Next Article in Journal
New UHPLC-QqQ-MS/MS Method for the Rapid and Sensitive Analysis of Ascorbic and Dehydroascorbic Acids in Plant Foods
Previous Article in Journal
Identifying Reducing and Capping Sites of Protein-Encapsulated Gold Nanoclusters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Telmisartan Organotin(IV) Complexes and their use as Carbon Dioxide Capture Media

1
Department of Chemistry, College of Science, Babylon University, Babil 51002, Iraq
2
Department of Chemistry, College of Science, Al-Nahrain University, Baghdad 64021, Iraq
3
Department of Optometry, College of Applied Medical Sciences, King Saud University, P.O. Box 10219, Riyadh 11433, Saudi Arabia
4
National Center for Petrochemicals Technology, King Abdulaziz City for Science and Technology, P.O. Box 6086, Riyadh 11442, Saudi Arabia
5
Department of Medical Instrumentation Engineering, Al-Mansour University College, Baghdad 64021, Iraq
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(8), 1631; https://doi.org/10.3390/molecules24081631
Submission received: 10 April 2019 / Revised: 22 April 2019 / Accepted: 24 April 2019 / Published: 25 April 2019

Abstract

:
Novel, porous, highly aromatic organotin(IV) frameworks were successfully synthesized by the condensation of telmisartan and an appropriate tin(IV) chloride. The structures of the synthesized organotin(IV) complexes were elucidated by elemental analysis, 1H-, 13C-, and 119Sn-NMR, and FTIR spectroscopy. The surface morphologies of the complexes were inspected by field emission scanning electron microscopy. The synthesized mesoporous organotin(IV) complexes have a Brunauer–Emmett–Teller (BET) surface area of 32.3–130.4 m2·g−1, pore volume of 0.046–0.162 cm3·g−1, and pore size of around 2.4 nm. The tin complexes containing a butyl substituent were more efficient as carbon dioxide storage media than the complexes containing a phenyl substituent. The dibutyltin(IV) complex had the highest BET surface area (SBET = 130.357 m2·g−1), the largest volume (0.162 cm3·g−1), and was the most efficient for carbon dioxide storage (7.1 wt%) at a controlled temperature (323 K) and pressure (50 bars).

Graphical Abstract

1. Introduction

The emission of greenhouse gases, for example, carbon dioxide (CO2), from the increased use of fossil fuels is responsible for the increasing temperatures on the earth. A high level of CO2 in the atmosphere leads to global warming [1,2,3], which causes a disturbance in nature’s equilibrium. High CO2 levels are responsible for drastic climate changes such as melting of the ice at both the north and south poles, an increase in sea level, floods, droughts, and drastic changes in the weather. In addition, a high level of CO2 in natural gas can lead to a significant reduction of the natural gas capacity and quality [4], and can damage gas pipes through corrosion [4]. Therefore, CO2 capture is vital for both the environment and industry, and has attracted the attention of researchers in both industry and academia [5]. The scientific community is under pressure to find alternative renewable sources of energy which do not contribute to increased levels of CO2 in the environment, or to develop new processes and materials that can reduce such harmful emissions from burning fossil fuels.
Some progress has been made in the development of new materials for the capture of greenhouse gases such as CO2 [6,7,8,9,10,11,12,13,14,15,16]. Recent technology has concentrated on the selective removal of CO2 from natural gas through absorption by chemicals [17]. For example, amines such as ammonia and ethanolamine can be used to absorb CO2 from natural gas [18,19]. However, amines are hazardous and have high volatility. Unlike amines, ionic liquids can be used to absorb CO2 at high temperatures [20]; however, such an approach suffers from high energy cost for the regeneration of the ionic liquids for reuse, corrosion of containers, and the use of a large volume of water [17]. A CO2 capture approach that involves the use of membrane separation using porous materials such as polymers, metal-organic frameworks, mixed matrixes, and inorganics has been developed [21,22]. Such materials are chemically stable, have polar surfaces, large pore sizes, and large surface areas [23,24,25,26]. In addition, the adsorption process is effective, simple, and environmentally friendly. However, it is very challenging and complex because it requires high pressure, multiple stages, and extensive recycling steps. Therefore, efforts have been made to develop novel adsorbents to selectively capture CO2 from natural gas. Many chemically stable porous organic polymers have been synthesized using simple procedures and used for the capture of CO2 [6,27]. The chemical-looping combustion technique can be used to selectively remove both CO2 and H2O from the gas stream in the presence of an oxygen carrier (e.g., metal oxides) [28]. Such a process is energy cost-effective, but requires a high pressure to operate. CO2 in air could also be removed through a direct air capture technique [29]. This process involves the use of a similar concept to that used in the adsorption technique. Various resins, amine-metal oxides, metal-supported carbonates, and aqueous hydroxides have been tested for the direct capture of CO2 [30]. Clearly, progress has been made in the selective removal of CO2 from natural gas, but there is still room for further improvement.
Recently, we have shown that polyphosphates can act as an efficient CO2 capture media [31]. In addition, we have investigated the synthesis and successful use of several organotin(IV) complexes as efficient photostabilizers for polymeric films [32,33] as part of our research on polymers [34,35,36,37,38,39,40,41]. In the current work, we report the synthesis of several mesoporous organotin(IV) complexes that contain both aliphatic (butyl) and aromatic (phenyl) substituents using simple and efficient procedures, and their successful use as CO2 capture media at 323 K and 50 bars. In addition, telmisartan is highly aromatic, containing both heterocycles and aryl rings and different aliphatic substituents (methyl and propyl groups) which are essential to increase the surface area and storage capacity. To the best of our knowledge, this is the first report for the use of organotin(IV) complexes as carbon dioxide capture media which turned to be efficient as for metal-organic frameworks.

2. Results and Discussion

2.1. Synthesis of Organotin(IV) Complexes 14

Four organotin(IV) complexes, 14, were synthesized from the reaction of telmisartan and the appropriate tin(IV) chloride salts. The reaction of an equimolar mixture of telmisartan and triphenyl(IV) chloride or tributyltin(IV) chloride in methanol under reflux for 8 h gave the corresponding (telmisartan)triorganotin(IV) complex 1 or 2 in 86% and 83% yield, respectively (Figure 1). Similarly, the reaction of telmisartan (two mole equivalents) and diphenyltin(IV) chloride or dibutyltin(IV) chloride in methanol under reflux for 8 h gave the corresponding bis(telmisartan)diorganotin(IV) complex 3 or 4 in 90% and 89% yield, respectively (Figure 2). Some of the physical data for organotin(IV) complexes 14, along with their elemental analyses, are shown in Table 1.
Table 1 shows that the melting point for triorganotin(IV) complex 2 was noticeably higher than that for complex 2. The variation in melting could be due the high stability of complex 2 compared with that for complex 1, since it contains the flexible butyl groups compared with the bulky phenyl groups in complex 1. In contrast, diorganotin(IV) complex 3, which contains two phenyl groups, has a higher melting point compared with that for complex 4, which contains two butyl groups. For the diorganotin(IV) complexes 3 and 4, only two substituents present and the steric hindrance becomes less important compared with that for triorganotin(IV) complexes 1 and 2.

2.2. FTIR Spectroscopy of Organotin(IV) Complexes 14

The FTIR spectra of complexes 14 show characteristic peaks within the 526–536 and 445–447 cm−1 region that correspond to the vibrations of Sn–C and Sn–O groups, respectively [42]. They also show strong absorption peaks (1685–1697 cm−1) corresponding to the vibrations of the carbonyl groups. The key FTIR spectral data of complexes 14 are shown in Table 2 (see Supplementary Materials for details).

2.3. NMR Spectroscopy of Organotin(IV) Complexes 14

The structures of organotin(IV) complexes 14 were confirmed by NMR spectroscopy (see Supplementary Materials for details). The NMR spectra show all the expected signals at the expected chemical shifts (Table 3). However, the 13C-NMR spectra of 14 show the overlap of various signals within the aromatic region (Table 4). The 119Sn-NMR spectra of 14 show the presence of singlet signals at the –185.0 to –276.0 ppm region (Table 3), which is significantly lower than that for the corresponding organotin(IV) salts. However, the chemical shift is dependent on the geometry of the complex [43,44], and these chemical shifts are consistent with the hypothesis of an increase in the tin atom coordination number within the complexes (i.e., tin nuclear shielding) [45].

2.4. Field Emission Scanning Electron Microscopy (FESEM) of Organotin(IV) Complexes 14

The morphology of the synthesized organotin(IV) complexes 14 was inspected by FESEM. The images (Figure 3) reveal that complexes 14 have homogeneous and porous structures. In addition, the images show the presence of tiny particle agglomerates, and the organotin(IV) complexes have different shapes and particles sizes. The particle sizes were calculated to be 24.56–34.13, 28.66–49.66, 23.50–32.94, and 19.68–51.47 nm for complexes 1, 2, 3, and 4, respectively. Organotin(IV) complex 3 has a lower porosity and more surface smoothness than the other organotin(IV) complexes.

2.5. Pure Gas Adsorption of Organotin(IV) Complexes 14

The physisorption isotherms of a gas can be determined either from the amount of gas removed in the gas phase or directly from the amount of gas uptake. The latter process is based on the gravimetric measurement of the adsorbent mass change [46]. The pore textural properties of complexes 14 were measured from the N2 adsorption–desorption graphs recorded at 77 K. In addition, the pore size of mesoporous complexes 14 were calculated from the adsorption–desorption isotherms using the Barrett–Joyner–Halenda (BJH) method. Organotin(IV) complex networks 14 show the formation of mesoporous structures and type III nitrogen sorption isotherms. Such isotherms have no identifiable monolayer formation [31]. The N2 isotherms and pore sizes of complexes 1–4 are shown in Figure 4, Figure 5, Figure 6 and Figure 7.
The mesopore size distribution can be calculated either from the desorption or adsorption branch of the isotherm. Organotin(IV) complexes 14 have small mesopores with consistent pore sizes (2.428–2.433 nm). Organotin(IV) complex 4 has the highest Brunauer–Emmett–Teller surface area (SBET = 130.357 m2·g−1) and the largest volume (0.162 cm3·g−1) of the organotin(IV) complexes. The porosity properties of complexes 14 are listed in Table 5.
The sorption of complexes 14 was investigated at a constant temperature (323 K) and pressure (50 bars). The sorption isotherms of CO2 and H2 in the presence of complexes 14 are shown in Figure 8, Figure 9, Figure 10 and Figure 11 and the gas uptakes are listed in Table 6. Complexes 14 showed a high CO2 uptake, possibly as a result of the strong van der Waals interaction between such complexes and CO2. The quantity of adsorbed CO2 was 18.2, 20.5, 16.5, and 35.0 cm3·g−1 for complexes 1, 2, 3, and 4, respectively. Clearly, complex 4 has the highest CO2 uptake capacity (7.1 wt%) of the organotin(IV) complexes, possibility because it has the largest BET surface area (SBET = 130.357 m2·g−1). In addition, strong hydrogen bonding and/or dipole-quadrupole interactions between CO2 and heteroatoms within the organotin(IV) complexes could take place [47]. Indeed, porous organic polymers containing oxygen, nitrogen, or sulfur atoms are efficient for the selective capture of CO2 over other gases such methane and nitrogen [48,49,50]. However, complexes 14 show very low adsorption for H2 (0.5–1.1 cm3·g−1) under identical conditions to those used for the CO2 uptake. Such behavior could be due to the weak interaction between the complexes and H2.
Complexes 14 have different geometries and substitution groups (phenyl and butyl), and therefore showed varied gas capture efficiency. It is clear that the butyl-substituted complexes (2 and 4) have better storage capacities than the phenyl-substituted complexes (1 and 3). The butyl group is flexible and leads to a larger surface area in complexes 2 and 4 (SBET = 68.357–130.357 m2·g−1) compared with that of phenyl-substituted complexes 1 and 3 (SBET = 32.374–46.338 m2·g−1). In complex 4, the flexible substitution unit (butyl group) points into the channels and leads to the formation of a molecular gate for gas adsorption [51]. It has been reported that the inclusion of alkyl chain substituents within the metal-organic frameworks can be used to tune the pores size to allow a better accommodation of gas molecules within the pores [12,52].

3. Materials and Methods

3.1. General

Chemicals were purchased from Merck (Schnelldorf, Germany). FTIR spectra (400–4000 cm−1) were recorded on an FTIR 8300 Shimadzu spectrophotometer (Tokyo, Japan) using KBr discs. An EM-017mth instrument was used to perform the elemental analyses. An MPD Mitamura Riken Kogyo apparatus (Tokushima, Japan) was used to determine melting points. 1H-(300 MHz), 13C-(75 MHz), and 119Sn-(107 MHz) NMR spectra were recorded on a Bruker DRX300 NMR spectrometer (Zurich, Switzerland). The FESEM images were recorded on a TESCAN MIRA3 FESEM system (Kohoutovice, Czech Republic) at an accelerating voltage of 26 kV. A Quantchrome chemisorption analyzer was used to record the nitrogen adsorption–desorption isotherms (77 K). The samples were dried at a vacuum oven 70 °C for 6 h under a flow of nitrogen. The pore volumes were determined at a relative pressure (P/P°) of 0.98. The BJH method was used to verify the pore size distributions. Gas uptakes were performed on an H-sorb 2600 high pressure volumetric adsorption analyzer (Beijing, China) at 50 bars and 323 K.

3.2. Synthesis of Triorganotin(IV) Complexes 1 and 2

A solution of telmisartan (0.51 g, 1.0 mmol) in chloroform (20 mL) was added slowly to a stirred solution of triphenyl or tributyltin chloride (1.0 mmol) in ethanol (10 mL) and the mixture was refluxed for 8 h. The solid obtained upon cooling was collected and recrystallized from methanol to give triorganotin(IV) 1 or 2.

3.3. Synthesis of Diorganotin(IV) Complexes 3 and 4

A solution of telmisartan (1.03 g, 2.0 mmol) in chloroform (30 mL) was added slowly to a stirred solution of diphenyl or dibutyltin chloride (1.0 mmol) and the mixture was refluxed for 8 h. The solid obtained upon cooling was collected and recrystallized from methanol to give diorganotin(IV) 3 or 4.

4. Conclusions

Four new organotin(IV) complexes containing telmisartan were synthesized and their structures were confirmed. The organotin(IV) complexes have predominantly mesoporous and heteroatom-rich structures. The performance and affinity of the tin complexes for carbon dioxide gas uptake were highly efficient compared with the performance and affinity for hydrogen gas uptake under the conditions used. The dibutyl organotin(IV) complex was the most efficient carbon dioxide storage medium, having a CO2 gas uptake up to 7.1 wt% at 323 K and 50 bar. However, the toxicity and degradability of such a complex, as well as the environmental accumulation of telmisartan, should be tested.

Supplementary Materials

The following are available online.

Author Contributions

Conceptualization and experimental design, A.G.H., K.J., E.Y., G.A.E.-H., and M.H.A.; Experimental work and data analysis, D.S.A; Writing, E.Y., G.A.E.-H., and D.S.A. All authors discussed the results and improved the final text of the paper.

Funding

The authors would like to extend their appreciation to the College of Applied Medical Sciences Research Centre and the Deanship of Scientific Research at King Saud University for their funding of this research, and to Al-Nahrain and Babylon Universities for the continued support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fanchi, J.R.; Fanchi, C.J. Energy in the 21st Century, 4th ed.; World Scientific Publishing: Singapore, Malaysia, 2016. [Google Scholar]
  2. Leung, D.Y.C.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev. 2014, 39, 426–443. [Google Scholar] [CrossRef] [Green Version]
  3. Ma, S.Q.; Zhou, H.-C. Gas storage in porous metal–organic frameworks for clean energy applications. Chem. Commun. 2010, 46, 44–53. [Google Scholar] [CrossRef] [PubMed]
  4. D’Alessandro, D.M.; Smit, B.; Long, J.R. Carbon dioxide capture: Prospects for new materials. Angew. Chem. Int. Ed. 2010, 49, 6058–6082. [Google Scholar] [CrossRef] [PubMed]
  5. Mastalerz, M.; Schneider, M.W.; Oppel, I.M.; Presly, O. A salicylbisimine cage compound with high surface area and selective CO2/CH4 adsorption. Angew. Chem. Int. Ed. 2011, 50, 1046–1051. [Google Scholar] [CrossRef] [PubMed]
  6. Ahmed, D.S.; El-Hiti, G.A.; Yousif, E.; Ali, A.A.; Hameed, A.S. Design and synthesis of porous polymeric materials and their applications in gas capture and storage: A review. J. Polym. Res. 2018, 25, 75. [Google Scholar] [CrossRef]
  7. Al-Mamoori, A.; Krishnamurthy, A.; Rownaghi, A.A.; Rezaei, F. Carbon capture and utilization update. Energy Technol. 2017, 5, 834–849. [Google Scholar] [CrossRef]
  8. Long, J.R.; Yaghi, O.M. The pervasive chemistry of metal–organic frameworks. Chem. Soc. Rev. 2009, 38, 1213–1214. [Google Scholar] [CrossRef]
  9. Eddaoudi, M.; Moler, D.B.; Li, H.; Chen, B.; Reineke, T.M.; O’Keeffe, M.; Yaghi, O.M. Modular chemistry: Secondary building units as a basis for the design of highly porous and robust metal−organic carboxylate frameworks. Acc. Chem. Res. 2001, 34, 319–330. [Google Scholar] [CrossRef]
  10. Tranchemontagne, D.J.; Mendoza-Cortés, J.L.; O’Keeffe, M.; Yaghi, O.M. Secondary building units, nets and bonding in the chemistry of metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1257–1283. [Google Scholar] [CrossRef]
  11. Johnson, J. Putting a lid on carbon dioxide. Carbon sequestration, clean-coal research mark government response to climate-change threat. Chem. Eng. News 2004, 82, 36–42. [Google Scholar] [CrossRef]
  12. Yong, Z.; Mata, V.; Rodrigues, A.E. Adsorption of carbon dioxide at high temperature—A review. Sep. Purif. Technol. 2002, 26, 195–205. [Google Scholar] [CrossRef]
  13. Férey, G. Hybrid porous solids: Past, present, future. Chem. Soc. Rev. 2008, 37, 191–214. [Google Scholar] [CrossRef] [PubMed]
  14. Yan, Y.; Yang, S.; Blake, A.J.; Schroder, M. Studies on metal–organic frameworks of Cu(II) with isophthalate linkers for hydrogen storage. Acc. Chem. Res. 2014, 47, 296–307. [Google Scholar] [CrossRef]
  15. Zhao, D.; Timmons, D.J.; Yuan, D.; Zhou, H.-C. Tuning the topology and functionality of metal−organic frameworks by ligand design. Acc. Chem. Res. 2011, 44, 123–133. [Google Scholar] [CrossRef]
  16. 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]
  17. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Mac Dowell, N.; Fernández, J.R.; Ferrari, M.-C.; Gross, R.; Hallett, J.P.; et al. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [Google Scholar] [CrossRef]
  18. Kong, Y.; Shen, X.; Fan, M.; Yang, M.; Cui, S. Dynamic capture of low-concentration CO2 on amine hybrid silsesquioxane aerogel. Chem. Eng. J. 2016, 283, 1059–1068. [Google Scholar] [CrossRef]
  19. Builes, S.; López-Aranguren, P.; Fraile, J.; Vega, L.F.; Domingo, C. Analysis of CO2 adsorption in amine-functionalized porous silicas by molecular simulations. Energy Fuels 2015, 29, 3855–3862. [Google Scholar] [CrossRef]
  20. Corvo, M.C.; Sardinha, J.; Casimiro, T.; Marin, G.; Seferin, M.; Einloft, S.; Menezes, S.C.; Dupont, J.; Cabrita, E.J. A rational approach to CO2 capture by imidazolium ionic liquids: Tuning CO2 solubility by cation alkyl branching. ChemSusChem 2015, 8, 1935–1946. [Google Scholar] [CrossRef]
  21. Japip, S.; Wang, H.; Xiao, Y.; Chung, T.S. Highly permeable zeolitic imidazolate framework (ZIF)-71 nano-particles enhanced polyimide membranes for gas separation. J. Membr. Sci. 2014, 467, 162–174. [Google Scholar] [CrossRef]
  22. Huang, Q.; Eić, M. Commercial adsorbents as benchmark materials for separation of carbon dioxide and nitrogen by vacuum swing adsorption process. Sep. Purif. Technol. 2013, 103, 203–215. [Google Scholar] [CrossRef]
  23. Thomas, A. Functional materials: From hard to soft porous frameworks. Angew. Chem. Int. Ed. 2010, 49, 8328–8344. [Google Scholar] [CrossRef] [PubMed]
  24. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.-H.; Long, J.R. Carbon dioxide capture in metal–organic frameworks. Chem. Rev. 2012, 112, 724–781. [Google Scholar] [CrossRef]
  25. 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]
  26. Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O.M. High-throughput synthesis of zeolitic imidazolate frameworks and application to CO2 capture. Science 2008, 319, 939–943. [Google Scholar] [CrossRef] [PubMed]
  27. Furukawa, H.; Yaghi, O.M. Storage of hydrogen, methane, and carbon dioxide in highly porous covalent organic frameworks for clean energy applications. J. Am. Chem. Soc. 2009, 131, 8875–8883. [Google Scholar] [CrossRef] [PubMed]
  28. Hossain, M.M.; de Lasa, H.I. Chemical-looping combustion (CLC) for inherent CO2 separations—A review. Chem. Eng. Sci. 2008, 63, 4433–4451. [Google Scholar] [CrossRef]
  29. Jones, C.W. CO2 capture from dilute gases as a component of modern global carbon management. Annu. Rev. Chem. Biomol. Eng. 2011, 2, 31–52. [Google Scholar] [CrossRef]
  30. Sanz-Pérez, E.S.; Murdock, C.R.; Didas, S.A.; Jones, C.W. Direct capture of CO2 from ambient air. Chem. Rev. 2016, 116, 11840–11876. [Google Scholar] [CrossRef]
  31. Ahmed, D.S.; El-Hiti, G.A.; Yousif, E.; Hameed, A.S.; Abdalla, M. New eco-friendly phosphorus organic polymers as gas storage media. Polymers 2017, 9, 336. [Google Scholar] [CrossRef] [PubMed]
  32. Ghazi, D.; El-Hiti, G.A.; Yousif, E.; Ahmed, D.S.; Alotaibi, M.H. The effect of ultraviolet irradiation on the physicochemical properties of poly(vinyl chloride) films containing organotin(IV) complexes as photostabilizers. Molecules 2018, 23, 254. [Google Scholar] [CrossRef]
  33. Ali, M.M.; El-Hiti, G.A.; Yousif, E. Photostabilizing efficiency of poly(vinyl chloride) in the presence of organotin(IV) complexes as photostabilizers. Molecules 2016, 21, 1151. [Google Scholar] [CrossRef]
  34. El-Hiti, G.A.; Alotaibi, M.H.; Ahmed, A.A.; Hamad, B.A.; Ahmed, D.S.; Ahmed, A.; Hashim, H.; Yousif, E. The morphology and performance of poly(vinyl chloride) containing melamine Schiff bases against ultraviolet light. Molecules 2019, 24, 803. [Google Scholar] [CrossRef]
  35. Alotaibi, M.H.; El-Hiti, G.A.; Hashim, H.; Hameed, A.S.; Ahmed, D.S.; Yousif, E. SEM analysis of the tunable honeycomb structure of irradiated poly(vinyl chloride) films doped with polyphosphate. Heliyon 2018, 4, e01013. [Google Scholar] [CrossRef]
  36. Hashim, H.; El-Hiti, G.A.; Alotaibi, M.H.; Ahmed, D.S.; Yousif, E. Fabrication of ordered honeycomb porous poly(vinyl chloride) thin film doped with a Schiff base and nickel(II) chloride. Heliyon 2018, 4, e00743. [Google Scholar] [CrossRef]
  37. Yousif, E.; Ahmed, D.S.; El-Hiti, G.A.; Alotaibi, M.H.; Hashim, H.; Hameed, A.S.; Ahmed, A. Fabrication of novel ball-like polystyrene films containing Schiff base microspheres as photostabilizers. Polymers 2018, 10, 1185. [Google Scholar] [CrossRef]
  38. Balakit, A.A.; Smith, K.; El-Hiti, G.A. Synthesis and characterization of a new photochromic alkylene sulfide derivative. J. Sulfur Chem. 2018, 39, 182–192. [Google Scholar] [CrossRef]
  39. Altaee, N.; El-Hiti, G.A.; Fahdil, A.; Sudesh, K.; Yousif, E. Screening and evaluation of poly(3-hydroxybutyrate) with Rhodococcus equi using different carbon sources. Arab. J. Sci. Eng. 2017, 42, 2371–2379. [Google Scholar] [CrossRef]
  40. Altaee, N.; El-Hiti, G.A.; Fahdil, A.; Sudesh, K.; Yousif, E. Biodegradation of different formulations of polyhydroxybutyrate films in soil. SpringerPlus 2016, 5, 762. [Google Scholar] [CrossRef]
  41. Yousif, E.; El-Hiti, G.A.; Haddad, R.; Balakit, A.A. Photochemical stability and photostabilizing efficiency of poly(methyl methacrylate) based on 2-(6-methoxynaphthalen-2-yl)propanoate metal ion complexes. Polymers 2015, 7, 1005–1019. [Google Scholar] [CrossRef]
  42. Akram, M.A.; Nazir, T.; Taha, N.; Adil, A.; Sarfraz, M.; Nazir, S.R. Designing, development and formulation of mouth disintegrating telmisartan tablet with extended release profile using response surface methodology. J. Bioequiv. Availab. 2015, 7, 262–266. [Google Scholar] [CrossRef]
  43. Pejchal, V.; Holeček, J.; Nádvorník, M.; Lyčka, A. 13C and 119Sn-NMR Spectra of some mono-n-butyltin(IV) compounds. Collect. Czech. Chem. Commun. 1995, 60, 1492–1501. [Google Scholar] [CrossRef]
  44. Shahid, K.; Ali, S.; Shahzadi, S.; Badshah, A.; Khan, K.M.; Maharvi, G.M. Organotin(IV) complexes of aniline derivatives. I. Synthesis, spectral and antibacterial studies of di- and triorganotin(IV) derivatives of 4-bromomaleanilic acid. Synth. React. Inorg. Met. Org. Chem. 2003, 33, 1221–1235. [Google Scholar] [CrossRef]
  45. Rehman, W.; Baloch, M.K.; Badshah, A.; Ali, S. Synthesis and characterization of biologically potent di-organotin(IV) complexes of mono-methyl glutarate. J. Chin. Chem. Soc. 2005, 52, 231–236. [Google Scholar] [CrossRef]
  46. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  47. Rabbani, M.G.; El-Kaderi, H.M. Template-free synthesis of a highly porous benzimidazole-linked polymer for CO2 capture and H2 storage. Chem. Mater. 2011, 23, 1650–1653. [Google Scholar] [CrossRef]
  48. Jin, Y.; Voss, B.A.; McCaffrey, R.; Baggett, C.T.; Noble, R.D.; Zhang, W. Microwave-assisted syntheses of highly CO2-selective organic cage frameworks (OCFs). Chem. Sci. 2012, 3, 874–877. [Google Scholar] [CrossRef]
  49. Yu, H.; Tian, M.; Shen, C.; Wang, Z. Facile preparation of porous polybenzimidazole networks and adsorption behavior of CO2 gas, organic and water vapors. Polym. Chem. 2013, 4, 961–968. [Google Scholar] [CrossRef]
  50. Katsoulidis, A.P.; Dyar, S.M.; Carmieli, R.; Malliakas, C.D.; Wasielewski, M.R.; Kanatzidis, M.G. Copolymerization of terephthalaldehyde with pyrrole, indole and carbazole gives microporous POFs functionalized with unpaired electrons. J. Mater. Chem. A 2013, 1, 10465–10473. [Google Scholar] [CrossRef]
  51. Zhang, W.; Wojtas, L.; Aguila, B.; Jiang, P.; Ma, S. Metal–metalloporphyrin framework modified with flexible tert-butyl groups for selective gas adsorption. ChemPlusChem 2016, 81, 714–717. [Google Scholar] [CrossRef]
  52. Meng, L.; Cheng, Q.; Kim, C.; Gao, W.-Y.; Wojtas, L.; Chen, Y.-S.; Zaworotko, M.J.; Zhang, X.P.; Ma, S. Crystal engineering of a microporous, catalytically active fcu topology MOF using a custom-designed metalloporphyrin linker. Angew. Chem. Int. Ed. 2012, 51, 10082–10085. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the telmisartan organotin (IV) complexes are available from the authors.
Figure 1. Synthesis of triorganotin(IV) complexes 1 and 2.
Figure 1. Synthesis of triorganotin(IV) complexes 1 and 2.
Molecules 24 01631 g001
Figure 2. Synthesis of diorganotin(IV) complexes 3 and 4.
Figure 2. Synthesis of diorganotin(IV) complexes 3 and 4.
Molecules 24 01631 g002
Figure 3. Field emission scanning electron microscopy (FESEM) images of 14.
Figure 3. Field emission scanning electron microscopy (FESEM) images of 14.
Molecules 24 01631 g003
Figure 4. N2 isotherms and pore diameters of 1.
Figure 4. N2 isotherms and pore diameters of 1.
Molecules 24 01631 g004
Figure 5. N2 isotherms and pore diameters of 2.
Figure 5. N2 isotherms and pore diameters of 2.
Molecules 24 01631 g005
Figure 6. N2 isotherms and pore diameters of 3.
Figure 6. N2 isotherms and pore diameters of 3.
Molecules 24 01631 g006
Figure 7. N2 isotherms and pore diameters of 4.
Figure 7. N2 isotherms and pore diameters of 4.
Molecules 24 01631 g007
Figure 8. Adsorption isotherms of CO2 and H2 for complex 1.
Figure 8. Adsorption isotherms of CO2 and H2 for complex 1.
Molecules 24 01631 g008
Figure 9. Adsorption isotherms of CO2 and H2 for complex 2.
Figure 9. Adsorption isotherms of CO2 and H2 for complex 2.
Molecules 24 01631 g009
Figure 10. Adsorption isotherms of CO2 and H2 for complex 3.
Figure 10. Adsorption isotherms of CO2 and H2 for complex 3.
Molecules 24 01631 g010
Figure 11. Adsorption isotherms of CO2 and H2 for complex 4.
Figure 11. Adsorption isotherms of CO2 and H2 for complex 4.
Molecules 24 01631 g011
Table 1. Physical properties and elemental analysis of 14.
Table 1. Physical properties and elemental analysis of 14.
Sn(IV) ComplexRColorYield (%)Melting Point (°C)Calcd. (Found; %)
CHN
1Phpale yellow86163–16570.93 (71.12)5.14 (5.13)6.49 (6.36)
2Buwhite83243–24567.25 (67.46)7.02 (7.11)6.97 (7.06)
3Phoff white90237–23972.06 (71.97)5.27 (5.36)8.62 (8.63)
4Buoff white89184–18670.53 (70.41)6.08 (5.98)8.89 (9.00)
Table 2. Key FTIR spectral data of complexes 1–4.
Table 2. Key FTIR spectral data of complexes 1–4.
Sn(IV) ComplexFTIR (ν, cm−1)
C=OC=NC=CSn–CSn–O
1168515411455526447
2169715401458528447
3169715431456536445
4169715361454536447
Table 3. 1H- and 119Sn-NMR spectral data (ppm, DMSO-d6) of complexes 14.
Table 3. 1H- and 119Sn-NMR spectral data (ppm, DMSO-d6) of complexes 14.
Sn(IV) Complex1H-NMR119Sn-NMR
11.00 (t, J = 7.6 Hz, 3H, Me), 1.83 (quintet, J = 7.6 Hz, 2H, CH2), 2.61 (s, 3H, Me), 3.17 (m, 2H, CH2), 3.81 (s, 3H, Me), 5.62 (s, 2H, CH2), 7.26–7.86 (m, 29H, Ar)–193.0
20.90 (t, J = 7.5 Hz, 9H, 3 Me), 0.99 (t, J = 7.7 Hz, 3H, Me), 1.23 (m, 6H, 3 CH2), 1.46 (m, 6H, 3 CH2), 1.58 (m, 6H, 3 CH2), 1.86 (quintet, J = 7.7 Hz, 2H, CH2), 2.63 (s, 3H, Me), 2.92 (t, J = 7.7 Hz, 2H, CH2), 3.39 (s, 3H, Me), 5.63 (s, 2H, CH2), 7.26–7.74 (m, 14H, Ar)–185.0
31.02 (t, J = 7.5 Hz, 6H, 2 Me), 1.83 (quintet, J = 7.5 Hz, 4H, 2 CH2), 2.63 (s, 6H, 2 Me), 2.93 (t, J = 7.5 Hz, 4H, 2 CH2), 3.40 (s, 6H, 2 Me), 5.63 (s, 4H, 2 CH2), 7.28–7.73 (m, 38H, Ar)–267.0
40.94–1.01 (m, 12H, 4 Me), 1.21–1.31 (m, 8H, 4 CH2), 1.79–1.83 (m, 8H, 4 CH2), 2.63 (s, 6H, 2 Me), 2.91 (t, J = 7.7 Hz, 4H, 2 CH2), 3.80 (s, 6H, 2 Me), 5.61 (s, 4H, 2 CH2), 7.21–7.86 (m, 28H, Ar)–242.5
Table 4. 13C-NMR Spectral data (ppm, DMSO-d6) of complexes 14.
Table 4. 13C-NMR Spectral data (ppm, DMSO-d6) of complexes 14.
Sn(IV) Complex13C-NMR
1168.3 (C=O), 156.7, 154.4, 143.1, 142.6, 141.3, 140.6, 137.0, 136.6, 136.4, 136.3, 135.2, 130.9, 129.8, 129.4, 128.6, 127.8, 126.8, 123.7, 122.7, 122.3, 119.1, 110.8, 109.7, 46.6 (CH2), 32.2 (Me), 29.2 (CH2), 21.2 (CH2), 16.9 (Me), 14.3 (Me)
2171.0 (C=O), 156.6, 154.5, 143.2, 142.5, 137.0, 136.4, 135.2, 130.8, 129.6, 129.2, 128.7, 128.1, 127.8, 126.8, 123.7, 122.6, 119.1, 110.4, 109.8, 46.0 (CH2), 32.2 (Me), 29.2 (CH2), 28.4 (CH2), 28.2 (CH2), 26.7 (CH2), 21.2 (CH2), 16.9 (Me), 14.3 (Me), 14.1 (Me)
3170.0 (C=O), 156.7, 154.4, 142.5, 141.5, 140.9, 140.7, 136.9, 136.4, 135.2, 132.7, 131.3, 130.8, 129.6, 129.2, 128.8, 127.7, 126.9, 123.7, 122.7, 122. 5, 119.0, 111.0, 109.9, 46.6 (CH2), 32.3 (Me), 29.2 (CH2), 21.2 (CH2), 16.9 (Me), 14.3 (Me)
4169.6 (C=O), 156.6, 154.5, 143.2, 142.9, 141.0, 140.8, 137.1, 136.3, 135.2, 133.1, 130.1, 129.8, 129.4, 128.6, 127.8, 126.8, 126.8, 123.7, 122.7, 119.1, 110.8, 109.4, 46.6 (CH2), 32.2 (Me), 30.8 (CH2), 29.2 (CH2), 27.3 (CH2), 26.1 (CH2), 21.2 (CH2), 16.9 (Me), 14.3 (Me), 13.9 (Me)
Table 5. Porosity properties of complexes 14.
Table 5. Porosity properties of complexes 14.
Sn(IV) ComplexSBET (m2·g−1) aVtotal (cm3·g−1) bPore Size (nm) c
146.3380.0622.433
268.4340.0972.428
332.3740.0462.432
4130.3570.1622.429
a Brunauer–Emmett–Teller (BET) surface area; b pore volume was calculated at a relative pressure (P/Po) of 0.98 from the nitrogen adsorption isotherm; c Barrett–Joyner–Halenda (BJH) average pore diameter was calculated from the desorption data.
Table 6. Gas uptake capacities at 323 K and 50 bars of complexes 14. a
Table 6. Gas uptake capacities at 323 K and 50 bars of complexes 14. a
Sn(IV) ComplexCO2 Uptake (cm3·g−1)CO2 Uptake (wt%)H2 Uptake (cm3·g−1)H2 Uptake (wt%)
118.23.61.10.009
220.54.00.70.006
316.53.30.50.006
435.07.11.10.013
a Data were collected by the volumetric gas sorption method.

Share and Cite

MDPI and ACS Style

Hadi, A.G.; Jawad, K.; Yousif, E.; El-Hiti, G.A.; Alotaibi, M.H.; Ahmed, D.S. Synthesis of Telmisartan Organotin(IV) Complexes and their use as Carbon Dioxide Capture Media. Molecules 2019, 24, 1631. https://doi.org/10.3390/molecules24081631

AMA Style

Hadi AG, Jawad K, Yousif E, El-Hiti GA, Alotaibi MH, Ahmed DS. Synthesis of Telmisartan Organotin(IV) Complexes and their use as Carbon Dioxide Capture Media. Molecules. 2019; 24(8):1631. https://doi.org/10.3390/molecules24081631

Chicago/Turabian Style

Hadi, Angham G., Khudheyer Jawad, Emad Yousif, Gamal A. El-Hiti, Mohammad Hayal Alotaibi, and Dina S. Ahmed. 2019. "Synthesis of Telmisartan Organotin(IV) Complexes and their use as Carbon Dioxide Capture Media" Molecules 24, no. 8: 1631. https://doi.org/10.3390/molecules24081631

Article Metrics

Back to TopTop