Next Article in Journal
Production Methods of Van der Waals Heterostructures Based on Transition Metal Dichalcogenides
Next Article in Special Issue
Synthesis and Crystal Structure of a Zn(II)-Based MOF Bearing Neutral N-Donor Linker and SiF62− Anion
Previous Article in Journal
Study on Dislocation-Dopant Ions Interaction in Ionic Crystals by the Strain-Rate Cycling Test during the Blaha Effect
Previous Article in Special Issue
Prussian Blue Analogue Mesoframes for Enhanced Aqueous Sodium-ion Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Incorporating the Thiazolo[5,4-d]thiazole Unit into a Coordination Polymer with Interdigitated Structure

Institut für Anorganische Chemie und Strukturchemie, Heinrich-Heine-Universität, Universitätsstraße 1, 40225 Düsseldorf, Germany
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(1), 30; https://doi.org/10.3390/cryst8010030
Submission received: 21 December 2017 / Revised: 5 January 2018 / Accepted: 6 January 2018 / Published: 12 January 2018
(This article belongs to the Special Issue Structural Design and Properties of Coordination Polymers)

Abstract

:
The linker 2,5-di(4-pyridyl)thiazolo[5,4-d]thiazole (Dptztz), whose synthesis and structure is described here, was utilized together with benzene-1,3-dicarboxylate (isophthalate, 1,3-BDC2−) for the preparation of the two-dimensional coordination network [Zn(1,3-BDC)Dptztz]·DMF (DMF = dimethylformamide) via a solvothermal reaction. Compound [Zn(1,3-BDC)Dptztz]·DMF belongs to the class of coordination polymers with interdigitated structure (CIDs). The incorporated DMF solvent molecules can be removed through solvent exchange and evacuation such that the supramolecular 3D packing of the 2D networks retains porosity for CO2 adsorption in activated [Zn(1,3-BDC)Dptztz]. The first sorption study of a tztz-functionalized porous metal-organic framework material yields a BET surface of 417 m2/g calculated from the CO2 adsorption data. The heat of adsorption for CO2 exhibits a relative maximum with 27.7 kJ/mol at an adsorbed CO2 amount of about 4 cm3/g STP, which is interpreted as a gate-opening effect.

Graphical Abstract

1. Introduction

Metal-organic frameworks are an intensively studied class of porous materials. Due to the immense quantity of possible inorganic and organic building units several applications are discussed (e.g., gas storage and separation, catalysis, sensing and heat transformation) [1,2,3,4,5,6,7,8,9,10,11,12,13]. Many different organic functionalities have been introduced into the frameworks either by a priori ligand functionalization or by post synthetic modification [14,15,16]. 4,4′-Biypridine based ligands have been used to construct a diverse set of different topologies (e.g., one-dimensional chains, ladders, two-dimensional and three-dimensional networks) [17]. 4,4′-Bipyridine ligands are also widely used in the synthesis of open network structures in combination with dicarboxylate ligands (e.g., terephthalate, isophthalate) via the so called mixed-ligand strategy [18,19]. Through functionalization of the organic ligands, the pore surface of those mixed-ligand MOFs can be tuned to enhance the selectivity in their sorption or sensing properties [20,21,22,23]. One famous family of mixed-ligands MOFs are the CIDs (coordination polymers with interdigitated structure) popularized by Kitagawa and co-workers. CIDs consist of an angular ligand (e.g., isophthalate, benzophenone-4,4′-dicarboxylate, azulene-1,6-dicarboxylate) and a 4,4′-bipyridine derivative and divalent transition metal ions. CIDs show very intriguing sorption properties due to their potential for functionalization and often inherent structural flexibility [24,25,26,27,28,29].
The heterocyclic thiazolo[5,4-d]thiazole (tztz) system (Figure 1) experienced a renaissance in the last decade after it was first prepared by Ketcham et al. in the 1960s [30]. The tztz unit was incorporated into photoactive materials due to its rigid and planar structure and electron deficiency. Both Maes et al. and Dessi et al. reviewed the synthetic procedures to obtain tztz-containing small molecules and polymers as well as their application in the field of organic electronics (e.g., OFETs, OSCs) [31,32]. In contrast, the tztz unit has been reported only in relatively few coordination compounds (15 hits in the CCDC database). The first examples were ruthenium and copper complexes with the doubly chelating 2,5-di(2-pyridyl)thiazolo[5,4-d]thiazole synthesized by Steel et al. [33]. Coordination polymers with 2,5-thiazolo[5,4-d]thiazoledicarboxylic acid (Figure 1) were obtained by Cheetham et al. with alkaline earth metals, whose connectivities vary with the cation size, and by Palmisano et al. with some transition metals, in which the ligand shows a N,O-chelating mode [34,35]. D’Alessandro et al. incorporated the donor-acceptor ligand N,N′-(thiazolo[5,4-d]thiazole-2,5-diylbis(4,1-phenylene))bis(N-(pyridine-4-yl)pyridin-4-amine into a two-dimensional zinc MOF and studied its electrochemical properties [36]. Recently the same group published the spectroelectrochemical properties of a ruthenium coordination complex with this ligand [37]. Additionally, Dai et al. synthesized tztz-linked microporous organic polymers, which show a high CO2:N2 selectivity [38].
Herein, we present the synthesis, structural analysis and the sorption properties of a new coordination polymer with interdigitated structure of the formula [Zn(1,3-BDC)Dptztz] consisting of Zn2+ ions, isophthalate and the 4,4′-bipyridine derivative 2,5-di(4-pyridyl)thiazolo[5,4-d]thiazole (Dptztz) (Figure 1, Scheme 1).

2. Materials and Methods

The chemicals used were obtained from commercial sources. No further purification has been carried out. CHN analysis was performed with a Perkin Elmer CHN 2400 (Perkin Elmer, Waltham, MA, USA). IR-spectra were recorded on a Bruker Tensor 37 IR spectrometer (Bruker Optics, Ettlingen, Germany) with ATR unit. Thermogravimetric analysis (TGA) was done with a Netzsch TG 209 F3 Tarsus (Netzsch, Selb, Germany) in the range from 20 to 700 °C, equipped with Al-crucible and applying a heating rate of 10 K·min−1 under nitrogen. The melting point was determined using a Büchi Melting Point apparatus B540. The powder X-ray diffraction pattern (PXRD) was obtained on a Bruker D2 Phaser powder diffractometer with a flat silicon, low background sample holder, at 30 kV, 10 mA for Cu-Kα radiation (λ = 1.5418 Å). Sorption isotherms were measured using a Micromeritics ASAP 2020 automatic gas sorption analyzer equipped with oil-free vacuum pumps (ultimate vacuum <10−8 mbar) and valves, which guaranteed contamination free measurements. The sample was connected to the preparation port of the sorption analyzer and degassed under vacuum until the outgassing rate, i.e., the rate of pressure rise in the temporarily closed manifold with the connected sample tube, was less than 2 μTorr/min at the specified temperature of 120 °C. After weighing, the sample tube was then transferred to the analysis port of the sorption analyzer. All used gases (He, N2, CO2) were of ultra-high purity (UHP, grade 5.0, 99.999%) and the STP volumes are given according to the NIST standards (293.15 K, 101.325 kPa). Helium gas was used for the determination of the cold and warm free space of the sample tubes. The heat of adsorption values were calculated using the ASAP 2020 v3.05 software. Water sorption isotherms were obtained volumetrically from a Quantachrome Autosorb iQ MP instrument equipped with an all-gas option. Prior to the sorption experiments, the compounds were degassed under dynamic vacuum.
2,5-Di(4-pyridyl)thiazolo[5,4-d]thiazole (Dptztz): 1.02 g (8.5 mmol) dithiooxamide and 2.0 mL (22 mmol, 2.6 eq) 4-pyridinecarboxaldehyde in 50 mL anhydrous DMF were refluxed for 6.5 h under nitrogen. During cooling the reaction mixture to room temperature, the product crystallized out in form of yellow prisms. The ligand was filtered and washed with a small amount of DMF and extensively with water. After drying in a vacuum oven at 60 °C overnight 1.82 g (6.1 mmol, 72 %) were obtained. 1H-NMR (300 MHz, DMSO-d6) δ [ppm]: 8.78 (d, 4JH,H = 6.12 Hz, 2H), 7.88 (d, 4JH,H = 6.12 Hz, 2H); MS (EI) m/z [rel. int.]: 296 (100%); 87.9 (91%); mp 319–322 °C.
[Zn(1,3-BDC)Dptztz]: 5.0 mg (0.017 mmol) of Dptztz were dissolved in 3 mL of hot DMF in a Pyrex tube. 5.4 mg (0.020 mmol) of Zn(NO3)2·4H2O and 2.8 mg (0.020 mmol) of isophthalic acid dissolved in 2 mL of DMF were added to the warm solution. The Pyrex tube was capped and placed into a preheated isothermal oven at 80 °C. After 12 h the first crystals appeared. After 3 days, the sample was removed from the oven and the solvent was directly exchanged with 3 × 3 mL of hot DMF. A light yellow crystal was selected to collect the single crystal data. Yield: 4 mg.
A larger amount of material was prepared by dissolving 100.4 mg (0.34 mmol) of Dptztz in 40 mL of hot DMF in a 100 mL Schott vial. Afterwards 88.4 mg (0.34 mmol) of Zn(NO3)·4H2O and 56.6 mg (0.34 mmol) of isophthalic acid dissolved in 10 mL of DMF were added and placed in an isothermal oven preheated at 120 °C. The sample was taken out after 3 days and the solvent was directly exchanged with 3 × 20 mL of hot DMF. Yield: 182.6 mg (90%). EA [%] calc. for: C22H12N4O4S2Zn C 50.25, H 2.30 N 10.70; found: C 50.89, H 2.93, N 11.51. IR (ATR) ṽmax [cm−1]: 3433, 1608, 1564, 1443, 1391, 1213, 1014, 833, 743, 724, 662, 619, 510.

Single Crystal X-ray Structures

Suitable crystals were carefully selected under a polarizing microscope, covered in protective oil and mounted on a 0.05 mm cryo loop. Data collection: Bruker Kappa APEX2 CCD X-ray diffractometer (Bruker AXS Inc., Madison, WI, USA) with microfocus tube, Mo-Kα radiation (λ = 0.71073 Å), multi-layer mirror system, ω-scans; data collection with APEX2 [39], cell refinement with SMART and data reduction with SAINT [39], experimental absorption correction with SADABS [40]. Structure analysis and refinement: All structures were solved by direct methods using SHELXL2014 [41,42]; refinement was done by full-matrix least squares on F2 using the SHELX-97 program suite. The hydrogen atoms for aromatic CH and for the amide group in DMF were positioned geometrically (C-H = 0.95 Å) and refined using a riding model (AFIX 43) with Uiso(H) = 1.2Ueq(C). The hydrogen atoms for CH3 of DMF were positioned geometrically (C-H = 0.98 Å) and refined using a riding model (AFIX 137) with Uiso(H) = 1.5Ueq(C). In [Zn(1,3-BDC)Dptztz] the thiazolothiazol (tztz) moiety was refined with a disorder model corresponding to a ring flip, which exchanges the S and N orientation, using PART n commands. The occupation factors of the S and N atoms were refined to about 0.904 for the A atoms and 0.096 for the B atoms. Thus, the disorder is relatively minor with only about 9.6% of the S and N atoms in the flipped position. The major atom tztz positions are designated as “A”, the minor ones as “B”. The minor positions could only be refined isotropically. The DMF crystal solvent molecule contained disorder in the two methyl groups, with two positions for each methyl group. This disorder does not give a perfect oriented Me2N group but we decided to keep the slightly disordered DMF molecule instead of removing its contribution with SQUEEZE. Each methyl group disorder was refined independently concerning the occupation factors. Crystal data and details on the structure refinement are given in Table 1. Graphics were drawn with DIAMOND [43]. Analyses on the supramolecular interaction were done with PLATON [44]. Phase purity and the representative nature of the single crystal was verified by positively matching the simulated and experimental powder X-ray diffractogram (PXRD) of the as-synthesized sample (Figure 2). CCDC 1812892 and 1812893 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html.

3. Results and Discussion

2,5-Di(4-pyridyl)thiazolo[5,4-d]thiazole (Dptztz) was synthesized according to the literature by the condensation of 4-pyridinecarboxaldehyde and dithioaxamide (Scheme 1) [45,46].
Single crystals of Dptztz were obtained after recrystallization from DMF in form of yellow prisms. Dptztz crystallizes in the monoclinic space group P21/c with half of the molecule in the asymmetric unit as the molecule sits on an inversion center (Figure 3). The molecule is almost planar with a dihedral angle between the pyridine ring and the tztz moiety of 13.65°.
Complementary CH⋯N hydrogen bonds between N1 and C1-H1 of adjacent Dptztz molecules form 1D strands which are parallel displaced by π-π interactions (Figure S1 in Supplementary Material).
Single crystals of the coordination network [Zn(1,3-BDC)Dptztz]·DMF were obtained after three days from a solvothermal reaction of Zn(NO3)2·4H2O, isophthalic acid and Dptztz in a molar ratio 1:1:1 in DMF at 80 °C. Due to the low solubility of Dptztz in common organic solvents, the reaction was carried out in a concentration of 3.4 × 10−3 mol/L and the mother liquor was directly exchanged with hot DMF after the crystallization process to remove unreacted Dptztz ligand. A larger amount of material for the sorption experiments was synthesized by scaling up the reaction by the factor of twenty in twice the concentration (6.8 × 10−3 mol/L) at 120 °C.
The crystal structure of the two-dimensional (2D) coordination network [Zn(1,3-BDC)Dptztz]·DMF was determined by single crystal diffraction analysis at 100 K. Compound [Zn(1,3-BDC)Dptztz] crystallizes in the triclinic space group P-1. The asymmetric unit consists of one Zn(II) ion, one molecule of the linkers 1,3-BDC2− and Dptztz, each, and a dimethylformamide (DMF) solvent molecule (Figure 4, Figure S2). One carboxylate group of 1,3-BDC2− connects two symmetry equivalent Zn(II) ions in a syn-syn-bis-monodentate coordination mode into a dinuclear unit with a Zn⋯Zn distance of 4.082 Å. The other carboxylate group chelates an adjacent Zn atom. Thereby the 1,3-BDC linkers bridge between neighboring dinuclear entities to form a one-dimensional double strand along the b-axis (Figure 5a). These double strands are pillared by Dptztz into a 2D coordination network structure (Figure 5b). The secondary building unit of the structure is the dinuclear unit {Zn2(O2C)4N4}. The 2D layers assemble through π-π interactions between isophthalate aryl rings and CH-π interactions between an isophthalate and a pyridyl-moiety of Dptztz of adjacent layers into a 3D supramolecular network (Figure 5c and Figure S3, Table S1). The 2D network in [Zn(1,3-BDC)Dptztz] is isotopic to the aforementioned CIDs (coordination polymers with interdigitated structure) studied by Kitagawa and co-workers [24,25,26,27,28,29].
The thermogravimetric analysis in Figure 6 shows a weight loss of 10.6% between 90 and 200 °C (calc. 12.2% for one DMF molecule per formula unit of [Zn(1,3-BDC)Dptztz]·DMF). The residual [Zn(1,3-BDC)Dptztz] framework shows stability up to 280 °C.
Prior to the sorption experiments the crystals of [Zn(1,3-BDC)Dptztz]·DMF were collected by suction filtration. Afterwards they were suspended in acetonitrile for three days to induce solvent exchange as part of the sample activation procedure. Additionally, the acetonitrile solvent was exchanged three times per day. Afterwards, the sample was degassed at 120 °C for 15 h under vacuum. The activated compound [Zn(1,3-BDC)Dptztz] shows no uptake of N2 at 77 K, which is in accordance with the observations by Kitagawa et al. for analogous CID structures [24,25,26,27,28,29]. For CO2—with its large polarizability and quadrupole moment—[Zn(1,3-BDC)Dptztz] shows a type I adsorption isotherm at 195 K with a maximum uptake of 138 cm3/g at 753 mmHg of CO2 (Figure 7). At higher absolutes pressures the desorption curve shows a small hysteresis, but at low pressures the hysteresis gap closes. This proves the microporous nature of [Zn(1,3-BDC)Dptztz]. Because [Zn(1,3-BDC)Dptztz] is non-porous towards N2 the CO2 data was used to calculated the BET surface area. The BET surface area from the CO2 adsorption isotherm is 417 m2/g (calculated from p/p0 = 0.00–0.04). The pore volume is 0.246 cm3/g at p/p0 = 0.5 calculated from the isotherm measured at 195 K. The calculated accessible surface area is 25.6% or 0.185 cm3/g calculated with PLATON from the DMF solvent-depleted structure. The measured pore volume is about 35% higher than the one calculated from the crystal structure data. This can be interpreted such that CO2 can create a larger interlayer volume through strong interaction with the highly polarized surface area at 195 K.
Additionally, the CO2 isotherms at 273 K and 293 K were measured (Figure 7). The data is presented in Table 3. From the CO2 isotherms at 273 K and 293 K, the heat of adsorption at zero coverage was derived as 26.2 kJ/mol. The heat of adsorption of CO2 for MOFs can vary from 20 kJ/mol to over 90 kJ/mol. A higher heat of adsorption is usually indicative for stronger framework CO2 interactions [47]. Representative values are 20kJ/mol for [Zn(1,4-BDC)(TED)], 30 kJ/mol for HKUST-1, 47 kJ/mol for Mg-MOF-74, 63 kJ/mol for MIL-100-Cr and 96 kJ/mol for mmen-Cu-BTTri [48,49,50,51,52].
The heat of adsorption curve (Figure 8) has a relative maximum at a quantity adsorbed of about 4 cm3/g STP with 27.7 kJ/mol. Afterwards the heat of adsorption decreases to 25.7 kJ/mol. For most MOF materials, the heat of adsorption curve decreases monotonically, since the adsorption sites with the highest affinity towards the adsorbate are occupied first and at higher loadings the adsorption sites usually have weaker affinities. Two MOF classes for which the heat of adsorption does not decrease monotonically are the MIL-53 and MIL-47 series. Férey et al. suggested that the transformation of MIL-53 from a closed or narrow-pore to the open or wide-pore phase is an endothermic process and that this process is balanced with the exothermic adsorption process. Subsequently, the MIL-53 MOFs also have a local maximum between 0 and 1 bar (see Figure S4 in Supplementary Material) [53]. Many CIDs also show gate-opening phenomena and/or an adsorbate specific expansion upon the adsorption process. Yet, to the best of our knowledge, no heat of adsorption curves for CIDs are published in the literature. But Pera–Titus and Farruseng calculated the phase transition energies for CID-21 and CID-22 (benzene and tetrazine spacer) to be 1.4 and 1.5 kJ/mol from the closed to the open phase for CO2 adsorption at 195 K, respectively [54]. These values are in good accordance with the difference between the heat of adsorption at zero coverage and the relative maximum of the curve in Figure 8. So it can be concluded that [Zn(1,3-BDC)Dptztz] shows a gate-opening effect for CO2, which is represented in a non-monotonic heat of adsorption curve.
Compound [Zn(1,3-BDC)Dptztz] gradually adsorbs H2O at 293 K with a maximum uptake of 121 mg/g at 0.9 p/p0 (Figure 9). This uptake equals 3.5 H2O molecules per asymmetric unit. The desorption curve shows a hysteresis, indicative for a strong interaction of H2O with the framework. [Zn(1,3-BDC)Dptztz] adsorbs CO2 at 195 K due to its large polarizability and quadrupole moment while N2 at 77 K is not adsorbed. The adsorption characteristics of [Zn(1,3-BDC)Dptztz] apparently depend on the interaction between the adsorbate and the framework und not only on the pore size. The sorption characteristics of [Zn(1,3-BDC)Dptztz] towards H2O with its pronounced hysteresis underpin these findings. It can be concluded that the decoration of the surface with the polarized and electron poor tztz moiety can alter sorption properties to become highly selective towards polarizable adsorbents. Further studies on different tztz-containing materials are underway in our institute.

4. Conclusions

The 4,4′-dipyridyl N,N′-donor ligand Dptztz with the central thiazolo[5,4-d]thiazole unit was successfully synthesized and its crystal structure was determined for the first time. With the linker Dptztz, the thiazolo[5,4-d]thiazole-unit was integrated into a solvent-filled coordination network of the formula [Zn(1,3-BDC)Dptztz]·DMF belonging to the class of coordination polymers with interdigitated structure (CIDs). Synthesis of the coordination network was carried out via a mixed ligand strategy in a solvothermal reaction. Interdigitation between the 2D layers to a 3D supramolecular network appears to be controlled by π-π interactions between isophthalate aryl rings and CH-π interactions between isophthalate and pyridyl moieties. Activated [Zn(1,3-BDC)Dptztz] shows a BET surface of 417 m2/g calculated from CO2 adsorption data, while N2 which unlike CO2 is not as polarizable and has no quadrupole moment is not adsorbed. The heat of adsorption for CO2 exhibits a relative maximum at a quantity adsorbed of about 4 cm3/g STP with 27.7 kJ/mol, which is interpreted as a gate-opening effect. This is the first report of the sorption characteristics of a tztz-functionalized porous MOF material.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4352/8/1/30/s1. Figure S1: Section of the packing diagram of 2,5-di(4-pyridyl)thiazolo[5,4-d]thiazole (Dptztz), showing (a) the complementary C1-H1⋯N1 hydrogen bonds and (b) the π-π interactions, thereby giving a supramolecular layer in the ab plane; Figure S2: Asymmetric unit of [Zn(1,3-BDC)Dptztz]·DMF showing the ring-flip disorder of the thiazolothiazol moiety; Figure S3: π-π and CH-π interactions between the 2D layers in [Zn(1,3-BDC)Dptztz]; Figure S4: Heat of adsorption curve of MIL-53-Cr; Table S1. Distances [Å] and angles [°] for the shortest π-π and CH-π supramolecular interactions between the 2D layers in [Zn(1,3-BDC)Dptztz].

Acknowledgments

The work was supported by the Federal German Ministry of Education and Research (BMBF) under grant-# 03SF0492C (Optimat).

Author Contributions

Simon Millan designed the experiments, synthesized the ligand and [Zn(1,3-BDC)Dptztz]. Gamall Makhloufi collected the single crystal data. Data analysis and measurements were performed by Simon Millan, while Christoph Janiak and Simon Millan wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705–714. [Google Scholar] [CrossRef] [PubMed]
  2. Maurin, G.; Serre, C.; Cooper, A.; Férey, G. The new age of MOFs and of their porous-related solids. Chem. Soc. Rev. 2017, 46, 3104–3107. [Google Scholar] [CrossRef] [PubMed]
  3. Furukawa, H.; Ko, N.; Go, Y.B.; Aratani, N.; Choi, S.B.; Choi, E.; Yazaydin, A.Ö.; Snurr, R.Q.; O’Keeffe, M.; Kim, J.; Yaghi, O.M. Ultrahigh porosity in metal-organic frameworks. Science 2010, 329, 424–428. [Google Scholar] [CrossRef] [PubMed]
  4. Adil, K.; Belmabkhout, Y.; Pillai, R.S.; Cadiau, A.; Bhatt, P.M.; Assen, A.H.; Maurin, G.; Eddaoudi, M. Gas/vapour separation using ultra-microporous metal–organic frameworks: Insights into the structure/separation relationship. Chem. Soc. Rev. 2017, 46, 3402–3430. [Google Scholar] [CrossRef] [PubMed]
  5. Li, J.-R.; Kuppler, R.J.; Zhou, H.-C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477–1504. [Google Scholar] [CrossRef] [PubMed]
  6. Dechnik, J.; Gascon, J.; Doonan, C.J.; Janiak, C.; Sumby, C.J. Mixed-matrix membranes. Angew. Chem. Int. Ed. 2017, 56, 9292–9310. [Google Scholar] [CrossRef] [PubMed]
  7. Lee, J.; Farha, O.K.; Roberts, J.; Scheidt, K.A.; Nguyen, S.T.; Hupp, J.T. Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450–1459. [Google Scholar] [CrossRef] [PubMed]
  8. Herbst, A.; Janiak, C. MOF catalysts in biomass upgrading towards value-added fine chemicals. CrystEngComm 2017, 19, 4092–4117. [Google Scholar] [CrossRef]
  9. Kitao, T.; Zhang, Y.; Kitagawa, S.; Wang, B.; Uemura, T. Hybridization of MOFs and polymers. Chem. Soc. Rev. 2017, 46, 3108–3133. [Google Scholar] [CrossRef] [PubMed]
  10. Lustig, W.P.; Mukherjee, S.; Rudd, N.D.; Desai, A.V.; Li, J.; Ghosh, S.K. Metal–organic frameworks: functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242–3285. [Google Scholar] [CrossRef] [PubMed]
  11. 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]
  12. Gangu, K.K.; Maddila, S.; Mukkamala, S.B.; Jonnalagadda, S.B. A review on contemporary Metal–Organic Framework materials. Inorg. Chim. Acta 2016, 446, 61–74. [Google Scholar] [CrossRef]
  13. Jeremias, F.; Fröhlich, D.; Janiak, C.; Henninger, S.K. Water and methanol adsorption on MOFs for cycling heat transformation processes. New J. Chem. 2014, 38, 1846–1852. [Google Scholar] [CrossRef]
  14. Almeida Paz, F.A.; Klinowski, J.; Vilela, S.M.F.; Tome, J.P.C.; Cavaleiro, J.A.S.; Rocha, J. Ligand design for functional metal-organic frameworks. Chem. Soc. Rev. 2012, 41, 1088–1110. [Google Scholar] [CrossRef] [PubMed]
  15. Tanabe, K.K.; Cohen, S.M. Postsynthetic modification of metal-organic frameworks—A progress report. Chem. Soc. Rev. 2011, 40, 498–519. [Google Scholar] [CrossRef] [PubMed]
  16. Islamoglu, T.; Goswami, S.; Li, Z.; Howarth, A.J.; Farha, O.K.; Hupp, J.T. Postsynthetic tuning of metal–organic frameworks for targeted applications. Acc. Chem. Res. 2017, 50, 805–813. [Google Scholar] [CrossRef] [PubMed]
  17. Biradha, K.; Sarkar, M.; Rajput, L. Crystal engineering of coordination polymers using 4,4′-bipyridine as a bond between transition metal atoms. Chem. Commun. 2006, 4169–4179. [Google Scholar] [CrossRef] [PubMed]
  18. Bhattacharya, B.; Ghoshal, D. Selective carbon dioxide adsorption by mixed-ligand porous coordination polymers. CrystEngComm 2015, 17, 8388–8413. [Google Scholar] [CrossRef]
  19. Haldar, R.; Maji, T.K. Metal-organic frameworks (MOFs) based on mixed linker systems: Structural diversities towards functional materials. CrystEngComm 2013, 15, 9276–9295. [Google Scholar] [CrossRef]
  20. Bae, Y.-S.; Mulfort, K.L.; Frost, H.; Ryan, P.; Punnathanam, S.; Broadbelt, L.J.; Hupp, J.T.; Snurr, R.Q. Separation of CO2 from CH4 using mixed-ligand metal–organic frameworks. Langmuir 2008, 24, 8592–8598. [Google Scholar] [CrossRef] [PubMed]
  21. Henke, S.; Schneemann, A.; Wütscher, A.; Fischer, R.A. Directing the breathing behavior of pillared-layered metal–organic frameworks via a systematic library of functionalized linkers bearing flexible substituents. J. Am. Chem. Soc. 2012, 134, 9464–9474. [Google Scholar] [CrossRef] [PubMed]
  22. Glomb, S.; Woschko, D.; Makhloufi, G.; Janiak, C. Metal–organic frameworks with internal urea-functionalized dicarboxylate linkers for SO2 and NH3 adsorption. ACS Appl. Mater. Interfaces 2017, 9, 37419–37434. [Google Scholar] [CrossRef] [PubMed]
  23. Takashima, Y.; Martínez, V.M.; Furukawa, S.; Kondo, M.; Shimomura, S.; Uehara, H.; Nakahama, M.; Sugimoto, K.; Kitagawa, S. Molecular decoding using luminescence from an entangled porous framework. Nat. Commun. 2011, 2, 168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Horike, S.; Tanaka, D.; Nakagawa, K.; Kitagawa, S. Selective guest sorption in an interdigitated porous framework with hydrophobic pore surfaces. Chem. Commun. 2007, 3395–3397. [Google Scholar] [CrossRef]
  25. Tanaka, D.; Nakagawa, K.; Higuchi, M.; Horike, S.; Kubota, Y.; Kobayashi, T.C.; Takata, M.; Kitagawa, S. Kinetic gate-opening process in a flexible porous coordination polymer. Angew. Chem. Int. Ed. 2008, 47, 3914–3918. [Google Scholar] [CrossRef] [PubMed]
  26. Fukushima, T.; Horike, S.; Inubushi, Y.; Nakagawa, K.; Kubota, Y.; Takata, M.; Kitagawa, S. Solid solutions of soft porous coordination polymers: Fine-tuning of gas adsorption properties. Angew. Chem. 2010, 122, 4930–4934. [Google Scholar] [CrossRef]
  27. Nakagawa, K.; Tanaka, D.; Horike, S.; Shimomura, S.; Higuchi, M.; Kitagawa, S. Enhanced selectivity of CO2 from a ternary gas mixture in an interdigitated porous framework. Chem. Commun. 2010, 46, 4258–4260. [Google Scholar] [CrossRef] [PubMed]
  28. Hijikata, Y.; Horike, S.; Sugimoto, M.; Sato, H.; Matsuda, R.; Kitagawa, S. Relationship between channel and sorption properties in coordination polymers with interdigitated structures. Chem. Eur. J. 2011, 17, 5138–5144. [Google Scholar] [CrossRef] [PubMed]
  29. Kishida, K.; Horike, S.; Nakagawa, K.; Kitagawa, S. Synthesis and adsorption properties of azulene-containing porous interdigitated framework. Chem. Lett. 2012, 41, 425–426. [Google Scholar] [CrossRef]
  30. Johnson, J.R.; Ketcham, R. Thiazolothiazoles. I. The reaction of aromatic aldehydes with dithiooxamide. J. Am. Chem. Soc. 1960, 82, 2719–2724. [Google Scholar] [CrossRef]
  31. Bevk, D.; Marin, L.; Lutsen, L.; Vanderzande, D.; Maes, W. Thiazolo[5,4-d]thiazoles - promising building blocks in the synthesis of semiconductors for plastic electronics. RSC Adv. 2013, 3, 11418–11431. [Google Scholar] [CrossRef]
  32. Reginato, G.; Mordini, A.; Zani, L.; Calamante, M.; Dessi, A. Photoactive compounds based on the thiazolo[5,4-d]thiazole core and their application in organic and hybrid photovoltaics. Eur. J. Org. Chem. 2016, 233–251. [Google Scholar] [CrossRef]
  33. Zampese, J.A.; Keene, F.R.; Steel, P.J. Diastereoisomeric dinuclear ruthenium complexes of 2,5-di(2-pyridyl)thiazolo[5,4-d]thiazole. Dalton Trans. 2004, 4124–4129. [Google Scholar] [CrossRef] [PubMed]
  34. Falcão, E.H.L.; Naraso; Feller, R.K.; Wu, G.; Wudl, F.; Cheetham, A.K. Hybrid organic–inorganic framework structures: Influence of cation size on metal–oxygen–metal connectivity in the alkaline earth thiazolothiazoledicarboxylates. Inorg. Chem. 2008, 47, 8336–8342. [Google Scholar]
  35. Aprea, A.; Colombo, V.; Galli, S.; Masciocchi, N.; Maspero, A.; Palmisano, G. Thiazolo[5,4-d]thiazole-2,5-dicarboxylic acid, C6H2N2O4S2, and its coordination polymers. Solid State Sci. 2010, 12, 795–802. [Google Scholar] [CrossRef]
  36. Rizzuto, F.J.; Faust, T.B.; Chan, B.; Hua, C.; D’Alessandro, D.M.; Kepert, C.J. Experimental and computational studies of a multi-electron donor–acceptor ligand containing the thiazolo[5,4-d]thiazole core and its incorporation into a metal–organic framework. Chem. Eur. J. 2014, 20, 17597–17605. [Google Scholar] [CrossRef] [PubMed]
  37. Hua, C.; Rizzuto, F.J.; Zhang, X.; Tuna, F.; Collison, D.; D’Alessandro, D.M. Spectroelectrochemical properties of a Ru(II) complex with a thiazolo[5,4-d]thiazole triarylamine ligand. New J. Chem. 2017, 41, 108–114. [Google Scholar] [CrossRef]
  38. Zhu, X.; Tian, C.; Jin, T.; Wang, J.; Mahurin, S.M.; Mei, W.; Xiong, Y.; Hu, J.; Feng, X.; Liu, H. Thiazolothiazole-linked porous organic polymers. Chem. Commun. 2014, 50, 15055–15058. [Google Scholar] [CrossRef] [PubMed]
  39. APEX2. SAINT, Data Reduction and Frame Integration Program for the CCD Area-Detector System, Bruker Analytical X-ray Systems; Data Collection Program for the CCD Area-Detector System: Madison, WI, USA, 1997–2006.
  40. Sheldrick, G.M. SADABS: Area-Detector Absorption Correction; University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  41. Sheldrick, G.M. Crystal structure refinement with SHElXL. Acta Crystallogr. Sect A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  42. Sheldrick, G. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  43. Brandenburg, K. DIAMOND, version 4.4; Crystal and Molecular Structure Visualization; Crystal Impact—K. Brandenburg & H. Putz Gbr: Bonn, Germany, 2009–2017.
  44. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. Sect. D—Biol. Crystallogr. 2009, 65, 148–155. [Google Scholar] [CrossRef] [PubMed]
  45. Knighton, R.C.; Hallett, A.J.; Kariuki, B.M.; Pope, S.J.A. A one-step synthesis towards new ligands based on aryl-functionalised thiazolo[5,4-d]thiazole chromophores. Tetrahedron Lett. 2010, 51, 5419–5422. [Google Scholar] [CrossRef]
  46. Hisamatsu, S.; Masu, H.; Azumaya, I.; Takahashi, M.; Kishikawa, K.; Kohmoto, S. U-shaped aromatic ureadicarboxylic acids as versatile building blocks: Construction of ladder and zigzag networks and channels. Cryst. Growth Des. 2011, 11, 5387–5395. [Google Scholar] [CrossRef]
  47. Das, A.; D’Alessandro, D.M. Tuning the functional sites in metal-organic frameworks to modulate CO2 heats of adsorption. CrystEngComm 2015, 17, 706–718. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Wu, H.; Emge, T.J.; Gong, Q.; Nijem, N.; Chabal, Y.J.; Kong, L.; Langreth, D.C.; Liu, H.; Zeng, H.; Li, J. Enhancing gas adsorption and separation capacity through ligand functionalization of microporous metal–organic framework structures. Chem. Eur. J. 2011, 17, 5101–5109. [Google Scholar] [CrossRef] [PubMed]
  49. Liang, Z.; Marshall, M.; Chaffee, A.L. CO2 adsorption-based separation by metal organic framework (Cu-BTC) versus zeolite (13X). Energy Fuels 2009, 23, 2785–2789. [Google Scholar] [CrossRef]
  50. Caskey, S.R.; Wong-Foy, A.G.; Matzger, A.J. Dramatic tuning of carbon dioxide uptake via metal substitution in a coordination polymer with cylindrical pores. J. Am. Chem. Soc. 2008, 130, 10870–10871. [Google Scholar] [CrossRef] [PubMed]
  51. Llewellyn, P.L.; Bourrelly, S.; Serre, C.; Vimont, A.; Daturi, M.; Hamon, L.; De Weireld, G.; Chang, J.-S.; Hong, D.-Y.; Kyu Hwang, Y.; et al. High uptakes of CO2 and CH4 in mesoporous metal-organic frameworks MIL-100 and MIL-101. Langmuir 2008, 24, 7245–7250. [Google Scholar] [CrossRef] [PubMed]
  52. McDonald, T.M.; D’Alessandro, D.M.; Krishna, R.; Long, J.R. Enhanced carbon dioxide capture upon incorporation of N,N′-dimethylethylenediamine in the metal-organic framework CuBTTri. Chem. Sci. 2011, 2, 2022–2028. [Google Scholar] [CrossRef]
  53. Bourrelly, S.; Llewellyn, P.L.; Serre, C.; Millange, F.; Loiseau, T.; Férey, G. Different adsorption behaviors of methane and carbon dioxide in the isotypic nanoporous metal terephthalates MIL-53 and MIL-47. J. Am. Chem. Soc. 2005, 127, 13519–13521. [Google Scholar] [CrossRef] [PubMed]
  54. Pera-Titus, M.; Farrusseng, D. Guest-induced gate opening and breathing phenomena in soft porous crystals: Building thermodynamically consistent isotherms. J. Phys. Chem. C 2012, 116, 1638–1649. [Google Scholar] [CrossRef]
Figure 1. Examples of tztz-containing ligands.
Figure 1. Examples of tztz-containing ligands.
Crystals 08 00030 g001
Scheme 1. Reaction scheme for the synthesis of 2,5-di(4-pyridyl)thiazolo[5,4-d]thiazole from 4-pyridinecarboxaldehyde and dithioaxamide.
Scheme 1. Reaction scheme for the synthesis of 2,5-di(4-pyridyl)thiazolo[5,4-d]thiazole from 4-pyridinecarboxaldehyde and dithioaxamide.
Crystals 08 00030 sch001
Figure 2. PXRD pattern of [Zn(1,3-BDC)Dptztz]·DMF (simulated (red), as-synthesized (black)).
Figure 2. PXRD pattern of [Zn(1,3-BDC)Dptztz]·DMF (simulated (red), as-synthesized (black)).
Crystals 08 00030 g002
Figure 3. Molecular structure of Dptztz (50% thermal ellipsoids, symmetry transformation 1 − x, 2 − y, 2 − z).
Figure 3. Molecular structure of Dptztz (50% thermal ellipsoids, symmetry transformation 1 − x, 2 − y, 2 − z).
Crystals 08 00030 g003
Figure 4. Extended asymmetric unit of [Zn(1,3-BDC)Dptztz]·DMF (50% thermal ellipsoids; symmetry transformations: i = −x + 2, −y, −z + 1; ii = x + 1, y, z + 1; iii = −x + 2, −y + 1, −z + 1; iv = x − 1, y, z − 1; v = −x + 1, −y, −z, vi = x, y − 1, z). For the slight ring-flip disorder of the thiazolothiazol moiety and the DMF solvent molecule, which is omitted here for clarity, see Figure S2 in Supplementary Material. See Table 2 for selected bond lengths and angles.
Figure 4. Extended asymmetric unit of [Zn(1,3-BDC)Dptztz]·DMF (50% thermal ellipsoids; symmetry transformations: i = −x + 2, −y, −z + 1; ii = x + 1, y, z + 1; iii = −x + 2, −y + 1, −z + 1; iv = x − 1, y, z − 1; v = −x + 1, −y, −z, vi = x, y − 1, z). For the slight ring-flip disorder of the thiazolothiazol moiety and the DMF solvent molecule, which is omitted here for clarity, see Figure S2 in Supplementary Material. See Table 2 for selected bond lengths and angles.
Crystals 08 00030 g004
Figure 5. (a) 1D double strand of Zn2+ and 1,3-BDC2− along the b-axis and (b) 2D coordination network structure in the bc plane and (c) supramolecular 3D packing of the 2D layers in [Zn(1,3-BDC)Dptztz]·DMF (H atoms in (a,b) and DMF solvent molecules are not shown for clarity). In (c) the 2D layers are colored alternately black and yellow for clarity.
Figure 5. (a) 1D double strand of Zn2+ and 1,3-BDC2− along the b-axis and (b) 2D coordination network structure in the bc plane and (c) supramolecular 3D packing of the 2D layers in [Zn(1,3-BDC)Dptztz]·DMF (H atoms in (a,b) and DMF solvent molecules are not shown for clarity). In (c) the 2D layers are colored alternately black and yellow for clarity.
Crystals 08 00030 g005aCrystals 08 00030 g005b
Figure 6. TGA curve of [Zn(1,3-BDC)Dptztz]·DMF·in the temperature range 26–600 °C with a heating rate of 10 K/min under N2 atmosphere.
Figure 6. TGA curve of [Zn(1,3-BDC)Dptztz]·DMF·in the temperature range 26–600 °C with a heating rate of 10 K/min under N2 atmosphere.
Crystals 08 00030 g006
Figure 7. CO2 adsorption (closed symbols) and desorption (open symbols) isotherms for activated [Zn(1,3-BDC)Dptztz] measured at 195 K (black), 293 K (red) and 273 K (blue).
Figure 7. CO2 adsorption (closed symbols) and desorption (open symbols) isotherms for activated [Zn(1,3-BDC)Dptztz] measured at 195 K (black), 293 K (red) and 273 K (blue).
Crystals 08 00030 g007
Figure 8. Heat of adsorption plot of CO2 adsorption for [Zn(1,3-BDC)Dptztz].
Figure 8. Heat of adsorption plot of CO2 adsorption for [Zn(1,3-BDC)Dptztz].
Crystals 08 00030 g008
Figure 9. Water sorption isotherms at 293 K for [Zn(1,3-BDC)Dptztz].
Figure 9. Water sorption isotherms at 293 K for [Zn(1,3-BDC)Dptztz].
Crystals 08 00030 g009
Table 1. Crystal data and refinement details.
Table 1. Crystal data and refinement details.
Dptztz[Zn(1,3 BDC)Dptztz]·DMF
Chemical formulaC14H8N4S2C22H12N4O4S2Zn·C3H7NO
Mr296.36598.94
Crystal system, space groupMonoclinic, P21/cTriclinic, P 1 ¯
Temperature (K)100100
a (Å)8.3873 (5)9.1388 (6)
b (Å)6.3140 (3)10.0354 (7)
c (Å)11.7170 (6)14.2804 (11)
α (°)9088.417 (4)
β (°)93.699 (3)88.236 (5)
γ (°)9075.636 (4)
V (Å3)619.21 (6)1267.86 (16)
Z22
μ (mm−1)0.4231.181
Crystal size (mm)0.10 × 0.05 × 0.050.10 × 0.05 × 0.01
Absorption correctionMulti-scan, wR2(int) was 0.1649 before and 0.0771 after correction. The Ratio of minimum to maximum transmission is 0.8473. The λ/2 correction factor is 0.0015.Multi-scan, wR2(int) was 0.1533 before and 0.0488 after correction. The Ratio of minimum to maximum transmission is 0.9318. The λ/2 correction factor is 0.0015.
Tmin, Tmax0.6330, 0.74710.6951, 0.7460
No. of measured, independent and observed reflections6837, 965, 847 [I > 2σ(I)]17151, 4743, 3696 [I > 2σ(I)]
Rint0.0490.045
(sin θ/λ)max (Å−1)0.6390.612
R, wR(F2), S [F2 > 2σ (F2)]
R, wR(F2), S [all data]
0.0284, 0.0675, 1.067
0.359, 0.0699, 1.067
0.0400, 0.0849, 1.055
0.0609, 0.0916, 1.055
No. of reflections9654743
No. of parameters91364
Δρmax, Δρmin (e·Å−3)0.238, -0.1820.645, –0.581
Table 2. Selected bond lengths and angles (Å, °) in [Zn(1,3-BDC)Dptztz].
Table 2. Selected bond lengths and angles (Å, °) in [Zn(1,3-BDC)Dptztz].
Zn–O12.0532 (18)Zn–O4iii2.2269 (19)
Zn–O2i2.0218 (18)Zn–N12.166 (3)
Zn–O3iii2.1569 (19)Zn–N4ii2.151 (3)
O1–Zn–O2i119.30 (7)O2i–Zn–N4ii89.34 (9)
O1–Zn–O3iii88.78 (7)O3iii–Zn–O4iii60.01 (7)
O1–Zn–O4iii148.39 (7)O3iii–Zn–N190.18 (9)
O1–Zn–N189.84 (9)O3iii–Zn–N4ii90.56 (9)
O1–Zn‒N4ii86.00 (9)O4iii–Zn–N194.77 (8)
O2i–Zn–O3iii151.83 (7)O4iii–Zn–N4ii89.21 (8)
O2i–Zn–O4iii91.82 (7)N1–Zn–N4ii175.76 (8)
O2i–Zn–N191.99 (9)
Symmetry transformations: i = −x + 2, −y, −z + 1; ii = x + 1, y, z + 1; iii = −x + 2, −y + 1, −z + 1.
Table 3. CO2 sorption data for [Zn(1,3-BDC)Dptztz].
Table 3. CO2 sorption data for [Zn(1,3-BDC)Dptztz].
Quantity Adsorbed (cm3/g, mmol/g, wt %)Total Pore Volume (cm3/g)
195 K138, 6.16, 27.1%0.246 1
273 K51.9, 2.32, 10.2%0.092 2
293 K35.5, 1.59, 7.0%0.061 3
1 at p/p0 = 0.50, 2 at p/p0 = 0.03, 3 at p/p0 = 0.017.

Share and Cite

MDPI and ACS Style

Millan, S.; Makhloufi, G.; Janiak, C. Incorporating the Thiazolo[5,4-d]thiazole Unit into a Coordination Polymer with Interdigitated Structure. Crystals 2018, 8, 30. https://doi.org/10.3390/cryst8010030

AMA Style

Millan S, Makhloufi G, Janiak C. Incorporating the Thiazolo[5,4-d]thiazole Unit into a Coordination Polymer with Interdigitated Structure. Crystals. 2018; 8(1):30. https://doi.org/10.3390/cryst8010030

Chicago/Turabian Style

Millan, Simon, Gamall Makhloufi, and Christoph Janiak. 2018. "Incorporating the Thiazolo[5,4-d]thiazole Unit into a Coordination Polymer with Interdigitated Structure" Crystals 8, no. 1: 30. https://doi.org/10.3390/cryst8010030

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