Anda di halaman 1dari 8

International Journal of Hydrogen Energy 33 (2008) 116 123 www.elsevier.

com/locate/ijhydene

CO2 activation of ordered porous carbon CMK-1 for hydrogen storage


Kaisheng Xia, Qiuming Gao , Shuqing Song, Chundong Wu, Jinhua Jiang, Juan Hu, Lu Gao
State Key Laboratory of High Performance Ceramics and Superne Microstructures, Shanghai Institute of Ceramics, Chinese Academy of Sciences, 1295 Dingxi Road, Shanghai 200050, Peoples Republic of China Received 12 July 2007; received in revised form 29 August 2007; accepted 29 August 2007 Available online 10 October 2007

Abstract Porous carbons with large surface areas and pore volumes, as well as relatively uniform micropore texture were prepared by CO2 activation of ordered porous carbon CMK-1 for hydrogen storage. Both activation temperature and time were studied. The structure and texture changes in the carbons were investigated by using XRD and nitrogen sorption. Under the optimum condition that treating with CO2 at 950 C for 6 h, a high H2 uptake of 2.19 wt% was obtained at 77 K and 1 bar. The capacity is almost twice that of CMK-1 and higher than that of the reported porous carbons synthesized by silica templates. The pores smaller than 1 nm were conrmed to be the most efcient for hydrogen sorption. Besides, high efciency of surface areas (4.50 106 mol/m2 ) and micropores (1.36 102 mol/cm3 ) were got for the H2 uptake at 77 K and 1 bar. 2007 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
Keywords: Porous carbon; Activation; Adsorption; Hydrogen storage

1. Introduction Hydrogen has been recognized as an ideal energy carrier, because it is clean and easily produced from renewable energy sources and contains higher chemical energy per mass than hydrocarbon fuels [1]. In the past several decades, efcient hydrogen storage and transportation have been the major concerns in the hydrogen technology. The method of storing as compressed hydrogen or liqueed hydrogen is unsuitable for wide application because of its low storage density or high cost. Chemisorption of hydrogen in the form of metals or complex hydrides is an interesting method, but these materials are relatively expensive and the reversible capacity is still far from the DOE target for application (6.5 wt% and 62 kg H2 /m3 ) [1]. In recent years, physisorption of hydrogen on carbon-based nanomaterials or other porous materials have attracted greatly scientic interests. Nanostructured carbons are among the major candidates of physisorption for their lightweight, abundant natural precursors and low cost. The activated carbons have been reported to have relatively high hydrogen storage capacity at 77 K [2,3], but the
Corresponding author. Tel.: +86 21 52412513; fax: +86 21 52413122.

E-mail address: qmgao@mail.sic.ac.cn (Q. Gao).

pore size distribution is generally wide and more than half of the total porous volumes come from macropores, which contribute less to the uptake of hydrogen. Although carbon nanotubes (CNTs) and carbon nanobers (CNFs) have been studied as hydrogen storage materials, but there are still disputes about their capacities of hydrogen adsorption [410]. Therefore, a more suitable hydrogen storage medium is urgently needed. Recently, porous carbons with well-ordered pore systems offered great potential in hydrogen storage [1114]. The carbons were obtained via the template method, which involved the introduction of suitable carbon precursors into the ordered pores of the template followed by carbonization and nally removal of the template [1517]. These carbon materials usually have large specic surface areas and high pore volumes, which are useful for effective physisorption of H2 . Besides, the ordered networks may provide fast transportation in the materials, a noticeable volume of micropores can efciently adsorb hydrogen, and the micro- and the mesoporosity can be adjusted by changing the template, the carbon precursor and the amount of carbon inltrated in the template [11]. At the same time, activation has also been reported to be useful for improving the pore structure and increasing the H2 uptake in new nanostructured carbon materials such as SWNTs [18] and CNFs [19]. Thus,

0360-3199/$ - see front matter 2007 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.ijhydene.2007.08.019

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123

117

activation of the ordered porous carbon is desirable for getting a higher H2 uptake. In this work, ordered porous carbon designated as CMK-1 [15] was synthesized by replication from MCM-48 silica, and then was activated with carbon dioxide under different conditions in order to gain a higher hydrogen storage capacity. The structural and morphological changes caused by the CO2 activation were investigated by using X-ray diffraction (XRD) and nitrogen sorption at 77 K. The H2 sorption densities of the materials were measured at 77 K under the pressure between 0 and 1 bar. Moreover, we attempted to nd out the key factors in hydrogen storage, which may help to understand the H2 adsorption mechanism and direct the design of carbon-based hydrogen adsorbents. 2. Experimental 2.1. Preparation of MCM-48 template MCM-48 was synthesized following the literatures [2023]. In a typical synthesis, 5.2 g of cetyltrimethylammonium bromide (CTAB, AR) was completed dissolved in 240 g of distilled water and 100 mL of ethanol. Then, 30 mL of ammonia solution (AR, 25 wt%) was added to the surfactant solution. The mixture was stirred for 30 min and then 6.8 g of tetraethoxyorthosilane (TEOS, AR) was added. The molar composition of the gel was 1.0 M TEOS: 12.5 M NH3 : 54 M EtOH: 0.4 M CTAB: 425 M H2 O. After being stirred for 12 h at room temperature, the resultant solid was separated by ltration, washed with deionized water, dried in air at ambient temperature, and calcined in air at 550 C for 6 h to obtain the MCM-48. 2.2. Synthesis of CMK-1 Ordered porous carbon CMK-1 was synthesized via a two step impregnation of the mesopores of MCM-48 with a solution of sucrose using an incipient wetness method [1517]. Briey, 1.0 g of the as-prepared MCM-48 was impregnated with an aqueous solution obtained by dissolving 1.1 g of sucrose and 0.14 g of H2 SO4 in 5.0 g of deionized water. The mixture was then dried at 100 C for 6 h, and subsequently at 160 C for 6 h. The silica sample, containing partially polymerized and carbonized sucrose, was treated again at 100 and 160 C after the addition of 0.65 g of sucrose, 90 mg of H2 SO4 and 5.0 g of deionized water. The sucrosesilica composite was then heated at 900 C for 4 h under nitrogen to complete the carbonization. The silica template was dissolved with 5 wt% hydrouoric acid at room temperature. The template-free carbon product thus obtained was ltered, washed with deionized water and ethanol, and dried. 2.3. Activation of CMK-1 The as-prepared CMK-1 was placed in the centre of a quartz tube in a tube furnace and heated (5 C/min) to the activation temperature under nitrogen, and then the gas stream was

switched to CO2 (50 cm3 / min) [19,24]. Finally, the sample was cooled to room temperature under nitrogen. The obtained samples were denoted as C850-2, C900-2 and C950-2 for 2 h treatment in the gas stream of CO2 at 850, 900 and 950 C, respectively, and C950-4 and C950-6 represented the samples obtained after 4 and 6 h activation in CO2 at 950 C. 2.4. Materials characterization and hydrogen adsorption Thermogravimetric analyses (TGA) were carried out with a STA-409PC/4/H Luxx simultaneous TG-DTA-DSC apparatus (Netzsch, Germany) with a heating rate of 10 C/min in owing air atmosphere. XRD patterns were recorded on a Rigaku D/MAX-2250V diffractometer using Cu K radiation (40 kV and 40 mA). The scanning rate was 0.6 / min for the low-angle XRD measurements. The packing density was determined by pressing a given amount of sample into a glass tube (inner diameter 0.6 cm) with a cylindrical rod. The measurement of the height of the rod with the tube empty and with the sample permits the evaluation of the volume of the material and then the packing density. The particle sizes of the carbons are determined by using Brokhavens Zeta Plus particle size analyzer. The results showed that the average particle sizes of the carbons are in the range of 0.50.7 m. Volumetric nitrogen and hydrogen sorption studies were taken at 77 K using a Micromeritics ASAP 2020 system. Before measurements, the samples were degassed below 1.33 Pa at 90 C for 1 h and heated (10 C/min) to 350 C for 10 h. The specic surface area (SBET ) was calculated by the BET method in the relative pressure range of 0.040.20. Total pore volume (Vt ) was calculated at relative pressure p/p0 = 0.98. The microporous volume (Vmi ) was determined by applying DubininRadushkevich (DR) analyses on the corresponding isotherms in the relative pressure range 104 102 . The volume of pores smaller than 1 nm (V<1 nm ) was determined by the cumulative pore volume in the relative pressure range 106 104 using HorvathKawazoe (HK) method. The meso- and micropore sizes of samples were analyzed by the BarrettJoynerHalenda (BJH) and HK methods, respectively. Finally, H2 adsorption was measured by volumetric method under the pressure range of 01 bar at 77 K by using the Micromeritics ASAP 2020 system. 3. Results and discussion 3.1. Structural and textural characterization of MCM-48 and CMK-1 TGA analyses of the ordered porous carbon CMK-1 revealed that all carbon burned out in air from room temperature to 800 C and negligible silica residue (< 2.0 wt%) was left, indicating that the silica were successfully removed by the treatment with HF. Fig. 1 shows typical XRD patterns for the silica template MCM-48 and the carbon CMK-1 obtained following the above procedures. The low-angle XRD pattern of MCM-48 indicates excellent structural order for the cubic crystallographic space group I a3d. However, the carbon CMK-1, which was

118

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123


Table 1 Textural properties of MCM-48 and CMK-1
Samples MCM-48 CMK-1
(211)

(110)

SBET (m2 /g) 1736 1788

Vt (cm3 /g) 0.97 1.06

DBJH (nm)a 2.4 2.4

Intensity (a.u.)

a Peak
CMK-1 (220)

maxima of mesopore size distribution by BJH method.

(211)

MCM-48 (220) (420) (332)

3 2 (deg.)

Fig. 1. XRD patterns of MCM-48 and CMK-1.

800 Quantity adsorbed (cm3/g STP) ) 700 600 500


dV/dw MCM-48 CMK-1

in the amount of adsorption at relative pressures higher than 0.30. The detailed BET specic surface areas and the total pore volumes are listed in Table 1. The ordered mesoporous silica MCM-48 and the CMK-1 carbon showed comparative large specic surface areas of 1736 and 1788 m2 /g, respectively. The total pore volumes corresponding to MCM-48 and CMK-1 were 0.97 and 1.06 cm3 /g, respectively. In addition to the mesopores with the diameter of 2.4 nm, the structure of CMK-1 contains abundant micropores which were formed in the amorphous carbon framework prior to the removal of silica wall, while the mesopores were formed upon the removal of the silica wall. The more detailed pore structure of CMK-1 will be presented in latter part. 3.2. Effect of activation temperatures Three temperatures of 850, 900 and 950 C were selected in our experiment for comparison to nd out the optimum activation temperature, since the appropriate temperature range of CO2 activation may be between 800 and 1000 C [19,24]. The mass loss and packing density of the carbons are included in Table 2. As a result of 2 h treatment, the carbon burn off increased with activation temperature and the packing density decreased from 0.29 for CMK-1 to 0.24 g/cm3 for C950-2. XRD patterns of the unactivated CMK-1 and activated samples C850-2, C900-2 and C950-2 are shown in Fig. 3. The obvious diffraction peaks of the (1 1 0), (2 1 1) and (2 2 0) reections indicate the highly ordered structure of CMK-1. The (1 1 0) and (2 1 1) peaks for the CMK-1 carbon were retained after the activation by CO2 and the intensities were slightly reduced along with the increase of activation temperature. These results suggest that the primary structure of CMK-1 was unchanged after the 2 h treatment in CO2 at different temperatures. At the same time, the diffraction peaks became broader and shifted to higher angles along with the increase of activation temperature. The differences may be due to the lattice contraction of the framework because of heating treatment and formation of more micropores during the activation process. Fig. 4a shows the nitrogen sorption isotherms for the ordered carbon CMK-1 and those activated with CO2 at different temperatures. Typical I/II isotherms were found for those samples. At relative pressure below 0.1, which was ascribed to micropore lling, sample C850-2 showed an obvious enhancement in adsorption relative to the as-prepared CMK-1. With the increase of activation temperature, a further increase of nitrogen adsorption was found in samples C900-2 and C950-2. At very low pressure C950-2 had the highest nitrogen sorption, which

400 300 200 100 0 0.0 0.2

Pore width (nm)

0.4 0.6 Relative pressure (P/Po)

0.8

1.0

Fig. 2. N2 adsorption/desorption isotherms at 77 K for MCM-48 and CMK-1. The inset shows the corresponding pore size distribution curves calculated using BJH method.

obtained by removing the silica wall after carbonization, exhibits a different pattern relative to MCM-48. Ryoo et al. [15] have ascribed this change to a systematic transformation to a new ordered pore structure in the resultant carbon CMK-1 after the removal of silica framework. Further textural information is provided by nitrogen adsorption isotherms (Fig. 2). The corresponding pore size distributions of MCM-48 and CMK-1 are also provided. After removal of the silica template, the nitrogen adsorption isotherm of CMK-1 showed a steadily increase in the relative pressure range of 00.3, the distinct adsorption step in the relative pressure range of 0.10.3 disappeared and the nitrogen adsorption isotherms changed from type IV of MCM-48 to type I/II of CMK-1. According to the IUPAC classication, the type I/II isotherms may be often observed on microporous solids. The CMK-1 carbon only showed a very small increase

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123


Table 2 Textural character and hydrogen sorption capacity of CMK-1 and those activated with CO2 at different temperatures
Samples Burn off (%) 0 12 18 28 Packing density (g/cm3 ) 0.29 0.27 0.27 0.24 SBET (m2 /g) 1788 1889 1986 2373 Vt (cm3 /g) 1.06 1.09 1.16 1.43 Vmi (cm3 /g) 0.53 0.61 0.66 0.81 V<1 nm (cm3 /g) 0.17 0.25 0.27 0.31 DBJH a (nm) 2.4 2.0 2.0 2.1 DHK b (nm) 0.74 0.72 0.71 0.68

119

H2 sorption (wt%) 1.12 1.57 1.79 1.97

CMK-1 C850-2 C900-2 C950-2


a Peak b The

maxima of mesopore size distribution by BJH method. median micropore sizes analyzed by the HK method.

(110)

CMK-1

1000 Quantity adsorbed (cm3/g STP)


(211)

C850-2 C900-2

800

C950-2

Intensity (a.u.)

(220) CMK-1

600

C850-2

400

C900-2

200

C950-2

0 0.0 0.2

3 2 (deg.)

0.4 0.6 Relative pressure (P/Po)

0.8

1.0

Fig. 3. XRD patterns of the as-prepared CMK-1 and those activated with CO2 at different temperatures.

CMK-1 C850-2
dV/dw 2

C900-2 C950-2

implies the highest volume of ultramicropores (D < 0.7 nm) and supermicropores (0.7 < D < 2 nm) [25] in the sample activated with CO2 at 950 C. This suggests that higher activation temperature is helpful for the forming of micropores. Along with the increase of relative pressures, a steady increase in adsorption capacity was caused by the monolayer/multilayer adsorption of nitrogen molecules in these interconnected pore channels. However, the carbons only showed a very small increase in the amount of adsorption at relative pressures higher than 0.30, which may be ascribed to the adsorption of nitrogen molecules on the mesopores (2 D < 50 nm) or macropores (50 nm D) [26]. Fig. 4b shows the pore distribution curves of the carbons. The pore size distributions analyzed by the HK method showed a distribution of micropores ranging between 0.4 and 0.8 nm. It can be seen that the micropores about 0.6 nm, which has been reported to be optimum pore size for hydrogen storage [12,13], increased clearly with the increase of activation temperature. Moreover, the pore sizes analyzed by the BJH method showed that the primary mesopore size was reduced from about 2.4 nm for the unactivated sample to about 2.0 nm for the activated samples. The textural properties derived from adsorption data are summarized in Table 2. All activated samples have high surface area

dV/dW

4 6 8 Pore width (nm)

10

0.4

0.5

0.6

0.7 0.8 Pore width (nm)

0.9

1.0

1.1

Fig. 4. N2 adsorption/desorption isotherms at 77 K (a) and pore size distribution (b) for the as-prepared CMK-1 and those activated with CO2 at different temperatures.

(SBET ) and total pore volume (Vt ). The surface areas increased gradually from 1788 for CMK-1 to 1889 for C850-2, 1986 for C900-2 and 2373 m2 /g for C950-2. Correspondingly, the total pore volumes increased from 1.06 to 1.43 cm3 /g. The carbons also exhibited an obvious enhancement of microporous volume from 0.53 to 0.81 cm3 /g. Remarkably, a particularly high microporosity with a proportion of Vmi (57%) was obtained in

120

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123

2.0 H2 adsorption capacity (wt.%)

CMK-1 C850-2 C900-2

(110)

1.6

C950-2

Intensity (a.u.)

(211)

1.2

(220) CMK-1 C950-2

0.8

0.4
C950-4

0.0 0.0

C950-6

0.2

0.4 0.6 Relative pressure (P/Po)

0.8

1.0
1 2 3 2 (deg.)
Fig. 6. XRD patterns for the as-prepared CMK-1 and those activated at 950 C for different times.

Fig. 5. Hydrogen adsorption isotherms of CMK-1 and those activated with CO2 at different temperatures.

sample C950-2. The great enhancements are probably due to the fact that during activation the closed pores are opened, new narrow micropores are formed, and the pre-existent is widened [24]. It is noteworthy that the volume of micropores smaller than 1 nm (V<1 nm ) mounted up quickly with the increase of activation temperature. This indicates that activation with CO2 can effectively enhance the formation of ultramicropores and supermicropores, which is perfectly supported by the pore size distribution data based on the HK method analyses (Fig. 4b). It should be noted that the activation can also bring a quantity of pores bigger than 2 nm and that capacity increases along with the increase of the activation temperature, which can be inferred from the pore volumes in Table 2. The hydrogen adsorption of the carbon materials was measured by volumetric method at 77 K over the pressure range from 0 to 1 bar. The H2 uptake isotherms of CMK-1 and samples activated at different temperature are shown in Fig. 5 with the H2 adsorption capacities listed in Table 2. The amount of hydrogen uptake increased signicantly from 1.12 for CMK-1 to 1.57 for C850-2, 1.79 for C900-2 and 1.97 wt% for C950-2. These improvements should be related to the changes of textural characters caused by the activation, such as enhancements of the surface areas and pore volumes especially in micropores. Obviously, sample C950-2 exhibits the highest hydrogen sorption capacity, so 950 C may be considered as the most appropriate activation temperature in this case. 3.3. Effect of activation time In order to achieve higher hydrogen sorption capacity, the effect of different activation time at 950 C on H2 uptake was studied. The carbon burn off during the activation process and the packing density of the resulted carbons are shown in Table 3. After 6 h activation, the carbon burn off was about 57% and the packing density decreased to 0.18 g/cm3 . Fig. 6 exhibits the XRD patterns of samples activated at 950 C for different times. For comparison, the curve of

CMK-1 is also shown. It is found that the intensity of XRD peaks of the activated samples became weaker than that of the precursor CMK-1. Along with the increase of activation time, the intensities of the (1 0 0) peaks were reduced and the (1 1 0) and (2 0 0) peaks were hardly seen. When the activation time was prolonged to 6 h, no obvious XRD peak could be observed for C950-6, which indicates that the ordered pore arrangement is destroyed. Fig. 7 exhibits the nitrogen sorption isotherms and pore size distribution curves of the ordered carbon CMK-1 and the activated samples C950-2, C950-4 and C950-6. It is seen that the isotherms can also be classied as type I/II. As shown in Fig. 7a, the nitrogen adsorption isotherms of all activated carbons at 950 C showed very high adsorption at low pressure relative to CMK-1, indicating the huge increase of micropores. At the same time, along with the increase of activation time, the isotherms of C950-4 and C950-6 showed higher nitrogen sorption than C950-2 at relative pressure above 0.1, which may be attributed to the increased adsorption amounts in small mesopores and a few macropores formed from the interparticle voids. The corresponding specic surface areas and the total pore volumes obtained from the nitrogen adsorption isotherms are listed in Table 3. The surface areas increased from 1788 for CMK-1 to 2741 m2 /g for C950-4, and the total pore volumes increased from 1.06 to 1.87 cm3 /g for C950-6, respectively. Similarly, a steady increase of microporous volume could be found for these carbons. Samples C950-4 and C950-6 exhibit the largest microporous volume (0.94 cm3 /g) and very high proportion of microporous volume, i.e., 54% and 50%, respectively. These textural properties such as BET surface area and microporous volume are comparable to the highest values ever reported [27,28] for porous carbons. It can also be found that the volume of pores smaller than 1 nm gradually increased along with the increase of activation time. For instance, V<1 nm for C950-4 was almost double that of CMK-1. More information about the pore structure can be found in Fig. 7b. Obviously,

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123


Table 3 Textural character and hydrogen sorption capacity of CMK-1 and those activated at 950 C for different times
Samples Burn off (%) 0 28 45 57 Packing density (g/cm3 ) 0.29 0.24 0.21 0.18 SBET (m2 /g) 1788 2373 2741 2696 Vt (cm3 /g) 1.06 1.43 1.73 1.87 Vmi (cm3 /g) 0.53 0.81 0.94 0.94 V<1 nm (cm3 /g) 0.17 0.31 0.33 0.32 DBJH (nm) 2.4 2.1 2.1 < 2.0 DHK (nm) 0.74 0.68 0.73 0.73

121

H2 sorption (wt%) 1.12 1.97 2.16 2.19

CMK-1 C950-2 C950-4 C950-6

2.4
1200 Quantity adsorbed (cm3/g STP) 1000 800 600 400 200 0 0.0
CMK-1

CMK-1

H2 adsorption capacity (wt.%)

C950-2 C950-4 C950-6

2.0 1.6 1.2 0.8 0.4 0.0

C950-2 C950-4 C950-6

0.2

0.4 0.6 Relative pressure (P/Po)

0.8

1.0

0.0

0.2

0.4 0.6 Relative pressure (P/Po)

0.8

1.0

dV/dw

CMK-1 C950-2 C950-4 C950-6

Fig. 8. Hydrogen adsorption isotherms of CMK-1 and those activated at 950 C for different times.

10

Pore Width (nm)

0.4

0.5

0.6

0.7 0.8 0.9 Pore width (nm)

1.0

1.1

Fig. 7. N2 adsorption/desorption isotherms at 77 K (a) and pore size distribution (b) for the as-prepared CMK-1 and those activated at 950 C for different times.

between C950-4 and C950-6. The highest amount of hydrogen uptake 2.19 wt% was obtained for C950-6, which is almost twice that of CMK-1, and is higher than that of the reported porous carbons synthesized by silica templates [11,29]. We attribute the notable increase of hydrogen storage capacity to the change of textural property caused by activation, such as the increases of specic surface areas and volumes of pores with suitable pore sizes. Besides, the calculated volumetric H2 adsorption capacities can be obtained by multiplying the gravimetric H2 sorption by packing density. The results show that the volumetric capacity increased from 3.25 for CMK-1 to 4.73 for C950-2 and 3.94 kg/m3 for C950-6. Although the values are still low, the capacities at high pressure may be desirable for pursuing the DOE target. 3.4. Effect of textural properties on H2 uptake

increasing the activation time may lead to more micropores in the ultramicropore region (0.40.7 nm). At the same time, the primary mesopore diameter by the BJH method decreased from 2.1 nm for C950-2 to less than 2.0 nm for C950-6. H2 sorption isotherms at 77 K from 0 to 1 bar of CMK-1 and those carbons activated at 950 C for different time are shown in Fig. 8 with the adsorption capacities listed in Table 3. It is seen that the hydrogen uptake increased but the rate of increase decreased along with the increase of activation time. No distinct difference of H2 uptake exists

dV/dW

As to the hydrogen storage at 77 K, physisorption mechanism is dominant and H2 uptake is governed by the textural properties of host material. Several literatures [14,27,29,30] showed a tight relationship existed between the specic surface areas and the hydrogen adsorption densities. At the same time, many studies [1113,3133] revealed that the micropores (D < 2 nm) or the even smaller pores were essential for hydrogen storage. In this study, we considered the specic surface area (SBET ) and the micropore volumes (Vmi , V<1 nm ) as the main factors in

122

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123

2.4 2.1 H2 sorption (wt %) 1.8 1.5 1.2 0.9 1500 1800 2100 2400 2700 Specific surface area (m2/g) ) 3000

2.4 2.0 H2 sorption (wt %) 1.6 1.2 0.8 0.4 0.0 0.0
Vmi V<1 nm

uptake was plotted as a function of volume of pores below 1 nm. It clearly indicates the small pores are more effective and contribute greatly to the total hydrogen adsorption capacity. However, the contribution of pores a little larger than 1 nm to the adsorption of hydrogen should not be ignored, because an appropriate amount of bigger pores may provide fast transportation in the material and play an important role in hydrogen storage at a little higher pressure [34]. Therefore, a high H2 uptake can be expected in the nanostructured carbons with large surface area and pore volume as well as a narrow pore size distribution less than 1 nm. Moreover, we considered the hydrogen uptake as amounts per unit specic surface area or microporous volume, which was obtained by dividing the gravimetric capacity by the SBET or Vmi . The hydrogen adsorption per unit SBET increased evidently from 3.15 106 for CMK-1 to 4.50 106 for C900-2 and 4.05 106 mol/m2 for C950-6. Similarly, the hydrogen adsorption per unit Vmi increased from 1.06 102 for CMK-1 to 1.36 102 for C900-2 and 1.17 102 mol/cm3 for C9506. It indicates that CO2 activation can remarkably enhance the efciency of surface areas and micropores in porous carbon CMK-1 and thus increase the hydrogen storage capacity. It is noteworthy that the carbons in this work have demonstrated a higher H2 density in Vmi at 77 K and 1 bar than that of activated carbon AX-21 (1.17 102 mol/cm3 ) [3] and zeolite-like carbon (1.15 102 mol/cm3 ) [28], though it is still lower than that of TiC-CDC (1.59 102 mol/cm3 ) [31]. We ascribe the high hydrogen adsorption density in micropores to the relatively uniform micropore structure in the carbons obtained by using ordered porous carbon CMK-1 as a starting material. 4. Conclusions In summary, we have shown that a promising hydrogen storage material can be obtained by CO2 activation of ordered porous carbon CMK-1, which was synthesized by replication using MCM-48 as template. The activation was carried out considering both the effects of temperature and time. The structure and texture changes were investigated by using XRD and nitrogen sorption at 77 K. Unlike the wide pore size distribution in traditional activated carbons, the carbon materials prepared here have exhibited relatively uniform micropore texture. Under the optimum condition that treating with CO2 at 950 C for 6 h, a high H2 uptake of 2.19 wt% was obtained at 77 K and 1 bar. The capacity is almost twice that of CMK-1, and is higher than that of the reported porous carbons synthesized by silica templates. Moreover, the pores smaller than 1 nm were conrmed to be the most efcient for hydrogen sorption. The evaluations of hydrogen sorptions per unit SBET and Vmi showed the high efciencies of surface areas and micropores in the carbons for H2 uptake at 77 K and 1 bar. Acknowledgments This work was nancially supported by Chinese Academy of Sciences (Bairen Project and Creative Foundation) and Shanghai Nanotechnology Promotion Center (no. 0652nm025).

0.2

0.4 0.6 0.8 Microporous volume (cm3/g) )

1.0

Fig. 9. Hydrogen adsorption capacity versus (a) the specic surface area and (b) the microporous volume based on the N2 adsorption at 77 K.

H2 sorption. The hydrogen adsorption capacities of all porous carbons as prepared at 77 K and 1 bar were plotted versus the specic surface areas (Fig. 9(a)) and the micropore volumes (Fig. 9(b)), respectively. It is expectable that the materials with higher surface area certainly provided more accessible adsorbing sites resulting in higher adsorption capacities. As shown in Fig. 9(a), the hydrogen storage capacity increased along with the increase of the total specic surface area. It indicates the important role of SBET on H2 uptake at 77 K, although no linear relation between them was observed in our experiment. This result is not exactly obey the Chahine rule, e.g., the correlation between BET surface area and hydrogen uptake in carbons of 1 wt% H2 per 500 m2 /g of BET surface area [30], because the correlation holds when using the maximum surface excess adsorption value or saturation value typically obtained at higher pressures. Similar correlation could be found between Vmi and H2 uptake. This suggests that not all pores smaller than 2 nm are essentially effective for hydrogen adsorption at 77 K. It is noteworthy that a strict linear increase was found when H2

K. Xia et al. / International Journal of Hydrogen Energy 33 (2008) 116 123

123

References
[1] Schlapbach L, Zttel A. Hydrogen-storage materials for mobile applications. Nature 2001;414:3538. [2] Kajiura H, Tsutsui S, Kadono K, Kakuta M, Ata M, Murakami Y. Hydrogen storage capacity of commercially available carbon materials at room temperature. Appl Phys Lett 2003;82:11057. [3] Mandoki NT, Dentzer J, Piquero T, Saadallah S, David P, Guterl CV. Hydrogen storage in activated carbon materials: role of the nanoporous texture. Carbon 2004;42:27447. [4] Zttel A, Sudan P, Mauron P, Kiyobayashi T, Emmenegger C, Schlapbach L. Hydrogen storage in carbon nanostructures. Int J Hydrogen Energy 2002;27:20312. [5] Hirscher M, Becher M. Hydrogen storage in carbon nanotubes. J Nanosci Nanotech 2003;3:317. [6] Ansn A, Callejas MA, Benito AM, Maser WK, Izquierdo MT, Rubio B. et al. Hydrogen adsorption studies on single wall carbon nanotubes. Carbon 2004;42:12438. [7] Pan WY, Zhang XF, Li S, Wu DH, Mao ZQ. Measuring hydrogen storage capacity of carbon nanotubes by high-pressure microbalance. Int J Hydrogen Energy 2005;30:71922. [8] Tibbetts GG, Meisner GP, Olk CH. Hydrogen storage capacity of carbon nanotubes, laments, and vapor-grown bers. Carbon 2001;39: 2291301. [9] Browning DJ, Gerrard ML, Lakeman JB, Mellor IM, Mortimer RJ, Turpin MC. Studies into the storage of hydrogen in carbon nanobers: proposal of a possible reaction mechanism. Nano Lett 2002;2(3): 2015. [10] Lueking AD, Yang RT, Rodriguez NM, Baker RTK. Hydrogen storage in graphite nanobers: effect of synthesis catalyst and pretreatment conditions. Langmuir 2004;20(3):71421. [11] Gadiou R, Saadallah SE, Piquero T, David P, Parmentier J, Guterl CV. The inuence of textural properties on the adsorption of hydrogen on ordered nanostructured carbons. Microporous Mesoporous Mater 2005;79:1218. [12] Fang BZ, Zhou HS, Honma I. Ordered porous carbon with tailored pore size for electrochemical hydrogen storage application. J Phys Chem B 2006;110:487580. [13] Guterl CV, Frackowiak E, Jurewicz K, Friebe M, Parmentier J, Bguin F. Electrochemical energy storage in ordered porous carbon materials. Carbon 2005;43:1293302. [14] Yang ZX, Xia YD, Sun XZ, Mokaya R. Preparation and hydrogen storage properties of zeolite-templated carbon materials nanocast via chemical vapor deposition: effect of the zeolite template and nitrogen doping. J Phys Chem B 2006;110:1842431. [15] Ryoo R, Joo SH, Jun S. Synthesis of highly ordered carbon molecular sieves via template-mediated structural transformation. J Phys Chem B 1999;103:77436. [16] Joo SH, Jun S, Ryoo R. Synthesis of ordered mesoporous carbon molecular sieves CMK-1. Microporous Mesoporous Mater 2001;4445:1538.

[17] Jun S, Joo SH, Ryoo R, Kruk M, Jaronice M, Liu Z. et al. Synthesis of new, nanoporous carbon with hexagonally ordered mesostructure. J Am Chem Soc 2000;122:107123. [18] Smith MR, Bittner EW, Shi W, Johnson JK, Bockrath BC. Chemical activation of single-walled carbon nanotubes for hydrogen adsorption. J Phys Chem B 2003;107:375260. [19] Blackman JM, Patrick JW, Arenillas A, Shi W, Snape CE. Activation of carbon nanobres for hydrogen storage. Carbon 2006;44:137685. [20] Monnier A, Schth F, Huo Q, Kumar D, Margolese D, Maxwell RS. et al. Cooperative formation of inorganicorganic interfaces in the synthesis of silicate mesostructures. Science 1993;261:1299303. [21] Kim JM, Kim SK, Ryoo R. Synthesis of MCM-48 single crystals. Chem Commun 1998;259. [22] Schumacher K, Ravikovitch PI, Chesne AD, Neimark AV, Unger KK. Characterization of MCM-48 materials. Langmuir 2000;16:464854. [23] Mbileni CN, Prinsloo FF, Witcomb MJ, Coville NJ. Synthesis of mesoporous carbon supports via liquid impregnation of polystyrene onto a MCM-48 silica template. Carbon 2006;44:147683. [24] Morales IF, Almazn MCA, Mendoza MP, Garca MD, Garzn FJL. PET as precursor of microporous carbons: preparation and characterization. Microporous Mesoporous Mater 2005;80:10715. [25] Roussel T, Pellenq RJM, Bienfait M, Guterl CV, Gadiou R, Bguin F. et al. Thermodynamic and neutron scattering study of hydrogen adsorption in two mesoporous ordered carbons. Langmuir 2006;22: 46149. [26] Li ZJ, Yan WF, Dai S. Surface functionalization of ordered mesoporous carbonsa comparative study. Langmuir 2005;21:119992006. [27] Kabbour H, Baumann TF, Satcher JH, Saulnier A, Ahn CC. Toward new candidates for hydrogen storage: high-surface-area carbon aerogels. Chem Mater 2006;18:60857. [28] Yang ZX, Xia YD, Mokaya R. Enhanced hydrogen storage capacity of high surface area zeolite-like carbon materials. J Am Chem Soc 2007;129:16739. [29] Pang JB, Hampsey JE, Wu ZW, Hu QY, Lu YF. Hydrogen adsorption in mesoporous carbons. Appl Phys Lett 2004;85:48879. [30] Panella B, Hirscher M, Roth S. Hydrogen adsorption in different carbon nanostructures. Carbon 2005;43:220914. [31] Gogotsi Y, Dash RK, Yushin G, Yildirim T, Laudisio G, Fischer JE. Tailoring of nanoscale porosity in carbide-derived carbons for hydrogen storage. J Am Chem Soc 2005;127:160067. [32] Bguin F, Kierzek K, Friebe M, Jankowska A, Machnikowski J, Jurewicz K. et al. Effect of various porous nanotextures on the reversible electrochemical sorption of hydrogen in activated carbons. Electrochim Acta 2006;51:21617. [33] Mckeown NB, Gahnem B, Msayib KJ, Budd PM, Tattershall CE, Mahmood K. et al. Towards polymer-based hydrogen storage materials: engineering ultramicroporous cavities within polymers of intrinsic microporosity. Angew Chem Int Ed 2006;45:18047. [34] Yushin G, Dash RK, Jagiello J, Fischer JE, Gogotsi Y. Carbide-derived carbons: effect of pore size on hydrogen uptake and heat of adsorption. Adv Funct Mater 2006;16:228893.

Anda mungkin juga menyukai