Next Article in Journal
Growth Mechanism of SmB6 Nanowires Synthesized by Chemical Vapor Deposition: Catalyst-Assisted and Catalyst-Free
Previous Article in Journal
Enhanced Phase Transition Properties of VO2 Thin Films on 6H-SiC (0001) Substrate Prepared by Pulsed Laser Deposition
Previous Article in Special Issue
A Review on Catalytic Nanomaterials for Volatile Organic Compounds VOC Removal and Their Applications for Healthy Buildings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient One-Pot Synthesis of a Hexamethylenetetramine-Doped Cu-BDC Metal-Organic Framework with Enhanced CO2 Adsorption

1
U.S Pakistan Centre for Advanced Studies in Energy, National University of Sciences and Technology, H-12, Islamabad 44000, Pakistan
2
School of Chemical and Materials Engineering, National University of Sciences and Technology, H-12, Islamabad 44000, Pakistan
3
School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff CF10 3AT, UK
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(8), 1063; https://doi.org/10.3390/nano9081063
Submission received: 14 May 2019 / Revised: 27 June 2019 / Accepted: 30 June 2019 / Published: 24 July 2019
(This article belongs to the Special Issue Advanced Nanomaterials for Pollutant Gases Reduction and Abatement)

Abstract

:
Herein we report a facile, efficient, low cost, and easily scalable route for an amine-functionalized MOF (metal organic framework) synthesis. Cu-BDC⊃HMTA (HMTA = hexamethylenetetramine) has high nitrogen content and improved thermal stability when compared with the previously reported and well-studied parent Cu-BDC MOF (BDC = 1,4-benzenedicarboxylate). Cu-BDC⊃HMTA was obtained via the same synthetic method, but with the addition of HMTA in a single step synthesis. Thermogravimetric studies reveal that Cu-BDC⊃HMTA is more thermally stable than Cu-BDC MOF. Cu-BDC⊃HMTA exhibited a CO2 uptake of 21.2 wt % at 273 K and 1 bar, which compares favorably to other nitrogen-containing MOF materials.

Graphical Abstract

1. Introduction

The world is currently facing the urgent and demanding challenges of saving and utilizing energy as efficiently as possible. Ever increasing carbon dioxide levels in the atmosphere are a serious threat to the environment [1]. Various carbon capture techniques have been explored to mitigate carbon dioxide levels in the atmosphere, including point source CO2 capture using advanced materials [2]. Metal−organic frameworks (MOFs) are an advanced class of microporous and often crystalline nanomaterials comprised of metal coordination sites bridged by organic linkers [3,4]. The resulting organic/inorganic hybrid 3-D networks that form often contain well-defined porosity, high surface area, and tunable chemical functionalities with potential for versatile applications in catalysis [5,6], separations [7], and gas storage [8]. With respect to the last of these, many studies have investigated the capture of carbon dioxide gas. Amine sites have an affinity towards carbon dioxide, are known to be highly effective at enhancing CO2 adsorption, and are amenable to use under dry or humid conditions. [9]. In this paper, we describe the synthesis, characterization and CO2 sorption of a hexamethylenetetramine-doped metal-organic framework. This study is an effort to incorporate hexamethylenetetramine within a Cu-BDC (BDC = 1,4-benzenedicarboxylate) framework, using an in-situ modification during synthesis, and to study the effect on carbon dioxide gas sorption capacity. Herein, we report a very straightforward method for modification of already reported Cu-BDC [10]. The strategy has several advantages. First, the entire synthetic procedure is quite simple. Second, this method is efficient, with the potential for high-yields, and hexamethylenetetramine is a low-cost chemical (£15.68/kg [11]). Third, a high nitrogen content can be achieved in the resulting Cu-BDC⊃HMTA material.

2. Materials and Methods

All the chemicals were purchased from Sigma Aldrich/Merck (St. Louis, MO, USA) and used as received.

2.1. Synthesis

To prepare Cu-BDC⊃HMTA, equimolar quantities (1:1:1) of Cu(NO3)2·6H2O (296 mg, 1 mmol), terephthalic acid (166 mg, 1 mmol) and hexamethylenetetramine (140 mg, 1 mmol) were dissolved in 10 mL DMF in a 50 mL beaker. The contents were ultrasonicated at 25 °C for 30 min, and then the solution was transferred to a 23 mL Teflon vial in a steel Parr vessel. The Parr vessel was sealed and heated in an oven at 110 °C for 24 h to yield greenish-blue crystals. The reaction mixture was decanted, the product washed three times with DMF (5 mL), and then three times with THF (5 mL). This yielded blue crystals. The sample was activated in a vacuum oven at 130 °C for 12 h before further analysis. The same synthesis strategy was used for obtaining Cu-BDC MOF, without the addition of HMTA [10]. This yielded blue crystals. A schematic reaction scheme for Cu-BDC⊃HMTA synthesis is shown in Figure 1.

2.2. Characterization

Powder X-ray diffraction (PXRD) patterns were collected on an X’PertPro Panalytical Chiller 59 diffractometer (Malvern, UK) using copper Kα (1.54 Å) radiation. Diffraction patterns were recorded in the 2θ range, from 4.00 to 39.09 degrees, with a 2θ step size of 0.017, and a scan per step of 34.9 s. Elemental analyses (N, C and H) of prepared Cu-BDC⊃HMTA samples were performed to confirm presence of amine in the prepared material, using a FlashSmart NC ORG elemental analyzer (Oxford, UK). Thermogravimetic analyses (TGA) were performed using a Perkin Elmer Pyris 1 thermo-gravimetric analyzer (Champaign, IL, USA). The temperature was increased from 25 °C to 700 °C at a heating rate of 5 °C min−1 under a flow of air (20 mL min−1). SEM images were collected using TESCAN/VEGA-3 equipment (Brno, Czech Republic). A SHIMADZU IR Affinitt-1S spectrometer (Kyoto, Japan) was used to obtain IR spectra.
Prior to CO2 sorption studies, the samples were degassed at 130 °C for 10 h and then back-filled with helium gas. CO2 adsorption experiments were performed on a Quantachrome Isorb-HP100 volumetric type sorption analyzer (Boynton Beach, FL, USA). The sample was tested for adsorption at two different temperatures: 0 °C and 25 °C, at pressures from 1–14 bar. N2 adsorption studies of the MOF were conducted to analyze surface area and pore volume using a Quantachrome Nova 2200e at −196 °C at a relative pressure of P/P⁰ = 0.1–1.0 and prior to the measurement, samples were degassed at 160 °C under vacuum for 11 h.

3. Results and Discussion

3.1. PXRD Patterns of Cu-BDC and Cu-BDC⊃HMTA

Powder X-ray diffraction patterns, for products of the synthesis of Cu-BDC and Cu-BDC⊃HMTA, were collected and are shown in Figure 2. Both synthesized materials show sharp diffraction peaks indicating predominantly crystalline material. The peak positions are in good agreement with the PXRD of Cu-BDC previously reported by Carson et al. in 2009, indicating successful synthesis of the Cu-BDC [10]. There are several additional peaks in the PXRD pattern of Cu-BDC⊃HMTA, most notably at 2θ = 8.31°. These cannot be ascribed to a simple mechanical mixture of Cu-BDC and HMTA as, firstly, HMTA is soluble in the organic solvents used to wash the reaction product, which makes this option unlikely. Secondly, the increased CO2 sorption observed in the Cu-BDC⊃HMTA product (described below) is not consistent with a simple physical mixture of Cu-BDC and HMTA, which we would anticipate having a lower gas uptake than the pure Cu-BDC alone. Thirdly, and most conclusively, the powder XRD pattern of pure HMTA is shown in Figure 2, and no HMTA peaks correspond to the new peaks observed in Cu-BDC⊃HMTA. We tentatively ascribe the additional peaks to well-ordered HMTA molecules binding to the axial copper sites in the framework. Inspecting the (previously reported) Cu-BDC crystal structure [10] shows that the 2D layers pack with significantly offset paddlewheels from one layer to the next, with copper centers separated by 6.331 Å (Cu–Cu distance). Looking approximately along the Cu–Cu axis, although still with some notable offset, the next available paddlewheel is 9.068 Å (Cu–Cu distance) away. HMTA has four potentially coordinating nitrogen atoms separated by ~2.47 Å (N–N) across a tricyclic, adamantane-like cage [12]. There are examples of HMTA bridging copper nodes in a metal-organic framework, but these typically involve much shorter Cu–Cu distances (5.761 Å in the cited example) with an angle between the Cu-N bonds of adjacent centres across the HMTA of ~108°, which is not possible in our Cu-BDC⊃HMTA framework [12]. Further indirect evidence for bonding of HMTA to the framework is that it is not readily removed by simple washing with organic solvents. It has not been possible at this stage to synthesize crystals of Cu-BDC⊃HMTA of sufficient size and quality to obtain the single-crystal X-ray structure.

3.2. FTIR of Cu-BDC and Cu-BDC⊃HMTA

Fourier transform infrared spectra (FTIR) collected for prepared materials confirm the presence of representative functional groups indicative of Cu-BDC MOF formation (Figure 3). Sharp peaks representative of symmetric and asymmetric stretching of carboxylates bonded to Cu are observed at 1521 cm−1 and 1362 cm−1 in the Cu-BDC sample [10]. Both materials show the presence of what is likely to be water (even after vacuum-oven drying the samples) in the form of a broad peak centered around 3400 cm−1, which is much more evident in the Cu-BDC sample than in the Cu-BDC⊃HMTA material and is likely due to the rapid uptake of atmospheric water when performing the measurement in air. The relatively reduced water content in the Cu-BDC⊃HMTA sample may indicate slower water adsorption as a result of pore-blocking by adsorbed HMTA and, in agreement with the binding of HMTA to copper nodes proposed above, the occupation of axial sites on the copper paddlewheels by HMTA which otherwise could rapidly adsorb water. In addition to peaks coincident with those of Cu-BDC MOF, the Cu-BDC⊃HMTA sample illustrates some new features. A characteristic peak for amine-containing functional groups is observed at 1089 cm−1, consistent with C–N bond stretching [13,14]. Peaks at 2915 and 2845 cm−1 can be ascribed to stretching vibrations of C–H bonds introduced by the incorporation of HMTA [14,15].

3.3. Thermal Stability of Cu-BDC and Cu-BDC⊃HMTA

The two materials were studied by TGA and the results are shown in Figure 4. For both MOFs there is less than 2% weight loss observed below 150 °C, indicating the pre-treatment has removed the majority of the residual solvent, and there is only minimal adsorbed moisture. The small weight loss in Cu-BDC between 170 °C and 320 °C (approx. 8%) is consistent with loss of surface adsorbed DMF [13]. For Cu-BDC, decomposition of the benzene dicarboxylate starts at about 375 °C, above which temperature there is rapid degradation to the metal oxide. Notably, there is multi-step degradation of Cu-BDC⊃HMTA, with mass losses from ~250 °C and 425 °C consistent with HMTA sublimation and thermal degradation, respectively. In Cu-BDC⊃HMTA, linker degradation appears to be overlapped with HMTA degradation. No further weight loss was observed above 450 °C for Cu-BDC MOF, and above 550 °C for Cu-BDC⊃HMTA. The Cu-BDC thermal degradation results in 25% metal oxide, and 74% linker + DMF in the initial mass, consistent with the expected metal:linker:DMF ratio of 1:1:1. The relative proportions of Cu-BDC⊃HMTA components cannot reliably be extracted from these data due to the difficulty disentangling the overlapping mass losses of HMTA and linker in this experiment.

3.4. Elemental Composition of Cu-BDC and Cu-BDC⊃HMTA

To confirm the chemical composition of both samples, elemental analysis and EDS were performed (Table 1). The empirical formulae calculated on the basis of EDS, and elemental analysis for Cu-BDC, and Cu-BDC⊃HMTA are: C11H11CuNO5 and C14H16CuN4O4, respectively. This is consistent with metal: linker: {DMF or HMTA, respectively} molar ratios of 1:1:1, and in line with each of the copper axial sites being occupied by HMTA in Cu-BDC⊃HMTA.

3.5. Morphology of Cu-BDC and Cu-BDC⊃HMTA

Scanning electron microscopy images of the prepared samples are shown in Figure 5. The Cu-BDC crystallites have a plate like structure, while Cu-BDC⊃HMTA shows a more regular flat rod-like structure.

3.6. CO2 Adsorption Studies of Cu-BDC MOF and Cu-BDC⊃HMTA

The CO2 adsorption capacity for both MOF materials was evaluated by monitoring pseudo equilibrium adsorption uptake. Samples were initially degassed at 130 °C for 12 h. 200 mg of each sample was used for three consecutive adsorption-desorption cycles at 273 K or 298 K with adsorbate pressure ranging between 1 to 14 bar. For Cu-BDC⊃HMTA, the CO2 uptake recorded at 1 bar was 4.8 mmol g−1 (21.1 wt %), and 2.1 mmol g−1 (9.24 wt %) at 273 K and 298 K, respectively (Figure 6). Notably, for Cu-BDC, without the amine modification, CO2 uptake was measured at 1 bar as only 1.2 mmol g−1 (5.28 wt %), and 0.8 mmol g−1 (3.53 wt %) at 273 K and 298 K, respectively. At 14 bar, the CO2 uptake at 273 °C, and 298 °C for Cu-BDC⊃HMTA is 12 mmol g−1 (52.8 wt %), and 9 mmol g−1 (39.6 wt %), respectively (Figure 6), again markedly higher than for Cu-BDC (17.4 wt %, and 13.2 wt % at 273 K and 298 K, respectively).

3.7. Surface Area and Porosity of Cu-BDC⊃HMTA MOF

The N2 adsorption isotherm for Cu-BDC⊃HMTA was recorded at 77 K (Figure 7B). The Langmuir and BET surface areas for Cu-BDC were found to be 868 m2/g and 708 m2/g, respectively, while Cu-BDC⊃HMTA revealed lower values of 683 m2/g (Langmuir), and 590 m2/g (BET) (Table 2). Although the introduction of the amine into Cu-BDC⊃HMTA reduces its surface area, the CO2 adsorption is increased. The presence of additional binding sites in MOFs by amine/amide incorporation has been shown to induce dispersion, and electrostatic forces that enhance CO2 gas adsorption (Table 2) [13]. The isosteric heat of CO2 adsorption (Qst) in Cu-BDC⊃HMTA was calculated from the adsorption isotherms at 273 and 298 K (Figure 7A) as 29.8 kJ mol−1. Such a moderate value is lower than many other MOFs (see Table 2), and is highly desirable because of the anticipated lower material regeneration energy demand.
Qst is the heat Q released in a constant temperature calorimeter when a differential amount of gas is adsorbed at constant pressure. The Van’t Hoff isobar equation relates Qst to adsorption isotherms at different temperatures. It is derived from equating the chemical potential of the adsorbed phase, and the gas phase, applying the Gibbs Helmholtz relation, and assuming that the vapor phase behaves like an ideal gas. From experimentally obtained isotherms at a constant amount adsorbed and two different temperatures (T1 and T2), Qst is obtained by following equation:
Q s t = R   ( ( l n P 1 l n P 2 ) ( 1 T 1 1 T 2 ) )
where R is ideal gas constant. Isosteric heats of adsorption for Cu-BDC⊃HMTA were calculated using 273 K and 298 K isotherms using the slope of a Van’t Hoff plot against the amount adsorbed. Here, the Qst value decreases with loading, indicating strong interaction between the quadrupole moment of carbon dioxide and the adsorbent surface.
The pore size distribution obtained from BET–BJH N2 adsorption shows micropores at around 8.3 Å in Cu-BDC, and larger pores at 14 Å that may originate from defects or inter-crystalline gaps [14]. Unsurprisingly, a narrow pore distribution of smaller pores around 7.1 Å is observed in the Cu-BDC⊃HMTA sample.
The uptake of CO2 is only part of the utility of these materials. They also need to have reproducible uptake on more than one sorption-desorption cycle. Cu-BDC revealed a significant loss in adsorption capacity over successive operations compared to Cu-BDC⊃HMTA (Figure 8). Here, adsorption capacity calculated at 298 K and 14 bar for Cu-BDC lowered by about 14% after three cycles from 3 to 2.6 mmol/g. This decrease in adsorption capacity over successive cycles was more prominent at higher temperature compared to lower temperature (273 K) adsorption. In contrast, Cu-BDC⊃HMTA demonstrated much lower percentage decline in CO2 uptake over three successive adsorption cycles (1.1% at 298 K and 0.83% at 273 K).

4. Conclusions

In summary, we report the simple modification of a Cu-BDC MOF during synthesis by the incorporation of a hexamethylenetetramine additive. The Cu-BDC⊃HMTA MOF material forms as a crystalline solid with rod-like crystallites. Thermogravimetric studies reveal that Cu-BDC⊃HMTA is more thermally stable than Cu-BDC MOF. Moreover, carbon dioxide adsorption studies for these samples reveal markedly better carbon dioxide uptake by the amine-modified framework (5.25 wt % for Cu-BDC, and 21.2 wt % for Cu-BDC⊃HMTA, respectively, at 273 K and 1 bar). The addition of nitrogen atoms by the incorporation of HMTA leads to the enhanced adsorption of CO2 gas, which we ascribe to favorable interactions [26] between CO2 molecules and the nitrogen-modified pores [27]. The modified MOF, Cu-BDC⊃HMTA, also displays enhanced cyclic stability and can be reused over three cycles. This study describes a cost-effective strategy for the incorporation of amine groups in MOF structures, for enhanced CO2 capture applications, using HMTA as a cheap additive. This serves as a low-cost alternative to expensive amine based ligands that are often custom-built to make MOFs for carbon dioxide capture. Future studies are needed to address the longer-term stability of Cu-BDC⊃HMTA and stability to contaminants.

Author Contributions

Conceptualization, T.L.E.; Data curation, M.A.; Investigation, A.A.; Methodology, N.I.; Supervision, T.N., T.L.E.; writing, all authors.

Funding

Funding for this project was provided by USPCAS-E, Cardiff University, NUST Pakistan and Higher Education, Pakistan and The APC was funded by NUST, Pakistan.

Acknowledgments

The authors are very grateful to Higher Education Commission, Pakistan for providing financial support to carry out some part of this research work at Cardiff University, UK. TLE gratefully acknowledges the Royal Society for the award of a University Research Fellowship (6866), and Cardiff University for funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bao, Z.; Yu, L.; Ren, Q.; Lu, X.; Deng, S. Adsorption of CO2 and CH4 on a magnesium-based metal organic framework. J. Colloid Interface Sci. 2011, 353, 549–556. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, J.; Thallapally, P.K.; McGrail, B.P.; Brown, D.R.; Liu, J. Progress in adsorption-based CO2 capture by metal-organic frameworks. Chem. Soc. Rev. 2012, 41, 2308–2322. [Google Scholar] [CrossRef] [PubMed]
  3. 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] [PubMed]
  4. Easun, T.L.; Nevin, A.N. Metal nodes and metal sites in metal–organic frameworks. Organomet. Chem. 2019, 42, 54–79. [Google Scholar]
  5. Li, J.R.; Sculley, J.; Zhou, H.C. Metal–Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869–932. [Google Scholar] [CrossRef] [PubMed]
  6. Han, D.; Jiang, F.L.; Wu, M.Y.; Chen, L.; Chen, Q.H.; Hong, M.C. A non-interpenetrated porous metal-organic framework with high gas-uptake capacity. Chem. Commun. 2011, 47, 9861–9866. [Google Scholar] [CrossRef]
  7. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [PubMed]
  8. Furukawa, H.; Muller, U.; Yaghi, O.M. “Heterogeneity within order” in metal-organic frameworks. Angew. Chem. Int. Ed. 2015, 54, 3417–3430. [Google Scholar] [CrossRef]
  9. Mason, J.A.; McDonald, T.M.; Bae, T.-H.; Bachman, J.E.; Sumida, K.; Dutton, J.J.; Kaye, S.S.; Long, J.R. Application of a high-throughput analyzer in evaluating solid adsorbents for post combustion carbon capture via multicomponent adsorption of CO2, N2, and H2O. J. Am. Chem. Soc. 2015, 137, 4787–4803. [Google Scholar] [CrossRef]
  10. Carson, C.G.; Hardcastle, K.; Schwartz, J.; Liu, X.; Hoffmann, C.; Gerhardt, R.A.; Tannenbaum, R. Synthesis and structure characterization of copper terephthalate metal-organic frameworks. Eur. J. Inorg. Chem. 2009, 16, 2338–2343. [Google Scholar] [CrossRef]
  11. Hexamethylenetetraamine; Sigma-Aldrich Corporation online (UK) pricing 11th April 2019 of £62.80 for 4 kg, ReagentPlus®, 99%. Available online: https://www.sigmaaldrich.com/catalog/product/sial/h11300?lang=en&region=PK (accessed on 28 April 2019).
  12. Ilyes, E.; Florea, M.; Madalan, A.M.; Haiduc, I.; Parvulescu, V.I.; Andruh, M. A Robust Metal–Organic Framework Constructed from Alkoxo-Bridged Binuclear Nodes and Hexamethylenetetramine Spacers: Crystal Structure and Sorption Studies. Inorg. Chem. 2012, 51, 7954–7956. [Google Scholar] [CrossRef] [PubMed]
  13. 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]
  14. Yang, Y.; Lin, R.; Ge, L.; Hou, L.; Bernhardt, P.; Rufford, T.E.; Wang, S.; Rudolph, V.; Wang, Y.; Zhu, Z. Synthesis and characterization of three amino-functionalized metal-organic frameworks based on the 2-aminoterephthalic ligand. Dalton Trans. 2015, 44, 8190–8197. [Google Scholar] [CrossRef]
  15. Hu, Z.; Nalaparaju, A.; Peng, Y.; Jiang, J.; Zhao, D. Modulated hydrothermal synthesis of UIO-66(Hf) type Metal Organic Frameworks for Optimal Carbon Dioxide Separation. Inorg. Chem. 2016, 55, 1134–1141. [Google Scholar] [CrossRef] [PubMed]
  16. Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M.J. Enhanced CO2 Binding Affinity of a High-Uptakerht-Type Metal−Organic Framework Decorated with Acylamide Groups. J. Am. Chem. Soc. 2011, 133, 748–751. [Google Scholar] [CrossRef] [PubMed]
  17. Cai, J.F.; Wang, H.Z.; Wang, H.L.; Duan, X.; Wang, Z.Y.; Cui, Y.J.; Yang, Y.; Chen, B.L.; Qian, G.D. An amino-decorated NbO-type metal–organic framework for high C2H2 storage and selective CO2 capture. RSC Adv. 2015, 5, 77417–77422. [Google Scholar]
  18. Xiong, S.; He, Y.; Krishna, R.; Chen, B.; Wang, Z. Metal–Organic Framework With Functional Amide Groups For Highly Selective Gas Separation. Cryst. Growth Des. 2013, 13, 2670–2674. [Google Scholar] [CrossRef]
  19. Zheng, B.; Yang, Z.; Bai, J.; Li, Y.; Li, S. High and selective CO2 capture by two mesoporous acylamide-functionalized rht-type metal–organic frameworks. Chem. Commun. 2012, 48, 7025–7028. [Google Scholar] [CrossRef]
  20. Duan, J.; Yang, Z.; Bai, J.; Zheng, B.; Li, Y.; Li, S. Highly selective CO2 capture of an agw-type metal–organic framework with inserted amides: Experimental and theoretical studies. Chem. Commun. 2012, 48, 3058–3060. [Google Scholar] [CrossRef]
  21. Park, J.; Li, J.R.; Chen, Y.P.; Yu, J.; Yakovenko, A.; Wang, Z.U.; Sun, L.B.; Balbuena, P.B.; Zhou, H.C. A Versatile Metal–Organic Framework For Carbon Dioxide Capture and Cooperative Catalysis. Chem. Commun. 2012, 48, 9995–9997. [Google Scholar] [CrossRef]
  22. Lu, Z.; Xing, H.; Sun, R.; Bai, J.; Zheng, B.; Li, Y. Water Stable Metal–Organic Framework Evolutionally Formed from a Flexible Multidentate Ligand with Acylamide Groups for Selective CO2 Adsorption. Cryst. Growth Des. 2012, 12, 1081–1084. [Google Scholar] [CrossRef]
  23. Wang, Z.; Zheng, B.; Liu, H.; Lin, X.; Yu, X.; Yi, P.; Yun, R. High Capacity Gas Storage by a Microporous Oxalamide Functionalized NbO-Type Metal–Organic Framework. Cryst. Growth Des. 2013, 13, 5001–5006. [Google Scholar] [CrossRef]
  24. Demessence, A.; D’Alessandro, D.M.; Foo, M.L.; Long, J.R. Strong CO2 binding in a water stable, triazolate-bridged metal-organic framework functionalized with ethylenediamine. J. Am. Chem. Soc. 2009, 131, 8784–8786. [Google Scholar] [CrossRef] [PubMed]
  25. Montoro, C.; Garcia, E.; Calero, S.; Perez-Fernandez, M.A.; Lopez, A.L.; Barea, E.; Navarro, J. Functionalization of MOF open metal sites with pendant amines for CO2 capture. J. Mater. Chem. 2012, 22, 10155–10158. [Google Scholar] [CrossRef]
  26. 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]
  27. Karmakar, A.; Desai, A.V.; Manna, B.; Joarder, B.; Ghosh, S.K. An Amide-Functionalized Dynamic Metal–Organic Framework Exhibiting Visual Colorimetric Anion Exchange and Selective Uptake of Benzene over Cyclohexane. Chem. A Eur. J. 2015, 21, 7071–7076. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Reaction scheme for Cu-BDC⊃HMTA synthesis.
Figure 1. Reaction scheme for Cu-BDC⊃HMTA synthesis.
Nanomaterials 09 01063 g001
Figure 2. PXRD patterns for Cu-BDC (black, reported by Carson et al., 2009), Cu-BDC (red, synthesized herein), Cu-BDC⊃HMTA (blue), HMTA (green), Cu(NO3)2·6H2O (lilac) and the H2BDC linker (dark yellow).
Figure 2. PXRD patterns for Cu-BDC (black, reported by Carson et al., 2009), Cu-BDC (red, synthesized herein), Cu-BDC⊃HMTA (blue), HMTA (green), Cu(NO3)2·6H2O (lilac) and the H2BDC linker (dark yellow).
Nanomaterials 09 01063 g002
Figure 3. FTIR spectra for Cu-BDC and Cu-BDC⊃HMTA.
Figure 3. FTIR spectra for Cu-BDC and Cu-BDC⊃HMTA.
Nanomaterials 09 01063 g003
Figure 4. TGA of Cu-BDC MOF (dashed line) and Cu-BDC⊃HMTA (solid line).
Figure 4. TGA of Cu-BDC MOF (dashed line) and Cu-BDC⊃HMTA (solid line).
Nanomaterials 09 01063 g004
Figure 5. SEM images at 10 µm for (a) Cu-BDC MOF and at 20 µm for (b) Cu-BDC⊃HMTA MOF.
Figure 5. SEM images at 10 µm for (a) Cu-BDC MOF and at 20 µm for (b) Cu-BDC⊃HMTA MOF.
Nanomaterials 09 01063 g005
Figure 6. CO2 adsorption isotherms for Cu-BDC and Cu-BDC⊃HMTA at 273 K and 298 K, as shown.
Figure 6. CO2 adsorption isotherms for Cu-BDC and Cu-BDC⊃HMTA at 273 K and 298 K, as shown.
Nanomaterials 09 01063 g006
Figure 7. (A) Isosteric heats of CO2 adsorption onto Cu-BDCHMTA. (B) N2 adsorption-desorption isotherms at 77 K. Adsorption is represented by hollow circles (Red = Cu-BDCHMTA, Green = Cu BDC), and desorption is marked by closed circles.
Figure 7. (A) Isosteric heats of CO2 adsorption onto Cu-BDCHMTA. (B) N2 adsorption-desorption isotherms at 77 K. Adsorption is represented by hollow circles (Red = Cu-BDCHMTA, Green = Cu BDC), and desorption is marked by closed circles.
Nanomaterials 09 01063 g007
Figure 8. CO2 adsorption in mmol/g calculated at 14 bar for Cu-BDC and Cu-BDC⊃HMTA at 273 K and 298 K as shown.
Figure 8. CO2 adsorption in mmol/g calculated at 14 bar for Cu-BDC and Cu-BDC⊃HMTA at 273 K and 298 K as shown.
Nanomaterials 09 01063 g008
Table 1. Elemental composition of Cu-BDC and Cu-BDC⊃HMTA.
Table 1. Elemental composition of Cu-BDC and Cu-BDC⊃HMTA.
Elemental CompositionCalculated by Elemental AnalyzerCalculated by EDS
MOF SampleCHNCONCu
Cu-BDC44.03
(44.07)
3.64
(3.69)
4.63
(4.67)
45.94
(44.0)
27.27
(26.62)
5.70
(4.67)
21.09
(21.13)
Cu-BDC⊃HMTA45.80
(45.77)
4.4
(4.39)
15.30
(15.26)
47.81
(45.77)
18.59
(17.41)
16.30
(15.26)
17.30
(17.27)
Note: Theoretical values in brackets and calculated values outside brackets.
Table 2. Surface area, CO2 uptake and Qst values for selected Cu-based MOFs.
Table 2. Surface area, CO2 uptake and Qst values for selected Cu-based MOFs.
MaterialBET (m2/g)Temperature (K)Pressure (bar)CO2 Adsorption (wt %)Qst (KJ mol−1)Reference
Cu (TATB)3360293--61[13]
[Cu3(TDPAT)]1938273125.842.2[14]
Cu 2 ( H 2 O ) 2 BDPO 2447273140.125.4[15]
[Cu4(μ4-O)Cl2(COO)4N4]26902731027.336.5[16]
Cu ( pia ) 2 ( SiF 6 ) ( EtOH ) 2 ( H 2 O ) 12 28529615.530[17]
[ Cu 3 ( BTB ) 6 ] n 328827320157-[18]
[ Cu 3 L 2 ( H 2 O ) 5 ] 2690273127.3-[19]
[ Cu 2 PDAI ( H 2 O ) ] 1372273128.626.3[20]
[ Cu 2 ( TCMBT ) ( bpp ) ( μ 3 OH ) ] · 6 H 2 O 8082982025.526.7[21]
[ Cu 2 ( BDPT 4 ) ( H 2 O ) 2 ] 1400273130.722.5[22]
en-CuBTTri345298158.290[23]
en@CuBTC-298119.530[24]
mmen-CuBTTri870298115.4-[25]
Cu-BDC7082731 (14)5.28 (17.4)-Present study
Cu-BDC⊃HMTA5902731 (14)21.2 (52.8)29.8Present study

Share and Cite

MDPI and ACS Style

Asghar, A.; Iqbal, N.; Noor, T.; Ali, M.; Easun, T.L. Efficient One-Pot Synthesis of a Hexamethylenetetramine-Doped Cu-BDC Metal-Organic Framework with Enhanced CO2 Adsorption. Nanomaterials 2019, 9, 1063. https://doi.org/10.3390/nano9081063

AMA Style

Asghar A, Iqbal N, Noor T, Ali M, Easun TL. Efficient One-Pot Synthesis of a Hexamethylenetetramine-Doped Cu-BDC Metal-Organic Framework with Enhanced CO2 Adsorption. Nanomaterials. 2019; 9(8):1063. https://doi.org/10.3390/nano9081063

Chicago/Turabian Style

Asghar, Aisha, Naseem Iqbal, Tayyaba Noor, Majid Ali, and Timothy L. Easun. 2019. "Efficient One-Pot Synthesis of a Hexamethylenetetramine-Doped Cu-BDC Metal-Organic Framework with Enhanced CO2 Adsorption" Nanomaterials 9, no. 8: 1063. https://doi.org/10.3390/nano9081063

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