TNP can also be effectively monitored with the help of the shifting direction of luminescent spectra (red shift) which has rarely been reported before. On the basis of the aforementioned, JLU-MOF71 can be considered as a potential luminescence sensor for detecting Fe3+ and TNP.



sensing,27−29 and drug delivery.30 Among these, luminescent metal−organic frameworks (LMOFs) have been widely studied in biological and chemical detection because of their unique advantages, such as short response time, handy operability, and high distinguishability.31−33 When there exist some functional sites (−OH, pyridine sites, amide groups) in the framework of LMOFs, it will tremendously enhance the sensing effect through luminescence quenching or enhancement.34−36 For example, Neppolian et al. reported an NH2functionalized Fe-MOF which could improve the sensing effects for explosives through hydrogen bonding between the amine group of MOF and nitro-aromatic compounds.37 Another fascinating example was reported by Gao and coworkers. In their work, an OH-functionalized Eu-MOF was synthesized, and the −OH group made a great contribution for sensing Fe3+.38 Recently, lots of LMOF materials were synthesized and used to sense metal cations or explosives.39−41 However, LMOFs with functional sites that exhibit high performance for detecting both metal cations and explosives have rarely been reported.42−45 Hence, it is imperative to design and synthesize such versatile materials which can not only save costs and make sensing more efficiently but also prevent cations and NAEs from doing harm to the environment and human body more effectively.

INTRODUCTION Iron ion is indispensable in metabolic processes for hemoglobin formation, brain and muscle function, and the synthesis processes of DNA and RNA.1,2 Both deficiency and overload of irons can cause serious health disorders such as insomnia, skin diseases, and declining immunity.3,4 With the rapid development of industry, iron ion gradually becomes one of the mainly inorganic contaminants. As a result, it will lead to fearful damage to both environment and human lives. Similarly, nitro-aromatic explosives (NAEs) such as nitrobenzene (NB), 3-nitrotoluene (3-NT), 2,4,6-trinitrophenol (TNP), 4-nitrophenol (4-NP), 4-nitrotoluene (4-NT), and 2,4-dinitrotoluene (2,4-DNT), due to their explosive nature, are listed as one of the hazards of antiterrorism or environmental security in different countries. Among NAEs, TNP has drawn great attention because of its strong explosive and harmfulness.5−7 Herein, it is highly imminent to monitor and track Fe3+ and TNP. Traditional detection methods, for instance, mass spectrometry (MS),8 atomic absorption spectrometry (AAS),9 ion mobility spectrum,10 and gas chromatography,11 are time-consuming, low sensitivity, and selectivity with high cost.12−17 Therefore, seeking for an effective method, which is used to detect contaminants such as Fe3+ and NAEs, has always been an exceedingly urgent issue. Metal−organic frameworks (MOFs) as emerging multifunctional-organic frameworks materials have been efficiently employed in numerous applications, such as gas storage and separation,18−21 heterogeneous catalysis,22−26 luminescence © XXXX American Chemical Society

Received: January 22, 2019

A

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. (a) Simplified diagrams of the coordination mode of TPHC, 4,4-bipy, and Zn2+. (b) Polyhedral view of the framework along the [001] direction. (c) The space-filling model of the channel along the [001] direction. (d) The formation of the 3D skeleton from 2D layers and the pillars of 4,4-bipy. Color scheme: carbon = gray, nitrogen = blue, oxygen = red, zinc = green, sodium = yellow, hydrogen = white.



Given the above considerations, an unexplored rigid and conjugated ligand (TPHC) and 4,4-bipy were selected to construct porous MOFs materials for studying luminescence sensing. The reasons we choose the TPHC ligand are as follows: (1) the ligand is rigid and conjugated which is benefiting for enhancing the luminescence of MOFs materials; (2) the ligand possesses many carboxylic acid groups which can coordinate with the analytes. These characteristic attributes are helpful for luminescence recognition. Interestingly, we successfully constructed a pillar-layered Zn-LMOF by using the mixed-linker strategy. 4,4-bipy as a pillar supports the 2D layers and forms JLU-MOF71. In addition, there exist many uncoordinated carboxylate sites in the channels of JLUMOF71. It can not only bind to metal ions but also give electrons to the electron-deficient nitro-aromatic explosives, which make it detect metal ions and nitro-aromatic explosives more effectively. Thus, the detecting performance of JLUMOF71 is commendable and the Ksv values of sensing Fe3+ and TNP are 5.77 × 104 M−1 and 8.47 × 104 M−1, respectively, which exceed many previous reported MOFs materials. Notably, TNP can also be distinguished by shifting wavelength (red shift), which has rarely been reported before. Therefore, JLU-MOF71 is proved to be an excellent candidate for monitoring Fe3+ and TNP, simultaneously.

EXPERIMENTAL SECTION

Materials and Methods. All solvents and reagents were procured from the pharmaceutical companies and employed without purification. Powder X-ray diffraction (PXRD) was performed with a Rigaku D/max-2550 diffractometer by Cu−Kα radiation (λ = 1.5418 Å). The thermogravimetric data (TGA) were recorded with a TGA Q500 thermogravimetric analyzer with a heating rate of 10 °C/ min between 25 and 800 °C. C, H, and N elemental analyses were carried out on a Vario MICRO (Elementar, Germany). N2, CO2, CH4, C2H6, and C3H8 gas adsorption experiments were recorded with Micromeritics ASAP 2420 and Micromeritics 3-Flex instruments. The luminescence spectra were obtained by a FLUOROMAX-4 spectrophotometer. The UV−vis spectra of liquid state were acquired on a SHIMADZU UV-2450. The PerkinElmer Optima 3300 Dv spectrometer was utilized for inductively coupled plasma (ICP) analyses. Synthesis of JLU-MOF71. A single crystal of JLU-MOF71 was synthesized via solvothermal reaction of Zn(NO3)2 6H2O (15 mg, 0.05 mmol), TPHC ligand (4 mg, 0.008 mmol), NaNO3 (2 mg, 0.023 mmol), 4,4-bipy (5 mg, 0.026 mmol), and HNO3 (0.1 mL) (2.2 mL HNO3 in 10 mL DMF) in DMF (1 mL) at 85 °C for 72 h. The colorless crystals were generated and washed with N,N-dimethylformamide (DMF), then dried in air (yield 68% based on TPHC). Elemental analysis data (%) for JLU-MOF71: C, 42.52; H, 3.63; N, 4.41, found: C, 42.89; H, 3.77; N, 4.06. The experimental PXRD pattern of JLU-MOF71 can match well with the simulated one, indicating the phase purity of the as-synthesized product (Figure S3). B

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. (a) Luminescence quenching efficiency of JLU-MOF71 in different metal ions. (b) Influence on the emission spectra of JLU-MOF71 decentralized in DMF by gradual addition of targeted Fe3+ (10 mM, 20 μL addition every time; inset: SV plots of targeted Fe3+). Single-Crystal X-ray Crystallography. The Bruker Apex II CCD diffractometer was applied for collecting X-ray data. It used a graphite monochromator with Mo−Kα (λ = 0.71073 Å) radiation at 20 °C. Full-matrix least-squares technique on F2 with the SHELXTL97 process was used to solve structure.46 All atoms (except H) were located by anisotropic technology. The formula was calculated from the TGA and C, H, and N elemental analyses. The integrated crystallographic data are listed in Table S1. The characteristic bond lengths and angles are shown in Table S2. Topology data for the JLUMOF71 were obtained by using TOPOS 4.0.47 Luminescence Sensing Experiments. The solid state luminescence of JLU-MOF71 and the TPHC ligand was performed at 25 °C. The luminescent sensing experiments were evaluated in DMF solutions. For cations sensing measurements, 2 mg ground samples were added into 2 mL diverse solutions of M(NO3)x (Mx+ = K+, Na+, Li+, Zn2+, Mg2+, Co2+, Ni2+, Pb2+, Cu2+, Fe3+, Cr3+, Al3+, 1 × 10−3 mol L−1), and the suspension was obtained by ultrasonication. For nitroaromatic explosives measurements, a range of suspensions of 3-NT, TNP, 4-NP, 2,4-DNT, NB, and 4-NT were prepared at the same concentration (1 × 10−4 mol L−1), and luminescence data were collected after ultrasonication for 30 min. The luminescence titration experiments of metal ions and nitro-aromatic explosives were carried out by adding DMF solution of Fe3+ and TNP step by step. The Ksv can be gained by the Stern−Volmer (SV) equation: I0/I = 1 + Ksv[C].48 I0 and I are luminescence intensity of the blank sample and the addition of analyte. Ksv is Stern−Volmer quenching constant. The equation L = 3Sb/K is applied to calculate limits of detection (LOD).49 Sb and K are standard deviation for blank specimens and slope of the Stern−Volmer, respectively.

clearly illustrates the forming process of JLU-MOF71 the by mixed-linker strategy. Each TPHC ligand connects to six Zn2+ ions, forming a 2D layer. Then the 2D layer is further linked by 4,4-bipy to generate a 3D framework. Gas Adsorption and Separation Behavior. The permanent porosity and surface area of JLU-MOF71 were explored by reversible N2 sorption measurements at 77 K, which displays a reversible type-I isotherm characteristic of microporous materials (Figure S5). The BET surface area and Langmuir surface area for JLU-MOF71 were calculated to be 322 and 436 m2 g−1, respectively. Because of its permanent porosity, the exploration of multifarious gas adsorption was carried out for JLU-MOF71 (Figures S6−S9). The CO2 adsorption amount for JLU-MOF71 is 47 and 26 cm3 g−1 at 273 and 298 K under 1 bar, respectively. Although the BET surface area of JLU-MOF71 is lower, its CO2 sorption capacity surpasses that of some MOFs materials with higher BET surface area, such as JLU-Liu31, MOF-892, and MOF893.50,51 Besides, the selectivity for the following gas pairs, CO2/CH4, C2H6/CH4, and C3H8/CH4, was also calculated (Figure S10). The selectivity of C3H8 over CH4 is 53, which is comparable to JLU-Liu6.52 Sensing of Cations. Taking advantage of exposed carboxylate groups in the framework of JLU-MOF71, luminescence exploration experiments were performed. First, the solid state luminescent properties of TPHC ligand and JLU-MOF71 were carried out at room temperature. The TPHC ligand exhibits luminescence emissions at 417 nm under the excitation wavelength of 360 nm, and the emission spectrum of JLU-MOF71 was shown at 449 nm under the same excitation (Figure S11). The results indicate that the luminescence of JLU-MOF71 can be mainly ascribed to the rigid and conjugated ligand because of its similar emission wavelengths. Then the luminescence sensing tests were carried out. As shown in Figure 2a, the luminescence of JLU-MOF71 is almost completely quenched in Fe3+, while the quenching percentages of other ions are almost negligible. It reveals that the detecting performance of JLU-MOF71 for Fe3+ is better than other cations. This may be due to the strong bonding between uncoordinated carboxylate sites and Fe3+. The luminescence titration experiment was further carried out to explore the recognition ability of JLU-MOF71 for Fe3+. As shown in Figure 2b, as the concentration increased, the luminescence intensity gradually decreased and the luminescence of JLU-MOF71 almost completely quenched when 200 μL of Fe3+ DMF solution (10−3 mol/L) was added. The Stern−Volmer quenching constant (Ksv) value, which is a



RESULTS AND DISCUSSION Single-crystal data show that JLU-MOF71 crystallizes in a monoclinic crystal system and shows a 3D framework with space group of P2/c. As depicted in Figure 1a, each zinc ion coordinates to three O atoms from three TPHC ligands and one N atom from one 4,4-bipy ligand and exhibits tetrahedral geometry. TPHC ligand is tortile with the dihedral angles of 55° (Figure S1a) and connects with six zinc ions which can be simplified as a pentahedron. Similarly, 4,4-bipy also possesses a distorted dihedral angle of 29° (Figure S1b) and links two zinc ions. It can be simplified as a linear rod. Therefore, the whole structure can be regarded as a 4,6-connected network that belongs to fsh topology with a Schläfli symbol of (43·63) (Figure 1b). As illustrated in Figure 1c, the space-filling model shows a straight hexagonal channel along the [001] direction with a diameter of about 8.5 Å × 6.6 Å (considering the van der Waals radius). Furthermore, there exist many uncoordinated carboxylic acid sites in this channel. Due to the coordination of uncoordinated carboxyl groups and sodium ions, the small square pores are blocked (Figure S2). Figure 1d C

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

solution. This may indicate the bonding between uncoordinated carboxylate groups and Fe3+. Finally, the UV−vis absorption peak of Fe3+ and the luminescence emission peak of JLU-MOF71 were partially overlapped (Figure S13). This proves that the efficient sensing effect for Fe3+ may be ascribed to the resonance energy transfer from the framework (donor) to the analyte (acceptor).55 Sensing of Nitro-Aromatic Explosives. It is well-known that the nitro-explosives are electron-deficient. If there appear exposed sites in the MOFs materials, the sensing effect for nitro-explosives will be improved. Given uncoordinated carboxyl sites in JLU-MOF71, the sensing experiments on nitro-explosives were carried out. As shown in Figure 4a, the luminescence emission intensity of JLU-MOF71 was almost completely quenched in TNP solution (10−4 mol/L). However, the quenching percentage of 4-NP is only 60%, and that of other NAEs does not change too much. Herein, JLU-MOF71 can be applied to detect TNP effectively. Luminescence titration experiments were employed for further sensing experiments (Figure 4b). The LOD and Ksv of TNP reached up to 0.64 μm and 8.47 × 104 M−1, respectively, which can be compared with some previously reported MOFs (Table S6). Besides, as can be seen from Figure 5, the emission

significant criterion for quantification of luminescence quenching efficiency, of JLU-MOF71 for sensing Fe3+ was 5.77 × 104 M−1. This value is higher than some other reported MOFs. Therefore, JLU-MOF71 is able to detect Fe3+ with excellent sensitivity. Furthermore, it was also observed that the quenching efficiency of JLU-MOF71 toward Fe3+ is hardly affected by other disturbing ions, which confirms that JLUMOF71 can sense Fe3+ selectively (Figure 3).

Figure 3. Luminescence intensities of JLU-MOF71 soaked in the pure individual DMF solutions of M(NO3)x (1 × 10−3 mol L−1; blue color) and mingled metal ions containing Fe3+ ion (1 × 10−3 mol L−1; red color) at an excitation wavelength of 360 nm.

LOD is another important criterion for assessing the sensing effect of the materials. It was calculated, and its value can reach up to 6.4 μm, which surpasses that of many reported MOFs materials (Table S4). Generally, the mechanism of ion luminescent quenching are (1) integrated framework collapses; (2) ions exchange between analyte and skeleton; (3) interaction between targeted ion and uncoordinated functional group from the framework; (4) fluorescence resonance energy transfer (FRET).53,54 Hence, auxiliary tests were performed to explore the possible luminescent quenching mechanisms of JLU-MOF71. The PXRD patterns of JLU-MOF71 are consistent with the JLU-MOF71 samples after sensing tests (Figure S12), which confirms that the luminescence quenching was not caused by the framework collapse. As shown in Table S5, when JLU-MOF71 was soaked in pure solution of DMF and the iron ion, the contents of zinc and sodium ions were basically unchanged. While the content of iron ion changed obviously after JLU-MOF71 being soaked in the iron ion

Figure 5. Luminescence intensities of JLU-MOF71 soaked in the pure DMF solutions of NAEs (1 × 10−4 mol L−1; blue color) and mingled NAEs containing TNP (1 × 10−4 mol L−1; red color) under an excitation wavelength of 360 nm.

intensity of JLU-MOF71 nearly unchanged when other NAEs were added, while it sharply decreased when TNP was added into suspensions. Interestingly, the maximum emission wavelength of JLU-MOF71 was red-shifted 30 nm compared to the

Figure 4. (a) Luminescence quenching efficiency of JLU-MOF71 in different NAEs. (b) Influence on the emission spectra of JLU-MOF71 decentralized in DMF by gradual addition of targeted TNP (1 mM, 20 μL addition every time; inset: SV plots of targeted TNP). D

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX

Inorganic Chemistry



emission wavelength before adding TNP (Figure S14), which can be also regarded as a useful method for distinguishing TNP from other NAEs.56,57 It indicates that JLU-MOF71 shows high anti-interference capacity in the existence of other competitive NAEs. Additionally, the regenerative capability is significant equipment for practical application. Thus, recycling performance has gained interests. JLU-MOF71 was dispersed in the DMF solution (10−4 M) of TNP for the first time test, and then the suspension after the test was centrifuged and washed several times with DMF solution for the next time test. Fortunately, the luminescence intensity and quenching efficiencies of JLU-MOF71 nearly remained unchanged after five cycles (Figure 6). This elucidates that JLU-MOF71 exhibits high sensitivity, selectivity, and reusability for sensing TNP.

Article

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b00204. Crystallographic data for JLU-MOF71, crystal structure figures, PXRD patterns, TGA curves, gases adsorption properties, and UV−vis adsorption spectra for JLUMOF71 (PDF) Accession Codes

CCDC 1889036 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 Authors

*E-mail: [email protected] (L.Z.). *E-mail: [email protected] (Y.L.). ORCID

Guanghua Li: 0000-0003-3029-8920 Yunling Liu: 0000-0001-5040-6816 Notes

The authors declare no competing financial interest.



Figure 6. Five cycle tests of JLU-MOF71 for sensing TNP.

ACKNOWLEDGMENTS The authors gratefully acknowledge the financial support of the National Natural Science Foundation of China (Nos. 21771078, 21671074, and 21621001), the 111 Project (B17020), and the National Key Research and Development Program of China (2016YFB0701100).

Quenching Mechanism of Explosives. In order to explain the underlying quenching mechanism of explosives, the PXRD patterns of JLU-MOF71 samples after soaking TNP were tested (Figure S15). The result indicates that the framework remains intact and does not substantially affect luminescence detection. As shown in Figure S16, there is an obvious overlap between the UV absorption peak of TNP and the luminescence emission peak of JLU-MOF71. Hence, the luminescence quenching may be due to the competition between analytes and organic ligands in the framework.58 Furthermore, the electron-deficient TNP can accept electrons from the framework of the electron-rich JLU-MOF71.59



REFERENCES

(1) Gogoi, C.; Biswas, S. A new quinoline based luminescent Zr(IV) metal-organic framework for the ultrasensitive recognition of 4nitrophenol and Fe(III) ions. Dalton Trans. 2018, 47, 14696−14705. (2) Barba-Bon, A.; Costero, A. M.; Gil, S.; Parra, M.; Soto, J.; Martínez-Máñez, R.; Sancenón, F. A new selective fluorogenic probe for trivalent cations. Chem. Commun. 2012, 48, 3000−3002. (3) Zhao, D.; Liu, X.-H.; Zhao, Y.; Wang, P.; Liu, Y.; Azam, M.; AlResayes, S. I.; Lu, Y.; Sun, W.-Y. 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. (4) Zhao, Q.; Li, F.; Huang, C. Phosphorescent chemosensors based on heavy-metal complexes. Chem. Soc. Rev. 2010, 39, 3007−3030. (5) Hu, J.; Wu, K.; Dong, S.; Zheng, M. A luminescent Cd(II)-MOF as recyclable bi-responsive sensor for detecting TNP and iron(III)/ silver(I) with high selectivity and sensitivity. Polyhedron 2018, 153, 261−267. (6) Salinas, Y.; Martínez-Máñez, R.; Marcos, M. D.; Sancenón, F.; Costero, A. M.; Parra, M.; Gil, S. Optical chemosensors and reagents to detect explosives. Chem. Soc. Rev. 2012, 41, 1261−1296. (7) Thomas, S. W.; Joly, G. D.; Swager, T. M. Chemical sensors based on amplifying fluorescent conjugated polymers. Chem. Rev. 2007, 107, 1339−1386. (8) Ingkaninan, K.; Hazekamp, A.; de Best, C. M.; Irth, H.; Tjaden, U. R.; van der Heijden, R.; van der Greef, J.; Verpoorte, R. The application of HPLC with on-line coupled UV/MS-biochemical detection for isolation of an acetylcholinesterase inhibitor from narcissus ‘Sir Winston Churchill’. J. Nat. Prod. 2000, 63, 803−806.



CONCLUSIONS In summary, we successfully constructed a pillar-layered ZnLMOF with exposed carboxylate sites by a mixed-linker strategy. 4,4-bipy as a pillar extended the 2D layers into the 3D framework. There exist many uncoordinated carboxylate sites in the framework of JLU-MOF71, which make a great contribution to detecting Fe3+ and TNP with high sensitivity and selectivity, simultaneously. In addition, the quenching constant (Ksv) values of JLU-MOF71 toward TNP and Fe3+ can reach up to 104 M−1 level, and the LODs are 6.4 and 0.64 μm, respectively. The above values precede those of many other reported MOFs materials. It is worth mentioning that a red shift occurred in the process of detecting TNP, which can also be regarded as a means of luminescence sensing. The energy and electron transfer play significant roles in the quenching mechanisms of Fe3+ and TNP. In brief, this work broadens the horizon of design and synthesis of MOFs with luminescence sensing performance. E

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

based on novel lanthanide metal-organic frameworks. Inorg. Chem. 2013, 52, 8082−8090. (29) Zhou, J.-M.; Shi, W.; Li, H.-M.; Li, H.; Cheng, P. Experimental studies and mechanism analysis of high-sensitivity luminescent sensing of pollutional small molecules and ions in Ln4O4 cluster based microporous metal-organic frameworks. J. Phys. Chem. C 2014, 118, 416−426. (30) Su, H.; Sun, F.; Jia, J.; He, H.; Wang, A.; Zhu, G. A highly porous medical metal-organic framework constructed from bioactive curcumin. Chem. Commun. 2015, 51, 5774−5777. (31) Hu, Z.; Deibert, B. J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815−5840. (32) Kan, W.-Q.; Wen, S.-Z. A pH-controlled fluorescent sensor and logic gate based on a 1D→2D→3D Cd(II)-containing coordination polymer. Polyhedron 2017, 128, 169−174. (33) Meng, X.; Wei, M.-J.; Wang, H.-N.; Zang, H.-Y.; Zhou, Z.-Y. Multifunctional luminescent Zn(II)-based metal-organic framework for high proton-conductivity and detection of Cr3+ ions in the presence of mixed metal ions. Dalton Trans. 2018, 47, 1383−1387. (34) Wong, K. L.; Law, G. L.; Yang, Y. Y.; Wong, W. T. A highly porous luminescent terbium-organic framework for reversible anion sensing. Adv. Mater. 2006, 18, 1051−1054. (35) Chen, B.; Wang, L.; Xiao, Y.; Fronczek, F. R.; Xue, M.; Cui, Y.; Qian, G. A luminescent metal-organic framework with lewis basic pyridyl sites for the sensing of metal ions. Angew. Chem. 2009, 121, 508−511. (36) Xie, D.; Ma, Y.; Gu, Y.; Zhou, H.; Zhang, H.; Wang, G.; Zhang, Y.; Zhao, H. Bifunctional NH2-MIL-88(Fe) metal-organic framework nanooctahedra for highly sensitive detection and efficient removal of arsenate in aqueous media. J. Mater. Chem. A 2017, 5, 23794−23804. (37) Karthik, P.; Pandikumar, A.; Preeyanghaa, M.; Kowsalya, M.; Neppolian, B. Amino-functionalized MIL-101(Fe) metal-organic framework as a viable fluorescent probe for nitroaromatic compounds. Microchim. Acta 2017, 184, 2265−2273. (38) Xu, H.; Dong, Y.; Wu, Y.; Ren, W.; Zhao, T.; Wang, S.; Gao, J. An -OH group functionalized MOF for ratiometric Fe3+ sensing. J. Solid State Chem. 2018, 258, 441−446. (39) 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. (40) Ye, J.; Bogale, R. F.; Shi, Y.; Chen, Y.; Liu, X.; Zhang, S.; Yang, Y.; Zhao, J.; Ning, G. A water-stable dual-channel luminescence sensor for UO22+ ions based on an anionic terbium(III) metal-organic framework. Chem. - Eur. J. 2017, 23, 7657−7662. (41) Yan, B. Lanthanide-functionalized metal-organic framework hybrid systems to create multiple luminescent centers for chemical sensing. Acc. Chem. Res. 2017, 50, 2789−2798. (42) Nagarkar, S. S.; Desai, A. V.; Ghosh, S. K. Engineering metalorganic frameworks for aqueous phase 2,4,6-trinitrophenol (TNP) sensing. CrystEngComm 2016, 18, 2994−3007. (43) Rachuri, Y.; Parmar, B.; Bisht, K. K.; Suresh, E. Mixed ligand two dimensional Cd(II)/Ni(II) metal organic frameworks containing dicarboxylate and tripodal N-donor ligands: Cd(II) MOF is an efficient luminescent sensor for detection of picric acid in aqueous media. Dalton Trans. 2016, 45, 7881−7892. (44) Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T. Luminescent metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1330−1352. (45) Chen, B.; Xiang, S.; Qian, G. Metal-organic frameworks with functional pores for recognition of small molecules. Acc. Chem. Res. 2010, 43, 1115−1124. (46) Sheldrick, G. M. A short history of shelx. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (47) Blatov, V. A.; Shevchenko, A. P.; Proserpio, D. M. Applied topological analysis of crystal structures with the program package topos Pro. Cryst. Growth Des. 2014, 14, 3576−3586.

(9) Bohdálková, L.; Bohdálek, P.; Břízová, E.; Pacherová, P.; Kuběna, A. A. Atmospheric metal pollution records in the Kovárš ká Bog (Czech Republic) as an indicator of anthropogenic activities over the last three millennia. Sci. Total Environ. 2018, 633, 857−874. (10) Xu, X.-Y.; Yan, B. Eu(III)-functionalized ZnO@MOF heterostructures: integration of pre-concentration and efficient charge transfer for the fabrication of a ppb-level sensing platform for volatile aldehyde gases in vehicles. J. Mater. Chem. A 2017, 5, 2215−2223. (11) Hu, J.; Cheng, T.; Dong, S.; Zhou, C.; Huang, X.; Zhang, L. Multifunctional luminescent Cd(II)-based metal-organic framework material for highly selective and sensitive sensing 2,4,6-trinitrophenol (TNP) and Fe3+ cation. Microporous Mesoporous Mater. 2018, 272, 177−183. (12) Nagarkar, S. S.; Desai, A. V.; Ghosh, S. K. A fluorescent metalorganic framework for highly selective detection of nitro explosives in the aqueous phase. Chem. Commun. 2014, 50, 8915−8918. (13) Guo, X.-Y.; Zhao, F.; Liu, J.-J.; Liu, Z.-L.; Wang, Y.-Q. An ultrastable zinc(II)-organic framework as a recyclable multiresponsive luminescent sensor for Cr(III), Cr(VI) and 4-nitrophenol in the aqueous phase with high selectivity and sensitivity. J. Mater. Chem. A 2017, 5, 20035−20043. (14) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal-organic frameworks. Chem. Rev. 2012, 112, 1126−1162. (15) 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. (16) Wang, L.; Fan, G.; Xu, X.; Chen, D.; Wang, L.; Shi, W.; Cheng, P. Detection of polychlorinated benzenes (persistent organic pollutants) by a luminescent sensor based on a lanthanide metalorganic framework. J. Mater. Chem. A 2017, 5, 5541−5549. (17) Wang, Y.; Jia, W.; Chen, R.; Zhao, X.-J.; Wang, Z.-L. Guestinduced SC-SC transformation within the first K/Cd heterodimetallic triazole complex: a luminescent sensor for high-explosives and cyano molecules. Chem. Commun. 2017, 53, 636−639. (18) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D.-W. Hydrogen storage in metal-organic frameworks. Chem. Rev. 2012, 112, 782−835. (19) Li, J.-R.; Ma, Y.; McCarthy, M. C.; Sculley, J.; Yu, J.; Jeong, H.K.; Balbuena, P. B.; Zhou, H.-C. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coord. Chem. Rev. 2011, 255, 1791−1823. (20) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (21) Murray, L. J.; Dincă, M.; Long, J. R. Hydrogen storage in metalorganic frameworks. Chem. Soc. Rev. 2009, 38, 1294−1314. (22) Ma, L.; Abney, C.; Lin, W. Enantioselective catalysis with homochiral metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1248−1256. (23) Aguado, S.; Canivet, J.; Farrusseng, D. Facile shaping of an imidazolate-based MOF on ceramic beads for adsorption and catalytic applications. Chem. Commun. 2010, 46, 7999−8001. (24) Dhakshinamoorthy, A.; Garcia, H. Catalysis by metal nanoparticles embedded on metal-organic frameworks. Chem. Soc. Rev. 2012, 41, 5262−5284. (25) Falkowski, J. M.; Sawano, T.; Zhang, T.; Tsun, G.; Chen, Y.; Lockard, J. V.; Lin, W. Privileged phosphine-based metal-organic frameworks for broad-scope asymmetric catalysis. J. Am. Chem. Soc. 2014, 136, 5213−5216. (26) Li, Y.-Z.; Wang, H.-H.; Yang, H.-Y.; Hou, L.; Wang, Y.-Y.; Zhu, Z. An uncommon carboxyl-decorated metal-organic framework with selective gas adsorption and catalytic conversion of CO2. Chem. - Eur. J. 2018, 24, 865−871. (27) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105−1125. (28) Zhou, J.-M.; Shi, W.; Xu, N.; Cheng, P. Highly selective luminescent sensing of fluoride and organic small-molecule pollutants F

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (48) Lin, R.-B.; Li, F.; Liu, S.-Y.; Qi, X.-L.; Zhang, J.-P.; Chen, X.-M. A noble-metal-free porous coordination framework with exceptional sensing efficiency for oxygen. Angew. Chem. 2013, 125, 13671−13675. (49) Buragohain, A.; Yousufuddin, M.; Sarma, M.; Biswas, S. 3D luminescent amide-functionalized cadmium tetrazolate framework for selective detection of 2,4,6-trinitrophenol. Cryst. Growth Des. 2016, 16, 842−851. (50) Yao, S.; Sun, X.; Liu, B.; Krishna, R.; Li, G.; Huo, Q.; Liu, Y. Two heterovalent copper-organic frameworks with multiple secondary building units: high performance for gas adsorption and separation and I2 sorption and release. J. Mater. Chem. A 2016, 4, 15081−15087. (51) Nguyen, P. T. K.; Nguyen, H. T. D.; Nguyen, H. N.; Trickett, C. A.; Ton, Q. T.; Gutiérrez-Puebla, E.; Monge, M. A.; Cordova, K. E.; Gándara, F. New metal-organic frameworks for chemical fixation of CO2. ACS Appl. Mater. Interfaces 2018, 10, 733−744. (52) Wang, D.; Zhao, T.; Cao, Y.; Yao, S.; Li, G.; Huo, Q.; Liu, Y. High performance gas adsorption and separation of natural gas in two microporous metal-organic frameworks with ternary building units. Chem. Commun. 2014, 50, 8648−8650. (53) Wang, B.; Yang, Q.; Guo, C.; Sun, Y.; Xie, L.-H.; Li, J.-R. Stable Zr(IV)-based metal-organic frameworks with predesigned functionalized ligands for highly selective detection of Fe(III) ions in water. ACS Appl. Mater. Interfaces 2017, 9, 10286−10295. (54) Wang, R.; Dong, X.-Y.; Xu, H.; Pei, R.-B.; Ma, M.-L.; Zang, S.Q.; Hou, H.-W.; Mak, T. C. W. A super water-stable europiumorganic framework: guests inducing low-humidity proton conduction and sensing of metal ions. Chem. Commun. 2014, 50, 9153−9156. (55) Liu, L.; Wang, Y.; Lin, R.; Yao, Z.; Lin, Q.; Wang, L.; Zhang, Z.; Xiang, S. Two water-stable lanthanide metal-organic frameworks with oxygen-rich channels for fluorescence sensing of Fe(III) ions in aqueous solution. Dalton Trans. 2018, 47, 16190−16196. (56) Pramanik, S.; Hu, Z.; Zhang, X.; Zheng, C.; Kelly, S.; Li, J. A systematic study of fluorescence-based detection of nitroexplosives and other aromatics in the vapor phase by microporous metal-organic frameworks. Chem. - Eur. J. 2013, 19, 15964−15971. (57) He, H.; Song, Y.; Sun, F.; Bian, Z.; Gao, L.; Zhu, G. A porous metal-organic framework formed by a V-shaped ligand and Zn(II) ion with highly selective sensing for nitroaromatic explosives. J. Mater. Chem. A 2015, 3, 16598−16603. (58) 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. (59) Lv, R.; Wang, J.; Zhang, Y.; Li, H.; Yang, L.; Liao, S.; Gu, W.; Liu, X. An amino-decorated dual-functional metal-organic framework for highly selective sensing of Cr(III) and Cr(VI) ions and detection of nitroaromatic explosives. J. Mater. Chem. A 2016, 4, 15494−15500.

G

DOI: 10.1021/acs.inorgchem.9b00204 Inorg. Chem. XXXX, XXX, XXX−XXX