You are on page 1of 12

View Article Online / Journal Homepage / Table of Contents for this issue

PERSPECTIVE www.rsc.org/ees | Energy & Environmental Science

New sorbents for hydrogen storage by hydrogen spillover – a review


Lifeng Wang and Ralph T. Yang*
Received 9th May 2008, Accepted 12th June 2008
First published as an Advance Article on the web 24th June 2008
DOI: 10.1039/b807957a

The utilization of hydrogen as an energy source or carrier for fuel-cell powered vehicles is limited by the
lack of a safe and effective hydrogen storage system. Recent advances in the application of hydrogen
spillover for hydrogen storage have provided a promising route for hydrogen storage. An overview of
the progress made on hydrogen storage by spillover at ambient temperature on various adsorbents,
including carbons, metal organic frameworks (MOFs) and other nanostructured materials, is given in
this review. New techniques for facilitating hydrogen spillover that were developed in our laboratory
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

are discussed, along with future directions.

1. Introduction cells, as well as for non-automobile transportation applications


such as motorcycles. The research targets in hydrogen storage
Hydrogen storage is the crucially missing link to a future capacities depend on the specific applications. The most chal-
‘‘hydrogen economy.’’ Hydrogen can be stored in compressed lenging targets are for automobile application.
tanks, in liquefied form, and in compressed tanks filled with Metal hydrides have been studied intensively for over four
sorbent materials. The most promising technique is the sorbent decades.7,8 Studies on adsorbents are more recent. Carbon
approach.1–3 The sorbent approach includes metal hydrides and materials, metal organic frameworks (MOFs) and other nano-
adsorbents.4 For transportation applications, the US Depart- structured and porous materials have received considerable
ment of Energy (DOE) has set 6.5 wt% and 62 kg H2 m3 as the interest due to their high surface areas.9–15 These materials
system capacity targets for on-board hydrogen storage in fuel cell have exhibited promising hydrogen storage capacities at 77 K.
applications in vehicles, at ambient temperature.3,5,6 The pressure However, among the currently available candidate storage
is not specified, but 100 atm has been a nominal pressure for materials, none is capable of meeting the DOE criteria for
research. As a reference, for a compact passenger vehicle pow- personal transportation vehicles at moderate temperatures and
ered by fuel cell, 4 kg H2 is needed for a driving range of 400 km. pressures.16–18 The hydrogen adsorption capacities at the ambient
To store hydrogen in gas tanks at the normal ambient temper- temperature on all known sorbents are below 0.6–0.8 wt% at
ature, 700 atm is needed for a compressed tank (of a reasonable 298 K and 100 atm. This is true for all sorbent materials
size), which is unacceptably high. To store hydrogen in the liquid including the MOFs and templated carbons. In fact, the capac-
form requires large energy consumption for liquefaction at 20 K ities for MOF-177 and MIL-101 are well below 0.7 wt% at 298 K
and it also suffers from the ‘‘boil-off’’ problem. Other examples and 100 atm, although these two materials are claimed to have
for mobile applications are storage for portable electronics such the highest BET surface areas and storage capacities at 77 K.19–24
as laptop computers and cell phones that are powered by fuel Based on experimental results and molecular orbital calcula-
tions, Yang et al. proposed to use the hydrogen ‘‘spillover’’
Department of Chemical Engineering, University of Michigan, Ann Arbor, approach for hydrogen storage at ambient temperature.25,26
Michigan 48109, USA. E-mail: yang@umich.edu; Fax: (+1 734) 764-7453 Hydrogen spillover may be broadly defined as the transport (i.e.,

Lifeng Wang received his BS Ralph T. Yang is Dwight F.


degree in 2002 and PhD degree Benton Professor of Chemical
in 2007 in Chemistry from Jilin Engineering at University of
University, China. From 2007, Michigan. He received his PhD
he has been a postdoctoral fellow from Yale University in 1971.
in the group of Professor Ralph He has published two books:
T. Yang at the University of ‘‘Gas Separation by Adsorption
Michigan. Processes’’ and ‘‘Adsorbents:
Fundamentals and Applica-
tions’’. He has received three
national awards from AIChE,
one ACS National Award,
and he is a member of the U.S.
Lifeng Wang Ralph T: Yang National Academy of Engi-
neering.

268 | Energy Environ. Sci., 2008, 1, 268–279 This journal is ª The Royal Society of Chemistry 2008
View Article Online

via surface diffusion) of dissociated hydrogen adsorbed or errors in the result. For example, in a typical setup, an artefact
formed on a first surface onto another surface. The first surface is of 2.6 wt% adsorption can be caused by a 1  C temperature rise.
typically a metal (that dissociates H2) and the second surface is Calibration was sometimes not done properly. LaNi5 (a known
typically the support on which the metal is doped. Hydrogen metal hydride) was often used for calibration. The hydrogen
spillover is a well documented phenomenon in the catalysis uptake in LaNi5 (to form LaNi5H 6, or 1.37 wt% gain) occurs at
literature, and has been known in the catalysis community for 2 bar at 298 K, in one step. Thus, the high pressure range (2–
over four decades, although it is still not well understood.27,28 100 atm), where the large errors occur, was not calibrated.
Much evidence has been shown in the literature on its roles Helium is typically used for calibration of dead spaces; ignoring
played in catalytic reactions. Very little has been studied on adsorption by He also causes significant errors.49 We have
hydrogen storage by spillover at ambient temperature, although established a calibration procedure using the commercially
it is also known to occur at such temperatures. To exploit spill- available superactivated carbon, AX-21 (with BET surface area
over for storage, among the key questions are whether spillover is 2600–2800 m2 g1), as the calibration material,30,49 which yields
reversible at ambient temperature and if the desorption rates at 0.6 wt% storage at 100 atm and 298 K, with a slightly concave
ambient temperature are fast enough for automotive applica- isotherm from 0–100 atm. The same result has been reported by
tions. several laboratories.50 When all errors were eliminated from the
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

By using the spillover approach, significant amounts of experimental measurements, the hydrogen storage capacities
hydrogen storage capacities at 298 K have been obtained on for all purified carbon nanotubes (both single-walled and multi-
a number of sorbents, including single-walled carbon nanotubes walled) were below 0.3 wt% at 298 K and 100 atm. More
(SWNTs) and multi-walled carbon nanotubes (MWNTs)25,29,30 recently, it was claimed that an exceptionally high hydrogen
and graphite nanofibers,30,31 activated carbons,32,33 MOFs,34,35 storage capacity of 6.9 wt% at 77 K and 20 bars over all known
covalent organic frameworks (COFs),23 zeolites,36 and meso- sorbents was obtained on a microporous carbon (with a BET
porous silica.37 It has been found in our studies that the hydrogen surface area of 3200 m2 g1) that was synthesized by templating
isotherms are reversible at ambient temperature and that the using zeolite beta.51 Our recent work showed that the hydrogen
discharge rates are fast enough to meet the DOE targets on storage capacity in a microporous carbon with an ultra-high
discharge rates. This review is aimed at summarizing the recent BET surface area of >3700 m2 g1 was still less than 1 wt% at
developments in hydrogen storage by spillover, as well as 298 K and 100 atm. On the basis of numerous theoretical studies
a fundamental understanding of the hydrogen spillover and careful experimental validation, several reports have
phenomena for hydrogen storage. The literatures on hydrogen claimed that at ambient temperature the hydrogen storage
storage in carbon, MOFs, metal hydride and other sorbents capacities in pure carbon materials are all well below 1 wt%.52–54
without using hydrogen spillover have been reviewed and However, recent results have shown that hydrogen storage in
discussed by many authors,3,4,7–15 which will not be focused on carbon materials can be significantly enhanced by spillover
in this review. techniques.

2.1.1. Spillover by physical mixing. Hydrogen storage in


2. Hydrogen storage by hydrogen spillover carbon materials can be increased by secondary spillover27 from
a supported catalyst.
2.1. Hydrogen storage on carbon by spillover
Mixing a catalytic material with an otherwise inert secondary
Since the first report on hydrogen storage in carbon nanotubes material has been used to demonstrate hydrogen spillover by
by Dillon et al.,38 carbon materials have attracted considerable Srinivas and Rao.55 They studied the effect of H2 spillover on
attention for hydrogen storage. High hydrogen adsorption benzene hydrogenation over a Pt/C catalyst and a Pt/C catalyst
capacities have been reported in various carbon nano- diluted with carbon. From chemisorption results a direct obser-
structures.39–42 However, the large amounts of hydrogen stored vation of hydrogen spillover at room temperature was made.
in carbon materials were disputed because of the difficulties in Lueking and Yang investigated hydrogen storage on various
obtaining reproducible adsorption capacities from different carbon materials (SWCNTs, MWCNTs, graphite nanofiber,
laboratories.2 This discrepancy in the hydrogen storage abilities activated carbon and carbon fibers) by secondary spillover.30 It
of carbon nanostructures is considered to be due to difficulties was found that a supported catalyst as a hydrogen source
in accurate measurements, impurities in the carbon samples and enhanced the overall hydrogen uptake of the carbon materials:
in the hydrogen gas, and poor understanding of the hydrogen by simple mixing of carbon nanotubes with supported palladium,
sorption mechanism.43–45 The controversy was partly or largely the hydrogen uptake on carbon nanotubes was increased by
caused by the large experimental errors involved in the high- a factor of three. The adsorption of spiltover hydrogen on carbon
pressure measurements of hydrogen storage.4 This is true for was the predominant factor in the magnitude of the overall
both volumetric and gravimetric measurements. The volumetric hydrogen uptake.30
technique is the most commonly used method for measuring Secondary spillover provides a method to clearly delineate the
high pressure isotherms.46 The experimental pitfalls in volu- role of carbon surface functionalities; maintaining the primary
metric measurements for hydrogen storage have been discussed metal–carbon material and varying the secondary carbon
in detail by Tibbetts et al.47 and Darkrim et al.48 A more detailed establishes a constant hydrogen source to the secondary
analysis and discussion on all issues on the accuracies of volu- material. This eliminates several variables inherent to primary
metric measurements at high pressures have been given by us spillover, including doping efficiencies, carbon–metal interface,
recently.49 A very small error in the experiment can cause large and metal content.

This journal is ª The Royal Society of Chemistry 2008 Energy Environ. Sci., 2008, 1, 268–279 | 269
View Article Online

2.1.2. Spillover assisted by carbon bridge building. The


hydrogen storage capacity in nanostructured and porous carbon
materials can be increased by physically mixing the carbon
materials with a supported catalyst. It was suggested that the
improved hydrogen storage capacity was due to the atomic
hydrogen spillover from the supported catalyst and hence
improved contacts between the spillover source and the
secondary receptor would facilitate the hydrogen spillover,
resulting in the enhanced hydrogen capacity.55–57
Conner and Falconer proposed that secondary spillover
requires intimate contact between the two unlike materials and
there may be an energy barrier to transfer hydrogen from one
material to another.28
Lueking and Yang studied the hydrogen storage on MWNTs
with various degree of catalyst removal (Ni, Mg).25 In that study,
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

they observed that removal of the catalyst decreased the uptake


from 0.6 wt% to below detection limits. Normalization by
metal content and temperature-programmed desorption studies
suggested hydrogen dissociation and subsequent spillover to the
MWNT. It was found that metal–support interactions were key
to the spillover.
To improve the contact between the spillover source and the
secondary receptor, a bridge-building technique was proposed by
Yang et al.58 The bridge building technique involves mixing
Fig. 1 Hydrogen spillover in a supported catalyst system: (a) adsorption
the catalyst (a small amount) and the sorbent with a small
of hydrogen on a supported metal particle; (b) the low-capacity receptor;
amount of carbon precursor such as glucose, followed by (c) primary spillover of atomic hydrogen to the support; (d) secondary
a temperature programmed heating protocol to carbonize the spillover to the receptor enhanced by a physical bridge; (e) primary and
precursor. By using this simple technique, carbon ‘‘nanobridges’’ secondary spillover enhancement by improved contacts and bridges.
can be built between a spillover source particle and a secondary (Reprinted with permission from ref. 32. Copyright 2005 American
receptor particle. Chemical Society).
The bridge-building techniques were first applied to two
typical carbon materials of SWNTs and superactivated carbon that the enhancement factor was a weak function of pressure.
(AX-21) due to their relative chemical stabilities and high surface Moreover, reversibility was demonstrated through desorption
areas. In this case, a supported catalyst (Pd/C) served as the and readsorption at 298 K.32
source of hydrogen atoms via dissociation and primary spillover
and AX-21 or SWNTs was the secondary spillover receptor. 2.1.3. Spillover by chemical doping. Hydrogen spillover
Fig. 1 depicts the spillover process in a supported catalyst system. through mixing a hydrogen dissociation source and a receptor is
The spillover source is shown in Fig. 1a, along with an adsorbed an effective method for enhancing the hydrogen storage capacity
hydrogen molecule. Fig. 1b is a schematic of the low-capacity on carbon sorbents. However, the mixing process involves many
receptor, which could be bundles of SWNTs or AX-21. The size factors, such as particle sizes, sample amounts, grinding time
of the hydrogen molecule precludes it from accessing the inter- and intensity. These factors will affect the particle sizes and the
stices in the SWNT bundle or additional micropores of AX-21. contacts between particles. This would result in different
The molecular size may also prevent transport of hydrogen connectivities between the particles, hence leading to poor
through any defects in an individual SWNT surface; thus, the reproducibility in sample and the storage capacity.
adsorption at endohedral sites is dependent upon diffusion to Another way to introduce metal onto the receptor material is
uncapped ends and further transport inside the tube to these by chemical-doping. Chemical doping has been developed
sites. Examples of primary and secondary spillover are shown extensively, particularly in catalysis. Compared with physical
schematically in parts c and d, respectively, of Fig. 1. Dissocia- mixing, chemical doping produces identical samples and is more
tion is assumed to take place on the metal particle, and atomic reproducible. Several studies showed that hydrogen uptake on
hydrogen spills over to the support. The atoms are transported nanostructured carbons (CNTs, active carbon, carbon nano-
to the receptor via diffusion across the bridge and can access fibers, etc.) was enhanced by doping with transition metals (Ni,
additional sites on the receptor. It is noted that the creation of Pd, Pt, etc.), and the maximum uptake could be achieved by
more intimate contacts between the metal particle and the optimizing adsorbents and treatment conditions.59–64
support may also contribute to primary spillover enhancement, Back et al. investigated the hydrogen storage capacities
as depicted in Fig. 1e. of Pd-doped sepiolite-derived carbon nanofibers (SDCNs) at
By using our bridge building technique, the hydrogen different Pd particle sizes and different Pd doping levels.59 It was
adsorption amounts were increased by a factor of 2.9 on the found that the hydrogen storage capacity was dependent on the
AX-21 receptor and 1.6 on the SWNT receptor at 298 K and Pd particle size and doping amount as the Pd particle surface
100 kPa. Similar results were also obtained at 10 MPa, indicating area and the carbon BET surface area changed with them. The

270 | Energy Environ. Sci., 2008, 1, 268–279 This journal is ª The Royal Society of Chemistry 2008
View Article Online

maximum hydrogen uptake of about 0.59 wt% was obtained at capacity of 0.6 wt% was obtained on the undoped AX-21 sample
298 K and 90 bar for the 3 wt% Pd-doped SDCN due to the at 298 K and 10 MPa. By ultrasound assisted doping of Pt, the
smaller Pd particle size and relatively high surface area of carbon. Pt/AX-21 sample had a hydrogen storage capacity of 1.2 wt%
Anson et al. prepared palladium-loaded single-walled carbon under the same conditions, which was enhanced by a factor of 2.
nanotubes and palladium-loaded MAXSORB activated carbon It is worth noting that the hydrogen storage capacity was lower
by reaction of the raw carbon support with Pd2(dba)3$CHCl3.60 on doped samples without using ultrasound. In this work, the
A maximum hydrogen capacity of 0.7 wt% was obtained at 90 doping of AX-21 in an acetone solution of H2PtCl6 was per-
bar in a palladium-loaded MAXSORB sample, while the formed under ultrasonication (at 100 W, 42 kHz) at room
capacity for the raw MAXSORB at the same pressure was temperature for 1 h. The main phenomena responsible for
0.42 wt%. They observed that the H/Pd atomic ratios in the ultrasound actions are cavitation and acoustic streaming, the
palladium-loaded samples were always higher than in the bulk effects of which on adsorption and desorption from solution
Pd. The spillover effect was considered as the cause for the high have been studied.67 The conditions of ultrasonication should be
H/Pd atomic ratios. optimized further for enhanced storage capacity.
Bettahar et al. investigated hydrogen storage on nickel cata- Our recent studies indicated that uniform nanoparticles of Ru,
lysts supported on amorphous activated carbon.61 They found Pt and Ni could be also effectively doped on porous carbon
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

that the hydrogen storage capacity depended on the metal supports by using the ultrasonication assisted doping method.
content and the temperature of pre-treatments.
Among the aforementioned catalysts, the reversible hydrogen 2.1.5. Spillover by plasma assisted doping. From the view-
capacities were all below 1.0 wt% at high pressure and room point of spillover, high dispersion of metals would enhance
temperature. More recently, Kim et al. reported a higher the utilization efficiency of metals for dissociation of hydrogen
hydrogen capacity of 2.8 wt% on Ni doped on multi-walled for hydrogen spillover. Back et al.59 and Yang et al.33 have
carbon nanotubes at a moderate temperature.64 But the observed that smaller particle size and high dispersion of parti-
hydrogen adsorbed on the material could only be released at cles enhanced the storage capacity. Recently, a new reduction
temperatures higher than 340 K (340–520 K). method assisted by plasma has received much attention for
preparing nanosized metal-doped materials.68–72 Plasma appli-
2.1.4. Spillover by ultrasonication assisted doping. Doping of cation in catalyst preparation has been previously reviewed.68
metal nanoparticles is one of the key factors for storage capacity Catalysts produced by this method had uniform metal particle
as well as for catalyst preparation. It still remains, for the most sizes and the particles were highly dispersed on the surface of the
part, as an art. The importance of doping at the point of zero supports. Li et al. doped Pt nanoparticles on Norit activated
charge (PZC, i.e., the pH at which the surface is neutral) has been carbon with post-doping plasma treatment, and investigated its
illustrated by Regalbuto and coworkers.65 Recently, Yang et al. hydrogen storage properties.66,73 It was discovered that plasma
showed that doping under ultrasonication was efficient.66 Yang treatment significantly enhanced the hydrogen storage capacity
and Li doped Pt nanoparticles on a superactivated carbon of the doped sample.
(AX-21) by using ultrasound-assisted impregnation of H2PtCl6 As shown in Fig. 3, the equilibrium hydrogen adsorption
in a solution of acetone.33 The TEM results revealed that the isotherms on Norit activated carbon (NAC), 3 wt% Pt doped
uniform Pt particles of approximately 2 nm were highly dispersed on Norit carbon prepared by hydrogen reduction, and 3 wt%
on the carbon support. As shown in Fig. 2, the hydrogen storage Pt doped on Norit carbon prepared by plasma reduction are

Fig. 2 High-pressure hydrogen isotherms at 298 K for pure AX-21 (-), Fig. 3 High-pressure hydrogen isotherms at 298 K for NAC (,),
and the Pt/AX-21 sample: adsorption (B), and desorption (:). H2-reduced Pt/NAC-H (A), and plasma-treated Pt/NAC-P sample:
(Reprinted with permission from ref. 33. Copyright 2007 American adsorption (B), and desorption (:). (Reprinted with permission from
Chemical Society). ref. 73. Copyright 2007 American Chemical Society).

This journal is ª The Royal Society of Chemistry 2008 Energy Environ. Sci., 2008, 1, 268–279 | 271
View Article Online

compared. The Norit carbon with a BET surface area of about for MOF-177 is around 3000 m2 g1. More recently, Férey et al.
1000 m2 g1 had a hydrogen storage capacity of 0.3 wt% at 298 K reported a nanoporous chromium terephthalate-based material
and 10 MPa. After doping the Norit carbon with 3 wt% Pt by (MIL-101) with the highest Langmuir surface area (4500–5900
hydrogen reduction, the hydrogen storage capacity was increased m2 g1) among all MOFs. It was reported that the hydrogen
by about 50%, due to spillover. For the Pt-doped carbon treated storage capacities in this material at 8 MPa were 6.1 wt% at 77
with plasma, the hydrogen storage capacity was increased by K, and 0.43 wt% at 298 K.19,20 Although these MOFs have
approximately 270% from the value of the plain carbon. From remarkable hydrogen capacities at 77 K, no significant hydrogen
TEM results and thermal stability tests, it was concluded that storage capacities were obtained with the MOFs at room
the Pt particles were highly dispersed and strongly anchored on temperature.23,24
the carbon supports. This large enhancement by plasma treat-
ment was attributed to high dispersion of Pt particles and 2.2.1. Spillover on IRMOFs
chemical bridges that were built between the Pt particles and the 2.2.1.1. Spillover by physical mixing. Significant hydrogen
carbon surface, which enhanced the spillover. storage capacities at ambient temperature on nanostructured
In the preliminary study, glow discharge Ar plasma was used. carbon materials have been achieved by using secondary spill-
The glow discharge was generated by applying 900 V to the over. It is expected that the hydrogen storage on MOFs at room
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

planar electrodes. The temperature of the plasma was near the temperature can also be much improved by hydrogen spillover
ambient temperature. Previous studies suggested that the ambient techniques. The hydrogen spillover was first applied to MOF-5
temperature during plasma treatment was the main reason for the (also known as IRMOF-1) and IRMOF-8, which were con-
high dispersion of metal particles. It is known that plasmas structed by linking tetrahedral [Zn4O]6+ clusters with linear
consist of ionized atoms and molecules as well as energized elec- carboxylates.34
trons. Other plasmas (microwave plasma, plasma spraying, radio For spillover experiments, a commercial catalyst containing
frequency, etc.) and other gases (H2, O2, Air and CO2 etc.) will 5 wt% Pt supported on active carbon was used for dissociation of
give similar or even better results than the glow discharge Ar H2. Active carbon was the primary receptor for hydrogen spill-
plasma. This would further reduce the costs of plasma treatment, over. The catalyst and the secondary spillover receptor (MOF-5
and consequently the costs of sorbents for hydrogen storage. or IRMOF-8) (at a weight ratio of 1 : 9) were ground together to
produce the physical mixture.
As shown in Fig. 4, pure MOF-5 had a hydrogen storage
2.2. Hydrogen storage on MOFs by spillover
capacity of 0.4 wt% at 298 K and 10 MPa. Similar results were
Metal–organic frameworks (MOFs) are porous materials con- also observed by Panella and Hirscher82 and Rowsell et al.75 By
structed by coordinate bonds between multidentate ligands and thoroughly mixing MOF-5 with Pt/AC catalyst, the H2 storage
metal atoms or small metal-containing clusters. MOFs can be capacity of MOF-5 was enhanced to 1.5 wt% at 10 MPa.
generally synthesized via self-assembly from different organic Considering the relatively low H2 uptake of Pt/AC (1.0 wt% at
linkers and metal nodules. Due to the variable building blocks, 10 MPa) and the small amount of Pt/AC in the mixture (10 wt%),
MOFs have very large surface areas, high porosities, uniform the hydrogen uptake on MOF-5 was increased by a factor of 3.3.
and adjustable pore sizes and well-defined hydrogen occupation The enhanced storage capacity was clear evidence for secondary
sites. These features make MOFs promising candidates for spillover of H atoms to MOF-5, as a secondary receptor. It is
hydrogen storage.13,14 noted that there was no apparent saturation value for the
Yaghi et al. first reported a high hydrogen adsorption of 4.5 physical mixture at 10 MPa. Reversibility was assessed by
wt% on Zn4O(bdc)3 (bdc ¼ 1,4-benzenedicarboxylate) at 77 K measuring the desorption branch down to 1 atm. The second
and less than 1 atm, which is known as MOF-5 or IRMOF-1.74 adsorption branch was in complete agreement with the first
The reported hydrogen capacities were lower in their subsequent adsorption branch.
studies.75 The novelty of this work provided a promising
candidate for hydrogen storage. A number of unique MOFs
have been synthesized and evaluated for their hydrogen storage
capacity.76–88 It has become clear that high surface area and pore
volume are important factors for high hydrogen uptakes on
MOFs at 77 K.77–80 The effects of surface area, pore volume and
heat of adsorption on hydrogen uptake in MOFs were recently
discussed by Snurr et al.81 Extensive work was directed toward
the synthesis of MOFs with high surface areas and pore
volumes. A series of isoreticular (meaning having the same
underlying topology) metal organic frameworks, Zn4O(L)3,
were constructed by changing the different linking zinc
oxide clusters with linear carboxylates L, so as to get a high
porosity.75–77 Recently, MOF-177 (Zn4O(BTB)2) was formed by
linking the same clusters with a trigonal carboxylate. MOF-177 Fig. 4 High-pressure hydrogen isotherms at 298 K for MOF-5. Dotted
was claimed to have a high Langmuir surface area of 5640 m2 line is prediction based on the weighted average of the mixture.
g1, and the highest hydrogen storage of 7.5 wt% H2 at 77 K and (Reprinted with permission from ref. 34. Copyright 2006 American
70 bar.21,22 The more meaningful surface area, BET surface area, Chemical Society).

272 | Energy Environ. Sci., 2008, 1, 268–279 This journal is ª The Royal Society of Chemistry 2008
View Article Online

Similarly remarkable enhancement of H2 storage capacity by


spillover was also observed on IRMOF-8. The H2 uptake on
pure IRMOF-8 was 0.5 wt% at 298 K and 10 MPa. By physically
mixing IRMOF-8 with Pt/AC, the hydrogen uptake on IRMOF-
8 was increased to 1.8 wt% under the same conditions. This
uptake is equivalent to 8.3 H2 (or 16.6 H atoms) per formula unit,
which is well below the saturation of IRMOF-8.74 Furthermore,
the adsorption was also totally reversible and did not show any
saturation up to 10 MPa, similar to results on MOF-5.

2.2.1.2. Spillover by bridge-building techniques. Secondary


spillover requires intimate contacts between the two unlike
materials because there exist tremendous physical/energy
barriers for surface diffusion of hydrogen atoms from one
material to another. It has been demonstrated that bridge-
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

building technique could facilitate secondary spillover and hence


increase the storage capacities in nanostructured carbon mate-
rials. Yang and Li further extended the bridge-building technique
to MOFs.35
This work was first focused on IRMOF-1 and IRMOF-8,
whose hydrogen uptakes had been proved to be enhanced via
secondary spillover induced by physical mixing. In previous
work, it was found that carbon bridges could be built between an Fig. 5 (a) Primary spillover of atomic hydrogen from Pt metal to the
activated carbon and a Pd/AC catalyst by carbonizing an added activated carbon support and secondary spillover to the MOF receptor
glucose precursor at 673 K.32,58 It is expected that carbon bridges that has limited contacts with the support. (b) Facilitated primary and
can also be formed by carbonization of a hydrocarbon precursor secondary spillover by using carbon bridges (dark black zone).
that was previously introduced into a physical mixture of the (Reprinted with permission from ref. 35. Copyright 2006 American
Chemical Society).
MOF and catalyst. However, it is difficult to build carbon
bridges in MOFs by using glucose precursor because MOFs are
thermally unstable at temperatures >573 K.74 Yang et al. used
sucrose instead of glucose as the carbon precursor. The ternary
physical mixture of sucrose, Pt/AC catalyst, and MOF was first
heated in He to 473 K (melting point of sucrose ¼ 463 K) for 3 h
to allow the sucrose to melt thoroughly and to fill the interstices
(by capillarity) between the particles of the catalyst and the MOF
(Fig. 5). Then the temperature was finally increased to 523 K to
carbonize the sucrose. This temperature program was effective
for building carbon bridges and avoiding the collapse of the
MOF structure. Thermogravimetric analysis (TGA) confirmed
the complete carbonization of sucrose by this heating protocol.
X-Ray diffraction (XRD) and N2 adsorption results indicated
that the structures of MOF materials were well retained during
the bridge-building treatment.
By building carbon bridges, the bridged sample exhibited
a hydrogen uptake of 3 wt% at 10 MPa. It represented an
enhancement factor of 2 compared with the physical mixture
without bridges. This enhancement indicated the built bridges Fig. 6 High-pressure hydrogen isotherms at 298 K for pure IRMOF-8
further facilitated secondary spillover of hydrogen atoms from (,), Pt/AC and IRMOF-8 physical mixture (1 : 9 weight ratio; A) and
the catalyst to the surface of the MOF material. Compared with for a bridged sample of Pt/AC-bridges-IRMOF-8: first adsorption (B),
desorption (:), and second adsorption (>). (Reprinted with permission
pure IRMOF-1, the hydrogen adsorption amount of bridged
from ref. 35. Copyright 2006 American Chemical Society).
IRMOF-1 has been enhanced by a factor of 7.5. These results
indicated that the creation of carbon bridges was crucially
important for achieving a high hydrogen adsorption capacity (Zn4O(C12H6O4)3), which is much lower than the total number of
by spillover. atoms in IRMOF-8, assuming 1 H per surface atom. Further-
Similar enhancement of hydrogen storage by spillover was also more, the ab initio molecular orbital calculation results indicated
observed on IRMOF-8. As shown in Fig. 6, by using the bridge- that it was favorable for one surface atom (such as oxygen) to
building technique, the H2 uptake was increased from 0.5 wt% adsorb two H atoms at the same time. Thus, the theoretical
on pure IRMOF-8 to 4 wt% at 10 MPa on bridged IRMOF-8. maximum storage would be at least 6.5 wt%. No apparent
This uptake amount is equivalent to 34 H per formula unit saturation value was approached for the bridged sample; as the

This journal is ª The Royal Society of Chemistry 2008 Energy Environ. Sci., 2008, 1, 268–279 | 273
View Article Online

isotherm was linear even at 10 MPa. The absence of a saturation were synthesized in the presence of H2O, so their structures are
value suggests that a further increase in capacity can be expected stable upon water adsorption. HKUST-1 is a copper benzene-
at pressures >10 MPa; for example, 6 wt% storage is expected at tricarboxylate porous material reported by Williams et al.89
15 MPa, a viable pressure for practical automotive applications. Recent research showed that it was a good adsorbent for H2 and
NO storage.90 The high gas capacities in HKUST-1 were due to
2.2.2. Spillover on MOF-177. Another interesting example is the presence of a metal site in the walls of the material that could
the storage capacity of MOF-177 by spillover at ambient interact strongly with gas molecules.91 MIL-101 was synthesized
temperature, which was expected to be promising, because from benzene-1,4-dicarboxylate and trimetric chromium(III)
MOF-177 was reported to have the highest hydrogen storage octahedral clusters by Férey’s group. It was reported to have the
(among all MOFs) of 7.5 wt% H2 at 77 K and 70 bar.21,22 Pure largest surface area (Langmuir surface area, 4500–5900 m2 g1)
MOF-177 showed a hydrogen storage capacity of only 0.62 wt% among all MOFs.19 A high hydrogen storage capacity of 3.75
at 298 K and 10 MPa. This value is among the highest hydrogen wt% was obtained on MIL-101 at 77 K and 2 MPa first, then it
storage capacities reported in MOFs at room temperature. By was reported that the hydrogen storage capacity in this material
physical mixing of MOF-177 with Pt/AC catalyst, the hydrogen at 8 MPa was 6.1 wt% at 77 K, and 0.43 wt% at 298 K.20
uptake on MOF-177 at 10 MPa was enhanced to 0.8 wt%. After Yang and Li investigated the hydrogen capacities of MIL-101
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

building bridges, the hydrogen uptake at 10 MPa on MOF-177 and HKUST (Fig. 7 and 8).23 The hydrogen uptakes on the pure
was increased to 1.5 wt%, representing an enhancement factor of MIL-101 and HKUST are 0.51 wt% and 0.35 wt%, respectively.
2.5 as compared with pure MOF-177. The enhancement was By simply mixing with a small amount of Pt/AC catalyst (at 9 : 1
lower than that on IRMOF-1 and IRMOF-8 samples under the mass ratio), their hydrogen uptakes were enhanced to 0.9 wt%
same conditions.35 This was due to the larger crystals of MOF- and 0.7 wt% up to 10 MPa. By building carbon bridges between
177 (hundreds of microns) than those of IRMOF-1 and IRMOF- the MOFs and the Pt/AC catalyst, the hydrogen capacities were
8 (500 nm). The mismatch in the sizes resulted in poor further enhanced to 1.5 wt% on bridged-MIL-101 and 1.1 wt%
connectivity between the hydrogen atom source (Pt/AC) and the on bridged-HKUST. There was no apparent saturation for the
MOF-177 receptor and consequently low hydrogen uptakes.24 bridged samples and the adsorption in the bridged samples was
Intimate contacts between the hydrogen dissociation source and fully reversible at 298 K, as found previously on the bridged
the receptor are required for facilitating the spillover of H atoms. IRMOFs.
Our recent studies also indicated that reducing crystal size of
MOF-177 helped to improve its hydrogen storage capacity.
The lower heat of adsorption on MOF-177 (Table 1) is another
reason for the lower enhancement. Although MOF-177,
IRMOF-1 and IRMOF-8 have the same metal clusters (Zn4O),
the metal content in MOF-177 is lower than those of IRMOF-1
and IRMOF-8, which leads to the lower heat of adsorption on
MOF-177. The effects of heats of adsorption on hydrogen
storage will be discussed in detail in section 2.3.2.

2.2.3. Spillover on moisture-stable MOFs. Most MOFs such


as aforementioned IRMOFs and MOF-177 are unstable in air
because the moisture would cause decomposition of frameworks.
This always leads to the poor reproducibility and storage
capacity in the MOFs samples. Much care must be taken to avoid
the collapse of MOFs. The stability of MOFs is a key factor in
their practical application, because moisture exposure is inevi-
table in practical applications. Stable MOFs with high hydrogen
capacities at room temperature are desirable. Fig. 7 High-pressure hydrogen adsorption at 298 K for the MIL-101
Recently, two kinds of new MOFs, MIL-10119 and HKUST- samples. A: pure MIL-101; B: Pt/AC and MIL-101 physical mixture;
1,89 have received more attention due to their relative stabilities in :: MIL-101–bridges–Pt/AC. (Reprinted with permission from ref. 23.
air when compared with other MOFs. HKUST-1 and MIL-101 Copyright 2008 Wiley).

Table 1 Comparison of MOFs, pure and bridged to Pt/AC catalyst, on their H2 storage properties. Bridged sample: 10 wt% Pt/AC (containing 5 % Pt),
10 wt% carbon bridges, and 80 wt% MOF

BET surface H2 at 77 K, H2 at 298 K, Bridged sample, H2 at


area/m2 g1 1 atm (wt%) 100 atm (wt%) 298 K, 100 atm (wt%) DH (bridged)/kJ mol1

IRMOF-8 548 1.4 0.4 2.2–4.0 21


COF-1 628 1.1 0.3 0.7 7
HKUST-1 1296 2.2 0.3 1.1 9
MIL-101 2930 1.8 0.5 1.5 13
MOF-177 3100 1.5 0.6 1.5 10

274 | Energy Environ. Sci., 2008, 1, 268–279 This journal is ª The Royal Society of Chemistry 2008
View Article Online

the overall heat of adsorption thus obtained and the binding


energy between the spiltover H and the MOF sites is not known,
but they are obviously related; i.e., a larger heat of adsorption
indicates a stronger binding energy. The DH values (for conve-
nience, referring to the absolute values) are somewhat related to
the contents of the metal clusters in the MOF. For example, COF
contains no metal, and shows the lowest DH. IRMOF-8 and
MOF-177 have the same metal clusters (Zn4O), and are directly
comparable. The metal content in IRMOF-8 is higher than that
in MOF-177, and the DH value is higher for IRMOF-8.
For pure MOFs, not only are surface area and pore volume
important factors for hydrogen storage capacity, framework
composition is another important factor. This is because inter-
actions other than the van der Waals interactions exist between
H2 and MOF. Electric charges (on metal oxide sites) play
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

a significant role in these systems.88 The effects of different


factors on hydrogen uptake in pure MOFs have been discussed
Fig. 8 High-pressure hydrogen adsorption at 298 K for the HKUST-1 by many authors.77–81
samples. A: pure HKUST-1; B: Pt/AC and HKUST-1 physical mixture; For the bridged samples, IRMOF-8 has the highest storage
:: HKUST-1–bridges–Pt/AC. (Reprinted with permission from ref. 23. capacity because of the large DH value; whereas MIL-101 and
Copyright 2008 Wiley).
MOF-177 also show high storage capacities, due to their large
BET surface area. Thus, it is clear that both surface area and the
Recently, Liu et al. studied the hydrogen storage properties DH values are important for storage by spillover.
of doped (physical mixing with Pt/C) and bridged MIL-101 and
MIL-53 samples under mild conditions.92 Both doped and 2.3.3. Spillover on zeolites. Zeolite is a type of microporous
bridged MIL-101 and MIL-53 samples showed enhanced solid used commercially in catalysis and gas separation.4,95
hydrogen storage capacities compared with pristine samples. The Recently, various types of zeolites have been investigated for
gravimetric storage capacities of bridged MIL-101 and MIL-53 hydrogen storage due to their high thermal stability, low cost and
samples were 1.14 and 0.63 wt% at 293 K and 5 MPa. adjustable composition.96–104 Zecchina et al. obtained a hydrogen
capacity of 1.28 wt% on H-SSZ-13 zeolite at 77 K and 0.92 bar.103
Langmi et al. reported a high gravimetric storage capacity of
2.3. Hydrogen storage on other adsorbents by spillover
2.19 wt% on Ca exchanged X zeolite at cryogenic temperatures
2.3.1. Spillover on covalent organic frameworks (COFs). and 15 bar.99 However, hydrogen adsorption on zeolites at room
Covalent organic frameworks (COFs) are constructed from light temperature is low. Kayiran and Darkrim105 and Weitkamp et al.
elements (H, B, C, N, and O) that form strong covalent reported that hydrogen capacities on zeolites were less than
bonds.93,94 Compared with MOFs, they have light weights and 0.5 wt% at room temperature and 60 bar.106
good thermal stabilities (up to 600  C). These features could be Recent studies have shown that the amount of hydrogen
suitable for gas storage. Recently, COF-1 and COF-5 were adsorbed on zeolites depended on the framework structure,
synthesized by condensation reactions of phenyl diboronic acid composition, and acid–base nature of the zeolites.96–107 The
{C6H4[B(OH)2]2} and hexahydroxytriphenylene [C18H6(OH)6]. cations in the zeolite create strong electric fields (and field
Yang et al. investigated the hydrogen capacity of COF-1.23 gradients) that will favor gas adsorption.108 It is expected that
The hydrogen uptake on pure COF-1 was 1.28 wt% at 77 K and 1 high H2 storage capacities could be obtained on zeolites with
atm and 0.26 wt% at 298 K and 10 MPa. By using the bridge- high cation densities.
building technique, the bridged COF-1 sample had a hydrogen Yang et al. reported that low-silica (thus, high charge density)
storage capacity of 0.7 wt%, which was enhanced by a factor of type X (LSX) zeolite (Si/Al ¼ 1) had a hydrogen storage capacity
2.6. The relatively low hydrogen capacity was due to the lower of 0.6 wt% at 298 K and 10 MPa.36 By hydrogen spillover,
surface area of COF-1. bridged Li–LSX had a hydrogen storage capacity of 1.6 wt% at
298 K and 10 MPa, which was enhanced by a factor of 2.6
2.3.2. Comparison of MOFs and COFs. The main hydrogen compared with that on pure Li–LSX. The storage capacity of
storage characteristics of the MOFs and COFs that were studied bridged Li–LSX is among the highest for zeolite materials at
in our laboratory can now be compared. Pure MOFs, COFs ambient temperature. Li–LSX zeolite was selected as the
and their counterparts bridged with 10 wt% Pt/AC catalyst (i.e., 5 hydrogen adsorbent because it had the largest number of charge-
wt% Pt/activated carbon) are compared.23,24,35 The results are compensating cations per unit cell among all faujasites.
summarized in Table 1.
The overall heats of adsorption (DH) shown in the table were 2.3.4. Spillover on mesoporous silica. Compared with micro-
obtained from the temperature dependence of the isotherms, i.e., porous zeolite materials, mesoporous silica has high surface area,
by using the Clausius–Clapeyron equation. The isosteric heats of large pore sizes and amorphous pore walls. These features of
adsorption declined with loading; the values shown in the table mesoporous silica facilitate the loading of nanoparticles in the
were those after leveling off. The detailed relationship between pore channels or tailoring of composition of pore walls.109,110

This journal is ª The Royal Society of Chemistry 2008 Energy Environ. Sci., 2008, 1, 268–279 | 275
View Article Online

Ramachandran et al. evaluated the feasibility of Al–MCM-41


for hydrogen storage.37 The hydrogen storage properties of
Al–MCM-41 with varying contents of aluminium and different
metal oxide doped Al–MCM-41 were compared. Their gas chro-
matographic studies indicated that hydrogen uptake in Al–MCM-
41 samples was strongly dependent on the density of Brønsted acid
sites and the amount of metal oxide. Hydrogen spillover was
mainly responsible for the enhanced hydrogen adsorption for
the doped Al–MCM-41. The studies indicate the potential of
mesoporous materials as good candidates to form and stabilize
nanomaterials within their pores. Mesoporous materials could
also be exploited for better adsorption properties by producing
mono-, bi-, and tri-metallic alloys for hydrogen storage.

3. Mechanistic and kinetics studies of hydrogen


Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

storage by spillover
3.1. Evidence of hydrogen spillover

Although hydrogen storage by spillover is a rather new field, the


phenomenon of ‘‘hydrogen spillover’’ has been known for a long
time.111 Atomic hydrogen spillover was first observed during
studies of ethylene hydrogenation via heterogeneous catalysis
by Sinfelt and Lucchesi.112 Further evidence was observed by
Khoobiar.113 They found the formation of the hydrogen bronze
HxWO3 from WO3 at room temperature after WO3 was exposed
to hydrogen in the presence of a platinum-containing catalyst.
The experimental result could only be interpreted by the diffu-
sion of atomic hydrogen from the platinum catalyst to WO3.
During the last four decades, hydrogen spillover has been
extensively studied particularly in the field of catalysis. The
voluminous literature on atomic hydrogen spillover has been
reviewed.27,28,114 Recently, detailed structural characterization of
the hydrogen-adsorbed species has added insight into the
mechanism of hydrogen adsorption. Previously, Schwarz inves-
tigated the hydrogen adsorption on an activated carbon mixed
with transition metal from room temperature to 77 K, and Fig. 9 Rates of adsorption (a) and desorption (b) during stepped
observed that the spillover effects were strongest at 77 K and increases and decreases in pressure, at 298 K, for bridged MIL-101
nearly nil at room temperature.115 This observation was not sample.
correct.116–122 Direct evidence for spillover of atomic hydrogen at
room temperature has been reported. Inelastic neutron scattering
In the kinetics study, the pressure was increased or decreased
studies have shown the atomic hydrogen spillover from Pt to
in steps, all at 298 K. Thus, the pressure was increased from 0, to
carbon at room temperature.116,117 Infrared spectroscopic study
5, 20, 40, 60, 80 and 100 atm (approximate values) during
provided evidence of spillover of atomic hydrogen from Au
adsorption, and then the steps were reversed in desorption. At
nanoparticles to TiO2 also at room temperature.118 Evidence has
each end pressure, the percent completion with respect to the
also been shown by deuterium isotope tracer studies at room
final (equilibrium) value was recorded. These are shown in Fig. 9.
temperature on Pt/TiO2.119,120 The temperature programmed
It is clear that the rates depended on the loading; the rates were
desorption (TPD) technique was extended to study the spillover
slower at higher loading. This is another manifestation of spill-
on Pt/Al2O3 by Chen and White.121 A number of theoretical
over and that the surface diffusion step was the rate-limiting step,
studies have been done on the mechanism and energetics of
as the distance for surface diffusion increased with loading.
spillover. The more recent DFT study by Chen et al. illustrated
facile pathways for spillover from a Pt particle onto a graphene
basal plane via ‘‘physisorbed H atoms.’’122 3.3. Mechanistic isotherm model for spillover storage
The mechanism of hydrogen spillover has also been investi-
3.2. Kinetics of adsorption and desorption
gated.123–134 It was found that hydrogen molecules were dissoci-
The kinetics of adsorption and desorption via spillover storage ated rapidly on metal sites and then diffused slowly away from
have been studied for both metal-doped carbons and catalyst- them to the receptor. Surface diffusion of hydrogen atoms has
bridged MOFs in our laboratory. Our recent kinetics studies on been proposed to be the rate-determining step in hydrogen
bridged MIL-101 are shown in Fig. 9, as a representative result. spillover.

276 | Energy Environ. Sci., 2008, 1, 268–279 This journal is ª The Royal Society of Chemistry 2008
View Article Online

Hattori et al. investigated the kinetics of hydrogen adsorption on in principle, be determined independently.134 This equation can
SO42–ZrO2, and WO3–ZrO2 supported platinum catalysts in the explain the nearly linear behavior of the isotherms that have been
temperature range 323–573 K.132,133 They observed very slow observed for all spillover sorbent systems investigated so far.
uptake rates of hydrogen on these two supported platinum samples. More recently, Lachawiec and Yang used a deuterium isotope
Yang et al. studied the mechanism of hydrogen adsorption on sequential dosing and TPD technique to study the mechanism
bridged IRMOF-8 samples at room temperature. A two- and kinetics of spillover of atomic hydrogen on Pt/AX-21 and
dimension sectional plot of the spherical model was proposed for IRMOF-8 at room temperature.138 The sorbent sample was
evaluating the diffusion time constants from the uptake rates, as dosed sequentially with H2 and D2 and the sample was then
shown in Fig. 10.134 In this model, H2 is dissociated on the surface quickly ‘‘frozen.’’ Temperature programmed desorption fol-
of the Pt particle, where equilibrium is maintained. The atomic lowed. TPD product distributions and HD formation showed
hydrogen undergoes surface diffusion onto the surface of the that desorption of the isotopes followed the reverse sequence of
activated carbon particle, with radius R1, followed by diffusion that of dosing, indicating reverse spillover during desorption.
onto the surface of the IRMOF-8 receptor. The average sphere of The HD was clearly formed by scrambling of the surface spilt-
diffusion for the IRMOF-8 receptor is taken as R2, which over H and D atoms.
depends on the connectivity through bridging as well as the ratio
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

of Pt/AC catalyst over receptor. Based on this model, it was


found that the diffusion time constant, D/R2, decreased sharply 4. Summary and outlook
with surface concentration. This result clearly indicates that the Significant hydrogen storage capacity has been obtained at
diffusion distance, R2, was increasing with surface concentration. ambient temperature by spillover. Further enhanced hydrogen
The concentration dependence for surface diffusion has been capacity and understanding of hydrogen spillover are still needed
discussed in detail and explained satisfactorily by models by
for meeting the DOE targets. Several aspects should be further
Higashi et al.135 and Yang et al.46,136,137 Surface diffusion is the
developed to improve the hydrogen capacity by spillover.
result of a hopping process, for both physical adsorption and
chemisorption of small molecules or atoms. Within the mono-
layer coverage, the surface diffusivity usually increases with 4.1. Heats of adsorption
surface concentration.
The mechanistic model was further extended to explain the Hydrogen storage by spillover is an atomic hydrogen adsorption
hydrogen adsorption on Pt/AX-21 systems.33 It was also process. The surface adsorption sites of the adsorbent determine
observed that the diffusion time constant (D/R2) decreased with the storage capacity by spillover. It is understood that an
surface concentration. The proposed mechanistic model for adsorbent with more and stronger surface adsorption sites for
hydrogen spillover could interpret the experimental results. hydrogen atoms should have higher hydrogen capacity by
Based on this mechanistic model, an equilibrium isotherm for spillover. Our recent work on a series of MOFs showed that the
the spillover system was derived by Yang and co-workers, and hydrogen uptakes were mainly related with the heats of
the isotherm was given as in eqn (1):134 adsorption of hydrogen on these adsorbents. Higher hydrogen
pffiffiffiffiffiffiffiffi uptakes were obtained on MOFs with higher heats of adsorption
K1 k1 PH2 (stronger surface adsorption sites) even some MOFs had rela-
q¼ pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi (1)
1 þ k2 PH2  K 0 k1 PH2 tively lower surface areas. For example, IRMOF-8 has the
highest storage capacity among the MOFs studied in our work. It
has a high heat of adsorption (20 kJ mol1) and a relative low
Here, q is the equilibrium amount adsorbed at hydrogen
BET surface area (about 600 m2 g1). The heat of adsorption for
pressure PH2, and the constants were defined in ref. 134 and can,
hydrogen on the adsorbent is the key factor in the hydrogen
adsorption by spillover.

4.2. Pore volume and surface area

It was suggested that the hydrogen uptake on an adsorbent at


77 K and high pressures was related to its pore volume. This is
easily understood because at 77 K and high pressures, the
hydrogen storage in an adsorbent is mainly by condensation of
H2 molecules into its porosity, while adsorption by spillover is by
the interaction of hydrogen atoms with the surface sites of the
adsorbent. Therefore, the surface area and not the pore volume
of the adsorbent should be an important factor in hydrogen
adsorption by spillover. An adsorbent with high surface area
Fig. 10 Two-dimension sectional plot of spherical model for hydrogen
would provide more adsorption sites for hydrogen adsorption via
spillover on bridged IRMOF-8 sample, including Pt/AC catalyst (AC: spillover. For example, Pt was doped on carbons with different
activated carbon support) bridged with a receptor sorbent. (Reprinted surface areas. The highest hydrogen uptake was obtained on Pt/
with permission from ref. 134. Copyright 2007 American Chemical templated carbon with the highest surface area among the
Society). carbons studied.

This journal is ª The Royal Society of Chemistry 2008 Energy Environ. Sci., 2008, 1, 268–279 | 277
View Article Online

4.3. Contacts between the source and the receptor 25 A. D. Lueking and R. T. Yang, J. Catal., 2002, 206, 165.
26 F. H. Yang and R. T. Yang, Carbon, 2002, 40, 437.
Contacts between the source and the receptor are very important 27 A. J. Robell, E. V. Ballou and M. Boudart, J. Phys. Chem., 1964, 68,
for hydrogen spillover. Enhanced storage capacity has been 2748.
28 W. C. Conner, Jr. and J. L. Falconer, Chem. Rev., 1995, 95, 759.
achieved by several methods such as bridge building, plasma 29 F. H. Yang, A. J. Lachawiec and R. T. Yang, J. Phys. Chem. B, 2006,
assisted doping and ultrasound assisted building. However, the 110, 6236.
contact effect is still least understood in spillover. It is suggested 30 A. D. Lueking and R. T. Yang, Appl. Catal., A, 2004, 265, 259.
that the enhanced contacts between the source and receptor 31 A. D. Lueking, R. T. Yang, N. M. Rodriguez and R. T. K. Baker,
Langmuir, 2004, 20, 714.
would facilitate the spillover and hence increase the storage 32 A. J. Lachawiec, G. Qi and R. T. Yang, Langmuir, 2005, 21, 11418.
capacity. Significant hydrogen uptakes could be expected by 33 Y. Li and R. T. Yang, J. Phys. Chem. C, 2007, 111, 11086.
enhancing contact strength between the source and receptor, 34 Y. Li and R. T. Yang, J. Am. Chem. Soc., 2006, 128, 726.
35 Y. Li and R. T. Yang, J. Am. Chem. Soc., 2006, 128, 8136.
creating more contact sites and option of the ratio of source and 36 Y. Li and R. T. Yang, J. Phys. Chem. B, 2006, 110, 17175.
receptor. 37 S. Ramachandran, J.-H. Ha and D. K. Kim, Catal. Commun., 2007,
8, 1934.
38 A. C. Dillon, K. M. Johns, T. A. Bekkedahl, C. H. Klang,
Acknowledgements D. S. Bethune and M. J. Heben, Nature, 1997, 386, 377.
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

39 C. Liu, Y. Y. Fan, M. Liu, H. T. Cong, H. M. Cheng and


The authors acknowledge the funding provided by the US M. S. Dresselhaus, Science, 1999, 286, 1127; P. Chen, X. Wu,
Department of Energy’s Office of Energy Efficiency and J. Lin and K. L. Tan, Science, 1999, 285, 91.
Renewable Energy within the Hydrogen Sorption Center of 40 A. Chambers, C. Park, R. T. K. Baker and N. M. Rodriguez,
J. Phys. Chem. B, 1998, 102, 4253.
Excellence (HS CoE), and funding from NSF. 41 Y. Ye, C. C. Ahn, C. Witham, B. Fultz, J. Liu, A. G. Rinzler,
D. Colbert, K. A. Smith and R. E. Smalley, Appl. Phys. Lett.,
References 1999, 74, 2307.
42 R. Zacharia, S.-u. Rather, S. W. Hwang and K. S. Nahm, Chem.
1 A. C. Dillon and M. J. Heben, Appl. Phys. A, 2001, 72, 133. Phys. Lett., 2007, 434, 286; C.-H. Chen and C.-C. Huang,
2 L. Schlapbach and A. Zuttel, Nature, 2001, 414, 353. Microporous Mesoporous Mater., 2008, 109, 549; B.-J. Kim,
3 S. Satyapal, J. Petrovic, C. Read, G. Thomas and G. Ordaz, Catal. Y.-S. Lee and S.-J. Park, J. Colloid Interface Sci., 2008, 318, 530.
Today, 2007, 120, 246. 43 A. Zuttel, Mater. Today, 2003, 6, 24.
4 R. T. Yang, Adsorbents: Fundamentals and Applications, Wiley, New 44 R. T. Yang, Carbon, 2000, 38, 623.
York, 2003. 45 M. Becher, M. Haluska, M. Hirscher, A. Quintel, V. Skakalova,
5 Department of Energy (DOE), A Multiyear Plan for the Hydrogen U. Dettlaff-Weglikovska, X. Chen, M. Hulman, Y. Choi, S. Roth,
R&D Program, US Department of Energy (Office of Power V. Meregalli, M. Parrinello, R. Strobel, L. Jorissen,
Delivery, Office of Power Technologies, Energy Efficiency and M. M. Kappes, J. Fink, A. Zuttel, I. Stepanek and P. Bernier, C.
Renewable Energy), August, 1999. R. Phys., 2003, 4, 1055.
6 S. Hynek, W. Fuller and J. Bentley, Int. J. Hydrogen Energy, 1997, 46 R. T. Yang, Gas Separation by Adsorption Processes, Butterworth,
22, 601. Boston and London, 1987. Paperback edition, Imperial College
7 G. Sandrock, S. Suda and L. Schlapbach, Hydrogen in Intermetallic Press, London, 1997.
Compounds II, Topics in Applied Physics, Springer-Verlag, 1992, vol. 47 G. G. Tibbetts, G. P. Meisner and C. H. Olk, Carbon, 2001, 39, 2291.
67((5)), pp. 197; G. Sandrock and Y. Yurum, Hydrogen Energy 48 F. Darkrim, P. Malbrunot and G. P. Tartaglia, Int. J. Hydrogen
Systems: Production and Utilization of Hydrogen and Future Energy, 2002, 27, 193.
Aspects, NATO ASI Series, Kluwer Academic Publishers, 1994, 49 A. J. Lachawiec, T. R. DiRamondo and R. T. Yang, Rev. Sci.
pp. 253. Instrum., 2008, 79, 063906.
8 S.-i. Orimo, Y. Nakamori, J. R. Eliseo, A. Zuttel and C. M. Jensen, 50 E. Poirier, R. Chahine and T. K. Bose, Int. J. Hydrogen Energy,
Chem. Rev., 2007, 107, 4111. 2001, 26, 831; M. K. Haas, J. M. Zielinski, G. Dantsin, C. G. Coe,
9 R. Q. Snurr, J. T. Hupp and S. T. Nguyen, AIChE J., 2004, 50, 1090. G. P. Pez and A. C. Cooper, J. Mater. Res., 2005, 20, 3214;
10 M. Dinca, A. Dailly, Y. Lin, C. M. Brown, D. A. Newmann and L. Zhou, Y. Sun and Y. Zhou, Chem. Eng. Commun., 2006, 193, 564.
J. R. Long, J. Am. Chem. Soc., 2006, 128, 16876. 51 Z. Yang, Y. Xia and R. Mokaya, J. Am. Chem. Soc., 2007, 129, 1673.
11 A. M. Seayad and D. M. Antonelli, Adv. Mater., 2004, 16, 765. 52 B. Panella, M. Hirscher and S. Roth, Carbon, 2005, 43, 2209;
12 F. Lamari Darkrim, P. Malbrunot and G. P. Tartaglia, Int. R. Dash, J. Chmiola, G. Yushin, Y. Gogotsi, G. Landisio,
J. Hydrogen Energy, 2002, 27, 193. J. Singer, J. Fischer and S. Kucheyev, Carbon, 2006, 44, 2489.
13 J. Rowsell and O. M. Yaghi, Angew. Chem., Int. Ed., 2005, 44, 4670. 53 M. Shiraishi, T. Takenobu, H. Kataura and M. Ata, Appl. Phys. A,
14 D. J. Collins and H.-C. Zhou, J. Mater. Chem., 2007, 30, 3154. 2004, 78, 947; E. Poirier, R. Chahine, P. Bénard, D. Cossement,
15 R. Ströbel, J. Garche, P. T. Moseley, L. Jörisen and G. Wolf, J. Power L. Lafi, E. Mélançon, T. K. Bose and S. Désilets, Appl. Phys. A,
Sources, 2006, 159, 781; P. Kowalczyk, R. Holyst, M. Terrones and 2004, 78, 961.
H. Terrones, Phys. Chem. Chem. Phys., 2007, 9, 1786. 54 B. Panella, M. Hirscher and B. Ludescher, Microporous Mesoporous
16 S. Suda and G. Sandrock, Z. Phys. Chem., 1994, 183, 149. Mater., 2007, 103, 230.
17 A. W. C. van den Berg and C. O. Areán, Chem. Commun., 2008, 668. 55 S. T. Srinivas and P. K. Rao, J. Catal., 1994, 148, 470.
18 F. E. Pinkerton, B. G. Wicke, C. H. Olk, G. G. Tibbetts, 56 M. Boudart, M. A. Vannice and J. E. Benson, Z. Phys. Chem. Neue Folge,
G. P. Meisner, M. S. Meyer and J. F. Herbst, J. Phys. Chem. B, 1969, 64, 171; R. B. Levy and M. Boudart, J. Catal., 1974, 32, 304.
2000, 104, 9460. 57 W. C. Neikam and M. A. Vannice, J. Catal., 1972, 27, 207.
19 G. Férey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, 58 R. T. Yang, Y. Li and A. J. Lachawiec, Chemical Bridges for
S. Surble and I. Margiolaki, Science, 2005, 309, 2040. Enhancing Hydrogen Storage by Spillover and Methods Forming the
20 M. Latroche, S. Surblé, C. Serre, C. Mellot-Draznieks, Same, US patent application, Serial No. 11/442.898, 2006.
P. L. Llewellyn, J.-H. Lee, J.-S. Chang, S. H. Jhung and G. Férey, 59 C. Back, G. Sandi, J. Prakash and J. Hranisavljevic, J. Phys. Chem.
Angew. Chem., Int. Ed., 2006, 45, 8227. B, 2006, 110, 16225.
21 H. Chae, D. Y. Siberio-Perez, J. Kim, Y. Go, M. Eddaoudi, 60 A. Anson, E. Lafuente, E. Urriolabeitia, R. Navarro, A. M. Benito,
A. Matzger, M. O’Keeffe and O. M. Yaghi, Nature, 2004, 427, 523. W. K. Maser and M. T. Martinez, J. Phys. Chem. B, 2006, 110, 6643.
22 H. Furukawa, M. Miller and O. M. Yaghi, J. Mater. Chem., 2007, 61 M. Zielinski, R. Wojcieszak, S. Monteverdi, M. Mercy and
17, 3197. M. M. Bettahar, Catal. Commun., 2005, 6, 777.
23 Y. Li and R. T. Yang, AIChE J., 2008, 54, 269. 62 R. Zacharia, K. Y. Kim, A. K. M. Fazle Kibria and K. S. Nahm,
24 Y. Li and R. T. Yang, Langmuir, 2007, 23, 12937. Chem. Phys. Lett., 2005, 412, 369; S.-u. Rather, R. Zacharia,

278 | Energy Environ. Sci., 2008, 1, 268–279 This journal is ª The Royal Society of Chemistry 2008
View Article Online

S. W. Hwang, M.-u.-d. Naik and K. S. Nahm, Chem. Phys. Lett., 95 M. E. Davis, Nature, 2002, 417, 813.
2007, 441, 261. 96 L. Regli, A. Zecchina, J. G. Vitillo, D. Cocina, G. Spoto,
63 D. Lupu, A. R. Biris, I. Misan, A. Jianu, G. Holzhuter and C. Lamberti, K. P. Lillerud, U. Olsbye and S. Bordiga, Phys.
E. Burkel, Int. J. Hydrogen Energy, 2004, 29, 97; R. Campesi, Chem. Chem. Phys., 2005, 7, 3197.
F. Cuevas, R. Gadiou, E. Leroy, M. Hirscher, C. Vix-Guterl and 97 F. Stphanie-Victoire, A. M. Goulay and E. C. de Lara, Langmuir,
M. Latroche, Carbon, 2008, 46, 206. 1998, 14, 7255.
64 H. S. Kim, H. Lee, K. S. Han, J. H. Kim, M. S. Song, M. S. Park, 98 V. B. Kazansky, J. Mol. Catal. A: Chem., 1999, 141, 83.
J. Y. Lee and J. K. Kang, J. Phys. Chem. B, 2005, 109, 8983. 99 H. W. Langmi, D. Book, A. Walton, S. R. Johnson, M. M. Al-
65 X. Hao, L. Quach, J. Korah, W. A. Spieker and J. R. Regalbuto, Mamouri, J. D. Speight, P. P. Edwards, I. R. Harris and
J. Mol. Catal. A, 2004, 219, 97. P. A. Anderson, J. Alloys Compd., 2005, 404–406, 637.
66 R. T. Yang, Y. Li and A. J. Lachawiec, Enhancing Hydrogen 100 V. B. Kazansky, V. Y. Borovkov, A. Serich and H. G. Karge,
Spillover and Storage, US patent application filed, Serial No. 11/ Microporous Mesoporous Mater., 1998, 22, 251.
820,954, 2007. 101 M. A. Makarova, V. L. Zholobenko, K. M. Alghefaili,
67 B. S. Schueller and R. T. Yang, Ind. Eng. Chem. Res., 2001, 40, 4912. N. E. Thompson, J. Dewing and J. Dwyer, J. Chem. Soc., Faraday
68 C. J. Liu, G. P. Vissokov and B. W. L. Jang, Catal. Today, 2002, 72, 173. Trans., 1994, 90, 1047.
69 J. C. Legrand, A. M. Diamy, G. Riahi, Z. Randriamanantenasoa, 102 M. G. Nijkamp, J. E. M. J. Raaymakers, A. J. van Dillen and
M. Polisset-Thfoin and J. Fraissard, Catal. Today, 2004, 89, 177. K. P. de Jong, Appl. Phys. A, 2001, 72, 619.
70 I. G. Koo, M. S. Lee, J. H. Shim, J. H. Ahn and W. M. Lee, J. Mater. 103 A. Zecchina, S. Bordiga, J. G. Vitillo, G. Ricchiardi, C. Lamberti,
Chem., 2005, 15, 4125. G. Spoto, M. Bjrgen and K. P. Lillerud, J. Am. Chem. Soc., 2005,
Published on 24 June 2008. Downloaded on 12/2/2018 10:07:45 AM.

71 X. L. Zhu, P. P. Huo, Y. P. Zhang and C. J. Liu, Ind. Eng. Chem. 127, 6361.
Res., 2006, 45, 8604. 104 H. W. Langmi, A. Walton, M. M. Al-Mamouri, S. R. Johnson,
72 J. J. Zhou, Y. P. Zhang and C. J. Liu, Langmuir, 2006, 22, 11388. D. Book, J. D. Speight, P. P. Edwards, I. Gameson, P. A. Anderson
73 Y. Li, R. T. Yang, C.-J. Liu and Z. Wang, Ind. Eng. Chem. Res., and I. R. Harris, J. Alloys Compd., 2003, 356–357, 710.
2007, 46, 8277. 105 S. B. Kayiran and F. L. Darkrim, Surf. Interface Anal., 2002, 34, 100.
74 N. L. Rosi, J. Eckert, M. Eddaoudi, D. T. Vodak, J. Kim, 106 J. Weitkamp, M. Fritz and S. Ernst, Int. J. Hydrogen Energy, 1995,
M. O’Keefee and O. M. Yaghi, Science, 2003, 300, 1127. 20, 967.
75 J. L. C. Rowsell, A. R. Millward, K. S. Park and O. M. Yaghi, J. Am. 107 J. G. Vitillo, G. Ricchiardi, G. Spoto and A. Zecchina, Phys. Chem.
Chem. Soc., 2004, 126, 5666; J. L. C. Rowsell and O. M. Yaghi, Chem. Phys., 2005, 7, 3948.
J. Am. Chem. Soc., 2006, 128, 1304. 108 C. O. Aren, O. V. Manoilova, B. Bonelli, M. R. Delgado,
76 O. M. Yaghi, M. O’Keeffe, N. Ockwig, H. K. Chae, M. Eddaoudi G. T. Palomino and E. Garrone, Chem. Phys. Lett., 2003, 370, 631.
and J. Kim, Nature, 2003, 423, 705. 109 J. S. Beck, J. C. Vartuli, W. J. Roth, M. E. Leonowicz, C. T. Kresge,
77 A. G. Wong-Foy, A. J. Matzger and O. M. Yaghi, J. Am. Chem. K. D. Schmitt, C. T.-W. Chu, D. H. Olson, E. W. Sheppard,
Soc., 2006, 128, 3494. S. B. McCullen, J. B. Higgins and J. L. Schlenker, J. Am. Chem.
78 K. M. Thomas, Catal. Today, 2007, 120, 389. Soc., 1992, 114, 10834.
79 X. Lin, J. Jia, X. Zhao, K. M. Thomas, A. J. Blake, G. S. Walker, 110 A. Corma, Chem. Rev., 1997, 97, 2373.
N. R. Champness, P. Hubberstey and M. Schröder, Angew. Chem., 111 H. Taylor, Annu. Rev. Phys. Chem., 1961, 12, 127.
Int. Ed., 2006, 45, 7358. 112 J. H. Sinfelt and P. J. Lucchesi, J. Am. Chem. Soc., 1963, 85, 3365.
80 J. Y. Lee, L. Pan, S. P. Kelly, J. Jagiello, T. J. Emge and J. Li, Adv. 113 S. Khoobiar, J. Phys. Chem., 1964, 68, 411.
Mater., 2005, 17, 2703. 114 G. M. Pajonk, Appl. Catal., A, 2000, 202, 157.
81 H. Frost, T. Düren and R. Q. Snurr, J. Phys. Chem. B, 2006, 110, 115 J. A. Schwarz, US Pat., 4 716 736, 1988.
9565; H. Frost and R. Q. Snurr, J. Phys. Chem. C, 2007, 111, 18794. 116 P. C. H. Mitchell, A. J. Ramirez-Cuesta, S. F. Parker, J. Tomkinson
82 B. Panella and M. Hirscher, Adv. Mater., 2005, 17, 538. and D. Thompsett, J. Phys. Chem. B, 2003, 107, 6838.
83 L. Pan, M. B. Sander, X. Y. Huang, J. Li, M. Smith, E. Bittner, 117 P. C. H. Mitchell, A. J. Ramirez-Cuesta, S. F. Parker and
B. Bockrath and J. K. Johnson, J. Am. Chem. Soc., 2004, 126, 1308. J. Tomkinson, J. Mol. Struct., 2003, 651–653, 781.
84 M. Dinca and J. R. Long, J. Am. Chem. Soc., 2005, 127, 9376; 118 D. A. Panayotov and J. T. Yates Jr., J. Phys. Chem. C, 2007, 111, 2959.
M. Dinca, A. F. Yu and J. R. Long, J. Am. Chem. Soc., 2006, 128, 119 D. D. Beck and J. M. White, J. Phys. Chem., 1984, 88, 2764.
8904; S. S. Kaye, A. Dailly, O. M. Yaghi and J. R. Long, J. Am. 120 D. D. Beck, A. O. Bawagan and J. M. White, J. Phys. Chem., 1984,
Chem. Soc., 2007, 129, 14176. 88, 2771.
85 X. B. Zhao, B. Xiao, A. J. Fletcher, K. M. Thomas, D. Bradshaw 121 H.-W. Chen and J. M. White, J. Mol. Catal., 1986, 25, 355.
and M. J. Rosseinsky, Science, 2004, 306, 1012. 122 L. Chen, A. C. Cooper, G. P. Pez and H. S. Cheng, J. Phys. Chem. C,
86 D. N. Dybtsev, H. Chun, S. H. Yoon, D. Kim and K. Kim, J. Am. 2007, 111, 18995.
Chem. Soc., 2004, 126, 32; M. P. Suh, J. W. Ko and H. J. Choi, 123 U. Roland, T. Braunschweig and F. Roessner, J. Mol. Catal. A:
J. Am. Chem. Soc., 2002, 124, 10976. Chem., 1997, 127, 61.
87 B. Kesanli, Y. Cui, M. R. Smith, E. W. Bittner, B. C. Bockrath and 124 M. Jaroniec and R. Mady, Physical Adsorption on Heterogeneous
W. Lin, Angew. Chem., Int. Ed., 2005, 44, 72; Y. Kubota, M. Takata, Solids, Elsevier, Amsterdam, 1988.
R. Matsuda, R. Kitaura, S. Kitagawa, K. Kato, M. Sakata and 125 W. Rudzinski and D. H. Everett, Adsorption of Gases on
T. C. Kobayashi, Angew. Chem., Int. Ed., 2005, 44, 920. Heterogeneous Surfaces, Academic Press, San Diego, 1992.
88 D. Sun, S. Ma, Y. Ke, D. J. Collins and H.-C. Zhou, J. Am. Chem. 126 A. J. Robell, E. V. Ballou and M. Boudart, J. Phys. Chem., 1964, 68,
Soc., 2006, 128, 3896; S. Ma and H.-C. Zhou, J. Am. Chem. Soc., 2748.
2006, 128, 11734; S. Ma, D. Sun, M. W. Ambrogio, J. A. Fillinger, 127 R. Kramer and M. Andre, J. Catal., 1979, 58, 287.
S. Parkin and H.-C. Zhou, J. Am. Chem. Soc., 2007, 129, 1858. 128 A. Borgschulte, R. J. Westerwaal, J. H. Rector, H. Schreuders,
89 S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen and B. Dam and R. Griessen, J. Catal., 2006, 239, 263.
I. D. Williams, Science, 1999, 283, 1148. 129 M. Boudart, A. W. Aldag and M. A. Vannice, J. Catal., 1970, 18, 46.
90 B. Xiao, P. S. Wheatley, X. B. Zhao, A. J. Fletcher, S. Fox, 130 M. Arai, M. Fukushima and Y. Nishiyama, Appl. Surf. Sci., 1996,
A. G. Rossi, I. L. Megson, S. Bordiga, L. Regli, K. M. Thomas 99, 145.
and R. E. Morris, J. Am. Chem. Soc., 2007, 129, 1203. 131 K. J. Sladek, E. R. Gilliland and R. F. Baddour, Ind. Eng. Chem.
91 C. Prestipino, L. Regli, J. G. Vitillo, F. Bonino, A. Damin, Fundam., 1974, 13, 100.
C. Lamberti, Zecchina, P. L. Solari, K. O. Kongshaug and 132 S. Triwahyono, T. Yamada and H. Hattori, Appl. Catal., A, 2003,
S. Bordiga, Chem. Mater., 2006, 18, 1337. 250, 65.
92 Y.-Y. Liu, J.-L. Zeng, J. Zhang, F. Xu and L.-X. Sun, Int. 133 N. Satoh, J. Hayashi and H. Hattori, Appl. Catal., A, 2000, 202, 207.
J. Hydrogen Energy, 2007, 32, 4005. 134 Y. Li, F. H. Yang and R. T. Yang, J. Phys. Chem. C, 2007, 111, 3405.
93 A. P. Cote, A. I. Benin, N. W. Ockwig, M. O’Keeffe, A. J. Matzger 135 K. Higashi, H. Ito and J. Oishi, J. At. Energy Soc. Jpn., 1963, 5, 846.
and O. M. Yaghi, Science, 2005, 310, 1166. 136 R. T. Yang, J. B. Fenn and G. L. Haller, AIChE J., 1973, 19, 1052.
94 M. El-Kaderi, J. R. Hunt, J. L. Mendoza-Cortés, A. P. Côté, 137 K. Kapoor and R. T. Yang, AIChE J., 1989, 35, 1735.
R. E. Taylor, M. O’Keefe and O. M. Yaghi, Science, 2007, 316, 268. 138 A. J. Lachawiec and R. T. Yang, Langmuir, 2008, 24, 6159.

This journal is ª The Royal Society of Chemistry 2008 Energy Environ. Sci., 2008, 1, 268–279 | 279

You might also like