• 沒有找到結果。

Chapter 4.Results and Discussion

4.2.6 Summary

As well known, it can be observed that the electrocatalytic activity of Pt nanoparticles on HNO3-MWNTs is higher than the others due to more amounts of dispersive Pt nanoparticles.

4.2.6 Summary

1. Various solvents, eg. HNO3, H2SO4, and KOH (800C, 6 hr), modified MWNTs can form the same functional group -COOH and –OH.

2. Well-dispersed functional groups on MWNTs could improve the efficiency of half-cell test.

3. The functional groups on MWNTs increase anchoring sites of Pt precursor to increase the efficiency of half-cell test.

4. The amount of functional groups in order to attract Pt on HNO3-MWNTs is more than on H2SO4-MWNTs and on KOH-MWNTs due to a strong oxidizing agent HNO3.

W/O KOH H2SO4 HNO3

-Fig. 4.18 Electrocatalytic activity is evaluated by the current peaks

W/O KOH H2SO4 HNO3

(1) Pt-CO+Pt-OH -> 2Pt+ CO

2+H++e

4.3. Multi-wall carbon nanotubes are modified by HNO

3

with Temperature (T), time (t), and concentration (conc.).

MWNTs are modified by HNO3, H2SO4, and KOH at high temperature in Section 4.2 and may produce the same functional groups, the alcoholic hydroxyl groups (-OH) and carboxylic acids (-COOH), to attract more Pt ions nucleating uniformly and densely.

Moreover, the amount of functional groups in order to attract Pt on HNO3-MWNTs is more than on H2SO4-MWNTs and on KOH-MWNTs due to a strong oxidizing agent HNO3. However, there are the other parameters, temperature, time, and concentration, in the function of chemical modification for MWNTs. In this work, we would find the best temperature (T) for HNO3-MWNTs may produce a lot of functional groups to increase Pt anchoring sites in Table. 4.4. Furthermore, for 2M HNO3, the highest temperature would be lower 1000C to maintain the concentration.

Table 4.4 HNO3-MWNTs with different T

Temperature (T) 800C 900C 1000C 2M HNO3

(b.p. ~1050C)

24 hr 24 hr 24 hr

14M HNO3

(b.p. ~1220C)

24 hr 24 hr 24 hr

4.3.1.1 Analysis of MWNTs morphology with T

Fig. 4.19 and Fig. 4.20 show that there is little damage for 14 M HNO3-MWNTs at 800C and 900C for 24 hr. On the contrary, Fig. 4.21 shows that there is a serious damage for 14 M HNO3-MWNTs at 1000C for 24 hr. The total loading amount of Pt would decrease if the amount of MWCNT as support were broken. Therefore, the temperature of chemical modification would be lower than 1000C.

4.3.1.2 Analysis of functional groups with T

We try to assay the functional groups with FTIR after MWNTs are modified by 2 M or by 14 M HNO3 with temperature. Fig. 4.22(a) (b) show that oxidization with HNO3 at 900C and 1000C successfully introducedcarboxylic acids (-COOH), carboxyl groups (-C=O), and alcoholic hydroxyl groups (-OH) on MWNTs surfaces. However, the functional groups of 2M and 14M HNO3-MWNTs at 800C would be difficult to be detected because of their weak response in the IR mode. It is suggested that temperature is a main factor if the oxidization of chemical reaction could proceed. Finally, 900C is the optimum temperature for 2M and 14M HNO3-MWNTs to form a lot of functional groups.

Fig. 4.19 Little damage for 14M HNO3-MWNTs at 800C

Fig. 4.20 Little damage for 14M HNO3-MWNTs at 900C

Fig. 4.21 A serious damage for 14M HNO3-MWNTs at 1000C

Fig. 4.22 (a) FTIR of 2M HNO3-MWNTs with different T

Fig. 4.22 (b) FTIR of 14M HNO3-MWNTs with different T

4000 3000 2000 1000

From Section 4.3.1, the modified MWNTs at 900C form a lot of functional groups for 2 M HNO3 and14 M HNO3. In the followed Section 4.3.2 and Section 4.3.3, we would find the other best parameters time (t) with high (14M) and low (2M) concentration (conc.) for HNO3-MWNTs may produce the most functional groups to increase Pt anchoring sites in Table. 4.5

Table 4.5 MWNTs are modified by HNO3 at 900C for 2M and 14M with t

900C Time (hr)

2M HNO3 00 06 12 18 24 48

14M HNO3 00 06 12 18 24 48

4.3.2.1 FTIR of 2M HNO

3

-MWNTs with t

The number of carboxylic acids (-COOH) would be expected by its intensity in the FTIR spectrum as MWNTs (0.01g) are diluted with a potassium bromide (KBr) dispersedly [32]. Fig. 4.23 shows that the number of carboxylic acids (-COOH) would increase with time gradually so more Pt could be anchored on MWNTs by more carboxylic acids (-COOH).

4.3.2.2 EDX analysis of Pt/2M HNO

3

-MWNTs

Pt loading on carbon nanotube would be increased because more functional groups, the carboxylic acids (-COOH), act nucleation sites and make active catalyst areas large.

The energy dispersive analysis in Fig. 4.24, Fig. 4.25, Fig. 4.26, Fig. 4.27 and Fig. 4.28 show that the amount of Pt loaded on MWNTs with reference to carbon can be evaluated qualitatively as 14.37 wt%, 16.72 wt%, 24.38 wt%, 28.21 wt%, and 26.99 wt% after the chemical modification of 2M HNO3 with time. As mentioned above, 2M HNO3-MWNTs from 12 hr to 24 hr may anchor the most amount of Pt in Fig. 4.29.

Therefore, it is called stable state for the most loading of Pt.

4.3.2.3 Half-cell test

The activity of catalyst is practically important in the study on fuel cell, which is usually evaluated by the current peaks with the dispersion and the amount of catalyst.

The electrocatalytic activity of Pt nanoparticles is obtained from the CV measurements performed in 1M methanol and 1M sulfuric acid electrolyte. Fig. 4.30 shows that the electrocatalytic activity of Pt nanoparticles is evaluated gradually by the current peaks of methanol oxidation. As a whole, it can be observed that the highest electrocatalytic activity of Pt nanoparticles for 2M HNO3-MWNTs range from 12 hr to 24 hr (stable state) due to the most amounts of dispersive Pt nanoparticles. However, the electrocatalytic activity of Pt nanoparticles for 2M HNO3-MWNTs decreases much significantly from 24 hr to 48 hr because MWNTs may be destroyed very much at 48 hr.

4.3.2.4 Effective activating area

The electrochemically active surface areas of the Pt nanoparticles are obtained from the CV measurements performed in 1M sulfuric acid electrolyte. A typical

reference platinum electrode. The hydrogen absorption and desorption peaks are clearly seen in the voltammogram at potentials between -0.1 and 0.2 V, which is consistent with those observed for platinized Pt. In detail, a redox peak at 0.5-0.7 V in all voltammograms can be attributed to the quinine and hydroquinone groups. Furthermore, there is a couple of redox peak at about 0.4 V, which may be associated with the C-O and C=O groups [41].

The electrochemically active surface area, Sact, in units of cm2/mg Pt, is calculated from the CV curves for Pt [42] by eq. (5)

Sact 210QH

= (5)

In which QH, in units of (mA/mg Pt) V, is the integrated area of the hydrogen adsorption region in the voltammogram and the charge for monolayer hydrogen adsorption on Pt equal to 210 μC/ cm2. To obtain the integrated area for the hydrogen adsorption peaks in Fig. 4.31, a horizontal line is drawn to correct the double-layer charging, and a vertical line is drawn to separate the molecular hydrogen region [42].

Thus, in Table 4.6, the electrochemically active surface areas of the Pt nanoparticles are evaluated gradually with the time of 2M HNO3-MWNTs due to the trend of the electrocatalytic activity of Pt nanoparticles.

Fig. 4.23 The number of -COOH would increase with t gradually

Fig. 4.24 The amount of Pt loading on raw MWNTs is 14.37 wt%

4000 3000 2000 1000

Fig. 4.25 The amount of Pt loading on 2M HNO3-MWNTs at 6 hr is 16.72 wt%

Fig. 4.28 The amount of Pt loading on 2M HNO3-MWNTs at 24 hr is 26.99 wt%

Fig. 4.29 2M HNO3-MWNTs from 12 hr to 24 hr may anchor the most Pt 100.00

Totals

2.23 26.99

Pt

97.77 73.01

C

Atomic%

Weight%

Element

0 5 10 15 20 25 0

10 20 30

40 Pt EDX

Pt weight %

Time (hr)

Stable state

Fig. 4.30 The currents peaks represent the activity of Pt with t

Fig. 4.31 Electrochemically active surface areas of Pt with t

Table 4.6 The value of electrochemically active surface areas of the Pt with the t Time of 2M HNO3 QH (mC) S pt (cm2)

00hr 0.28363 1.35064

06hr 0.43399 2.0666

12hr 0.44241 2.10672

18hr 0.52727 2.51081

24hr 0.6042 2.877

-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 -0.01

0.00 0.01

Current density (A/cm 2 )

Potential (V)

no treatment 2M HNO

3 06hr 2M HNO

3 12hr 2M HNO

3 18hr 2M HNO

3 24hr

4.3.3.1 FTIR of 14M HNO

3

-MWNTs with t

The number of carboxylic acids (-COOH) would be expected by its intensity in the FTIR spectrum as MWNTs (0.01g) are diluted with a potassium bromide (KBr) dispersedly [32]. Fig. 4.32 shows that the number of carboxylic acids (-COOH) for 14M HNO3-MWNTs would increase with time more fast than for 2M HNO3-MWNTs so more Pt nanoparticles could be anchored on MWNTs in the short time.

4.3.3.2 EDX analysis of Pt/14M HNO

3

-MWNTs

Pt loading on carbon nanotube would be increased because more functional groups, the carboxylic acids (-COOH), act nucleation sites and make active catalyst areas large.

The energy dispersive analysis in Fig. 4.33, Fig. 4.34, Fig. 4.35, Fig. 4.36 and Fig. 4.37 show that the amount of Pt loaded on MWNTs with reference to carbon can be evaluated qualitatively as 15.38 wt%, 26.52 wt%, 27.94 wt%, 25.79 wt%, and 25.95 wt% after the chemical modification of 14M HNO3 with time. As mentioned above, 14M HNO3-MWNTs from 6 hr to 24 hr may anchor the most amount of Pt in Fig. 4.38.

Thus, it is called stable state for the most loading of Pt.

4.3.3.3 Half-cell test

The activity of catalyst is practically important in the study on fuel cell, which is usually evaluated by the current peaks with the dispersion and the amount of catalyst.

The electrocatalytic activity of Pt nanoparticles is obtained from the CV measurements performed in 1M methanol and 1M sulfuric acid electrolyte. Fig. 4.39 shows that the

of methanol oxidation. As a whole, it can be observed that the highest electrocatalytic activity of Pt nanoparticles for 14M HNO3-MWNTs range from 6 hr to 24 hr (stable state) due to the most amounts of dispersive Pt nanoparticles. However, like the decay of the electrocatalytic activity for 2M HNO3-MWNTs, the electrocatalytic activity of Pt nanoparticles for 14M HNO3-MWNTs decreases much significantly from 24 hr to 48 hr because MWNTs may be destroyed very much at 48 hr.

4.3.3.4 Effective activating area

The electrochemically active surface areas of the Pt nanoparticles are obtained from the CV measurements performed in 1M sulfuric acid electrolyte. A typical voltammogram is shown in Fig. 4.40, in which the potential is expressed versus that of a reference platinum electrode. It is the same as 2M HNO3-MWNTs to obtain the integrated area for the hydrogen adsorption peaks in Fig. 4.40. Thus, a horizontal line is drawn to correct the double-layer charging, and a vertical line is drawn to separate the molecular hydrogen region [42]. Moreover, in Table 4.7, the electrochemically active surface areas of the Pt nanoparticles for 14M HNO3-MWNTs from 6 hr to 24 hr are in the stable state due to the trend of the electrocatalytic activity of Pt nanoparticles.

4.3.4 Summary

1. Temperature is one of the main factors if MWNTs modified by HNO3 form functional groups -COOH and –OH.

2. 2M and 14M HNO3-MWNTs at 800C are unfavorable to form functional groups.

3. The modified MWNTs at 1000C could oxidize and destroy the surface of MWNTs very much.

4. 900C is the optimum temperature for 2M and 14M HNO3-MWNTs to form functional groups.

5. 2M HNO3-MWNTs form a lot of functional groups (stable state) from 12hr to 24hr.

6. 14M HNO3-MWNTs form a lot of functional groups (stable state) from 6hr to 24hr.

2. Higher concentration of HNO3 allows the surface of MWNTs to form functional groups quickly.

Fig. 4.32 The number of -COOH would increase with t gradually

Fig. 4.33 The amount of Pt loading on raw MWNTs is 15.38 wt%

4000 3000 2000 1000

Fig. 4.34 The amount of Pt loading on 14M HNO3-MWNTs at 6 hr is 26.52 wt%

Fig. 4.37 The amount of Pt loading on 14M HNO3-MWNTs at 24 hr is 25.95 wt%

Fig. 4.38 14M HNO3-MWNTs from 6 hr to 24 hr may anchor the most Pt 100.00

Totals

2.11 25.95

Pt

97.89 74.05

C

Atomic%

Weight%

Element

0 5 10 15 20 25

0 10 20 30

40 Pt EDX

Time (hr)

Pt weight %

Stable state

Fig. 4.39 The currents peaks represent the activity of Pt with t

Fig. 4.40 Electrochemically active surface areas of Pt with t

Table 4.7 The value of electrochemically active surface areas of the Pt with the t Time of 14M HNO3 QH (mC) Spt(cm2)

00hr 0.1304 0.621

06hr 0.42083 2.004

12hr 0.5032 2.3962

18hr 0.48 2.286

24hr 0.49685 2.366

-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 -0.005

0.000 0.005 0.010

Current density (A/cm

2

)

Potential (V)

no treatment 14M HNO3 06hr 14M HNO3 12hr 14M HNO3 18hr 14M HNO3 24hr

4.4 Analysis of 14M HNO

3

-MWNTs

This chemical modification of MWNTs enabled us to obtain a high dispersion of Pt nanoparticles, even when a large metal load is required on the supports with small surface areas, as in the case of electrocatalysts for fuel cells. Hence, there are two positive factors to rise the electrocatalytic activity and the effective activating surface areas very much. However, there is no significant rise for Pt nanoparticles on MWNTs modified by HNO3. For the reason, there may be some negative factors to reduce its electrocatalytic activity. In the other hand, we try to explain the stable state with analyzing 14M HNO3-MWNTs. Moreover, TEM images as evidences to support the hypothetical model.

4.4.1 Analysis of mean Pt nanoparticle size

The size of Pt nanoparticle is also a factor for electrocatalytic activity in the same weight. Furthermore, the smaller Pt nanoparticles in the same weight represent the more effective activating surface areas. Then, the mean Pt nanoparticle size is analyzed by XRD and TEM.

The Pt nanoparticles on raw-MWNTs and on 14M HNO3-MWNTs are shown in Fig. 4.41. Both Pt on raw-MWNTs and on 14M HNO3-MWNTs displays the characteristic patterns of Pt fcc diffraction. The 2θ values of the (111) peak is about 39.80 and the Pt (200) diffraction is about 46.40. The broader diffraction peaks for the catalysts also led to smaller average metal particle size as calculated by the Scherrer equation (1) [43].

wavelength (1.54056 Ǻ for Cu Kα1 radiation), B is the peak broadening, and θB is the angle corresponding to the peak maximum. The calculation results are estimated the average size of about 5.3226 nm for Pt/raw-MWNTs and about 6.76 nm for Pt/14M HNO3-MWNTs in Table. 4.8.

In the other approach, the mean size of Pt nanoparticles on raw-MWNTs and 14M HNO3-MWNTs are analyzed by TEM. The particle size distributions of the metal on MWNTs are obtained by directly measuring the sized of 60 randomly chosen particles in the magnified TEM images. Fig.4.42 and Fig.4.43 show that the average diameters of 5.8802 nm for Pt/raw-MWNTs and 7.29 nm for Pt/14M HNO3-MWNTs are accompanied by relatively narrow particle size distributions (4~10 nm). The estimated average size of Pt nanoparticles is in good agreement with the XRD measurements.

Finally, from XRD and TEM measurements, it is observed that, for 14M HNO3-MWNTs, the diameter of the metal particles deposited is larger than that of raw-MWNTs. By the way, the particle size of Pt may be correlated with the oxidation of MWNTs, which indicates that the efficient deposition of Pt nanoparticles is due to a strong interaction between the metal salt precursor and the functional groups of the MWNTs. Chemical functional groups, namely –COOH and –OH derived from chemical oxidation processes, act as anchoring sites for metal nanoparticles. These –COOH sites may induce the impregnation of larger particles which is a negative factor to rise the electrocatalytic activity and the effective activating surface areas. Hence, there is no significant rise for Pt nanoparticles on MWNTs modified by HNO3.

4.4.2 Analysis of MWNTs morphology

Fig. 4.44 (a) shows that raw MWNTs display hollow tubes with amorphous and crystalline layers. It is known that 14M HNO3-MWNTs not only opens the closed tips of the tubes, but creates functional groups, -COOH, on the surface to attract metal ions to nucleate. Fig. 4.44 (b) shows that the amorphous surface layers and the cap of MWNTs are removed first with 14M HNO3 due to its non-crystalline structure. The open cap of MWNTs may allow Pt to get into the inner surface of tubes by the other method in Fig. 4.45. Then, even crystalline layer may be broken for a longer time. In other words, the functional groups on the surface layer of 14M HNO3-MWNTs may be removed for a longer time. However, the total number of functional groups may keep a balance on 14M HNO3-MWNTs after a certain time. In this work, we take a hypothetical model for stable state in Fig. 4.46 to explain it with TEM image. Although functional groups on the surface layer of 14M HNO3-MWNTs may be removed, new functional groups would be formed on the new surface layer which is inner the removed surface layer. Therefore, the total number of functional groups may keep a balance after some time.

4.4.3 Summary

1. The amount of functional groups is in the equilibrium during the stable state.

2. COO-H+ attract more Pt precursors to nucleate Pt particle so it produce the larger particle against its efficiency.

Fig. 4.41 XRD patterns of Pt on raw-MWNTs and 14M HNO3-MWNTs

Table 4.8 The average size of Pt/raw-MWNTs and Pt/14M HNO3-MWNTs

time K λ Kα1 B θB size (Ǻ)

0 29.74 1.54 1.08 39.8 53.226

6 29.74 1.54 0.88 39.8 67.74

12 29.74 1.54 0.884 39.7 67.34

18 29.74 1.54 0.8832 39.93 67.6

24 29.74 1.54 0.882 39.7 67.5

20 30 40 50 60

2 theta degree

Intesit y (a.u.)

C(101)

C(002)

Pt(200) Pt(111)

no treatment

900C 14M HNO3 06hr 900C 14M HNO3 12hr 900C 14M HNO3 18hr 900C 14M HNO3 24hr

Fig. 4.42 The average diameters for Pt/raw-MWNTs is 5.8802 nm

Fig. 4.43 The average diameters for Pt/14M HNO3-MWNTs is 7.29 nm

Fig. 4.44 (a) raw MWNTs (b) 14M HNO3-MWNTs

Fig. 4.45 (a) raw MWNTs with Pt (b) 14M HNO3-MWNTs with Pt

Fig. 4.46 Hypothetical model for stable state of 14M HNO3-MWNTs

8.09 nm

(a) (b)

Chapter 5 Conclusions

1. Various solvents, eg. HNO3, H2SO4, and KOH (800C, 6 hr), modified MWNTs can form the same functional group -COOH and –OH.

2. Well-dispersed functional groups on MWNTs could improve the efficiency of half-cell test.

3. The functional groups on MWNTs increase anchoring sites of Pt precursors to increase the efficiency of half-cell test.

4. The amount of functional groups in order to attract Pt on HNO3-MWNTs is more than on H2SO4-MWNTs and on KOH-MWNTs due to a strong oxidizing agent HNO3.

5. Temperature is one of the main factors if MWNTs modified by HNO3 form functional groups -COOH and –OH.

6. 2M and 14M HNO3-MWNTs at 800C are unfavorable to form functional groups.

7. The modified MWNTs at 1000C could oxidize and destroy the surface of MWNTs very much.

8. 900C is the optimum temperature for 2M and 14M HNO3-MWNTs to form functional groups.

9. 2M HNO3-MWNTs form a lot of functional groups (stable state) from 12hr to 24hr.

10. 14M HNO3-MWNTs form a lot of functional groups (stable state) from 6hr to 24hr.

11. Higher concentration of HNO3 allows the surface of MWNTs to form functional groups quickly.

12. The amount of functional groups is in the equilibrium during the stable state.

13. COO-H+ attract more Pt precursorsto nucleate Pt particle so it produce the larger particle against its efficiency.

Reference:

1. M.S. Wilson, Membrane catalyst layer for fuel cells, U.S. Pat. No.5,234,777 (1993).

2. Y.-G. Chun, C.-S. Kim, D.-H. Peck, D.-R. Shin, J. Power Sources 71 (1998) 174.

3. S.J. Lee, S. Mukerjee, J. McBreen, Y.W. Rho, Y.T. Kho, T.H. Lee, Electrochim. Acta 43 (1998) 3693.

4. O.J. Murphy, G.D. Hitchens, D.J. Manko, J. Power Sources 47 (1994) 353.

5. E.A. Ticianelli, C.R. Derouin, A. Redondo, S. Srinivasan, J. Electrochem. Soc. 135 (1988) 2209.

6. X. Cheng, B. Yi, M. Han, J. Zhang, Y. Qiao, J. Yu, J. Power Sources 79 (1999) 75.

7.M.S. Wilson, S. Gottesfeld, J. Appl. Electrochem. 22 (1992) 1.

8. M. Uchida, Y. Fukuoka, Y. Sugawara, N. Eda, A. Ohta, J. Electrochem. Soc. 143 (1996) 2245.

9. S.H. Joo, S.J. Choi, I. Oh, J. Kwak, Z. Liu, O. Terasaki, R. Ryoo, Nature 412 (2001) 169.

10. C.A. Bessel, K. Laubernds, N.M. Rodriguez, R. Terry, K. Baker, J. Phys. Chem. B 105 (2001) 1115.

11. N.M. Rodriguez, M.-S. Kim, R. Terry, K. Baker, J. Phys. Chem. 98 (1994) 13108.

12. Y.-C. Liu, X.-P. Qiu, Y.-Q. Huang, W.-T. Zhu, J. Power Sources 111 (2002) 160.

13. B.C. Satishkumar, E.M. Vogl, A. Govindaraj, C.N.R. Rao, J. Phys. D 29 (1996) 3173.

14. B.C. Satishkumar, A. Govindaraj, J. Mofokeng, G.N. Subbanna, C.N.R. Rao, J. Phys. B 29 (1996) 4925.

15. W. Li, C. Liang, J. Qiu, W. Zhou, H. Han, Z. Wei, G. Sun, Q. Xin, Carbon 40 (2002) 791.

16. G. Che, B.B. Lakshmi, C.R. Martin, E.R. Fisher, Langmuir 15 (1999) 750.

17. V. Lordi, N. Yao, J. Wei, Chem. Mater. 13 (2001) 733.

18. G. Che, B.B. Lakshmi, C.R. Martin, E.R. Fisher, Langmuir 15 (1999) 346.

19. G. Girishkumar, K. Vinodgopal, P.V. Kamat, J. Phys. Chem. B 108 (2004) 19960.

21.Elzbieta Frackowiak, Grzegorz Lota, Thomas Cacciaguerra, Francois Beguin, Electrochem.

Commun. 8 (2002) 129.

22. B. Rajesh, K.R. Thampi, J.M. Bonard, N. Xanthopoulos, H.J. Mathieu, B. Viswanathan, J. Phys. Chem. B 107 (2003) 2701.

23. Z. He, J. Chen, D. Liu, H. Tang, W. Deng, Y. Kuang, Mater. Chem. Phys. 85 (2004) 396.

24. Z. He, J. Chen, D. Liu, H. Zhou, Y. Kuang, Diam. Relat. Mater. 13 (2004) 1764.

25. D.J. Guo, H.L. Li, J. Electroanal. Chem. 573 (2004) 197.

26. Z. Liu, X. Lin, J.Y. Lee, W. Zhang, M. Han, L.M. Gan, Langmuir 18 (2002) 4054.

27. B. Rajesh, V. Karthik, S. Karthikeyan, K.R. Thampi, J.M. Bonard, B. Viswanathan, Fuel 81 (2002) 2177.

28. C. Wang, M. Waje, X. Wang, J.M. Tang, R.C. Haddon, Y. Yan, Nano Lett. 4 (2004) 345.

29. B. Xue, P. Chen, Q. Hong, J.Y. Lin, K.L. Tan, J. Mater. Chem. 11 (2001) 2378.

30. R.Q. Yu, L.W. Chen, Q.P. Liu, J.Y. Lin, K.L. Tan, S.C. Ng, H.S.O. Chan, G.Q. Xu, T.S.A.

Hor, Chem. Mater. 10 (1998) 718.

31. T. Freelink, W. Visscher, and J.A.R.van Veen, Surf. Sci. 335 (1995) 353.

32. Ki Chul Park, Takuya Hayashi, Hiroshi Tomiyasu, Morinobu Enbo, and Mildred S.

Dresselhaus, J. Mater. Chem. 15 (2005) 407.

33. 汪建民,材料分析,中國材料學會,台灣新竹 (1998).

34. L.Q. Jiang, L. Gao, Carbon 41 (2003) 2923.

35. B. Rohland, Marlies Pietrzak, S. Moller, Mihaela-C. Bunescu, Marion Wienecke and T. Barfels, Fullerenes, Nanotubes, and Carbon Nanostructures 13 (2005) 511.

36. H.J. Wang, H.Yu, F. Peng, P. Lv, Electrochem. Commun. 8 (2006) 499.

37. Yong-Tae Kim, Tadaoki Mitani, J. Catalysis 283 (2006) 394.

38. T.I.T. Okpalugo, P.Papakonstantinou, H.Murphy, J.McLaughlin, N.M.D. Brown and T.McNally, Fullerenes, Nanotubes, and Carbon Nanostructures 13 (2005) 477.

39. T.Teranishi, M. Hosoe, T. Tanaka, and M.Miyake, J. Phys. Chem. B 103 (1999) 3818.

40. 衣寶廉,燃料電池-原理與應用,五南圖書出版公司 台灣 (2004).

41. Min Chen and Yangchuan Xing, Langmuir 21 (2005) 9334.

42. Woods, R. Electroanal. Chem. 9 (1976) 1.

43. B.E. Warren, X-ray Diffraction, Addison-Wesley, Reading, MA, 1996.

個人簡歷

姓名:莊方慈 性別:男

出生日期:民國 71 年 2 月 16 日 籍貫:台北市

住址:台北市北投區知行路 91 巷 5 號 3 樓

學歷:國立清華大學材料科學與工程學系學士 (89.9-93.6)

學歷:國立清華大學材料科學與工程學系學士 (89.9-93.6)

相關文件