Sie sind auf Seite 1von 10

International Journal of Chemical & Petrochemical Technology (IJCPT) ISSN 2277-4807 Vol.

3, Issue 3, Aug 2013, 17-26 TJPRC Pvt. Ltd.

MICROPOROUS ORGANIC FRAMEWORK POLYMERS FOR HYDROGEN STORAGE APPLICATIONS


FADI IBRAHIM Research Scholar, Department of Chemistry, Abdul Latif -Thanian, Al-Ghanim School Kuwait, Kuwait

ABSTRACT
A microporous organic framework polymers (OFPs 7-12) based on a polyimide framework were developed as a new H2 storage mateials. OFP shows a reversible H2 storage capacity reach to 3.94 wt% at 10 bar and 77K. This polymers consider to overcome the low gas storage challeng related to fuel nanotechnology applications.

KEYWORDS: Microporous, Framework Polymer, Hydrogen Adsorption INTRODUCTION


The efficient Hydrogen storage has been identified as one of the challenging tasks to accomplish the goals for hydrogen economy. However, the use of hydrogen as an onboard fuel involves many concerns which include easy storage and release, transportation, and safety issues.1,2 In particular, hydrogen storage is envisaged as a major concern for meeting DOE targets both in terms of volumetric as well as gravimetric density for vehicular applications.3 The major criteria in the development of new hydrogen storage materials are improved energy storage density, kinetics and cycle life, using readily available elements at a reasonable cost, low heat of formation, low cost, low weight, high stability against O2, and moisture for long cycle life.3 This includes use of metal hydrides, complex hydrides, chemical storage, carbon based nanoporous materials and Microporous and mesoporous sorbents for hydrogen storage. For this purpose many materials were explored including the carbon nanotubes4-5 (CNT),activated carbons6-7, zeolites
8-9

, metal hydrides10 materials and other MOFs.11-13 Recently several attempts focussed on porous polymers with

intrinsic microporosity (PIM) such as PIM1, CTC, HATN and Trip-PIM were reported.14-17 The storage capacity of PIMs is around 0.95 to 1.63 wt% H 2 at 77 K and 1 bar. This values are generally promising for the development of microporous polymers. These organic polymers composed of light elements and can be an ideal substrate for storing and transportaion. Organic polymers offer a wide range of advantages over MOF, zeolites and other microporous materials as an onborad storage materials. The main attaraction is the low intrinsic density (composed of light elemenets like C, H, N, O, while MOF and zeolites composed of heavier elements like metals) and other merits including good chemical, comparable thermal and structural stability.14 Other class of porous polymers reported were Hyper crosslinked polymers 18-19 (hcps). They were prepared by extensive crosslinking reactions which can pevent the polymer chain entanglement to give porous material having high surface areas with permanent porosity. The hydrogen adsorption of HCPs around 1.28-1.5 at 77K and 1 bar. The possibility of the preparation of a framework material was confirmed by making a model compound 2 directly from a reaction between four molar equivalents of di-tert-butyl catechol with the 4,4`-bis-(4,5,6,7-tetrafluoro)-2,2`-bis-(triafluoromethyl)-benzidine isoindole1,3-dione 1 (scheme 1). The reaction product was characterised by mass spectroscopy, NMR and elemental analyses which indicate that all the eight F atoms were replaced by four catechol units. This highly efficient nucleophilic aromatic substitution reaction (yield >82%) is clearly suitable for the preparation of the target organic framework polymer.

18

Fadi Ibrahim

F F

O O

OH

K2CO3, DMF

+
CF3 F3C OH

80 oC, 24 h
F3C

CF3

F F

O O

Scheme 1: Synthesis of Model Compound (2) The polymer framework was synthesised by the polycondensation of different monomers with TTSBI at 80 oC (Scheme 2).

O O

O O

N O O O

Ar

N O O O

HO HO F F F F

(i)

N
Ar N

O HO
O
O O

O O O

O O O O O

HO
F F
O O O O O O O O O O

OH
F F

O O

(ii) (i) & (ii)


K2CO3, DMF/ 120 oC , 24 h.

OH

Ar
O O O O N O O

(0.92 nm)
O O

Ar
N O O O O

O O O

O O O O

CF3

Ar =

H3C CF3

CH3

OFP-7

OFP-9

OFP-11

CF3 CF3 O

CF3

CF3

OFP-8

OFP-10

OFP-12

Scheme 2: Synthesis and Idealised Structure of the Organic Framework Polymer (OFPs 7-12)

RESULTS AND DISCUSSIONS


The insoluble yellow fluorescent powder was characterised by IR, solid state 13C NMR, elemental and XPS analysis. The IR spectroscopic analysis confirms the structural identification of OFP-3 by the appearance of new dibenzodioxane link related band and the absence of bands corresponding to fluorinated aromatics as supported by the XPS analysis (<1% residual fluorine). The FTIR spectra of organic framework polymers clearly indicate the presence of imide functional groups. As shown in Figure 1, the characteristic bands related to the imide stretching (C=O symmetric and asymmetric bands in the range of 1722-1792) found in the spectrum of all the three polymers. All the spectra quite similar in appearance and in good agreement with the data of polyimides.

Microporous Organic Framework Polymers for Hydrogen Storage Applications

19

Figure 1: FTIR Spectra of Monomer (1), the Model Compound and the OFP-10 However, the bands slightly shifted to lower wave number in the case of all OFPs in comparison to the corresponding fluoro-monomers, which can be attributed to the inductive effect of fluorine atoms. More over both spectra are quite similar and can be superimposed nicely excluding the bands related to the tert-butyl groups at 1413 cm-1. Another feature of monomer 1 is the presence of strong bands observed at 1499 & 1511, which are characteristic of the fluorine-containing C=C. In the case of 2 and OFP-10 the bands due to C=C appeared as a single broad band and shifted to low wave number, which is a clear indication of change in the C=C environment (1458 & 1467) through the dibenzodioxane link formation. The assignment for the aromatic F-containing C=C and C-F stretching spectral data was made on the basis of some known aromatic fluoro-compounds and the precursor molecules. XPS Analysis X-ray photoelectron spectroscopy is an additional technique which was used to detect the presence of fluorine and other elements within OFP-8 as shown in Figures 2 and 3. The XPS survey spectra of monomer IM-8 and OFP-8 reveals the presence of carbon, nitrogen, oxygen and fluorine. The binding energy of the N1s at 400.2 eV confirms the type of nitrogen bonding to carbon as represented by the structure OFP-4. Moreover it is clear from the spectrum that the base catalysed polymerisation condition does not degrade the imide units as supported by IR and XPS. However, the polymerisation condition did not affect the type of nitrogen bonding in OFP-8, and it therefore appears at the same binding energy. The binding energy of the F1s at 687.97 eV confirms the presence of fluorine in both compound IM-8 monomer and OFP-8. On the other hand, fluorine analysis of OFP-8 shows only a trace amount (<1 wt. %) compared to monomer IM-8. This can be considered as a further support that the fluorine atoms have been replaced by the oxygen of the dioxane links and consequently confirm the proposed structure of the OFP-8. Trace amount of fluorine found in OFP-8 could be attributed to the unreacted fluorine containing-end groups and other impurities in terms of entrapped fluorinated byproducts from base catalysed reaction as shown in Figure 3.

Figure 2: XPS Survey Spectrum of 1,4-Bis(4,5,6,7-Tetrafluorophthalimido)-2,3,5,6- Tetramethylbenzene IM-8

20

Fadi Ibrahim

Figure 3: XPS Survey Spectrum of OFP-8 The synthesis part mainly focussed on a novel approach to create microporous polymer framework with high volumetric capacity. The best systems were framework materials prepared by the polycondensation (Scheme 1). Based on a model reaction with the diteritiarybutyl catechol with fluro and chloroimides. Both the model reactions give tetrasubstituted product with different yields. The fluoro analogoes give good yield in a short time. The reaction route and the best monomers employed are shown in scheme 1. The mcroporous rigid polymer framework was synthesised by the polycondensation of octafluorodiimide with the spirocatechol at 120 oC. The prodcut obtained was yellow fluoresecent powder. This material offers high. Also we employed different solvents such as DMF and DEMAC and the best system known to be DMF. The solid state
13 13

C NMR

(75 MHz) spectrum (Figure 4) of OFP-11, was examined and the signal positions were located precisely with the help of C NMR spectra of monomers and precursor molecules

Figure 4: Comparison of 13C NMR Spectra of OFP-11 The signals marked as 1 (4 carbons, belongs to monomer 1) and 2 (8 carbons, belongs to spirobiscatechol unit) represent the methyl groups as depicted in the scheme 3 of the proposed repeat unit. The signal at 42.99 (marked as 3) corresponding to methylene groups of the spirobiscatechol unit, represent a total carbons of four. There are two types of quarternary carbons come under one signal at 57.76 marked as 4 which represent a total of 6 carbons. The signal appeared at 111.16 marked as 5 represents two types of aromatic carbon (4 + 8 carbons). The overlapped signals between 132.8 and 147.8 representing 30 aromatic carbons within the repeat unit. Based on the information retrieved by the 13C NMR spectra of the precursor molecules, the overlapping resonances can be confidently assigned to the four different carbon environments as marked as 6,7,8, 9 in the Scheme 3. The signal marked as 10 appeared at 161.70 clearly represent the carbonyl carbon (4 carbons). Thus all the signals in the spectrum accounted for the proposed idealised structure of the OFP-11, which further supports the highly efficient polycondensation process, through the formation of dibenzodioxane links. The solid state 13C NMR (75 MHz) spectrum of OFP-7 shown in Figure 5.

Microporous Organic Framework Polymers for Hydrogen Storage Applications

21

Figure 5: Solid State Left: 13C NMR Spectrum of OFP-7 Additionally, the comparison between the intensity of individual signals supports the proposed repeat unit OFPs. For example, the signals in the spectrum are also following the ratio which fulfils the proposed structure as shown in scheme 3.

Scheme 3: Proposed Repeat Unit of the Ideal Structure of OFPs 7-12 Generally, the hydrogen sorption capacity of materials can be enhanced by increasing surface area. In addition, the adsorption is completely reversible, and there is no significant hysteresis. The repeatability of the hydrogen adsorption was also checked and found to be constant even after t he repeated cycles. The comparison of BET surface area calculated from nitrogen adsorption isotherm and hydrogen adsorption isotherm (using Langmuir equation) of OFPs are summarized in Table 1. Table 1: Comparison of Surface Area and Nitrogen Uptake (wt.%) at 77 K of Organic Framework Polymer (OFPs) with Selected Other High Performance Materials OFPs OFP-7 OFP-8 OFP-9 OFP-10 OFP-11 OFP-12 SABET (m2/g) 637 1120 1340 1188 760 765 PvMicro (cm3/g) 0.39 0.91 1.00 0.74 0.48 0.72 H2 Uptake at 1 Bar 0.92 1.89 2.47 2.23 1.43 1.23

The nitrogen adsorption measurements (Figure 6) give a BET surface area reach to 1340 m2/g. On comparing various organic polymers and MOFs we observed that the storage capacity is only poorly correlated to the BET surface area. Nitrogen adsorption measurements give a BET surface area of 1252 m2/g for OFP-6, which is higher than the Trip-

22

Fadi Ibrahim

PIM18 but less than the HCPs19. The adsorption/desorption isotherms clearly exhibit pronounced hysteresis up to low partial pressure which is typically observed for microporous materials.

Figure 6: N2 Adsorption/Desorption Isotherms at 77 K for OFPs The high surface area of OFP-3 can be due to its rigid nonlinear architecture. The rigidity arises from the restricted rotation as a result of the dioxane linkage formation and the nonlinearity created by the spirobiscatechol. In addition the four methyl substituents hinder the rotation about C-N bond of the imide groups. In particular, the polymer constrained by a fixed network structure freezes any possible structural relaxations leading to creation of microporous structure. The comparison of hydrogen adsorption for OFP-4, OFP-6 and other microporous materials is summarized in Table 2. The storage of hydrogen is achieved via physical adsorption of hydrogen onto the surface of the material. It is very important to note that the hydrogen physisorption is completely reversible and this behaviour is always consistent with microporous materials. The mechanism of hydrogen uptake by any physisorption based -carriers is the monolayer adsorption; therefore, materials of high specific surface area and of subnanometer pore size are preferable. 20 Table 2: Comparison of the Hydrogen Uptake at 77 K Material OFP-9 OFP-10 Trip-PIM HCP IRMOF-8 SABET/ (m2/g) 1340 1188 1056 1466 1818 H2 % Uptake at (1 Bar) 2.47 1.23 1.65 1.28 1.50 H2 Uptake at (10 Bar) 3.94 3.53 2.71 2.75 3.6 Ref 18 19 20

The micropore size distribution for OFP-3 as calculated by the Horvath-Kawazoe (HK) method. The HK method is suitable for microporous materials and this equation relates the adsorption potential with the micropore size and allows each amount adsorbed at a relative pressure to be expressed in terms of the width of a slit shaped pore. 21 Our analysis shows maximum pore size in the range of 0.5-0.6 nm. The small pore diameter is especially advantageous for H 2 storage. X-ray diffraction revealed that OFP-11 is an amorphous material. High resolution transmission electron microscopy (HRTEM) was employed for the spatial mapping of the pores. The framework organization leads to worm hole type pore structures. High resolution transmission electron microscopy (HRTM) was emloyed for the spatial mapping of the pores (Figure 7). The framework organization leads to worm hole type pore structures.

Microporous Organic Framework Polymers for Hydrogen Storage Applications

23

Figure 7: HRTM Image of OFP-4 Hydrogen adsorption/desorption isotherms were measured at 77 K using a ASAP 2020 and 99.999% pure H2 over the pressure range 0-10 bar (these measurements were carried out at Micromeritics Analytical Laboratory, USA). The framework polymer OFP-11 afforded a hydrogen storage capacity of 1.43 wt.% at 1 bar which is comparable with the other reported microporous polymers.12-16 However, at 10 bar OFP-9 adsorbs nearly 3.94 % by weight of material as shown in Figure 8.

Figure 8: Hydrogen Adsorption/Desorption Isotherm of at 77 K for OFP-11 In addition the shape of the isotherm indicates that adsorption has not reached saturation and further significant hydrogen uptake could occur at higher pressures. This level of uptake is higher than all the reported organic polymers and some MOFs. More over small micropores such as that indicated for OFP-9 can adsorb hydrogen effectively.20 These ultramicropores allow the dihydrogen molecule to interact with multiple portions of the framework thereby increasing the interaction energy. In addition, the adsorption is completely reversible and there is no significant hysteresis which is consistent with the physisorption of hydrogen on a microporous material.

ACKNOWLEDGEMENTS
I'm grateful for Dr.Saad Makhseed and for financial support from the College of Graduate Studies for my PhD candidacy, and the facilities used from ANALAB, Kuwait University.

EXPERIMENTAL
Materials and Methods All the chemicals were of reagent grade purity and used without further purification. The dry solvent Dimethylformamide (DMF) and DEMAC were purchased from Aldrich Co. The finely grounded anhydrous potassium carbonate was used after drying at 200 oC. 1H NMR spectra (400 MHz) were recorded using Bruker DPX 400. FTIR spectra were recorded on a Perkin Elmer System 2000 FTIR. Elemental analyses were carried out using LECO Elemental Analyzer CHNS 932. Mass analyses were done on a VG Autospec-Q. Solid state NMR measurements were carried out on a Bruker Avance 300 spectrometer equipped with a cross polarization magic angle spinning (CP/MAS) probe and a fully automated pneumatic unit for sample spinning. X-ray photoelectron spectroscopy (XPS) measurement was carried out in

24

Fadi Ibrahim

a VG ESCALAB 200 instrument. High resolution transmission electron Microscopy (HRTEM: JEOL Model JEM- 3010 [300 KV]). TGA and DSC analyses were carried out on Shimadzu TGA-50 and Shimadzu DSC-50. Molecular structures were created using the Spartan programme and the Semiemperical molecular orbital methods (PM3 and MNDO) were used to simulate molecular structure. Surface area analysis was carried out using ASAP 2010 Micromeritics Sorptometer. Hydrogen sorption measurements were done using the Micromeritics ASAP 2020. 2,2`-Bis-(4,5,6,7-Tetrafluoro)-2-(2-Methyl-(1,1`)-Binaphthalenyl-2yl)Isoindole-1,3-Dione, IM-11 Tetrafluorophthalic anhydride was added to a stirred solution of 4,5,6,7-tetrafluoro)-2-(2-methyl-(1,1`)binaphthalenyl-2yl in glacial acetic acid and refluxed for 12 h. The white solid obtained 26 was further refluxed in acetic anhydride for 12 h. After cooling, the precipitated product was filtered off and washed with petroleum ether to give a light yellow solid. Yield 82%; m.p >300 oC; 1H NMR (DMSO-d6, 400 MHz, v ppm): 8.12-8.10 (d, 2H), 8.02-8.00 (d, 2H), 7.617.57 (t, 2H), 7.52-7.50 (d, 2H), 7.39-7.35 (t, 2H), 7.09-7.06 (d, 2H); 13C NMR (100 MHz) (DMSO-d6): 167.91, 145.32, 142.71, 140.72, 139.59, 127.42, 126.40, 124.11, 121.92, 118.18, 115.52; IR (KBr) /cm _1: 1728 (imide). CHN calculated for C36H12F8N2O4 (676.37): C, 63.90; H, 1.77; N, 4.14. Found: C, 63.63; H, 1.58; N, 4.06; MS (EI): m/z (%) 676 (M+, 100%). OFP-11 A mixture of 27 (0.50 g, 0.83 mmol) and TTSBI (0.55 g, 1.66 mmol) in dry DMF (100 ml) and K 2CO3 (6 g, 43.12 mmol) was stirred at 120 oC for 24 hrs. The reaction mixture was quenched with deionised water, filtered and washed repeatedly with methanol. Further purification was done using soxhlet extraction by methanol, ethanol, acetone and THF. The resulting polymer was yellow fluorescent powder 39 was dried under vacuum oven at 100 oC for 12 hrs. Yield 72%; m.p. >300 oC; 13C NMR (100 MHz): 162.28, 150.23, 142.71, 140.72, 131.26, 127.42, 126.40, 124.11, 121.92, 118.18, 113.26, 57.90, 42.31, 28.72; IR (KBr) /cm_1: 1783 and 1754 (imide) 1508, 952. CHN calculated for C78H52N2O12 (1208): C, 77.48; H, 4.30; N, 2.31. Found: C, 77.11; H, 4.10; N, 2.04. BET surface area = 760 m 2/g; total pore volume = 0.48 cm3/g.

REFERENCES
1. 2. 3. A. W. C. van den Berg and C. O. Arean, Chem. Commun., 2008 668-681. M. Bououdina, D. Grant and, G. Walker, Int. J. Hydrogen Energy, 2006, 31, 177-182. D. K. Slattery and M. D. Hampton, Proceedings of the 2002 US DOE Hydrogen Program Review, NREL/CP610-32405, Florida. 4. 5. A. Zuttel, Materials Today, 2003, 6, 24-33. A. Ansn, M. A. Callejas, A. M. Benito, W. K. Maser, M. T. Izquierdo, B. Rubio, J. Jagiello, M. Thommes, J. B. Parra and M. T. Martnez, Carbon, 2004, 42 12431248; A. Ansn, M. Benham, J. Jagiello, M. A. Callejas, A. M. Benito, W. K. Maser, A. Zuttel, P. Sudan and M. T. Martnez, Nanotechnology, 2004, 15, 1503-1508. 6. H. Kajiura, S. Tsutsui, K. Kadono, M. Kakuta, M. Ata and Y. Murakami, Appl. Phys. Lett., 2003, 82, 1105-1107; K. Kadono, H. Kajiura and M. Shiraishi, Appl. Phys. Lett., 2003, 83, 3392-3394.

Microporous Organic Framework Polymers for Hydrogen Storage Applications

25

7.

J. Weitkamp, M. fritz and S. Ernst, Int. J. Hydrogen Energy, 1995, 20, 967-970; H.W. Langmi, A. Walton, M. M. Al-Mamouri, S. R. Johnson, D. Book, J. D. Speight, P. P. Edwards, I. Gameson, P. A. Anderson and I. R. Haris, J. Alloy. Compd., 2003, 356-357, 710-715.

8. 9.

B. Bogdanovi and M. Schwickardi, J. Alloy. Compd.,1997, 253- 254, 1-9. J. L. C. Rowsell, A. R. Millward, K. S. Park and O. M. Yaghi, J. Am. Chem. Soc., 2004, 126, 5666-5667; A. Daily, J. J. Vajo and C. C. Ahn, J. Phys. Chem. B., 2006, 110, 1099-1101.

10. B. Panella, M. Hirscher, H. Putter and U. Muller, Adv. Funct. Mater., 2006, 16, 520-524; J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 1304-1315. 11. B. Chen, N. W. Ockwig, A. R. Millward, D. S. Contreras and O. M. Yaghi, Angew. Chem., Int. Ed., 2005, 44, 4745-4749; X. Lin, J. Jia, X. Zhao, K. M. Thomas, A. J. Blake, G. S. Walker, N. R. Champnes, P. Hubberstey and M. Schroder, Angew. Chem., Int. Ed., 2006, 45, 7358-7364. 12. N. B. McKeown, B. S. Ghanem, K. J. Msayib, P. M. Budd, C. E. Tattershall, K. Mahmood, S. Tan, D. Book, H. W. Langmi and A. Walton, Angew. Chem. Int. Ed., 2006 45, 1804-1807. 13. B. S. Ghanem, N.B. McKeown, K.D.M. Harris, Z. Pan, P.M. Budd, A. Butler, J. Selbie, D. Book and A. Walton, Chem. Commun., 2007, 67-69. 14. J-Y. Lee, C. D. Wood, D. Bradshaw, M. J. Rosseinky and A. I. Cooper, Chem. Commun., 2006, 2670-2672; C. D. Wood, B. Tan, A. Trewin, H. Niu, D. Bradshaw, M. J. Rosseinsky, Y. Z. Khimyak,N. L. Campbell, R. Kirk, E. Stockel and A. I. Cooper, Chem. Mater., 2007, 19, 2034-2048. 15. J. Germain, J. Hradil, J. M. J. Frechet and F. Svec, Chem., Mater., 2006, 18, 4430-4435. 16. J. Germain, J. M. J. Frechet and F. Svec, J. Chem, Mater., 2007, 17, 4989-4997. 17. N. Texier-mandoki, J. Dentzer, T.Piquero, S. Saadallah, P. David and C. Vix-Guterl, Carbon, 2004, 42, 27442747. 18. Ghanem BS, Msayib KJ, McKeown NB, Harris KDM, Pan Z, Budd PM, Butler A, Selbie J, Book D, Walton A. Chem Commun, 2007, 6769. 19. Wood CD, Tan B, Trewin A, Niu HJ, Bradshaw D, Rosseinsky MJ, Khimyak YZ, Campbell NL, Kirk R, Stockel E, Cooper AI.. Chem Mater,. 2007, 19, 20342048. 20. Hidetoshi K., Makoto Y., Kazuhiro T., and Ken O. (1997). Gas permselectivity of carbonized polypyrrolone membrane. Chem. Commun. 1051-1052.

APPENDICES
OFP-7 Yield 65%; m.p. >300 oC; 13C NMR (100 MHz): 28.22, 41.28, 49.36, 55.83, 113.44, 138.43, 146.12, 160.77; IR (KBr)/cm_1: 1786 and 1757 (imide) 1508, 954. CHN calculated for C74H56N2O12 (1164): C, 76.28; H, 4.81; N, 2.40. Found: C, 75.84; H, 4.21; N, 2.19. BET surface area = 637 m 2/g; total pore volume = 0.39 cm3/g.

26

Fadi Ibrahim

OFP-8 Yield 85%; m.p. >300 oC; 13C NMR (100 MHz): 167.12, 145.23, 143.65, 142.85, 141.72, 130.36, 128.59, 122.13, 112.63, 109.68, 118.32, 66.35, 57.91, 39.98, 29.65; IR (KBr) /cm_1: 1777 and 1749 (imide) 1508, 951. CHN calculated for C73H48N2O12F6 (1258): C, 69.63; H, 3.81; N, 2.22. Found: C, 68.32; H, 3.54; N, 3.15. BET surface area = 1120 m 2/g; total pore volume = 0.81 cm3/g. OFP-9 Yield 84%; m.p. >300 oC; 13C NMR (100 MHz): 167.42, 145.38, 142.77, 141.68, 136.26, 131.89, 129.83, 123.68, 122.65, 120.62, 109.23, 66.60, 15.21; IR (KBr) /cm_1: 1774 and 1748 (imide) 1508, 952. CHN calculated for C75H46N2O12F6 (1280): C, 70.31; H, 3.59; N, 2.18. Found: C, 70.01; H, 3.28; N, 2.02. BET surface area = 1340 m 2/g; total pore volume = 1.00 cm3/g. OFP-10 Yield 85%; m.p. >300 oC; 13C NMR (100 MHz): 167.78, 145.86, 142.41, 132.42, 130.59, 128.42, 125.40, 118.11, 117.92, 117.18, 112.61, 57.98, 40.43, 29.57; IR (KBr) /cm_1: 1792 and 1738 (imide) 1507, 949. CHN calculated for C72H40N2O12F6 (1238): C, 69.78; H, 3.23; N, 2.26. Found: C, 69.20; H, 3.14; N, 3.01. BET surface area = 1188 m2/g; total pore volume = 0.74 cm3/g. OFP-11 Yield 72%; m.p. >300 oC; 13C NMR (100 MHz): 162.28, 150.23, 142.71, 140.72, 131.26, 127.42, 126.40, 124.11, 121.92, 118.18, 113.26, 57.90, 42.31, 28.72; IR (KBr) /cm_1: 1783 and 1754 (imide) 1508, 952. CHN calculated for C78H52N2O12 (1208): C, 77.48; H, 4.30; N, 2.31. Found: C, 77.11; H, 4.10; N, 2.04. BET surface area = 760 m 2/g; total pore volume = 0.48 cm3/g. OFP-12 Yield 74%; m.p. >300 oC; 13C NMR (100 MHz): 27.57, 40.94, 55.55, 119.42, 133.8, 138.36, 146.00, 160.49; IR (KBr) /cm_1: 1783 and 1752 (imide) 1508, 955. CHN calculated for C 72H46N2O14 (1162): C, 74.35; H, 3.95; N, 2.40. Found: C, 74.11; H, 3.61; N, 2.21. BET surface area = 765 m 2/g; total pore volume = 0.72 cm3/g.

Das könnte Ihnen auch gefallen