DOI:
10.1039/C6RA04077B
(Paper)
RSC Adv., 2016,
6, 30048-30055
New nitrogen-rich azo-bridged porphyrin-conjugated microporous networks for high performance of gas capture and storage†
Received
15th February 2016
, Accepted 15th March 2016
First published on 18th March 2016
Abstract
A series of new conjugated microporous polymers (Azo-1, Azo-2 and Azo-3) based on a nitrogen-rich porphyrin building unit and an azo bond linkage were synthesized by KOH assisted condensation. These materials were characterized by Fourier transform infrared spectroscopy (FT-IR), solid-state 13C NMR, XPS, scanning electron microscopy (SEM), high-resolution transmission electron microscopy (TEM), X-ray diffraction (XRD), thermogravimetric analysis (TGA), and tested for gas (N2, CO2 and H2) adsorption. It was revealed that the azos presented the formation of porous polymer networks affording amorphous particles with rough surfaces and irregular morphology with excellent thermal stability under nitrogen conditions. The Brunauer–Emmett–Teller (BET) model of the N2 adsorption gave apparent surface area ranges of 520–675 m2 g−1. The results from non-local density functional theory (NL-DFT) calculations suggested a pore size distribution between 1.6 and 4.0 nm. The gas (CO2, H2) adsorption isotherms demonstrated outstanding CO2 uptake up to 17.5 wt% (3.98 mmol g−1 for Azo-2) and moderate H2 storage. The isosteric heats of adsorption (Qst) are high, with values of 36–37 kJ mol−1 for the azo polymers. Moreover, the azo-polymer networks exhibited excellent selectivity with CO2/N2 up to 64.3 for Azo-2 at 273 K/1 bar. It was suggested that the nitrogen-rich active sites of the polymers play an important role for CO2 capture and storage.
Introduction
Porous polymers1–3 generally possess permanent porosity, low mass densities, synthetic diversification, and high physicochemical stability, which make them highly competitive in gas capture, storage and separation applications.4–6 Especially for H2 storage7–9 and CO2 storage and separation,10–12 porous polymers are considered as promising materials. Increasing CO2 concentration in the atmosphere mainly caused by the rapid consumption of fossil fuels13 has partly led to the global climate change and some environmental issues,14 which have drawn great attentions and concerns. Consequently, CO2 capture and storage is of great interest to meet the energy and environmental demands.
In order to enhance the adsorption interactions and selectivity of CO2 capture, nitrogen has been doped into many porous materials, such as carbon-based materials,15–19 microporous polymer networks.20 Great efforts have been taken to introduce functional amines into the carbon networks via organic grafting or direct heat treatment, however, these technologies by post-introduction of amines suffer from some drawbacks of time consuming, high-temperature need, and/or high cost. Recently, N-rich porous polymer networks demonstrate excellence in CO2 caption and conversion. Coskun group21 reported a nanoporous polymers incorporating N-heterocyclic carbenes for CO2 capture and conversion. The resulting polymer networks possess a Brunauer–Emmett–Teller (BET) surface area of 475 m2 g−1 and CO2 uptake of ∼9.55 wt% (2.1 mmol g−1) at 298 K/1 bar.
Conjugated micro- and mesoporous polymers (CMPs) that permit the linking of building blocks in a π-conjugated fashion are a class of porous materials with 3D or 2D networks and high surface areas,22 which have attracted much attention. CMPs is an interesting platform for the molecular design of porous skeletons because they allow to manipulate their networks and to tune the porous parameters. One of the remarkable characteristics of CMPs is their crosslinked 3D molecular skeletons.23 These structural features provide capability of mass transport of the gases into the strictly microporous insides.
Up to date, one of the most characteristic features of CMPs is the rather broad diversity of π-units.22 Molecular building blocks, ranging from simple phenyl units to extended arenes, heterocyclic aromatic units and large macrocycles, have been successfully exploited for the synthesis of porous polymers. Recently, porphyrin and its derivatives have become one of the major building blocks24–26 for porous polymer because they possess large planar π-conjugated structure, peculiar electronic properties, good photochemical effect, thermal stability,27,28 and so on. Liu, et al.29 reported a class of metal functional microporous covalent triazine network prepared using a metalloporphyrin for enhanced CO2 uptake. It presented a very high BET specific area up to ∼1500 m2 g−1, and excellent CO2 uptake capability of 3.17 mmol g−1, (13.9 wt%). More recently, a porphyrin polymer with benzimidazole linkages30 was synthesized to form a rigid amorphous network that exhibited selective CO2 adsorption with capability of 12.1 wt% (2.76 mmol g−1) and a CO2/N2 selectivity of 72 at 273 K/1 bar operation, evident the nanoporous porphyrin networks for gas capture and storage.
Furthermore, some azo-bridged porous polymers have been developed for CO2 capturing properties. Azo-linked porous organic framework (a 2D layered Azo-POF-2) exhibited high CO2/N2 selectivity of 76 at 273 K.31 Arab et al. reported32 a series of new azo-linked polymers (ALPs), with surface area of 412–801 m2 g−1, demonstrated CO2 uptake capacities of up to 2.94 mmol g−1 at 298 K/1 bar and good CO2/N2 selectivity (56). Another class of ALPs, synthesized by homocoupling of aniline-like building units in the presence of copper(I) bromide and pyridine, possessing higher surface area of 862–1235 m2 g−1, presented higher CO2 uptake capacities of up to 5.37 mmol g−1 at 237 K/1 bar.33 The investigation shows that the porosity parameters (surface area, pore volume and pore size) play key roles in CO2 uptake capacity and CO2 selectivity.
In this paper, combining the advantages of CMPs with the porphyrin units, a series of unprecedented, electron-rich conjugated microporous polymers (Azo-1, Azo-2, and Azo-3) based on nitrogen-rich porphyrin building units and azo bond linkage was synthesized by KOH assisted condensation. The azo-polymer networks were characterized by solid-state NMR spectroscopy, infrared spectroscopy, and elemental analysis to obtain their structural and chemical composition. Thermogravimetric analysis showed that all azos networks exhibit a high thermal stability under N2. Despite their moderate specific BET equivalent surface areas (520–675 m2 g−1), they featured remarkable CO2 uptakes up to 17.52 wt% (3.98 mmol g−1) and excellent selectivity (CO2/N2 ratio 15/1) at 273 K/1 bar, and moderate capability for H2 storage (1.15 wt%, 5.75 mmol g−1) at 77 K/1 bar. High CO2 adsorption is expected to occur because of the Lewis acid–base interactions between the active electron-rich aromatic constituents and the electron-poor CO2 molecules.
Experimental
Materials
Pyridine (99.5%), acetic anhydride (98.5%), potassium hydroxide (90%), and propionic acid (99%) were purchased from Sinopharm Chemical Reagent Co. Ltd. p-Nitrobenzaldehyde (99%) was obtained from Adamas. p-Phenylenediamine (99%) was purchased from Sigma-Aldrich. Melamine (99%) and N,N-dimethylformamide (DMF) (99.8%) were obtained from J&K Chemical. Pyrrole (99%) and benzidine (98%) were purchased from Energy Chemical. Pyrrole was freshly distilled before used, and the other chemicals were used as received.
Synthetic procedures
Synthesis of 5,10,15,20-tetrakis(4-nitrophenyl)-21H,23H-porphine (TNPP). TNNP synthesis was followed the experimental procedure in literature.34 p-Nitrobenzaldehyde (22.0 g, 145 mmol) and acetic anhydride (24.0 mL, 254 mmol) were dissolved in propionic acid (600 mL). The solution was then refluxed, to which freshly distilled pyrrole (10.0 mL, 144 mmol) was slowly added. After refluxing for 30 min, the resulting mixture was cooled to the room temperature. Then the precipitate was collected by filtration, washed with distilled water and methanol, and dried under vacuum. The resulting powder was dissolved in pyridine (160 mL) which was refluxed for 1 h. After cooling, the precipitate was collected by filtration and washed with acetone to give 4.5 g TNPP as a purple crystal in 15.6% yield. Elem. anal. calcd for C44H26N8O8 (%): C, 66.50; N, 16.11. Found: C, 65.62; N, 13.38. HRMS (EI-MS): C44H26N8O8 for [M + H]+, calculated 795.1874; found 795.1947.
Synthesis of Azo-1. TNPP (0.3176 g, 0.4 mmol), p-phenylenediamine (0.0865 g, 0.8 mmol) and KOH (0.22 g, 3.94 mmol) were dissolved in DMF (25 mL). The reaction mixture was heated to 150 °C and stirred for 24 h under a nitrogen atmosphere. The reaction mixture was cooled down to room temperature, added with 150 mL of distilled water and stirred for 1 h. Black precipitate was filtered off and washed several times with distilled water, acetone and THF in order to remove any unreacted monomers and KOH. Subsequently, black precipitates were dried at 60 °C under vacuum overnight to yield Azo-1: 0.26 g. Yield was 64% with elemental analysis of C: 72.08; H: 4.50; N: 14.92 (%).
Synthesis of Azo-2 and Azo-3. The procedure is similar to the synthetic route of Azo-1. The details of the procedure and product data are provided in ESI.†
Characterization methods and testing of gas sorption
Characterization methods. Fourier transform infrared spectroscopy (FT-IR) was performed on a Spectrum 100 (Perkin Elmer, Inc., USA) spectrometer with a scan range of 4000–400 cm−1. The sample powders were pulverized with KBr, and pressed into disks. Solid-state 13C cross polarization/magic angle spinning nuclear magnetic resonance (CP/MAS NMR) analysis was conducted on a Bruker AVANCE III 300 Spectrometer. Meanwhile, the cross polarization/magic angle spinning nuclear magnetic resonance (CP/MAS NMR) experiments were achieved by incorporating the total side-band suppression (TOSS) pulse. Samples were spun at 5 kHz in 4 mm rotor within a MAS probe. An acquisition time of 20 ms, a contact time of 1 ms and a 6.5 μs pre-scan delay were used. The recycle time was 2 s to obtain fully relaxed spectra. Chemical shifts were externally referenced to adamantane at 38.48 ppm. X-ray photoelectron spectroscopy (XPS) was performed using an AXIS Ultra DLD system (Kratos Co., Japan) with Al Kα radiation as the X-ray source. Thermogravimetric analysis (TGA) of the samples was undertaken using a Q5000IR (TA Instruments, USA) thermogravimetric analyzer with a heating rate of 20 °C min−1 from room temperature up to 800 °C under nitrogen flow. Scanning electron microscopy (SEM) was performed using an FEI Sirion-200 (FEI Co., USA) field emission scanning electron microscope. Transmission electron microscopy (TEM) observations were carried out with a JEOL-2100 (JEOL Ltd., Japan) electron microscope at an operating voltage of 200 kV. X-ray diffraction (XRD) measurements were performed using a Rigaku D/Max 2500 X-ray diffractometer with Cu Kα radiation (k = 1.54 Å) at a generator voltage of 40 kV and a generator current of 50 mA with a scanning speed of 2° min−1 from 2° to 35°. HRMS was performed on Q-Tof Premier (Waters, USA).
Testing of gas sorption. The gas sorption isotherms were measured via an Autosorb-iQA3200-4 (ASIQA3200-4) sorption analyzer (Quantatech Co., USA). Nitrogen adsorption/desorption isotherms were measured at 77 K using a Quantachrome ASIQA3200-4 gas sorption analyser. Carbon dioxide (CO2) measurements were measured at 273 and 298 K using a Quantachrome ASIQA3200-4 extended sorption analyser fitted with a chiller circulator dewar. Hydrogen (H2) sorption isotherm measurements were analyzer at 77 and 87 K. Prior to the measurement, the samples were degassed in vacuum 150 °C for 4 h. The isosteric heats of adsorption were calculated using the standard calculation routines provided with the ASIQA3200-4 software. The pore size distribution was calculated from non-local density functional theory (NL-DFT).
Results and discussion
Synthesis
The strategy for the synthesis of azo-bridged porphyrin-based conjugated microporous polymers (Azo-1, Azo-2, and Azo-3) is presented in Scheme 1S.† From previous report,35 azo compounds can be prepared by simple synthetic protocols. However, we choose azo condensation reaction to connect the building blocks and synthesis of conjugated polymers for the following reasons: (1) azo condensation is an efficient, well-established process and does not require expensive catalysts; (2) it can introduce azo bonds into the conjugated structure that can produce electron-rich conjugated networks; (3) it promotes the binding between the electron-rich azo series and the electron-poorer CO2.
Structural and chemical characterization
FT-IR spectra of Azo-1, Azo-2, and Azo-3 show the characteristic –N
N– stretching band at around 1598, 1597 and 1598 cm−1 (λN
N: 1630–1580 cm−1)36 along with the respective bands for aromatic rings (see Fig. 1), which is not appearing in the building block TNPP. Furthermore, several weak bands are observed in the region from 1300 to 1520 cm−1. Compared with the bands located in TNPP, the bands of azo series are much weaker, indicating that the degree of polymerization is high. The bands located at 1510 and 1344 cm−1 correspond to N–O stretching mode,35 suggesting most terminal nitro groups were reacted.
 |
| Fig. 1 FT-IR spectra of Azo-1, Azo-2, Azo-3, and TNPP. | |
In the solid-state 13C CP/MAS NMR spectra (Fig. 2), Azo-1, Azo-2, and Azo-3 gave signals of chemical shift at 151.6, 144.2, 131.2, and 118.1 ppm, which were consistent with those of azo-linked aromatic polymers,37 attributing to the carbon atoms of the porphyrin rings and phenyl linkers.38,39 The carbons bonded to the azine nitrogen (–N
C–) appear at 151.6 ppm whereas those bonded to the imino nitrogen (–NH–) appear at 118.1 ppm, suggesting that the downfield carbon resonance should be assigned to the α-pyrrolenine-like ring and the high-field resonance to the (1H)-pyrrole-like ring.40 As expected, a typical peak at around 151.6 ppm, attributed to the azo-linked aromatic carbons (–Car–N
N–Car–),41 is found for Azo-1, Azo-2, and Azo-3.
 |
| Fig. 2 Solid-state 13C NMR spectra of Azo-1 (black), Azo-3 (red) and Azo-3 (blue). Asterisks represent spinning sidebands. | |
XPS was performed to analyze the elemental composition and nitrogen bonding configurations in azo series. The XPS survey spectrum shows the presence of carbon, nitrogen, and oxygen distinctly compared with the theoretical vales of 850 eV, 399 eV, and 533 eV,42 respectively (see Fig. S1†). The bonding configurations of nitrogen in azos are revealed by the N 1s core level spectra (Fig. 3a–c). All of the nitrogen responses of Azo-1 and Azo-2, giving 398.0, 399.9, and 401.0 eV, correspond to the imine nitrogen43 (–C
N–) (N1), azo nitrogen (Ph–N
N–Ph) (N2), and pyrrolic nitrogen (–NH–) (N3), respectively.44,45 The nitrogen responses of Azo-3 can be deconvoluted into five types: 398.0, 399.9, 401.0, and 405.7, corresponding to the imine nitrogen (–C
N–) (N1), azo nitrogen (Ph–N
N–Ph) (N2), pyrrolic nitrogen (–NH–) (N3), and nitro nitrogen (Ph–NO2) (N4), respectively (Fig. 3d). The integral areas of these five peaks for azos were used to calculate the content of nitrogen species (N1–N4, Fig. S1†), from which we can find the ratio of N4 for Azo-3 is the highest, suggesting that the degree of polymerization is lower than the other two polymer networks.
 |
| Fig. 3 N 1s XPS spectra (a–c) and different nitrogen species for azos (a: Azo-1, b: Azo-2, c: Azo-3), and (d) different structures containing nitrogen (N1–4). | |
TGA was performed on Azo-1, Azo-2, and Azo-3 to determine the thermal stability and to confirm the absence of guest molecules inside the pores (see Fig. S2†). Azo-1 shows mass loss around 0.30% at temperature 132 °C, which is possibly caused by loss of volatiles (such as moisture) or trapped solvent molecules in the polymer networks when compared to that of TNPP. Similarly, Azo-2 shows mass loss around 2.41% at temperature 100 °C, and Azo-3 shows mass loss around 0.99% at 133 °C. Due to the decomposition of the network, a gradual weight loss of 5% for Azo-1, Azo-2, Azo-3 was observed after 317 °C, 135 °C, 305 °C, respectively and about 10% weight loss at 500 °C for all. Overall, Azo-1, Azo-2, and Azo-3 retain 65.63%, 68.05%, 66.97% of their mass respectively at 800 °C as revealed by TGA experiments, similar to that of TNNP. However, azo series presents less weight loss between 400 and 500 °C, which may be ascribable to the high degree of polymerization. These results highlight that Azo-1/2/3 have excellent thermal stability in a nitrogen atmosphere.
Morphological and crystallographic characterization
The morphology and microstructure of the Azo-1, Azo-2, and Azo-3 were investigated by SEM and TEM. All of the polymer networks illustrated similar morphology; Fig. 4 shows the representative SEM and TEM images of Azo-1 (more see Fig. S3†). Similar to other microporous polymers,46,47 Fig. 4a shows the azo polymer networks are wrinkles, and composing of loose agglomerates of tiny particles with rough surface and irregular shapes. For the sake of clarity, an enlarged TEM images are presented in Fig. 4b. The white region corresponds to the nanopore channels, while the dark region is attributed to the polymer networks skeleton. It clearly shows that the pores in the polymer networks are quite uniform and the polymer networks demonstrate typical microporous materials.
 |
| Fig. 4 A typical (a) SEM image and (b) TEM image of Azo-1 polymer. | |
In order to elucidate the crystallinity of the azo series, these polymer networks were also characterized by XRD (see Fig. S4†), intense peaks in the XRD patterns of Azo-1, Azo-2 and Azo-3 were not found except some minor peaks, indicating that Azo-1, Azo-2 and Azo-3 are mostly amorphous, which was consistent with previous SEM and TEM images showing litter crystallinity within the polymer networks. Indeed, conjugated microporous polymers are known to be amorphous solids.14
Studies of gas sorption and storage
Gas sorption. The permanent porosity in the networks was investigated by sorption analysis using N2, CO2, and H2 as the adsorbate molecules. Fig. 5a shows the N2 adsorption and desorption isotherms for Azo-1/2/3, while Table 1 summarizes the porous properties of a few comparable polymer networks. Azo-1, Azo-2, and Azo-3 displayed type II isotherms with type I character at higher relative pressures according to IUPAC classification. In Fig. 5a, one can find all of the desorption isotherms of the three networks exhibit some extent hysteresis, suggesting the presence of some mesoporous within the materials. The calculated Brunauer–Emmett–Teller (BET) specific surface areas of the azo series are 571 m2 g−1, 675 m2 g−1, and 520 m2 g−1, and the Langmuir surface areas 668 m2 g−1, 730 m2 g−1, and 610 m2 g−1 for Azo-1, Azo-2, and Azo-3, respectively. From the robust porous features of the three materials presented in Table 1, we found that total pore volumes of the azo series are 1.342 cm3 g−1, 1.686 cm3 g−1, 1.383 cm3 g−1 for Azo-1, Azo-2, and Azo-3, respectively. These values are comparable or higher than that of previously reported microporous materials constructed through azo condensation.37
 |
| Fig. 5 (a) N2 adsorption (solid) and desorption (hollow) isotherms of Azo-1, Azo-2 and Azo-3 at 77 K; (b) pore size distribution of azo series. | |
Table 1 Surface properties and gas uptake for Azo-1/2/3
Polymer network |
SBET [m2 g−1] |
SLangmuir [m2 g−1] |
dAVa [nm] |
SMICRO [m2 g−1] |
VMICRO [cm3 g−1] |
VTOTb [cm3 g−1] |
VMICRO/VTOT |
H2 uptake [wt%] |
CO2 uptake [wt%] |
77 K |
87 K |
273 K |
298 K |
Note: average pore diameter. Note: total pore volume at P/P0 = 0.99. |
Azo-1 |
571 |
668 |
3.47 |
338 |
0.156 |
1.342 |
0.12 |
0.86 |
0.76 |
9.42 |
7.11 |
Azo-2 |
675 |
730 |
2.72 |
437 |
0.186 |
1.686 |
0.11 |
1.15 |
0.80 |
17.52 |
8.97 |
Azo-3 |
520 |
610 |
2.89 |
376 |
0.164 |
1.383 |
0.12 |
0.97 |
0.71 |
12.63 |
6.60 |
Fig. 5b shows the pore size distribution (PSD) curves for the three polymer networks as calculated using NL-DFT. Azo-1 exhibits average pore diameter centering at around 3.47 nm with a major peak at 1.6 nm, and a minor peak in the 3.8–4.0 nm region. Azo-2 presents average pore diameter centering at around 2.72 nm with a major peak at 1.6 nm, and a minor in the 3.8–4.0 nm region. Azo-3 gives average pore diameter centering at around 2.89 nm with a major peak at 1.6 nm, and a minor in the 3.8–4.0 nm region. In addition, all of the networks show weak peaks in the size range of 2–4 nm. The PSD analysis agrees with the results of the N2 isotherms (Fig. 5a), suggesting that the azo polymers feature some mesopores within the micropore structures. Since types I isotherm corresponds to microporous materials, and type IV is attributed for mesoporous materials. So the nitrogen isotherms of azos show type I with type IV character.48 The contribution of microporosity to the networks can be calculated as the ratio of the micropore volume, VMICRO, over the total pore volume, VTOTAL. The ratios indicate networks Azo-1/2/3 possess a similar ratio of 0.11–0.12. The results along with TEM images indicate that Azo-1/2/3 are typical microporous materials. The ratio of Azo-2 is similar to that of Azo-1/3, but the absolute micropore volume per g of Azo-2 is larger than that of Azo-1/3, which may explain its higher N2 adsorption, and the higher BET specific surface area.
The large surface area, microporous structure and narrow pore size distribution of the polymer networks and nitrogen-enriched networks intrigue us to further explore their application in CO2 capture and separation from other gases. The CO2 sorption of the polymer networks were investigated by a volumetric method at 273 and 298 K (Fig. 6). For all the three polymers, the CO2 uptake appears a rapid rise in the initial stage, implying that CO2 molecule has favorable interaction with the polymer skeleton. Moreover, it is noted that the uptake of CO2 continually increases with the relative pressure in the experimental range, suggesting that a higher adsorption capacity can be achieved at the high pressure. Fig. 6a shows the CO2 uptakes by Azo-1/2/3 are 94.2 mg g−1, 175.2 mg g−1, and 126.3 mg g−1 at 273 K and 1 bar, respectively. Azo-2 obtains the highest CO2 uptake (3.98 mmol g−1, 17.52 wt% at 273 K/1 bar), which may be ascribed to its larger micropore volume as shown in previous BET values. Additionally, it was found that Azo-3 exhibits a higher CO2 adsorption capacity than Azo-1 even though its apparent BET surface area is lower. This difference may arise from pore size distribution effect of azo and the difference in micropore volume (VMICRO for Azo-1, Azo-3 is 0.156, 0.164 cm3 g−1, respectively). It has been reported that the micropore volume has a stronger effect on the gas uptakes of a material than the total pore volume and the specific surface area.49 On the other hand, it is expected that the CO2 uptake of the azo networks decrease at 298 K (which are 71.1 mg g−1 for Azo-1, 89.7 mg g−1 for Azo-2, and 66.0 mg g−1 for Azo-3, see Fig. 5c). These values are still comparable to or better than that of very high surface area porous polymers (Table S1 in ESI†).
 |
| Fig. 6 Gas adsorption studies. (a) CO2 adsorption isotherms of Azo-1, Azo-2, and Azo-3 at 273 K (a) and 298 K (c). H2 adsorption isotherms of Azo-1, Azo-2, and Azo-3 at 77 K (b) and 87 K (d). Isosteric heats of adsorption of Azo-1, Azo-2, and Azo-3 for CO2 (e) and H2 (f). | |
We have further studied the hydrogen (H2) sorption performance of the azo series in powder form. The H2 sorption of the network polymers were investigated by the volumetric method at 77 and 88 K. Fig. 6b and d shows the H2 adsorption isotherms up to a maximum H2 pressure of 1 bar. At 77 K, the H2 uptakes are 8.6 mg g−1 for Azo-1, 11.5 mg g−1 for Azo-2, and 9.7 mg g−1 for Azo-3 (Fig. 6b). Azo-2 exhibits the largest H2 uptake of 5.75 mmol g−1 at 77 K/1 bar (1.15 wt%). Uptakes of H2 are in the order of Azo-2 > Azo-3 > Azo-1, as same as the order of CO2 sorption of these networks. This can also be explained by the pore size distribution effects and the difference in micropore volume. As expected, the H2 uptakes of networks decrease at 87 K to 7.6 mg g−1 for Azo-1, 8.0 mg g−1 for Azo-2, and 7.1 mg g−1 for Azo-3 (Fig. 6d). Compared to the reported porous networks (Table S2 in ESI†), the capability of H2 uptakes is moderate.
Heats of adsorption. To further understand the pore surface characteristics of the materials and the adsorption process, the isosteric heats of adsorption (Qst) for CO2 and H2 were investigated. Qst of CO2 was calculated using the Clausius–Clapeyron equation from the sorption data collected at 273 and 298 K. Fig. 6e shows Qst of CO2 for Azo-1/2/3. At the adsorption onset, the Qst values for Azo-1, Azo-2 and Azo-3 are 37, 36, and 36 kJ mol−1, respectively. The values are higher than that of other porous materials, such as high surface area MPIs (30.4–34.8 kJ mol−1),50 electron-rich PECONFs (29–34 kJ mol−1),51 imine-linked PPFs (21.8–29.2 kJ mol−1),52 and comparable to or higher than microporous polymers with organic ammonium ions or Lewis basic amine in pores (30–55 kJ mol−1).53,54 The heat of adsorption drops by 40% (Azo-1), 30.1% (Azo-2) and 37.5% (Azo-3) during the CO2 uptake experiments, suggesting that, if the pore volumes of the materials can be further increased, a significantly enhanced CO2 gas uptake can probably be achieved even at low CO2 pressure. The Clausius–Clapeyron equation for heat adsorption is given as below:55 |
 | (1) |
where P, T, R and C are the pressure, temperature at the equilibrium state, the gas constant, and equation constant, respectively.From the sorption data collected at 77 and 87 K, Qst of H2 adsorption was calculated using the eqn (1) too. Fig. 6f shows Qst of Azo-1, Azo-2 and Azo-3 are of 6.2, 8.5 and 8.8 kJ mol−1 respectively at the adsorption onset. The heat of adsorption decreases slightly with increased hydrogen loading. Compared to the isosteric heats of adsorption for CO2, adsorption of H2 at the surfaces of the azo networks is much less, mainly because of little active sites at the surfaces of the materials. Physisorption of the H2 contributes the most for its adsorption.
Gas selectivity. To assess the potential applications of these networks in post combustion CO2 capture under ambient pressure, single component gas adsorption isotherms were carried out with CO2 or N2 under 273 K and up to 1 bar (Fig. 7). The selectivity for CO2 over N2 was calculated by the ratios of the initial slopes of CO2 and N2 adsorption isotherms based on the pressure below 0.10 bar and the ideal adsorbed solution theory.56 Azo-1, Azo-2 and Azo-3 give CO2/N2 selectivity of 42.1, 64.3 and 44.5, respectively, based on the initial 5 data points (<0.1 bar) of the isotherm curves, indicating greatly increased selectivity of Azo-2 for CO2 uptake over N2. The high CO2 adsorption capacity and CO2/N2 adsorption selectivity are the two key parameters for evaluating the potential of porous materials in carbon dioxide capture applications. The excellent selective adsorption of CO2 may pave the way to CO2 capture and separation, as well as air-cleaning applications by the microporous materials.
 |
| Fig. 7 Adsorption isotherms of CO2 and N2 gases at 273 K for Azo-1 (a), Azo-2 (b), and Azo-3 (c). | |
Conclusions
A series of new conjugated microporous polymers, Azo-1, Azo-2, and Azo-3 was successfully synthesized through azo condensation of TNPP with p-phenylenediamine, benzidine and melamine, respectively. Their chemical structures were confirmed by FT-IR and solid-state 13C CP/MAS NMR measurements. TGA measurements showed excellent thermal stability of these networks. XRD studies indicated amorphous structure of the polymers. Pore size distributions revealed a mainly microporous network structure for Azo-1, Azo-2, and Azo-3 with BET surface areas of 571 m2 g−1, 675 m2 g−1, and 520 m2 g−1, respectively, whereas Azo-2 distinguishes by its higher micropore volume. Under the present reaction conditions, pore sizes between 1.6 and 4 nm were obtained. Owing to the narrow pore size distribution and nitrogen-rich pore surfaces, azo series exhibit exceptionally high CO2 uptake. Thanks to the bigger nano-micropore volume, Azo-2 showed the highest BET surface area up to 675 m2 g−1 and a remarkable CO2 uptake of 3.98 mmol g−1 (or 17.52 wt%) at 273 K and 1 bar pressure. The heat of adsorption for CO2 was calculated to be 36–37 kJ mol−1, which is larger than that of many reported porous materials. Moreover, azos showed excellent adsorption selectivity with CO2/N2 up to 64.3 for Azo-2 at 273 K/1 bar. The H2 storage capability of the Azo-1/2/3 is moderate through physisorption of the azo polymers. The above results indicate that the reported azo porous polymer networks are promising candidates as adsorbents for CO2 capture, separation and storage, and will be exploited for CO2 caption in industrial (e.g. combustion gas) and environmental air-cleaning.
Acknowledgements
The authors thank the financial support from National Natural Science Foundation of China (51403126). ZZ and JW acknowledge the support from the JSNN at UNCG.
Notes and references
- A. Thomas, P. Kuhn, J. Weber, M. M. Titirici and M. Antonietti, Macromol. Rapid Commun., 2009, 30, 221–236 CrossRef CAS PubMed.
- D. Wu, F. Xu, B. Sun, R. Fu, H. He and K. Matyjaszewski, Chem. Rev., 2012, 112, 3959–4015 CrossRef CAS PubMed.
- Z. Xiang and D. Cao, J. Mater. Chem. A, 2013, 1, 2691–2718 CAS.
- Q. Chen, M. Luo, P. Hammershøj, D. Zhou, Y. Han, B. W. Laursen, C.-G. Yan and B.-H. Han, J. Am. Chem. Soc., 2012, 134, 6084–6087 CrossRef CAS PubMed.
- D. Yuan, W. Lu, D. Zhao and H. C. Zhou, Adv. Mater., 2011, 23, 3723–3725 CrossRef CAS PubMed.
- W. Lu, D. Yuan, J. Sculley, D. Zhao, R. Krishna and H.-C. Zhou, J. Am. Chem. Soc., 2011, 133, 18126–18129 CrossRef CAS PubMed.
- J. Germain, J. M. Fréchet and F. Svec, Small, 2009, 5, 1098–1111 CrossRef CAS PubMed.
- O. K. Farha, A. Ö. Yazaydın, I. Eryazici, C. D. Malliakas, B. G. Hauser, M. G. Kanatzidis, S. T. Nguyen, R. Q. Snurr and J. T. Hupp, Nat. Chem., 2010, 2, 944–948 CrossRef CAS PubMed.
- M. Pumera, Energy Environ. Sci., 2011, 4, 668–674 CAS.
- D. M. D'Alessandro, B. Smit and J. R. Long, Angew. Chem., Int. Ed., 2010, 49, 6058–6082 CrossRef PubMed.
- G. Férey, C. Serre, T. Devic, G. Maurin, H. Jobic, P. L. Llewellyn, G. De Weireld, A. Vimont, M. Daturi and J.-S. Chang, Chem. Soc. Rev., 2011, 40, 550–562 RSC.
- R. Dawson, E. Stöckel, J. R. Holst, D. J. Adams and A. I. Cooper, Energy Environ. Sci., 2011, 4, 4239–4245 CAS.
- 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.
- N. B. McKeown and P. M. Budd, Chem. Soc. Rev., 2006, 35, 675–683 RSC.
- J. Wang, I. Senkovska, M. Oschatz, M. R. Lohe, L. Borchardt, A. Heerwig, Q. Liu and S. Kaskel, ACS Appl. Mater. Interfaces, 2013, 5, 3160–3167 CAS.
- G. Sethia and A. Sayari, Carbon, 2015, 93, 68–80 CrossRef CAS.
- G.-P. Hao, W.-C. Li, D. Qian and A.-H. Lu, Adv. Mater., 2010, 22, 853–857 CrossRef CAS PubMed.
- K. Huang, S.-H. Chai, R. T. Mayes, G. M. Veith, K. L. Browning, M. A. Sakwa-Novak, M. E. Potter, C. W. Jones, Y.-T. Wu and S. Dai, Chem. Commun., 2015, 51, 17261–17264 RSC.
- G. Sethia and A. Sayari, Energy Fuels, 2014, 28, 2727–2731 CrossRef CAS.
- N. Popp, T. Homburg, N. Stock and J. Senker, J. Mater. Chem. A, 2015, 3, 18492–18504 CAS.
- S. N. Talapaneni, O. Buyukcakir, S. H. Je, S. Srinivasan, Y. Seo, K. Polychronopoulou and A. Coskun, Chem. Mater., 2015, 27, 6818–6826 CrossRef CAS.
- Y. Xu, S. Jin, H. Xu, A. Nagai and D. Jiang, Chem. Soc. Rev., 2013, 42, 7965–8178 RSC.
- A. I. Cooper, Adv. Mater., 2009, 21, 1291–1295 CrossRef CAS.
- Y. Liu, X. Guo, N. Xiang, B. Zhao, H. Huang, H. Li, P. Shen and S. Tan, J. Mater. Chem., 2010, 20, 1140–1146 RSC.
- X. Feng, L. Chen, Y. Dong and D. Jiang, Chem. Commun., 2011, 47, 1979–1981 RSC.
- N. U. Day, C. C. Wamser and M. G. Walter, Polym. Int., 2015, 64, 833–857 CrossRef CAS.
- W. M. Campbell, K. W. Jolley, P. Wagner, K. Wagner, P. J. Walsh, K. C. Gordon, L. Schmidt-Mende, M. K. Nazeeruddin, Q. Wang and M. Grätzel, J. Phys. Chem. C, 2007, 111, 11760–11762 CAS.
- W. Zheng, N. Shan, L. Yu and X. Wang, Dyes Pigm., 2008, 77, 153–157 CrossRef CAS.
- X. Liu, H. Li, Y. Zhang, B. Xu, S. A, H. Xia and Y. Mu, Polym. Chem., 2013, 4, 2445–2448 RSC.
- V. S. P. K. Neti, J. Wang, S. Deng and L. Echegoyen, RSC Adv., 2015, 5, 10960–10963 RSC.
- J. Lu and J. Zhang, J. Mater. Chem. A, 2014, 2, 13831–13834 CAS.
- P. Arab, E. Parrish, T. Islamoglu and H. M. El-Kaderi, J. Mater. Chem. A, 2015, 3, 20586–20594 CAS.
- P. Arab, M. G. Rabbani, A. K. Sekizkardes, T. İslamoğlu and H. M. El-Kaderi, Chem. Mater., 2014, 26, 1385–1392 CrossRef CAS.
- M. Yuasa, K. Oyaizu, A. Yamaguchi and M. Kuwakado, J. Am. Chem. Soc., 2004, 126, 11128–11129 CrossRef CAS PubMed.
- H. A. Patel, S. Hyun Je, J. Park, D. P. Chen, Y. Jung, C. T. Yavuz and A. Coskun, Nat. Commun., 2013, 4, 1357 CrossRef PubMed.
- P. Narayanan, Essentials Of Biophysics, Anshan, 2000 Search PubMed.
- H. A. Patel, S. H. Je, J. Park, D. P. Chen, Y. Jung, C. T. Yavuz and A. Coskun, Nat. Commun., 2013, 4, 1357 CrossRef PubMed.
- X. Feng, L. Liu, Y. Honsho, A. Saeki, S. Seki, S. Irle, Y. Dong, A. Nagai and D. Jiang, Angew. Chem., 2012, 124, 2672–2676 CrossRef.
- A. Modak, M. Nandi, J. Mondal and A. Bhaumik, Chem. Commun., 2012, 48, 248–250 RSC.
- L. Frydman, A. C. Olivieri, L. E. Diaz, A. Valasinas and B. Frydman, J. Am. Chem. Soc., 1988, 110, 5651–5661 CrossRef CAS.
- S. Xie, A. Natansohn and P. Rochon, Macromolecules, 1994, 27, 1489–1492 CrossRef CAS.
- A. P. Hitchcock and C. E. Brion, J. Electron Spectrosc. Relat. Phenom., 1980, 18, 1–21 CrossRef CAS.
- A. P. Dementjev, A. de Graaf, M. C. M. van de Sanden, K. I. Maslakov, A. V. Naumkin and A. A. Serov, Diamond Relat. Mater., 2000, 9, 1904–1907 CrossRef CAS.
- L. Silipigni, G. De Luca, Q. Quattrone, L. M. Scolaro, G. Salvato and V. Grasso, J. Phys.: Condens. Matter, 2006, 18, 5759–5772 CrossRef CAS.
- W. Lian, Y. Sun, B. Wang, N. Shan and T. Shi, J. Serb. Chem. Soc., 2012, 77, 335–348 CrossRef CAS.
- Y. Yang, Q. Zhang, Z. Zhang and S. Zhang, J. Mater. Chem. A, 2013, 1, 10368–10374 CAS.
- M. Mastalerz, M. W. Schneider, I. M. Oppel and O. Presly, Angew. Chem., Int. Ed., 2011, 50, 1046–1051 CrossRef CAS PubMed.
- G. Crini and P.-M. Badot, Sorption Processes and Pollution: Conventional and Non-conventional Sorbents for Pollutant Removal from Wastewaters, Presses Univ. Franche-Comté, 2010 Search PubMed.
- A. P. Katsoulidis and M. G. Kanatzidis, Chem. Mater., 2012, 24, 471–479 CrossRef CAS.
- G. Li and Z. Wang, Macromolecules, 2013, 46, 3058–3066 CrossRef CAS.
- P. Mohanty, L. D. Kull and K. Landskron, Nat. Commun., 2011, 2, 401 CrossRef PubMed.
- Y. Zhu, H. Long and W. Zhang, Chem. Mater., 2013, 25, 1630–1635 CrossRef CAS.
- A. Demessence, D. M. D'Alessandro, M. L. Foo and J. R. Long, J. Am. Chem. Soc., 2009, 131, 8784–8786 CrossRef CAS PubMed.
- Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li, Angew. Chem., 2012, 124, 3420–3423 CrossRef.
- V. Krungleviciute, L. Heroux, A. D. Migone, C. T. Kingston and B. Simard, J. Phys. Chem. B, 2005, 109, 9317–9320 CrossRef CAS PubMed.
- X. Zhu, C.-L. Do-Thanh, C. R. Murdock, K. M. Nelson, C. Tian, S. Brown, S. M. Mahurin, D. M. Jenkins, J. Hu, B. Zhao, H. Liu and S. Dai, ACS Macro Lett., 2013, 2, 660–663 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: Additional experimental data for materials. See DOI: 10.1039/c6ra04077b |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.