Next Article in Journal
Atom Economical Multi-Substituted Pyrrole Synthesis from Aziridine
Previous Article in Journal
Wearable Capacitive Pressure Sensor for Contact and Non-Contact Sensing and Pulse Waveform Monitoring
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Selective CO2 Adsorption and Photothermal Conversion Study of an Azo-Based Cobalt-MOF Material

1
Henan Key Laboratory of Function-Oriented Porous Materials, College of Chemistry and Chemical Engineering, Luoyang Normal University, Luoyang 471934, China
2
Hubei Nuclear Industry Geological Bureau, Xiaogan 432003, China
3
School of Materials Science and Engineering, Zhengzhou University, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2022, 27(20), 6873; https://doi.org/10.3390/molecules27206873
Submission received: 16 September 2022 / Revised: 11 October 2022 / Accepted: 11 October 2022 / Published: 13 October 2022
(This article belongs to the Section Inorganic Chemistry)

Abstract

:
A new metal–organic framework (MOF), [Co2(L)2(azpy)]n (compound 1, H2L = 5-(pyridin-4-ylmethoxy)-isophthalic acid, azpy = 4,4′-azopyridine), was synthesized by a solvothermal method and further characterized by elemental analysis, IR spectra, thermogravimetric analysis, single-crystal and powder X-ray diffraction. The X-ray single-crystal diffraction analysis for compound 1 indicated that two cis L22− ligands connected to two cobalt atoms resulted in a macrocycle structure. Through a series of adsorption tests, we found that compound 1 exhibited a high capacity of CO2, and the adsorption capacity could reach 30.04 cm3/g. More interestingly, under 273 K conditions, the adsorption of CO2 was 41.33 cm3/g. In addition, when the Co-MOF was irradiated by a 730 nm laser, rapid temperature increases for compound 1 were observed (temperature variation in 169 s: 26.6 °C), showing an obvious photothermal conversion performance. The photothermal conversion efficiency reached 20.3%, which might be due to the fact that the parallel arrangement of azo units inhibited non-radiative transition and promoted photothermal conversion. The study provides an efficient strategy for designing MOFs for the adsorption of CO2 and with good photothermal conversion performance.

Graphical Abstract

1. Introduction

The emission of CO2 will cause a rise in the global average temperature, which will lead to a series of ecological and environmental problems [1,2,3,4]. At present, there are two feasible methods for reducing CO2 concentrations. One is to develop new sources of energy that can be recycled, and the other is to use CO2 as a starting material to convert it into useful chemical products. However, disadvantages such as high cost and high energy consumption are the challenges in solving the above problems. Researchers have been trying to explore new recyclable adsorption materials for the adsorption and separation of CO2. Metal–organic frameworks (MOFs) as functional materials, due to their ease of design, diverse structure, chemical modifiability, high porosity and specific surface area, have been widely investigated in various fields such as fluorescent sensing [5,6,7,8,9,10,11,12], heterogeneous catalysis [13,14,15,16,17,18], drug delivery [19,20,21,22,23,24], gas storage and separation [25,26,27,28,29,30,31,32]. Modifying the metal sites and organic ligands is conducive to improving the adsorption and separation ability of the materials [33,34,35,36,37,38].
To date, numerous MOFs have been designed for the adsorption and separation of CO2 [39,40,41,42,43,44,45]. Luo et al. made many outstanding efforts in the field of gas adsorption [44,45,46,47,48,49]. For example, they designed a porous DMOF based on a photosensitive diarylethene ligand, and the photochromic behavior was controlled by UV irradiation with visible light for the first time [50], with up to a 75% CO2-desorption capacity. Additionally, they constructed another metal–organic framework material with a highly photosensitive structure by substituting diarylethene ligands for azo ligands, ECUT-15 [47]. Under UV irradiation, it still performed its general CO2 adsorption capability and had a highly selective adsorption of CO2 over N2, CH4, O2 and CO. At instantaneous conditions, the CO2 capture/release performance was up to 78%. In addition, M-MOF-74 series were reported by Matzger A. J. et al. and Long J. R. et al., respectively [51,52]. The results showed that M-MOF-74 had a highly selective adsorption capacity for CO2 under low pressure conditions due to the strong coordination between the unsaturated metal ions and CO2 in the structure There is an increasing number of MOFs supplied for small molecule gas adsorption and purification. However, the development of this research remains quite slow.
In addition, photothermal therapy has been demonstrated to have potential for cancer treatment because of the forceful killing of diseased lesions with high selectivity under near-infrared laser irradiation [53,54]. Except for purely inorganic materials or purely organic materials displaying the function of converting NIR light to heat, MOFs have also shown effective photothermal conversion capacity, which has been demonstrated by various ruthenium MOFs and Zn-MOFs based on π–π stacking interactions [55,56]. Although a variety of MOF complexes with near-infrared photothermal conversion properties have been achieved, reports about the photothermal effects induced by conjugation effects based on azo bonds are rare.
Herein, we report a new cobalt-based MOF, compound 1, formulated as [Co2(L)2(azpy)]n (H2L = 5-(pyridin-4-ylmethoxy)-isophthalic acid, azpy = 4,4′-azopyridine), which was synthesized under solvothermal conditions, and its crystal structure was determined by single-crystal X-ray diffraction (Scheme 1). Compound 1 was characterized by the IR spectrum, elemental analysis, PXRD, solid-state UV–Vis absorption and thermogravimetric analysis. In addition, compound 1 exhibited a high selectivity for CO2 under 273 K conditions. The adsorption capacity could reach 41.33 cm3/g. In addition, obvious photothermal conversion performance was observed under 730 nm laser irradiation accompanied by a high conversion efficiency of 20.3% (temperature change: 53.6 °C).

2. Results

2.1. Crystal Structural Description

The single-crystal X-ray diffraction analysis indicated that compound 1 crystallized in a monoclinic P21/c space group. In the asymmetric unit (Figure 1), compound 1 contained two crystallographically independent Co(II) ions, namely Co1 and Co2 ions, and each Co2+ ion had a hexahedral coordination environment with four carboxylic oxygen atoms from three different H2L ligands, one pyridine nitrogen atom from the azpy ligand and one pyridine nitrogen atom from the L2 ligand in the opposite direction. Careful analyzing of this structure showed that a methylene group serving as a flexible site was present in the H2L ligand, which caused the formation of a cage-type unit, in which the separation between two phenyl units was 6.92 Å. Furthermore, these cage-type units were bridged by multiple azpy ligands, finally resulting in the generation of a novel three-dimensional framework structure. Notably, a very large pore structure was observed in this 3D framework, showing a 6.3 Å × 11.7 Å window. Platon calculated a solvent-accessible void space of 2042.1 Å3 (equal to 40.6% of the cell volume), showing the potential porous MOFs of compound 1. Moreover, the azpy ligands were arranged in parallel, in which the distance between two azo units was 3.42 Å, reflecting obvious π–π stacking interaction between the azpy ligands. In addition, an eight-membered ring was found based on two cis H2L ligands by the connection of two cobalt atoms (Figure 2). Additionally, the bond lengths were Co1 − O1 = 2.088(2) Å, Co1 − O2 = 2.300(19) Å, Co1 − O5 = 1.998(19) Å, Co1 − O7 = 2.035(18) Å, Co1 − N5 = 2.182(3) Å and Co1 − N6 = 2.148(3) Å. Furthermore, the adjacent non-bonding distance of Co1–Co2 was 4.35 Å.

2.2. PXRD and Thermal Properties

The simulated and experimental PXRD patterns of compound 1 are shown in Figure 3. Their peak positions corresponded well with each other, indicating the phase purity of the solids, although some small shifts were observed, which might be attributed to the cis–trans isomerism of the azo bonds [57,58].
In addition, the thermal stability of compound 1 was also investigated under a nitrogen atmosphere in the temperature range of 20–800 °C. From the TG curve (Figure 4), we found that the first weight loss occurred from 20 to 183 °C, which corresponded to the loss of the free water and DMF molecules. Additionally, there was almost no weight loss from 183 to 221 °C. Upon further heating above 221 °C, the successive losses corresponded to the collapse of the skeleton (352 °C) and the decomposition of the organic ligands (383 °C). In addition, the solid-state UV–Vis absorption of compound 1 was also explored carefully from 250 to 800 nm with a UV-vis spectrophotometer. The result clearly reflected a strong absorption from 300 to 550 nm. A careful structural analysis of this compound showed that the strong and wide-range UV-Vis absorption might be related to the absorption of azpy ligands due to the fact that the azo bond is a good chromophore, which can effectively absorb photoelectrons (Figure 4a) [9].

2.3. Adsorption Properties

The high chemical stability and thermal stability of compound 1 prompted us to study the gas adsorption performances. As shown in Figure 5a,b, the porosity of the activated materials of compound 1 was manifested by N2 adsorption at 77 K, expressing a fully reversible type-I isotherm, a signature characteristic of microporous materials, and giving a BET of 402.6 m2/g, a uniform pore size of 1.0 nm and a total pore volume of 0.21 cm3/g. Furthermore, this proper porosity and desirable aperture encouraged us to investigate the adsorption selectivity of CO2/CH4/CO in detail. At 298 K, the adsorption isotherms of the single components CO2, CH4 and CO were collected, as presented in Figure 6a. Compound 1 showed a unique adsorption capacity for CO2 compared to CH4 and CO. Furthermore, the adsorption amount of CO2 (30.04 cm3/g) was higher than that of CH4 (8.71 cm3/g) and CO (9.12 cm3/g). Finally, we further explored the adsorption capacity of compound 1 of CO2 under the influence of different temperatures (Figure 6b). The experimental results showed that the adsorption value could reach 41.33 cm3/g at 273 K. All the above experimental studies indicated that compound 1 could be used as a potential adsorbing material for CO2 in the future (Table 1).

2.4. Photothermal Conversion Properties

To our knowledge, π–π stacking interactions can trigger active non-radiative pathways and the inhibition of the radiative transition process, leading to an effective photothermal (PT) conversion phenomenon [59,60,61,62,63]. Furthermore, all the UV–vis spectra of compound 1 and two ligands displayed obvious near infrared absorption at 730 nm (Figure 4a). In addition, the TGA of compound 1 proved that compound 1 was able to maintain good structural stability until 221 °C, telling us that the temperature rise induced by the photo-thermal conversion was probably not the cause of the change in the structural framework (Figure 4b). Thus, the near-infrared photothermal conversion performance of compound 1 was performed earnestly under irradiation from a 730 nm near-infrared laser. In contrast, the temperature change experiments of the ligands H2L and azpy were also explored carefully under the same conditions as that of compound 1.
Firstly, crystalline compound 1 was irradiated at a power of 0.6 W/cm2 under a 730 nm laser accompanied by the conditions of the external temperature being stable. Interestingly, a rapid temperature rise was observed. The temperature rose to 75.0 °C from 21.4 °C in only 109 s. Once the 730 nm near-infrared laser was turned off, the temperature declining process was seen clearly by the infrared imaging camera, reflecting the good photothermal conversion properties of compound 1 (Figure 7). Thus, the near-infrared photothermal conversion efficiency of compound 1 was calculated based on the reduced temperature data according to the calculation method described below (Equations (1)–(4)), showing a high conversion efficiency of 20.3%, reflecting the excellent near-infrared photothermal performance of Co−MOF, i.e., compound 1. In contrast, illumination experiments of the ligands H2L and azpy under a 730 nm laser were also carried out. The results showed that the temperature increase in the ligands H2L and azpy was extremely weak and the temperature changes were 4.5 °C (from 22.1 to 26.6 °C) and 3.0 °C (from 22.4 to 25.4 °C), which was negligible (Figure 8). A careful analysis of structure 1 showed that the π–π stacking interaction between the azpy ligands could trigger non-radiative transition and induce obvious photothermal conversion effects.
Based on the total energy balance of the system:
i m i c p , i d T d t = Q s Q loss
η = hS ΔTmax/I (1 − 10−A)
θ = (T − Tsurr)/(Tmax − Tsurr)
In Equations (1)–(3) above, mi (0.03 g) is the sample mass, cp,i (1.926 J (g°C)−1) is the heat capacity of the system components and A is the absorbance of the samples at a wavelength of 730 nm (0.634). Qs is the photothermal energy input by the near-infrared laser irradiation sample and Qloss is the heat energy lost to the surrounding environment. I is the laser power (0.6 W cm−2). When the temperature reaches its maximum, the system is in equilibrium.
Qs = Qloss = hS ΔTmax
where h is the heat transfer coefficient and S is the surface area of the container. I is the laser power (0.6 W cm−2).
τ s = i m i c p , i / hS , thus d θ d t = Q s τ s hS Δ Tmax θ τ s , and when the laser is off, Qs = 0, τs = −t/lnθ. The value of τs was obtained from the cooling curve of sample. The photothermal conversion efficiency η was 20.3%.

3. Experiment

3.1. Materials and Methods

All reagents and solvents were analytical grade and were obtained from commercial sources and used without further purification. Elemental analyses (C, H and N) were performed on a Vario EL III elemental analyzer. The infrared spectrum was measured with KBr pellets in the 4000–400 cm−1 region on a Nicolet 170SX spectrometer. Thermogravimetric analyses were performed on an SDT 2960 thermal analyzer from room temperature to 800 °C at a heating rate of 20 K min−1 under nitrogen flow. Powder X-ray diffraction (PXRD) data were collected on a Rigaku D/Max-2500PC diffractometer with Cu/Kα radiation (λ = 1.5406 Å) over a 2θ range of 5−50° with a scan speed of 5°/min at room temperature. CO2 (99.99%), N2 (99.99%), CH4 (99.99%) and CO (99.99%) gas adsorption measurements were performed using an ASAP 2020 system. UV–visible absorption spectra were recorded on a Lambda 750 s UV–vis spectrophotometer.

3.2. Synthesis of Compound 1

A mixture of H2L (27 mg, 0.10 mmol), azpy (18 mg, 0.10 mmol) and Co(NO3)2·6H2O (29 mg, 0.10 mmol) were dissolved in DMF (6 mL) and H2O (2 mL) in a screw-capped vial. The vial was capped and heated at 100 °C for 72 h. Orange block crystals of 1 were obtained in high yields (92%) based on the H2L ligand. Elemental analysis was calculated (%) for 1 C38H26Co2N6O10 (844.52): C, 54.04; H, 3.10; N, 9.95. It was found as: C, 54.07; H, 3.11; N, 9.50. The IR for 1 was: 3397 m, br, 3075 m, 2929 m, 2353 w, 1666 m, 1569 s, 1394 s, 1250 m, 1117 w, 1057 w, 920 vs, 845 w, 782 m, 727 m, 673 vs, 622 vs, 567 s, 528 vs, 479 vs.

3.3. Single-Crystal X-ray Structure Determination

The single-crystal X-ray diffraction analysis of 1 was carried out on a Rigaku Saturn 724 CCD diffractomer (Mo-Kα, λ = 0.71073 Å) at room temperature. The structure was solved by a direct method with SHELXS-97 [64] and refined by the full-matrix least-squares method on F2 with anisotropic displacement parameters for all non-H atoms (SHELXL-97) [65]. An empirical absorption correction was applied by the SADABS program [66]. The hydrogen atoms were assigned with common isotropic displacement factors and included in the final refinement with the use of geometrical restraints. The crystallographic data and selected bond lengths and angles for 1 are listed in Table 2.
Crystallographic data for the structural analysis were deposited at the Cambridge Crystallographic Data Center, and CCDC No. 2,178,932 for compound 1 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 10 September 2022).

4. Conclusions

To conclude, we indeed constructed a new cobalt-based MOF [Co2(L)2(azpy)]n (H2L = 5-(pyridin-4-ylmethoxy)-isophthalic acid, azpy = 4,4′-azopyridine) by a solvothermal method. Notably, compound 1 had relative high CO2 adsorption capacities, namely 30.04 cm3/g and 41.33 cm3/g at 298 K and 273 K, respectively. However, the adsorption capacity of other gases was very weak (CH4: 8.71 cm3/g and CO: 9.12 cm3/g). Interestingly, compound 1 showed a rapid temperature change from 20.4 to 47.0 °C under a 730 nm laser radiation and the photothermal conversion efficiency could reach 20.6%, reflecting the good photothermal conversion properties of compound 1. This study will promote the development of new photothermal conversion materials and gas adsorption materials.

Author Contributions

L.-L.D. and D.-X.Z.: synthesis of the title complex, formal analysis, IR-related experiment. L.-L.D.: writing—review and editing. T.-T.Z. and T.C.: completed cultivation and testing of single crystals. J.-L.S. and J.-Z.Z.: completed formal analysis and gas adsorption. X.-Y.L. and M.-Y.L.: completed UV–Vis absorption spectra and photothermal conversion experiment. S.-R.L.: supervision, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Henan Province (No. 212300410209), Key Scientific Research Projects of Higher Education of He’nan Province (Grant 22A150016), the Shanghai Science and Technology Committee (19DZ2270100) and Heluo Young Talent Lifting Project (2022HLTJZC09).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fan, Y.L.; Zhang, H.P.; Yin, M.J.; Krishna, R.; Feng, X.F.; Wang, L.; Luo, M.B.; Luo, F. High adsorption capacity and selectivity of SO2 over CO2 in a metal–organic framework. Inorg. Chem. 2021, 60, 4–8. [Google Scholar] [CrossRef] [PubMed]
  2. Gu, S.F.; Xiong, X.H.; Gong, L.L.; Zhang, H.P.; Xu, Y.; Feng, X.F.; Luo, F. Classified encapsulation of an organic dye and metal–organic complex in different molecular compartments for white-light emission and selective adsorption of C2H2 over CO2. Inorg. Chem. 2021, 60, 8211–8217. [Google Scholar] [CrossRef] [PubMed]
  3. Belmabkhout, Y.; Bhatt, P.M.; Adil, K.; Pillai, R.S.; Cadiau, A.; Shkurenko, A.; Maurin, G.; Liu, G.; Koros, W.J.; Eddaoudi, M. Natural gas upgrading using a fluorinated MOF with tuned H2S and CO2 adsorption selectivity. Nat. Energy 2018, 3, 1059–1066. [Google Scholar] [CrossRef]
  4. Qin, J.H.; Xiao, Z.; Xu, P.; Li, Z.H.; Lu, X.Y.; Yang, X.G.; Lu, W.W.; Ma, L.F.; Li, D.S. Anionic porous Zn-Metalated porphyrin metal–organic framework with PtS topology for gas-phase photocatalytic CO2 reduction. Inorg. Chem. 2022, 61, 13234–13238. [Google Scholar] [CrossRef] [PubMed]
  5. Sun, T.Y.; Gao, Y.B.; Du, Y.Y.; Zhou, L.; Chen, X. Recent advances in developing lanthanide metal–organic frameworks for ratiometric fluorescent sensing. Front. Chem. 2020, 8, 624592. [Google Scholar] [CrossRef]
  6. 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]
  7. Li, Y.L.; Zhao, Y.; Wang, P.; Kang, Y.S.; Liu, Q.; Zhang, X.D.; Sun, W.Y. Multifunctional metal–organic frameworks with fluorescent sensing and selective adsorption properties. Inorg. Chem. 2016, 55, 11821–11830. [Google Scholar] [CrossRef]
  8. Liu, X.B.; Lin, H.; Xiao, Z.Y.; Fan, W.D.; Huang, A.; Wang, R.M.; Zhang, L.L.; Sun, D.F. Multifunctional lanthanide–organic frameworks for fluorescent sensing, gas separation and catalysis. Dalton Trans. 2016, 45, 3743–3749. [Google Scholar] [CrossRef]
  9. Li, T.; Dang, L.; Zhao, C.; Lv, Z.; Yang, X.; Zhao, Y.; Zhang, S. A self-sensitized Co (II)-MOF for efficient visible-light-driven hydrogen evolution without additional cocatalysts. J. Solid State Chem. 2021, 304, 122609–122614. [Google Scholar] [CrossRef]
  10. Qin, J.; Xu, P.; Huang, Y.; Xiao, L.; Lu, W.; Yang, X.; Ma, L.; Zang, S. High loading of Mn(II)-metalated porphyrin in MOF for photocatalytic CO2 reduction in gas−solid condition. Chem. Commun. 2021, 57, 8468–8471. [Google Scholar] [CrossRef]
  11. Qin, J.-H.; Qin, W.-J.; Xiao, Z.; Yang, J.-K.; Wang, H.-R.; Yang, X.-G.; Li, D.-S.; Ma, L.-F. Efficient energy-transfer-induced high photoelectric conversion in a dye-encapsulated ionic pyrene-based metal–organic framework. Inorg. Chem. 2021, 60, 18593–18597. [Google Scholar] [CrossRef] [PubMed]
  12. Chang, X.H.; Qin, J.H.; Han, M.L.; Ma, L.F.; Wang, L.Y. Exploring the structural diversities and magnetic properties of copper(II) and manganese (II) complexes based on 5-methoxyisophthalate and flexible bis (imidazole) ligands. CrystEngComm 2014, 16, 870–882. [Google Scholar] [CrossRef]
  13. Zhou, Z.; Wang, Y.; Peng, F.; Meng, F.; Zha, J.J.; Ma, L.; Du, Y.H.; Peng, N.; Ma, L.F.; Zhang, Q.H.; et al. Intercalation-activated layered MoO3 nanobelts as biodegradable nanozymes for tumor-specific photo-enhanced catalytic therapy. Angew. Chem. Int. Ed. 2022, 61, e202115939. [Google Scholar]
  14. Yang, X.G.; Zhai, Z.M.; Lu, X.M.; Ma, L.F.; Yan, D.P. Fast crystallization-deposition of orderly molecule level heterojunction thin films showing tunable up-conversion and ultrahigh photoelectric response. ACS Cent. Sci. 2020, 6, 1169–1178. [Google Scholar] [CrossRef]
  15. Yang, X.G.; Qin, J.H.; Huang, Y.D.; Zhai, Z.M.; Ma, L.F.; Yan, D.P. Highly enhanced UV-vis-NIR light harvesting and photoelectric conversion of pyrene MOF by encapsulation of D-π-A cyanine dye. J. Mater. Chem. C 2020, 8, 17169–17175. [Google Scholar] [CrossRef]
  16. Chang, X.H.; Zhao, Y.; Han, M.L.; Ma, L.F.; Wang, L.Y. Five Cd(II) coordination polymers based on 2,3′,5,5′-biphenyltetracarboxylic acid and N-donor coligands: Syntheses, structures and fluorescent properties. CrystEngComm 2014, 16, 6417–6424. [Google Scholar] [CrossRef]
  17. Chang, X.H.; Qin, W.J.; Zhang, X.Y.; Jin, X.; Yang, X.G.; Dou, C.X.; Ma, L.F. Angle-Dependent Polarized Emission and Photoelectron Performance of Dye-Encapsulated Metal-Organic Framework. Inorg. Chem. 2021, 60, 10109–10113. [Google Scholar] [CrossRef]
  18. Yang, X.; Zhai, Z.; Lu, X.; Qin, J.; Li, F.; Ma, L. Hexanuclear Zn(II)-Induced Dense π-Stacking in a Metal–Organic Framework Featuring Long-Lasting Room Temperature Phosphorescence. Inorg. Chem. 2020, 59, 10395–10399. [Google Scholar] [CrossRef]
  19. Yang, X.; Lu, X.; Zhai, Z.; Qin, J.; Chang, X.; Han, M.; Li, F.; Ma, L. π-Type halogen bonding enhanced the long-lasting room temperature phosphorescence of Zn(ii) coordination polymers for photoelectron response applications. Inorg. Chem. Front. 2020, 7, 2224–2230. [Google Scholar] [CrossRef]
  20. Zhang, H.P.; Gong, L.L.; Yin, M.J.; Xiong, X.H.; Zhang, Q.Y.; Feng, X.F.; Luo, F.; Carney, J.B.; Yue, Y.F. Efficient organic iodide capture by a mesoporous bimetallic-organic framework. Cell Rep. Phys. Sci. 2022, 3, 100830. [Google Scholar] [CrossRef]
  21. Yang, X.G.; Lu, X.M.; Zhai, Z.M.; Zhao, Y.; Liu, X.Y.; Ma, L.F.; Zang, S.Q. Facile synthesis of micro-scale MOF host-guest with long-last phosphorescence and enhanced optoelectronic performance. Chem. Commun. 2019, 55, 11099–11102. [Google Scholar] [CrossRef] [PubMed]
  22. Qin, J.; Zhang, H.; Sun, P.; Huang, Y.; Shen, Q.; Yang, X.; Ma, L. Ionic liquid induced highly dense assembly of porphyrin in MOF nanosheets for photodynamic therapy. Dalton Trans. 2020, 49, 17772–17778. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, Y.; Yang, X.; Lu, X.; Yang, C.; Fan, N.; Yang, Z.; Wang, L.; Ma, L. {Zn6} cluster based metal-organic framework with enhanced room-temperature phosphorescence and optoelectronic performances. Inorg Chem. 2019, 58, 6215–6221. [Google Scholar] [CrossRef] [PubMed]
  24. Dang, L.; Li, T.; Cui, Z.; Sui, D.; Ma, L.; Jin, G. Selective construction and stability studies of molecular trefoil knot and solomon link. Dalton Trans. 2021, 50, 16984–16989. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, Y.; Yang, G.P.; Zhang, P.F.; Ma, L.L.; Wang, J.M.; Li, G.P.; Wang, Y.Y. Microporous Cd(II) Metal–Organic Framework for CO2 Catalysis, Luminescent Sensing, and Absorption of Methyl Green. Cryst. Growth Des. 2021, 21, 2734–2743. [Google Scholar] [CrossRef]
  26. Liu, J.; Hang, M.; Wu, D.; Jin, J.; Cheng, J.G.; Yang, G.P.; Wang, Y.Y. Fine-tuning the porosities of the entangled isostructural Zn(II)-based metal–organic frameworks with active sites by introducing different N-auxiliary ligands: Selective gas sorption and efficient CO2 conversion. Inorg. Chem. 2020, 59, 2450–2457. [Google Scholar] [CrossRef] [PubMed]
  27. Xue, D.X.; Wang, Q.; Bai, J.F. Amide-functionalized metal–organic frameworks: Syntheses, structures and improved gas storage and separation properties. Coord. Chem. Rev. 2019, 378, 2–16. [Google Scholar] [CrossRef]
  28. Dang, L.; Li, T.; Zhao, C.; Zhang, T.; Ye, X.; Sun, X.; Wang, H.; Ma, L. Supramolecular Rh6 catalytic system promoting directed [4+4] cycloaddition reaction of anthracene under UV irradiation. J. Solid State Chem. 2022, 306, 122785–122792. [Google Scholar] [CrossRef]
  29. Li, J.; Xiong, L.; Fu, L.; Bo, W.; Du, Z.; Feng, X. Structural diversity of Mn(II) and Cu(II) complexes based on 2-carboxyphenoxyacetate linker: Syntheses, conformation comparison and magnetic properties. J. Solid State Chem. 2022, 305, 122636. [Google Scholar] [CrossRef]
  30. Li, J.; Zhang, Y.; Du, Z.; Feng, X. One-pot solvothermal synthesis of mononuclear and oxalate-bridged binuclear nickel compounds: Structural analyses, conformation alteration and magnetic properties. Inorg. Chim. Acta 2022, 530, 120697. [Google Scholar] [CrossRef]
  31. Daglar, H.; Gulbalkan, H.C.; Avci, G.; Aksu, G.O.; Altundal, O.F.; Altintas, C.; Erucar, I.; Keskin, S. Effect of metal–organic framework (MOF) database selection on the assessment of gas storage and separation potentials of mofs. Angew. Chem. Int. Ed. 2021, 60, 7828–7837. [Google Scholar] [CrossRef] [PubMed]
  32. Dhainaut, J.; Bonneau, M.; Ueoka, R.; Kanamori, K.; Furukawa, S. Formulation of metal–organic framework inks for the 3D printing of robust microporous solids toward high-pressure gas storage and separation. ACS Appl. Mater. Interfaces 2020, 12, 10983–10992. [Google Scholar] [CrossRef] [PubMed]
  33. Li, J.R.; Sculley, J.; Zhou, H.C. Metal–organic frameworks for separations. Chem. Rev. 2012, 112, 869–932. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, J.P.; Zhang, Y.B.; Lin, J.B.; Chen, X.M. Metal azolate frameworks: From crystal engineering to functional materials. Chem. Rev. 2012, 112, 1001–1033. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, H.R.; Yang, X.G.; Qin, J.H.; Ma, L.F. Long-lived room temperature phosphorescence of organic–inorganic hybrid systems. Inorg. Chem. Front. 2021, 8, 1942–1950. [Google Scholar] [CrossRef]
  36. Wang, H.; Meng, W.; Wu, J.; Ding, J.; Hou, H.; Fan, Y. Crystalline central-metal transformation in metal-organic frameworks. Coord. Chem. Rev. 2016, 307, 130–146. [Google Scholar] [CrossRef]
  37. Dang, L.L.; Li, T.T.; Zhang, T.T.; Zhao, Y.; Chen, T.; Gao, X.; Ma, L.F.; Jin, G.X. Highly selective synthesis and near-infrared photothermal conversion of metalla-Borromean ring and [2]catenane assemblies. Chem. Sci. 2022, 13, 5130–5140. [Google Scholar] [CrossRef]
  38. Wang, Y.F.; Li, S.H.; Ma, L.F.; Geng, J.L.; Wang, L.Y. Syntheses, crystal structures, and magnetic studies of two cobalt(II) coordination polymers based on concurrent ligand extension. Inorg. Chem. Commun. 2015, 62, 42–46. [Google Scholar] [CrossRef]
  39. Liu, X.T.; Jia, Y.Y.; Zhang, Y.H.; Ren, G.J.; Feng, R.; Zhang, S.Y.; Zaworotko, M.J.; Bu, X.H. A new Co(ii) metal–organic framework with enhanced CO2 adsorption and separation performance. Inorg. Chem. Front. 2016, 3, 1510–1515. [Google Scholar] [CrossRef]
  40. Pal, A.; Chand, S.; Das, M.C. A water-stable twofold interpenetrating microporous MOF for selective CO2 adsorption and separation. Inorg. Chem. 2017, 56, 13991–13997. [Google Scholar] [CrossRef]
  41. Yuan, J.Q.; Li, J.T.; Che, S.T.; Li, G.H.; Liu, X.Y.; Sun, X.D.; Zou, L.F.; Zhang, L.R.; Liu, Y.L. Two unique copper cluster-based metal–organic frameworks with high performance for CO2 adsorption and separation. Inorg. Chem. Front. 2019, 6, 556–561. [Google Scholar] [CrossRef]
  42. Dang, L.L.; Zhang, T.T.; Chen, T.; Zhao, Y.; Zhao, C.C.; Aznarez, F.; Sun, K.X.; Ma, L.F. Coordination assembly and NIR photothermal conversion of Cp*Rh-based supramolecular topologies based on distinct conjugated systems. Org. Chem. Front. 2022, 9, 5505–5515. [Google Scholar] [CrossRef]
  43. Dang, L.-L.; Chen, T.; Zhang, T.-T.; Li, T.-T.; Sun, J.L.; Zhang, K.J.; Ma, L.-F. Size-induced highly selective synthesis of organometallic rectangular macrocycles and heterometallic cage based on half-sandwich rhodium building block. Molecules 2022, 27, 3756. [Google Scholar] [CrossRef] [PubMed]
  44. Guo, L.J.; Feng, X.F.; Gao, Z.; Krishna, R.; Luo, F. Robust 4d–5f bimetal–organic framework for efficient removal of trace SO2 from SO2/CO2 and SO2/CO2/N2 mixtures. Inorg. Chem. 2021, 60, 1310–1314. [Google Scholar] [CrossRef]
  45. Fan, C.B.; Gong, L.L.; Huang, L.; Luo, F.; Krishna, R.; Yi, X.F.; Zheng, A.M.; Zhang, L.; Pu, S.Z.; Feng, X.F.; et al. Significant enhancement of C2H2/C2H4 separation by a photochromic diaryletheneunit: A temperature- and Light-Responsive separation switch. Angew. Chem. Int. Ed. 2017, 56, 7900–7906. [Google Scholar] [CrossRef]
  46. Xu, Z.Z.; Xiong, X.H.; Xiong, J.B.; Krishna, R.; Li, L.B.; Fan, Y.L.; Luo, F.; Chen, B.L. A robust Th-azole framework for highly efficient purification of C2H4 from a C2H4/C2H2/C2H6 mixture. Nat. Commun. 2020, 11, 3163. [Google Scholar] [CrossRef]
  47. Luo, F.; Fan, C.B.; Luo, M.B.; Wu, X.L.; Zhu, Y.; Pu, S.Z.; Xu, W.Y.; Guo, G.C. Photoswitching CO2 capture and release in a photochromic diarylethene metal–organic framework. Angew. Chem. Int. Ed. 2014, 53, 9298–9301. [Google Scholar] [CrossRef]
  48. Dang, L.-L.; Zhang, T.-T.; Li, T.-T.; Chen, T.; Zhao, Y.; Zhao, C.-C.; Ma, L.-F. Stable Zinc-Based Metal-Organic Framework Photocatalyst for Effective Visible-Light-Driven Hydrogen Production. Molecules 2022, 27, 1917. [Google Scholar] [CrossRef]
  49. Pei, X.R.; Dang, L.L.; Zhang, T.T.; Chen, T.; Ren, F.X.; Liu, S.R. Stable nickel-based metal–organic framework containing thiophene/diimidazole units for effective near-infrared photothermal conversion. Catalysts 2022, 12, 777. [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]
  51. Wang, H.; Qin, J.; Huang, C.; Han, Y.; Xu, W.; Hou, H. Mono/bimetallic water-stable lanthanide coordination polymers as luminescent probes for detecting cations, anions and organic solvent molecules. Dalton Trans. 2016, 45, 12710–12716. [Google Scholar] [CrossRef]
  52. Mason, J.A.; Sumida, K.; Herm, Z.R.; Krishna, R.; Long, J.R. Evaluating metal–organic frameworks for post-combustion carbon dioxide capture via temperature swing adsorption. Energy Environ. Sci. 2011, 4, 3030–3040. [Google Scholar] [CrossRef]
  53. Lyu, Y.; Fang, Y.; Miao, Q.Q.; Zhen, X.; Ding, D.; Pu, K.Y. Intraparticle molecular orbital engineering of semiconducting polymer nanoparticles as amplified theranostics for in vivo photoacoustic imaging and photothermal therapy. ACS Nano 2016, 10, 4472–4481. [Google Scholar] [CrossRef] [PubMed]
  54. Pu, K.; Shuhendler, A.J.; Jokerst, J.V.; Mei, J.G.; Gambhir, S.S.; Bao, Z.N.; Rao, J.H. Semiconducting polymer nanoparticles as photoacoustic molecular imaging probes in living mice. Nat. Nanotechnol. 2014, 9, 233–2393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Yan, T.; Li, Y.Y.; Gu, Q.Y.; Li, J.; Su, J.; Wang, H.Y.; Zuo, J.L. A Tetrathiafulvalene/Naphthalene Diimide-Containing Metal–Organic Framework with fsc Topology for Highly Efficient Near-Infrared Photothermal Conversion. Inorg. Chem. 2022, 61, 3078–3085. [Google Scholar] [CrossRef]
  56. Liu, Y.G.; Liu, G.L.; Peng, T.; Gu, C.; Gu, C.; Li, J.J.; Liu, X.Q.; Sun, L.B. Near-infrared light triggered release of ethane from a photothermal metal-organic framework. Chem. Eng. J. 2021, 420, 130490. [Google Scholar] [CrossRef]
  57. Lyndon, R.; Konstas, K.; Ladewig, B.P.; Southon, P.D.; Kepert, C.J.; Hill, M.R. Dynamic photo-switching in metal–organic frameworks as a route to low-energy carbon dioxide capture and release. Angew. Chem. Int. Ed. 2013, 52, 3695–3698. [Google Scholar] [CrossRef]
  58. Chen, B.L.; Ma, S.Q.; Eric, J.; Hurtado, E.J.; Lobkovsky, E.B.; Zhou, H.C. A triply interpenetrated microporous metal-organic framework for selective sorption of gas molecules. Inorg. Chem. 2007, 46, 8490–8492. [Google Scholar] [CrossRef]
  59. Yu, W.; Qiu, F.-Y.; Luo, S.-T.; Shi, H.-T.; Yuan, G.; Wei, X. Coordination Assembly and Host–Guest Chemistry of a triply Interlocked [2]Catenane. Inorg. Chem. Front. 2021, 8, 2356–2364. [Google Scholar] [CrossRef]
  60. Gao, X.; Cui, Z.; Shen, Y.-R.; Liu, D.; Lin, Y.-J.; Jin, G.-X. Synthesis and near-infrared photothermal conversion of discrete supramolecular topologies featuring half-sandwich [Cp*Rh] Units. J. Am. Chem. Soc. 2021, 143, 17833–17842. [Google Scholar] [CrossRef]
  61. Wang, Y.; Zhu, W.; Du, W.; Liu, X.; Zhang, X.; Dong, H.; Hu, W. Cocrystals strategy towards materials for near-infrared photothermal conversion and imaging. Angew. Chem. Int. Ed. 2018, 57, 3963–3967. [Google Scholar] [CrossRef]
  62. Liu, J.C.; Wang, J.F.; Han, Q.; Guan, P.S.; Liu, L.L.; Chen, L.J.; Zhao, J.W.; Streb, C.; Song, Y.F. Multicomponent self-assembly of a giant heterometallic polyoxotungstate supercluster with antitumor activity. Angew. Chem. Int. Ed. 2021, 60, 11153–11157. [Google Scholar] [CrossRef] [PubMed]
  63. Zhou, Z.; Li, X.Q.; Hu, T.T.; Xue, B.L.; Chen, H.; Ma, L.F.; Liang, R.Z.; Tan, C.L. Molybdenum-Based Nanomaterials for Photothermal Cancer Therapy. Adv. NanoBiomed Res. 2022. [Google Scholar] [CrossRef]
  64. Sheldrick, G.M. SHELXS-97, Program for the Solution of Crystal Structure; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  65. Sheldrick, G.M. SHELXL-97, Program for the Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  66. Sheldrick, G.M. SADABS Siemens Area Correction Absorption Program; University of Göttingen: Göttingen, Germany, 1994. [Google Scholar]
Scheme 1. Chemical structures of azpy and H2L in this work.
Scheme 1. Chemical structures of azpy and H2L in this work.
Molecules 27 06873 sch001
Figure 1. Single-crystal X-ray structure of 1. Coordination environment of Co2+ ions in 1 (a); coordination geometry of ligand L22− (b); view of parallel arrangement mode of azpy in 1 (c); view of 2D structure of 1 (d).
Figure 1. Single-crystal X-ray structure of 1. Coordination environment of Co2+ ions in 1 (a); coordination geometry of ligand L22− (b); view of parallel arrangement mode of azpy in 1 (c); view of 2D structure of 1 (d).
Molecules 27 06873 g001
Figure 2. View of the 3D framework of compound 1.
Figure 2. View of the 3D framework of compound 1.
Molecules 27 06873 g002
Figure 3. The PXRD patterns of compound 1 (a); infrared spectroscopy (IR) of compound 1 (b).
Figure 3. The PXRD patterns of compound 1 (a); infrared spectroscopy (IR) of compound 1 (b).
Molecules 27 06873 g003
Figure 4. The solid-state UV–Vis absorption of compound 1, ligand azpy and H2L (a); thermogravimetric curves of compound 1 (b).
Figure 4. The solid-state UV–Vis absorption of compound 1, ligand azpy and H2L (a); thermogravimetric curves of compound 1 (b).
Molecules 27 06873 g004
Figure 5. The N2 adsorption at 77 K of compound 1 (a); the distribution of pore size of compound 1 (b).
Figure 5. The N2 adsorption at 77 K of compound 1 (a); the distribution of pore size of compound 1 (b).
Molecules 27 06873 g005
Figure 6. The adsorption isotherms of CO2, CH4 and CO for compound 1 (a); The adsorption isotherms of CO2 at 273 K and 298 K for compound 1 (b).
Figure 6. The adsorption isotherms of CO2, CH4 and CO for compound 1 (a); The adsorption isotherms of CO2 at 273 K and 298 K for compound 1 (b).
Molecules 27 06873 g006
Figure 7. (a) Photothermal conversion images of compound 1 and ligands H2L and azpy; (b) Diagram of photothermal conversion measurement.
Figure 7. (a) Photothermal conversion images of compound 1 and ligands H2L and azpy; (b) Diagram of photothermal conversion measurement.
Molecules 27 06873 g007
Figure 8. (a) Photothermal conversion curve of compound 1 and ligands azpy and H2L; (b) fitting the linear line of lnθ−t in the solid state of compound 1 under laser irradiation of 0.6 W/cm2.
Figure 8. (a) Photothermal conversion curve of compound 1 and ligands azpy and H2L; (b) fitting the linear line of lnθ−t in the solid state of compound 1 under laser irradiation of 0.6 W/cm2.
Molecules 27 06873 g008
Table 1. A comparison table indicating the CO2 adsorption properties of several novel materials.
Table 1. A comparison table indicating the CO2 adsorption properties of several novel materials.
MOFAdsorption (CO2)TemperatureRef (doi)
ZSM-51.630 mmol/g363 KJ. Porous Mater.2022, 29, 19.
SBA-161.630 mmol/g363 KJ. Porous Mater.2022, 29, 19.
Zn3(OH)(p-cdc)2.50.586 mmol/g298 KSci. Total Environ.2020, 707, 135090.
PEI-MCM-415.495 mmol/g373 KDoi:10.1080/09593330.2021.1958012.
HPS−PEHA-704.5 mmol/g343 KMaterials2016, 9(10), 835.
Cu-OATA2.236 mmol/g298 KACS Appl. Mater. Interfaces2021, 13, 33188.
Cd-OATA0.509 mmol/g298 KACS Appl. Mater. Interfaces2021, 13, 33188.
MOF-17733.5 mmol/g298.15 KEnviron. Sci. Pollut. Res.2019, 26, 36214.
Mg-MOF-743.7 mmol/g77 KACS. Omega2021, 6, 7739.
MCFs0.45 mmol/g77 KACS. Omega2021, 6, 7739.
{[Co2(2,6-NDC)2(L)2]·xG}n 3.35 mmol/g273 KInorg. Chem.2019, 58, 11553.
{[Co2(2,6-NDC)2(L)2]·xG}n2.70 mmol/g295 KInorg. Chem.2019, 58, 11553.
[Co2(L)2(azpy)]n1.845 mmol/g273 KThis work
Table 2. Crystallographic data and experimental details for compound 1.
Table 2. Crystallographic data and experimental details for compound 1.
Complex1
Empirical formula C38H26Co2N6O10·3[H2O]
Formula weight 898.55
Temperature/K 293(2)
Crystal system monoclinic
Space group P21/c
a/Å 20.0316(9)
b/Å 16.1623(5)
c/Å 15.9700(6)
α/°90
β/°103.454(5)
γ/°90
Volume/Å3 5028.5(3)
Z4
ρcalc/g cm−3 1.187
µ/mm−1 0.717
F(000)/e1840.0
2θ range for data collection/° 6.584 to 50.054
Reflections collected 22293
Independent reflections 8824 [Rint = 0.0404, Rsigma = 0.0615]
Data/restraints/parameters 8824/0/505
Goodness-of-fit on F2 1.029
Largest diff. peak/hole/, e Å−30.35/−0.38
Final R indexes [I> =2σ (I)] R1 = 0.0481, wR2 = 0.1098
Final R indexes [all data] R1 = 0.0662, wR2 = 0.1182
R = [∑ ||F0| − |Fc||/∑|F0|], RW = ∑W [|F02 − Fc2|2/∑W (|Fw|2)2]1/2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dang, L.-L.; Zong, D.-X.; Lu, X.-Y.; Zhang, T.-T.; Chen, T.; Sun, J.-L.; Zhao, J.-Z.; Liu, M.-Y.; Liu, S.-R. The Selective CO2 Adsorption and Photothermal Conversion Study of an Azo-Based Cobalt-MOF Material. Molecules 2022, 27, 6873. https://doi.org/10.3390/molecules27206873

AMA Style

Dang L-L, Zong D-X, Lu X-Y, Zhang T-T, Chen T, Sun J-L, Zhao J-Z, Liu M-Y, Liu S-R. The Selective CO2 Adsorption and Photothermal Conversion Study of an Azo-Based Cobalt-MOF Material. Molecules. 2022; 27(20):6873. https://doi.org/10.3390/molecules27206873

Chicago/Turabian Style

Dang, Li-Long, De-Xi Zong, Xiao-Yan Lu, Ting-Ting Zhang, Tian Chen, Jiu-Long Sun, Jiu-Zhou Zhao, Meng-Yang Liu, and Shui-Ren Liu. 2022. "The Selective CO2 Adsorption and Photothermal Conversion Study of an Azo-Based Cobalt-MOF Material" Molecules 27, no. 20: 6873. https://doi.org/10.3390/molecules27206873

Article Metrics

Back to TopTop