검색 전체 메뉴
PDF
맨 위로
OA 학술지
Influence of oxyfluorination on activated carbon nanofibers for CO2 storage
  • 비영리 CC BY-NC
  • 비영리 CC BY-NC
ABSTRACT
Influence of oxyfluorination on activated carbon nanofibers for CO2 storage
KEYWORD
carbon nanofibers , CO2 storage , porosity , oxyfluorination , electrospun
  • 1. Introduction

    Emissions from the combustion of fossil fuels are contributing to an increase in the con-centration of carbon dioxide (CO2) in the atmosphere. It is generally accepted that this in-crease in atmospheric CO2 may be causing global climate change. Researchers have reported that global CO2 emissions from fossil fuels increased 29% between 2000 and 2008 and 41%between 1990 and 2008. According to a new study, the current concentration of CO2 in the atmosphere is at its highest now in at past 2 million years [1]. Consequently, CO2 capture from large point sources has recently received attention as a potential means of mitigating fossil fuel CO2 emissions. Current or proposed methods for capturing CO2 from flue gas have included absorption, adsorption, cryogenic distillation, and membrane separation. However, commercial CO2 capture technology existing today is very expensive and energy intensive. Improving the technologies for CO2 adsorption is necessary in order to achieve low-energy rewards [2].

    CO2 adsorption is typically used as a final polishing step in a hybrid CO2 capture system [3]. Ideally, CO2 adsorbents should have the following characteristics: a high specific surface area, well developed micro- and mesopores, and many active sites on the surfaces, such as amine functional groups and basic metal oxides [4-6]. Up to the present, different types of adsorbents have been used for CO2 adsorption, such as zeolites, alumina, mesoporous silica, CaO, and porous carbons [7-11]. In particular, porous carbons have been preferred for CO2 adsorption due to their highly de-veloped porosity, extended surface area, surface chemistry, and thermal stability. Previous studies have also used activated carbons, activated carbon fibers (ACF), carbon molecular sieves, and carbon nanotubes as adsorbents for CO2 adsorption [12-14].

    In general, the adsorption capacity on porous carbon materi-als depends on the surface area and pore structures of adsor-bents. Additionally, the surface carbon/adsorbates interactions have an influence on the adsorption capacity [15-17]. Surface modification is able to change this interaction, and numerous surface modification methods have been published in the litera-ture. Of these methods, fluorination is one of the most effective methods for modifing carbon surfaces [18-21]. Because fluorine gas has a strong reactivity, it can easily react with surface ca-rbon, even at room temperature. Furthermore, the surface nature of activated carbons can be controlled, regardless of whether they are hydrophobic or hydrophilic, by direct fluorination and oxyfluorination [22].

    We modified activated electrospun CFs with fluorine gas, and characterized the changes in its surface properties and pore structure before and after oxyfluorination. Afterwards, we investigated the interaction of adsorbent?adsorbate from CO2 adsorption isotherms of oxyfluorinated activated elec-trospun CFs at 273 K. From these results, we determined the relationship between the storage capacity and the pore struc-tures of the ACF samples. This paper reports our preliminary results on the catalytic effect of ACFs modified by F2 and its potential benefits.

    2. Experimental Methods

       2.1. Materials

    Polyacrylonitrile (PAN, d = 1.184, Aldrich) was used as a car-bon source. N,N-dimethyl formamide (DMF) was adopted as a solvent, because its boiling point is 426 K and it has a conductiv-ity (conductivity = 10.90 μS/cm, dipole moment = 3.82 Debye) sufficiently high for electrospinning. In order to develop the pore structure, potassium hydroxide (KOH) was used as a chemical activation agent. Fluorine (Messer Grieheim GmbH, 99.8%) and oxygen (99.99%) gases were used for the oxyfluorination treat-ment [23,24]. The prepared polymer solution (PAN/DMF, 10 wt%) was ejected from a syringe tip onto an aluminum-foil-cov-ered collector using an electrospinning apparatus. A schematic diagram of the electrospinning apparatus was depicted in our previous paper [25]. Electrospinning was performed under the following conditions: feeding rate of polymer solution: 1 mL/h; supplied voltage: 18 kV; tip to collector distance: 10 cm; collec-tor rpm: 150. The stabilization of the electrospun materials was carried out in air by heating up to 523 K at a heating rate of 1 K/min, and finally, samples were treated at 523 K for 8 h. Carbon-ization of the stabilized electrospun materials was carried out under a nitrogen atmosphere with the following conditions to carbonize the oxidized fibers: heating rate: 10 K/min; reaction temperature: 1323 K; holding time: 1 h nitrogen feeding rate: 100 mL/h [26]. The final product is called an electrospun carbon fiber in this study.

    The KOH solutions (6 M) were prepared as chemical acti-vation agents. The CF sample was placed on an alumina boat at a ratio of 10 mL/g (KOH solution/CF) in a steel pipe to carry out the chemical activation. Activation was conducted at 1023 K for 3 h in an argon atmosphere. The heating rate was 5 K/min, and the feed rate of the argon gas was 40 mL/min. After chemical activation, samples were washed with distilled water several times to remove residual potassium. These were then dried at 383 K overnight. These activated carbon nanofibers are called ACF.

       2.2. Surface modification by oxyfluorination of ACFs

    As a pretreatment step, ACFs were heated at 70℃ for 24 h to remove impurities. The oxyfluorination was carried out at 1 bar for 5 min with oxygen: fluorine gas ratios of 7:3, 5:5 and 3:7. These oxyfluorinated samples are labeled oxyfluorination effects of activated carbon nanofiber (OFACF)-1 (O2:F2 = 7:3), OFACF-2 (O2:F2 = 5:5) and OFACF-3 (O2:F2 = 3:7). More de-tails on this reaction are provided in a previous study [26].

       2.3. Characterization of prepared samples

    The X-ray photoelectron spectroscopy (XPS) spectra of the samples were obtained using a MultiLab 2000 spectrometer (Thermo Electron Corporation, England) in order to investigate the elements present in the samples. Al Ka (1485.6 eV) was used as the X-ray source with a 14.9-keV anode voltage, a 4.6-A fila-ment current and a 20-mA emission current. All samples were treated at 10-9 mbar to remove impurities. The survey spectra were obtained at a 50-eV pass energy and a 0.5-eV step size.

    The textural properties of the all the samples were investi-gated. Samples were degassed at 423 K for 3 h, and then, ni-trogen adsorption was carried out at 77 K using a Brunauer?Emmett?Teller (BET) apparatus (Micromeritics ASAP 2020) to investigate the specific surface area, total pore volume, pore size distribution and micropore fraction.

    The CO2 adsorption capacity was measured by CO2 isother-mal adsorption at 273 K and 1 atm. Furthermore, the parameters of the Dubinin?Astakhov equation:

    image

    were obtained from CO2 adsorption. p, p0, V, V0, n, R, E0, and ββ are the equilibrium pressure, the saturation vapor pressure of the gas at the analysis temperature (T), the volume adsorbed at equilibrium pressure, the monolayer capacity, the Astakhov exponent, the gas constant, characteristic energy, and the affinity coefficient of the analysis gas, respectively.

    3. Results and Discussion

       3.1. Surface properties of oxyfluorinated ACFs

    XPS analysis was very useful for characterizing the surface chemical structures of the oxyfluorinated ACFs in the previous

    [Table 1.] XPS surface element analysis of as-received and OFACFs

    label

    XPS surface element analysis of as-received and OFACFs

    studies [27-29]. Characterization of the surface chemical com-position of the oxyfluorinated ACFs was carried out using XPS analysis. Table 1 lists the atomic ratio of each element on the surface of the ACFs. The surface carbon concentration of the samples decreased dramatically after oxyfluorination, whereas the fluorine atomic ratio of the samples significantly increased. In addition, the O1s peaks of the oxyfluorinated ACFs slightly increased with the decrease in fluorine partial pressure, but the O1s peak of sample OFACF-3 decreased because there was in-sufficient fluorine gas to react with the oxygen gas.

    Fig. 1 shows the elemental survey of pristine and oxyfluori-nated ACFs. In the ACF sample, C1s and O1s peaks represent the carbon atoms from ACFs and the oxygen atoms from the H2O adsorption on ACFs from the ambient atmosphere, respective-ly. The peaks of O1s and F1s confirmed the introduction of both oxygen and fluorine atoms, respectively, on the oxyfluorinated ACFs. The peak intensities varied depending on the composi-tion of the reactive gases. In the C1s spectra of oxyfluorinated ACFs, the peak intensity increased with decreasing fluorination pressure. Fluorine gas appeared to react with the carbon more actively than with the oxygen, resulting in the breakage of the sp2 carbon aromatic structure. The carbon-oxygen single and double-bond peak intensities increased with increasing oxygen content in the reactive gas [30]. In sample OFACF-2, the total content of carbon-oxygen single and double bonds increased by over 20%. These carbon?oxygen bonds contributed to the increased hydrophilicity of the ACFs; while in the OFACF-3 sample, the total content of carbon-oxygen single- and double-bonds decreased because there was insufficient fluorine gas to react with the oxygen gas. The fluorinated carbon bonds were observed as semi-ionic and covalent carbon-fluorine bonds [31]. Even though the intensities of all of the fluorinated carbon bonds increased as the fluorine content in the reactive gas increased, the covalent carbon-fluorine bond was dominant.

    The O1s and F1s peaks were also investigated as supplementary data for interpreting the C1s peaks. The variation in the O1s peaks corresponded well with the variation of the C1s peaks upon oxy-fluorination. The carbon-oxygen single and double bonds were produced more favorably by oxyfluorination. The variation in the F1s peaks also corresponded well with the variation of the C1s peaks upon oxyfluorination. The covalent carbon-fluorine bonds were more dominant when the fluorine content in the reactive gas was increased.

    The changes in the chemical bonds of the ACFs after oxyfluo-rination were investigated by C1s deconvolution, and the results are depicted in Fig. 2. Table 2 lists the C1s spectra and assignments of the oxyfluorinated ACFs. The C1s peaks of oxyfluorinated ACFs was deconvoluted to several pseudo-Voigt functions (sums of the Gaussian and Lorentzian functions) using a peak-analysis program obtained from Unipress Co., USA. The pseudo-Voigt function is given by [32]:

    image

    where F(E) is the intensity at energy E, H is the peak height, E0 is the peak center and S is a shape function related to the symmetry and the Gaussian?Lorentzian mixing ratio. The components ob-served for ACF were C(1), C(2), C(3), and C(4) at 284.5, 285.7, 86.9, and 288.1 eV, corresponding respectively to the sp2 car-bon, sp3 carbon, carbon-oxygen single- and double-bonds. After oxyfluorination, the C(5) and C(6) components were observed in OFACF-1, OFACF-2, and OFACF-3 samples at 289.3 and 290.5 eV, corresponding to semi-covalent and covalent carbon-fluo-rine, respectively. All of these results are in accord with those from other studies [19,21,27-29,33-35]. The concentration of C(3) and C(4) increased as the O2 to F2 ratio increased in the gas mixture, whereas the ratio only decreased in the OFACF-3 sample because there was insufficient fluorine gas to react with

    [Table 2.] C1s spectra and assignment of the OFACFs

    label

    C1s spectra and assignment of the OFACFs

    the oxygen gas, as previously mentioned. However, C(5) and C(6) decreased with a decreasing F2 ratio in the gas. The effect of the functional groups discussed in this section will be further discussed in later sections.

       3.2. Textural properties

    A summary of the textural properties is presented in Table .3 KOH activation has the effect of increas-ing specific surfacearea, total pore volume, and micropore volume while increasing the concentration of the used chemical agent solution. The nitrogen isotherms of the samples are shown in Fig. 3. All of the curves have an inflection point, which is called a ‘knee,’ around P/P0 (0.01). This point indicates that monolayer adsorption is complete. Thus, it is possible to calculate the specific surface area at this point using the BET equation. Generally, in the low-pressure region, less than around 0.1 P/P0, the steep curve of the isotherm indicates the development of the micropore struc-

    [Table 3.] Textural properties of the un-oxyfluorinated and OFACFs

    label

    Textural properties of the un-oxyfluorinated and OFACFs

    ture, and the change in the relatively high-pressure region indi-cates the improvement of meso/macro pores [36]. In the case of OFACF-1, OFACF-2 and OFACF-3, a well-developed mi-croporous property was confirmed by the striking initial change in the low relative pressure range [37]. In our previous work, it was shown that KOH activated carbon nanofibers have highly

    developed micropores, and oxyfluorination treatment of ACFs does not significantly affect the pore structure under the condi-tions used [26].

    The pore size distribution was evaluated by applying a non-local density functional theory (NLDFT) model, as shown in Fig. 4. The NLDFT model was applied under the assumption that the slit-shaped pores have uniformly dense carbon walls and that the adso-rbate is a fluid of hard spheres [38]. Additionally, the interaction between adsorbent and adsorbate was considered on their elemental spe-cies by a soft program installed on the BET apparatus. The pore width was distributed mainly in the micopore range, especially in the range of 0.7-1.6 nm. Based on the infor-mation provided by other groups, this pore size distribu-tion would be very beneficial for CO2 uptake, because it has been predicted that the optimized pore size for enhanc-ing the capacity of CO2 storage in carbon materials is 0.8-1.1 nm [39,40].

       3.3. Capacity of CO2uptake

    Adsorption equilibrium isotherms of CO2 on all samples at 273 K are plotted in Fig. 5. It can be seen that the OFACFs had a better performance in CO2 adsorption than the pris-tine ACFs. This may occur because the filling of micropores with CO2 gas is increase at that pressure [41]. This clearly indicates that modification by KOH activation and oxyfluo-rination treatment increase the affinity of ACFs toward CO2. Comparing ACFs, a higher CO2 uptake can be observed in the case of OFACF-2. The cause is surely the more highly-developed pore structure. Based on the investigation of oxyfluorination effects for enhancing CO2 storage, the CO2 uptake of the oxyfluorinated samples was studied. In all pres-sure ranges, oxyfluorination effects can be observed, showing a higher capacity of CO2uptake than raw ACF. It is around 11% (OFACF-1 and OFACF-3) and 30% (OFACF-2) higher, as compared with ACF sample, respectively. Overall, the CO2 uptakes of the samples were 11.2 (ACF), 14.2 (OFACF-1), 16.2 (OFACF-2) and 13.9 wt% (OFACF-3).

       3.4. Suggested mechanism for the improved effects of oxyfluorinated ACFs

    The mechanism of CO2 storage in the slit pores of oxyfluorinated ACFs is suggested in Fig. 6. The CO2 molecule has a non-polar structure. The CO2 molecules near the carbon pores are affected by the semi-ionic interaction of oxygen groups, which has the lone-pair electron, causing the attraction of the electrons in the CO2 mol-ecules,as shown in Fig. 6. This reaction might play an important role as a guide for enhancing the CO2 storage capacity. Eventually, the CO2 gas can be stored in the slit pores of carbon. Some of the residual CO2 molecules in the carbon silt pores can be also affected by oxygen group effects, such as the grabbing effects due to semi ionic interaction, resulting in the high efficiency of CO2 storage. All of these reactions were carried out without any chemical bonds between the CO2 molecules and the carbon pores, so the desorption of CO2 is easily reversible for ease of use.

    4. Conclusions

    In conclusion, we have shown that the CO2 adsorption be-haviors of ACFs are directly related to oxyfluorination. Samples were oxyfluorinated for surface modification to investigate the semi-ionic interaction effect of enhancing the capacity of CO2 storage by oxygen group. The resulting samples were highly microporous, over 85%, with a high specific surface area greater than 1200 m2/g and a pore size distribution in the range of 0.7-1.6 nm, optimal for CO2 storage. Eventually, the CO2 uptake increased up to 16.2 wt% due to the highly developed micropore structure and semi-ionic interaction effect of oxyfluorination.

참고문헌
  • 1. Reay DS, Dentener F, Smith P, Grace J, Feely RA (2008) Global nitrogen deposition and carbon sinks. [Nature Geosci] Vol.1 P.430 google cross ref
  • 2. (2007) Tracking Industrial Energy Efficien-cyand CO2 Emissions: In Support of the G8 Plan of Action: Energy Indicators google
  • 3. Aaron D, Tsouris C (2005) Separation of CO2 from flue gas: a review. [Sep Sci Technol] Vol.40 P.321 google cross ref
  • 4. Meng L, Cho KS, Park SJ (2009) CO2 adsorption of amine functionalized activated carbons. [Carbon Lett] Vol.10 P.221 google
  • 5. Meng L, Cho KS, Park SJ (2010) Effect of heat treatment on CO2 adsorp-tionof ammonized graphite nanofibers. [Carbon Lett] Vol.11 P.34 google cross ref
  • 6. Kim BJ, Cho KS, Park SJ (2010) Copper oxide-decorated porous carbons for carbon dioxide adsorption behaviors. [J Colloid Interface Sci] Vol.342 P.575 google cross ref
  • 7. Yang H, Xu Z, Fan M, Gupta R, Slimane RB, Bland AE, Wright I (2008) Progress in carbon dioxide separation and capture: a review. [J Environ Sci] Vol.20 P.14 google cross ref
  • 8. Chatti R, Bansiwal AK, Thote JA, Kumar V, Jadhav P, Lokhande SK, Biniwale RB, Labhsetwar NK, Rayalu SS (2009) Amine loaded ze-olitesfor carbon dioxide capture: amine loading and adsorption studies. [Microporous Mesoporous Mater] Vol.121 P.84 google cross ref
  • 9. Caixin L, Lei Z, Jiguang D, Qing M, Hongxing D, Hong H (2008) Surfac-tant-aided hydrothermal synthesis and carbon dioxide adsorption behavior of three-dimensionally mesoporous calcium oxide single-crystallites with tri- tetra- and hexagonal morphologies. [J PhysChem C] Vol.112 P.19248 google cross ref
  • 10. Plaza MG, Pevida C, Arias B, Fermoso J, Arenillas A, Rubiera F, Pis JJ (2008) Application of thermogravimetric analysis to the evaluation of aminated solid sorbents for CO2 capture. [J Therm Anal Calorim] Vol.92 P.601 google cross ref
  • 11. Siriwardane RV, Shen MS, Fisher EP, Poston JA (2001) Adsorption of CO2 on molecular sieves and activated carbon. [Energy Fuels] Vol.15 P.279 google cross ref
  • 12. Tenney CM, Lastoskie CM (2006) Molecular simulation of carbon diox-ide adsorption in chemically and structurally heterogeneous porous carbons. [Environ Prog] Vol.25 P.343 google cross ref
  • 13. Moon SH, Shim JW (2006) A novel process for CO2/CH4 gas separa-tion on activated carbon fibers-electric swing adsorption. [J Col-loidI nterface Sci] Vol.298 P.523 google cross ref
  • 14. Hu YH, Ruckenstein E (2006) Applicability of Dubinin-Astakhov equation to CO2 adsorption on single-walled carbon nanotubes. [Chem Phys Lett] Vol.425 P.306 google cross ref
  • 15. Zhao XB, Xiao B, Fletcher AJ, Thomas KM (2005) Hydrogen adsorption on functionalized nanoporous activated carbons. [J Phys Chem B] Vol.109 P.8880 google cross ref
  • 16. Celzard A, Perrin A, Albiniak A, Broniek E, Mareche JF (2007) The effect of wetting on pore texture and methane storage ability of NaOH activated anthracite. [Fuel] Vol.86 P.287 google cross ref
  • 17. Chingombe P, Saha B, Wakeman RJ (2005) Surface modification and characterisation of a coal-based activated carbon. [Carbon] Vol.43 P.3132 google cross ref
  • 18. Touhara H, Okino F (2000) Property control of carbon materials by fluo-rination. [Carbon] Vol.38 P.241 google cross ref
  • 19. Lee YS (2007) Syntheses and properties of fluorinated carbon materi-als. [J Fluorine Chem] Vol.128 P.392 google cross ref
  • 20. Tressaud A, Durand E, Labrugere C (2004) Surface modification of sev-eral carbon-based materials: comparison between CF4 rf plasma and direct F2-gas fluorination routes. [J Fluorine Chem] Vol.125 P.1369 google cross ref
  • 21. Lee YS, Lee BK (2002) Surface properties of oxyfluorinated PAN-based carbon fibers. [Carbon] Vol.40 P.2461 google cross ref
  • 22. Lee YS, Kim YH, Hong JS, Suh JK, Cho GJ (2007) The adsorption prop-erties of surface modified activated carbon fibers for hydrogen storages. [Catal Today] Vol.120 P.420 google cross ref
  • 23. Im JS, Park SJ, Kim T, Lee YS (2009) Hydrogen storage evaluation based on investigations of the catalytic properties of metal/metal oxides in electrospun carbon fibers. [Int J Hydrogen Energy] Vol.34 P.3382 google cross ref
  • 24. Im JS, Yun J, Lim YM, Kim HI, Lee YS (2010) Fluorination of electros-pun hydrogel fibers for a controlled release drug delivery system. [Acta Biomater] Vol.6 P.102 google cross ref
  • 25. Im JS, Park SJ, Lee YS (2007) Preparation and characteristics of electro-spun activated carbon materials having meso- and macropores. [J Colloid Interface Sci] Vol.314 P.32 google cross ref
  • 26. Im JS, Park SJ, Lee YS 2009 The metal-carbon-fluorine system for im-proving hydrogen storage by using metal and fluorine with differ-ent levels of electronegativity. [Int J Hydrogen Energy] Vol.34 P.1423 google cross ref
  • 27. Ho KKC, Lee AF, Bismarck A (2007) Fluorination of carbon fibres in atmospheric plasma. [Carbon] Vol.45 P.775 google cross ref
  • 28. Bismarck A, Tahhan R, Springer J, Schulz A, Klapotke TM, Zell H, Michaeli W (1997) Influence of fluorination on the properties of carbon fibres. [J Fluorine Chem] Vol.84 P.127 google cross ref
  • 29. Tressaud A, Durand E, Labrugere C, Kharitonov AP, Kharitonova LN (2007) Modification of surface properties of carbon-based and poly-meric materials through fluorination routes: from fundamental re-search to industrial applications. [J Fluorine Chem] Vol.128 P.378 google cross ref
  • 30. Park SJ, Kim BJ (2005) Ammonia removal of activated carbon fibers pro-ducedby oxyfluorination. [J Colloid Interface Sci] Vol.291 P.597 google cross ref
  • 31. Chamssedine F, Claves D (2008) Selective substitution of fluorine atoms grafted to the surface of carbon nanotubes and application to an oxyfluorination strategy. [Carbon] Vol.46 P.957 google cross ref
  • 32. Kauffman DR, Sorescu DC, Schofield DP, Allen BL, Jordan KD, Star A (2010) Understanding the sensor response of metal-deco-ratedcarbon nanotubes. [Nano Lett] Vol.10 P.958 google cross ref
  • 33. Xu B, Wang X, Lu Y (2006) Surface modification of polyacrylonitrile-based carbon fiber and its interaction with imide. [Appl Surf Sci] Vol.253 P.2695 google cross ref
  • 34. Crassous I, Groult H, Lantelme F, Devilliers D, Tressaud A, Labru-gere C, Dubois M, Belhomme C, Colisson A, Morel B (2009) Study of the fluorination of carbon anode in molten KF-2HF by XPS and NMR investigations. [J Fluorine Chem] Vol.130 P.1080 google cross ref
  • 35. Ma K, Chen P, Wang B, Cui G, Xu X (2010) A study of the effect of oxygen plasma treatment on the interfacial properties of carbon fi-ber/epoxy composites. [J Appl Polym Sci] Vol.118 P.1606 google cross ref
  • 36. Gregg SJ, Sing KSW (1982) Adsorption Surface Area and Porosity. google
  • 37. Im JS, Woo SW, Jung MJ, Lee YS (2008) Improved capacitance charac-teristics of electrospun ACFs by pore size control and vanadium catalyst. [J Colloid Interface Sci] Vol.327 P.115 google cross ref
  • 38. Ravikovitch PI, Neimark AV (2001) Characterization of nanoporous ma-terialsfrom adsorption and desorption isotherms. [Colloids Surf A: Physicochem Eng Aspects] Vol.11 P.187-188 google cross ref
  • 39. Celzard A, Albiniak A, Jasienko-Halat M, Mareche JF, Furdin G (2005) Methane storage capacities and pore textures of active carbon sundergoing mechanical densification. [Carbon] Vol.43 P.1990 google cross ref
  • 40. Yeon SH, Osswald S, Gogotsi Y, Singer JP, Simmons JM, Fischer JE, Lillo-Rodenas MA, Linares-Solano A (2009) Enhanced methane stor-age of chemically and physically activated carbide-derived carbon. [J Power Sources] Vol.191 P.560 google cross ref
  • 41. Im JS, Park SJ, Kim TJ, Kim YH, Lee YS (2008) The study of control-lingpore size on electrospun carbon nanofibers for hydrogen adsorption. [J Colloid Interface Sci] Vol.318 P.42 google cross ref
이미지 / 테이블
  • [ Table 1. ]  XPS surface element analysis of as-received and OFACFs
    XPS surface element analysis of as-received and OFACFs
  • [ Fig. 1. ]  Elemental-survey data of the un-oxyfluorinated and oxyfluori-nated activated carbon nanofibers (OFACFs).
    Elemental-survey data of the un-oxyfluorinated and oxyfluori-nated activated carbon nanofibers (OFACFs).
  • [ Fig. 2. ]  C1s deconvolution of the un-oxyfluorinated and oxyfluorinated activated carbon nanofibers (OFACFs). (a) ACF, (b) OFACF-1, (c) OFACF-2, and (d)OFACF-3.
    C1s deconvolution of the un-oxyfluorinated and oxyfluorinated activated carbon nanofibers (OFACFs). (a) ACF, (b) OFACF-1, (c) OFACF-2, and (d)OFACF-3.
  • [ Table 2. ]  C1s spectra and assignment of the OFACFs
    C1s spectra and assignment of the OFACFs
  • [ Table 3. ]  Textural properties of the un-oxyfluorinated and OFACFs
    Textural properties of the un-oxyfluorinated and OFACFs
  • [ Fig. 3. ]  Adsorption isotherms of nitrogen on the un-oxyfluorinated and oxyfluorinated activated carbon nanofibers (OFACFs).
    Adsorption isotherms of nitrogen on the un-oxyfluorinated and oxyfluorinated activated carbon nanofibers (OFACFs).
  • [ Fig. 4. ]  Non-local density functional theory pore size distribution. OFACF: oxyfluorinated activated carbon nanofiber.
    Non-local density functional theory pore size distribution. OFACF: oxyfluorinated activated carbon nanofiber.
  • [ Fig. 5. ]  Adsorption equilibrium isotherms of CO2on all samples. OFACF: oxyfluorinated activated carbon nanofiber.
    Adsorption equilibrium isotherms of CO2on all samples. OFACF: oxyfluorinated activated carbon nanofiber.
  • [ Fig. 6. ]  Suggested mechanism for the improved effects of oxyfluori-natedactivated carbon nanofibers.
    Suggested mechanism for the improved effects of oxyfluori-natedactivated carbon nanofibers.
(우)06579 서울시 서초구 반포대로 201(반포동)
Tel. 02-537-6389 | Fax. 02-590-0571 | 문의 : oak2014@korea.kr
Copyright(c) National Library of Korea. All rights reserved.