You are on page 1of 5

Available online at www.sciencedirect.

com

Separation and Purification Technology 58 (2007) 219–223

Hydrogen storage in platelet graphite nanofibers


Chao-Wei Huang, Hung-Chih Wu, Yuan-Yao Li ∗
Department of Chemical Engineering, National Chung Cheng University, Chia-Yi 62102, Taiwan, ROC

Abstract
Platelet graphite nanofibers (PGNFs) were synthesized by thermal decomposition from a mixture containing polymer and catalyst. The hydrogen
storage capacity of the material was measured using the volumetric method carried out at 298 K at a pressure up to 4.83 MPa. The results show
that the PGNF possess a 3.3 wt.% hydrogen storage capacity. The gaseous molecules were adsorbed between two graphene planes in the PGNFs
and residual amounts of hydrogen were possession within PGNFs while pressure decreased to ambient conditions. X-ray diffraction was used to
understand the PGNF structural modifications before and after hydrogen uptake. We found that the PGNF interlayer undergoes expansion and
returns back to the initial value before and after hydrogen uptake.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Platelet graphite nanofibers; Hydrogen storage; X-ray diffraction

1. Introduction which graphene layers are parallel to the fiber axis [also called
multi-walled carbon nanotubes (MWCNT)] [8]; (d) cup-stacked
Since the discovery of fullerene [1] and carbon nanotubes graphite nanofibers in which the graphene layers tilt and truncate
(CNTs) [2], these materials have been intensively studied conically along the fiber length [27]. A variety of GNFs, in terms
because of their small size and extraordinary properties. Numer- of diameter, length, and graphene layer arrangement can be
ous carbon nano-structures have been synthesized by changing controlled by synthesis conditions using various hydrocarbons,
the preparation conditions, including carbon nano-onions [3], metal catalysts [8,28,29], and temperature [30,31].
carbon nanocapsules [4,5], carbon discs [6], carbon nanorods Because ever-stricter environmental regulations put great
[7], graphite nanofiber (GNF) [8], etc. These materials have pressure on energy companies and automobile manufacturers
been intensively studied and used in advanced devices such as to develop cleaner burning fuels and cleaner running vehicles,
single electron transistors [9], quantum wires [10], flat panel dis- various environmentally friendly fuels and related technologies
plays [11–13], nano-forceps [14], scanning probe microscopes are being developed to replace fossil fuels. For example, numer-
[15–17] and energy storage (hydrogen, methane) [18–20]. The ous studies have been carried out on hydrogen applications as
fabrication strategies are based on arc-discharge [2,4], laser abla- clean energy fuels. The development of hydrogen storage mate-
tion [21], chemical vapor deposition (CVD) [8,22], flame [5], rials capable of meeting the U.S. Department of Energy (DOE)
templated synthesis [23], and high-pressure experiments [24]. storage target of 6.5 wt.% could reduce the dependence on fossil
GNF is a novel form of filamentous graphite obtained and fuels. The GNFs is one of the candidates as a hydrogen storage
investigated by Murayama and Maeda in 1990 [25]. In fact, material because of its unique structure, which exhibits virtu-
it was first reported in 1973 [26], but the microstructure was ally open edges and an interlayer spacing between the graphene
not thoroughly studied due to the limitations of the analyti- platelet (≥3.35 Å), acting like a slit-shaped pore hence possess-
cal instruments at that time. GNF is classified into four types: ing the ideal configuration for use in hydrogen storage, whose
(a) platelet graphite nanofibers (PGNF) in which graphene lay- kinetic diameter is only 2.89 Å [18]. Gupta et al. reported [32] the
ers are perpendicular to the fiber axis; (b) herringbone graphite lateral expansion of graphene layers and increasing of the inter-
nanofibers (HBGNF) in which graphene layers tilt at an angle layer spacing during hydrogen storage process. Theoretically,
to the fiber axis; (c) tubular graphite nanofibers (TGNF) in hydrogen can be adsorbed onto the surface and then incorpo-
rated/intercalated between GNF graphene layers or hydrogen
dissociation onto graphite edge sites [33]. Browing et al. [19]
∗ Corresponding author. Tel.: +886 5 2720411x33403; fax: +886 5 2721206. have also noted that the graphite edge sites in the GNF struc-
E-mail address: chmyyl@ccu.edu.tw (Y.-Y. Li). ture may play a role in the dissociation of hydrogen leading to

1383-5866/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.seppur.2007.07.032
220 C.-W. Huang et al. / Separation and Purification Technology 58 (2007) 219–223

Fig. 1. Schematic diagram of the P–C–T isothermal system: an automatic gas reaction controller.

subsequent interacalation. This involves the initial dissociation gen adsorption was performed using an ASAP 2020 system at
of hydrogen, believed to be catalyzed by carbon edge sites, which 77 K. The sample was degassed in a vacuum at 473 K for 4 h
constitute the majority of the nanofiber surface, a property which prior to the measurements. Surface area was calculated using
is probably an important contributory factor toward their high the Brunauer–Emmett–Teller (BET) equation [45]. The PGNF
hydrogen storage capacities. structural modifications were measured using XRD (Shimadzu,
The hydrogen storage capacity of GNFs can be measured XRD-6000) used Cu K␣ radiation from a rotating anode X-ray
by volumetric [18,19,32,34–39], gravimetric (TG) [40,41], and source.
thermal desorption spectroscopy (TDS) [42,43]. Chambers et The hydrogen storage capacity was measured using a pressure
al. [18] reported that GNFs are capable of sorbing and retain- composition temperature (P–C–T) isothermal system manufac-
ing in excess of 20 L (STP) of hydrogen per gram (∼67 wt.%) tured by the Advanced Materials Corporation. Fig. 1 shows a
when the GNFs are exposed to the gas at pressures of 120 atm simplified sketch of this apparatus. During a typical experiment,
at 293 K. Park et al. [34] reported that adequate GNF pretreat- approximately 100 mg of PGNFs were initially loaded into the
ment to remove chemisorbed gases and allow hydrogen access sample chamber while heating under vacuum at 423 K for 4 h,
to the GNF pores is a critical step in obtaining high GNF storage removing any physisorbed water. Following this the system was
capacity (up to 40 wt.%). These two reports remain contested, as allowed to cool to room temperature, after which hydrogen was
several other laboratories have not been able to reproduce these admitted and uptake determined. Initial testing included a blank
results [35–37,40]. Nonetheless, in recent years, the hydrogen experiment, without sample, which showed that the stainless
storage capacity reported in the literature for GNFs continues to steel vessel did not give rise to hydrogen adsorption, while also
emerge showing less than 1 wt.% up to 15 wt.% [19,32,39,41]. ensuring that the instrument was leak free. Hydrogen was then
To date, the reported hydrogen storage capacities have been dif- introduced into the reservoir container and subsequently allowed
ferent with a lack of reproducibility. Therefore, further studies access to the sample chamber for interaction with the nanofibers.
are needed to better understand the factors that influence the The system pressure was increased from an initial value up to
hydrogen storage capacity (as well as the measurements) and 4.83 MPa at 298 K and measured at regular intervals. The PGNF
mechanisms for hydrogen storage within GNFs. hydrogen storage capacity was calculated from the changes in
In this study, we have attempted to realize the hydrogen pressure following material interaction with the gas.
storage capacity of PGNFs synthesized using thermal decompo-
sition of poly(ethylene glycol) (PEG) containing catalyst under
3. Results and discussion
a nitrogen atmosphere [44] and to understand the structural
modifications in the PGNFs before and after hydrogen uptake.
3.1. Structural characteristics of the PGNFs
The PGNFs were characterized using high-resolution transmis-
sion electron microscopy (HR-TEM), field-emission scanning
Fig. 2(a) and (b) are SEM and HR-TEM images of synthe-
electron microscopy (FE-SEM), X-ray diffraction (XRD) and
sized PGNFs, respectively. It is clearly observed that the PGNFs
hydrogen uptake measurements using the volumetric method.
exhibited a diameter of about 20–40 nm and a few micrometers
in length. The as-produced GNF has a platelet structure with a
2. Experiment high degree of graphitization and well-ordered graphene layers
perpendicular to the fiber axis. It also shows large numbers of
PGNFs were prepared by thermal decomposition from a mix- nanopores, in which only edge sites are exposed, on the surface
ture containing PEG and nickel chloride (NiCl2 ) at 1023 K of the PGNFs. These nanopores can act like slit-shaped pores,
under a nitrogen atmosphere. The PGNF preparation process hence possessing the ideal arrangement for hydrogen storage.
details were described in our previous work [44]. The samples The selected area electron diffraction (SAED) pattern inserted
were observed using FE-SEM (HITACHI S-4800) for mor- in Fig. 2(b) also shows PGNFs with high degree of graphitiza-
phology analysis while HR-TEM (PHILIPS Tecnai F20) was tion. The SAED pattern can be used to determine the spacing of
employed to understand the structure and crystal forms. Nitro- the graphene layers by measuring the distance from the center
C.-W. Huang et al. / Separation and Purification Technology 58 (2007) 219–223 221

Fig. 3. P–C–T curves of the PGNF hydrogen storage capacity.

an atomic layer. However, surface areas accessible to nitrogen


(kinetic diameter is 3.64 Å) may not necessarily be accessible
to the hydrogen molecule (kinetic diameter is 2.89 Å). Thus
hydrogen uptake does not always correlate to BET surface area,
particularly for nanocarbons [19,41,47]. Accordingly, it can be
thought that the difference between 3.3 and 0.28 wt.% might
be evidence of hydrogen intercalation within the graphene lay-
ers. Dresselhaus et al. reported that the intercalant must initially
find exposed graphene edges. These sites then provide pathways
for more intercalant to move inward, filling the space between
the basal planes [33]. Consequently, the possibility of hydrogen
atom intercalation into the graphene layers might be considered
and increase the hydrogen storage capacity.

3.3. XRD analysis

Because the amount of hydrogen desorption is less than


that which is adsorpted, it has been speculated the graphene
Fig. 2. (a) SEM and (b) HR-TEM images of PGNFs. interlayer was expanded during the adsorption process and an
unexpectedly high hydrogen retention capacity is found within
spot to the diffracted spot. The d-spacing of PGNFs was about GNFs [32–34]. To determine whether graphene layer expansion
3.362 Å, being close to that of graphite. The BET surface area occurs after the hydrogen uptake process, we used the XRD tech-
of PGNFs was 123 m2 /g. nique to understand the structural modifications in PGNFs. Fig. 4

3.2. Hydrogen storage

PGNFs hydrogen storage was carried out using the volumet-


ric method at 298 K with the pressure up to 4.83 MPa. Fig. 3
shows the P–C–T isothermal curves of the PGNFs hydrogen
storage capacity. The result shows that the PGNFs possess a
3.3 wt.% hydrogen storage capacity. It was thought that because
of the weak (van der Waals) platelet bonding, the large number
of non-rigid wall nanopores can expand to accommodate hydro-
gen in a multilayer configuration [18]. Zuttel et al. reported that
the maximum amount of potential hydrogen adsorbed accord-
ing to the model was calculated to be 2.28 × 10−3 wt.% uptake
for each m2 of surface area for the adsorption of a mono-
layer of hydrogen at the surface [46]. This analysis requires
a flat graphene sheet and ignores the microporosity contribu-
tion. Therefore, hydrogen storage capacity only can be achieved Fig. 4. XRD patterns of (a) as-produced, (b) immediately after hydrogen uptake
about 0.28 wt.% by covering the outer PGNFs surface with and (c) a week after PGNF hydrogen uptake.
222 C.-W. Huang et al. / Separation and Purification Technology 58 (2007) 219–223

shows the XRD results of (a) as-produced sample, (b) imme- [11] D.W.A. Heer, A. Chatelain, D. Ugarte, A carbon nanotube field-emission
diately after hydrogen uptake, and (c) a week after hydrogen electron source, Science 270 (1995) 1179–1180.
uptake. It was clearly found that hydrogen induces an expansion [12] S. Fan, M.G. Chapline, N.R. Franklin, T.W. Tombler, A.M. Cassell, H. Dai,
Self oriented regular arrays of carbon nanotubes and their field emission
of the graphene layer d-spacing from 3.362 Å (prior to adsorp- properties, Science 283 (1999) 512–514.
tion and consistent with the result calculated using SAED) to [13] N.S. Lee, D.S. Chung, I.T. Han, J.H. Kang, Y.S. Choi, H.Y. Kim, S.H. Park,
3.398 Å (immediately after hydrogen uptake) and then decrease Y.W. Jin, W.K. Yi, M.J. Yun, J.E. Jung, C.J. Lee, J.H. You, S.H. Jo, C.G.
to 3.38 Å (a week after hydrogen uptake under ambient condi- Lee, J.M. Kim, Application of carbon nanotubes to field emission displays,
tions). The hydrogen uptake leads to graphene layer expansion Diam. Relat. Mater. 10 (2001) 265–270.
[14] P. Kim, C.M. Lieber, Nanotube nanotweezers, Science 286 (1999)
and an increase in the d-spacing of the structure. The PGNF 2148–2150.
d-spacing seems to undergo an expansion and return to the ini- [15] H. Dai, J.H. Hafner, A.G. Rinzler, D.T. Colbert, R.E. Smalley, Nan-
tial value during and after hydrogen uptake. This may indicate otubes as nanoprobes in scanning probe microscopy, Nature 384 (1996)
that some hydrogen still remains within the structure. Park et al. 147–150.
[34] reported these changes can be rationalized according to the [16] C.L. Cheung, J.H. Hafner, T.W. Odom, K. Kim, C.M. Lieber, Growth
and fabrication with single-walled carbon nanotube probe microscopy tips,
notion that as hydrogen enters the region between the graphene Appl. Phys. Lett. 76 (2000) 3136–3138.
layers, a concomitant expansion of the interlayer takes place to [17] C.L. Cheung, J.H. Hafner, C.M. Lieber, Carbon nanotube atomic force
accommodate the gas molecules, being slowly released over a microscopy tips: direct growth by chemical vapor deposition and appli-
period of time. cation to high-resolution imaging, Proc. Natl. Acad. Sci. USA 97 (2000)
3809–3813.
[18] A. Chambers, C. Park, R.T.K. Baker, N.M. Rodriguez, Hydrogen storage
4. Conclusions in graphite nanofibers, J. Phys. Chem. B 102 (1998) 4253–4256.
[19] D.J. Browning, M.L. Gerrard, J.B. Lakeman, I.M. Mellor, R.J. Mor-
This study obtained high degree of graphitization and large timer, M.C. Turpin, Studies into the storage of hydrogen in carbon
numbers of open edges in PGNFs prepared from the ther- nanofibers: proposal of a possible reaction mechanism, Nano Lett. 2 (2002)
mal decomposition of a mixture containing PEG and NiCl2 . 201–205.
[20] E. Bekyarova, K. Murata, M. Yudasaka, D. Kasuya, S. Iijima, H. Tanaka, H.
Hydrogen storage in PGNFs, measured using the volumetric Kahoh, K. Kaneko, Single-wall nanostructured carbon for methane storage,
method at 298 K and a pressure up to 4.83 MPa, possessed J. Phys. Chem. B 107 (2003) 4681–4684.
a 3.3 wt.% hydrogen storage capacity. The d-spacing of the [21] A. Thess, R. Lee, P. Nikolaev, H. Dai, P. Petit, J. Robert, C. Xu, Y.H.
PGNFs underwent an expansion and returned back to the ini- Lee, S.G. Kim, A.G. Rinzler, D.T. Colbert, G.E. Scuseria, D. Tománek,
J.E. Fischer, R.E. Smalley, Crystalline ropes of metallic carbon nanotubes,
tial value confirmed by XRD analysis. The graphene layer
Science 273 (1996) 483–487.
d-spacing expansion was induced from 3.362 Å (prior to adsorp- [22] Z.F. Ren, Z.P. Huang, J.W. Xu, J.H. Wang, P. Bush, M.P. Siegal, P.N.
tion) to 3.398 Å (immediately after hydrogen uptake) and then Provencio, Synthesis of large arrays of well-aligned carbon nanotubes on
decreased to 3.38 Å (a week after hydrogen uptake at ambient glass, Science 282 (1998) 1105–1107.
conditions). As a result, a residual amount of hydrogen may [23] S.H. Joo, S.J. Choi, I. Oh, J. Kwak, Z. Liu, O. Terasaki, R. Ryoo, Ordered
nanoporous arrays of carbon supporting high dispersions of platinum
remain in PGNFs, which perturbed the PGNF graphene layer
nanoparticles, Nature 412 (2001) 169–172.
structure. [24] P.F. McMillan, New materials from high-pressure experiments, Nat. Mater.
1 (2002) 19–25.
References [25] H. Murayama, T. Maeda, A novel form of filamentous graphite, Nature 345
(1990) 791–793.
[1] H.W. Kroto, J.R. Heath, S.C. Obrien, R.F. Curl, R.E. Smalley, C-60- [26] H.P. Boehm, Carbon from carbon monoxide disproportionation on nickel
buckminsterfullerene, Nature 318 (1985) 162–163. and iron catalysts: morphological studies and possible growth mechanisms,
[2] S. Iijima, Helical microtubules of graphitic carbon, Nature 354 (1991) Carbon 11 (1973) 583–590.
56–58. [27] M. Endo, Y.A. Kim, T. Hayashi, Y. Fukai, K. Oshida, M. Terrones, T.
[3] D. Ugarte, Curling and closure of graphitic networks under electronbeam Yanagisawa, S. Higaki, M.S. Dresselhaus, Structural characterization of
irradiation, Nature 359 (1992) 707–709. cup-stacked-type nanofibers with an entirely hollow core, Appl. Phys. Lett.
[4] Y. Saito, K. Nishikubo, K. Kawabata, T. Matsumoto, Carbon nanocap- 80 (2002) 1267–1269.
sules and single-layered nanotubes produced with platinum-group metals [28] C. Park, R.T.K. Baker, Catalytic behavior of graphite nanofiber supported
(Ru, Rh, Pd, Os, Ir, Pt) by arc discharge, J. Appl. Phys. 80 (1996) 3062– nickel particles. 2. The influence of the nanofiber structure, J. Phys. Chem.
3067. B 102 (1998) 5168–5177.
[5] T.C. Liu, Y.Y. Li, Synthesis of carbon nanocapsules and carbon nanotubes [29] O.C. Carneiro, N.M. Rodriguez, R.T.K. Baker, Growth of carbon nanofibers
by an acetylene flame method, Carbon 44 (2006) 2045–2050. from the iron-copper catalyzed decomposition of CO/C2 H4 /H2 mixtures,
[6] A. Krishnan, E. Dujardin, M.M.J. Treacy, J. Hugdah, S. Lynum, T.W. Ebbe- Carbon 43 (2005) 2389–2396.
sen, Graphitic cones and the nucleation of curved carbon surfaces, Nature [30] O.C. Carneiro, M.S. Kim, J.B. Yim, N.M. Rodriguez, R.T.K. Baker, Growth
388 (1997) 451–454. of graphite nanofibers from the iron-copper catalyzed decomposition of
[7] Y. Gogotsi, J.A. Libera, N. Kalashnikov, M. Yoshimura, Graphite polyhe- CO/H2 mixtures, J. Phys. Chem. B 107 (2003) 4237–4244.
dral crystals, Science 290 (2000) 317–320. [31] A. Tanaka, S.H. Yoon, I. Mochida, Preparation of highly crystalline
[8] N.M. Rodriguez, A. Chambers, R.T.K. Baker, Catalytic engineering of nanofibers on Fe and Fe–Ni catalysts with a variety of graphene plane
carbon nanostructures, Langmuir 11 (1995) 3862–3866. alignments, Carbon 42 (2004) 591–597.
[9] S.J. Tans, A.R.M. Verschueren, C. Dekker, Room-temperature transistor [32] B.K. Gupta, O.N. Srivastava, Further studies on microstructural characteri-
based on a single carbon nanotube, Nature 393 (1998) 49–52. zation and hydrogenation behavior of graphitic nanofibres, Int. J. Hydrogen
[10] S.J. Tans, M.H. Devoret, H. Dai, A. Thess, R.E. Smalley, L.J. Geerligs, C. Energy 26 (2001) 857–862.
Dekker, Individual single-wall carbon nanotubes as quantum wires, Nature [33] M.S. Dresselhaus, G. Dresselhaus, Intercalation compounds of graphite,
386 (1997) 474–477. Adv. Phys. 30 (1981) 139–326.
C.-W. Huang et al. / Separation and Purification Technology 58 (2007) 219–223 223

[34] C. Park, P.E. Anderson, A. Chambers, C.D. Tan, R. Hidalgo, N.M. [41] A.D. Lueking, L. Pan, D.L. Narayanan, C.E.B. Clifford, Effect of expanded
Rodriguez, Further studies of the interaction of hydrogen with graphite graphite lattice in exfoliated graphite nanofibers on hydrogen storage, J.
nanofibers, J. Phys. Chem. B 103 (1999) 10572–10581. Phys. Chem. B 109 (2005) 12710–12717.
[35] C.C. Ahn, Y. Ye, B.V. Ratnakumar, C. Witham, R.C. Bowman, B. Fultz, [42] A.C. Dillon, K.M. Jones, T.A. Bekkedahl, C.H. Kiang, D.S. Bethune, M.J.
Hydrogen desorption and adsorption measurements on graphite nanofibers, Heben, Storage of hydrogen in single-walled carbon nanotubes, Nature 386
Appl. Phys. Lett. 73 (1998) 3378–3380. (1997) 377–379.
[36] M. Rzepka, E. Bauer, G. Reichenauer, T. Schliermann, B. Bernhardt, K. [43] K. Shindo, T. Kondo, M. Arakawa, Y. Sakurai, Hydrogen adsorp-
Bohmhammel, E. Henneberg, U. Knoll, H.E. Maneck, W. Braue, Hydrogen tion/desorption properties of mechanically milled activated carbon, J. Alloy
storage capacity of catalytically grown carbon nanofibers, J. Phys. Chem. Compd. 359 (2003) 267–271.
B 109 (2005) 14979–14989. [44] C.W. Huang, Y.Y. Li, In situ synthesis of platelet graphite nanofibers from
[37] G.G. Tibbetts, G.P. Meisner, C.H. Olk, Hydrogen storage capacity of thermal decomposition of poly(ethylene glycol), J. Phys. Chem. B 10
carbon nanotubes, filaments, and vapor-grown fibers, Carbon 39 (2001) (2006) 23242–23246.
2291–2301. [45] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular
[38] E. Poirier, R. Chahine, T.K. Bose, Hydrogen adsorption in carbon nanos- layers, J. Am. Chem. Soc. 60 (1938) 309–319.
tructures, Int. J. Hydrogen Energy 26 (2001) 831–835. [46] A. Zuttel, P. Sudan, P. Mauron, P. Wenger, Model for the hydrogen
[39] Y.Y. Fan, B. Liao, M. Liu, Y.L. Wei, M.Q. Lu, H.M. Cheng, Hydrogen adsorption on carbon nanostructures, Appl. Phys. A: Mater. 78 (2004)
uptake in vapor-grown carbon nanofibers, Carbon 37 (1999) 1649–1652. 941–946.
[40] R. Ströbel, L. Jörissen, T. Schliermann, V. Trapp, W. Schütz, K. Bohmham- [47] A.D. Lueking, R.T. Yang, N.M. Rodriguez, R.T.K. Baker, Hydrogen stor-
mel, G. Wolf, J. Garche, Hydrogen adsorption on carbon materials, J. Power age in graphite nanofibers: effect of synthesis catalyst and pretreatment
Sources 84 (1999) 221–224. conditions, Langmuir 20 (2004) 714–721.

You might also like