Detection of Pesticides in Aqueous Medium and in Fruit Extracts

Department of Chemistry, Siksha-Bhavana, Visva-Bharati University , Santiniketan 731235 , West Bengal , India. Inorg. Chem. , 2018, 57 (19), pp 12155â...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/IC

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

Detection of Pesticides in Aqueous Medium and in Fruit Extracts Using a Three-Dimensional Metal−Organic Framework: Experimental and Computational Study Debal Kanti Singha,†,‡ Prakash Majee,‡ Saurodeep Mandal,‡ Sudip Kumar Mondal,*,‡ and Partha Mahata*,† †

Department of Chemistry, Jadavpur University, Kolkata 700032, India Department of Chemistry, Siksha-Bhavana, Visva-Bharati University, Santiniketan 731235, West Bengal, India



Inorg. Chem. Downloaded from pubs.acs.org by WESTERN UNIV on 09/17/18. For personal use only.

S Supporting Information *

ABSTRACT: A new, three-dimensional cadmium based metal− organic framework [Cd3(PDA)1(tz)3Cl(H2O)4]·3H2O {PDA = 1,4-phenylenediacetate and tz = 1,2,4-triazolate}, 1, has been successfully synthesized using slow diffusion method at room temperature. The structure of compound 1 has been determined using single crystal X-ray diffraction. The triazolate ligands connect three different types of octahedral Cd2+ ions to form a two-dimensional structure. The chloride ion and PDA ligands connect the two-dimensional layers to form a three-dimensional structure. The phase purity of 1 was confirmed by powder X-ray diffraction, thermogravimetric analysis, and IR spectroscopy. Aqueous dispersion of compound 1 gives intense luminescence emission at 290 nm upon excitation at 225 nm. This emission was used for the luminescence based detection of pesticides, especially azinphos-methyl, chlorpyrifos, and parathion in aqueous medium. The selectivity of pesticide detection remains unaltered even in the presence of surfactant molecules. The mechanisms of luminescence quenching were successfully explained by the combination of absorption of excitation light, resonance energy transfer, and the possibility of electron transfer. Experimental findings are also well supported by the density functional theory calculations. Selectivity of pesticides detection in real samples such as apple and tomato juice has also been observed.



INTRODUCTION The rapid increase of population demands the production of vegetables and fruits in a short time.1 For this purpose, the use of toxic organophosphorus pesticides is growing in cultivation, animal farming, and aquafarming to boost the production, as the pesticides protect these living objects from harmful microorganisms.2−7 To combat against some vector-borne diseases like dengue and malaria and for wiping out the unwanted plants, pesticides have been recently used.8 It has been observed that nearly 1% of the total applied pesticides literally are used by the targeted species and rest of the 99% applied pesticides are bioaccumulated through the food chain.9,10 Azinphos-methyl, chlorpyrifos, and parathion are widely used pesticides among the organophosphorus pesticides used in recent times.11−15 Accidental spillages during transport and storage and agriculture runoff from treated land are the main sources of pesticide contamination. Besides that, inhalation through the respiratory system and skin penetration upon dermal exposure are the well-defined paths of pesticide toxicity. Pesticides are also a major concern to the farmers who mix and apply these pesticides, as pesticides can be absorbed through the layers of the epidermis at different penetration rates.16,17 Actually, the organophosphorus pesticides permanently damage acetylcholinesterase enzyme which is respon© XXXX American Chemical Society

sible for proper functioning of the nervous system by hydrolyzing the neurotransmitter acetylcholine and maintaining the proper neurotransmission within cholinergic networks.18−27 Due to their high toxicity, the U.S. Environmental Protection Agency (EPA) categorized organophosphorus pesticides as toxicological class I (extremely toxic).28,29 Therefore, for the sake of environmental protection and public health, it is important to design a method for the rapid detection of organophosphorus pesticides in aqueous medium and in real samples. Detection of organophosphorus pesticides has been done using various methods such as liquid chromatography triple quadruple-tandem mass spectrometry,30−32 gas chromatography−nitrogen phosphorus detection,33,34 micellar electrokinetic chromatography,35 potentiometry,36,37 capillary electrophoresis,38−40 flow injection spectophotometric analysis,41 and cyclic voltametry with electrochemical liquid-phase microextraction.42,43 Although these methods are very efficient with respect to sensitivity and limit of detection, they have some disadvantages like expensive, complicated sample preparation methods and requirement of longer time for analysis, Received: June 26, 2018

A

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Inorganic Chemistry



complicated instrumentations and need for highly trained personnel, and also they have some portability issues during infield use. In the recent past, luminescence based detection methods have gained remarkable interest for their various beneficial characteristics such as efficient selectivity and sensitivity, portability, fast response time, and usefulness in both medium solution and solid state.44−49 Luminescence based pesticide detections have already been explored with different types of materials such as polymer,50−52 coordination compounds,53−63 quantum dots,64,65 and nanometarials.66−68 However, these types of materials have some disadvantages such as complicated preparation processes, poor photostability, harmfulness toward the environment, and the absence of molecular organization.69−72 Metal−organic frameworks (MOFs) are a relatively new type of compound with diverse structural characteristics formed through the self-assembly of metal ions and organic ligands.73−76 During the last two decades, research on MOFs has been popularized based on their properties and useful applications originated from their structures.77−100 Recently, luminescence properties of MOFs have been used very efficiently for detection experiments based on the structure of the MOFs and molecular level interactions between MOFs and analytes through the chemical bonding and noncovalent type of interactions along with energy and electron transfer processes.101−104 During the past decade, for the detection of small molecules, hazardous organic compounds, cations, anions, pH, and temperature, many researchers have used luminescent MOF compounds.105−130 Detection of pesticides in liquid medium using luminescent MOFs has also been reported in a few cases.131−135 For real life applications, the detection of pesticide in aqueous medium is important as pesticides may mix with water from the field of cultivation and aquafarm. For this purpose, two luminescent MOF compounds based on Eu and Tb have been used by Zhang and co-workers for the first time detection of diethyl chlorophosphate (DCP) pesticide in water. 136 Recently we have reported a solvothermally synthesized Cd based MOF for the detection of pesticides in water.137 However, in the above two cases the selectivity of pesticides detection has not been examined in real samples which is highly desirable for practical application. Additionally, during the synthesis of MOF, minimization of heat used and consequently the minimization of energy input are highly desirable from a techno-economic point of view.138 Besides that, the synthesis should be carried out at ambient temperature and pressure so as to avoid the high energy consumption and pressure associated risks during solvothermal synthesis.139,140 On the basis of the above points, a MOF [Cd3(PDA)1(tz)3Cl(H2O)4]·3H2O, 1, with a three-dimensional structure has been successfully synthesized using a slow layer diffusion technique, and its luminescence property is utilized for the pesticide detection in aqueous medium and real samples such as apple and tomato extract. This is the first observation, to our knowledge, where a MOF is utilized for selective detection of pesticide in aqueous medium along with the presence of micelle and fruit/vegetable extracts. The experimental results are well supported by theoretical calculations. Here, we present the synthesis, characterization, structure, and pesticide detection application of 1.

Article

EXPERIMENTAL SECTION

Materials. For the synthesis, CdCl2·H2O (98%, Merck), 1,4phenylenediacetic acid (97%, Alfa Aesar), 1,2,4-triazole (98%, SigmaAldrich), NaOH (97%, Merck), and methanol (Merck, 99.5%) were used as received. For the detection experiments, azinphos-methyl (AM) (97.8%, Sigma-Aldrich), parathion (PT) (99.7%, SigmaAldrich), chlorpyrifos (CP) (99.3%, Sigma-Aldrich), diazinon (DZ) (98.5%, Sigma-Aldrich), endosulfan (ES) (99.4%, Sigma-Aldrich), malathion (MT) (99.1%, Sigma-Aldrich), dichlorvos (DV) (99.8%, Sigma-Aldrich), hexadecyl trimethylammonium bromide (CTAB) (99%, Sigma-Aldrich), and sodium dodecyl sulfate (SDS) (98.5%, Sigma-Aldrich) were used as received. Double distilled water was used for both synthesis and detection experiments. Synthesis. Na2PDA was synthesized by a simple solvent evaporation reaction. For this purpose, 1,4-phenylenediacetic acid (1.005 g, 5 mmol) and NaOH (0.4 g, 10 mmol) were dissolved in 20 mL of water in a 100 mL beaker. The mixture was heated at 100 °C along with continuous starring until it solidified. Compound 1 was synthesized at room temperature via a layer diffusion method. For this, an aqueous solution (7.5 mL) of Na2PDA (0.375 mmol, 0.0893 g) was mixed with methanol solution (7.5 mL) of 1,2,4-triazole (0.375 mmol, 0.0261 g), and the mixture was stirred for 60 min to make a homogeneous solution. On the other hand, an aqueous solution of Cd2+ ions was prepared by dissolving CdCl2·H2O (1 mmol, 0.2054 g) in 20 mL of water in a narrow tube. Two milliliters of the ligand solution (mixture of Na2PDA and 1,2,4-triazole) was slowly and carefully layered over 2 mL of aqueous solution of Cd2+ ions in a capped test tube. After 7 weeks, colorless needle-shaped crystals suitable for single crystal diffraction were collected from the test tube and washed with a water−methanol mixture, and dried in air (yield 85% based on Cd). Elemental analysis of 1: calcd (%): C 21.46, H 3.13, N 14.08; found (%): C 21.35, H 3.05, N 14.74. Initial Characterizations. Powder X-ray diffraction (XRD) of 1 was carried out using Rigaku Miniflex diffractometer (see Supporting Information, Figure S1), and the patterns were compared with the simulated XRD pattern generated based on the structure determined using the single-crystal XRD to confirm the phase purity. IR spectrum was collected on a Shimadzu FTIR-8400S spectrophotometer in the solid state (see Supporting Information, Figure S2), and the results are summarized in Table S1 (see Supporting Information). Thermogravimetric analysis (TGA) were carried out in nitrogen atmosphere (PerkinElmer instrument STA 6000, flow rate = 20 mL min−1 and heating rate = 20 °C min−1) in the temperature range 40− 825 °C (see Supporting Information, Figure S3). From the TGA data, the weight loss of 6.1% (calculated 6.04%) up to 125 °C may be due to the removal of the water of crystallization. The weight loss above 350 °C is due to the decomposition of the framework. Single-Crystal Structure Determination. The details of instrumentation of single crystal structure determination are provided in Supporting Information (see p S3), and details of the structure solution and final refinement are given in the Table S2 (see Supporting Information). CCDC 1851539 contains the crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Center (CCDC) via www.ccdc.cam.ac.uk/data_request/cif. Photoluminescence Study. Room temperature photoluminescence properties of 1 dispersed in water were carried out using a PerkinElmer LS-55 spectrofluorometer. The aqueous dispersions of 1 were prepared by mixing 2 mg of 1 into 4 mL water, and it was ultrasonicated for 45 min. After this, 3 mL of this solution was diluted to 43 mL by water. Finally, 2 mL of this dispersion was used for the photoluminescence measurements using quartz cuvette. One millimolar pesticide solution was prepared in acetonitrile solvent, and the pesticide solution (analyte) was added into a quartz cuvette containing an aqueous dispersion of 1 using a micropipette. A Shimadzu UV 3101PC spectrophotometer was used to record the UV−vis spectra of aqueous solution of all the analytes. B

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



RESULTS AND DISCUSSION Structure. Single crystal structure determination shows that compound 1 has a monoclinic crystal system with P21/c space group. The asymmetric unit of 1 consists of three crystallographically independent Cd2+ ions, three 1,2,4-triazolates (tz), one 1,4-phenylenediacetate (PDA), one chloride ion, four coordinated water molecules, and three lattice water molecules (Figure 1). The octahedral Cd(1)2+ ion is coordinated by three

Figure 1. Ball and stick representation of asymmetric unit of [Cd3(PDA)1(tz)3Cl(H2O)4]·3H2O, 1.

oxygen atoms of carboxylate groups of PDA ligands, two triazolate nitrogen atoms, and a water molecule (see Supporting Information, Figure S4a). The Cd(2)2+ ion is also octahedrally coordinated by four triazolate nitrogen atoms and two water molecules (see Supporting Information, Figure S4b), whereas Cd(3)2+ ion is octahedrally coordinated by three triazolate nitrogen atoms, two chloride ions, and a water molecule (see Supporting Information, Figure S4c). All the three triazolates have similar connectivity where each triazolate connects three Cd2+ ions (see Supporting Information, Figure S5). Out of the two carboxylate groups of PDA ligand, one has bidentate connectivity with the Cd(1)2+ ion, and other one has monodentate connectivity with the Cd(2) 2+ ion (see Supporting Information, Figure S6). The Cd−O, Cd−N, and the Cd−Cl bonds have average distances of 2.423, 2.291, and 2.639 Å, respectively. The bond angles (O/N/Cl−Cd−O/N/Cl) vary in the range of 52.5(2)− 174.51(17)°. The selected bond distances and bond angles are listed in Table S3 and Table S4, respectively (see Supporting Information). In compound 1, the connectivity between all the three octahedral Cd2+ ions and all the three triazolate ligands formed a two-dimensional structure in the ab plane (Figure 2a). The two-dimensional structures are connected with each other through the edge sharing of two Cd(3)-octahedra via two Cl− ions. Figure 2b shows the three-dimensional arrangement through the connectivity between two 2D triazolate connected layers in the ac plane. The three-dimensional structure is further stabilized by the inter two-dimensional connectivity through the Cd(1)octahedra via PDA ligands (Figure 3). The lattice water molecules [O(100), O(200), and O(300)] occupy the channel created by the three-dimensional connectivity (shown in yellow in Figure 3). Luminescence Based Pesticides Detection Behavior. The luminescence spectra of compound 1 upon excitation at 225 nm in aqueous dispersion exhibits a strong emission at 290

Figure 2. (a) Figure shows the two-dimensional triazolate connected structure of [Cd3(PDA)1(tz)3Cl(H2O)4]·3H2O, 1, in the ab plane. (b) Three dimensional connectivity between two two-dimensional layers through Cl− ions shown in the ac plane. In both the figures, hydrogen atoms of tz are not shown for clarity.

nm (see Supporting Information, Figure S7), which was consequence of π* → π and π* → n transitions of aromatic dicarboxylate ligands of compound 1. To investigate the performance of 1 to detect traces of pesticides, we have performed luminescence quenching experiments with the accumulative addition of very dilute pesticides solution (in acetonitrile) to the water dispersion of compound 1. The experimental emission spectra in the presence of azinphosmethyl and chlorpyrifos (up to 30 μM) in compound 1 are shown in Figure 4. It can be observed that a significant quenching of luminescence occurred with the addition of azinphos-methyl and chlorpyrifos. In the case of azinphosmethyl at a concentration of 30 μM, luminescence intensity quenched almost 90%, whereas in the case of chlorpyrifos almost 52% quenching of luminescence intensity occurred. The quenching effect of azinphos-methyl and chlorpyrifos encouraged the investigation of the quenching behavior of compound 1 in the presence of a few other pesticides (see Supporting Information, Figure S8). Parathion, diazinon, C

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

when azinphos-methyl solution was added to the compound 1, the luminescence spectra of the compound exhibits a large amount of decrease in luminescence intensity with keeping the luminescence response of compound 1 toward azinphosmethyl unaffected. This process was repeated for another two cycles (see Supporting Information, Figure S15). To provide a better view, we have plotted [(I/I0) × 100] vs concentration of analyte added (see Supporting Information, Figure S16). The selectivity of the compound 1 toward the detection of azinphos-methyl in the presence of parathion and chlorpyrifos was confirmed by the stepwise quenching of luminescence intensity. The detection limit of azinphos-methyl was determined by performing luminescence quenching experiment using an extremely dilute solution of azinphos-methyl. It was found that compound 1 can detect azinphos-methyl in water at a concentration as low as 25 nanomolar (≡ 8 ppb) (see Supporting Information, Figure S17). The quenching efficiency of different pesticides can be quantified by fitting the plot of intensity ratio (I0/I) vs concentration ([A]) of pesticides with the Stern−Volmer (SV) equation: (I0/I) = Ksv[A] + 1, where I and I0 represent the luminescence intensity with and without the pesticide, and Ksv is the Stern−Volmer constant (M−1). The plot of azinphosmethyl deviates from linearity at higher concentration, whereas for all the other pesticides it was linear (see Figure 5). Deviation from linearity is the indication of involvement of multiple mechanism of quenching. Static quenching may occur if there is strong molecular level interaction through Hbonding or π−π stacking between compound 1 and the pesticides. The dynamic quenching may also be involved possibly because of the energy and electron transfer between the compound 1 and pesticides. Besides these processes, the inner filter effect is also responsible for the quenching of luminescence. For better understanding the sensitivity, we have calculated KSV values for chlorpyrifos, parathion, diazinon, malathion, endosulfan, and dichlorvos, which were found to be 3.16 × 104 M−1, 3.17 × 104 M−1, 0.47 × 104 M−1, 0.27 × 104 M−1, 0.18 × 104 M−1, and 0.13 × 104 M−1, respectively (see Figure 6a). But in the case of azinphos-methyl, the SV plot was found to be nonlinear, and in the higher concentration it is curved upward. So in this case, multiple mechanisms of quenching are operative. So to distinguish the effect of static

Figure 3. Three-dimensional overall structure of [Cd3(PDA)1(tz)3Cl(H2O)4]·3H2O, 1, shown in the bc plane. Note the presence of lattice water molecules within the channel. Hydrogen atoms of PDA and tz are not shown for clarity.

endosulfan, malathion, and dichlorvos are the other pesticides for which we have performed luminescence titration experiments in a similar fashion (see Supporting Information, Figures S9−S13). The luminescence intensity remained almost unaltered in the presence of these pesticides except in the case of parathion, where only a small quenching was observed (see Supporting Information, Figure S14). In the case of parathion, nearly 49% luminescence quenching is observed at a concentration of 30 μM. As azinphos-methyl (AM), chlorpyrifos (CP), and parathion (PT) showed ∼90%, 52%, and 49% of luminescence quenching, we have performed the competitive quenching experiment among this pool of pesticides. For this purpose, we have first measured the luminescence spectrum of the aqueous dispersion of compound 1. Then to this two-step addition of parathion was performed using 2.5 μM solution of parathion in each step followed by the addition of total 5 μM solution chlorpyrifos solution in two successive steps. Thereafter the two-step addition of azinphos-methyl was carried out by the utilization of a 2.5 μM solution of azinphosmethyl in each step following a similar procedure, and the respective luminescence spectrum was recorded. The luminescence intensity of compound 1 changes to a small extent in the steps in which parathion and chlorpyrifos were added. But

Figure 4. Emission spectra of 1 dispersed in water upon incremental addition of acetonitrile solution of (a) azinphos-methyl and (b) chlorpyrifos (λex = 225 nm). Final concentration of azinphos-methyl and chlorpyrifos in the medium is indicated in the legend. D

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 7. Emission spectra of 1 dispersed in CTAB upon incremental addition of acetonitrile solution of azinphos-methyl (λex = 225 nm). The final concentration of analytes solution is as below: (a) compound 1 in CTAB, (b) a + 5 μM AM, (c) a + 10 μM AM, (d) a + 15 μM AM, (e) a + 20 μM AM, (f) a + 25 μM AM, (g) a + 30 μM AM [where AM stands for azinphos-methyl].

Figure 5. Plot of [(I0/I) − 1)] of compound 1 vs concentration of pesticides ([A]). I0 and I are luminescence intensities in the absence and presence of pesticides, respectively.

Following a similar experimental protocol, we have also checked the sensing of azinphos-methyl in SDS micellar medium. In this case, 77% quenching of luminescence intensity is observed (see Supporting Information, Figure S18). So, the sensitivity of compound 1 for azinphos-methyl remains almost unaffected by surfactant molecules in the medium. Finally, it was tested in real samples for the potential application of the detector in real-life pesticide detection. As azinphos-methyl is widely applied in apple and tomato production,30,141−144 we have prepared the apple and tomato juice extract and examined the sensitivity of azinphos-methyl detection in these juice solutions. The detailed procedure for the apple and tomato juice extract preparation is provided in the Supporting Information. Pesticides Detection in Apple and Tomato Extracts. At first, the luminescence spectra of compound 1 was measured by taking 2 mL of aqueous dispersion of compound 1 with 100 μL of apple extract. Then luminescence titration experiments with the stepwise addition of azinphos-methyl were performed (see Figure 8). Here quenching of 71% of the

and dynamic quenching, we have used the modified form of the

Stern−Volmer

equation

(

I0 I

)

−1 ×

1 [A]

=

(KD + KS) + KDKS[A] to determine KD and KS, where KD and KS are the dynamic and static quenching constant respectively (see Figure 6b). The so obtained KD and KS values are 3.2 × 104 M−1 and 10.8 × 104 M−1 respectively. We have also checked the efficiency of pesticide detection in micellar medium as the surfactant molecules are often used to dissolve the pesticides. For this purpose we have checked the quenching efficiency of azinphos-methyl detection above the CMC of hexadecyl trimethylammonium bromide (CTAB) and sodium dodecyl sulfate (SDS) surfactant. Here we have taken compound 1 in the cuvette in the presence of CTAB, and the corresponding luminescence spectrum was measured. Then the luminescence quenching experiment was done by adding azinphos-methyl pesticide step-by-step, as was done earlier. In this case, quenching of 87% of the initial intensity was observed at a concentration of 30 μM azinphos-methyl (see Figure 7).

Figure 6. (a) Plot of I0/I of 1 (at 290 nm) vs concentration of pesticide [A] where I0 and I are luminescence intensity of compound 1 in the absence and presence of pesticide, respectively and (b) plot of

(

I0 I

)

−1 ×

1 [A]

vs [A], where I0 and I are luminescence intensity of compound 1 in

the absence and presence of azinphos-methyl pesticide, respectively. E

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry intensity was observed at a concentration of 30 μM of azinphos-methyl.

resonance energy transfer from 1 to pesticides. The possible light absorption, emission, and other photophysical processes are shown in Figure 9.

Figure 8. Emission spectra of the compound 1 dispersed in apple extract upon addition of acetonitrile solution of azinphos-methyl. The final concentration of analytes solution is as below: (a) 2 mL aqueous dispersion of compound 1 + 100 μL apple extract, (b) a + 2.5 μM AM, (c) a + 5 μM AM, (d) a + 10 μM AM, (e) a + 15 μM AM, (f) a + 20 μM AM, (g) a + 25 μM AM, (h) a + 30 μM AM [where AM stands for azinphos-methyl].

Figure 9. Schematic of various energy/electron transfer processes in the absence of pesticide (upper panel) and in the presence of pesticide (lower panel): 1. UV light excites ligand center; 2. Ligand based luminescence; 3. Energy transfer from ligand to analyte; 4. Electron transfer from ligand to analytes; 5. Absorption of UV light by the pesticide; 6. Pesticide based luminescence. Presence of pesticide absorbs a part of hυ as well as takes up energy (due to 3 and 4) from the ligand center excited state, resulting in large quenching of luminescence.

The quenching experiments for the selective detection of azinphos-methyl were also done in the presence of tomato extract following the above experimental procedure (see Supporting Information, Figure S19). In this case, almost 80% quenching of luminescence intensity is observed. However, it is relevant to test the stability of compound 1 in the presence of CTAB and pesticide in aqueous medium. For this purpose, we have checked the stability of compound 1 by soaking the ground crystals in water, aqueous CTAB solution and in the presence of acetonitrile solution (30 μM) of azinphos-methyl for 24 h in three separate experiments. Thereafter, the well filtered and dried powders were ground well, and the powder X-ray diffraction experiments were done on a Bruker D8 Discover instrument using Cu Kα radiation. The powder XRD pattern was found to be well matched with the simulated PXRD pattern of compound 1. This indicates that the material remains stable in aqueous solutions, in CTAB and in the presence of azinphos-methyl (see Supporting Information, Figure S20). To explain the experimental luminescence quenching efficiencies of the pesticides such as azinphos-methyl, we have to consider two important criteria: (i) the close proximity due to molecular level association between the azinphosmethyl and the compound 1, (ii) finding out the possible photophysical processes from the emission spectra of the compound and absorption spectra of the pesticides. The Hbonding and/or the π−π stacking interaction between 1 and azinphos-methyl may occur with the aromatic ligands (PDA and tz) of 1 and the aromatic part of the azinphos-methyl pesticide, which is evident from the infrared spectra of 1 in the presence of azinphos-methyl (see Supporting Information, Figure S21). These interactions probably cause static quenching of the luminescence intensity. In addition to the static quenching, three other processes occur: (i) inner filter effect, (ii) electron transfer from 1 to pesticides, and (iii)

Computational Studies. We have performed density functional theory (DFT) calculations to find out the HOMO− LUMO energy gap of the ligand moiety of compound 1 and the pesticide molecules, which may add further evidence to the proposed mechanism of quenching. All the structure optimization and calculations were done in the gas phase with the basis set 6-311G+(d,p) using Gaussian 16 (see Figure 10).145 The calculated energies are provided in Figure 10.

Figure 10. HOMO and LUMO energies for optimized H2PDA ligand and selected pesticides. Optical gap calculation using DFT [Basis set 6-311G+(d,p) using Gaussian 16]. F

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

luminescence quenching behavior. Moreover, it selectively recognized azinphos-methyl, even in micellar environments in which pesticides are generally mixed to make them watersoluble. Besides that compound 1 could recognize the pesticides even in real samples such as apple extract and tomato extract. The mechanisms of the luminescence quenching were found to be the combination of static quenching, electron transfer, resonance energy transfer, and excitation energy absorption. Experimental findings are also well supported by DFT calculations. So compound 1 could be an alternative pesticide detector in aqueous medium and also for real samples.

The lower the energy LUMO level of the pesticide than that of the ligand, the greater is the chance of electron transfer from ligand to pesticide. Following this, the electron transfer efficiency is in the order parathion > azinphos-methyl > chlorpyrifos > endosulfan > diazinon > malathion, whereas electron transfer is not possible in the case of dichlorvos. This computation studies pointed out that the electron transfer may be an important path for the luminescence quenching in the presence of pesticides, but it is not the only mechanism for the observed luminescence quenching. The UV−vis absorption spectra (Figure 11) of the pesticides in aqueous medium can give some fruitful reason behind the



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01767. PXRD pattern, TGA curve, IR spectra, additional figures, luminescence based titrations plots (PDF) Accession Codes

CCDC 1851539 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.



Figure 11. Absorption spectra of different pesticides (in same concentration) and luminescence spectra of compound 1. The rectangular shaped dashed line indicates the overlapping region of luminescence spectra of compound 1 and second absorption band of the pesticides indicating the possibility of resonance energy transfer.

AUTHOR INFORMATION

Corresponding Authors

*(P.M.) E-mail: [email protected]. *(S.K.M.) E-mail: [email protected]. ORCID

Prakash Majee: 0000-0002-9906-7867 Partha Mahata: 0000-0002-0341-890X

observed quenching of luminescence. Azinphos-methyl strongly absorbs a range of wavelength in the UV region centered at 226 nm. Chlorpyrifos also possesses a weak absorption band close to this region. However, the other five pesticides have a very small absorbance at 225 nm. Therefore, it is evident that azinphos-methyl and chlorpyrifos can absorb a portion of the excitation light by competing with the ligands of compound 1. From Figure 11, it is also observed that azinphos-methyl, parathion, and chlorpyrifos show a second absorption band, which is well overlapped with the luminescence spectra of 1 centered at 290 nm. This is a clear indication of the possibility of resonance energy transfer between the compound 1 and these three pesticides, which consequently reduces the luminescence intensity. The huge quenching in the case of azinphos-methyl is due to the better overlap, compared to the overlaps in the case of parathion and chlorpyrifos. So in addition to static quenching, a combined effect of resonance energy transfer, absorption of excitation light, and electron transfer was involved for the observed luminescence quenching.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by an INSPIRE Faculty Research Grant (IFA12-CH-69) [P.M.] of Department of Science and Technology (DST), Government of India, and Fast Track Project Grant (SB/FT/CS-138/2014) [S.K.M.] of Science and Engineering Research Board (SERB), Government of India. P.M. thanks Jadavpur University for a JU research grant. P. Majee thanks CSIR, New Delhi, for JRF fellowship.



REFERENCES

(1) Chen, J. Y.; Lin, Y. J.; Kuo, W. C. Pesticide residue removal from vegetables by ozonation. J. Food Eng. 2013, 114 (3), 404−411. (2) Fan, X.-D.; Zhang, W.-L.; Xiao, H.-Y.; Qiu, T.-Q.; Jiang, J.-G. Effects of ultrasound combined with ozone on the degradation of organophosphorus pesticide residues on lettuce. RSC Adv. 2015, 5 (57), 45622−45630. (3) Ormad, M. P.; Miguel, N.; Lanao, M.; Mosteo, R.; Ovelleiro, J. L. Effect of Application of Ozone and Ozone Combined with Hydrogen Peroxide and Titanium Dioxide in the Removal of Pesticides From Water. Ozone: Sci. Eng. 2010, 32 (1), 25−32. (4) Wang, L.; Liang, Y.; Jiang, X. Analysis of Eight Organophosphorus Pesticide Residues in Fresh Vegetables Retailed in Agricultural Product Markets of Nanjing, China. Bull. Environ. Contam. Toxicol. 2008, 81 (4), 377−382.



CONCLUSION We have successfully synthesized a new MOF at room temperature using a slow diffusion method. A three-dimensional structure containing lattice water molecules inside the channels was established by single crystal X-ray data. The luminescence property of compound 1 was utilized for the detection of pesticides in water medium studying the G

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (5) Liu, Q.; Jiang, X.; Zhang, Y.; Zheng, L.; Jing, W.; Liu, S.; Sui, G. A novel test strip for organophosphorus detection. Sens. Actuators, B 2015, 210, 803−810. (6) Lu, Y.; Sun, Q.; Hu, B.; Chen, X.; Miao, R.; Fang, Y. Synthesis and sensing applications of a new fluorescent derivative of cholesterol. New J. Chem. 2016, 40 (2), 1817−1824. (7) Damalas, C. A.; Eleftherohorinos, I. G. Pesticide Exposure, Safety Issues, and Risk Assessment Indicators. Int. J. Environ. Res. Public Health 2011, 8 (5), 1402−1419. (8) Vikrant, K.; Tsang, D. C. W.; Raza, N.; Giri, B. S.; Kukkar, D.; Kim, K.-H. Potential Utility of Metal−Organic Framework-Based Platform for Sensing Pesticides. ACS Appl. Mater. Interfaces 2018, 10 (10), 8797−8817. (9) Kim, K.-H.; Kabir, E.; Jahan, S. A. Exposure to pesticides and the associated human health effects. Sci. Total Environ. 2017, 575, 525− 535. (10) Zhu, X.; Zheng, H.; Wei, X.; Lin, Z.; Guo, L.; Qiu, B.; Chen, G. Metal-organic framework (MOF): a novel sensing platform for biomolecules. Chem. Commun. 2013, 49 (13), 1276−1278. (11) Agarwal, S.; Tyagi, I.; Gupta, V. K.; Dehghani, M. H.; Bagheri, A.; Yetilmezsoy, K.; Amrane, A.; Heibati, B.; Rodriguez-Couto, S. Degradation of azinphos-methyl and chlorpyrifos from aqueous solutions by ultrasound treatment. J. Mol. Liq. 2016, 221, 1237−1242. (12) Cacciatore, L. C.; Nemirovsky, S. I.; Verrengia Guerrero, N. R.; Cochón, A. C. Azinphos-methyl and chlorpyrifos, alone or in a binary mixture, produce oxidative stress and lipid peroxidation in the freshwater gastropod Planorbarius corneus. Aquat. Toxicol. 2015, 167, 12−19. (13) Funari, R.; Della Ventura, B.; Schiavo, L.; Esposito, R.; Altucci, C.; Velotta, R. Detection of Parathion Pesticide by Quartz Crystal Microbalance Functionalized with UV-Activated Antibodies. Anal. Chem. 2013, 85 (13), 6392−6397. (14) Funari, R.; Della Ventura, B.; Carrieri, R.; Morra, L.; Lahoz, E.; Gesuele, F.; Altucci, C.; Velotta, R. Detection of parathion and patulin by quartz-crystal microbalance functionalized by the photonics immobilization technique. Biosens. Bioelectron. 2015, 67, 224−229. (15) Hossain, M. M.; Kim, C. S.; Cha, H. J.; Lee, H. J. Amperometric Detection of Parathion and Methyl Parathion with a Microhole-ITIES. Electroanalysis 2011, 23 (9), 2049−2056. (16) Mishra, R. K.; Vinu Mohan, A. M.; Soto, F.; Chrostowski, R.; Wang, J. A microneedle biosensor for minimally-invasive transdermal detection of nerve agents. Analyst 2017, 142 (6), 918−924. (17) MacFarlane, E.; Carey, R.; Keegel, T.; El-Zaemay, S.; Fritschi, L. Dermal Exposure Associated with Occupational End Use of Pesticides and the Role of Protective Measures. SH W 2013, 4 (3), 136−141. (18) Zhang, W.; Ge, X.; Tang, Y.; Du, D.; Liu, D.; Lin, Y. Nanoparticle-based immunochromatographic test strip with fluorescent detector for quantification of phosphorylated acetylcholinesterase: an exposure biomarker of organophosphorus agents. Analyst 2013, 138 (18), 5431−5436. (19) Jang, Y. J.; Kim, K.; Tsay, O. G.; Atwood, D. A.; Churchill, D. G. Update 1 of: Destruction and Detection of Chemical Warfare Agents. Chem. Rev. 2015, 115 (24), PR1−PR76. (20) Kaur, R.; Kaur, N. A novel cation ensembled fluorescent organic nanoparticle for selective detection of organophosphorus insecticides. Dyes Pigm. 2017, 139, 310−317. (21) Nouanthavong, S.; Nacapricha, D.; Henry, C. S.; Sameenoi, Y. Pesticide analysis using nanoceria-coated paper-based devices as a detection platform. Analyst 2016, 141 (5), 1837−1846. (22) Ferri, D.; Barba-Bon, A.; Costero, A. M.; Gavina, P.; Parra, M.; Gil, S. An Au(iii)-amino alcohol complex for degradation of organophosphorus pesticides. RSC Adv. 2015, 5 (129), 106941− 106944. (23) Chong, H.; Ching, C. B. Development of Colorimetric-Based Whole-Cell Biosensor for Organophosphorus Compounds by Engineering Transcription Regulator DmpR. ACS Synth. Biol. 2016, 5 (11), 1290−1298.

(24) Cossi, P. F.; Beverly, B.; Carlos, L.; Kristoff, G. Recovery study of cholinesterases and neurotoxic signs in the non-target freshwater invertebrate Chilina gibbosa after an acute exposure to an environmental concentration of azinphos-methyl. Aquat. Toxicol. 2015, 167, 248−256. (25) Zhang, N.; Si, Y.; Sun, Z.; Li, S.; Li, S.; Lin, Y.; Wang, H. Labon-a-drop: biocompatible fluorescent nanoprobes of gold nanoclusters for label-free evaluation of phosphorylation-induced inhibition of acetylcholinesterase activity towards the ultrasensitive detection of pesticide residues. Analyst 2014, 139 (18), 4620−4628. (26) Marrs, T. C. Toxicology of Organophosphate Nerve Agents. In Chemical Warfare Agents; John Wiley & Sons, Ltd, 2007; pp 191−221. (27) Bajgar, J. Organophosphates/Nerve Agent Poisoning: Mechanism of Action, Diagnosis, Prophylaxis, And Treatment. In Advances in Clinical Chemistry; Elsevier, 2004; Vol. 38, pp 151−216. (28) Fahimi-Kashani, N.; Hormozi-Nezhad, M. R. Gold-Nanoparticle-Based Colorimetric Sensor Array for Discrimination of Organophosphate Pesticides. Anal. Chem. 2016, 88 (16), 8099−8106. (29) Tiwari, N.; Asthana, A.; Upadhyay, K. Kinetic-spectrophotometric determination of methyl parathion in water and vegetable samples. Spectrochim. Acta, Part A 2013, 101, 54−58. (30) Montemurro, M.; Siano, G. G.; Culzoni, M. J.; Goicoechea, H. C. Automatic generation of photochemically induced excitationemission-kinetic four-way data for the highly selective determination of azinphos-methyl in fruit juices. Sens. Actuators, B 2017, 239, 397− 404. (31) Portolés, T.; Cherta, L.; Beltran, J.; Hernández, F. Improved gas chromatography−tandem mass spectrometry determination of pesticide residues making use of atmospheric pressure chemical ionization. J. Chromatogr. A 2012, 1260, 183−192. (32) Walorczyk, S.; Drożdżyński, D. Improvement and extension to new analytes of a multi-residue method for the determination of pesticides in cereals and dry animal feed using gas chromatography− tandem quadrupole mass spectrometry revisited. J. Chromatogr. A 2012, 1251, 219−231. (33) Mahpishanian, S.; Sereshti, H.; Baghdadi, M. Superparamagnetic core−shells anchored onto graphene oxide grafted with phenylethyl amine as a nano-adsorbent for extraction and enrichment of organophosphorus pesticides from fruit, vegetable and water samples. J. Chromatogr. A 2015, 1406, 48−58. (34) Mahpishanian, S.; Sereshti, H. Three-dimensional graphene aerogel-supported iron oxide nanoparticles as an efficient adsorbent for magnetic solid phase extraction of organophosphorus pesticide residues in fruit juices followed by gas chromatographic determination. J. Chromatogr. A 2016, 1443, 43−53. (35) Soisungnoen, P.; Burakham, R.; Srijaranai, S. Determination of organophosphorus pesticides using dispersive liquid−liquid microextraction combined with reversed electrode polarity stacking modemicellar electrokinetic chromatography. Talanta 2012, 98, 62−68. (36) Fernando, J. C.; Rogers, K. R.; Anis, N. A.; Valdes, J. J.; Thompson, R. G.; Eldefrawi, A. T.; Eldefrawi, M. E. Rapid detection of anticholinesterase insecticides by a reusable light addressable potentiometric biosensor. J. Agric. Food Chem. 1993, 41 (3), 511− 516. (37) Ristori, C.; Del Carlo, C.; Martini, M.; Barbaro, A.; Ancarani, A. Potentiometric detection of pesticides in water samples. Anal. Chim. Acta 1996, 325 (3), 151−160. (38) Chang, P.-L.; Hsieh, M.-M.; Chiu, T.-C. Recent Advances in the Determination of Pesticides in Environmental Samples by Capillary Electrophoresis. Int. J. Environ. Res. Public Health 2016, 13 (4), 409. (39) Malik, A. K.; Faubel, W. A Review of Analysis of Pesticides Using Capillary Electrophoresis. Crit. Rev. Anal. Chem. 2001, 31 (3), 223−279. (40) El Rassi, Z. Capillary electrophoresis of pesticides. Electrophoresis 1997, 18 (12−13), 2465−2481. (41) Zheng, Y.-H.; Hua, T.-C.; Sun, D.-W.; Xiao, J.-J.; Xu, F.; Wang, F.-F. Detection of dichlorvos residue by flow injection calorimetric H

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry biosensor based on immobilized chicken liver esterase. J. Food Eng. 2006, 74 (1), 24−29. (42) Raghu, P.; Kumara Swamy, B. E.; Madhusudana Reddy, T.; Chandrashekar, B. N.; Reddaiah, K. Sol−gel immobilized biosensor for the detection of organophosphorous pesticides: A voltammetric method. Bioelectrochemistry 2012, 83, 19−24. (43) Wang, M.; Li, Z. Nano-composite ZrO2/Au film electrode for voltammetric detection of parathion. Sens. Actuators, B 2008, 133 (2), 607−612. (44) Lakowicz, J. R. Principles of Fluorescence Spectroscopy, 2006. (45) Snow, E. S.; Perkins, F. K.; Houser, E. J.; Badescu, S. C.; Reinecke, T. L. Chemical Detection with a Single-Walled Carbon Nanotube Capacitor. Science 2005, 307 (5717), 1942−1945. (46) Singha, D. K.; Mahata, P. Luminescent coordination polymerfullerene composite as a highly sensitive and selective optical detector for 2,4,6-trinitrophenol (TNP). RSC Adv. 2015, 5 (36), 28092− 28097. (47) Singha, D. K.; Majee, P.; Mondal, S. K.; Mahata, P.; Eu-Doped Y-Based, A. Luminescent Metal-Organic Framework as a Highly Efficient Sensor for Nitroaromatic Explosives. Eur. J. Inorg. Chem. 2015, 2015, 1390−1397. (48) Shustova, N. B.; Cozzolino, A. F.; Reineke, S.; Baldo, M.; Dincă, M. Selective Turn-On Ammonia Sensing Enabled by HighTemperature Fluorescence in Metal−Organic Frameworks with Open Metal Sites. J. Am. Chem. Soc. 2013, 135 (36), 13326−13329. (49) 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 (10), 2881−2885. (50) Zeng, X.; Ma, J.; Luo, L.; Yang, L.; Cao, X.; Tian, D.; Li, H. Pesticide Macroscopic Recognition by a Naphthol-Appended Calix[4]arene. Org. Lett. 2015, 17 (12), 2976−2979. (51) Yao, J.; Fu, Y.; Xu, W.; Fan, T.; Gao, Y.; He, Q.; Zhu, D.; Cao, H.; Cheng, J. Concise and Efficient Fluorescent Probe via an Intromolecular Charge Transfer for the Chemical Warfare Agent Mimic Diethylchlorophosphate Vapor Detection. Anal. Chem. 2016, 88 (4), 2497−2501. (52) Southard, G. E.; Van Houten, K. A.; Murray, G. M. Soluble and Processable Phosphonate Sensing Star Molecularly Imprinted Polymers. Macromolecules 2007, 40 (5), 1395−1400. (53) Raj, P.; Singh, A.; Kaur, K.; Aree, T.; Singh, A.; Singh, N. Fluorescent Chemosensors for Selective and Sensitive Detection of Phosmet/Chlorpyrifos with Octahedral Ni2+ Complexes. Inorg. Chem. 2016, 55 (10), 4874−4883. (54) Barba-Bon, A.; Costero, A. M.; Gil, S.; Sancenon, F.; MartinezManez, R. Chromo-fluorogenic BODIPY-complexes for selective detection of V-type nerve agent surrogates. Chem. Commun. 2014, 50 (87), 13289−13291. (55) Knapton, D.; Burnworth, M.; Rowan, S. J.; Weder, C. Fluorescent Organometallic Sensors for the Detection of ChemicalWarfare-Agent Mimics. Angew. Chem., Int. Ed. 2006, 45 (35), 5825− 5829. (56) Dennison, G. H.; Johnston, M. R. Mechanistic Insights into the Luminescent Sensing of Organophosphorus Chemical Warfare Agents and Simulants Using Trivalent Lanthanide Complexes. Chem. - Eur. J. 2015, 21 (17), 6328−6338. (57) Kanagaraj, K.; Affrose, A.; Sivakolunthu, S.; Pitchumani, K. Highly selective fluorescent sensing of fenitrothion using per-6-aminoβ-cyclodextrin:Eu(III) complex. Biosens. Bioelectron. 2012, 35 (1), 452−455. (58) Zhang, H.; Hua, X.; Tuo, X.; Chen, C.; Wang, X. Polystyrene microsphere-based lanthanide luminescent chemosensor for detection of organophosphate pesticides. J. Rare Earths 2012, 30 (12), 1203− 1207. (59) Schwierking, J. R.; Menzel, L. W.; Menzel, E. R. Organophosphate Nerve Agent Detection with Europium Complexes. Sci. World J. 2004, 4, 948−955. (60) Hussein, B. H. M.; Khairy, G. M.; Kamel, R. M. Fluorescence sensing of phosdrin pesticide by the luminescent Eu(III)- and

Tb(III)-bis(coumarin-3-carboxylic acid) probes. Spectrochim. Acta, Part A 2016, 158, 34−42. (61) Azab, H. A.; Khairy, G. M.; Kamel, R. M. Time-resolved fluorescence sensing of pesticides chlorpyrifos, crotoxyphos and endosulfan by the luminescent Eu(III)−8-allyl-3-carboxycoumarin probe. Spectrochim. Acta, Part A 2015, 148, 114−124. (62) Azab, H. A.; Duerkop, A.; Mogahed, E. M.; Awad, F. K.; Abd El Aal, R. M.; Kamel, R. M. Fluorescence and Electrochemical Sensing of Pesticides Methomyl, Aldicarb and Prometryne by the Luminescent Europium-3-Carboxycoumarin Probe. J. Fluoresc. 2012, 22 (2), 659− 676. (63) Shunmugam, R.; Tew, G. N. Terpyridine−Lanthanide Complexes Respond to Fluorophosphate Containing Nerve Gas GAgent Surrogates. Chem. - Eur. J. 2008, 14 (18), 5409−5412. (64) Zor, E.; Morales-Narváez, E.; Zamora-Gálvez, A.; Bingol, H.; Ersoz, M.; Merkoçi, A. Graphene Quantum Dots-based Photoluminescent Sensor: A Multifunctional Composite for Pesticide Detection. ACS Appl. Mater. Interfaces 2015, 7 (36), 20272−20279. (65) Zhao, Y.; Ma, Y.; Li, H.; Wang, L. Composite QDs@MIP Nanospheres for Specific Recognition and Direct Fluorescent Quantification of Pesticides in Aqueous Media. Anal. Chem. 2012, 84 (1), 386−395. (66) Zhang, K.; Yu, T.; Liu, F.; Sun, M.; Yu, H.; Liu, B.; Zhang, Z.; Jiang, H.; Wang, S. Selective Fluorescence Turn-On and Ratiometric Detection of Organophosphate Using Dual-Emitting Mn-Doped ZnS Nanocrystal Probe. Anal. Chem. 2014, 86 (23), 11727−11733. (67) Aragay, G.; Pino, F.; Merkoçi, A. Nanomaterials for sensing and destroying pesticides. Chem. Rev. 2012, 112, 5317. (68) Liu, D.; Chen, W.; Wei, J.; Li, X.; Wang, Z.; Jiang, X. A Highly Sensitive, Dual-Readout Assay Based on Gold Nanoparticles for Organophosphorus and Carbamate Pesticides. Anal. Chem. 2012, 84 (9), 4185−4191. (69) Shen, Y.-W.; Hsu, P.-H.; Unnikrishnan, B.; Li, Y.-J.; Huang, C.C. Membrane-Based Assay for Iodide Ions Based on Anti-Leaching of Gold Nanoparticles. ACS Appl. Mater. Interfaces 2014, 6 (4), 2576− 2582. (70) Gole, B.; Song, W.; Lackinger, M.; Mukherjee, P. S. Explosives Sensing by Using Electron-Rich Supramolecular Polymers: Role of Intermolecular Hydrogen Bonding in Significant Enhancement of Sensitivity. Chem. - Eur. J. 2014, 20 (42), 13662−13680. (71) Singha, D. K.; Majee, P.; Mondal, S. K.; Mahata, P. Visible detection of explosive nitroaromatics facilitated by a large stokes shift of luminescence using europium and terbium doped yttrium based MOFs. RSC Adv. 2015, 5 (123), 102076−102084. (72) Singha, D. K.; Majee, P.; Mondal, S. K.; Mahata, P. Selective Luminescence-Based Detection of Cd2+ and Zn2+ Ions in Water Using a Proton-Transferred Coordination Polymer-Amine Conjugate Pair. ChemistrySelect 2017, 2 (11), 3388−3395. (73) Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O. M. Topological Analysis of Metal−Organic Frameworks with Polytopic Linkers and/or Multiple Building Units and the Minimal Transitivity Principle. Chem. Rev. 2014, 114 (2), 1343−1370. (74) O’Keeffe, M.; Yaghi, O. M. Deconstructing the Crystal Structures of Metal−Organic Frameworks and Related Materials into Their Underlying Nets. Chem. Rev. 2012, 112, 675−702. (75) Natarajan, S.; Mahata, P. Metal-organic framework structures how closely are they related to classical inorganic structures? Chem. Soc. Rev. 2009, 38 (8), 2304−2318. (76) Cook, T. R.; Zheng, Y.-R.; Stang, P. J. Metal−Organic Frameworks and Self-Assembled Supramolecular Coordination Complexes: Comparing and Contrasting the Design, Synthesis, and Functionality of Metal−Organic Materials. Chem. Rev. 2013, 113 (1), 734−777. (77) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal−Organic Frameworks. Chem. Rev. 2012, 112 (2), 724−781. (78) Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. High-Throughput Synthesis of Zeolitic I

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Imidazolate Frameworks and Application to CO2 Capture. Science 2008, 319 (5865), 939−943. (79) Yang, J.; Trickett, C. A.; Alahmadi, S. B.; Alshammari, A. S.; Yaghi, O. M. Calcium l-Lactate Frameworks as Naturally Degradable Carriers for Pesticides. J. Am. Chem. Soc. 2017, 139 (24), 8118−8121. (80) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1477−1504. (81) Nugent, P.; Belmabkhout, Y.; Burd, S. D.; Cairns, A. J.; Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. Porous materials with optimal adsorption thermodynamics and kinetics for CO2 separation. Nature 2013, 495 (7439), 80−84. (82) Kurmoo, M. Magnetic metal-organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1353−1379. (83) Mahata, P.; Natarajan, S.; Panissod, P.; Drillon, M. Quasi-2D XY Magnetic Properties and Slow Relaxation in a Body Centered Metal Organic Network of [Co4] Clusters. J. Am. Chem. Soc. 2009, 131 (29), 10140−10150. (84) He, C.; Liu, D.; Lin, W. Nanomedicine Applications of Hybrid Nanomaterials Built from Metal−Ligand Coordination Bonds: Nanoscale Metal−Organic Frameworks and Nanoscale Coordination Polymers. Chem. Rev. 2015, 115 (19), 11079−11108. (85) Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Férey, G.; Morris, R. E.; Serre, C. Metal−Organic Frameworks in Biomedicine. Chem. Rev. 2012, 112 (2), 1232−1268. (86) Chughtai, A. H.; Ahmad, N.; Younus, H. A.; Laypkov, A.; Verpoort, F. Metal-organic frameworks: versatile heterogeneous catalysts for efficient catalytic organic transformations. Chem. Soc. Rev. 2015, 44 (19), 6804−6849. (87) Huang, Y.-B.; Liang, J.; Wang, X.-S.; Cao, R. Multifunctional metal-organic framework catalysts: synergistic catalysis and tandem reactions. Chem. Soc. Rev. 2017, 46 (1), 126−157. (88) Ji, P.; Manna, K.; Lin, Z.; Urban, A.; Greene, F. X.; Lan, G.; Lin, W. Single-Site Cobalt Catalysts at New Zr8(μ2-O)8(μ2-OH)4 MetalOrganic Framework Nodes for Highly Active Hydrogenation of Alkenes, Imines, Carbonyls, and Heterocycles. J. Am. Chem. Soc. 2016, 138 (37), 12234−12242. (89) Deenadayalan, M. S.; Sharma, N.; Verma, P. K.; Nagaraja, C. M. Visible-Light-Assisted Photocatalytic Reduction of Nitroaromatics by Recyclable Ni(II)-Porphyrin Metal−Organic Framework (MOF) at RT. Inorg. Chem. 2016, 55 (11), 5320−5327. (90) Wu, Z.-L.; Wang, C.-H.; Zhao, B.; Dong, J.; Lu, F.; Wang, W.H.; Wang, W.-C.; Wu, G.-J.; Cui, J.-Z.; Cheng, P. A Semi-Conductive Copper−Organic Framework with Two Types of Photocatalytic Activity. Angew. Chem., Int. Ed. 2016, 55 (16), 4938−4942. (91) Hasegawa, Y.; Nakanishi, T. Luminescent lanthanide coordination polymers for photonic applications. RSC Adv. 2015, 5 (1), 338−353. (92) Wang, M.-S.; Guo, S.-P.; Li, Y.; Cai, L.-Z.; Zou, J.-P.; Xu, G.; Zhou, W.-W.; Zheng, F.-K.; Guo, G.-C. A Direct White-LightEmitting Metal−Organic Framework with Tunable Yellow-to-White Photoluminescence by Variation of Excitation Light. J. Am. Chem. Soc. 2009, 131 (38), 13572−13573. (93) Gong, Q.; Hu, Z.; Deibert, B. J.; Emge, T. J.; Teat, S. J.; Banerjee, D.; Mussman, B.; Rudd, N. D.; Li, J. Solution Processable MOF Yellow Phosphor with Exceptionally High Quantum Efficiency. J. Am. Chem. Soc. 2014, 136 (48), 16724−16727. (94) Duan, T.-W.; Yan, B. Hybrids based on lanthanide ions activated yttrium metal-organic frameworks: functional assembly, polymer film preparation and luminescence tuning. J. Mater. Chem. C 2014, 2 (26), 5098−5104. (95) Ramaswamy, P.; Wong, N. E.; Shimizu, G. K. H. MOFs as proton conductors - challenges and opportunities. Chem. Soc. Rev. 2014, 43 (16), 5913−5932. (96) Yoon, M.; Suh, K.; Natarajan, S.; Kim, K. Proton Conduction in Metal−Organic Frameworks and Related Modularly Built Porous Solids. Angew. Chem., Int. Ed. 2013, 52 (10), 2688−2700.

(97) Sadakiyo, M.; Yamada, T.; Kitagawa, H. Proton Conductivity Control by Ion Substitution in a Highly Proton-Conductive Metal− Organic Framework. J. Am. Chem. Soc. 2014, 136 (38), 13166−13169. (98) Ramaswamy, P.; Wong, N. E.; Gelfand, B. S.; Shimizu, G. K. H. A Water Stable Magnesium MOF That Conducts Protons over 10−2 S cm−1. J. Am. Chem. Soc. 2015, 137 (24), 7640−7643. (99) Horike, S.; Umeyama, D.; Kitagawa, S. Ion Conductivity and Transport by Porous Coordination Polymers and Metal−Organic Frameworks. Acc. Chem. Res. 2013, 46 (11), 2376−2384. (100) Huang, R.-W.; Wei, Y.-S.; Dong, X.-Y.; Wu, X.-H.; Du, C.-X.; Zang, S.-Q.; Mak, T. C. W. Hypersensitive dual-function luminescence switching of a silver-chalcogenolate cluster-based metal− organic framework. Nat. Chem. 2017, 9, 689. (101) Mahata, P.; Mondal, S. K.; Singha, D. K.; Majee, P. Luminescent rare-earth-based MOFs as optical sensors. Dalton Trans. 2017, 46 (2), 301−328. (102) Cheng, J.; Zhou, X.; Xiang, H. Fluorescent metal ion chemosensors via cation exchange reactions of complexes, quantum dots, and metal-organic frameworks. Analyst 2015, 140 (21), 7082− 7115. (103) Cui, Y.; Zhu, F.; Chen, B.; Qian, G. Metal-organic frameworks for luminescence thermometry. Chem. Commun. 2015, 51 (35), 7420−7431. (104) 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 (11), 3242−3285. (105) Hu, Z.; Deibert, B. J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43 (16), 5815−5840. (106) 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 Compouds. J. Am. Chem. Soc. 2011, 133, 4153−4155. (107) Wanderley, M. M.; Wang, C.; Wu, C.-D.; Lin, W. A Chiral Porous Metal−Organic Framework for Highly Sensitive and Enantioselective Fluorescence Sensing of Amino Alcohols. J. Am. Chem. Soc. 2012, 134 (22), 9050−9053. (108) Singha, D. K.; Mahata, P. Highly Selective and Sensitive Luminescence Turn-On-Based Sensing of Al3+ Ions in Aqueous Medium Using a MOF with Free Functional Sites. Inorg. Chem. 2015, 54 (13), 6373−6379. (109) Song, J. H.; Kim, Y.; Lim, K. S.; Kang, D. W.; Lee, W. R.; Hong, C. S. Phase Transformation, Exceptional Quenching Efficiency, and Discriminative Recognition of Nitroaromatic Analytes in Hydrophobic, Nonporous Zn(II) Coordination Frameworks. Inorg. Chem. 2017, 56 (1), 305−312. (110) Bhattacharyya, S.; Chakraborty, A.; Jayaramulu, K.; Hazra, A.; Maji, T. K. A bimodal anionic MOF: turn-off sensing of CuII and specific sensitization of EuIII. Chem. Commun. 2014, 50 (88), 13567− 13570. (111) 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 (48), 20908−20915. (112) Cui, Y. J.; Yue, Y. F.; Qian, G. D.; Chen, B. L. Luminescent Functional Metal-Organic Frameworks. Chem. Rev. 2012, 112, 1126. (113) Zhao, D.; Cui, Y.; Yang, Y.; Qian, G. Sensing-functional luminescent metal-organic frameworks. CrystEngComm 2016, 18 (21), 3746−3759. (114) Harbuzaru, B. V.; Corma, A.; Rey, F.; Jorda, J. L.; Ananias, D.; Carlos, L. D.; Rocha, J. A Miniaturized Linear pH Sensor Based on a Highly Photoluminescent Self-Assembled Europium(III) MetalOrganic Framework. Angew. Chem., Int. Ed. 2009, 48, 6476. (115) Li, G.-P.; Liu, G.; Li, Y.-Z.; Hou, L.; Wang, Y.-Y.; Zhu, Z. Uncommon Pyrazoyl-Carboxyl Bifunctional Ligand-Based Microporous Lanthanide Systems: Sorption and Luminescent Sensing Properties. Inorg. Chem. 2016, 55 (8), 3952−3959. J

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (116) Jiang, H. L.; Feng, D.; Wang, K.; Gu, Z. Y.; Wei, Z.; Chen, Y. P.; Zhou, H. C. An exceptionally stable, porphyrinic Zr metal-organic framework exhibiting pH-dependent fluorescence. J. Am. Chem. Soc. 2013, 135, 13934. (117) Asha, K. S.; Bhattacharjee, R.; Mandal, S. Complete Transmetalation in a Metal−Organic Framework by Metal Ion Metathesis in a Single Crystal for Selective Sensing of Phosphate Ions in Aqueous Media. Angew. Chem., Int. Ed. 2016, 55 (38), 11528− 11532. (118) Cui, Y. J.; Xu, H.; Yue, Y. F.; Guo, Z. Y.; Yu, J. C.; Chen, Z. X.; Gao, J. K.; Yang, Y.; Qian, G. D.; Chen, B. L. A Luminescent MixedLanthanide Metal-Organic Framework Thermometer. J. Am. Chem. Soc. 2012, 134, 3979. (119) Rao, X.; Song, T.; Gao, J.; Cui, Y.; Yang, Y.; Wu, C.; Chen, B.; Qian, G. A Highly Sensitive Mixed Lanthanide Metal−Organic Framework Self-Calibrated Luminescent Thermometer. J. Am. Chem. Soc. 2013, 135 (41), 15559−15564. (120) Zhao, D.; Rao, X.; Yu, J.; Cui, Y.; Yang, Y.; Qian, G. Design and Synthesis of an MOF Thermometer with High Sensitivity in the Physiological Temperature Range. Inorg. Chem. 2015, 54 (23), 11193−11199. (121) Miyata, K.; Konno, Y.; Nakanishi, T.; Kobayashi, A.; Kato, M.; Fushimi, K.; Hasegawa, Y. Chameleon Luminophore for Sensing Temperatures: Control of Metal-to-Metal and Energy Back Transfer in Lanthanide Coordination Polymers. Angew. Chem., Int. Ed. 2013, 52 (25), 6413−6416. (122) Xiao, Y.; Cui, Y.; Zheng, Q.; Xiang, S.; Qian, G.; Chen, B. A microporous luminescent metal-organic framework for highly selective and sensitive sensing of Cu2+ in aqueous solution. Chem. Commun. 2010, 46 (30), 5503−5505. (123) Zhu, Y.-M.; Zeng, C.-H.; Chu, T.-S.; Wang, H.-M.; Yang, Y.Y.; Tong, Y.-X.; Su, C.-Y.; Wong, W.-T. A novel highly luminescent LnMOF film: a convenient sensor for Hg2+ detecting. J. Mater. Chem. A 2013, 1 (37), 11312−11319. (124) 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 (65), 9153− 9156. (125) Zhao, Y.; Xu, X.; Qiu, L.; Kang, X.; Wen, L.; Zhang, B. Metal− Organic Frameworks Constructed from a New Thiophene-Functionalized Dicarboxylate: Luminescence Sensing and Pesticide Removal. ACS Appl. Mater. Interfaces 2017, 9 (17), 15164−15175. (126) Wang, H.; Lustig, W. P.; Li, J. Sensing and capture of toxic and hazardous gases and vapors by metal??organic frameworks. Chem. Soc. Rev. 2018, 47 (13), 4729−4756. (127) Li, H.-Y.; Wei, Y.-L.; Dong, X.-Y.; Zang, S.-Q.; Mak, T. C. W. Novel Tb-MOF Embedded with Viologen Species for MultiPhotofunctionality: Photochromism, Photomodulated Fluorescence, and Luminescent pH Sensing. Chem. Mater. 2015, 27 (4), 1327− 1331. (128) Dong, X.-Y.; Wang, R.; Wang, J.-Z.; Zang, S.-Q.; Mak, T. C. W. Highly selective Fe3+ sensing and proton conduction in a waterstable sulfonate-carboxylate Tb-organic-framework. J. Mater. Chem. A 2015, 3 (2), 641−647. (129) 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 (47), 15964−15971. (130) Rocha, J.; Carlos, L. D.; Paz, F. A. A.; Ananias, D. Luminescent multifunctional lanthanides-based metal-organic frameworks. Chem. Soc. Rev. 2011, 40 (2), 926−940. (131) Kumar, P.; Paul, A. K.; Deep, A. Sensitive chemosensing of nitro group containing organophosphate pesticides with MOF-5. Microporous Mesoporous Mater. 2014, 195, 60−66. (132) Kumar, P.; Paul, A. K.; Deep, A. A luminescent nanocrystal metal organic framework for chemosensing of nitro group containing organophosphate pesticides. Anal. Methods 2014, 6 (12), 4095−4101.

(133) Zheng, X.; Zhou, L.; Huang, Y.; Wang, C.; Duan, J.; Wen, L.; Tian, Z.; Li, D. A series of metal-organic frameworks based on 5-(4pyridyl)-isophthalic acid: selective sorption and fluorescence sensing. J. Mater. Chem. A 2014, 2 (31), 12413−12422. (134) Wen, L.; Xu, X.; Lv, K.; Huang, Y.; Zheng, X.; Zhou, L.; Sun, R.; Li, D. Metal−Organic Frameworks Constructed from d-Camphor Acid: Bifunctional Properties Related to Luminescence Sensing and Liquid-Phase Separation. ACS Appl. Mater. Interfaces 2015, 7 (7), 4449−4455. (135) Tao, C.-L.; Chen, B.; Liu, X.-G.; Zhou, L.-J.; Zhu, X.-L.; Cao, J.; Gu, Z.-G.; Zhao, Z.; Shen, L.; Tang, B. Z. A highly luminescent entangled metal-organic framework based on pyridine-substituted tetraphenylethene for efficient pesticide detection. Chem. Commun. 2017, 53 (72), 9975−9978. (136) Zheng, S.-R.; Chen, R.-L.; Liu, Z.-M.; Wen, X.-L.; Xie, T.; Fan, J.; Zhang, W.-G. Construction of terpyridine-Ln(iii) coordination polymers: structural diversity, visible and NIR luminescence properties and response to nerve-agent mimics. CrystEngComm 2014, 16 (14), 2898−2909. (137) Singha, D. K.; Majee, P.; Mondal, S. K.; Mahata, P. Highly Selective Aqueous Phase Detection of Azinphos-Methyl Pesticide in ppb Level Using a Cage-Connected 3D MOF. ChemistrySelect 2017, 2 (20), 5760−5768. (138) Noh, H.; Kung, C.-W.; Islamoglu, T.; Peters, A. W.; Liao, Y.; Li, P.; Garibay, S. J.; Zhang, X.; DeStefano, M. R.; Hupp, J. T.; Farha, O. K. Room Temperature Synthesis of an 8-Connected Zr-Based Metal−Organic Framework for Top-Down Nanoparticle Encapsulation. Chem. Mater. 2018, 30 (7), 2193−2197. (139) Reinsch, H. Green” Synthesis of Metal-Organic Frameworks. Eur. J. Inorg. Chem. 2016, 2016 (27), 4290−4299. (140) Ramos-Fernandez, E. V.; Grau-Atienza, A.; Farrusseng, D.; Aguado, S. A water-based room temperature synthesis of ZIF-93 for CO2 adsorption. J. Mater. Chem. A 2018, 6 (14), 5598−5602. (141) Schulz, R.; Hahn, C.; Bennett, E. R.; Dabrowski, J. M.; Thiere, G.; Peall, S. K. C. Fate and Effects of Azinphos-Methyl in a FlowThrough Wetland in South Africa. Environ. Sci. Technol. 2003, 37 (10), 2139−2144. (142) Mercade, J. V.; Montoya, A. A monoclonal antibody-based ELISA for the analysis of azinphos-methyl in fruit juices. Anal. Chim. Acta 1997, 347 (1), 95−101. (143) Neidert, E.; Saschenbrecker, P. W. Occurrence of pesticide residues in selected agricultural food commodities available in Canada. J. AOAC Int. 1996, 79 (2), 549−566. (144) Athanasopoulos, P. E.; Pappas, C. Effects of fruit acidity and storage conditions on the rate of degradation of azinphos methyl on apples and lemons. Food Chem. 2000, 69 (1), 69−72. (145) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 16; Gaussian, Inc.: Wallingford, CT, 2016.

K

DOI: 10.1021/acs.inorgchem.8b01767 Inorg. Chem. XXXX, XXX, XXX−XXX