Вы находитесь на странице: 1из 13

REVIEW

Recent Advances in Hydrogen Storage in Metal-Containing Inorganic Nanostructures and Related Materials
By Abdul M. Seayad and David M. Antonelli*
An overview of recent advances in the application of non-carbonaceous nanostructured and composite materials in hydrogen storage is presented in this review. The main focus is on complex hydrides, nongraphitic nanotubes, and other porous composite and framework materials, since carbon nanotubes have been the subject of numerous other reviews. Recent advances in the area of alanates show a promising reversible absorption capability of up to 5 %, closing in on the projected Department of Energy (DOE) target of 6 %. Non-carbon nanotubes mainly showed a sorption capacity of 13 wt.-%, although a promising level of 4.2 wt.-% is shown by boron nitride nanotubes after collapse of their walls. Other interesting materials included here are lithium nitride and porous metallo-organic frameworks.

1. Introduction
The demand for an efficient and clean fuel alternative has increased in recent years[1] and is expected to become more pronounced in the future,[2,3] since many automotive companies have announced the potential appearance of commercially available fuel cell vehicles as soon as 2010. Hydrogen is considered one of the best alternative fuels[46] due its abundance, easy synthesis, and non-polluting nature when used in fuel cells.[7,8] However, the main concern is the efficient storage and transport of this highly flammable gas.[9] The expected hazards involved in storing and using gaseous hydrogen in high-pressure vessels has triggered research worldwide on more safe and efficient ways for hydrogen storage for stationary or mobile applications.[10] While stationary hydrogen fuel cells are less exacting in terms of component sizes and storage capacities, extremely efficient hydrogen storage methods, from both the gravimetric and volumetric standpoints, are essential for vehicular applications.[11] A reversible hydro-

[*]

[+]

Prof. D. M. Antonelli, Dr. A. M. Seayad[+] Department of Chemistry and Biochemistry University of Windsor Windsor, Ontario, N9B 3P4 (Canada) E-mail: danton@uwindsor.ca Present address: Max-Planck Institut fr Kohlenforschung, Kaiser-Wilhelm-Platz 1, D-45470 Mlheim an der Ruhr, Germany.

gen sorption capacity of 56 wt.-% at 100 C and 0.1 MPa are targeted for automotive applications. The main challenges in the field of hydrogen storage are to devise new materials or combinations of materials to exhibit 1) high volumetric/gravimetric[12] capacity, 2) fast sorption kinetics at near-ambient temperatures, and 3) high tolerance to recycling. One of the most promising classes of materials for hydrogen storage are nanostructured composites, because they have dramatically different chemical, physical, thermodynamic, and transport properties as compared to their bulk counterparts. Due to the wide range of compositions, the ability to tailor pore and grain sizes, and the capacity to intimately weave two or more phases together at the nanometer level, nanophase composite materials may open the window to greater hydrogen-storage capacities and lower kinetic adsorption barriers as compared to coarse-grained materials.[13] Carbon nanotubes (CNTs)[14] continue to be an important category of nanomaterials for possible applications in hydrogen storage,[1517] displaying an acceptable capacity range of 24 wt.-% under suitable conditions;[18] however, recent interest has shifted away from this area because performance under operating conditions is not as efficient as originally anticipated. For example, single-walled carbon nanotubes after ultra-sonication show hydrogen uptake at room temperature, although this storage may be due to metal particles incorporated during the sonication treatment. Reactive high-energy ball-milling of graphite leads to a material with a high hydrogen-loading capacity, but the temperatures required for hy 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2004, 16, No. 910, May 17

DOI: 10.1002/adma.200306557

765

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

drogen release are far too high for applications.[19] In view of today's knowledge, the amount of hydrogen required for automotive applications greatly exceeds that offered by carbon nanotubes.[20] Other carbon structures[21,22] and doped CNTs have improved hydrogen sorption capacities, so applications may yet be on the horizon.[23] Several reviews have recently appeared on this subject and interested readers may refer to them.[2427] In this review we present recent developments in hydrogen storage using nanostructured and composite materials other than carbon nanotubes, as these materials show great promise in this exciting and rapidly expanding area of research. The main focus is on nanocrystalline complex hydrides, inorganic nanotubes such as boron nitride (BN), titanium sulfide (TiS2) and molybdenum sulfide (MoS2), as well as lithium nitride, metalorganic frameworks, and low-valency mesoporous titanium.

2. Metal Hydrides
Metal hydrides are a promising means of effectively storing hydrogen due to their high storage capacities at low pressures, whilst they also maintaining volumetric densities comparable to that of liquid hydrogen.[28] The highest volumetric densities of hydrogen in any material are found in metal hydrides (Fig. 1), which are defined as a concentrated single-phase compounds involving the host metal and hydrogen.[29]

Hydrides can be broadly classified as 1) ionic hydrides in which hydrogen exists as H (e.g., MH, M is mainly alkali or alkaline earth metals), 2) covalent hydrides in which hydrogen shares the electron pair with non-metals or atoms with similar electronegativities (e.g., H2O, H2S, SiH4, hydrocarbons, etc.) and metal hydrides in which hydrogen acts as a metal and are formed mainly with transition metals including the rare earth and actinide series. Metallic hydrides form a wide range of stoichiometric and non-stoichiometric compounds by direct interaction of hydrogen with metals.[30] The hydrogen atom enters the metal lattice to form a solid solution and the metal hydride begins to crystallize when the local hydrogen concentration exceeds a certain limit. The rate of entry and reaction[31] depends in part on the metal particle size, and for this reason more efficient hydrogen sorption capacity can be achieved with nanosized metals as compared to their bulk counterparts. Metal hydrides containing only one metal have limited practical applications in hydrogen storage because of the high thermodynamic stability of these compounds. For this reason, a wide range of alloys with two or more metals have been investigated over the past several decades in order to find a material that meets the practical requirements. These alloys show more complex thermodynamics and phase diagrams than the pure metal hydrides, and may therefore lead to a material with commercial hydrogen storage applications.[32] A wide range of nanocrystalline metal hydrides can be prepared by mechanochemical process such as high-energy ball-

David Antonelli was born in Chicago in 1963 and raised in the United States, Great Britain, and Canada. He completed his BSC at the University of Alberta in 1987 and his Ph.D. in organometallic chemistry with Martin Cowie at the University of Alberta in 1990. He was an NSERC postdoctoral fellow in organometallic chemistry at Oxford University with Malcolm Green in 1991, and at Caltech in 199293 with John E. Bercaw. He later studied nanoporous materials as a postdoctoral associate with Prof. Jackie Ying at the Massachusetts Institute of Technology Department of Chemical Engineering from 19941996. He is currently an Associate Professor in Chemistry at the University of Windsor and directs a research group focused on the electronic and catalytic properties of nanoporous materials with variable oxidation states and conducting molecular wires in the pores. Prof. Antonelli is the author of over 50 publications and was the winner of several awards, including the Ontario Premier's Research Excellence Award, a Royal Society of Britain Research Award, and a prestigious NSERC AGENO award. Abdul Majeed Seayad was born and brought up in Kerala, India. After completing an M.Sc. in Chemistry at the University of Kerala (1995), he joined the homogeneous catalysis division of the National Chemical Laboratory, Pune, India as a CSIR (India) research fellow (19952000) for his doctoral thesis under the guidance of Dr. R. V. Chaudhari, and received his Ph.D. degree in Chemistry in 2000 from the University of Pune. Between 2001 and 2003 he worked as an Alexander von Humboldt research fellow in the research group of Professor Mathias Beller at Leibniz-Institut fr Organische Katalyse (IfOK), Universitt Rostock e.V., Rostock, Germany. He then worked with Professor David M. Antonelli, University of Windsor, Canada as a postdoctoral research fellow working with hydrogen-storage materials. Presently he is working as a postdoctoral research fellow at the Max-Planck Institut fr Kohlenforschung, Mlheim, Germany in the research group of Professor Benjamin List. His research interests include C1 chemistry, catalytic organic synthesis and hydrogen-storage materials. 766
 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 910, May 17

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

Figure 1. Gravimetric and volumetric densities, and corresponding specific energies and energy densities, of a variety of different hydrogen-storage media, including two common fossil fuels: gasoline and propane. The FreedomCar targets for 2005, 2010, and 2015 are indicated within the shaded regions. The gravimetric hydrogen densities for pressurized hydrogen gas and cryogenic liquid hydrogen (diamond symbols) include the mass of the storage container, whereas the values reported for the metal hydrides, complex hydrides, and other hydrogen-storage materials are based on the absolute (theoretical) amount of hydrogen. (Reproduced with permission from [2b]. Copyright Elsevier 2003.)

milling.[3336] This technique provides easy preparation of samples,[37] flexible grain-size control, as well as easy scale-up,[38] and the nanosized metal hydrides[39] thus-formed show higher hydrogen sorption capacities compared to the bulk hydride materials. Once ball-milled, these metal hydrides and intermetallic compounds[40] generally need to be activated, and the first cycles of hydrogenation and dehydrogenation should be performed under relatively high temperatures and pressures. The hydrogen-sorption kinetics were found to be increased by ball-milling in the presence of certain organic additives.[4144] For example, when magnesium and graphite are milled in the presence of benzene,[45] cyclohexane,[46] or tetrahydrofuran,[47,48] the hydriding characteristics were increased as compared to that of magnesium alone, or magnesium and graphite milled together. During this milling process, the graphite layered structure breaks apart to form graphite lamellae of approximately 20 nm dispersed on magnesium particles. When the pure metals (Ti, V, or Mg) or intermetallics such as FeTi, Mg2Ni, etc., or the nanocomposites such as MgV[4951] are milled for about 30 min in the presence of 10 wt.-% graphite, the initial hydrogenation kinetics were found to in-

crease considerably.[52] A recent review on this subject (nanocrystalline materials for hydrogen storage) is presented by Huot[53] and interested readers may refer to this for further reading on the preparation, properties, and hydrogen-storage capabilities of various nanostructured metal hydrides and related intermetallic compounds. From recent research it is apparent that the main disadvantages of metal hydrides, even in their nanocrystalline form, are their low gravimetric hydrogen content, the higher temperature needed for desorption of hydrogen, and problems associated with their regeneration. Other problems associated with their use include the cost, low specific uptake by weight in many cases, unfavorable kinetics requiring heating cycles, and susceptibility to contamination by impurities.

2.1. Complex Hydrides Complex hydrides are inorganic salt-like compounds of anions such as [BH4], [AlH4], stabilized mainly with light metal cations. The hydrogen in the complex hydrides is located at 767

Adv. Mater. 2004, 16, No. 910, May 17

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

the corners of a tetrahedron with the metal in the center, and the negative charge of the anion, [BH4] and [AlH4], is compensated by a cation such as Li or Na. The hydride complexes of boron, the tetrahydroborates M(BH4), and of aluminum, the tetrahydroaluminate M(AlH4), are perhaps the most promising hydrogen-storage materials discovered to date, as they have shown promising levels of reversible adsorption according to the DOE targets of 56 wt.-%. In their pure bulk form they are thermodynamically quite stable and decompose only at elevated temperatures, often above the melting point. However, the decomposition temperature can be reduced and the hydrogen adsorption efficiency improved by using various dopants or reducing the grain size. Some of the recent developments in complex alanates and borates are reviewed below. 2.1.1. Aluminum Hydrides Aluminum hydrides of light alkali metals are all very promising hydrogen-storage materials. This is because they all contain a high weight percentage of hydrogen, much of which can be removed and replaced continuously. One drawback of alkali aluminum hydrides, however, is that they are not always easy to prepare. Na3AlH6 can be synthesized by reaction of NaH and NaAlH4 in heptane at 165 K and 140 bar of hydrogen[54] or by direct reaction of sodium and aluminum in toluene (438 K and 350 bar hydrogen).[55] Similarly, Na2LiAlH6 can be prepared from LiAlH4 and NaH in toluene under 300 bar of hydrogen at 160 C.[56] It can also be formed by the reaction of NaAlH4, LiH, and NaH in heptane under hydrogen pressure.[57] NaAlH4 prepared and purified in different ways leads to materials of different morphologies and grain sizes. For example, Bogdanovic et al.[58] have shown that different methods of isolation of NaAlH4 from a tetrahydrofuran (THF) solution of commercial Na alanate leads to crystallization of NaAlH4 of different particle sizes and shapes, all of which were characterized by scanning electron microscopy (SEM) investigations (Fig. 2). Large NaAlH4 crystals of about 50 25 lm size and 0.8 m2 g1 specific surface area are obtained through ether precipitation (Fig. 2a), due to low rate of nucleation and uniform crystal growth. Precipitation by pentane (Fig. 2b) produced NaAlH4 as an agglomerate of particles of average 10 20 lm size and specific surface of approx. 2.5 m2 g1 with a small amount in the range of 50 lm. Still finer NaAlH4 particles of approx. 510 lm size (Fig. 2c) were obtained by pouring solutions of NaAlH4 in THF into pentane. These two fine Na alanate samples (obtained through pentane precipitation) showed a higher hydrogen-sorption capacity compared to that of the large-grained samples when doped with Ti particles. Further reduction of the size of Na alanates to the nanoscale is expected to be advantageous. Most of these preparations require filtration, washing, and drying to obtain a purified product. Hence a low temperature and pressure preparation method that gives high yield without purification steps may be more suitable for commercial applications. Huot et al.[59] have shown that nanocrystalline Na3AlH6 and Na2LiAlH6 can easi-

Figure 2. Scanning electron microscopy (SEM) images of NaAlH crystals obtained by precipitation of NaAlH from THF solutions by addition of ether (a) or pentane (b), or by pouring THF solutions of NaAlH into pentane (c). (Reproduced with permission from [58]. Copyright Elsevier 2000.)

5,5 5 4,5 Capacity [wt %] 4 3,5 3 2,5 2 1,5 1 0,5 0


0 4 8 12 16 20 24

Cycle-No.

Figure 3. Hydrogen-storage capacity level in a 25 cycle test of NaAlH4 doped with 2 mol-% TiN nanoparticles. Dehydrogenation at 120/180 C and normal pressure; hydrogenation at 100 C and 10085 bar pressure.

ly be produced by energetic ball-milling of NaH, LiH, and NaAlH4 in stoichiometric composition. Thus, high-energy ball-milling can produce nanoscale hydrides quickly and easily with potentially improved performances as compared to the bulk materials. The mechanism and dynamics of hydrogen desorption have been extensively studied. Thermal decomposition of NaAlH4 at higher temperatures takes place in two steps to give NaH,

768

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 910, May 17

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

Al, and H2. In principle the first step can give 3.7 wt.-% H2, and up to 5.5 wt.-% in the second step: NaAlH4 1/3 Na3AlH6 + 2/3 Al + H2 (3.7 wt.-% H2) 1/3 Na3AlH6 NaH + Al + 3/2 H2 (1.8 wt.-% H2) (1) (2)

9.3 wt.-%. Studies have shown that it decomposes in two major steps (Eqs. 6,7). The first decomposition temperature is 163 C and the resulting residue at 200 C consists of MgH and Al, which continues to release hydrogen in the temperature range of 240380 C and transforms into a mixture of Al and the intermetallic compound Al3Mg2 (Eq. 8) at still higher temperatures (400 C).[72] Mg(AlH4)2 MgH2 + 2 Al + 3 H2 MgH2 Mg + H2 2 Al + Mg 1/2 Al2Mg3 + 1/2 Al (6) (7) (8)

Decomposition of NaH to Na and hydrogen requires a still higher temperature. The first two steps of decomposition can be accelerated and the decomposition temperature can be substantially reduced by doping with other metal cations such as Ti3+, Ti4+, Zr4+, Fe3+, etc., as demonstrated by Bogdanovic et al.[57,58,63,64] and others.[6062] X-ray diffraction (XRD) analysis and solid-state NMR spectroscopy shows that the majority of phases involved are NaAlH4, Na3AlH6, Al, and NaH.[63] Other traces of unidentified phases are also observed, one of which has been tentatively assigned to an AlTi alloy. By variation of NaAlH4 particle sizes, dopants (catalysts), and doping procedures, kinetics and the de- and rehydrogenation stabilities within different cycles can be substantially improved,[58] underscoring the importance of nanocrystalline processing in this field. Recently Bogdanovic et al.[64] reported that doping Na alanate with TiN nanoparticles could substantially reduce the temperature required for decomposition as well as the hydrogenation time required for practical purposes, and the desorbed hydrogen could reach close to the theoretical limit. However, the hydrogenation capacity for these materials is found to slightly decrease over the course of several cycles before stabilizing (Fig. 3). Anton[65] has recently studied the effect of a wide range of different dopants on the hydrogen-sorption capacity and kinetics of Na alanate. In general, it was found that the amount and type of dopant had a substantial effect on these parameters, with Ti providing the best results. In other work, it was shown that hydrogenation kinetics can be improved by the addition of 10 wt.-% carbon to sodium alanates.[66] XRD[67] and microstructural[68] characterization were performed on a series of doped alanates to understand the mechanism of action of these catalysts; the results of this study, however, were ambiguous.[69] Recently Vajeeston et al.[70] attempted to elucidate the pressure-induced structural changes occurring during hydrogen adsorption, but the detailed mechanism of this process was not completely resolved. In contrast to NaAlH4, KAlH4 smoothly decomposes without a transition metal catalyst (Eqs. 35) to give ~ 3.5 wt.-% (about 80 % of the theoretical) at a temperature range of 250350 C and is reversible.[71] 3 KAlH4 K3AlH6 + 2 Al + 3 H2 (H = 2.9 wt.-%) K3AlH6 3 KH + Al + 3/2 H2 (H = 1.4 wt.-%) 3 KH 3 K + 3/2 H2 (H = 1.4 wt.-%) (3) (4) (5)

In the first decomposition step, 6.6 wt.-% of hydrogen is released. Synthesis of Mg(AlH4)2 is generally achieved via a metathesis reaction between sodium alanate and magnesium chloride.[73,74] Peak-shape analysis of the XRD pattern indicates that the magnesium alanate produced is a nanocrystalline material with a mean grain size of 30 nm. The main difference in the decomposition of Mg(AlH4)2 as compared to other Li and Na alanates is that it transforms into a non-alanate metal hydride during the first decomposition step. Doping with TiCl3 and reducing the grain size by ball-milling was not found to be beneficial to the dehydrogenation temperature or the kinetics in this case. The formation of stable MgH as an intermediate, or Al2Mg3 as an end product, negatively influences the re-adsorption of hydrogen and is the main drawback of this system. The effects of grain size and doping need to be further investigated before promising reversible hydrogen-storage behavior is achieved. 2.1.2. Borohydrides Borohydride complexes with suitable alkali or alkaline earth metals are a promising class of compounds for hydrogen storage. The hydrogen content can reach values of up to 18 wt.-% for LiBH4. Thermal analysis shows mainly three decomposition peaks for LiBH4 (Eqs. 911).[75] The first peak occurs at around 100 C and corresponds to a structural transition from orthorhombic to polycrystalline with a small liberation (0.3 wt.-%) of hydrogen. A fusion is then observed around 270 C without liberation of hydrogen. At 320 C the first significant hydrogen desorption peak occurs as the material liberates an additional 1 wt.-% of hydrogen. The second desorption peak begins at 400 C and reaches its maximum around 500 C. The total amount of hydrogen desorbed up to the temperature of 600 C is 9 wt.-%,[76] which corresponds exactly to half of the hydrogen in the starting compound. The end product has the nominal composition LiBH2''.[77] LiBH4 LiBH4e + 1/2 (e)H2 (structural transition at T = 108 C) LiBH4e LiBH2 + 1/2 (1e)H2 (first hydrogen peak starting at T = 200 C)

(9)

Magnesium alanate is another interesting complex hydride, which has a theoretical hydrogen-storage capacity of

(10)

Adv. Mater. 2004, 16, No. 910, May 17

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

769

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

LiBH2 LiH + B + 1/2 H2 (second hydrogen peak starting at T = 453 C)

(11)

Mixing LiBH4 with SiO2 powder (25:75 by weight) lowers the desorption temperature so that 9 wt.-% of hydrogen is liberated below 400 C, reaching a maximum 13.5 wt.-% at higher temperatures.[78] The thermal-desorption spectrum (Fig. 4) in this case also exhibits three hydrogen desorption peaks.

Figure 4. Thermal desorption spectra of LiBH4. The sample was heated after evacuation at room temperature with a heating rate of 2 K min1. The gas flow was measured as a function of time and the desorbed hydrogen was computed from the integrated gas flow: a) pure LiBH4 and b) LiBH4 mixed with SiO2 as catalyst. (Reproduced with permission from [78]. Copyright Elsevier 2003.)

NaBH4 solutions are non-flammable and stable in air for months. d H2 generation only occurs in the presence of selected catalysts and can even occur at 0 C. d The only other product in the gas stream is water vapor and other products can be recycled. d H2 generation rates are easily controlled. d Volumetric and gravimetric H2 storage efficiencies are high. These points demonstrate that aqueous alkaline borohydride is promising from a commercial standpoint in situations calling for a high-density source of hydrogen. While the generation of hydrogen from sodium borohydride is not reversible and leads to borate waste materials, it is an excellent choice for one-time applications due to its high volumetric hydrogen content, which is greater than compressed hydrogen in gas cylinders. In conclusion, it appears that reducing grain sizes and adding small amounts of dopants to metal hydrides are effective ways to improve performance of many metal hydrides in hydrogen-storage conditions. In addition to this, alkaline solutions of borohydrides may also offer practical advantages in certain situations where the emphasis is on one-time use of a fuel course without the requirement of reversibility. From the results presented in this review it is apparent that this is a burgeoning area of research with great commercial promise, and that further development is expected to yield even better results.
d

The addition of SiO2 powder catalyzes the decomposition reaction of LiBH4 and lowers the temperature for all three hydrogen-desorption features. A total of 4.5 wt.-% of the hydrogen remains as LiH in the decomposed product. All attempts to synthesize LiBH4 from the elements at elevated temperatures up to 650 C and 150 bar H2 pressure failed. LiBH4 is generally synthesized from the reaction of ethyllithium with diborane[79] or the direct reaction of the corresponding metal with diborane in ether solvents under suitable conditions.[80] In most of the reports, the grain size distribution is not available. However, because of the results obtained in other hydride systems, it can be easily seen that preparing nanoscale lithium borohydrides would be advantageous for improved hydrogen-sorption properties. Solid and nanocrystalline borohydrides of sodium[81] and potassium[82] are less explored, but molecularly dispersed aqueous borohydride solution is a safe method for liberating hydrogen in a controllable way and represents an alternative method to the use of solid-phase systems.[83] The spontaneous hydrolysis of borohydrides in aqueous solution is in general inhibited by the addition of NaOH, and these solutions can effectively liberate hydrogen in the presence of catalysts such as Ru[84] or Ni.[85] In principle, aqueous NaBH4 solutions can generate 7.3 wt.-% hydrogen and up to 6.5 wt.-% hydrogen liberation is achieved at only 40 C by using 10 wt.-% NaOH solution. Generating hydrogen from alkaline NaBH4 solutions in this way has many attractive advantages, such as:

3. Nanotubes
Nanotubes are another important category of promising materials for hydrogen storage. Major developments in non-carbonaceous nanotubes, including boron nitride (BN), titanium sulfide (TiS2), and molybdenum sulfide (MoS2), are reviewed in this section.

3.1. Boron Nitride Nanotubes Boron nitride (BN) nanotubes are interesting materials, as they are isoelectronic and isostructural to carbon nanotubes. Recently, Ma et al.[86] prepared two kinds of BN nanotubes, multiwalled and bamboo BN nanotubes, using chemical vapor deposition (CVD) by pyrolyzing a B-N-O precursor at 2000 K under a N2/NH3 atmosphere.[87,88] These nanotubes were found to absorb hydrogen at levels up to 1.8 and 2.6 wt.-% respectively at about 10 MPa. This compares well to the negligible 0.2 wt.-% absorption in conventional BN powders as shown in Figure 5. The low hydrogen absorption in multiwalled nanotubes is mainly due to the fact that these have closed ends (Fig. 6a) and the absorption can therefore only take place on the outside surface and interstitial sites of the tubes and bundles. On the other hand, bamboo nanotubes, which are regarded as

770

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 910, May 17

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

Figure 5. Hydrogen adsorption as a function of pressure in multiwalled BN nanotubes and bamboo nanotubes as compared to bulk BN powder at 10 MPa. (Reproduced with permission from [86]. Copyright American Chemical Society 2002.)

Figure 7. a) TEM image of BN nanotubes after heating at 1500 C for 6 h in the presence of platinum; the arrow points to a platinum nanoparticle. b) High-resolution TEM image of the collapsed BN nanotubes. (Reproduced with permission from [90]. Copyright American Chemical Society 2002.)

CVD method using a mixture of B2O2 and Mg) to 1500 C on the surface of platinum plates. In contrast, BN nanomaterials prepared in the presence of LaB6 and Pd/boron powder show a hydrogen storage capacity of 3 wt.-%.[91]

3.2. Titanium Sulfide Nanotubes (TiS2)


Figure 6. The morphologies of BN nanotubes: a) multiwall nanotubes and b) bamboo-like nanotubes. Scale bar: 100 nm. (Reproduced with permission from [86]. Copyright American Chemical Society 2002.)

polymerized nanobells (Fig. 6b), usually exhibit a more defective structure and have many open-edge layers on the exterior surface that may contribute to a higher hydrogen-absorption capacity. The mechanism of absorption is believed to be chemical, and about 70 % of the absorbed hydrogen is retained after depressurizing. The slow equilibration time (approx. 4 h) also suggests that the adsorption is mainly due to a chemical interaction. Highly ball-milled nanocrystalline BN powders can also absorb up to 2.6 wt.-% of hydrogen.[89] A higher hydrogen-absorption capacity of up to 4.2 wt.-% at 10 MPa was observed for BN nanotubes with collapsed walls.[90] The structural collapse, as shown in the transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images (Fig. 7), significantly increases the surface area from 254.2 to 789.1 m2 g1. This indicates that surface area is more important in hydrogen uptake than mesoporosity. These materials are prepared by heating BN nanotubes (prepared by the
Adv. Mater. 2004, 16, No. 910, May 17

Titanium sulfide is an interesting material for hydrogen storage, since foreign atoms can be easily intercalated in between the STiS layers that are held by van der Waals' interactions. This gives facile compositional flexibility. Chen et al.[92] have synthesized multiwalled TiS2 nanotubes with uniform open-ended tubular structures (Fig. 8) with an outer diameter of ~ 30 nm, an inner diameter of ~ 10 nm, and an interlayer spacing of ~ 0.57 nm. These nanostructures are composed of nanocrystalline TiS2 with a hexagonal structure. These materials reversibly absorb 2.5 wt.-% hydrogen at 25 C and about 4 MPa. The absorption capacity was found to decrease with rising temperature, as shown in Figure 9. TiS2 nanotubes absorb hydrogen both chemically (40 %) and physically (60 %).

3.3. Molybdenum Sulfide Nanotubes (MoS2) Nanotubes of the type MS2 (M = Mo, W)[9395] are analogous to pure carbon nanotubes. MoS2 nanotubes were synthesized by direct reaction of (NH4)2MoS4[96] and hydrogen.[97] Polycrystalline (NH4)2MoS4 is first ball-milled in an atmosphere of hydrogen. The fine powder is then transferred onto
 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

771

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW
Figure 8. TEM (a,b) and HRTEM (c) images of the as-synthesized TiS2 nanotubes. (Reproduced with permission from [92]. Copyright American Chemical Society 2003.)

Figure 9. PCT curves for hydrogen absorption and desorption of TiS2 nanotubes at 25, 75, and 125 C. (Reproduced with permission from [92]. Copyright American Chemical Society 2003.)

the surface area of the nanotubes increased, mainly because of defects induced in the nanotube multiwalls as evidenced form SEM and HRTEM analysis. The HRTEM images demonstrate that the nanotube tip is completely open, as shown in Figure 11. The outer diameter of a typical hollow tube is ~ 25 nm, while the inner diameter is ~ 10 nm. The average distance between each two neighboring fringes (c/2) is 0.63 nm, which corresponds to the interlayer (002) d-spacing of the 2H-MoS lattice. It is also shown that after KOH treatment, more defects are introduced around the nanotube. The KOH-treated MoS2 nanotubes showed higher hydrogen-sorption capacity as compared to the untreated and the polycrystalline MoS2. This is shown in Figure 12.[98] The adsorption and desorption is claimed to be highly reversible at 25 C and the specific surface area of the KOH treated material is approx. 28 m2 g1 as compared to 22 and 3.6 m2 g1 in the untreated and polycrystalline MoS2, respectively.

an alumina substrate and sintered in floating hydrogen/thiophene at the relatively low temperature of 400 C for 1 h to form the wire-like nanotubes in 90 % purity with lengths of several hundred nanometers. These are shown along with polycrystalline MoS2 in Figure 10. When treated with KOH,

4. Miscellaneous Materials
In this section we review a small number of materials that cannot be included in the above sections. In general, metal

772

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 910, May 17

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

Figure 10. SEM images of MoS2: a) polycrystalline, b) nanotubes without KOH treatment, and c) nanotubes after KOH treatment. (Reproduced with permission from [98]. Copyright American Chemical Society 2003.)

nitrides, organic framework materials, and porous oxides absorb only small amounts of hydrogen, however the examples from each of these three classes below are all potential competitors to metal hydride and nanotube technology. This makes them particularly intriguing, as they awaken new interest in classes of materials that were thought to be uninteresting from the standpoint of hydrogen-storage performance.

4.1. Lithium Nitride Lithium nitride (Li3N) has a maximum theoretical hydrogen-sorption capacity of 11.5 wt.-%. It was reported as early as 1910 by Dafert and Miklauz[99] that the reaction between Li3N and H2 generates Li3NH4, which is a mixture of 2 LiH and LiNH2,[100] as per the equation below Li3N + 2 H2 Li3NH4 (2 LiH + LiNH2) (12)

Figure 11. a) TEM and b) HRTEM images of MoS2 nanotubes without KOH treatment, and c) HRTEM image of KOH-treated MoS nanotubes. (Reproduced with permission from [98]. Copyright American Chemical Society 2003.)

Hu and Ruckenstein[101,102] later showed that the complete recovery of Li3N from the hydrogenated compounds is a difficult process that requires high temperatures (above 430 C) and long times during which sintering occurs and leads to inefficient recovery of Li3N. For this reason, the reversible storage capacity of Li3N was thought to be limited to about 5 wt.-%. In a recent reinvestigation of this system, Chen et al.[103] reported that lithium nitride (Li3N) is a promising candidate for reversible hydrogen storage. Figure 13 shows the absorption

Figure 12. The hydrogen adsorption amount versus pressure of polycrystalline MoS2, and nanotubes without and after KOH treatment at 25 C. (Reproduced with permission from [98]. Copyright American Chemical Society 2003.)

Figure 13. Weight variations during hydrogen absorption and desorption processes over Li3N samples; Abs: absorption; Des: desorption. (Reproduced with permission from [103]. Copyright Nature Publishing Group 2002.)

Adv. Mater. 2004, 16, No. 910, May 17

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

773

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

desorption characteristics of a fresh Li3N sample. The absorption starts at a temperature of around 100 C and a rapid weight gain is observed in the temperature range of 170 210 C. Total hydrogen absorption of about 9.3 wt.-% was reported after maintaining the sample at 255 C for half an hour. Substantial absorption can also be obtained below 200 C if sufficient time is provided. About 6.3 wt.-% of hydrogen was desorbed below 200 C under vacuum (105 mbar), and the remaining hydrogen could only be desorbed at elevated temperatures (above 320 C). No ammonia formation[102] was detected under these conditions. Unlike most of the metal hydrides, which exhibit one plateau in the pressurecomposition (PC) isotherm, Li3N has two. The first one has a rather low equilibrium pressure (below 0.07 bar) and the second plateau is sloped (Fig. 13); the overall equilibrium pressure is below 0.2 bar at 195 C, 0.5 bar at 230 C, and 1.5 bar at 255 C. The hydrogen absorbed during the first plateau might not be easily desorbed, and should correspond to the high temperature desorbed portion. XRD measurements of samples with different degrees of hydrogenation showed clear phase and composition changes during the hydrogen absorption and desorption. Pristine hexagonal Li3N phase shifted to face-centered cubic (fcc) lithium imide and hydride phases after being half-hydrogenated. The fully hydrogenated Li3N sample is composed of body-centered tetragonal (bct) lithium amide (LiNH2) and an enhanced lithium hydride phase.[104] Phase changes in desorption follow a path that is almost the reverse of that observed during adsorption. Hence it can be deduced that the hydrogen sorption in Li3N may occur as per the following equation (Eq. 13) and the molar ratio of absorbed hydrogen to Li3N is four, about 11.5 wt.-% Li3N + 2 H2 Li2NH + LiH + H2 LiNH2 + 2 LiH (13)

The hydrogenated Ca2NH sample is composed of CaNH and CaH2, indicating that hydrogen is stored according to the following the reaction: Ca2NH + H2 CaNH + CaH2 (14)

All of these nitrides showed relatively fast kinetics in hydrogen storage. For a 500 mg sample, almost all hydrogen (for Li2NH and Ca2NH) or a substantial amount of hydrogen (for Li3N) can be absorbed within 10 min under 30 bar of hydrogen and at temperatures differing for each compound. The desorption rate strongly depends on temperature and hydrogen pressure. Though the nitride systems offer excellent potential, in order to meet practical applications at moderate temperatures with improved chemical stability, further improvements are required. It is conceivable that nanocrystalline grains and doped composite lithium and calcium nitride materials may offer substantially improved performance over the pure bulk phases, with much lower adsorption temperatures, much closer to the practical goal.

4.2. Microporous MetalOrganic Frameworks Rosi et al.[105] recently reported an interesting metalloorganic framework (MOF)[107] material with hydrogen-sorption capacities at 78 K or ambient temperature under safe pressures (up to 20 bar). These materials are crystalline metalorganic frameworks with cubic cavities of uniform size and internal structure. This work is of special interest because the majority of microporous framework materials composed of metal oxides (zeolites, etc.) have only moderate to poor hydrogen-storage capacities. The material MOF-5[107,108] (Fig. 14a), in which inorganic [OZn4]6+ groups are joined to an octahedral array of [O2CC6H4CO2]2 (1,4-benzenedicarboxylate, BDC) groups to form porous cubic framework, showed 4.5 wt.-% hydrogen absorption at 78 K and moderate pressures. However, only 1 wt.-% absorption was achieved at

A similar hydrogen-storage phenomenon was observed in a related CaNH system. A reversibly hydrogen-storage capacity of 1.9 wt.-% (theoretical maximum is 2.1 wt.-%) is achieved for Ca2NH over a temperature range of 350600 C.

Figure 14. Single-crystal X-ray structures of MOF-5 (a), IRMOF-6 (b), and IRMOF-8 (c) illustrated for a single cube fragment of their respective cubic three-dimensional extended structure. On each of the corners is a cluster [OZn4(CO2)6] of an oxygen-centered Zn4 tetrahedron that is bridged by six carboxylates of an organic linker (Zn: blue polyhedron; O: red spheres; C: black spheres). The large yellow spheres represent the largest sphere that would fit in the cavities without touching the van der Waals' atoms of the frameworks. Hydrogen atoms have been omitted. (Reproduced with permission from [105]. Copyright American Association for the Advancement of Science 2003.)

774

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 910, May 17

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

room temperature and 20 bar pressure, and the absorption capacity was found to increase with pressure as shown in Figure 15.

age applications. These materials possess the high surface area and porosity of nanotubes and framework structures, with accessible variable oxidation states that may be important in adsorption of hydrogen by chemical pathways. High surface area mesoporous titanium oxide[111] showed only below 1 wt.-% absorption at 200 C and 20 bar pressure. However these materials showed a promising reversible absorption level of 35 wt.-% when suitably reduced by bis(toluene)titanium, bis(toluene)vanadium, or a variety of lithium fullerides.[112] The mechanism of this adsorption is currently under investigation and further optimization of adsorption levels and temperatures is ongoing. During hydrogen cycling the porosity is lost, which may suggest that the role of the porosity is to allow uniform surface reduction and initial access of hydrogen to all of these sites. The poor performance of the unreduced materials underscores the importance of surface reduction in this system. Further work is in progress to optimize the conditions for maximum reversible absorption capacity. Suitably modified titanium oxide nanotubes[113115] may also be promising as an alternative for efficient hydrogen storage and need to be explored along with other non-carbon nanotubes[116119] and nanomaterials.[120122]

5. Summary
Suitable nanocrystalline materials such as metal hydrides or complex hydrides are attracting increased interest and showing promise for onboard hydrogen-storage applications. Control of composition and grain size and/or porosity as well as the synergistic effect of metals such as Ti and Fe or Ti and Zr, have been shown to play key roles in enhancing reversible hydrogen storage in a wide variety of metal alanate and porous inorganic systems. However, issues such as their hydrogenation kinetics, grain growth, and phase separation that serve to reduce the storage capacity need to be considered further. Various non-carbon nanotube materials are also worth analyzing for further improvements with respect to preparation, mechanism of hydrogen sorption, sorption kinetics, storage capacity, etc. Further, possibilities of easily prepared mesoporous and organic framework materials should be explored as alternative materials for hydrogen-storage applications. A summary of all the materials discussed here and their theoretical and observed hydrogen-sorption capacities is presented in Table 1. The main concerns and barriers for mobile applications of most of these materials are the cost, weight and volume, efficiency, durability, refueling time, codes and standards, and life-cycle and efficiency analyses. In considering the overall hydrogen economy, its impact on atmosphere[123] and climate change[124] due to the hydrogen leakage are also needs to be considered. The efficient collection/recycle of water exhaust from automotives should also be taken into account.
Received: December 1, 2003 Final version: March 18, 2004

Figure 15. Hydrogen gas-sorption isotherm for MOF-5 at a) 78 K and b) 298 K. (Reproduced with permission from [105]. Copyright American Association for the Advancement of Science 2003.)

The MOF-5 structure motif and related compounds are ideal for gas absorption, because the linkers are isolated from each other and accessible from all sides to sorbate gas molecules. The scaffolding-like nature of MOF-5 and its derivatives leads to extraordinarily high apparent surface areas (25003000 m2 g1) for these structures. In other similar structures, such as IRMOF-6 (Fig. 14b, with cyclobutylbenzene linker) and IRMOF-8 (Fig. 14c, with naphthalene linker), the specific H2 uptake is approximately doubled and quadrupled, respectively, compared to MOF-5 at room temperature and 20 bar pressure. The hydrogen-absorption capacity of these structures at room temperature is comparable to that of carbon nanotubes at cryogenic temperatures and can be fine tuned by modifying the structure with suitable linkers. 4.3. Mesoporous Transition Metal Oxides Recently, we became interested in mesoporous transition metal oxides[109,110] as a possible material for hydrogen-storAdv. Mater. 2004, 16, No. 910, May 17

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

775

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

Table 1. Summary of reversible hydrogen-storage capacity of various nanostructured materials.


H2 sorption capacity [wt.-%] Theoretical Complex hydrides NaAlH4 KAlH4 Mg(AlH4)2 LiBH4 Nanotubes BN TiS2 MoS2 Nitrides Li3N Ca2NH Microporous MOF 7.5 5.7 9.3 18 11.3 2.1 Maximum Achieved 5.3 [64] 3.5 [71] 6.6 [72] 13.3 [78] 4.2 [90] 2.5 [92] 1.2 [98] 9.3 [103] 1.9 [103] 4.5 [105]

Material

[1] D. Sperling, M. A. DeLuchi, Annu. Rev. Energy 1989, 14, 375. [2] a) C. C. Elam, C. E. G. Padr, G. Sandrock, A. Luzzi, P. Lindblad, E. F. Hagen, Int. J. Hydrogen Energy 2003, 28, 601. b) J. A. Ritter, A. D. Ebner, J. Wang, R Zidan, Mater. Today 2003, September, 18. [3] M. Conte, A. Iacobazzi, M. Ronchetti, R. Vellone, J. Power Sources 2001, 100, 171. [4] M. A. Paevey, Fuel from Water, Merit, Louisville, KY 1998. [5] P. Hoffmann, The Forever Fuel, Westview Press, Boulder, CO 1981. [6] R. M. Dell, D. A. J. Rand, J. Power Sources 2001, 100, 2. [7] B. D. McNicol, D. A. J. Rand, K. R. Williams, J. Power Sources 2001, 100, 47. [8] A. B. Stambouli, E. Traversa, Renewable Sustainable Energy Rev. 2002, 6, 297. [9] J. M. Ogden, Annu. Rev. Energy Environ. 1999, 24, 227. [10] L. Schlapbach, A. Zttel, Nature 2001, 414, 353. [11] G. D. Berry, S. M. Aceves, Energy Fuels 1998, 12, 49. [12] S. Louis, Hydrogen in Intermetallic Compounds I, in: Topics in Applied Physics, Vol. 63, Springer, Berlin 1988, p. 350. [13] A. Zaluska, L. Zaluski, J. O. Strom-Olsen, J. Alloys Compd. 2000, 298, 125. [14] C. N. R. Rao, A. Govindaraj, Acc. Chem. Res. 2002, 35, 998. [15] C. Dillon, K. M. Jenes, T. A. Bekkedehi, C. H. Kiang, D. S. Bethune, M. J. Heben, Nature 1997, 386, 377. [16] P. Chen, X. Wu, J. Lin, K. L. Tan Science 1999, 285, 91. [17] F. E. Pinkerton, B. G. Wicke, C. H. Olk, G. G. Tibbetts, G. P. Meisner, M. S. Meyer, J. F. Herbst, J. Phys. Chem. B 2000, 104, 9460. [18] B. Liu, Y. Y. Fan, M. Liu, H. T. Cong, H. M. Cheng, M. S. Dresselhaus, Science 1999, 286, 1127. [19] R. Strobel, L. Jorissen, T. Schliermann, V. Trapp, W. Schutz, K. Bohmhammel, G. Wolf, J. Garche, J. Power Sources 1999, 84, 221. [20] M. Hirscher, M. Becher, M. Haluska, F. v. Zeppelin, X. Chen, U. Dettlaff-Weglikowska, S. Roth, J. Alloys Compd. 2003, 356357, 433. [21] A. Chambers, C. Park, R. T. Baker, N. M. Rodriguez, J. Phys. Chem. B 1998, 102, 4253. [22] V. V. Simonyan, J. K. Johnson, J. Alloys Compd. 2002, 330332, 659. [23] F. L. Darkrim, P. Malbrunot, G. P. Tartaglia, Int. J. Hydrogen Energy 2002, 27, 193. [24] R. G. Ding, G. Q. Lu, Z. F. Yan, M. A. Wilson, J. Nanosci. Nanotechnol. 2001, 1, 7. [25] J. E. Fischer, Chem. Innovation 2000, 30, 21. [26] M. Hirscher, M. Becher, J. Nanosci. Nanotechnol. 2003, 3, 3. [27] H.-M. Cheng, Q.-H. Yang, C. Liu, Carbon 2001, 39, 1447.

[28] J. J. Reilly, Z. Phys. Chem. Neue Fol. 1979, 117, 155. [29] L. Schlapbach, I. Anderson, J. P. Burger, in Electronic and Magnetic Properties of Metals and Ceramics Part III, Vol. 3B. (Ed: K. H. J. Buschow), VCH, Weinheim, Germany 1994, pp. 271331. [30] T. P. Perng, J. K. Wu, Mater. Lett. 2003, 57, 3437. [31] W. B. Jung, K. S. Nahm, W. Y. Lee, Int. J. Hydrogen Energy 1990, 15, 641. [32] M. Yamaguchi, E. Akiba, in Electronic and Magnetic Properties of Metals and Ceramics Part II, Vol. 3B (Ed: K. H. J. Buschow), VCH, Weinheim, Germany 1994, pp. 333398. [33] J. S. Benjamin, Sci. Am. 1976, 235, 40. [34] D. R. Maurice, T. H. Courtney, Metall. Trans. A 1990, 21, 289. [35] L. Lue, M. O. Lai, Mechanical Alloying, Kluwer Academic, Boston 1997. [36] P. R. Sony, Mechanical Alloying Fundamentals and Applications, Cambridge International Science Publishing, Cambridge, UK 2000. [37] R. A. Varin, T. Czujko, Mater. Manuf. Process 2002, 17, 129. [38] C. C. Koch, Nanostruct. Mater. 1997, 9, 3. [39] J. Huot, G. Liang, R. Schulz, Appl. Phys. A: Mater. Sci. Process. 2001, 72, 187. [40] P. Dantzer, Mater. Sci. Eng. 2002, A329331, 313. [41] H. Imamura, T. Takahashi, R. Galleguillos, S. Tsuchiya, J. Less-Common Met. 1983, 89, 251. [42] H. Imamura, J. Less-Common Met. 1989, 153, 16. [43] H. Imamura, Y. Takesue, S. Tabata, N. Shigetomi, Y. Sakata, S. Tsuchiya, Chem. Commun. 1999, 2277. [44] H. Imamura, S. Tabata, Y. Takesue, N. Sakata, S. Kamazaki, Int. J. Hydrogen Energy 2000, 25, 837. [45] H. Imamura, Y. Takesue, T. Akimoto, Tabata, J. Alloys Compd. 1999, 293295, 564. [46] H. Imamura, N. Sakasai, T. Fujinaga, J. Alloys Compd. 1997, 253254, 34. [47] H. Imamura, N. Sakasai, Y. Kajii, J. Alloys Compd. 1996, 232, 218. [48] H. Imamura, N. Sakasai, J. Alloys Compd. 1995, 231, 810. [49] G. Liang, J. Huot, S. Boily, A. V. Neste, R. Schutz, J. Alloys Compd. 1999, 292, 247. [50] G. Liang, J. Huot, S. Boily, A. V. Neste, R. Schutz, J. Alloys Compd. 1999, 291, 295. [51] G. Liang, J. Huot, S. Boily, A. V. Neste, R. Schutz, J. Alloys Compd. 2000, 305, 239. [52] S. Bouaricha, J.-P. Dodelet, D. Guay, J. Hout, R. Schutz, J. Alloys Compd. 2001, 325, 245. [53] J. Huot, Nanocrystalline Materials for Hydrogen Storage, in Nanoclusters and Nanocrystals (Ed: H. S. Nalwa), American Scientific Publishers, Stevenson Ranch, CA 2003, Ch. 2. [54] L. I. Zakharkin, V. V. Gavrilenko, Dokl. Akad. Nauk. SSSR 1962, 145, 793. [55] E. C. Ashby, P. Kobetz, Inorg. Chem. 1966, 5, 1615. [56] P. Claudy, B. Bonnetot, J.-P. Bastide, J.-M. Letoffe, Mater. Res. Bull. 1982, 17, 1499. [57] B. Bogdanovic, M. Schwickardi, J. Alloys Compd. 1997, 253254, 1. [58] B. Bogdanovic, R. A. Brand, A. Marjanovic, M. Schwickardi, J. Tolle, J. Alloys Compd. 2000, 302, 36. [59] J. Huot, S. Boily, V. Guther, R. Schulz, J. Alloys Compd. 1999, 304, 383. [60] R. A. Zedan, S. Takara, A. G. Hee, C. M. Jensen, J. Alloys Compd. 1999, 285, 119. [61] C. M. Jensen, R. A. Zedan, N. Mariels, A. G. Hee, C. Hagen, Int. J. Hydrogen Energy 1999, 24, 461. [62] G. Sandrock, K. Gross, G. Thomas, C. Jensen, D. Meeker, S. Takara, J. Alloys Compd. 2002, 330332, 696. [63] B. Bogdanovic, M. Felderhoff, M. Germann, M. Hartel, A. Pommerin, F. Schuth, C. Weidenthaler, B. Zibrowius, J. Alloys Compd. 2003, 350, 246. [64] B. Bogdanovic, M. Felderhoff, S. Kaskel, A. Pommerin, K. Schlichte, F. Schth, Adv. Mater. 2003, 15, 1012. [65] D. L. Anton, J. Alloys Compd. 2003, 356357, 400.

776

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2004, 16, No. 910, May 17

A. M. Seayad, D. M. Antonelli/Hydrogen Storage in Inorganic Nanostructures

REVIEW

[66] A. Zaluska, L. Zaluski, J. O. Strom-Olsen, J. Alloys Compd. 2000, 298, 125. [67] K. J. Gross, G. Sandrock, G. J. Thomas, J. Alloys Compd. 2002, 330332, 691. [68] G. J. Thomas, K. J. Gross, N. Y. C. Yang, C. Jensen, J. Alloys Compd. 2002, 330332, 702. [69] K. J. Gross, G. J. Thomas, C. M. Jensen, J. Alloys Compd. 2002, 330332, 683. [70] P. Vajeeston, P. Ravindran, R. Vidya, H. Fjellvag, A. Kjekshus, Appl. Phys. Lett. 2003, 80, 2257. [71] H. Morioka, K. Kakizaki, S.-C. Chung, A. Yamada, J. Alloys Compd. 2003, 353, 310. [72] M. Fichtner, O. Fuhr, O. Kircher, J. Alloys Compd. 2003, 356357, 418. [73] E. Wiberg, R. Bauer, Z. Naturforsch. 1950, 5b, 397. [74] M. Fichtner, O. Fuhr, J. Alloys Compd. 2002, 345, 286. [75] E. M. Fedneva, V. L. Alpatova, V. I. Mikheeva, Russ. J. Inorg. Chem. 1964, 9, 826. [76] A. Zttel, S. Rentsch, P. Fischer, P. Wenger, P. Sudan, P. Mauron, C. Emmenegger, J. Alloys Compd. 2003, 356357, 515. [77] D. S. Stasinevich, G. A. Egorenko, Russ. J. Inorg. Chem. 1968, 13, 341. [78] A. Zttel, P. Wenger, S. Rentsch, P. Sudan, P. Mauron, C. Emmenegger, J. Power Sources 2003, 118, 1. [79] H. J. Schlesinger, H. C. Brown, J. Am. Chem. Soc. 1940, 62, 3429. [80] H. J. Schlesinger, H. C. Brown, H. R. Hoekstra, L. R. Rapp, J. Am. Chem. Soc. 1953, 75, 199. [81] Y. Wu, R. M. Mohring, Prepr. Symp. - Am. Chem. Soc., Div. Fuel Chem. 2003, 48, 940. [82] C. Weixiang, T. Zhiyan, G. Hetong, Dianchi 1998, 28, 157. [83] R. Aliello, M. A. Matthews, D. L. Reger, J. E. Collins, Int. J. Hydrogen Energy 1998, 23, 1103. [84] a) S. C. Amendola, S. L. Sharp-Goldman, M. S. Janjua, M. T. Kelly, P. J. Petillo, M. A. Binder, J. Power Sources 2000, 85, 186. b) S. C. Amendola, S. L. Sharp-Goldman, M. S. Janjua, M. T. Kelly, P. J. Petillo, M. A. Binder, Int. J. Hydrogen Energy 2000, 25, 969. [85] D. Hua, Y. Hanxi, A. Xinping, C. Chuansin, Int. J. Hydrogen Energy 2003, 28, 1095. [86] R. Ma, Y. Bando, H. Zhu, T. Sato, C. Xu, D. Wu, J. Am. Chem. Soc. 2002, 124, 7672. [87] T. Ishii, T. Sato, Y. Sekikawa, M. Iwata, J. Cryst. Growth 1981, 52, 285. [88] R. Ma, Y. Bando, T. Sato, Adv. Mater. 2002, 14, 366. [89] P. Wang, S. Orimo, T. Matsushima, H. Fujii, G. Majer, Appl. Phys. Lett. 2002, 80, 318. [90] C. Tang, Y. Bando, X. Ding, S. Qi, D. Golberg, J. Am. Chem. Soc. 2002, 124, 14 550. [91] T. Oku, M. Kuno, Diamond Relat. Mater. 2003, 12, 840. [92] J. Chen, S.-L. Li, Z.-L. Tao, Y.-T. Shen, C.-X. Cui, J. Am. Chem. Soc. 2003, 125, 5284.

[93] R. Tenne, L. Margulis, M. Genut, G. Hodes, Nature 1992, 360, 444. [94] M. Hershfinkel, L. A. Gheber, V. Volterra, J. L. Hutchison, R. Tenne, J. Am. Chem. Soc. 1994, 116, 1914. [95] Y. Feldman, E. Wasserman, D. J. Srolovitz, R. Tenne, Science 1995, 267, 222. [96] W. H. Pan, M. E. Leonowicz, E. I. Stiefel, Inorg. Chem. 1983, 22, 672. [97] J. Chen, N. Kuriyama, H. Yuan, H. T. Takeshita, T. Sakai, J. Am. Chem. Soc. 2001, 123, 11 813. [98] J. Chen, S. L. Li, Z. L. Tao, J. Alloys Compd. 2003, 356357, 413. [99] F. W. Dafert, R. Miklauz, Monatsh. Chem. 1910, 31, 981. [100] O. Ruff, H. Goeres, Chem. Ber. 1910, 44, 502. [101] Y. H. Hu, E. Ruckenstein, Ind. Eng. Chem. Res. 2003, 42, 5135. [102] Y. H. Hu, E. Ruckenstein, J. Phys. Chem. A 2003, 107, 9737. [103] P. Chen, Z. Xiong, J. Luo, J. Lin, K. L. Tan, Nature 2002, 420, 302. [104] Power Diffraction File TM Data Sets: 149, International Center for Diffraction Data (ICDD), Pennsylvania, PA 1999. [105] N. L. Rosi, J. Eckert, M. Eddaoudi, D. T. Vodak, J. Kim, M. O'Keeffe, O. M. Yaghi, Science 2003, 300, 1127. [106] M. Eddaoudi, D. B. Moler, H. Li, B. Chen, T. M. Reineke, M. O'Keeffe, O. M. Yaghi, Acc. Chem. Res. 2001, 34, 319. [107] H. Li, M. Eddaoudi, M. O'Keeffe, O. M. Yaghi, Nature 1999, 402, 276. [108] M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O'Keeffe, O. M. Yaghi, Science 2002, 295, 469. [109] X. He and D. M. Antonelli, Angew. Chem. Int. Ed. 2002, 41, 214. [110] J. Y. Ying, C. P. Mehnert, M. S. Wong, Angew. Chem. Int. Ed. 1999, 38, 56. [111] Y. Wang, X. Tang, L. Yin, W. Huang, Y. R. Hacohen, A. Gedanken, Adv. Mater. 2000, 12, 1183. [112] a) M. Vettraino, M. Trudeau, A. Y. H. Lo, R. W. Schurko, D. M. Antonelli J. Am. Chem. Soc. 2002, 124, 9567. b) U. S. Patent Pending. [113] T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Adv. Mater. 1999, 11, 1307. [114] P. Hoyer, Langmuir 1996, 12, 1411. [115] Y. Zhu, H. Li, Y. Koltypin, Y. R. Hacohen, A. Gedanken, Chem. Commun. 2001, 2616. [116] V. V. Pokropivnyi, Powder Metall. Met. Ceram. 2001, 40, 485. [117] V. V. Pokropivnyi, Powder Metall. Met. Ceram. 2001, 40, 582. [118] V. V. Pokropivnyi, Powder Metall. Met. Ceram. 2002, 41, 123. [119] A. L. Ivanovskii, Russ. Chem. Rev. 2002, 71, 175. [120] D. Li, Y. Xia, Nano Lett. 2003, 3, 555. [121] K. Tanaka, Thin Solid Films 1999, 341, 120. [122] S. Kaskel, K. Schlichte, G. Chaplais, M. Khanna, J. Mater. Chem. 2003, 13, 1496. [123] T. K. Tromp, R.-L. Shia, M. Allen, J. M. Eiler, Y. L. Yung, Science 2003, 300, 1740. [124] M. G. Schultz, T. Diehl, G. P. Brasseur, W. Zittel, Science 2003, 302, 624.

______________________

Adv. Mater. 2004, 16, No. 910, May 17

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

777

Вам также может понравиться