DOI:
10.1039/C5RA26617C
(Paper)
RSC Adv., 2016,
6, 13749-13756
Amine post-functionalized POSS-based porous polymers exhibiting simultaneously enhanced porosity and carbon dioxide adsorption properties†
Received
13th December 2015
, Accepted 26th January 2016
First published on 28th January 2016
Abstract
A novel hybrid porous polymer (HPP-1) is synthesized using octavinylsilsequioxane and 2,7-dibromo-9-fluorenone as monomers via Heck reaction. Subsequently, HPP-1 is post-functionalized by the conversion of deliberately introduced free ketone moieties into amine functionalities. Compared with HPP-1, the resulting material, HPP-1-amine, showed enhanced CO2 uptake from 0.63 mmol g−1 (HPP-1) to 1.01 mmol g−1 (HPP-1-EDA, EDA = ethylenediamine) and 0.72 mmol g−1 (HPP-1-HDA, HDA = hexamethylenediamine) at 298 K and 1 bar. Furthermore, the porosity of HPP-1-amine is not compromised, but enhanced with the BET specific surface area increasing from 529 m2 g−1 (HPP-1) to 651 m2 g−1 (HPP-1-EDA) and 615 m2 g−1 (HPP-1-HDA). Compared with HPP-1, the selectivity of CO2 over N2 of HPP-1-EDA increases nearly twice while that of HPP-1-HDA slightly decreases, indicating their potential applications in CO2 storage and capture. Although the enhanced extent of CO2/N2 selectivity is not as high as other analogues, we provide a new possibility for post-synthetic amine functionalization of porous polymers for simultaneously enhancing porosity and CO2 adsorption properties. The retained porosity won't limit their applications in other areas, such as heavy metal ion adsorption, ion channels and catalysis, etc.
Introduction
Functional porous polymers are an emerging class of fascinating materials with considerable technological relevance because of their widespread application promise in many fields, such as gas storage, gas separation, heterogeneous catalysis, light-harvesting, sensors, and energy storage and conversion, etc.1–6 Admittedly, functionality plays a decisive role in their applications. Functionality is traditionally achieved by selecting functional monomers through direct polymerization methodology.1 For example, monomers bearing amino,7 carboxyl,7,8 hydroxyl,7,9 and metal–organic groups10 have been utilized as building blocks to construct novel porous polymers exhibiting significant improvement of the existing properties or new properties compared to analogous materials. Although functional monomers are easy to design and select, their synthetic procedures are usually tedious. Moreover, polymerization occasionally fails due to incompatibility between functional groups and catalysts in several metal-catalyzed reactions.
In contrast, another important strategy to achieve functionality, i.e., post-functionalization or post-synthetic modification (PSM) on an established porous polymers can overcome the disadvantage. While this approach has been well-developed in metal–organic frameworks (MOFs),11 the exploitation in porous polymers is relatively limited but has developed rapidly in recent years. Several reactions, including amination,12 nitration,13 click reactions,14 protonation15 and coordination,16 have been successfully applied to post-functionalize porous polymers. Similar to the former strategy, the incorporation of new functional groups into porous polymers by post-functionalization has led to materials with enhanced or specific functionality, such as enhanced CO2 adsorption and selectivity,12,13 catalytic activity16 and conductivity,17 etc. For example, Du et al. reported a class of post-modified polymers of intrinsic microporosity (PIMs) showing super-permeable characteristics and outstanding CO2 separation performance by [2 + 3] cycloaddition of nitrile group with NaN3 to form tetrazole rings.18 Bipyridine-containing conjugated microporous polymers (CMPs) have been post-metallated with rhenium, rhodium and iridium to form metal–organic CMPs showing excellent catalytic activity for reductive amination.19 Kiskan et al. reported post-functionalized phenolphthalein-based CMPs showing tunable CO2 adsorption and photochemical activity for the photopolymerization of methyl methacrylate by merely tuning the pH value, and the post-functionalization process is switchable.15 However, this strategy suffers an obvious drawback, i.e., the porosity will be compromised, even lost because of the occupation of pore volume by functional groups, which may limit the applications of these materials in other fields, such as heavy metal ions adsorption and catalysis.20
In the present study, we select a polyhedral oligomeric silsesquioxane (POSS) based hybrid porous polymer with a ketone functionality (HPP-1) as an established material for amine post-functionalization by the conversion of deliberately introduced free ketone moieties into amine functionalities (see Scheme 1). The selection of POSS-based porous polymer is motivated by specific characteristics of POSS units, including the ideal inorganic–organic hybrid structure, rigidity, polyfunctionality and high thermal stability, etc.21,22 The selected ketone units are derived from 2,7-dibromo-9-fluorenone (Br-FO), which is a commercially available monomer and economical. Moreover, the amine post-functionalization process, i.e., the imine condensation reaction between ketone and amines is facile and mild. We explored the effect of the modification on the pore structure of the network and found that the porosity of the resultant amine-functionalized materials (HPP-1-amine) is not compromised, even increase to some extent. Moreover, the introduction of amine groups in the networks leads to materials with enhanced CO2 capacity and CO2/N2 selectivity.
|
| Scheme 1 Synthetic routes of hybrid porous polymer, HPP-1 and amine post-functionalized products, HPP-1-EDA and HPP-1-HDA. (i) Pd(OAc)2/P(o-CH3Ph)3, DMF/Et3N, 100 °C, 48 h; (ii) formic acid, MeOH, 80 °C, 72 h. | |
Results and discussion
Synthesis and characterization
Scheme 1 shows the synthetic routes of amine post-functionalized POSS-based porous polymers. HPP-1 was first synthesized from octavinylsilsesquioxanes (OVS) and Br-FO via Heck reaction, which was performed at 100 °C in N,N-dimethyl formamide for 48 h in the presence of Pd(OAc)2/P(o-CH3Ph)3 as the catalyst.23,24 Then the amine post-functionalization process was performed at 80 °C with a catalytic amount of formic acid by dispersing HPP-1 in the amine–methanol mixture,25 thus yielding the amine-functionalized products, HPP-1-amine (amine = EDA, ethylenediamine; HDA, hexamethylenediamine).
The process was monitored by Fourier-transformed infrared spectroscopy (FT-IR) according to the change of the characteristic νCO band from the fluorenone moieties and νCN band after amine post-functionalization. Taking HPP-1-EDA as an example, the intensity of νCO band at ca. 1720 cm−1 largely decreased and the peak at ca. 1640 cm−1 assigned to νCN band appeared when the process time was 24 h. When the time was extended to 72 h, the νCO band with weak intensity could still be observed (Fig. 1 and S1†), indicating that the full conversion of the ketone in HPP-1 to the imine moiety in HPP-1-EDA didn't occur. This finding was also found in HPP-1-HDA (Fig. S2†). The conversion efficiencies are 68.6% and 52.2% for HPP-1-EDA and HPP-1-HDA, respectively, calculated from the intensity decrease degree of νCO band using the unchanged νCC band at 1603 cm−1 as the reference. We speculate that the relatively low conversion efficiencies can be caused by two reasons. The first one is that the imine condensation reaction in this study is a kind of heterogeneous reaction due to the insolubility of HPP-1 in common solvents, unlike the high efficiencies found in the homogeneous reactions for preparing small organic molecules.25 Another one is that organic amine molecules couldn't reach the ketone groups embedded in the closed pores, which may exist in the porous network. In addition, the characteristic Si–O–Si stretching vibration peaks for those materials were observed at ca. 1130 cm−1.
|
| Fig. 1 The FT-IR spectra enlarged from 2000 cm−1 to 800 cm−1 of HPP-1, HPP-1-EDA at 24 h and 72 h. | |
The reaction process and the identification of the local structures of HPP-1 and HPP-1-amine were further determined by the solid-state 13C CP/MAS NMR and 29Si NMR spectroscopy. The signal at δ = 191 ppm attributed to the ketone carbon atoms (CO) largely decreased, but can be still detected after amine functionalization, while a new signal at δ = 163 ppm assigned to the imine carbon atoms (CN) appeared (Fig. 2). These results revealed the amine-functionalization process successfully happened but the conversion was incomplete, consistent with the FT-TR results. The resonances arising from other carbon atoms can be ascribed as shown in Fig. 2.
|
| Fig. 2 Solid-state 13C CP/MAS NMR spectra of HPP-1, HPP-1-EDA and HPP-1-HDA. Asterisk at ca. 49 ppm denotes the residual methanol derived from the extraction process. | |
The 29Si NMR spectra showed three signals at δ = ∼−61, ∼−70 and ∼−77 ppm, which were attributed to the T1, T2 and T3 units [Tn: CSi(OSi)n(OH)3−n] (Fig. 3). Scheme 2 shows the possible frameworks of HPP-1 and HPP-1-amine (HPP-1-EDA as an example). For HPP-1, a few POSS cages collapsed during the synthesis and T1 and T2 units were formed. Such destruction could be attributed to the distortion of POSS units to link the rigid FO units and the presence of triethylamine (Et3N) and acid; Et3N served as the acid adsorbent and acid, i.e., HBr was produced in the Heck reaction. This finding is consistent with our previous reports.23,24 Moreover, it is found that such destruction continued to occur in the amine post-functionalization process and new T1 and T2 units were formed. This can be proven by the decreased T3/(T3 + T2 + T1) ratio from 0.58 (HPP-1) to 0.48 (HPP-1-EDA) and 0.50 (HPP-1-HDA) calculated from the relative intensity of the T3, T2 and T1 signals. The consequent destruction for HPP-1-amine was obviously due to the addition of alkylamines, which are a kind of organic base. In addition, the appearance of peak from −90 to −110 ppm may be ascribed to Qn (Si(OSi)n(OH)4−n) species26 and suggests that Si–C bonds have been partially cleaved. However, most of the POSS cages remained in the porous networks. Except FT-IR and NMR measurements, elemental analysis is another important means to prove the successful conversion; the nitrogen content increased from 0 (HPP-1) to 4.59 wt% and 3.50 wt% for HPP-1-EDA and HPP-1-HDA, respectively.
|
| Fig. 3 Solid-state 29Si NMR spectra of HPP-1, HPP-1-EDA and HPP-1-HDA. The T3/(T3 + T2 + T1) ratios were calculated from the relative intensity of the T3, T2 and T1 signals. | |
|
| Scheme 2 Possible frameworks of HPP-1 and HPP-1-amine (HPP-1-EDA as an example) showing the destruction of POSS units, the formation of T1 and T2 units and possible hydrogen bond interactions. | |
Thermal stability and morphology
The thermal stabilities of the polymers were determined by thermogravimetric analysis (TGA) under N2 at 10 °C min−1 (Fig. S3†). HPP-1 exhibited high thermal decomposition temperature (Td,5 wt%) of up to 545 °C, which is higher than other POSS-based porous materials.23,24,26,27 As expected, much lower decomposition temperatures were found in HPP-1-EDA and HPP-1-HDA with Td,5 wt% of ca. 330 °C and 300 °C. The weight loss occurring at <100 °C is attributed to some residual water/solvent in the porous networks, which can be coordinated to free amines through hydrogen bonding. After that, slower degradation could be attributed to the alkyl amine groups. This finding is consistent with a previous report.28 Because of initial loss of some residual water/solvent and alkyl amine groups, the char yield of HPP-1 was higher than those of HPP-1-EDA and HPP-1-HDA.
Consistent with other POSS-based porous polymers, HPP-1 and HPP-1-amine were amorphous and no long-range crystallographic order was observed in powder X-ray diffraction (PXRD) measurements (Fig. S4†). Field-emission scanning electron microscopy (FE-SEM) measurements revealed that there is no obvious change of morphologies with irregular shapes after amine post-functionalization (Fig. 4a and S5†). High-resolution transmission electron microscopy (HR-TEMS) images are shown in Fig. 4b and S6.† The translucent TEM images are indicative of porous, not dense, structures of materials with relatively uniform pore diameters but no evidence of long-range ordering. The results are consistent with other POSS-based porous materials.22–24,26,27
|
| Fig. 4 FE-SEM (a) and HR-TEM (b) images of HPP-1 as an example. | |
Porosity
The porosity of the polymers was investigated by nitrogen adsorption and desorption experiments at 77 K and Table 1 shows the detailed porosity of these polymers. All the polymers appeared as IUPAC type I N2 isotherms with some type IV isotherm characteristics at higher relative pressures (Fig. 5a). All the isotherms showed a sharp uptake at low relative pressure and a gradually increasing uptake at higher relative pressures, suggesting the presence of micropores and mesopores within the networks. The Brunauer–Emmett–Teller surface areas (SBET) of HPP-1, HPP-1-EDA and HPP-1-HDA were calculated to be 529, 651 and 615 m2 g−1. Total pore volumes (Vtotal) were 0.37, 0.46 and 0.47 cm3 g−1. It is significantly noted that the porosity increased along with amine introduction, unlike the compromised porosity found in other amine-tethered porous materials.12,28,29 We speculate that this unusual phenomenon can be explained by two reasons. First, new destruction of POSS cages happened during the amine-functionalized process evidenced by the decreased T3/(T3 + T2 + T1) ratios as mentioned above. The fact that the destruction contributing to increase the porosity has been proved by Chaikittisilp et al., obtaining a POSS-containing network, PSN-5, with an ultrahigh SBET of 2500 m2 g−1, representing the highest surface area for the siloxane-based porous materials.27 For those materials with no destruction of siloxane cleavage, dense silica and organic matrixes were formed to stabilize the porous framework and the surface areas of such materials are limited to a certain value. On the contrary, the destruction of the siloxane parts prevents the formation of dense silica and organic matrixes, leading to an increased surface area.
Table 1 Porosity data of HPP-1 and HPP-1-amine
Sample |
Surface area/m2 g−1 |
Pore volume/cm3 g−1 |
CO2 uptakee/mmol g−1 |
N2 uptakef/mmol g−1 |
Qstg/kJ mol−1 |
Sh |
SBETa |
Smicrob |
Vtotalc |
Vmicrod |
Vmicro/Vtotal |
273 K |
298 K |
298 K |
298 K |
Surface area calculated from N2 adsorption isotherm using the BET method. Microporous surface area calculated from N2 adsorption isotherm using t-plot method. Total pore volume calculated at P/P0 = 0.99. Micropore volume derived using the t-plot method based on the Halsey thickness equation. At 1 bar. At 1 bar. At zero coverage. Selectivity is calculated from Henry equation. |
HPP-1 |
529 |
275 |
0.37 |
0.14 |
0.38 |
1.15 |
0.63 |
0.034 |
34 |
27.7 |
HPP-1-EDA |
651 |
282 |
0.46 |
0.13 |
0.28 |
1.69 |
1.01 |
0.036 |
49 |
49.3 |
HPP-1-HDA |
615 |
224 |
0.47 |
0.11 |
0.23 |
1.23 |
0.72 |
0.035 |
40 |
25.2 |
|
| Fig. 5 (a) Nitrogen adsorption (closed symbols) and desorption (open symbols) isotherms for HPP-1, HPP-1-EDA and HPP-1-HDA. For clarify, HPP-1-EDA and HPP-1-HDA are shifted vertically by 150 cm3 g−1 and 300 cm3 g−1; (b) pore size distribution curves of HPP-1, HPP-1-EDA and HPP-1-HDA. | |
Second, free activity of alkyl amine chains in the pores may be restricted by hydrogen bond interactions. As shown in Scheme 2, during the formation of HPP-1 framework and amine post-functionalization process, several Si–OH groups were produced due to the destruction of POSS units and hydrogen bonds could be formed between hydrogen bond donors, including –NH2 and –OH groups, and hydrogen bond acceptors, including –NH2, –OH, and even imine and residual ketone groups. Possible hydrogen bonds included O–H⋯O(H)Si, O–H⋯N(H2)C, O–H⋯OC, (H)N–H⋯O(H)Si, (H)N–H⋯OC, (H)N–H⋯N(H2)C and (H)N–H⋯NC, etc. These various hydrogen bond interactions could efficiently limit the free activity of alkyl amine chains in the pore and retain more free pore volume. Unfortunately, because the products are cross-linked materials and insoluble in common solvents, it is difficult to prove the presence of hydrogen bonds determined by traditional techniques, such as IR and 1H NMR. However, the presence of hydrogen bonds have been widely found in many silanol compounds.30,31 For example, in our previous report, we found that the hydrogen bond interactions existed in incompletely condensed silsesquioxanes (POSS-mono-ol, POSS-diol and POSS-triol).31
Along with the enhanced surface area, the pore size also appeared no large change, in contrast to the reduced pore size occupied by the bulky amine groups found in other amine-tethering products.12,28,29 All the porous networks showed similar pore size distribution with major micropore diameters at 1.4 nm and 1.8 nm, and mesopore diameters at 2.8 nm and 3.8 nm before and after the amine-tethering process (Fig. 5b). Additionally, lower porosity for HPP-HDA than HPP-1-EDA is apparently due to the larger space volume of HDA than EDA.
Carbon dioxide sorption
To evaluate their potential applications in CO2 capture, CO2 adsorption experiments were carried out. As expected, the introduction of alkylamine groups in HPP-1 resulted in materials with enhanced CO2 sorption properties at low pressures (Fig. 6a and S8†). The CO2 uptakes of HPP-1, HPP-1-EDA and HPP-1-HDA were 1.15, 1.69 and 1.23 mmol g−1 at 273 K, and 0.63, 1.01 and 0.72 mmol g−1 at 298 K, which were measured up to 1 bar. Apparently, one factor leading to the enhancement is the increased porosity for the materials with higher surface area and pore volume after amine post-functionalization. Another vital factor is the isosteric heat (Qst) of CO2 adsorption, which represents the affinity between the sorbent and CO2. Indeed, the Qst at low coverage, calculated from CO2 isotherms collected at 273 K and 298 K employing the Clausius–Clapeyron equation, increased substantially from 34 kJ mol−1 for the original HPP-1 to 49 kJ mol−1 and 40 kJ mol−1 for the amine-functionalized HPP-1-EDA and HPP-1-HDA at zero coverage (Fig. 6b), further confirming the great efficiency of our amine-functionalization strategy on ketone sites. This finding is consistent with the successful linking the amine on aldehyde anchors.28 Moreover, unlike the classical amine supported materials (e.g., mesoporous silica and alumina),32 it is not necessary to heat for reactivating HPP-1-amine between variable temperature adsorption measurements, thus suggesting potentially diminishing energy penalty for adsorption-desorption cycles. The CO2 cyclic capacity of HPP-1-EDA and HPP-1-HDA was also determined by saturating CO2 up to 1 bar at 273 K followed by a high vacuum for 100 min at 80 °C. It is found that there was no apparent loss in capacity after 10 cycles (Fig. S9†), thus indicating that adsorption during each cycle was completed and these materials possessed good stability for regeneration.
|
| Fig. 6 (a) CO2 adsorption isotherms of HPP-1, HPP-1-EDA and HPP-1-HDA at 273 K (closed symbols) and 298 K (open symbols), and N2 adsorption of HPP-1-EDA at 298 K as an example; (b) isosteric heats of carbon dioxide adsorption of HPP-1, HPP-1-EDA and HPP-1-HDA. | |
The high Qst of CO2 adsorption for the HPP-1-amine motivated us to assess the potential as prospective separation agents because the isosteric heat also plays a major role in determining the adsorption selectivity. Thus we measured the N2 adsorption isotherms of the materials at 298 K. The results showed that HPP-1-amine adsorbed more N2 than HPP-1 at all pressure between 0 and 1.0 bar, which may be due to the additional polarizing sites and the enhancement in specific surface area. One of the common methods to calculate the gas adsorption selectivity involves the use of Henry equation.33 To calculate the selectivity factor (S) for adsorption of CO2 over N2 from the corresponding Henry's constants, a nonlinear fitting of the adsorption isotherm was performed using the Toth model (Fig. S10–S12†). The calculated values are listed in Table 1. Compared with the CO2/N2 selectivity of HPP-1 (S: 27.7), nearly twofold increase and slight decrease were found in those of HPP-1-EDA (S: 49.3) and HPP-1-HDA (S: 25.2), respectively. The values are higher or comparable to the corresponding values of some MOFs34 and nitrogen-rich porous organic polymers.35 For HPP-1-EDA, enhanced CO2/N2 selectivity confirms that our unique amine functionalization process is an effective mean to not only improve the CO2 capacity but also promote its selectivity of CO2 over N2. This finding is consistent with other amine-functionalized porous materials.12,28,29 However, the enhanced extent of the selectivity is not as high as other analogues; for example, some of them even showed a more than tenfold selectivity enhancement.12,29 For HPP-1-HDA, its CO2 and N2 capacities were also improved. However, the enhanced extent of CO2 capacity doesn't exceed that of N2 capacity, thus leading to a slight decline of selectivity. We speculated these results can be explained by two reasons. First, the surface areas after amine functionalization were improved and the resulting products adsorb more N2 as shown in the N2 isotherm at 298 K while adsorbing more CO2. Second, the conversion of ketone into imine is incomplete (68.6% and 52.2% for HPP-1-EDA and HPP-1-HDA) and the amine loading capacity is not very high. However, there is still some upside potential for the conversion by optimizing the reaction, which may promote the selectivity. Additionally, the surface areas of HPP-1-EDA and HPP-1-HDA are still relatively low, which is not also beneficial to adsorb CO2.
Although the values of CO2 selectivity are lower than some amine-functionalized porous materials in recent reports,12,29 herein we provide a new possibility for post-synthetic amine functionalization of porous materials for improving the CO2 capacity and selectivity without compromising the porosity. This retaining porosity won't limit the resulting materials to be applied in other areas, such as heavy metal ion adsorption, ion channels and catalysis, etc. For example, for the heavy metal ion adsorption using porous materials as solid absorbents, two important factors should be considered: (i) the species and content of functional groups representing the potential loading capacity of the metal ions; and (ii) the porosity indicating the ease or difficulty of metal ions accessing into the pore.20 For common post-functionalized porous materials, the compromised porosity made the process of metal ions accessing into the pores more difficult. On the contrast, herein the enhanced surface area and pore volume of our amine-functionalized materials might make the process easier and thus effective heavy metal ions capacity would be achieved. It is similar when using in catalysis. Current materials could be used as support materials to load metal ion (such as palladium) and thus act as catalyst. It is known that the catalytic efficiency commonly depend on the content of metal ion. The retaining porosity also might make the process of metal ion loading on the networks easier and lead to higher metal ion loading capacity.
Conclusions
In summary, we reported novel amine post-functionalized POSS-based porous polymers (HPP-1-amine) by the conversion of deliberately introduced free ketone moieties into amine functionalities using a hybrid porous polymer (HPP-1) containing POSS and fluorenone units as the support material, which is synthesized based on octavinylsilsequioxane and 2,7-dibromo-9-fluorenone via Heck reaction. We found that the resultant materials exhibited simultaneously enhanced carbon dioxide adsorption properties and enhanced porosity, unlike the compromised porosity found in other amine post-functionalized porous materials. The CO2 uptake increased from 0.63 mmol g−1 (HPP-1) to 1.01 mmol g−1 (HPP-1-EDA) at 298 K and 1 bar, while the BET specific surface area increased from 529 m2 g−1 (HPP-1) to 651 m2 g−1 (HPP-1-EDA). Compared with HPP-1, the selectivity of CO2 over N2 of HPP-1-EDA increased nearly twice while that of HPP-1-HDA slightly decreased, indicating their potential applications in CO2 capture. Although the enhanced extent of CO2 selectivity is not as high as other analogues, the retaining porosity won't limit the applications of the resulting materials in other areas. Further investigations will focus on three aspects: (i) increasing the porosity of the starting porous polymers containing POSS and fluorenone units by altering the monomer species and reaction conditions; (ii) optimizing the amine functionalization method and enhancing the amine loading capacity; (iii) utilizing the amine-functionalized porous polymers in various areas, including CO2 capture, heavy metal ion adsorption, ion channels and catalysis, etc.
Experimental section
Materials
Unless otherwise noted, all reagents were obtained from commercial suppliers and used without further purification. OVS was synthesized by the previous report.36 N,N-Dimethylformamide (DMF) was first dried over CaH2 at 80 °C for 12 h, distilled under vacuum pressure and stored with 4 Å molecule sieves prior to use. Triethylamine (Et3N) was dried over CaH2 and used freshly.
Characterization
Fourier transform infrared (FTIR) spectra was recorded on a Bruker Tensor27 spectrophotometer. Solid-state 13C cross-polarization/magic-angle-spinning (CP/MAS) NMR and 29Si MAS NMR spectra were performed on Bruker AVANCE-500 NMR Spectrometer operating at a magnetic field strength of 9.4 T. The resonance frequencies at this field strength were 125 and 99 MHz for 13C NMR and 29Si NMR, respectively. A Chemagnetics 5 mm triple-resonance MAS probe was used to acquire 13C and 29Si NMR spectra. 29Si MAS NMR spectra with high power proton decoupling were recorded using a π/2 pulse length of 5 μs, a recycle delay of 120 s and a spinning rate of 5 kHz. Elemental analyses were conducted using an Elementar vario EL III elemental analyzer.
Field-emission scanning electron microscopy (FE-SEM) experiments were performed by using HITACHI S4800 Spectrometer. The high-resolution transmission electron microscopy (HR-TEM) experiments were performed by using a JEM 2100 electron microscope (JEOL, Japan) with an acceleration voltage of 200 kV.
Thermogravimetric analyses were performed with a Mettler Toledo SDTA-854 TGA system in nitrogen at a heating rate of 10 °C min−1 to 800 °C. Powder X-ray diffraction (PXRD) were performed using a Riguku D/MAX 2550 diffractometer under Cu-Kα radiation, 40 kV, 200 mA with a scanning rate of 10° min−1 (2θ).
Nitrogen sorption isotherm measurements were performed on a Micro Meritics surface area and pore size analyzer. Before measurement, samples were degassed at 100 °C for least 12 h. A sample of ca. 100 mg and a UHP-grade nitrogen (99.999%) gas source were used in the nitrogen sorption measurements at 77 K and collected on a Quantachrome Quadrasorb apparatus. BET surface areas were determined over a P/P0 range from 0.01 to 0.20. Nonlocal density functional theory (NL-DFT) pore size distributions were determined using the carbon/slit-cylindrical pore mode of the Quadrawin software. Carbon dioxide (CO2) adsorption isotherms at 298 K and 273 K, and nitrogen adsorption isotherms at 298 K were measured on a Micrometrics ASAP 2020. Prior to the measurements, the samples were degassed at 150 °C for at least 10 h.
Synthesis of hybrid porous polymer, HPP-1
OVS (632 mg, 1 mmol), palladium acetate (90 mg, 0.4 mmol) and tris(2-methylphenyl)phosphine (193 mg, 0.94 mmol) were dissolved in a DMF/Et3N solution (45 mL/15 mL) under argon. The mixture was stirred and bubbled by argon under stirring for 0.5 h at room temperature and 2,7-dibromo-9-fluorenone (Br-FO) (1.35 g, 4 mmol) was added. The resulting solution was then heated at 100 °C for 48 h and cooled to room temperature. The mixture was filtered and the residue was washed with THF, chloroform, water, methanol and acetone. Further purification of the sample was carried out by extraction with THF for 24 h and methanol for 24 h. The product was recovered by filtration, and dried under reduced pressure at 70 °C for 48 h. HPP-1 was afforded as a red-brown solid (1.47 g). Yield: 109%. The yield was calculated based on hypothetical 100% polycondensation. Elemental analysis calc. (wt%) for C68H40Si8O12: C 61.05, H 3.01; found C 54.70, H 3.64.
Synthesis of HPP-1-EDA
HPP-1 (0.30 g), ethylenediamine (3.00 g) and two drops of formic acid were added into 50 mL of MeOH. The resulting mixture was stirred and refluxed for 72 h. After cooling to room temperature, the mixture was filtrated and the residue was washed with MeOH three times. The product was dried under vacuum at 70 °C for 24 h and afforded as a brick-red solid (0.32 g). Elemental analysis: found C 58.45, H 4.46, N 4.59.
Synthesis of HPP-1-HDA
The synthesis procedure of HPP-1-HDA was similar to that of HPP-1-EDA except that ethylenediamine (3.00 g) was replaced by hexamethylenediamine (3.00 g). Elemental analysis: found C 60.19, H 4.91, N 3.50.
Acknowledgements
This research was supported by the National Natural Science Foundation of China (21502105, 21574075, 21274080 and 21274081), the National Science Foundation of Shandong Province (2015ZRE27053 and ZR2015BQ008), Shandong special fund for independent innovation and achievements transformation (No. 2014ZZCX01101) and the Fundamental Research Funds of Shandong University (2014GN009).
References
- D. Wu, F. Xu, B. Sun, R. Fu, H. He and K. Matyjaszewski, Chem. Rev., 2012, 112, 3959–4015 CrossRef CAS PubMed ; R. Dawson, A. I. Cooper and D. J. Adams, Prog. Polym. Sci., 2012, 37, 530–563 CrossRef ; Y. Xu, S. Jin, H. Xu, A. Nagai and D. Jiang, Chem. Soc. Rev., 2013, 42, 8012–8031 RSC ; T. Muller and S. Bräse, RSC Adv., 2014, 4, 6886–6907 RSC ; U. H. F. Bunz, K. Seehafer, F. L. Geyer, M. Bender, I. Braun, E. Smarsly and J. Freudenberg, Macromol. Rapid Commun., 2014, 35, 1466–1496 CrossRef PubMed ; P. J. Waller, F. Gandara and O. M. Yaghi, Acc. Chem. Res., 2015, 48, 3053–3063 CrossRef PubMed .
- A. I. Cooper, Adv. Mater., 2009, 21, 1291–1295 CrossRef CAS ; X. Du, Y. L. Sun, B. E. Tan, Q. F. Teng, X. J. Yao, C. Y. Sun and W. Wang, Chem. Commun., 2010, 46, 970–972 RSC ; A. Thomas, Angew. Chem., Int. Ed., 2010, 49, 8328–8344 CrossRef PubMed ; J. T. A. Jones, T. Hasell, X. Wu, J. Bacsa, K. E. Jelfs, M. Schmidtmann, S. Y. Chong, D. J. Adams, A. Trewin, F. Schiffman, F. Cora, B. Slater, A. Steiner, G. M. Day and A. I. Cooper, Nature, 2011, 474, 367–371 CrossRef PubMed .
- A. Kumar, D. G. Madden, M. Lusi, K.-J. Chen, E. A. Daniels, T. Curtin, J. J. Perry and M. J. Zaworotko, Angew. Chem., Int. Ed., 2015, 54, 14372–14377 CrossRef CAS PubMed ; Z. Xiang, R. Mercado, J. M. Huck, H. Wang, Z. Guo, W. Wang, D. Cao, M. Haranczyk and B. Smit, J. Am. Chem. Soc., 2015, 137, 13301–13307 CrossRef PubMed ; C. H. Lau, K. Konstas, A. W. Thornton, A. C. Y. Liu, S. Mudie, D. F. Kennedy, S. C. Howard, A. J. Hill and M. R. Hill, Angew. Chem., Int. Ed., 2015, 54, 2669–2673 CrossRef PubMed ; Y. Liao, J. Weber and C. F. J. Faul, Macromolecules, 2015, 48, 2064–2073 CrossRef .
- M. Rose, ChemCatChem, 2014, 6, 1166–1182 Search PubMed ; Q. Sun, Z. Dai, X. Meng, L. Wang and F.-S. Xiao, ACS Catal., 2015, 5, 4556–4567 CrossRef CAS ; T.-T. Liu, Z.-J. Lin, P.-C. Shi, T. Ma, Y.-B. Huang and R. Cao, ChemCatChem, 2015, 7, 2340–2345 CrossRef .
- K. Zhang, B. Tieke, F. Vilela and P. J. Skabara, Macromol. Rapid Commun., 2011, 32, 825–830 CrossRef CAS PubMed ; A. Patra and U. Scherf, Chem.–Eur. J., 2012, 18, 10074–10080 CrossRef PubMed ; X. M. Liu, Y. H. Xu and D. L. Jiang, J. Am. Chem. Soc., 2012, 134, 8738–8741 CrossRef PubMed ; J.-X. Jiang, A. Trewin, D. J. Adams and A. I. Cooper, Chem. Sci., 2011, 2, 1777–1781 RSC ; J. L. Novotney and W. R. Dichtel, ACS Macro Lett., 2013, 2, 423–426 CrossRef ; D. Wang, W. Yang, L. Li, S. Feng and H. Liu, RSC Adv., 2014, 4, 59877–59884 RSC ; W. Yang, X. Jiang and H. Liu, RSC Adv., 2015, 5, 12800–12806 RSC .
- Y. Kou, Y. Xu, Z. Guo and D. Jiang, Angew. Chem., Int. Ed., 2011, 50, 8753–8757 CrossRef CAS PubMed ; X. Zhuang, F. Zhang, D. Wu, N. Forler, H. Liang, M. Wagner, D. Gehrig, M. R. Hansen, F. Laquai and X. Feng, Angew. Chem., Int. Ed., 2013, 52, 9668–9672 CrossRef PubMed .
- R. Dawson, D. J. Adams and A. I. Cooper, Chem. Sci., 2011, 2, 1173–1177 RSC .
- J. Weber, N. Y. Du and M. D. Guiver, Macromolecules, 2011, 44, 1763–1767 CrossRef CAS .
- A. P. Katsoulidis and M. G. Kanatzidis, Chem. Mater., 2011, 23, 1818–1824 CrossRef CAS ; S.-H. Jia, X. Ding, H.-T. Yu and B.-H. Han, RSC Adv., 2015, 5, 71095–71101 RSC .
- Z. Xie, C. Wang, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2011, 133, 2056–2059 CrossRef CAS PubMed ; R. K. Totten, Y.-S. Kim, M. H. Weston, O. K. Farha, J. T. Hupp and S. T. Nguyen, J. Am. Chem. Soc., 2013, 135, 11720–11723 CrossRef PubMed .
- Z. Wang and S. M. Cohen, Chem. Soc. Rev., 2009, 38, 1315–1329 RSC .
- W. Lu, J. P. Sculley, D. Yuan, R. Krishna, Z. Wei and H.-C. Zhou, Angew. Chem., Int. Ed., 2012, 51, 7480–7484 CrossRef CAS PubMed ; S. Sung and M. P. Suh, J. Mater. Chem. A, 2014, 2, 13245–13249 Search PubMed ; W. Lu, M. Bosch, D. Yuan and H.-C. Zhou, ChemSusChem, 2015, 8, 433–438 CrossRef PubMed ; B. Satilmis, M. N. Alnajrani and P. M. Budd, Macromolecules, 2015, 48, 5663–5669 CrossRef ; C. H. Lau, K. Konstas, C. M. Doherty, S. Kanehashi, B. Ozcelik, S. E. Kentish, A. J. Hill and M. R. Hill, Chem. Mater., 2015, 27, 4756–4762 CrossRef ; D. Lee, C. Zhang and H. Gao, Macromol. Chem. Phys., 2015, 216, 489–494 CrossRef .
- C. Shen and Z. Wang, J. Phys. Chem. C, 2014, 118, 17585–17593 CAS .
- B. Kiskan and J. Weber, ACS Macro. Lett., 2012, 1, 37–40 CrossRef CAS ; H. Lim and J. Y. Chang, Macromolecules, 2010, 43, 6943–6945 CrossRef .
- B. Kiskan, M. Antonietti and J. Weber, Macromolecules, 2012, 45, 1356–1361 CrossRef CAS .
- P. Zhang, Z.-A. Qiao, X. Jiang, G. M. Veith and S. Dai, Nano Lett., 2015, 15, 823–828 CrossRef CAS PubMed ; A. M. Shultz, O. K. Farha, J. T. Hupp and S. T. Nguyen, Chem. Sci., 2011, 2, 686–689 RSC .
- H. Bildirir, I. Osken, T. Ozturk and A. Thomas, Chem.–Eur. J., 2015, 21, 9306–9311 CrossRef CAS PubMed .
- N. Du, H. B. Park, G. P. Robertson, M. M. Dal-Cin, T. Visser, L. Scoles and M. D. Guiver, Nat. Mater., 2011, 10, 372–375 CrossRef CAS PubMed .
- J.-X. Jiang, C. Wang, A. Laybourn, T. Hasell, R. Clowes, Y. Z. Khimyak, J. Xiao, S. J. Higgins, D. J. Adams and A. I. Cooper, Angew. Chem., Int. Ed., 2011, 50, 1072–1075 CrossRef CAS PubMed .
- N. Saman, K. Johari and H. Mat, Ind. Eng. Chem. Res., 2014, 53, 1225–1233 CrossRef CAS ; H. Cui, Y. Qian, Q. Li, Z. Wei and J. Zhai, Appl. Clay Sci., 2013, 72, 84–90 CrossRef .
- R. M. Laine and M. F. Roll, Macromolecules, 2011, 44, 1073–1109 CrossRef CAS ; D. B. Cordes, P. D. Lickiss and F. Rataboul, Chem. Rev., 2010, 110, 2081–2173 CrossRef PubMed .
- W. Chaikittisilp, A. Sugawara, A. Shimojima and T. Okubo, Chem. Mater., 2010, 22, 4841–4843 CrossRef CAS ; W. Chaikittisilp, A. Sugawara, A. Shimojima and T. Okubo, Chem.–Eur. J., 2010, 16, 6006–6014 CrossRef PubMed ; M. F. Roll, J. W. Kampf, Y. Kim, E. Yi and R. M. Laine, J. Am. Chem. Soc., 2010, 132, 10171–10183 CrossRef PubMed ; I. Nischang, O. Brüggemann and I. Teasdale, Angew. Chem., Int. Ed., 2011, 50, 4592–4596 CrossRef PubMed ; Y. Wu, D. Wang, L. Li, W. Yang, S. Feng and H. Liu, J. Mater. Chem. A, 2014, 2, 2160–2167 Search PubMed .
- D. Wang, L. Xue, L. Li, B. Deng, S. Feng, H. Liu and X. Zhao, Macromol. Rapid Commun., 2013, 34, 861–866 CrossRef CAS PubMed .
- D. Wang, W. Yang, L. Li, X. Zhao, S. Feng and H. Liu, J. Mater. Chem. A, 2013, 1, 13549–13558 Search PubMed ; D. Wang, W. Yang, S. Feng and H. Liu, Polym. Chem., 2014, 5, 3634–3642 RSC .
- P. Papanikolaou, J. Mohanraj, A. Czapik, M. Gdaniec, G. Accorsi and P. Akrivos, Dalton Trans., 2013, 42, 3357–3365 RSC ; K. Venkatesan, S. Dhivya, J. Rethavathi and S. Narasimhan, J. Chem. Pharm. Res., 2012, 4, 4477–4483 Search PubMed .
- Y. Peng, T. Ben, J. Xu, M. Xue, X. Jing, F. Deng, S. Qiu and G. Zhu, Dalton Trans., 2011, 40, 2720–2724 RSC .
- W. Chaikittisilp, M. Kubo, T. Moteki, A. Sugawara-Narutaki, A. Shimojima and T. Okubo, J. Am. Chem. Soc., 2011, 133, 13832–13835 CrossRef CAS PubMed .
- V. Guillerm, Ł. J. Weseliński, M. Alkordi, M. I. H. Mohideen, Y. Belmabkhout, A. J. Cairns and M. Eddaoudi, Chem. Commun., 2014, 50, 1937–1940 RSC .
- T. M. McDonald, D. M. D'Alessandro, R. Krishna and J. R. Long, Chem. Sci., 2011, 2, 2022–2028 RSC ; N. Planas, A. L. Dzubak, R. Poloni, L.-C. Lin, A. McManus, T. M. McDonald, J. B. Neaton, J. R. Long, B. Smit and L. Gagliardi, J. Am. Chem. Soc., 2013, 135, 7402–7405 CrossRef CAS PubMed ; Q. Yan, Y. Lin, C. Kong and L. Chen, Chem. Commun., 2013, 49, 6873–6875 RSC ; Y. Lin, H. Lin, H. Wang, Y. Suo, B. Li, C. Kong and L. Chen, J. Mater. Chem. A, 2014, 2, 14658–14665 Search PubMed .
- M. Fukawa, T. Sato and Y. Kabe, Chem. Commun., 2015, 51, 14746–14749 RSC ; Y. Kawakami, M. Fukawa, A. Yanase, Y. Furukawa, Y. Nagata, E. Suzuki, T. Horikawa and Y. Kabe, J. Organomet. Chem., 2015, 799–800, 265–272 CrossRef CAS ; S.-i. Kondo, N. Okada, R. Tanaka, M. Yamamura and M. Unno, Tetrahedron Lett., 2009, 50, 2754–2757 CrossRef ; S. Spirk, M. Nieger, F. Belaj and R. Pietschnig, Dalton Trans., 2009, 163–167 RSC .
- H. Liu, S.-i. Kondo, R. Tanaka, H. Oku and M. Unno, J. Organomet. Chem., 2008, 693, 1301–1308 CrossRef CAS .
- B. Dutcher, M. Fan and A. G. Russell, ACS Appl. Mater. Interfaces, 2015, 7, 2137–2148 Search PubMed ; C. Chen, J. Kim and W.-S. Ahn, Korean J. Chem. Eng., 2014, 31, 1919–1934 CrossRef CAS .
- E. Neofotistou, C. D. Malliakas and P. N. Trikalitis, Chem.–Eur. J., 2009, 15, 4523–4527 CrossRef CAS PubMed ; B. Wang, A. P. Côté, H. Furukawa, M. O'Keeffe and O. M. Yaghi, Nature, 2008, 453, 207–211 CrossRef PubMed .
- R. Banerjee, H. Furukawa, D. Britt, C. Knobler, M. O'Keeffe and O. M. Yaghi, J. Am. Chem. Soc., 2009, 131, 3875–3877 CrossRef CAS PubMed ; A. Demessence, D. M. D'Alessandro, M. L. Foo and J. R. Long, J. Am. Chem. Soc., 2009, 131, 8784–8786 CrossRef PubMed .
- S. Hug, M. B. Mesch, H. Oh, N. Popp, M. Hirscher, J. Senkerd and B. V. Lotsch, J. Mater. Chem. A, 2014, 2, 5928–5936 CAS .
- R. H. Baney, M. Itoh, A. Sakakibara and T. Suzuki, Chem. Rev., 1995, 95, 1409–1430 CrossRef CAS .
Footnote |
† Electronic supplementary information (ESI) available: FT-IR spectra, TGA curves, XRD patterns and BET plots of HPP-1, HPP-EDA and HPP-1-HDA, FE-SEM and HR-TEM images of HPP-EDA and HPP-1-HDA, CO2 adsorption–desorption isotherms of HPP-1-EDA and HPP-1-HDA at 273 K and 298 K, ten cycles of CO2 uptakes of HPP-1-EDA and HPP-1-HDA at 273 K, Toth model fitting of CO2 and N2 adsorption isotherms of HPP-1, HPP-EDA and HPP-1-HDA at 298 K. See DOI: 10.1039/c5ra26617c |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.