12
Published on 09 May 2007. Downloaded by New York University on 07/10/2014 22:06:54. View Article Online / Journal Homepage / Table of Contents for this issue

Hydrogen storage: the remaining scientific and technological challenges

  • Upload
    ulrich

  • View
    213

  • Download
    1

Embed Size (px)

Citation preview

This paper is published as part of the high-profile series of PCCP special issues on Alternative Fuel Technologies.

Guest edited by Joachim Maier (MPI Guest edited by Joachim Maier (MPI Stuttgart), Dirk Guldi (Universität Erlangen-Nürnberg), and Adriano Zecchina (University of Torino), and published in selected 2007 print issues of PCCP, all papers are collected online on a dedicated website:

www.rsc.org/pccp/altfuel

Visit the website for both cutting edge research papers and authoritative review articles by leaders in a range of fields of critical importance to the world today

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

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

Hydrogen storage: the remaining scientific and technological challenges

Michael Felderhoff,*aClaudia Weidenthaler,*

aRittmar von Helmolt

band

Ulrich Eberleb

Received 31st January 2007, Accepted 17th April 2007

First published as an Advance Article on the web 9th May 2007

DOI: 10.1039/b701563c

To ensure future worldwide mobility, hydrogen storage in combination with fuel cells for on-

board automotive applications is one of the most challenging issues. Potential solid-state solutions

have to fulfil operating requirements defined by the fuel cell propulsion system. Important

requirements are also defined by customer demands such as cost, overall fuel capacity, refuelling

time and efficiency. It seems that currently none of the different storage solid state materials can

reach the required storage densities for a hydrogen-powered vehicle. New strategies for storage

systems are necessary to fulfil the requirements for a broad introduction of automotive fuel cell

powertrains to the market. The combination of different storage systems may provide a possible

solution to store sufficiently high amounts of hydrogen.

1. Introduction

One of the most crucial points concerning the implementation

of hydrogen-based propulsion systems is the on-board storage

of hydrogen. Although showing some drawbacks such as

higher cost and lower energy density compared to a gasoline

or diesel tank, hydrogen storage is much closer to automotive

cost and performance figures than a large-scale battery system.

Additionally, for some methods (in particular liquid hydrogen

LH2 and compressed gaseous CGH2), refuelling is possible in

less than 5 min, compared to the recharging times of many

hours for high-voltage batteries. Utilizing other hydrogen

storage techniques, refuelling times of less than 1 h are

achievable. Thus, the application of hydrogen and fuel cells

as electrical energy source has attracted researchers for a long

time. However, even after large efforts have been made both in

the scientific and industrial world during the last 10 years, a

ubiquitous solution for the on-board storage of hydrogen has

not yet been found. The specific requirements of different

applications such as vehicles, portable or stationary devices

need the development of completely different hydrogen sto-

rage approaches. The physical storage methods of hydrogen

under cryogenic temperatures or high-pressures (35 to

70 MPa) can only be used in systems where larger amounts

of hydrogen in the range of several kg have to be stored. The

related technology elements are too complex for incorporation

into consumer electronics products or power tools. In these

cases only chemical methods for the storage and release of

hydrogen from metal hydrides or from hydrolysis reactions

with complex hydrides can be used. In recent years, a very

large number of publications dealing with the different storage

methods have been published. Often the implementation of a

storage method into technical systems is even more an en-

gineering challenge than a question of the hydrogen storage

capacity of the utilized materials or the storage density of the

system. Long-term storage, heat conductivity problems, heat

management or the kinetics of the reloading and deloading

processes of metal hydrides are only some examples. Some of

these problems, often ignored in the literature, are discussed in

this review.

The PEM type (proton exchange or polymer electrolyte

membrane) fuel cell, originally invented in 1962 by General

Electric, was used during NASA’s Gemini space missions.

New electrolyte materials for this type of fuel cell offered the

possibility of a more compact and lightweight fuel cell system

and respective propulsion system. During the 1990s, several

car companies began to develop the PEM fuel cell for passen-

ger car propulsion systems. At that time, the motivation for

the development was mainly based on emissions reduction.

Very similar to pure battery-electric vehicles, all emissions

occur during fuel production, whereas the only local waste

product generated during driving is just water vapour. Though

power densities have already significantly increased over re-

cent years, fuel cells still are not fully competitive with today’s

high-performance internal combustion engines. But recently,

the power density of this kind of drivetrain became sufficient

for the propulsion of a wide variety of passenger car archi-

tectures. Now, cost reduction, robustness and durability have

to be addressed, as well as the issue of hydrogen storage. A

consensus within the automotive industry has emerged that

fuel cells and hydrogen are the ultimate long-term technology

solution; but to achieve full-scale commercialisation, technical

improvements are still required. In the meantime, continuous

improvements in internal combustion engines, and transition

technologies (such as the hybridization of powertrains), and

the introduction of renewable fuels (such as ethanol or syn-

thetic fuels produced from biomass) are beginning to diversify

the portfolio of powertrain and fuel options.

Since most of the world’s car manufacturers invest in

extensive research activities in hydrogen storage and PEM

aMax-Planck-Institut fur Kohlenforschung, Kaiser-Wilhelm Platz 1,45470 Mulheim/Ruhr, Germany. E-mail: [email protected]; [email protected]

bGM Fuel Cell Activities, Hydrogen & Fuel Cell Research Strategy(Europe), IPC MK-01, 65423 Ruesselsheim, Germany. E-mail:[email protected]; [email protected]

This journal is �c the Owner Societies 2007 Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 | 2643

INVITED ARTICLE www.rsc.org/pccp | Physical Chemistry Chemical Physics

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

fuel cell developments, the main focus of this review is on

hydrogen storage methods for automotive applications under

the boundary conditions of a fuel cell drivetrain. The concept of

hydrogen internal combustion engines (ICE) shall be men-

tioned just briefly. Although having some advantages concern-

ing infrastructure implications (dual-fuel engines—gasoline and

hydrogen—are feasible), and technology maturity, there exist

significant challenges regarding vehicle efficiencies and the H2

supply rates from the storage system to the IC engine. A fuel

cell powertrain has a 36% efficiency in the European Driving

Cycle (i.e. GM HydroGen3, 4.6 kg H2, 400 km range), the

corresponding diesel engine shows EDC efficiency values of

22%. That translates reciprocally into a higher fuel consump-

tion for the diesel–IC variant. To ensure the same vehicle range,

a hypothetical hydrogen-powered IC engine (assuming the

same efficiency as the conventional variant) would thus require

a storage system about 1.65 times larger than the one of a fuel

cell system. A similar relationship and a similar factor applies

for the hydrogen rate that the tank system has to supply to the

powertrain. Both points are very difficult to tackle: first, the

available space for the tank system on board of the vehicle is

very limited. Second, a higher extraction rate poses an even

more unfavourable heat management challenge on the hydro-

gen system. Also, the heat management challenge during

refueling is significantly more demanding for an IC engine

when a re-filling time of less than 5 min is targeted. Both heat

management issues will be addressed and discussed in detail in

the next section. For those reasons in general (and efficiency,

respectively, fuel consumption in particular), most car manu-

facturers concentrate their efforts on hydrogen fuel cell vehicles.

2. On-board hydrogen storage options by physical

methods

Before discussing different storage options, it has to be made

clear that all values for gravimetric and volumetric energy

densities may correspond either to (a) just a materials ap-

proach or (b) a systems approach including all required

components and mounts. The definitions for both approaches

(a) and (b) should not be confused. From an automotive

perspective, the second approach is preferable; also the target

values provided by the US Department of Energy (DOE) are

developed and published on that systems basis.1,2

In this section, conventional methods, as well as the frame-

work and pre-conditions for pure solid-state and hybrid

storage technologies will be discussed in detail. The material-

based issues will be addressed extensively in the following

sections.

At present, there are two major physical options for on-

board hydrogen storage:

(1) CGH2 compressed gaseous hydrogen (35–70 MPa and

room temperature).

(2) Cryogenic (LH2 liquid hydrogen at 20 to 30 K,

0.5–1 MPa).

In the past, both options (1) and (2) have been implemented

by the automotive industry in many prototype vehicles. For

being competitive in range to conventional gasoline or diesel-

based systems, typically 4–7 kg of hydrogen have to be stored

on board. This remains a serious issue for the vehicle integra-

tion since the density of hydrogen is rather low (Fig. 1), even

when compressed to 70 MPa. Furthermore, cylindrical vessels

are required for these two and most of the other types of

storage systems. In existing vehicle architectures, without

modifications, there is no space for hydrogen storage modules

which could provide sufficient range. Thus, rear body mod-

ifications are necessary to integrate the hydrogen tank. In an

extreme case, one could even imagine concepts where the car is

built around the hydrogen storage device. In other cases, for

vehicle packaging reasons, the fuel storage comprises not only

a single but two or more pressure vessels. The design of such a

vessel (type IV, hence using polymer liner materials, instead

of a metallic liner for a type III vessel) is shown in detail in

Fig. 2.3

The volumetric storage density of CGH2 systems, both on a

system (Table 1)4 and a material level is rather low.5 The

respective energy densities for a hydrogen tank comprising a

single vessel correspond to values of about 0.048 kg H2 per kg

tank weight and 0.023 kg H2 per litre tank volume. Together

with the necessity for a cylindrical vessel design (caused by the

operating pressures of 35–70 MPa), the integration of such a

tank into existing car architectures remains to be an important

challenge. But despite all limitations of the CGH2 technology,

so far, this option yields the best overall technical performance

and shows the highest maturity for automotive applications.

In the past, liquid hydrogen was also regarded to be a viable

option for the automotive industry. However, due to the very

low phase-change enthalpy of about 0.45 MJ kg�1 hydrogen

between the liquid and gaseous state, the related technology

challenges could not be solved satisfyingly. Due to the low

operating temperature of 20 to 30 K and the large temperature

difference to the environment (300 K) an unavoidable heat

input (see below) of 2 to 3 W, consisting of three components,

occurs.6 The three different effects are:

(1) Thermal conduction, (2) convection, and (3) thermal

radiation

Among these, the thermal conduction (1) through pipes,

cables and mountings to the inner storage vessel and the

thermal radiation (2) from the environment to the cryogenic

liquid are dominant. Since a very good (hence very low)

surface-to-volume ratio S/V is needed, low overall values of

Fig. 1 Mass density of hydrogen under various conditions.

2644 | Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 This journal is �c the Owner Societies 2007

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

heat flows can only be achieved when working with cylindrical

tank structures (compare to CGH2). Due to their even further

unfavourable ratio for the same hydrogen capacity, so called

conformable tanks with more complex shapes show inherently

a worse thermal performance. A cylindrical LH2 system is

shown in Fig. 3a. In any case, the use of a very efficient multi-

layer vacuum super insulation (see Fig. 3b) consisting of

approximately 40 layers of metal foil is required. Wrapping

these foils around the cylindrical parts and the dome areas (as

well as the in- and outlets for H2 and the mountings) is time-

consuming and highly demanding. For conformable tanks,

this point becomes even more difficult and the thermal insula-

tion thus gets less effective. The remaining significant heat

input leads to an evaporation of the liquid phase that even-

tually causes a pressure rise. The time period between putting

the vehicle into an idle or parking mode and the venting

process is normally called ‘‘dormancy’’. Usually, when a

pressure level of approximately 1 MPa is reached, a valve

has to be opened and hydrogen is eventually vented. Depend-

ing on the application, hydrogen will either be diffused into

air, catalytically burned, or captured. Typically, the time range

for these processes last several days. After that, hydrogen is

continuously lost to the environment. The amount of hydro-

gen released is known as boil-off gas and can be calculated by

multiplying the heat flow and the time period, divided by the

H2 heat of evaporation. Both the length of the dormancy

period and the amount of H2 lost are considered to be crucial

for the down-selection process of hydrogen storage systems.

A related issue are the cooling-down losses during re-filling

hydrogen. The complete transfer line has to be cooled down to

about 20 K and therefore evaporation occurs. Although great

efforts are made, these losses also cannot be reduced to zero

and remain significant. Thereby, a so called ‘‘cold finger’’ (a

jacketed pipe) is introduced into the nozzle to establish a

connection path at cryogenic temperature levels between the

filling station and the vehicle. Before the hydrogen flow starts,

this line undergoes a helium purge procedure to remove all

contained air.

Both effects, on-board- and infrastructure-related, lead to

unacceptable hydrogen losses. The complexity of the LH2

storage system together with the challenge to reduce the

amount of boil-off gas leads to overall LH2 system costs which

are—at large scale—not favourable over those of CGH2

systems. Despite the fact that the volumetric storage density

of complete LH2 systems is slightly higher compared to CGH2

systems, most car manufacturers do not see strong advantages

in packaging that might outweigh the disadvantages men-

tioned above. Besides, the design flexibility of LH2 tank

systems is not really superior to CGH2 systems.

Additionally, the energy required to liquefy hydrogen al-

ready consumes 30% of the chemical energy stored based on

the net calorific value or lower heating value (LHV) of 120 MJ

per kg H2. For a simple comparison with CGH2 it is sufficient

to consider an ideal gas compressed under isothermal condi-

tions: by integrating over the ideal gas law, this results in

mechanical energy of about 8 MJ per kg H2 (or 7% of the

LHV) required to reach a pressure level of 70 MPa. On the

other hand, a real compression process is far from being

isothermal, and also technical efficiencies have to be consid-

ered. Hence, it has to be stated that an energy amount

corresponding to about 15% of the LHV is necessary to reach

a 70 MPa level in a real-world technology environment (and

12% of LHV for 35 MPa). However, these values are still

significantly lower than the technical liquefaction energy of

hydrogen (30% LHV, Fig. 4). The gravimetric and volumetric

hydrogen densities (based on the systems approach) of current

CGH2 tanks are shown in Table 1.

It seems that all relevant alternatives have to beat these

figures in most of the categories. The solid-state materials

based options and their challenges before implementation will

be discussed in the next section of this publication. But what

Fig. 2 Type IV compressed gaseous hydrogen vessel. Reprinted from ref. 5 with permission from Elsevier.

Table 1 Benchmarking of hydrogen storage technologies: compar-ison with existing 70 MPa high-pressure storage in carbon-fiber vessels(cost figure, compare to the European SRA target)4

Benchmark system 70 MPa CGH2Capacity 6 kg H2

Volumetric energy density 260 L, 0.023 kg L�1

Gravimetric energy density 125 kg, 0.048 kg kg�1

Shape CylindricalProduction cost @ large volumes 2000 Euro (from EU

strategic research agenda)Boil-off losses Not existingExtraction efficiency 100%Max. extraction rate 42 g H2 s

�1

Refilling time 3 minRefilling efficiency 495%Heat exchanger capability 0 kW

This journal is �c the Owner Societies 2007 Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 | 2645

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

are the boundary conditions in a perfect-world scenario?

Compared to the challenges of the CGH2 system, there are

limitations of available volume from vehicle packaging needs

in existing car architectures. Potential solid-state solutions

have to fulfil operating requirements defined by the fuel cell

propulsion system e.g. extraction rate, supply pressure to the

fuel cell system of about 0.5 MPa, as well as operating

temperature. Important requirements are also defined by

customer demands such as cost, overall fuel capacity, refueling

time and efficiency.

Fig. 3 (a) Liquid hydrogen vessel. (b) Schematic drawing of a multi-layer vacuum super insulation. Reprinted from ref. 5 with permission from

Elsevier.

Fig. 4 Isothermal compression of hydrogen, assumption of required mechanical energy, using an ideal-gas behaviour, a comparison to technical

values and to the required energy for the technical liquefaction energy of hydrogen is provided.

2646 | Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 This journal is �c the Owner Societies 2007

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

3. Physisorption

3.1 Introduction

For the condensation of hydrogen weak van der Waals inter-

actions of hydrogen molecules are the driving forces. The heat

of condensation (0.9 kJ mol�1 H2) of hydrogen is rather low.

This is reflected by the low boiling point of hydrogen (20.4 K).

Physisorption of hydrogen on surfaces is dominated by

slightly stronger van der Waals interactions between hydrogen

molecules and the surface of the adsorbent. The heat of

adsorption for porous materials is in the range of about

4–10 kJ mol�1 H2.7 Interactions between hydrogen molecules

are only of importance at temperatures between the boiling

point and the critical temperature (33.25 K)8 of hydrogen.

Only in this temperature range, a liquid phase can be expected.

A potentially very high hydrogen uptake inside pores as a

result of capillary condensation is therefore considered to be

impossible. For this reason and since the van der Waals forces

are rather weak, no significant amount of hydrogen can be

stored at ambient temperature and pressure. Physisorption

results in the formation of a hydrogen monolayer on the

adsorbent surface due to the stronger interaction of the

hydrogen with the surface.9 Adsorption at a temperature equal

or higher than the boiling point (20.4 K) of hydrogen thus only

leads to the adsorption in a monolayer.10 Hence, a multi-layer

coverage at 77 K (or even at room temperature) cannot really

be expected because storage temperatures are far above the

boiling point of liquid hydrogen.11 This is indicated by the fact

that hydrogen adsorption isotherms are typically showing the

shape of a Langmuir isotherm (i.e. monolayer adsorption).12

For a monolayer coverage at 77 K, the mass of hydrogen

adsorbed under saturation conditions is generally propor-

tional to the specific surface area determined by the BET

method.9,13 As an example, graphene sheets with a specific

surface area of 1315 m2 g�1, adsorb a maximum amount of 2

wt% of hydrogen.9 The excess amount of hydrogen adsorbed

at 77 K from the gas phase in saturation is thus 1.5 (�0.5) �10�3 wt% m�2 g.

3.2 Carbon materials

The intrinsic properties of carbon materials, i.e. their low

densities, high porosities and high specific surface areas make

them interesting as adsorbents for hydrogen. In the 1990s

extraordinary high hydrogen uptakes for different carbon

materials were published by several groups. Hydrogen adsorp-

tion capacities up to 10 wt% due to condensation of hydrogen

inside narrow single wall nanotubes (SWNT) were reported,

even at ambient conditions.14,15 Unfortunately, the reproduc-

tion of these spectacular high uptake capacities by other

groups failed. Instead uptake capacities of only one third of

the reported ones were observed, and only at cryogenic

temperatures.16 For graphitic nanofibers capacities of about

67 wt% had been reported.17,18 However, those very promis-

ing capacities of graphitic nanofibers could never be con-

firmed,19–22 and may be attributed to erroneous

measurements.

Even though the early reports on hydrogen uptake of

carbon materials, as mentioned above, were very promising,

in the following only relatively low uptakes could be ob-

served.13,23–26 For SWNT a hydrogen capacity below 1 wt%

was achieved at 80 bar and room temperature.16 At lower

pressure of about 2 MPa only 0.1 wt% uptake was found.26

Low capacities were also measured for activated carbons

(maximum capacity 1.6 wt%) close to ambient conditions.

Zuttel et al. investigated the hydrogen storage capacity of

more than 60 carbon samples at room temperature.27 Depend-

ing on the type of graphite and the specific surface areas of the

carbon materials, the reversible storage capacity ranged from

0.04–0.46 wt%.

There are very controversial reports concerning the influ-

ence of micropore volume, size of micropores and surface area

on the hydrogen uptake. Even though there is a broad range of

hydrogen uptake for a wide variety of porous and nanostruc-

tured carbons, correlations between the sorption capacity,

specific surface area and/or micropore volume are reported.

A very comprehensive study on hydrogen adsorption of more

than 30 different porous materials (carbons, silica, alumina,

and MOFs) with respect to the specific surface area, total pore

volume and micropore volume was recently published.28 The

systematic analysis of all data obtained at 77 K and 0.1 MPa

indicates a clear dependence of the H2 uptake on the size of the

micropores and the micropore volume. In general, the hydro-

gen uptake seems to be limited by the adsorbate density, the

pore structure of the adsorbent, and the pore volume of the

narrowest pores. Materials with very high pore volumes do not

necessarily adsorb much hydrogen. This is assigned to lower

interaction energy of hydrogen in wide pores compared with

smaller micropores. Hydrogen storage is dominated by small

pores with a narrow size distribution and therefore a certain

scatter (dependent on the pore size distribution) of the storage

capacities around the general proportional trend line is ob-

served.29 Small pores (o1 nm) are most efficient for hydrogen

storage while mesopores (420 nm) do not contribute much to

the excess hydrogen capacity.

Hydrogen storage in carbon nanostructures close to ambi-

ent conditions seems to be limited to values far below those set

as a requirement of the DOE and the automotive industry due

to physical reasons. Unfortunately, there is no way to over-

come the fundamental laws of physics and it is thus futile to

claim higher capacities than theoretically possible.

So far, there are only limited possibilities of increasing the

adsorption capacities to fulfil the requirements for mobile

applications. One possible strategy to increase the hydrogen

uptake is the adjustment of the physical conditions for the

storage process. The physical procedures are either to decrease

the adsorption temperature to about 80 K, and/or to increase

the hydrogen pressure. Another way is to tune the properties

of the adsorbent which is mainly limited to an increase of the

surface area and/or an increase of the micropore volume of the

adsorbent.

3.3 Metal–organic frameworks (MOFs)

Metal–organic frameworks (MOFs) appeared as highly pro-

mising storage materials for hydrogen. These materials com-

bine both high surface areas and large micropore volumes. In

1989 Hoskins and Robson proposed a new class of solid

This journal is �c the Owner Societies 2007 Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 | 2647

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

polymeric materials later known as MOFs.30 Inorganic build-

ing units are connected by organic linkers such as carboxylates

forming a 3D network. One of the most prominent members

with a structure consisting of four Zn4O(CO2)6 units con-

nected by benzene rings is called MOF-5.31 Four ZnO4 tetra-

hedra are joined by benzene dicarboxylate linkers resulting in

a 3D cubic framework with interconnected pores. The high

microporosity and a very high specific surface area make

MOFs interesting for physisorption processes. In early studies,

hydrogen uptakes of 1.98 wt% at 77 K and 0.1 MPa for

[Zn3(bpdc)3bpy] � 4DMF �H2O and 1.74 wt% for

[Co3(bpdc)3bpy] � 4DMF �H2O were reported.32 Higher up-

takes of 2.47 wt% at 1 bar and 77 K were obtained for

Cu–MOF (MOF-505).33 Actually, for MOF-5 exhibiting a

very high specific BET surface area of 2296 m2 g�1 an uptake

of 5.1 wt% at 5 MPa and 77 K could be independently

reproduced.34

Recently, for MOF-177 with a framework of zinc acetate

units linked by 1,3,5-benzenetribenzoate (BTB) with an esti-

mated Langmuir surface area of 4500 m2 g�1,35 a storage

capacity of 7.5 wt% at 7 MPa and 77 K was measured.36 This

is by far one of the highest surface areas and therewith the

highest hydrogen storage capacity for MOFs. The capacities

are even better than for high-performance activated carbons

such as AX-21 and Maxsorb MSC-30. Not only the surface

area but also the nature of organic linkers and the inorganic

units seem to have an influence on the storage capacity of the

MOFs at low pressures (before the saturation value of the

isotherm is reached).

3.4 Hydrogen storage in zeolites

The crystal structure of zeolites is defined by channels and

cavities which, if they are interconnected, form a pore system

large enough for diffusion of molecules or ions. Zeolites

possess large micropore volumes making them potential can-

didates for the storage of hydrogen. For zeolites two ways for

storing hydrogen are discussed: encapsulation and adsorp-

tion.37 The process first involves the diffusion of hydrogen

molecules into channels and cages (voids) of the structure. One

of the very first proposing zeolites as material for the encap-

sulation of hydrogen was Fraenkel.38,39 In the following years,

many different ion-exchanged zeolites were tested as potential

hydrogen storage materials.40–42 Trapping hydrogen is only

possible by increasing temperature and/or pressure. Under

ambient conditions, the molecules stay entrapped in the voids

by diffusion limitation. Vitillo et al. give a survey on zeolites

studied for the encapsulation of hydrogen.37 The uptake of

hydrogen by adsorption is below 0.5 wt% at ambient condi-

tions and below 2 wt% at 77 K and elevated pressures. The

authors calculated the maximum amount of hydrogen stored

either by encapsulation or by adsorption. Geometric con-

straints such as, for example, available volume and framework

flexibility restrict the hydrogen storage capacities of zeolites to

2.86 wt%. The theoretical calculations are supported by

experimental results. Unfortunately, almost no hydrogen can

be stored at room temperature and even at cryogenic tem-

peratures, the capacities are low. The influence of the frame-

work structure and the type of exchangeable cations on the

capacity was the topic of several publications. For all systems,

at low hydrogen loadings, the cell volume of the zeolite

decreases due to attractive forces between adsorbent and

adsorbate. After reaching a minimum, the volume increases

with increasing loading. This is due to a complete filling of the

free volume of the zeolite. Even starting with different geome-

tries, the filling curves all follow the same trend. The nature of

the extra framework cations seems to have an influence on the

hydrogen uptake.43 For zeolite A and RHO pore blocking by

large cations is the factor restricting the uptake of hydrogen.

More open structures as zeolite X and Y show no pore

blocking effects. However the highest uptake, observed for

Ca-exchanged zeolite X is about 2.2 wt% at 77 K and 1.5

MPa. Li-containing low silica zeolites show a capacity of 1.5

wt% at 77 K and 0.1 MPa.44 However, even though at first

glance microporous zeolites seemed to be promising materials

for hydrogen storage, the capacities and experimental require-

ments are unfavorable to make them real storage materials.

3.5 Summary of physisorption

As this short review of the literature shows, there is much

experimental evidence for the very limited hydrogen adsorp-

tion capacity of carbon materials, MOFs or zeolites close to

ambient conditions, confirming the theoretical limitations due

to physical reasons. It is a moot question whether it is

advisable to put too much effort in further detailed investiga-

tions of the hydrogen uptake of carbon materials at room

temperature. Even materials with higher surface areas are far

from providing sufficiently high storage capacities of more

than 5 wt% on a materials basis at room temperature. A very

comprehensive analysis of hydrogen adsorption capacities of

different porous adsorbents shows the general trend of hydro-

gen uptake vs. specific surface very clearly (Fig. 5).7 Cryogenic

storage approaches can increase the adsorption capacities of

adsorbers significantly or are the main requirement for ad-

sorption at all. However, there remains the open question

Fig. 5 Variation of hydrogen adsorption capacity at 0.1 MPa and

77 K with specific surface area for different porous materials. Carbon

materials: solid symbols, silicas, aluminas, and zeolites: cross symbols,

MOFs open symbols. (Reprinted from ref. 7 with permission from

Elsevier).

2648 | Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 This journal is �c the Owner Societies 2007

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

whether extremely low temperatures necessary for adsorption

of hydrogen fulfil the requirements for a reasonable energy

balance of the storage process, especially at the site of the

filling station. Another challenge, as for any solid state absor-

ber, is the kinetics of both dehydrogenation and rehydrogena-

tion processes and the related engineering burden to achieve

re-filling times shorter than 5 min.

A problem which is very often ignored concerns the techni-

cal problem of heat released by the storage process. A simple

calculation shows the inherent challenges of hydrogen adsorp-

tion using a liquid-nitrogen based heat exchanger to ensure an

operating temperature of 77 K. The heat of adsorption of

hydrogen on a surface is, depending on the adsorbent, in the

range between 4–10 kJ mol�1 H2. During the storage process

of 6 kg H2 heat in the order of magnitude of 12–30 MJ is

produced. When keeping the operating temperature constant,

this large amount of heat can only be removed by the

evaporation of liquid nitrogen. For nitrogen, the heat of

vaporization is 5.6 kJ mol�1 N2. To remove the heat released

during the adsorption of hydrogen high amounts of liquid

nitrogen are necessary. For 4 kJ mol�1 H2, 2200 mol N2

corresponding to 80 kg liquid nitrogen are required. About

200 kg of liquid nitrogen would be necessary if the heat of

adsorption would be close to the higher values. This extremely

large quantity of liquid nitrogen required for cooling purposes

causes severe engineering challenges. Therefore, more sophis-

ticated heat management technologies and tank operating

strategies (compared to a conventional 77 K LN2 dewar) are

required to provide a technically convenient way for storing

acceptable amounts of hydrogen. On a materials basis, about

10 wt% excess hydrogen capacity (at 77 K and 20 bar) and a

volumetric hydrogen density of minimum 35 g L�1 are needed

for such a system.

4. Hydrogen storage in chemical hydrides

4.1 Introduction

Chemical and metal hydrides exhibit an impressive volumetric

hydrogen density on a materials basis.5 To be competitive with

the very short refueling times of about two minutes for a

gasoline or diesel-based propulsion system, hydrogen based

systems have to overcome engineering challenges. Considering

a 6 kg H2 tank system (Table 1) utilizing a solid state absorber

M with an enthalpy of formation DH of about 20 MJ kg�1 H2

(typical for many hydrides), a thermal load of 120 MJ would

has to be compensated during refuelling:

MþH2 !MH2 þ DH

This leads to an average heat exchanger power of more than

600 kW. Such a high-performance device is not imaginable to

be installed on-board a vehicle due to cost, volume and weight

reasons. Typically, values less than 100 kW would be reason-

able for an automotive application. Assuming the driving

operating mode under full load, it is necessary to guarantee

a H2 supply rate of 2 g s�1 to the fuel cell propulsion. This

involves a heat management challenge of about 40 kW to be

supplied to the tank system.

Another restriction is that many solid state absorber sys-

tems, in particular many hydride systems, require operating

pressures of just below or above 10 MPa (at least during

refuelling). Hence an additional pressure container made of

advanced components such as carbon composites is necessary.

The perfect metal hydride system for an automotive applica-

tion store has to store more than 6 wt% hydrogen near room

temperature and with an equilibrium pressure of approxi-

mately 1 MPa. From the van’t Hoff equation the reaction

enthalpy and equilibrium pressure can easily be calculated

using the standard entropy of 130 J mol�1 K�1 for H2. At an

equilibrium pressure of 1 MPa of hydrogen the reaction

enthalpy DH is 33 kJ mol�1 H2 (Table 2). For such a system,

an amount of heat of 100 MJ would be released during the

refilling process with 6 kg H2. Table 2 describes roughly the

expected amount of heat released during the storage of 6 kg H2

in a solid storage system, depending on the stability of the

metal hydrides. The amount of material needed is calculated

for three different hydrogen storage capacities. A storage

material with an equilibrium pressure of 0.1 MPa and a

storage capacity of 4 wt% H2 releases 118 MJ heat during

the refilling process. An amount of 150 kg storage material is

necessary for this storage process. The same heat is released

from a material with a storage capacity of 8 wt% H2, but the

amount of hydride needed is reduced to 75 kg. These simple

calculations show that a reduction of the amount of heat is

only possible in combination with more unstable hydrogen

storage materials. An unstable metal hydride with an equili-

brium pressure of 300 bars releases only 75 MJ heat. This is

1/3 lower in comparison to a material with an equilibrium

pressure of 0.1 MPa. The heat management during the refilling

process of a solid state absorber is the crucial problem. As a

consequence, lowering of the heat produced is only possible if

less stable metal hydrides are used. But in this case materials

with an equilibrium pressure around 30 MPa at room tem-

perature are the materials of choice. Additionally, the operat-

ing temperature of a target material in the perfect-world

scenario has to be limited to 343 K for a low-temperature

PEM fuel cell. Such a temperature level could be provided by

using the waste heat of the fuel cell system. To serve higher

operating temperatures, hydrogen has to be converted directly

into heat by catalytic burning or indirectly by using an

electrical heater. Implementing one of these two options would

decrease the effective fuel capacity of the tank system and

lower the vehicle range due to the fact that this amount of

hydrogen is not available for the drivetrain supply.

Table 2 Heat release during the hydrogenation of metal hydrogensystems depending on the stability and the hydrogen storage capacityof the hydride used for the storage of 6 kg H2

Pressure/MPa

DH/kJ mol�1

H2

Heat release for 6 kg H2/MJ

4 wt%H2

6 wt%H2

8 wt%H2

0.1 39 118 118 1181 33 100 100 10010 28 84 84 8430 25 75 75 75Amount of storage materialrequired/kg

150 100 75

This journal is �c the Owner Societies 2007 Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 | 2649

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

The hydrogen content of more than 6 wt% excludes all

interstitial metal hydrides (LaNi5, TiFe) as storage materials,

because of their low storage capacities. On the other hand,

materials with high hydrogen content (MgH2) have high

decomposition temperatures which cannot be reached with a

PEM fuel cell.

An often proposed way out of this dilemma is the decom-

position of hydrogen-rich compounds such as sodium boro-

hydride, ammonia borane, or alane. Although these compounds

might be quite different, the challenges regarding their utiliza-

tion are similar. These days, the automotive industry has

evaluated these materials to be not viable for three or four

reasons, respectively:

(1) Inherent system complexity concerning the handling of

the decomposition process and the storage of the waste

material on-board.

(2) Infrastructure implications (a fuel cartridge capable of

storing 5 kg of hydrogen would weigh at least 50 kg when a

10% material storage density is assumed. Furthermore, the

vessel itself and other system components would add weight

and volume.

(3) Recycling and energy issues: typically, the waste material

ends in a deep thermodynamic sink, therefore in most cases no

reasonable cost- and energy-effective recycling process exists.

(4) Safety issues when a storage material is operated far

from its equilibrium conditions (may not apply to all concepts

mentioned above).

Since the first publication from Bogdanovic and Schwick-

ardi about Ti-doped NaAlH445 new research activities in

hydrogen storage systems have started. Complex metal hy-

drides, amides and thermodynamic tailored systems are now

the focus of research.

4.2 Borohydrides

Complex borohydrides, as for example the most commonly

used NaBH4, are materials with high hydrogen content (10.8

wt% for NaBH4). However, most of the borohydrides show

unfavourable thermodynamic properties and therefore cannot

be used as reversible hydrogen storage materials under accep-

table technical conditions. The thermal decomposition tem-

perature of NaBH4 around 673 K is much too high for PEM-

fuel cell applications (eqn (1)).

NaBH4 �!400 �C

NaHþ Bþ 1:5H2 ð1Þ

Compared to the complex aluminium hydride compounds the

thermal decomposition of complex borohydrides occurs in a

one step mechanism without the formation of any intermedi-

ate complex. The final products of the decomposition are

binary metal hydrides and elemental boron. Sometimes poly-

nuclear boron hydrides are formed. The formation of traces of

boron hydrogen compounds (BH3 or the dimerization product

B2H6) during the thermal decomposition can be a problem,

due to poisoning of the fuel cell catalyst and the destruction of

the membrane. Alternatively to the thermal decomposition

complex sodium, borohydrides can be used for the production

of hydrogen in a hydrolysis reaction with water.46,47 In this

case half of the released hydrogen originates from the water

and increases the hydrogen capacity of the whole system. Low

amounts of sodium hydroxide stabilize the NaBH4–water

solution and this mixture can be stored for a long time without

decomposition. For the decomposition a catalyst such as a Ru

compound is necessary. The final product of the hydrolysis of

NaBH4 is sodium metaborate NaBO2 (eqn (2)). The solubility

of NaBO2 in water is only 26 g/100 ml at 293 K, whereas

NaBH4 has a solubility of 55 g/100 ml at 293 K. To prevent

precipitation from the solution and blocking of active sides of

the catalyst, the NaBH4 concentration used must be lower

than the maximum solubility of the metaborate.48 A typical

composition of a commercial NaBH4 aqueous solution (Mil-

lennium Cell) is 20% NaBH4 and 1% NaOH for stabilization

of the solution. This low concentration of NaBH4 reduces the

storage capacity to 4 wt% for the whole hydride system.

NaBH4 þ 2H2O! 4H2 þNaBO2 ð2Þ

One significant barrier to the broad introduction of NaBH4 as

a hydrogen releasing material is the high price for the desorbed

hydrogen and the regeneration of the hydride from the

metaborate solution. The metaborate NaBO2 is a thermody-

namic sink with a heat of formation of DH=�1058 kJ mol�1.

For NaBH4 the heat of formation DH is �191 kJ mol�1. The

DDH for the preparation of NaBH4 starting from NaBO2 is

about 900 kJ mol�1. This energy amount is required for the

regeneration process in terms of chemical or other types of

energy and makes the process quite expensive. In practice the

NaBO2 is dissolved in water and cannot regenerate to NaBH4

in a water solution. A high additional amount of energy is

necessary for boiling off the water and drying the metaborate.

One additional problem could be the residue of sodium

hydroxide in the reaction mixture. Different studies for the

regeneration of NaBH4 from the NaBO2 or Na2B4O7 have

been published. At temperatures between 623–973 K and

hydrogen pressures of 7 MPa the reaction of NaBO2 with

MgH2 or MgSi regenerates NaBH4. Depending on the reac-

tion time, up to 98% yield can be obtained.49 Starting from a

boron oxide compound, NaBH4 can be prepared by the

reduction with MgH2 in a ball-milling process.50 The regen-

eration processes of NaBH4 are not clean processes. The by-

products MgO and/or SiO2 have to be separated and regen-

erated. However, all these regeneration processes are far away

from any industrial application. From all these disadvantages

it seems that NaBH4 is an interesting hydrogen carrier materi-

al for very special applications. But the high price and the

impossibility of an on-board rehydrogenation of the material

excludes the material for automotive applications. One other

interesting application of NaBH4–water solution is the usage

and development of the direct boron hydride fuel cell, where

the direct anodic oxidation of borohydride provides a more

negative potential than a H2 PEM fuel cell.51 However, there

are still ongoing research activities in this field.

4.3 Complex aluminium hydrides

Complex aluminium hydrides are attractive materials for

hydrogen storage because of their high hydrogen content,

reaching more than 10 wt% in LiAlH4 on a materials basis.

Complex aluminium hydrides are used in thermal decomposi-

tion reactions for the production of hydrogen. None of the

complex aluminium hydrides can be used in water solutions

2650 | Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 This journal is �c the Owner Societies 2007

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

like NaBH4, because all these materials react vigorously with

water and other protic solvents.

Most of the complex aluminium hydrides decompose in a

three step mechanism as shown for sodium aluminium hydride

in eqn (3). During the decomposition, Al metal and the

hexahydride Na3AlH6 as intermediate product are observed.

In the case of NaAlH4 the temperature of the second decom-

position step is too high for any usage in combination with

low-temperature PEM fuel cells with a working temperature

around 353 K. This reduces the usable hydrogen content to 3.6

wt% for the NaAlH4. Sometimes, like for Mg(AlH4)2 the

intermediate product cannot be observed and the product of

the decomposition is the alkaline or alkaline earth metal

hydride.

Apart from the hydrogen content of the complex systems

the thermodynamics of these materials is more important.

Unfortunately most of the known complex aluminium hy-

drides have unfavourable thermodynamic properties for re-

versible de- and rehydrogenation. For example, the first

decomposition step from the tetrahedral LiAlH4 to the octa-

hedral Li3AlH6 complex hydride is exothermic, which means

that LiAlH4 is a thermodynamically metastable material at

room temperature with an equilibrium pressure far away from

technical conditions.52,53 Because the decomposition is kineti-

cally hindered, LiAlH4 can be handled under normal condi-

tions. It has been shown that LiAlH4 decomposes slowly at

room temperature or immediately under the influence of a

catalyst.54 The same reasons excluded Mg(AlH4)2 as a rever-

sible hydrogen storage material. It decomposes in a one step

reaction to MgH2 and Al releasing the theoretical amount of 7

wt% of hydrogen. The decomposition enthalpy of this materi-

al is in the range of 0 kJ mol�1 Mg(AlH4)2 which again is much

too low for a reversible material.55–57

3NaAlH4 !Na3AlH6 þ 2Alþ 3H2

Na3AlH6 ! 3NaHþAlþ 1:5H2

3NaH! 3Naþ 1:5H2

ð3Þ

At present NaAlH4 is the only important reversible complex

metal hydride with high hydrogen content (5.6 wt% for the

first two decomposition steps), a decomposition temperature

near the working temperature of a PEM fuel cell and potential

low prices for the compounds NaH and Al metal, which are

the basic materials for the preparation of NaAlH4. To improve

the kinetics of the decomposition and the more important

rehydrogenation the reaction must be catalysed by addition of

titanium compounds. Starting from NaH, Al and TiCl3 as

catalyst and ball-milling under hydrogen pressure, refilling

times lower than 10 min for more than 90% of the hydrogen

content can be achieved.58 But temperatures of 403 K and

10 MPa of hydrogen are necessary. With other transition

metal chlorides e.g. ScCl3 or CeCl3 higher reversible hydrogen

contents and lower rehydrogenation times can be reached.59,60

But in all cases catalyst amounts in the range of 2–4 mol% and

hydrogen pressures around 10 MPa are necessary.

In combination with a PEM fuel cell NaAlH4 can release

hydrogen only from the first decomposition step, when the

waste heat of the fuel cell is used for heating up the alanate

tank. To overcome this problem, high-temperature PEM fuel

cells (i.e. based on polybenzimidazole membranes) could be

used.61 If such a novel membrane could be operated at

temperatures up to 473 K (and would show the same 353 K

performance values of current Nafion-related membranes), the

excess heat from the fuel cell could be used to generate

hydrogen from material from both decomposition steps of

the NaAlH4 material. At a working temperature of 423 K the

dissociation pressure of the first step is 6 MPa and 0.3 MPa for

the second step.62 These values would be high enough to

provide a sufficient supply pressure for the fuel cell system.

The other main challenge is the heat release during the filling

process of a deloaded material as stated in the introduction of

this chapter.

4.4 Lithium amide/imide

Other interesting hydrogen storage systems with high hydro-

gen content are the amide/imide compounds of the light

elements. Starting from Li3N this material can store two moles

of hydrogen with a theoretical capacity of more than 10 wt%

on a materials basis (eqn (4)).63

Li3Nþ 2H2 ! Li2 NHþ LiHþH2 ! LiNH2 þ 2LiH ð4Þ

Over a temperature range from 373 up to 523 K 9.3 wt% of

hydrogen can be achieved. But the equilibrium pressure of this

material is very low with only 0.15 MPa at a temperature of

528 K. During the last few years extensive investigations have

shown that the amide, hydride and composition of the mixture

can be varied over a wide range of elements and compositions.

One other interesting example is the ternary Li–Mg–N–H

system.64,65 According to eqn (5) this system can reversible

desorb and absorb 2 mol of hydrogen, with a theoretical

hydrogen storage capacity of 5.5 wt%. In reality, hydrogen

storage capacities around 4.5 wt% at temperatures up to 473

K for the reloading and deloading process are observed.

MgðNH2Þ2 þ 2LiH ! Li2MgN2H2 þ 2H2 ð5Þ

The equilibrium pressure was determined from a van’t Hoff

plot to be 0.1 MPa close to 360 K. This is much closer to

technical requirements, but it does not fit with the working

conditions of a conventional state-of-the-art PEM fuel cell

which needs a hydrogen pressure of approximately 0.5 MPa.

One important issue related to applications of the amide/

imide system in combination with a fuel cell is the production

of small amounts of ammonia during the decomposition step.

Ammonia reacts with the acid groups of the fuel cell mem-

brane and therefore reduces the proton conductivity of the

system.66 On the other hand, a fuel cell system is not very

tolerant towards ammonia contaminations since the pH value

of the fuel cell is shifted and therefore its operating conditions

are altered unfavourably. The order of magnitude of the

ammonia impurities in the desorbed H2 gas from the Li–N–H

system with ammonia was estimated via Raman spectroscopy

to be 0.1% at any temperature up to 673 K in a closed

system.67 Recently it was found that NH3 concentrations of

180 ppm at 453 K and 720 ppm at 513 K were produced during

the self-decomposition of amides formed by the reaction of the

educts in the system: 2 LiNH2 + MgH2.68

This journal is �c the Owner Societies 2007 Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 | 2651

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

4.5 Thermodynamic tailored systems

Thermodynamic tailoring is a well known chemical concept to

stabilize or destabilize hydride systems. With the addition of a

second component new reaction pathways with different ther-

modynamic properties are opened. One typical example is the

2:1 mixture of LiBH4 and MgH2, which decomposes accord-

ingly to eqn (6) to lithium hydride and magnesium boride.69

2 LiBH4 þMgH2 ! 2LiHþMgB2 þ 4H2 ð6Þ

LiBH4 contains more than 18 wt% of hydrogen on a materials

basis, but the decomposition temperature is in the range of

653 K and the laws of thermodynamics do not allow the

hydrogenation of the mixture LiH + B after a thermal

decomposition of the boron hydride under acceptable

technical conditions. With the addition of MgH2 to the

lithium borohydride the right side of the reaction is stabilized

or, in other words, LiBH4 is destabilized. The result is a

reduction of the decomposition enthalpy for the system about

25 kJ mol�1 of H2 in comparison to the pure LiBH4. The

decomposition temperature is still too high and the equili-

brium pressure still too low for a technical application. This

example shows in principle how the thermodynamics of a

metal hydride system can be influenced with the addition of a

second compound. Once again the heat effect is still important

for a thermodynamically tailored system and the same heat as

that in alanates is released if such a system will meet the

technical requirements.

4.6 Aluminium hydride, AlH3

A thermodynamic unstable but kinetically stabilized hydrogen

storage material is AlH3 with a hydrogen storage capacity of

10 wt% on a materials basis. This is twice the storage capacity

of Ti-doped NaAlH4. Above 373 K the H2 evolution of

undoped AlH3 is high enough to exceed the flow target for a

50 kW fuel cell.70 But the main problem of AlH3 is the

regeneration of the material. AlH3 cannot be prepared in a

direct synthesis starting from Al metal and hydrogen under

reasonable technical conditions. Regeneration must be done

outside the tank system using conventional organometallic

syntheses. Such reaction pathways are expensive and excluded

AlH3 for wide use in automotive applications.

4.7 Conclusion: hydrides

For all known solid state materials it must be noted that the

heat production during the refilling process poses a severe

challenge that due to thermodynamic restrictions may be

difficult to be overcome. This heat depends only on the

amount of stored hydrogen and is from a first approximation

independent of the solid material. A lower heat amount is

combined with a more unstable metal hydride system and as a

result the operating pressure inside the tank system must be

much higher.

The ‘‘right’’ solid state material for hydrogen storage for the

automotive application is unknown to date. The target mate-

rial has to offer better properties in most (but not in all)

categories than the competing CGH2 technology. In particu-

lar, a higher volumetric density on a systems level and a lower

operating pressure compared to 70 MPa would be highly

anticipated. For different applications and requirements, a

large variety of solid state materials are available. Irreversible

materials, like NaBH4, which produce hydrogen through

hydrolysis reactions have their uses in niche applications. They

will probably not be used in large-scale automotive applica-

tions due to the infrastructure and recycling implications. In

this case, reversible storage compounds are the materials of

choice. Apart from the more than 30 years old classical metal

hydrides, none of the systems with higher hydrogen content

have reached maturity for commercial applications. Ongoing

research on the known systems and the search for new

materials as well as engineering aspects are important to bring

solid state hydrogen storage materials into the market.

5. Summary

It seems that currently none of the different storage solid state

materials can reach the required storage densities for a fuel-cell

powered vehicle. The state-of-the-art 70 MPa CGH2 technol-

ogy has been established as the benchmark by the automotive

industry. The development of storage systems which combine

chemical and physical methods, so-called hybrid approaches

(i.e. the combination of a classical hydride with a 35 MPa

pressure vessel), are potential solutions. What are the lessons

to be learned from the properties of the known material classes

and are therefore the objectives for future research:

(1) Heat of formation has to be reduced to as low as

thermodynamically possible.

(2) Operating temperature should be limited to 343 K.

(3) Operating pressure should be limited to values less than

5 MPa for cryogenic temperatures or elevated temperatures

(up to 343 K).

(4) Operating pressure should be less than 35 MPa for

room-temperature applications using low DH hydrides.

These points should be used as orientation values for any

breakthrough materials. If such a target material could be

discovered, it would simplify the automotive packaging chal-

lenges significantly, especially when addressing an optimized

trade-off between the integration of the storage system into an

existing mass-production architecture and the consideration of

a purpose-built vehicle optimized for hydrogen as a fuel.

References

1 U. Eberle, G. Arnold and R. von Helmolt, J. Power Sources, 2006,154, 456–460.

2 Grand Challenge for Basic and Applied Research in HydrogenStorage 2003 (http://www.eere.energy.gov/hydrogenandfuelcells/pdfs/1_milliken_final.pdf).

3 Quantum Technologies, Inc., Irvine.4 SRA Strategic Research Agenda of the European Hydrogen andFuel Cell Technology Platform (https://www.hfpeurope.org/).

5 R. von Helmolt and U. Eberle, J. Power Sources, 2007, 165,833–843.

6 J. Zhang, T. S. Fisher, P. V. Ramachandran, J. G. Gore and I.Mudawar, J. Heat Transfer, 2005, 127, 1391–1399.

7 K. M. Thomas, Catal. Today, 2007, 120, 389–398.8 W. B. Leung, N. H. March and H. Motz, Phys. Lett. A, 1976, 56,425–426.

9 A. Zuttel, Naturwissenschaften, 2004, 91, 157–172.10 S. Brunauer, P. H. Emmett and E. Teller, J. Am. Chem. Soc., 1938,

60, 309–319.11 L. Zhou, Renew. Sust. Energ. Rev., 2005, 9, 395–408.

2652 | Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 This journal is �c the Owner Societies 2007

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online

12 D. K. Ross, Vacuum, 2006, 80, 1084–1089.13 R. Strobel, J. Garche, P. T. Moseley, L. Lorissen and G. Wolf, J.

Power Sources, 2006, 159, 781–801.14 A. C. Dillon, K. M. Jones, T. A. Bekkedahl, C. H. Klang, D. S.

Bethune, C. H. Klang, D. S. Bethune and M. J. Heben, Nature,1997, 386, 377–379.

15 Y. Ye, C. C. Ahn, C. Witham, B. Fultz, J. Liu, A. G. Rinzler, D.Colbert, K. A. Smith and R. E. Smally, Appl. Phys. Lett., 1999, 74,2307–2309.

16 M. Becher, M. Haluska, M. Hirscher, A. Quintel, V. Skakalova, U.Dettlaff-Weglikovska, X. Chen, M. Hulman, Y. Choi, S. Roth, V.Meregalli, M. Parinello, R. Strobel, L. Jorissen, M. M. Kappes, J.Fink, A. Zuttel, I. Stepanek and P. Bernier, C. R. Physique, 2003,4, 1055–1062.

17 A. Chambers, C. Park, R. T. K. Baker and N. M. Rodriguez, J.Phys. Chem. B, 1998, 102, 4253–4256.

18 C. Park, P. E. Anderson, A. Chambers, C. D. Tan, R. Hidalgo andN. M. Rodriguez, J. Phys. Chem. B, 1999, 103, 10572–10581.

19 M. Hirscher, M. Becher, M. Haluska, U. Dettlaff-Weglikowska, A.Quintel, G. S. Duesberg, Y. M. Choi, P. Downes, M. Hulman, S.Roth, I. Stepanek and P. Bernier, Appl. Phys., 2001, A72,129–132.

20 M. Hirscher, M. Becher, M. Haluska, A. Quintel, V. Skakalova, Y.M. Choi, U. Dettlaff-Weglikowska, S. Roth, I. Stepanek, P.Bernier, A. Leonhardt and J. Fink, J. Alloys Compd., 2002,330–332, 654–658.

21 M. Hirscher, M. Becher, M. Haluska, F. v. Zeppelin, X. H. Chen,U. Dettlaff-Weglikowska and S. Roth, J. Alloys Compd., 2003,356–357, 433–437.

22 M. Hirscher and M. Becher, J. Nanosci. Nanotechnol., 2003, 3,3–17.

23 Y. P. Zhou, K. Feng, Y. Sun and L. Zhou, Chem. Phys. Lett.,2003, 380, 526–529.

24 S. Orimo, A. Zuttel, L. Schlapbach, G. Majer, T. Fukunaga and H.Fujii, J. Alloys Compd., 2003, 356–357, 716–719.

25 G. G. Tibbett, G. P. Meisner and C. H. Olk, Carbon, 2001, 39,2291–2301.

26 A. Anson, M. Benham, J. Jagiello, M. A. Callejas, A. M. Benito,W. K. Maser, A. Zuttel, P. Sudan and M. T. Martinez, Nanotech-nology, 2005, 15, 1503–1508.

27 A. Zuttel, P. Sudan, P. Mauron, T. Kiyobayashi, C. Emmeneggerand L. Schlapbach, Int. J. Hydrogen Storage, 2002, 27, 203–212.

28 M. G. Nijkamp, J. E. M. J. Raaymakers, A. J. van Dillen and K. P.de Jong, Appl. Phys. A, 2001, 72, 619–623.

29 Y. Gogotsi, R. K. Dash, G. Yushin, T. Yildirim, G. Laudisio andJ. E. Fischer, J. Am. Chem. Soc., 2005, 127, 16006–16007.

30 B. F. Hoskins and R. Robson, J. Am. Chem. Soc., 1989, 111,5962–5964.

31 O. M. Yaghi, M. O’Keeffe, N. W. Ockwig, H. K. Chae, M.Eddaoudi and J. Kim, Nature, 2003, 423, 705–714.

32 J. Y. Lee, L. Pan, S. R. Kelly, J. Jagiello, T. J. Emge and J. Li, Adv.Mater., 2005, 17, 2703.

33 B. Chen, N. W. Ockwig, A. R. Millward, D. S. Contreras and O.M. Yaghi, Angew. Chem., Int. Ed., 2005, 44, 4745–4749.

34 B. Panella, M. Hirscher, H. Puttner and U. Muller, Adv. Funct.Mater., 2006, 16, 520–524.

35 H. K. Chae, D. Y. Siberio-Perez, J. Kim, Y. B. Go, M. Eddaoudi,A. J. Matzger, M. O’Keeffe and O. M. Yaghi, Nature, 2004, 427,523–527.

36 A. G. Wong-Foy, A. J. Matzger and O. M. Yaghi, J. Am. Chem.Soc., 2006, 128, 3494–3495.

37 J. G. Vitillo, G. Ricchiardi, G. Spoto and A. Zecchina, Phys.Chem. Chem. Phys., 2005, 7, 3948–3954.

38 D. Fraenkel and J. Shabtai, J. Am. Chem. Soc., 1977, 99,7074–7076.

39 D. Fraenkel, J. Chem. Soc., Faraday Trans., 1981, 77, 2029–2039.40 J. H. Yoon and N. H. Heo, J. Phys. Chem., 1992, 96, 4997–5000.41 J. Weitkamp, M. Fritz and S. Ernst, Int. J. Hydrogen Energy, 1995,

20, 967–970.42 L. Regli, A. Zecchina, J. G. Vitillo, D. Cocina, G. Spoto, C.

Lamberti, K. P. Lillerud, U. Olsbye and S. Bordiga, Phys. Chem.Chem. Phys., 2005, 7, 3197–3203.

43 H. W. Langmi, D. Book, A. Walton, S. R. Johnson, M. M. Al-Mamouri, J. D. Speight, P. P. Edwards, I. R. Harris and P. A.Anderson, J. Alloys Compd., 2005, 404–406, 637–641.

44 Y. W. Li and R. T. Yang, J. Phys. Chem. B, 2006, 110,17175–17181.

45 B. Bogdanovic and M. Schwickardi, J. Alloys Compd., 1997,253–254, 1–9.

46 S. C. Amendola, S. L. Sharp-Goldman, M. Saleem Janjua, N. C.Spencer, M. T. Kelly, P. J. Petillo and M. Binder, Int. J. HydrogenEnergy, 2000, 25, 969–975.

47 J.-H. Wee, J. Power Sources, 2006, 155, 329–339.48 Y. Shang and R. Chen, Energy Fuels, 2006, 20, 2142–2148.49 Y. Kojima and T. Haga, Int. J. Hydrogen Energy, 2003, 28,

989–993.50 Z. P. Li, N. Morigazaki, B. H. Liu and S. Suda, J. Alloys Compd.,

2003, 349, 232–236.51 Z. P. Li, B. H. Liu, K. Arai and S. Suda, J. Alloys Compd., 2005,

404–406, 648–652.52 J. A. Dilts and E. C. Ashby, Inorg. Chem., 1972, 11, 1230–1236.53 P. Claudy, B. Bonnetot, J. M. Lettoffe and G. Turck, Thermochim.

Acta, 1978, 27, 213–221.54 T. N. Dymova, D. P. Aleksandrov, V. N. Konoplev, T. A. Silina

and A. S. Sizareva, Russ. J. Coord. Chem., 1994, 20, 279–285.55 P. Claudy, B. Bonnetot and J. M. Lettoffe, J. Therm. Anal., 1979,

15, 119–128.56 M. Mamatha, B. Bogdanovic, M. Felderhoff, A. Pommerin, W.

Schmidt, F. Schuth and C. Weidenthaler, J. Alloys Compd., 2006,407, 78–86.

57 Y. Kim, E.-K. Lee, J.-H. Shim, Y. W. Cho and K. B. Yoon, J.Alloys Compd., 2006, 422, 283–287.

58 J. M. Bellosta von Colbe, M. Felderhoff, B. Bogdanovic, F. Schuthand C. Weidenthaler, Chem. Commun., 2005, 4732–4734.

59 B. Bogdanovic, M. Felderhoff, A. Pommerin, F. Schuth and N.Spielkamp, Adv. Mater., 2006, 18, 1198–1201.

60 T. Wang, J. Wang, A. D. Ebner and J. A. Ritter, J. Alloys Compd.,2006, DOI: 10.1016/j.jallcom.2006.10.072.

61 J. O. Jensen, Q. Li, R. He. C. Pan and N. J. Bjerrum, J. AlloysCompd., 2005, 404–406, 653–656.

62 B. Bogdanovic, R. A. Brand, A. Marjanovic, M. Schwickardi andJ. Tolle, J. Alloys Compd., 2000, 302, 36–58.

63 P. Cheng, Z. Xiong, J. Luo, J. Lin and K. L. Tan, Nature, 2002,420, 302–304.

64 Y. Nakamori and S. Orimo, Mater. Sci. Eng., B, 2004, 108, 48–50.65 Z. Xiong, J. Hu, G. Wu, P. Chen, W. Luo, K. Gross and J. Wang,

J. Alloys Compd., 2005, 398, 235–239.66 F. A. Uribe, S. Gottesfeld and T. A. Zawodzinski, J. Electrochem.

Soc., 2002, 149, A293–A296.67 S. Hino, T. Ichikawa, N. Ogita, M. Udagawa and H. Fujii, Chem.

Commun., 2005, 3038–3040.68 W. Luo and K. Stewart, J. Alloys Compd., 2006, DOI: 10.1016/

j.jallcom.2006.09.057.69 J. J. Vajo, S. L. Skeith and F. Mertens, J. Phys. Chem. B, 2005,

109, 3719–3722.70 J. Graetz and J. J. Reilly, Scr. Mater., 2007, 56, 835–839.

This journal is �c the Owner Societies 2007 Phys. Chem. Chem. Phys., 2007, 9, 2643–2653 | 2653

Publ

ishe

d on

09

May

200

7. D

ownl

oade

d by

New

Yor

k U

nive

rsity

on

07/1

0/20

14 2

2:06

:54.

View Article Online