Next Article in Journal
Morphological and Crystallographic Characterization of Primary Zinc-Rich Crystals in a Ternary Sn-Zn-Bi Alloy under a High Magnetic Field
Previous Article in Journal
Incompatibility Stresses and Lattice Rotations Due to Grain Boundary Sliding in Heterogeneous Anisotropic Elasticity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Lanthanide Coordination Polymers as Luminescent Sensors for the Selective and Recyclable Detection of Acetone

1
Key Laboratory of Comprehensive Utilization of Mineral Resource in Ethnic Regions, Key Laboratory of Resource Clean Conversion in Ethnic Regions, Education Department of Yunnan, School of Chemistry & Environment, Yunnan Minzu University, Kunming 650500, China
2
Key Laboratory of Medicinal Chemistry for Natural Resource Education Ministry, School of Chemical Science and Technology, Yunnan University, Kunming 650091, China
*
Author to whom correspondence should be addressed.
Crystals 2017, 7(7), 199; https://doi.org/10.3390/cryst7070199
Submission received: 23 May 2017 / Revised: 22 June 2017 / Accepted: 28 June 2017 / Published: 5 July 2017

Abstract

:
Three new isostructural lanthanide coordination polymers {[Ln(L)2·2H2O]·Cl·4H2O}, Ln = La (LaL 1), Tb (TbL 2), Eu (EuL 3), L = 4-carboxy-1-(4-carboxybenzyl)pyridinium, have been synthesized under hydrothermal conditions and characterized by single crystal X-ray diffraction, IR, TG, PXRD, and luminescence. The solid-state luminescence properties of those Ln-CPs were investigated, realizing the zwitterionic ligand (L) is an excellent antenna chromophore for sensitizing both Tb3+ and Eu3+ ions. We utilized TbL 2 as a representative chemosensor to consider the potential luminescence sensing properties in different solvent suspension, which has the potential to serve as the first case of a luminescent Ln-CP material based on the zwitterionic type of organic ligand for selective and recyclable sensing of acetone in methanol solution.

1. Introduction

Over the last decade, lanthanide-based metal–organic coordination polymers (Ln-CPs), as novel luminescent functional materials, have attracted a good deal of attention. They have attracted this attention not only because of their diverse architectures, but also because of their potential luminescent sensor applications for detecting cations [1,2], anions [3,4], small molecules [5], gases and vapors [6], biomolecules [7], temperature [8,9], and so on [10,11]. Compared with other luminescent sensors, Ln-CPs based luminescent sensing materials have been a rapidly developing area due to their outstanding merits such as high surface area, easily designed crystal structures, stable frameworks, and permanent porosity, as well as exposed active sites [12,13,14]. Recently, the zwitterionic type of organic ligands have drawn our attention [15,16], which simultaneously bear positive and negative charges with a certain separated distance on the coordination skeleton endowing special physical properties, for example, improved adsorption selectivity of gases or vapors [17]. However, to date, the Ln-CPs materials reported are usually based on the common organic ligands, while the study of Ln-CPs based on the zwitterionic types of organic ligands have rarely been reported [18]. Therefore, further systematic investigation on the Ln-CPs based on the zwitterionic types of organic ligands is absolutely necessary.
Acetone is a highly volatile organic solvent, which could stimulate cornea or metabolism disorder and have a toxic effect on the human body [19]. Considering the extensive application and potential harm of acetone, it is urgent to develop new luminescent sensors to detect it. To date, only a few CPs have been reported for the sensing of acetone [20,21,22,23], most of which were constructed by d10 transition metal ions (Zn and Cd) and π-conjugated ligands, whereas coordination polymer materials based on lanthanide ions have been rarely published [24]. The advantages of those lanthanide materials over the transition metal luminescent materials include the sharp and strong emissions, large stokes shifts, high quantum yields, and long luminescence lifetimes, which originate from f–f transitions of lanthanide ions [25]. Taking into account these kinds of sensor materials are still defective, the synthesis and development of new materials based on lanthanide CPs for fluorescent high selective detection of acetones is still a major challenge.
As an effective combination of the above mentioned aspects, we selected the zwitterionic types of organic ligand 4-carboxy-1-(4-carboxybenzyl)pyridinium chloride (H2LCl) as ligand to construct three isostructrual lanthanide coordination polymers {[Ln(L)2·2H2O]·Cl·4H2O}, Ln = La (LaL 1), Tb (TbL 2), Eu (EuL 3). Those materials have the potential to serve as the first example of a luminescent Ln-CP material based on the zwitterionic type of organic ligand for selective and recyclable sensing of acetone in methanol solution.

2. Results

2.1. Crystal Structures

2.1.1. Descriptions of the Crystal Structures of {[La(L)2·2H2O]·Cl·4H2O} (1)

The X-ray single-crystal diffraction studies reveal that compound LaL 1 crystallizes in the Pī space group of the triclinic system. As shown in Figure 1a, each La3+ center is eight-coordinated by six oxygen atoms (O1, O5, O8A, O4C, O2B, and O3D) from different L ligands and two oxygen atoms (O9 and O10) from two coordinated water molecules. The asymmetric unit of LaL 1 contains one crystallographically independent La3+, two deprotonated L ligands, two coordinated water molecules, one free Cl ion, and four guest water molecules. There are two kinds of coordination modes for the two carboxylate groups of L ligand; one adopts bidentate bridging coordination fashion μ21η1 and the other one shows monodentate bridging coordination mode μ11. The two adjacent La3+ ions are connected together through four carboxylate groups of four L ligands in bidentate bridging coordination mode to form a dinuclear unit (Figure 1), then neighboring dinuclear subunits are bridged by these four L ligands in opposite direction to form 1D beaded chain. When adjacent, these chains are further connected by other 1D beaded chains, building by L ligands in unidentate bridging coordination mode, to form a 2D network (Figure 1). The angle between these two chains is 78.71°. The 3D supramolecular architecture of compound LaL 1 is obtained from the interlayer π···π stacking interactions between neighboring phenyl rings and benzene rings in adjacent layers (centroid−centroid: 3.534(2) Å, 3.888(3) Å and 3.623(2) Å). (Figure 1)

2.1.2. PXRD Analysis

To determine whether the crystal structures are truly representative of the bulk materials tested in property studies, powder X-ray diffraction (PXRD) experiments were carried out for compounds 13. The PXRD experimental and as-simulated patterns of compounds 13 are shown in the Figure 2. Failing to obtain crystals suitable for single-crystal crystallography, we were unable to determine the structures of TbL 2 and EuL 3, but the PXRD patterns of TbL 2 and EuL 3 are in good agreement with that of LaL 1, with only minor shifts in peak positions, indicating that TbL 2 and EuL 3 are isomorphous with LaL 1, and indicating that the crystal structures are truly representative of the bulk crystal products. The differences in intensity may be owing to the preferred orientation of the crystal samples.

2.2. Solid-State Luminescence of 13

The solid-state luminescence properties and the excitation spectrum of the H2LCl ligand and compounds 13 were recorded at room temperature as shown in Figure 3 and Figure S1, respectively. As shown in Figure 3, complex LaL 1 presents a broad emission band centered at 409 nm (λex = 332 nm), which is assigned to the typical π*–π transition of ligands (Figure 3). The blue shift compared to the ligand emission is presumably due to a conformational change in the ligand upon binding with the metal ions [20,21,22,23]. Upon excitation at 332 nm, complex TbL 2 showed four characteristic emission bands of Tb3+ ion, centered at 489, 544, 582, and 619 nm (Figure 3), and these emissions are attributed to the f–f electronic transitions 5D47FJ (J = 6, 5, 4 and 3), respectively. Among these transitions, 5D47F5 (544 nm) green emission is a dominating intensity. While being excited at 332 nm, complex EuL 3 displays emissions at 590, 613, and 653 nm, which are attributed to the 5D07FJ (J = 1, 2, and 3) transitions of the Eu3+ ion, with a maximum at 5D07F2 transition (613 nm) (Figure 3). Notably, there was no emission band of the ligand observed in the emission spectra of the compounds, indicating efficient energy transfer from the ligand to the Tb3+ ion or Eu3+ ion. These results clearly indicate that the L ligand is an excellent antenna chromophore for sensitizing both Tb3+ and Eu3+ ions.

2.3. Organic Small Molecule Sensing

The high luminescence intensities of those CPs prompt us to utilize TbL 2 as a representative to consider the potential luminescence sensing properties in different solvent suspension. Firstly, the finely ground sample of TbL 2 (10 mg) is immersed in different common organic solvents (5 mL), treated by ultrasonication for 60 min, and then aged for 24 h to generate stable suspensions before the fluorescence study. The solvents used are methanol, trichloromethane, tetrahydrofuran, dichloromethane, cyclohexane, n-hexane, acetonitrile (CH3CN), N,N-dimethylacetamide (DMA), N,N-dimethylformamide (DMF), and acetone. Interestingly, the most interesting feature is that its luminescent emission spectrum, monitored at 613 nm (5D47F3), is largely dependent on the solvent molecules, particularly in the case of acetone, which exhibit the most significant quenching effects (Figure 4). Such solvent-dependent luminescence properties are of interest for the sensing of acetone solvent molecules. To explain the reason for the quenching effect, the absorption spectra of methanol and acetone are investigated, which reveals that acetone has a wide absorption range from 308 to 360 nm, while methanol exhibits no absorption (Figure S2). The absorbing band of acetone overlays part of the absorption band of sensor TbL 2, which may lead to energy transfer occuring between the sensor TbL 2 and the acetone molecules. Due to the intermolecular solute-solvent interactions between TbL 2 and acetone, the energy absorbed by TbL 2 is transferred to acetone molecules, resulting in a decrease in the luminescent intensity, which is similar to the previously reported CP-based sensors [22,23].
For better understanding of the sensing ability of TbL 2 for acetone, the sensing properties of TbL 2 for acetone were also investigated by recording the emissive spectra of the 2-methanol suspension with the gradual addition of acetone. As shown in Figure 4, the luminescence intensity monitored at 613 nm decreases clearly with the increasing acetone concentration. Moreover, the quenching effect can be rationalized for the low concentrations by the Stern–Volmer equation: Io/I = 1 + Ksv × [M], where Io and I are the luminescence intensities of ethanol suspensions of complex TbL 2 before and after the addition of acetone, respectively; Ksv is the quenching constant, and [M] is the molar concentration of acetone [20,21,22,23,24]. On the basis of the quenching experimental data, the linear correlation coefficient (R2) in the Ksv curve of TbL 2 with addition of acetone is 0.9877 (Figure 4), suggesting that the quenching effect of TNP on the fluorescence of TbL 2 fits well the Stern–Volmer model. The Ksv is 2.0263 × 104 M−1 at low concentrations of acetone, which is comparable to those of previously reported CP-based sensors [20,21,22,23,24].
Additionally, we also found that TbL 2 can be regenerated and reused for five numbers of cycles by centrifuging the dispersed crystals in methanol after sensing and washing several times with methanol (Figure 5). The quenching efficiencies of every cycle are basically unchanged through monitoring the emission spectra of TbL 2 dispersed in the presence of 200 μM acetone in methanol. Furthermore, the PXRD patterns of the initial sample and recovered by centrifuging of the dispersed crystals in methanol after sensing and washing several times with methanol indicate the high stability of this compound (Figure S3). The result reveals that TbL 2 could be applied as a fluorescence sensor for acetone with high selectivity and recyclability.

3. Experimental Section

CCDC 1498357 (1) contains the supplementary crystallographic data for this paper. This data can be obtained free of charge from the Cambridge Crystallographic Data Center via the internet at http://www.ccdc.cam.ac.uk/data_request/cif, or by e-mailing [email protected].

3.1. Material and Methods

All of the reagents and solvents used in this paper were purchased from Adamas-beta Co., Ltd. (Shanghai, China) and used as received without further purification, unless otherwise indicated. The NMR spectra were recorded on a Bruker DRX400 (1H: 400 MHZ, 13C: 100 MHZ), chemical shifts (δ) are expressed in ppm, and J values are given in Hz, and deuterated DMSO was used as solvent. IR spectra were recorded on a FT-IR Thermo Nicolet Avatar 360 using KBr pellet. The phase purity of the samples was investigated by powder X-ray diffraction (PXRD) measurements carried out on a Bruker D8-Advance diffractometer equipped with CuKα radiation (λ = 1.5406 Å) at a scan speed of 1°/min. Thermal stability studies were carried out on a NETSCHZ STA-449C thermoanalyzer with a heating rate of 10 °C/min under a nitrogen atmosphere. All fluorescence measurements were performed on an Edinburgh Instrument F920 spectrometer. C, H, and N were determined on an Elementar Vario III EL elemental analyzer.

3.2. Synthetic Procedures

3.2.1. Synthesis of 4-Carboxy-1-(4-Carboxybenzyl)Pyridinium Chloride (H2LCl)

A mixture of isonicotinic acid (1.231 g, 10 mmol) and methyl 4-(bromomethyl)benzoate (2.291 g, 10 mmol) in CH3CN (50 mL) was refluxed for 8 h. After the mixture was cooled down to room temperature, the resulting precipitate was filtered to provide a white solid which was dissolved and refluxed in 100 mL 2 N HCl aqueous solution for 2 h. After the solution cooled down to ambient temperature, the white precipitate formed were collected by filtration and washed with ether (20 mL × 3) to afford H2LCl (2.712 g, 92%) which was identical to data in the literature [15,26]. 1H NMR (400 Hz, DMSO-d6): δ = 9.46 (d, 2H, ArH, J = 6.4 Hz), 8.52 (d, 2H, ArH, J= 6.4 Hz), 7.99 (d, 2H, ArH, J = 8.4 Hz), 7.67 (d, 2H, ArH, J = 8.4 Hz), 6.12 (s, 2H, CH2); 13C NMR (100 Hz, DMSO-d6): δ = 167.2, 163.8, 146.9, 146.5, 139.0, 132.1, 130.45, 129.5, 128.2, 63.2. (Figure S4).

3.2.2. Synthesis of {[La(L)2·2H2O]·Cl·4H2O} (LaL 1)

LaCl3·6H2O (0.3 mmol), H2LCl (0.1 mmol), NaOH (0.2 mmol) and H2O (10 mL) were added to a 15 mL Telfon-lined stainless steel autoclave, and the solution was heated at 150 °C for 24 h, then cooled down to room temperature at a rate of 2 °C/h. Colorless block crystals of the product were collected by filtration and washed with water several times, then dried in air. The obtained yield based on the H2LCl was 22.5%. Main IR (KBr, cm−1): 3382, 3113, 3051, 1621, 1540, 1397, 1172, 1131 (Figure S5). The TG curve for LaL 1 shows a gradual weight loss 13.8% between 90 and 280 °C, which can be ascribed to the removal of two lattice water molecules and four free water molecules (Figure S6). The phase purity of the product was confirmed by PXRD experiments (Figure 2). Anal. Calcd for C28H32ClLaN2O8 (698.9337): C, 48.12; H, 4.61; N, 4.01. Found: C, 48.05; H, 4.69; N, 3.98.

3.2.3. Synthesis of {[Tb(L)2·2H2O]·Cl·4H2O} (TbL 2)

All attempts to get single crystals of TbL 2 by different methods were in vain. The procedure was similar to that of complex LaL 1 except that TbCl3·6H2O was used instead of LaCl3·6H2O. Yield: 19.2% (based on the ligand). The phase purity of the product was confirmed by PXRD experiments (Figure 2). Main IR (KBr, cm−1): 3395, 3113, 3047, 1625, 1544, 1397, 1176, 1131. (Figure S5). Anal. Calcd for C28H32ClN2O8Tb (718.9437): C, 46.78; H, 4.49; N, 3.90. Found: C, 46.71; H, 4.53; N, 3.87.

3.2.4. Synthesis of {[Eu(L)2·2H2O]·Cl·4H2O} (EuL 3)

All attempts to get single crystals of EuL 3 by different methods were in vain. The procedure was similar to that of complex LaL 1 except that EuCl3·6H2O was used instead of LaCl3·6H2O. Colorless precipitates were isolated in a yield of 13.5% (based on the ligand). The phase purity of the product was confirmed by PXRD experiments (Figure 5). Main IR (KBr, cm−1): 3391, 3109, 3051, 1621, 1548, 1397, 1176, 1131. (Figure S5). Anal. Calcd for C28H32ClEuN2O8 (711.9837): C, 47.24; H, 4.53; N, 3.93. Found: C, 47.19; H, 4.58; N, 3.91.

3.3. Crystallography

Crystallographic data were collected at 296 K on a Bruker Smart AXS CCD diffractometer with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å) using ω-scan technique. Cell parameters were retrieved using SMART software and refined with SAINT on all observed reflections. Absorption corrections were applied with the program SADABS. Structure LaL 1 was solved by direct methods using SHELXS-97 [27] and refined on F2 by full-matrix least-squares procedures with SHELXL-97 [28]. All nonhydrogen atoms were located in different Fourier syntheses and finally refined with anisotropic displacement parameters. Hydrogen atoms attached to the organic moieties were either located from the difference Fourier map or fixed stereochemically. The final chemical formula was estimated and combined with the TGA results. Details of the crystallographic data collection and refinement parameters are summarized in Table 1. Main bond lengths and angles are presented in Table 2.

4. Conclusions

In summary, we have designed three isostructural lanthanide coordination polymers (Ln-CPs) to explore chemosensors. Firstly, the solid-state luminescence properties of those Ln-CPs were investigated, realizing the zwitterionic ligand (L) is an excellent antenna chromophore for sensitizing both Tb3+ and Eu3+ ions. In addition, TbL 2 as a representative chemosensor, material could not only serve as a chemosensor for acetone but also can be regenerated and reused for at least five cycles. Owing to a few luminescent CP, sensors for detecting acetone have been realized and reported [20,21,22,23], to the best of our knowledge, those materials have the potential to serve as the first case of a luminescent Ln-CP material based on the zwitterionic type of organic ligand for selective and recyclable sensing of acetone in methanol solution. Further studies are under way in our lab.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4352/7/7/199/s1. Figure S1: The solid-state excitation spectra of ligand H2LCl, complex 1, 2, and 3. Figure S2: The UV/Vis absorption spectra for acetone and methanol. Figure S3: PXRD patterns of 1 after detection of acetone (after five cycles reused). Figure S4: 1H NMR (400 Hz, DMSO-d6) and 13C NMR (100 Hz, DMSO-d6) spectra of H2LCl Ligand. Figure S5: The IR spectra of ligand H2LCl, complex 1, 2, and 3. Figure S6: The TGA diagram of 1 as a representative.

Acknowledgments

This work was supported by the Yunnan Provincial Department of Education funded projects (2017ZZX085), the National Natural Science Foundation of China (Projects 21371151) and the Yunnan Province doctoral newcomer Award (YN2015015 and YN2014017).

Author Contributions

Kaimin Wang designed the experiments and wrote the paper; Yulu Ma carried out the synthesis experiments; Huaijun Tang performed X-ray structure determination and Yulu Ma analyzed the results; Huaijun Tang performed X-ray powder diffraction analysis; Yulu Ma performed the other experiments. All authors took part in the discussion processes and have approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jin, J.-C.; Wu, J.; Yang, G.-P.; Wu, Y.-L.; Wang, Y.-Y. A microporous anionic metal–organic framework for a highly selective and sensitive electrochemical sensor of Cu2+ ions. Chem. Commun. 2016, 52, 8475–8478. [Google Scholar] [CrossRef] [PubMed]
  2. Hu, Z.; Deibert, B.J.; Li, J. Luminescent metal–organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815–5840. [Google Scholar] [CrossRef] [PubMed]
  3. Cao, C.-S.; Hu, H.-C.; Xu, H.; Qiao, W.-Z.; Zhao, B. Two solvent-stable MOFs as a recyclable luminescent probe for detecting dichromate or chromate anions. CrystEngComm 2016, 18, 4445–4451. [Google Scholar] [CrossRef]
  4. Chen, B.; Wang, L.; Zapata, F.; Qian, G.; Lobkovsky, E.B. A Luminescent Microporous Metal−Organic Framework for the Recognition and Sensing of Anions. J. Am. Chem. Soc. 2008, 130, 6718–6719. [Google Scholar] [CrossRef] [PubMed]
  5. Shi, N.; Zhang, Y.; Xu, D.; Song, C.; Jin, X.; Liu, D.; Xie, L.; Huang, W. π-System based coordination polymer hollow nanospheres for the selective sensing of aromatic nitro explosive compounds. New J. Chem. 2015, 39, 9275–9280. [Google Scholar] [CrossRef]
  6. Barea, E.; Montoro, C.; Navarro, J.A.R. Toxic gas removal—metal–organic frameworks for the capture and degradation of toxic gases and vapours. Chem. Soc. Rev. 2014, 43, 5419–5430. [Google Scholar] [CrossRef] [PubMed]
  7. Qin, L.; Lin, L.-X.; Fang, Z.-P.; Yang, S.-P.; Qiu, G.-H.; Chen, J.-X.; Chen, W.-H. A water-stable metal–organic framework of a zwitterionic carboxylate with dysprosium: A sensing platform for Ebolavirus RNA sequences. Chem. Commun. 2016, 52, 132–135. [Google Scholar] [CrossRef] [PubMed]
  8. Shen, X.; Yan, B. Polymer hybrid thin films based on rare earth ion-functionalized MOF: Photoluminescence tuning and sensing as a thermometer. Dalton Trans. 2015, 44, 1875–1881. [Google Scholar] [CrossRef] [PubMed]
  9. Du, P.-Y.; Gu, W.; Liu, X. Highly selective luminescence sensing of nitrite and benzaldehyde based on 3d–4f heterometallic metal–organic frameworks. Dalton Trans. 2016, 45, 8700–8704. [Google Scholar] [CrossRef] [PubMed]
  10. Xu, Z.-X.; Ma, Y.-L.; Zhang, L.-S.; Zhang, J. A couple of Co(II) enantiomers constructed from semirigid lactic acid derivatives. Inorg. Chem. Commum. 2016, 73, 115–118. [Google Scholar] [CrossRef]
  11. Xu, Z.-X.; Ma, Y.-L.; Lv, G.-L. Homochiral coordination polymers with helixes and metal clusters based on lactate derivatives. J. Solid State Chem. 2017, 249, 210–214. [Google Scholar] [CrossRef]
  12. Suckert, S.; Germann, L.S.; Dinnebier, R.E.; Werner, J.; Näther, C. Synthesis, Structures and Properties of Cobalt Thiocyanate Coordination Compounds with 4-(hydroxymethyl)pyridine as Co-ligand. Crystals 2016, 6, 38. [Google Scholar] [CrossRef]
  13. Semitut, E.; Komarov, V.; Sukhikh, T.; Filatov, E.; Potapov, A. Synthesis, Crystal Structure and Thermal Stability of 1D Linear Silver(I) Coordination Polymers with 1,1,2,2-Tetra(pyrazol-1-yl)ethane. Crystals 2016, 6, 138. [Google Scholar] [CrossRef]
  14. Nakanishi, T.; Sato, O. Synthesis, Structure, and Magnetic Properties of New Spin Crossover Fe(II) Complexes Forming Short Hydrogen Bonds with Substituted Dicarboxylic Acids. Crystals 2016, 6, 131. [Google Scholar] [CrossRef]
  15. Wang, K.-M.; Du, L.; Ma, Y.-L.; Zhao, J.-S.; Wang, Q.; Yan, T.; Zhao, Q.-H. Multifunctional chemical sensors and luminescent thermometers based on lanthanide metal–organic framework materials. CrystEngComm 2016, 18, 2690–2700. [Google Scholar] [CrossRef]
  16. Ma, Y.; Du, L.; Wang, K.; Zhao, Q. Synthesis, Crystal Structure, Luminescence and Magnetism of Three Novel Coordination Polymers Based on Flexible Multicarboxylate Zwitterionic Ligand. Crystals 2017, 7, 32. [Google Scholar] [CrossRef]
  17. Pan, M.; Du, B.-B.; Zhu, Y.-X.; Yue, M.-Q.; Wei, Z.-W.; Su, C.-Y. Highly Efficient Visible-to-NIR Luminescence of Lanthanide (III) Complexes with Zwitterionic Ligands Bearing Charge-Transfer Character: Beyond Triplet Sensitization. Chem. Eur. J. 2016, 22, 2440–2451. [Google Scholar] [CrossRef] [PubMed]
  18. Wen, R.-M.; Han, S.-D.; Ren, G.-J.; Chang, Z.; Li, Y.-W.; Bu, X.-H. A flexible zwitterion ligand based lanthanide metal–organic framework for luminescence sensing of metal ions and small molecules. Dalton Trans. 2015, 44, 10914–10917. [Google Scholar] [CrossRef] [PubMed]
  19. Li, Y.; Song, H.; Chen, Q.; Liu, K.; Zhao, F.-Y.; Ruan, W.-J.; Chang, Z. Two coordination polymers with enhanced ligand-centered luminescence and assembly imparted sensing ability for acetone. J. Mater. Chem. A 2014, 2, 9469–9473. [Google Scholar] [CrossRef]
  20. Shi, Y.-X.; Hu, F.-L.; Zhang, W.-H.; Lang, J.-P. A unique Zn (II)-based MOF fluorescent probe for the dual detection of nitroaromatics and ketones in water. CrystEngComm 2015, 17, 9404–9412. [Google Scholar] [CrossRef]
  21. Hua, J.-A.; Zhao, Y.; Kang, Y.-S.; Lu, Y.; Sun, W.-Y. Solvent-dependent zinc (II) coordination polymers with mixed ligands: selective sorption and fluorescence sensing. Dalton Trans. 2015, 44, 11524–11532. [Google Scholar] [CrossRef] [PubMed]
  22. Gu, F.; Chen, H.; Han, D.; Wang, Z. Metal–organic framework derived Au@ ZnO yolk–shell nanostructures and their highly sensitive detection of acetone. RSC Adv. 2016, 6, 29727–29733. [Google Scholar] [CrossRef]
  23. Liu, X.-J.; Zhang, Y.-H.; Chang, Z.; Li, A.-L.; Tian, D.; Yao, Z.-Q.; Jia, Y.-Y.; Bu, X.-H. A Water-Stable Metal–Organic Framework with a Double-Helical Structure for Fluorescent Sensing. Inorg. Chem. 2016, 55, 7326–7328. [Google Scholar] [CrossRef] [PubMed]
  24. Hao, J.-N.; Yan, B. Ln3+ post-functionalized metal–organic frameworks for color tunable emission and highly sensitive sensing of toxic anions and small molecules. New J. Chem. 2016, 40, 4654–4661. [Google Scholar] [CrossRef]
  25. Barry, D.E.; Caffrey, D.F.; Gunnlaugsson, T. Lanthanide-directed synthesis of luminescent self-assembly supramolecular structures and mechanically bonded systems from acyclic coordinating organic ligands. Chem. Soc. Rev. 2016, 45, 3244–3274. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, J.-Y.; Wang, K.; Li, X.-B.; Gao, E.-Q. Magnetic coupling and slow relaxation of magnetization in chain-based MnII, CoII, and NiII coordination frameworks. Inorg. Chem. 2014, 53, 9306–9314. [Google Scholar] [CrossRef] [PubMed]
  27. Sheldrick, G.M. SHELXS-97, Program for the Solution of Crystal Structures; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  28. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Coordination environment of the La3+ ion center. Symmetry codes: (A) A –x + 1, −y + 1, −z + 1; (B) –x + 2, -y + 2, −z + 2; (C) –x + 1, −y + 1, −z + 2; (D) x + 1, y + 1, z. (b) Two adjacent La3+ ions are bridged by four carboxylate groups from four L ligands. (c) 2D network of compound 1. (d) 3D supramolecular structure of compound 1.
Figure 1. (a) Coordination environment of the La3+ ion center. Symmetry codes: (A) A –x + 1, −y + 1, −z + 1; (B) –x + 2, -y + 2, −z + 2; (C) –x + 1, −y + 1, −z + 2; (D) x + 1, y + 1, z. (b) Two adjacent La3+ ions are bridged by four carboxylate groups from four L ligands. (c) 2D network of compound 1. (d) 3D supramolecular structure of compound 1.
Crystals 07 00199 g001aCrystals 07 00199 g001b
Figure 2. The PXRD patterns of LaL 1, TbL 2, EuL 3.
Figure 2. The PXRD patterns of LaL 1, TbL 2, EuL 3.
Crystals 07 00199 g002
Figure 3. The solid-state emisssion spectra of (a) the H2LCl ligand with λex = 302 nm, (b) LaL 1, with λex = 332 nm, (c) TbL 2 with λex = 332 nm, (d) EuL 3 with λex = 332 nm.
Figure 3. The solid-state emisssion spectra of (a) the H2LCl ligand with λex = 302 nm, (b) LaL 1, with λex = 332 nm, (c) TbL 2 with λex = 332 nm, (d) EuL 3 with λex = 332 nm.
Crystals 07 00199 g003
Figure 4. (a) Comparisons of the luminescence intensity of TbL 2 in different solvent suspensions at room temperature (λex = 332 nm). (b) The luminescence emission spectra titration of TbL 2-dispersed in methanol with the addition of different concentrations of acetone at room temperature (λex = 332 nm). (c) Stern–Volmer plot of Io/I versus the concentration of acetone in the 2-menthanol suspension (insert: enlarged view of a selected area).
Figure 4. (a) Comparisons of the luminescence intensity of TbL 2 in different solvent suspensions at room temperature (λex = 332 nm). (b) The luminescence emission spectra titration of TbL 2-dispersed in methanol with the addition of different concentrations of acetone at room temperature (λex = 332 nm). (c) Stern–Volmer plot of Io/I versus the concentration of acetone in the 2-menthanol suspension (insert: enlarged view of a selected area).
Crystals 07 00199 g004
Figure 5. The quenching and recyclability test of TbL 2 upon addition of 200 μM acetone in methanol solution. The luminescence was recovered by centrifuging the dispersed crystals in methanol after sensing and washing several times with methanol.
Figure 5. The quenching and recyclability test of TbL 2 upon addition of 200 μM acetone in methanol solution. The luminescence was recovered by centrifuging the dispersed crystals in methanol after sensing and washing several times with methanol.
Crystals 07 00199 g005
Table 1. Crystal data and structure refinement for 1.
Table 1. Crystal data and structure refinement for 1.
Compound1
Chemical formulaC28H32N2O14ClLa
Formula mass794.92
Crystal systemTriclinic
a11.3001(7)
b11.7346(7)
c14.6280(9)
α/°100.2340(10)
β/°93.7830(10)
γ/°109.1450(10)
Unit cell volume/Å31787.09(19)
Space group
Z2
DX/Mg m−31.343
μ/mm−11.318
Reflections with I > 2σ(I)7215
Independent reflections8114
F(000)720
Rint0.0245
GOF on F21.094
R1, wR2 [I > 2σ(I)]0.0327, 0.0955
R1, wR2 [all data]0.0375, 0.1000
Residuals/e Å−30.793, −0.638
CCDC number1498357
Table 2. Elected bond lengths (Å) and angles (°) for compound 1.
Table 2. Elected bond lengths (Å) and angles (°) for compound 1.
Bond lengths (Å) for 1
La(1)-O(5)2.448(3)O(2)-C(6)1.254(4)
La(1)-O(8)A2.490(3)O(3)-C(14)1.246(4)
La(1)-O(1)2.504(3)O(4)-C(14)1.260(4)
La(1)-O(2)B2.508(3)O(5)-C(20)1.237(5)
La(1)-O(4)C2.512(2)O(6)-C(20)1.252(6)
La(1)-O(3)D2.562(2)O(7)-C(28)1.247(5)
La(1)-O(10)2.569(3)O(8)-C(28)1.269(5)
La(1)-O(9)2.605(3)N(1)-C(4)1.356(5)
La(1)-O(4)D2.959(2)N(1)-C(3)1.380(5)
O(1)-C(6)1.244(4)N(1)-C(7)1.502(4)
Bond angles (°) for 1
O(5)-La(1)-O(8)A91.37(11)O(3)D-La(1)-O(10)68.16(9)
O(5)-La(1)-O(1)145.38(9)O(5)-La(1)-O(9)71.04(10)
O(8)A-La(1)-O(1)79.47(10)O(8)A-La(1)-O(9)69.88(10)
O(5)-La(1)-O(2)B79.90(9)O(1)-La(1)-O(9)133.44(9)
O(8)A-La(1)-O(2)B139.19(10)O(2)B-La(1)-O(9)69.58(10)
O(1)-La(1)-O(2)B127.76(9)O(4)C-La(1)-O(9)75.85(9)
O(5)-La(1)-O(4)C144.80(9)O(3)D-La(1)-O(9)138.14(10)
O(8)A-La(1)-O(4)C88.20(9)O(10)-La(1)-O(9)127.61(10)
O(1)-La(1)-O(4)C68.85(8)O(5)-La(1)-O(4)D119.81(9)
O(2)B-La(1)-O(4)C77.69(9)O(8)A-La(1)-O(4)D146.12(9)
O(5)-La(1)-O(3)D79.70(9)O(1)-La(1)-O(4)D67.36(8)
O(8)A-La(1)-O(3)D141.40(9)O(2)B-La(1)-O(4)D65.49(8)
O(1)-La(1)-O(3)D86.90(9)O(4)C-La(1)-O(4)D73.97(8)
O(2)B-La(1)-O(3)D76.46(9)O(3)D-La(1)-O(4)D46.25(7)
O(4)C-La(1)-O(3)D120.19(8)O(10)-La(1)-O(4)D101.20(8)
O(5)-La(1)-O(10)74.02(10)O(9)-La(1)-O(4)D129.85(9)
O(8)A-La(1)-O(10)73.27(9)C(6)-O(1)-La(1)139.4(2)
O(1)-La(1)-O(10)71.36(9)C(6)-O(2)-La(1)B136.0(2)
O(2)B-La(1)-O(10)138.94(10)C(4)-N(1)-C(3)120.5(3)
O(4)C-La(1)-O(10)138.48(9)C(4)-N(1)-C(7)119.4(3)
O(3)D-La(1)-O(10)68.16(9)C(3)-N(1)-C(7)120.1(3)

Share and Cite

MDPI and ACS Style

Wang, K.; Ma, Y.; Tang, H. Lanthanide Coordination Polymers as Luminescent Sensors for the Selective and Recyclable Detection of Acetone. Crystals 2017, 7, 199. https://doi.org/10.3390/cryst7070199

AMA Style

Wang K, Ma Y, Tang H. Lanthanide Coordination Polymers as Luminescent Sensors for the Selective and Recyclable Detection of Acetone. Crystals. 2017; 7(7):199. https://doi.org/10.3390/cryst7070199

Chicago/Turabian Style

Wang, Kaimin, Yulu Ma, and Huaijun Tang. 2017. "Lanthanide Coordination Polymers as Luminescent Sensors for the Selective and Recyclable Detection of Acetone" Crystals 7, no. 7: 199. https://doi.org/10.3390/cryst7070199

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop