Você está na página 1de 16

7.

17

Ammonia Synthesis: State of the Bellwether Reaction

A Hellman, Chalmers University of Technology, Gothenburg, Sweden


K Honkala, University of Jyvaskyla, Jyvaskyla, Finland
S Dahl, Technical University of Denmark, Kgs. Lyngby, Denmark
CH Christensen, DONG Energy, Fredericia, Denmark
JK Nrskov, SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, Menlo Park, CA, USA;
Stanford University, Stanford, CA, USA
2013 Elsevier Ltd. All rights reserved.

7.17.1
7.17.2
7.17.3
7.17.4
7.17.4.1
7.17.5
7.17.6
7.17.7
References

7.17.1

Introduction
The Ammonia Synthesis Process
Reaction Mechanism, Active Sites, and Reaction Kinetics
Micro-Kinetic Models of Catalytic Ammonia Synthesis
Activity Trends
Discovery of New Catalysts for Ammonia Synthesis
New Perspectives in Ammonia Synthesis
Conclusion

Introduction

It has been argued that ammonia synthesis from the elements is


among the most important discoveries of the last century. What
is the most important invention of the twentieth century? Aeroplanes, nuclear energy, space flight, television, and computers
will be the most common answers. Yet none of these can match
the synthesis of ammonia from its elements.1 The industrial
HaberBosch process enables the mass production of fertilizers
that is needed to feed the continuously growing population of
the Earth (Figure 1). Ammonia production currently consumes
more than 1% of all energy produced globally.2 This is noteworthy because the energy efficiency of ammonia synthesis has
been improved dramatically since its discovery. Today, the
majority of all ammonia produced in the world is used directly
in the fertilizer industry, but new applications also are emerging,
such as the use of ammonia in the removal of nitrogen oxides
from exhaust gases by selective catalytic reduction (SCR). Still,
the volume of such applications is limited and, within the
foreseeable future, ammonia production is expected to follow
population growth unless new breakthrough applications are
developed, for instance, the use of ammonia as an energy
carrier.3
Any method for ammonia synthesis must address the necessary activation of the extraordinarily strong triple bond in the N2
molecule. For instance, N2 activation in the gas phase requires
extreme conditions. However, such a method for nitrogen fixation was actually commercialized during the early twentieth
century, but its energy consumption was enormous.4 Nature
relies on an alternative route, where enzymes catalyze the stepwise hydrogenation of the nitrogen molecule, which allows for
nitrogen fixation at ambient conditions.5,6 Mankind, however,
relies exclusively on the HaberBosch process for the supply of
ammonia not covered by the enzyme process. Today, the energy
consumption for this process is remarkably low compared with
other means of ammonia production, and the technology is well
suited for large-scale manufacturing.7,8

Comprehensive Inorganic Chemistry II

459
460
462
465
468
470
471
473
473

The fundamentals of the technology behind the Haber


Bosch process were originally developed by Fritz Haber9 and
Carl Bosch10 almost a century ago. First, Haber determined the
thermodynamic equilibrium of ammonia under atmospheric
pressure and high temperature, 1000  C, with the help of an
iron (Fe) catalyst.11 Even though the amount of ammonia
produced was small, its success showed that ammonia could
be produced directly from gaseous N2 and H2. Later, Haber
suggested more optimal operational conditions: pressures
around 150200 atm and temperatures around 500  C. Following this development, Bosch solved several technical problems related to operating a chemical plant at these extreme
conditions and, within his team, created the technology
required to commercialize ammonia production. The scientific
community soon recognized their pioneering work. In 1918,
Fritz Haber received The Nobel Prize in Chemistry for the
synthesis of ammonia from its elements, and in 1931, Carl
Bosch was awarded The Nobel Prize in Chemistry in recognition of his contributions to the invention and development of
chemical high pressure methods. Hence, the developments in
physical chemistry and chemical engineering in the early part
of the last century played a crucial role in the development of
the HaberBosch process, and vice versa.
During the period 19091912, Alwin Mittasch12 conducted, for the first time ever, a large-scale screening experiment to find a substitute to Habers more exotic osmium- and
uranium-based catalysts preferentially employed at that time.
In all, some 3000 catalyst compositions were tested in about
20 000 small-scale experiments. Eventually, he arrived at an
Fe-based catalyst, which actually has a composition very similar to that of the catalyst used industrially today.12 Although in
the same experiments ruthenium (Ru) was found to be an
interesting catalyst candidate,13 it was not until the 1970s14,15
that Ru was widely recognized as the best elementary-metal
catalyst for ammonia synthesis.16,17 Today, it is known that, at
low temperature and close to the thermodynamic equilibrium,
Ru exhibits considerably higher activity than Fe.16,1821

http://dx.doi.org/10.1016/B978-0-08-097774-4.00725-7

459

460

Ammonia Synthesis: State of the Bellwether Reaction

7.17.2

100

The Ammonia Synthesis Process

80

5
4

60

3
40
2
20

1
0
1900

Consumption of nitrogen fertilizer


(megatons of nitrogen)

World population (billions)

0
1925

1950
Year

1975

Figure 1 The sudden growth in the global consumption of nitrogen


fertilizer during the twentieth century was caused by the increase in the
world population. The strong correlation between the two indicates that
the HaberBosch process is of paramount importance for increasing
the population. Reproduced from V. Smil, Nature 1999, 400, 415, with
permission Copyright 1999 Scientific American, Inc. All rights
reserved.

However, because of the higher price of Ru and a shorter


catalyst lifetime, the dominance of iron-based catalysts has
only recently been challenged by promoted Ru catalysts2224
in a few installations.
A simple view of heterogeneous catalysis is that the catalyst
provides a surface onto which the reactants temporarily adsorb
and, during this process, some intramolecular bonds are weakened sufficiently for old bonds to be broken and new bonds to
be created. The bonds between the reaction intermediates and
the catalyst must be weak enough so that the products can be
released. In the HaberBosch process, reactant gas molecules N2
and H2 are dissociated and adsorbed onto well-dispersed, nanosized metal particles, and ammonia is formed by the stepwise
hydrogenation of nitrogen atoms on the catalyst surface. The
catalytic cycle is closed by the desorption of ammonia. The relative simplicity and importance of ammonia synthesis have made
it the textbook example or the bellwether reaction in heterogeneous catalysis. As such, it is commonly used as a test reaction to
develop new concepts and ideas (see also Chapter 7.01).2529
The ultimate goal of catalysis research is to design and tune
the activity, selectivity, and lifetime of catalysts by controlling
their structural properties at the atomic level. The identification
of concepts to achieve this goal is still one of the key issues of
research in catalysis. A variety of different strategies have been
put forward, most of which are based on structurereactivity
relationships, taking different aspects of these exceedingly
complex systems into account.
In the historical development of surface science and the
goal of atomic understanding of heterogeneous catalysis,
ammonia synthesis also has played a central role.25,30 This
development has allowed the design of new ammonia catalysts
based on atomistic insight20 and, lately, made it possible to
understand and quantitatively describe ammonia synthesis
from first-principle calculations.31

In the HaberBosch process, ammonia is synthesized from a gas


containing H2 and N2 at approximately stoichiometric ratio.
N2 3H2 P 2NH3  99:22 kJ mole1
Importantly, ammonia synthesis is an equilibrium-limited
exothermic reaction at relevant reaction conditions, and it is
required that the synthesis gas must be free of any reactive
contaminants that poison the catalyst. It is mainly these
requirements that determine the optimal process design of an
ammonia synthesis plant and an ammonia reactor.
A modern implementation of the HaberBosch process
provides a highly efficient method for producing ammonia in
large quantities. There are many slightly different process
designs,32 all of which are based on synthesis gas production
from a carbonaceous feedstock, steam, and air, followed by
ammonia synthesis in a recycle loop. A standard process
scheme utilizing natural gas as feedstock is shown schematically in Figure 2.
The first step in the process is desulphurization of the
natural gas feed in order to avoid sulfur poisoning of the
downstream catalysts. The second step is steam reforming,
where the natural gas is reacted with steam to produce an
equilibrium mixture of H2, CO, CO2, and CH4. This reaction
is endothermic, so it is carried out in tubular reactors placed in
a heated furnace in order to supply the heat for the reaction
and to maximize the equilibrium content of the desired products H2 and CO. Following this, air is added to supply the N2
required for ammonia production, while the O2 from the air
converts the remaining CH4 in an exothermic reaction that
increases the temperature and the H2 and CO content further.
The reaction mixture is then cooled and followed by a two-step
watergas shift process, where CO is reacted with steam and
transformed into H2 and CO2. Since all oxygen-containing
molecules poison the ammonia synthesis catalyst, CO2 is subsequently removed by absorption and a final CO/CO2 cleanup
is carried out, for example, by the methanation reaction.
Before sending the synthesis gas into the ammonia reactor,
it is compressed and, at a point before it reaches the ammonia
catalyst, water is removed by condensation. As mentioned
above, ammonia synthesis is an equilibrium-limited exothermic reaction, and the ammonia product is favored by low
temperature and high pressure. The high reaction temperature
necessary for obtaining sufficiently high reaction rate, together
with constraints on the synthesis pressure, excludes full conversion of the synthesis gas.32 This means that most of the N2
and H2 passes unreacted through the reactor, and must be
recycled in the loop after being cleaned for ammonia. Since
the synthesis gas that is fed into the loop contains some inert
gases (CH4 and Ar), a purge is also needed in the loop to avoid
too large a build-up of these diluents.
Optimal heat management is important for obtaining an
energy-efficient ammonia synthesis process. The coupling
between the exothermic process steps (secondary reforming,
watergas shift, ammonia synthesis, and cooling steps) and the
endothermic process steps (steam reforming water evaporation
and heating steps) is achieved with suitable heat exchangers
and proper steam generation and usage. This integration of

Ammonia Synthesis: State of the Bellwether Reaction

461

optimal operating line, occasionally also called the maximum


rate line, illustrates the temperatures at which the maximum rate
is reached at given pressure and gas composition. Different synthesis loop configurations have been proposed to make the
actual operation line follow the optimal operating line as closely
as possible. Figure 3 shows how it is done in an internally cooled
reactor with a countercurrent flow of synthesis gas in the cooling
tubes or in an indirectly cooled reactor, where the gas is cooled
by heat exchange between the individual beds. The resulting
operating lines are also shown. These two reactor designs have
been extensively used in the ammonia synthesis industry.
The Tennessee Valley Authority reactor that was common from
1930 to 1965 typifies the internally cooled reactor. Today, most
new industrial ammonia synthesis reactors are adiabatic two- or
three-bed radial-flow arrangements with indirect cooling.

the processes, together with the high-pressure conditions


(100200 bar) for the HaberBosch process, means that very
large plants are most efficient and that the huge global ammonia production is predominantly carried out in a few hundred
centralized plants.
When designing the high-pressure ammonia reactor, an
important criterion is to keep the reactor and catalyst volume
small. As the temperature in an adiabatic reactor will increase as
ammonia is synthesized by 1418  C for each percent of ammonia produced,2,34 some kind of cooled reactor must be employed
to reach a reasonable ammonia concentration. To illustrate the
reactor design process, the so-called optimal operation lines are
an essential aid. Using only a single catalyst, it is well known that
the minimum catalyst volume will be obtained in a reactor where
the optimal operating line is followed as closely as possible. The

CO2
Air

H2O

450
500 C

0 C

Natural gas
400
500 C

300 C

350 C

NH3

300 C

1200 C

KMR
KM

400 C
200 C

1000 C

800 C

325 C

100 C

220 C

Desulphurization

Primary
reforming tubular

Secondary
reforming

Shift

Purge
synthesis

Methanator

% NH3

% NH3

Figure 2 Process scheme of ammonia synthesis. (Courtesy of Haldor Topsoe AS.)

350

450
550
Temperature (C)

350

450
550
Temperature (C)

Figure 3 The schematic picture of an internally cooled reactor and of an adiabatic three-bed reactor, cooled by two heat exchangers. The equilibrium
and operational lines are shown in the ammonia concentration versus temperature plot. The corresponding operation lines for that reaction in that
particular reactor is also shown. Reprinted with permission from Jacobsen, C. J. H.; Dahl, S.; Boisen, A.; Clausen, B. S.; Topse, H.; Logadottir, A.;
Nrskov, J. K., J. Catal. 2002, 205, 382387.

Ammonia Synthesis: State of the Bellwether Reaction

In such reactors, the ammonia concentration and the temperature vary in the two (or three) beds, as shown in Figure 3. The
position of the optimal operation lines depends on the type of
catalyst; however, the differences are small and the lines are
generally positioned 3060  C below the equilibrium curve.
The absolute reaction rate is much more dependent on the
catalyst of choice and, in Section 7.17.5, we will discuss how
the reactor volume can also be minimized by choosing different
catalysts in different parts of the reactor.

7.17.3 Reaction Mechanism, Active Sites, and


Reaction Kinetics
In heterogeneously catalyzed reactions, like ammonia synthesis,
the catalytic action comes from the interaction of reactants and
reaction intermediates with the active surface sites of a solid
catalyst. The direct link between activity and the number of active
sites and the output of products from a catalyst has been a spur to
establish an understanding of surface reactions at the atomic
level.35 However, it has taken more than half a century for the
surface-science community to develop techniques that can routinely probe the atomic processes of interest.36,37 Throughout the
1960s to the early 1990s, thousands of well-defined single-crystal
surface structures were studied with various surface-science
techniques, such as low-energy electron diffraction, Auger electron spectroscopy, and x-ray photoelectron spectroscopy (XPS).
These studies were followed by experimental and theoretical
investigations of the nature of the chemical bonding of adsorbates at surfaces and interfaces. Detailed electronic structures of
surface chemical bonds were obtained by spectroscopic techniques, such as ultraviolet and XPS, and x-ray emission spectroscopy. First-principles calculations helped to rationalize the
experimental findings and to provide a consistent framework
for understanding surface chemical bonding.38,39 For example,
the d-band model correlates the binding strength of an adsorbate
to a transition-metal surface and the local d-electron density of
the binding site.28 Along with the experimental data, the d-band
model provides answers to fundamental questions such as why
bulk gold is noble while small gold clusters are not,40,41 and why
surface defect sites (steps and kinks) are more chemically active
than sites of higher coordination on flat surfaces.42
Long before surface-science experiments4348 provided direct
observations of the reaction intermediates, the mechanism of
ammonia synthesis was basically resolved. Already as early as
1933, Emmett et al.49,50 found strong experimental evidence that
N2 dissociation over Fe catalysts is the rate-limiting step for
ammonia synthesis under industrially relevant conditions.
About 30 years later, surface-science studies revealed a detailed
picture of the N2 dissociation process.4548,51,52 Following the
identification of the rate-determining step, the connection
between ammonia synthesis at elevated temperatures and pressures and ultra-high vacuum surface-science measurements on
iron-based catalysts (single-crystal surfaces) was established both
experimentally4548,5355 and theoretically (Figure 4).25,5659
The first surface-science studies on ammonia synthesis
focused on N2 dissociation and subsequent ammonia formation over various Fe crystal surfaces at low pressures. Based
on these results, it was established that the dissociation reaction is structure sensitive and the open Fe(111) and Fe(211)

0.5N2 + 1.5H2

17

0
Potential energy/kJ mol-1

462

Fe
~21

46

0.5N2-* + 1.5H2

50

-100

NH3
~41

~33 NH -*
3

-200

NH2-* + H-*

259
106

-300

N-* + 3H-*

NH-* + 2H-*

Figure 4 The reaction energy diagram for ammonia synthesis on an


iron surface as inferred from surface-science measurements. Energies
are in kJ mol1. Reprinted with permission from Hinrichsen, O. Catal.
Today 1999, 53, 177188.

surfaces are the most active.61 This was attributed to the access
to the so-called C7 active site (surface atoms with seven
nearest neighbors, Figure 5). Activity measurements at high
pressure on similar single-crystal surfaces showed that Fe
crystal faces with C7 sites have significantly higher catalytic
activity for ammonia synthesis than other Fe crystal faces
(Figure 5).54 These findings, together with the investigations
of elementary reaction steps45,62 and the effects of catalyst
support and additives,55,62 were important in establishing
an understanding of industrial ammonia synthesis based on
surface science.
First-principles studies63,64 show that the adsorption energy
of N on Fe(100) is higher compared to the more close-packed
(111) and (110) surfaces. Especially, the adsorbed phase c
(2 2)-N/Fe(100) is particularly stable, indicating that this
phase will form under high pressure of nitrogen. Consequently,
at higher coverage of N, the surfaces (111) and (110) might very
well reconstruct to form a stable c(22)-N/Fe(100) overlayer.
Furthermore, the dissociation of N2 on Fe(111) has, essentially,
two different pathways: one with a low energy barrier but with a
large entropic barrier, and the other with a high energy barrier.64
The low-energy-barrier path includes several precursor states,
which results in a complicated dissociation path. Under normal
ammonia synthesis conditions, the low-barrier, high-entropy
path will dominate, but at the highest temperatures, the highbarrier process may become more effective.
After studying ammonia synthesis over iron, the attention
turned to ruthenium, which is a more active catalyst than iron,
especially at high ammonia concentrations. This gave rise to
industrial interest in ruthenium-based ammonia catalysts in
the 1990s.
In the first attempts to measure the N2 dissociation rate on
different Ru surface facets, the process was found to be structure insensitive.65 As with the case of iron, it was possible to
model the activity of technical catalysts at high-pressure conditions based on the surface-science results,60 so there seemed
to be some consistency. However, there was one problem;
experimentally, the activation energy for the dissociation reaction was found to be around 0.40 eV on the Ru(0001) surface,
which disagrees with a calculated barrier of 1.36 eV.66 The
answer to this discrepancy was found in a joint experimental
and theoretical effort,67 where dissociative adsorption of N2

Ammonia Synthesis: State of the Bellwether Reaction

Figure 6(a) highlights the difference in sticking coefficient


observed on these surfaces, and Figure 6(b) compares the
calculated N2 dissociation barriers over stepped and flat surfaces. The barrier on a step site is very close to the experimentally observed value on bare Ru, which strongly supports the
interpretation that steps (or sites with similar characteristics)
dominate the N2 dissociation.
The conclusion was that, to efficiently break the strong
triple bond, Ru steps should expose a threefold hollow site
and a bridge site close together and form an active site for
dissociation. In this configuration, atoms of a nitrogen molecule do not simultaneously bind to the same Ru surface atom
at the transition state during dissociation (see the structure at
the bottom of Figure 6(b)). This has later been confirmed by
other studies.69,70 According to the terminology developed by

over a closed packed Ru(0001) surface was revisited. Measurements were done on both clean and Au-passivated surfaces. In
the latter case, small amounts of Au (12% on the surface)
were deposited on the surface, where they preferentially decorate reactive step sites. The results show that the presence of Au
reduces the N2 dissociation rate almost completely, demonstrating the high structure sensitivity of N2 dissociation on Ru.
Based on the measurements, it was concluded that the reaction
takes place solely at the step sites, with a dissociation barrier of
0.4 eV. The rate on the planar terrace was measured to be at
least nine orders of magnitude smaller than that at the steps at
500 K. The first ab initio calculations supported this and they
gave a corresponding 1.5-eV barrier difference. Laser-assisted
associative desorption of N2 from Ru(0001) also confirmed
the reactivity difference between step and terrace atoms.68

Fe(111)

Fe(210)

C4

C4

C7

C5
C8

C7

14
T = 400 C

12

20 atm 3:1 H2:N2

Mol NH3/cm2 s 10-9

463

Fe(100)

10

C4
C8

8
6

Fe(211)

Fe(110)

C5

C6

C7

2
0
(111) (211) (100) (210) (110)
Surface orientation
The first layer

The second layer

The third layer

Figure 5 Produced ammonia over different Fe facets (left) and the corresponding structures (right). Reprinted with permission from Somorjai, G. A.;
Li, Y. Proc. Natl. Acad. Sci. 2011, 108, 917924.

-10
Ru(0001)
Ea = 0.4 0.1 eV

-12

-13

TS

2.0

DE [eV/N2]

log S0

-11

1.0

0.0

N2,ad
2Nad

12% Au on Ru(0001)
Ea = 1.3 0.2 eV

-1.0

-14
0.0015

0.0020

0.0025
1/T (1/K)

0.0030

0.0035

Figure 6 (Left) An Arrhenius plot of measured thermal sticking coefficients of N2 on a clean Ru(0001) surface and the same surface covered with
0.010.02 mL of gold. (Right) Results from density-functional calculations comparing N2 dissociation on a terrace and at a step on Ru(0001). The upper
curve shows the adsorption and transition-state energies for the dissociation on the terrace, whereas the lower curve shows the same energies at the step.
The energy zero is taken to be the energy of the N2 molecule in the gas phase. Reprinted figure with permission from Dahl, S.; Logadottir, A.; Egeberg, R. C.;
Larsen, J. H.; Chorkendorff, I.; Tornqvist, E.; Nrskov, J. K. Phys. Rev. Lett. 1999, 83, 1814. Copyright 1999 by the American Physical Society.

464

Ammonia Synthesis: State of the Bellwether Reaction

Van Hardeveld and Van Montfoort,71 these active sites are


called B5 sites. As can be seen in the top part of Figure 6(b),
on the terrace, one Ru surface atom is bonded to both nitrogen
atoms, which leads to the high barrier. The higher reactivity of
step atoms compared to terrace atoms is due to the lower
coordination number and higher lying d-states, which, alone,
however, cannot explain the large difference in the activation
barrier. The size of this electronic effect can be estimated from
the difference in adsorption energies of nitrogen atoms at the
right of the energy diagrams in Figure 6(b). It is possible to
obtain a lower activation barrier for N2 dissociation at the step
sites without affecting the adsorption strength of the nitrogen
atoms, which is the reason why the steps totally dominate the
ammonia synthesis reaction. This is seen in Figure 7(a), where
the calculated potential energy diagrams for ammonia synthesis over flat and stepped Ru surfaces are shown. The comparison of rate constants presented in Figure 7(b) shows that N2
dissociation over the step edge is a rate-limiting step in ammonia synthesis, despite the fact that NH2 hydrogenation has
higher activation energy. Had the nitrogen atoms been bound
much stronger to the step sites, the hydrogenation steps would
have been more uphill in energy, making them rate limiting.
Thus, the full benefit of the low activation barrier for N2
dissociation on the step sites would not have been so large.
This enormous difference in reactivity between steps and
terraces has two consequences. First, it means that the rate of
N2 dissociation on Ru(0001) is completely dominated by as little
as a fraction of a percent of steps on the surface. The steps, thus,
dominate all thermal surface-science experiments of N2 adsorption on single-crystal surfaces. Since the number of step sites does
not vary much between different surface facets, this explains why
the reaction was initially found to be structure insensitive.
Second, it means that the ammonia synthesis reaction should
be extremely structure sensitive on Ru. For nanoparticles, the step
density depends strongly on the particle size, suggesting that the
catalytic activity of Ru catalysts for ammonia synthesis should be
dependent on the catalyst particle size.

Jacobsen et al. 73 were first to count the relative number of


B5-type sites present on Ru nanoparticles. By simple arguments,
it was estimated that the maximum probability for B5 sites was
obtained for particles in the range of 1.82.5 nm, and for larger
particles, the probability for B5 sites decreased monotonically.
For smaller particle sizes, the B5-type sites disappear, simply
because the crystals have to be of a certain size before these
sites are formed. Honkala et al. 31 calculated the surface energies
of the most relevant facets of Ru and used these to create
an atomistic Wulff construction for several Ru particles with
different diameters. After including edge effects, the B5 sites
appear along the (0001)/(1011) intersection, and an estimate
for the number of active sites per particle size can be conducted
(Figure 8). It was concluded that there exists a lower limit to
the optimal ruthenium crystal size, which could explain the
measurements74,75 for promoted and unpromoted Ru/MgO catalysts, and for promoted Ru/MgAl2O4 catalysts.73
An even more advanced estimate was presented by
Gavnholt and Schiotz.70 They took into account all cluster
sizes between 1.5 and 5 nm in diameter, and also included
the temperature dependence on the structure. The obtained
values for the number of sites are quite low, with a peak
value between 1.5 and 2.5 mol g1 at particle diameter 3.5
nm and 30 mol g1 at particle diameter 2 nm. These values
are approximately an order of magnitude lower than the estimates made by Honkala et al.31 However, other sites included
in the study by Gavnholt and Schiotz70 gave significant contributions to the overall rate. The graphs at 300 and 700 K in
Figure 9 indicate a maximum in the catalytic activity per volume catalyst at a cluster diameter of 3 nm, even though the
curve is not so smooth. At 300 K, it is clearly seen that all the
catalytic activity comes from step sites A and B, whereas step
site D starts contributing at 700 K. At 1200 K, the picture is a bit
clearer and smoother. There still seems to be an optimal cluster
diameter at 3 nm. The existence of such a maximum fits well
with the experimental observation that the catalytic activity
of a sample can increase after sintering of the smallest Ru

200
N2

100
E tot (eV/molecule)

700 K
500 C
350 C

TSterrace
3H2

-1.00

k4

NH3
0
NH3

k1
0

-2.00
-3.00

-4.00
0

TSstep

-5.00

*
*
*
*
*
*
*
3
3
6H + 5H + 4H + 3H + 3H + 2H + H NH NH
+3 +3 +3 +
+ +2
3
*
*
*
*
H
*
*
2
3
3
*
N
*
N 2 2N 2N NH NH 2 NH 3 NH NH + N H 3
*+
N
+
+
+ N *+
+
* H2*
N* N* N*
H
N
N
H2

H2

H2

0.001

0.0015

0.002

0.0025

0.003

0.0035

-1
1/T [K ]

Figure 7 (left) The comparison of potential energy landscape for ammonia formation over terrace and step sites on Ru(0001). The plot shows
that the step effect is less pronounced for the hydrogenation step, which results from the fact that transitions states are similar on both surfaces. (Right)
The comparison of the rate constants for NH2 hydrogenation and N2 dissociation over stepped Ru. Despite that the hydrogenation barrier is higher
than the dissociation barrier, N2 dissociation is by far the slowest step due to the entropic effects. The rate constants were estimated with harmonic
transition-state theory. Reprinted with permission from Logadottir, A.; Nrskov, J. K. J. Catal. 2003, 220, 273279.

465

Ammonia Synthesis: State of the Bellwether Reaction

1 nm

Step atom

2.5

Step A

step A
step B
step C
step D

Energy/eV

1.5
1
0.5
0

Step B

-0.5
-1

Small
crystal

1.5

300 K

Step C

0.1

Edge atoms
Step D

Active sites

0.01
0

4
6
8
Crystal size (nm)

10

Figure 8 A Wulff-shaped Ru particle, with an average diameter of 2.9 nm.


Atoms that belong to active B5 sites are shown in red. Also depictured is a
TEM image of a supported Ru particle with a step. The fraction of edge
atoms and active sites on small Ru crystals are shown relative to the total
number of atoms as a function of crystal size. Reprinted with permission
from Honkala, K.; Hellman, A.; Remediakis, I. N.; Logadottir, A.; Carlsson,
A.; Dahl, S.; Christensen, C. H.; Nrskov, J. K. Science 2005, 307, 555558
and Jacobsen, C. J. H.; Dahl, S.; Hansen, P. L.; Tornqvist, E.; Jensen, L.;
Topsoe, H.; Prip, D. V.; Moenshaug, P. B.; Chorkendorff, I., J. Mol. Catal. A
Chem. 2000, 163, 1926.

particles.73 Furthermore, it is seen that the step site D also


contributes to the total activity at this higher temperature,
due to fact that it has a greater presence on the clusters than
the other two and the fact that the difference in barrier height
becomes less important as the temperature increases. The
ammonia synthesis typically runs at a temperature of around
700 K in industrial plant.

7.17.4 Micro-Kinetic Models of Catalytic Ammonia


Synthesis
As described earlier, N2 dissociation is the rate-limiting step in
catalytic ammonia synthesis. The full reaction follows a LangmuirHinshelwood (LH) mechanism, where H2 is also dissociatively adsorbed on a surface before adsorbed nitrogen is
stepwise hydrogenated to form ammonia (see Figure 7). Each
reaction step obeys microscopic reversibility, and the full set
elementary reaction steps is:
H2 2* P 2H*
N2 2* P 2N*
N* H* P NH* *
NH* H* P NH2 * *
NH2 * H* P NH3 * *
NH3 * P NH3 *
where * and X* correspond to an empty site and an adsorbed
species X, respectively (see also Chapter 7.02).

2.5
3
3.5
4
Bond length / angstrom

700 K

4.5

1200 K

Activity per volume (arbitrary unit)

Ru(0001)
surface

Activity per volume (arbitrary unit)

Fraction of total atoms

D
D
B
B
B

2.0

3.0

4.0

Diameter/nm

5.0 2.0

3.0

4.0

Diameter/nm

5.0 2.0

3.0

4.0

5.0

Diameter/nm

Figure 9 Gavnholt and Schiotz70 define four different step sites.


The calculated activation barriers range between 0.5 and 2 eV.
Depending on the operational temperature, the contribution to the
ammonia production varies between the different step sites. Reprinted
figures with permission from Gavnholt, J.; Schiotz, J. Phys. Rev. B
2008, 77, 035404, Copyright 2008 by the American Physical Society.

Based on the reaction mechanism, it is possible to construct a


micro-kinetic model.76,77 Such models are extensively applied to
develop and verify an atomistic understanding of catalysis. The
success stories include industrially relevant catalytic reactions,
such as methanol synthesis78 and decomposition,79,80 water
gas shift,81 ethylene oxidation,82,83 and ammonia synthesis.84,85
These examples show that the micro-kinetic models can bridge
the temperature, pressure, and material gaps that exist between
fundamental surface-science studies and industrially relevant
catalysis (see also Chapter 7.03).
A key step in constructing a micro-kinetic model is to formulate an expression for a catalytic rate per active site based on
describing the elementary reaction steps.76,77,86 Usually, it is
necessary to make some approximations in order to obtain a
solvable model. Mean-field micro-kinetic models, where interactions between adsorbates are neglected, are, in many cases,
adequate for a quantitative description of the reaction rate. To
analyze trends for the activity of different catalysts, the meanfield models have distinct advantages, since the additional
assumptions, such as the inclusion of a rate-determining reaction step and the steady-state approximation, make the model
entirely analytical.87 However, during typical reaction conditions, the coverage of reactant, intermediate, and product adsorbates can be substantial, which can challenge the mean-field
assumption. To overcome this limitation, one can employ
extensions, such as mean-field expressions for lateral interaction
or even quasi-chemical approximations that include some local
structure information of the adsorbates.88 The other possibility
is to combine the micro-kinetic formulation with Monte Carlo
(MC) simulations of the surface coverages.31 Ammonia

466

Ammonia Synthesis: State of the Bellwether Reaction

synthesis, owing to N2 dissociation being the rate-determining


step and at equilibrium with the products, is well suited for such
a combination.31
By combining activation barriers and adsorption energies
from density functional theory (DFT) calculations72 and a
micro-kinetic model, it is possible to predict the relative reaction rates over different catalyst materials.67 This methodology
has been particularly successful for ammonia synthesis.19,31,72
The obtained results have provided the foundation for the
volcano curve16 and have, furthermore, given the basis for a
rational design approach to new catalysts. For more information, see Section 7.17.5.
Micro-kinetic models of ammonia synthesis over ironbased catalysts show a remarkable agreement with laboratory
measurements, both over single crystals and industrial
catalysts,56,58,76,86,89,90 see Figure 10(a) for an early example.
Different micro-kinetic models can give good overall agreement with experimental results; therefore, the success of a
micro-kinetic model cannot be used as proof for a particular
reaction mechanism.77,91 Because of this, it is very important
to take as many input parameters as possible from independent measurements, instead of obtaining them by fitting to the
overall rate data.
The predictive power of micro-kinetic models combined
with surface-science techniques are perhaps best exemplified
by the quantitative micro-kinetic model for a multipromoted
iron catalyst developed by Sehested et al.89 The parameters are
close to those derived from surface-science experiments and
based on Langmuiran adsorption. Here, the activity of a multipromoted iron catalyst is measured at total pressures from 1 to
100 bar and at temperatures from 320 to 440  C, with the
hydrogen to nitrogen ratio varied by a factor of 10. The predictions of the model were compared to activity measurements
conducted during the study, and also to independent measurements of Nielsen.26

The micro-kinetic modeling of ammonia synthesis is easier


on Ru than on Fe,84,85 partly due the lack of molecular precursor
states of N2. Dahl and coworkers developed a micro-kinetic
model for nonpromoted ruthenium based predominantly on
surface-science observations.84,85 The main assumptions are
that N2 dissociation is a rate-determining step, and only the
step sites of an Ru(0001) surface are active in the reaction. The
model describes very well the rates of ammonia formation over a
ruthenium single crystal and over a supported Ru-based catalyst,
as shown in Figure 11. The input parameters for adsorption and
activation energies were, to a large extent, based on surfacescience results for the relevant species on ruthenium, although
some estimates were needed for parameters that are not directly
available from measurements. The model is also in excellent
agreement with an N2 isotope scrambling experiment over an
Ru/MgO catalyst (Figure 11). The fact that the micro-kinetic
results show good agreement with many widely different experimental results strongly suggests that the model gives a fundamentally correct picture of ammonia synthesis over ruthenium.
Given the relative simplicity of the reaction mechanism on
Ru, it is natural that ammonia synthesis was the very first heterogeneous catalytic reaction that was fully characterized from first
principles. Its productivity was predicted for a supported technical catalyst with a surprisingly high level of accuracy.31,91,93 All
energy-related input parameters for the micro-kinetic model
were calculated with DFT. These include adsorption and activation energies, interaction energies, and entropy contributions.
The only input from the catalyst is the average particle size
distribution determined by high-resolution transmission electron microscopy. An MC technique was employed to determine
adsorbate coverages employing pair interactions from DFT
calculations, which takes the model beyond the ordinary
mean-field approximation. The absolute productivities are in
semi-quantitative agreement with the experimental data measured over an industrial, high-surface catalyst at various

10.0
Calculated ammonia concentration (%)

Calculated exit NH3 mole fraction

10-1

10-2

1 atm

10-3
10-3

150 atm
300 atm

10-2
10-1
Experimental exit NH3 mole fraction

1.0

0.1
0.1

1.0
Measured ammonia concentration (%)

10.0

Figure 10 Ammonia concentration predicted from two micro-kinetic models,56,89 plotted as a function of the measured ammonia concentration.
Reprinted figure with permission from Stoltze, P.; Nrskov, J. K. Phys. Rev. Lett. 1985, 55, 2502. Copyright 1985 by the American Physical Society;
Sehested, J.; Jacobsen, C. J. H.; Tornqvist, E.; Rokni, S.; Stoltze, P. J. Catal. 1999, 188, 8389.

Ammonia Synthesis: State of the Bellwether Reaction


temperatures, flows, and ratios; see Figure 12. For the dependence of flow, temperature, and H2:N2 ratio, the first-principles
model even provides quantitative agreement with the experimental data; see Figure 13.

467

In addition, an explanation of why productivity can be


predicted with such a good accuracy, although the activation
energies of individual elementary steps are less accurate, was
provided by Honkala et al.31 It turns out that the network of

10
7000

5000

x (29N2) (ppm)

Model output (NH3-%)

6000

0.1

4000
3000
2000
1000

0.01
0.01

0.1

0
450

10

500

550

600

Experimental output (NH3-%)

650

700

750

800

T (K)

1
0.1
0.01

Model ammonia output (NH3-%)

0.001
0.001
0.01
0.1
1
10
Experimental
ammonia
output
(NH
-%)
(a)
3

(c)

Model ammonia output (NH3-%)

10

10
1
0.1
0.01

0.001
0.001
0.01
0.1
1
10
Experimental
ammonia
output
(NH
-%)
(b)
3

10

100

1
0.1
0.01
0.001
0.001

0.01

0.1

10

Experimental ammonia output (NH3-%)

Model nitrogen output (N2-%)

Model ammonia output (NH3-%)

Figure 11 Comparison between measured and calculated NH3 production for the Ru/MgA1204 catalyst: pressure, 1100 bar; H2:N2 ratio, 6:11:4;
temperature, 320440  C. The solid line is the calculated outlet concentration of 29N2 when an inert gas containing 0.67% 29N2 and 0.60% 30N2 is
passed over 138 mg of catalyst containing 20 mmol g1 of active sites. The flow is 50 Nml min1 and the reactor is treated as a plug flow reactor.
The conditions are equal to the ones used by Hinrichsen et al.92 over the Ru/MgO catalyst, and the open circles are their results. Reprinted with
permission from Dahl, S.; Sehested, J.; Jacobsen, C. J. H.; Tornqvist, E.; Chorkendorff, I. J. Catal. 2000, 192, 391399.

(d)

10
1
0.1
0.01
0.001
0.0001
0.0001 0.001 0.01

0.1

10

100

Experimental nitrogen output (N2-%)

Figure 12 The calculated ammonia synthesis/decomposition output as calculated from the first-principles model and compared to measurements on
an industrially supported Ru-based catalyst. Reprinted with permission from Hellman, A.; Honkala, K.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.;
Dahl, S.; Christensen, C. H.; Nrskov, J. K. Surf. Sci. 2009, 603, 17311739.

468

Ammonia Synthesis: State of the Bellwether Reaction

Productivity (NH3-%)

consecutive reactions is self-regulating in the sense that the


errors in the different parts of the reaction cancel each other.
This is related to the so-called compensation effect,87 which
gives hope that one may more generally be able to calculate
catalytic rates directly from first principles.
In the present discussion about reaction mechanism and
active sites in ammonia synthesis, we have neglected the fact
that all ammonia synthesis catalysts are promoted by alkali
metals, which is central for preparing the most active catalysts.
Alkali-metal promoters like K and Cs added to ruthenium
catalysts increase the activity for ammonia synthesis. In addition, the reaction order of ammonia changes from negative to
about zero, whereas the reaction order for hydrogen decreases
and can end up at 1.85 It is well known that alkali metals
adsorbed on a metal surface donate electrons to the surface.
Theoretical calculations have shown that the presence of alkali
metal on metal surfaces lowers the N2 dissociation barrier
through a direct electrostatic interaction.64 The rate will
increase due to the higher dissociation rate and smaller inhibition by NH*, and the reaction order for ammonia will
increase as an effect of the latter. The model suggested by
Dahl et al.85 can qualitatively explain all the changes observed
when ruthenium catalysts are promoted with alkali metal due
to the rather high coverage of NHx, x 13, during synthesis
over the nonpromoted catalysts. So, as was the case for step
sites compared to terrace sites, the success of alkali-metal

promoters is based on their ability to lower the dissociation


barrier of the rate-limiting N2 dissociation step without
increasing the binding of N-containing species to the surface.

7.17.4.1

Activity Trends

In this section, it is shown how knowledge of the reaction


mechanism and structure of the active site has led to a fundamental understanding of the trends in catalytic ammonia synthesis activity of transition metals.
In 1911, Sabatier formulated a principle of what constitutes
a good catalyst.94 He states that a catalyst should neither interact too strongly or too weakly with the reactant. The reason is
that desorption from a reactive metal catalyst is slow and will
increase on less reactive metals. However, on very noble
metals, the large energy barrier for dissociation decreases the
dissociation rate. Thus, the best catalyst must be a compromise
between the two extremes.
Sabatiers principle provides the rationale behind the volcano curves95 often encountered in heterogeneous catalysis.
The volcano curves display the turnover frequency (or some
other activity-related property) against a property, such as the
heat of adsorption of the reactant by the catalyst. The typical
appearance looks like a triangle or an inverted parabola, and as
seen in Figure 14, this also holds for ammonia synthesis over
different transition metals.

Measured
Calculated

Measured
Calculated

Measured
Calculated

1
1
0.1
0.1
0.01
580 600 620 640 660 680 700 720

Temperature (K)

(a)

40

60

80 100 120 140 160 180 200

Flow (ml/min [STP])

(b)

4:1

1:1

1:3

1:5

Ratio (N2:H2)

(c)

Rate (STP ml NH3 / h)

100
Ru
10
1.0

Os
Rh

0.01

Re
Mo

0.1

Co
Ni

Pt
40

(a)

Fe

Ir

80

120

H: (kcal / metal atom)

Log k (molecules / cm2/s)

Figure 13 Ammonia productivity as a function of (a) temperature, (b) flow, and (c) ratio. The synthesis conditions are (a) N2:H2 (1:3), 30 Nml min1
(STP); (b) 713 K, N2:H2 (1:3); (c) 713 K, 30 N ml min1 (STP), all at a total pressure of 50 bar. Reprinted with permission from Hellman, A.; Honkala, K.;
Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C. H.; Nrskov, J. K. Surf. Sci. 2009, 603, 17311739.

160

16

Ru

14

Fe

12

Ni

Re

10
Pt

8
0

(b)

Co

Rh

40

80

120

160

H: (kcal / metal atom)

Figure 14 The rate of (a) ammonia synthesis and (b) ammonia decomposition as a function of the heat of chemisorption over various metals.96
Reproduced from Aika, K.; Yamazaki, K.; Ozaki, A. Chem. Lett. (Jap.) 1973, 2, 161.

Ammonia Synthesis: State of the Bellwether Reaction

The fundamental reason for the existence of volcano curves


in heterogeneous catalysis is the general scaling that exists
between the interaction strength of surfaces with different
reactants, products, reaction intermediates, and transition
states.97,98 In chemistry, linear correlations between activation
(free) energies and reaction (free) energies are widespread,
dating back to Brnsted in 192899 and to Evans and Polanyi
a decade later.100 In heterogeneous catalysis, such relations
have been assumed to hold.28,101 However, establishing accurate linear relationships over a sufficient range had to wait for
DFT calculations to become accurate enough. Figure 15(a)
shows the first published so-called BrnstedEvansPolanyi
(BEP) relationship based on DFT calculations between the reaction energy for dissociative adsorption for N2 (the nitrogen
adsorption energy) and the N2 dissociation barrier on different
transition metals.19 The slope of the BEP lines in Figure 15 is
0.9. The reason for why this value is close to 1 is that the
transition state for N2 dissociation is very final-state-like. Therefore, the transition-state energy essentially follows the nitrogen
adsorption energy from one metal to the next. Figure 15(b)
shows the BEP relation over a much wider range, that is, almost
12 eV in the chemisorption energy.102 This clearly indicates that
the occurrence of a BEP relation is a general manifestation of
the interaction of an adsorbate with the electronic structure of a
transition-metal surface (see also Chapter 7.15).
The BEP curves in Figure 15 are key to understanding the
trends in the ammonia synthesis activity of different transitionmetal catalysts, since a good catalyst for ammonia synthesis is
characterized by a low activation energy for N2 dissociation and
weak bonding of the most abundant surface intermediates,
which all bond to the surface via a nitrogen atom.19 For this
reason, the BEP relation does not only provide a method for
estimating activation barriers from adsorption energies, but it
governs the ammonia synthesis activity completely, as it determines the relation between the rate of the slowest step and the
amount of free sites on the surface, which is limiting when the
ammonia product gives rise to a high surface coverage on more
reactive surface sites.
From Figure 15, it is clear that the BEP relation is not
restricted to a single active site. The two lines represent terrace
6.0

and step sites, respectively. The linearity of the plots is a result


of a fixed geometry of the adsorption site for N and the transition state when the electronic structure of the transition
metals give rise to variation in both energies. That the two
lines have the same slope shows the transition state is finalstate-like to the same extent. The parallel shift of the two lines
is a manifestation of the geometrical effect, in the present case,
the requirement for dissociation sites with at least five metal
atoms at the step site, as discussed previously.
By combining a micro-kinetic model for the ammonia synthesis with some assumptions on how the adsorption energies
of NHx and H scale with the N adsorption energy and, finally,
the calculated BEP relation, the catalytic activity as a function
of nitrogen-binding strength can be calculated.19 Figure 16
displays the results; they clearly form a volcano curve with Fe
and Ru near the maximum, closely resembling what is found
experimentally, as seen when comparing to Figure 14. When
using the micro-kinetic model to analyze the surface coverage
and N2 dissociation rate as a function of the nitrogen adsorption energy, it is clear how the maximum in activity is a
compromise between a high N2 dissociation rate (low barrier
for N2 dissociation) and a large number of free sites (low N and
NH coverage) where the reaction can occur. On the right side
of the volcano curve, the rate is low due to a too high barrier for
N2 dissociation and, on the left side, the rate is low due to a too
high coverage of N and NH, leaving too few free sites for N2
dissociation, which is still the rate-limiting step here.
The effect of catalyst promotion with alkali metals can
easily be incorporated into the model based on the ability to
reduce the N2 dissociation barrier and lower the binding
strength of NHx species to the surface.18 The result of the
analysis is that promotion is most effective for the best nonpromoted catalysts and that promotion will always be essential
for obtaining an optimal ammonia synthesis catalyst.
It has now been established that the ability to describe
catalytic trends for different surfaces as a function of a single
key adsorption energy parameter (the descriptor) not only
holds for ammonia synthesis.103 The method is very general.
For instance, the BEP relation has been established for many
reactions and, for some reactions sharing some classes of

Cu(111)

5.0

Pd(211)step

3.0

ETS [eV]

Ea [eV/N2]

Fit to pure metals


Pure metals fcc(211)
Decorated steps
Bulk alloys

Pd(111)

4.0

fcc-Fe(111)

2.0
1.0

469

Ru(0001)
Fe(110)
Fe/Ru(0001)

0.0 Mo(110)
-1.0
-2.0
-3.0

Ru(0001)step
Fe(110)step
Fe(111)

Mo(110)step

-2
-2.0

-1.0

0.0

1.0

DE [eV/N2]

2.0

3.0

4.0

-6

-4

-2

Ediss (N2) [eV]

Figure 15 Correlation between the energy in the transition state and the dissociative adsorption energy of nitrogen. Reprinted with permission
from Logadottir, A.; Rod, T. H.; Nrskov, J. K.; Hammer, B.; Dahl, S.; Jacobsen, C. J. H. J. Catal. 2001, 197, 229231 and Munter, T. R.; Bligaard, T.;
Christensen, C. H.; Norskov, J. K. Phys. Chem. Chem. Phys. 2008, 10, 5202.

470

Ammonia Synthesis: State of the Bellwether Reaction

20.0

CoMo
10

Ru

10
10
10
10
10

15.0

Fe

-1

TOF [s1]

TOF(s-1)

10

Co3Mo3N

17.5

Os

Co
-2

-3

Mo

12.5
Ru

10.0
7.5
5.0

Ni

-4

2.5

Fe

-5

-100.0 -75.0 -50.0

-25.0

0.0

25.0

50.0

75.0

[DE-DE(Ru)](kJ/mol N2)
Figure 16 Turnover frequencies for ammonia synthesis as a function of
the adsorption energy of nitrogen. Reprinted from Jacobsen, C. J. H.;
Dahl, S.; Clausen, B. S.; Bahn, S.; Logadottir, A.; Nrskov, J. K. J. Am.
Chem. Soc. 2001, 123, 84048405, with permission Copyright 2001
American Chemical Society.

reaction, a universal behavior has been established, for example, the dissociative adsorption of the diatomic molecules O2,
N2, NO, and CO98 and (de)hydrogenation reactions.97 Furthermore, linear scaling relations between the adsorption energies of adsorbates that bind in a similar fashion to a surface has
been established,104 for example, the adsorption energy of NH,
NH2 that binds via the N atom to the surface scales with the
adsorption energy of N, and the slope of these linear scaling
relations can be approximately derived from simple bond
order conservation arguments.104
By combining BEP relations, scaling relations with microkinetic models, it has now been possible to construct volcano
curves for a number of reactions.105 For some reactions, it is
necessary to use not only one but two descriptors, for example,
reactions where CO is hydrogenated, since the reaction intermediates bind via both C and O, and the adsorption energy of
these does not correlate linearly. However, this still results in a
large reduction of parameters that need to be known in order
to predict the activity of a catalyst in a given reaction.

7.17.5 Discovery of New Catalysts for Ammonia


Synthesis
The ability to use computational methods to screen new catalyst materials is very valuable, since the preparation and testing
of new catalysts is a time-consuming task. The full kinetic
description of a given catalyst from theoretical calculation is a
quite demanding task, and screening a large number of systems
using a procedure that requires such an approach for each
system is, at the moment, too time-consuming, even with the
best computers. Therefore, a great achievement is that a direct
link can be created between measured catalytic activity and
easily-calculable parameter(s) describing the surface site, for
example, the adsorption energy of N on a step site. In this
section, it will be shown how this idea can be used to search
for more optimal catalysts and, furthermore, how it leads to a

3
4
5
6
NH3 concentration (%)

Figure 17 Measured turnover frequencies for promoted Ru, Co3Mo3N,


and Fe catalysts. Reprinted from Jacobsen, C. J. H.; Dahl, S.; Clausen,
B. S.; Bahn, S.; Logadottir, A.; Nrskov, J. K. J. Am. Chem. Soc. 2001, 123,
84048405, with permission Copyright 2001 American Chemical Society.

method to find the most optimal loading of an ammonia


synthesis reactor with the best combination of catalysts.
One of the first examples of where the results from
electronic-structure calculations were exploited to find a new
catalyst is ammonia synthesis. The volcano curve in Figure 16
presents a starting point, as it shows that none of the pure
transition metals are located at the top of the volcano curve.
Ruthenium is nearest to the maximum, but it is a scarce and
expensive metal. The volcano curve also reveals what is needed
in order to obtain a more optimal catalyst: a material with an
active site that has optimal nitrogen adsorption energy must be
found. It is natural to assume that this can be achieved by
creating an active site where two different metals from either
side of the volcano are present.
Following this interpolation principle, it was realized that
Mo-Co-based materials are suitable candidates to be tested.
The ammonia synthesis turnover frequency turned out to be
much better over a mixed-metal Co3Mo3N catalyst than over
its constituents.20 It is even better than the optimal Fe catalyst
at all reaction conditions, and also better than Ru at low NH3
concentrations; see Figure 17.
Central to the interpolation principle working experimentally is that the mixed site must be stable under reaction conditions. This might not work due to the preferential segregation
of one metal to the surface or phase separation of the alloy
under the reaction conditions, which can be induced by strong
binding of one of the components to the reaction intermediates. This could have been expected for the Co-Mo system,
since N binds considerably stronger to the Mo than to Co;
however, importantly, the ordering energy of the ternary
nitride prevents this.20
This rational approach for the computational screening of
materials for better catalysts is of general validity and has now
been used for a number of catalytic reactions, as reviewed
recently.105 The approach can be used at different levels. In
the simplest form, the interpolation principle is used to suggest
new catalysts candidates, while in more time-consuming
schemes, the volcano curve descriptors are calculated for a
large number of possible active sites, exposing different

Ammonia Synthesis: State of the Bellwether Reaction


elements.106 In this way, it is possible to take into account
discrepancies from BEP and scaling relations.
The position of the maximum of the volcano curve is
not the same at different reaction conditions. As an example,
Figure 18 shows how the position of the volcano curve changes
as the approach to equilibrium changes. At low ammonia concentrations, Fe is the best catalyst among the pure metals, while
Ru is the best choice near equilibrium concentrations of ammonia. This is consistent with experimental observations and results
from the fact that Fe catalysts and other catalysts on the left side
of the volcano are strongly inhibited by ammonia, which gives
rise to a high coverage of N and NHx species and, therefore, the
number of free sites available for N2 dissociation decrease with
increasing ammonia concentration. Since the coverage of Ncontaining species in equilibrium ammonia is low for catalysts
on the right side of the volcano curve, their activities are not
affected significantly by the approach to equilibrium.
Based on the volcano curves and the simplified assumption
that the density of active sites in different catalysts is the same,
an optimal value of the nitrogen-binding energy can be calculated, giving rise to the so-called optimal catalyst curves, which
can be used to choose the catalysts to load into a given ammonia synthesis reactor for the optimal ammonia output.33 In the
inlet of the reactor, a reactive catalyst like Fe is preferred, while
as the ammonia concentration increases, less reactive catalysts
are preferred. This implies that it is optimal to load an infinite
number of different catalysts with carefully chosen nitrogenbinding energies. In practice, this is, of course, not possible.
Generally, the smallest possible total catalyst volume results
when the reactor is designed to follow the optimal operating
line (see Section 7.17.2), and simultaneously use catalysts
with nitrogen-binding energies as close to the optimum as
possible.33 Consequently, it means that, with our current
understanding of the ammonia synthesis reaction, it is possible

to link density functional calculations with industrial reactor


design and catalyst selection.33
It is, of course, important to balance the potential savings in
the reactor volume against the cost of available catalysts and, in
practice, iron catalysts are totally dominating the market due
to their low price. One attempt was made to take advantage of
the high activity of ruthenium catalysts close to equilibrium. In
the Kellogg advanced ammonia process,107,108 a promoted
carbon-supported ruthenium catalyst is used in the last bed
of the process, where the ammonia concentration is high. This
part of the reactor is operated at lower pressure than normal
and with a H2:N2 ratio below 3 to further promote the activity
of Ru. However, it is doubtful that the benefits of this process
design can compensate for the higher cost and shorter lifetime
of the Ru/C catalysts compared to the traditional Fe catalyst.
From a price perspective, the Co3Mo3N catalyst is more attractive than an Ru catalyst; however, so far, a process using the
catalyst has not been realized. One reason for this is that it is
the product of activity per active site and the density of these
that determines the technically relevant activity of a catalyst,
and, so far, it has been difficult to prepare the Co3Mo3N
catalyst with a high surface area, since high-temperature nitrification is needed in order to produce the catalyst.

7.17.6

New Perspectives in Ammonia Synthesis

Depleting fossil fuel resources together with severe environmental consequences caused by the combustion of carboncontaining molecules enforce us to look for new energy solutions. There are several possible options for currently employed
hydrocarbons both for storage and for use as energy carriers,109
of which the most prominent ones are hydrogen, methanol,
ethanol, and methane. Also, synthetic hydrocarbons could

h = 5%
Fe 20%
Ru
Os

102

471

Fe

Ru
Os

101
90%
TOF (s-1)

100
10-1

Co

Mo

Mo

Co

10-2
10-3
10-4
10-5
-100

-50
0
EN*-EN* (Ru) (kJ/mol N2)

50

-100

-50
0
EN*-EN* (Ru) (kJ/mol N2)

50

Figure 18 (Left panel): calculated volcano curves at 420  C, 80 bar, 2:1 H2:N2, equilibrium 17.4%NH3. (Right panel): calculated volcano curves at
450  C, 200 bar, 3:1 H2:N2; equilibrium 25.4% NH3. In both cases, the volcano curves are calculated at ammonia concentrations corresponding to a
90, 20, and 5% approach to equilibrium. Reprinted with permission from Jacobsen, C. J. H.; Dahl, S.; Boisen, A.; Clausen, B. S.; Topse, H.;
Logadottir, A.; Nrskov, J. K. J. Catal. 2002, 205, 382387.

Ammonia Synthesis: State of the Bellwether Reaction

gradually replace fossil energy carriers, with the further benefit of


keeping the present infrastructure intact.
Ammonia has a high hydrogen content, making it a viable
candidate for serving as a hydrogen source for standard hydrogen
fuel cells. It can also be used in internal combustion engines and
direct ammonia fuel cells. All these together make ammonia a
promising option for new energy solutions.3,109 While nitrogen
is taken from the air, the HaberBosch process requires a hydrogen source. By far the most common source is currently natural
gas, but coal, petroleum coke, or heavy petroleum fractions can
be used to produce hydrogen as well. Electrocatalytic water splitting offers a sustainable way to obtain hydrogen, where any
source of electricity can be employed, including hydro, wind,
and solar cell power, or even nuclear power.
The ability to become a liquid at moderate pressure allows
ammonia to actually store more hydrogen per unit volume than
compressed hydrogen or cryogenic liquid hydrogen. However,
before ammonia can be used as a transportation fuel, design
standards for on-board ammonia fuel tanks must be established,
as well as procedures for ammonia transfer from storage to vehicle tanks. Compared to gasoline, ammonia has higher ignition
energy, higher flashpoint, and a narrower explosive range when
mixed with air, thus, an explosion or fire would be less likely with
a ruptured ammonia tank than with a gasoline tank. Both gasoline and diesel fuel can contain carcinogenic components in their
vapors, while ammonia does not. Furthermore, combusted
ammonia does not produce greenhouse gases nor soot, since it
does not contain carbon. However, the downside is the toxicity of
ammonia: at low concentration, vapors are highly irritating, with
sharp suffocating odor, and at high concentration, it is actually
life threatening. It has been shown that a convenient way to
alleviate this safety problem is to reversibly store the ammonia
with high density in an ammine complex such as Mg(NH3)6Cl2,
Ca(NH3)8Cl2, Mn(NH3)6Cl2, and Ni(NH3)6Cl2.110 The ammonia is bound in these compounds with a strength that makes the
partial pressure of ammonia so low that the storage material can
be handled safely at room temperature and allows the desorption
of ammonia at reasonable temperatures. Hence, although alternatives exist, ammonia has several interesting properties that
make it a possible option for special applications in energy technology. These properties include high energy density, a wellestablished synthesis, and high hydrogen content.
Ammonia decomposition into nitrogen and hydrogen, a
reverse process to ammonia formation, will be a key reaction
in any future ammonia-based energy infrastructure. Today,
ammonia decomposition plays a role in two different areas
related to energy and environmental science, namely, to provide carbon-free hydrogen for fuel cells and to remove ammonia from the reformate of internal gasification combined cycle
power plants.
Much of our current understanding of ammonia synthesis
can be directly employed to build an understanding of ammonia
decomposition.21 Owing to deviations in operating conditions,
for example, pressure, temperature, and gas composition,
ammonia formation and decomposition processes present
some clear and important differences. For instance, the volcano
curve for ammonia decomposition shifts towards catalyst materials that bind to nitrogen weaker, that is, towards more noble
metals. Similar to the ammonia synthesis reaction, the position
of the maximum of the volcano in ammonia decomposition

sensitively depends also on the reaction conditions. Interestingly, it is observed that the optimal ammonia synthesis catalyst
is never the optimal ammonia decomposition catalyst; see
Figure 19.21 This does not indicate that the principle of microscopic reversibility does not hold, but, rather, highlights that
widely different reaction conditions in ammonia synthesis and
decomposition result in very different optimal binding energies
for the two reactions, except, of course, at equilibrium. Also, for
ammonia decomposition, a combination of first-principle calculations, micro-kinetics, and surface science can lead to the
discovery of new catalysts formulations.111
One can imagine several scenarios where it would be advantageous to produce ammonia locally on a smaller scale, even if the
production costs would be higher than in the HaberBosch process. This point of view is becoming more relevant as ammonia is
used in the SCR process for removing nitrogen oxides from diesel
engine exhaust, with the recent proposal of using ammonia salts
as an energy carrier,3 or with the increased political focus on
transporting safety for hazardous chemicals. Hinnemann and
Nrskov have performed a theoretical study on the possibility to
produce ammonia under ambient temperatures and pressures. As
mentioned in the introduction, microorganisms exist in nature,
which use the enzyme nitrogenase to form ammonia from protons, electrons, and atmospheric nitrogen. In the enzyme, the
active site is an MoFe7S9 cluster, which catalyzes the reaction:
N2 8H 8e P 2NH3 H2 ;
where the nitrogen molecule is hydrogenated stepwise. The
source of energy for this reaction consists of at least 16 ATP
molecules.6 These presumably increase the chemical potential
of the electrons. It has been hypothesized that part of the
enzyme functions just like a battery.6
Since protonation reactions in particular appear to be so
well-described within the DFT methodology, it offers hope that
it is adequate also for describing electrochemical ammonia
synthesis.112 In an electrochemical cell, ammonia synthesis
on a ruthenium electrode in aqueous solution cannot proceed
via the same reaction mechanism as the HaberBosch process,
because all available ruthenium step sites for N2 dissociation
will be poisoned by strongly adsorbing oxygen atoms and OH
groups.112 The calculations show that the barrier for N2 dissociation is too high on the close-packed terraces at low electrochemical temperature. This leaves the biomimetic mechanism
-0.4
Optimal DE (eV)

472

-0.6

Synthesis

Decomposition

-0.8
-1.0
-1.2
-1.4
-1.6
0.001

0.01
0.1
1
10
Ammonia concentration (%)

100

Figure 19 Dissociative N2 adsorption energy of optimal catalysts for


ammonia synthesis/decomposition at 773 K, 1 bar, and 3:1 H2/N2.
Equilibrium corresponds to ca. 0.13% ammonia. Reprinted with
permission from Boisen, A.; Dahl, S.; Nrskov, J. K.; Christensen, C. H.,
J. Catal. 2005, 230, 309312.

5(

4(
+

e
+

(H

*N

H
N

*N

*NHNH
*NHNH2

H
*N

U=0V

2N

*N

2(

3(

2H

6(
*N

e
+

e
+

+
2

*N

U = -1.08 V

*N
N
H

Free energy [eV]

6(

Ammonia Synthesis: State of the Bellwether Reaction

Reaction coordinate
Figure 20 Free energy for the associative mechanism of electrochemical
ammonia synthesis on a flat Ru(0001) surface obtained from DFT
calculations of the binding energy and vibrational frequencies, as well as
entropy of the gas molecules. For an electrolyte with pH0 at 300 K, this
also gives the free energy for the electrochemical reaction when there is no
applied bias, U 0 V. The most difficult step is the addition of the first
proton to the adsorbed N2 molecule. With an applied potential of U
1.08 V, all the elementary steps involve either no change or a decrease in
the free energy. Reprinted with permission from Srensen, R. Z.;
Hummelshj, J. S.; Klerke, A.; Reves, J. B.; Vegge, T.; Nrskov, J. K.;
Christensen, C. H. J. Am. Chem. Soc. 2008, 130, 86608668.

strongly favored, especially because the tendency for direct


protonation of the nitrogen molecule can be tuned directly
by varying the voltage over the electrochemical cell. Figure 20
shows the full free-energy reaction diagram for the biomimetic
ammonia synthesis route at zero potential with respect to the
standard hydrogen electrode. The plot demonstrates that the
first hydrogenation process has, by far, the highest barrier. An
energy diagram is also shown at a potential of 1.08 V. At such
a low potential, all protonation reactions are thermodynamically downhill. If such a negative potential were applied,
ammonia could be produced electrochemically.112

7.17.7

Conclusion

We have, throughout this chapter, shown many of the reasons


why ammonia synthesis has been, and still is, the bellwether
reaction in heterogeneous catalysis. In particular, we have
described how, in recent years, it has been possible to provide
a full understanding of the ammonia synthesis reaction at the
atomic level through the combined use of experiments and
quantum mechanical electronic structure calculations. This
level of understanding led to the understanding of catalytic
trends, which was used for the rational design of a new catalyst
for ammonia synthesis. It is now clear that the approach is of
general validity and it has been used successfully to design new
catalysts for a number of reactions.
Finally, we have reviewed how ammonia synthesis over
ruthenium in industrial conditions can be understood only
from input from first-principle calculations. The fact that we
can treat one reaction on one surface in such enormous

473

theoretical detail reveals promise. There is a linear correspondence between the computational power and the number of
catalyst materials that we can screen for a given reaction at the
current high level of accuracy. With the historical increase in
available computer power, it is difficult to remain skeptical of
the importance of theory in the field of catalysis, already in the
near future. It can be envisioned that, soon, theoretical modeling will not only be used for reproducing the experimentally
known facts about a given reaction on a given surface, but
could become the standard choice as the first starting point
when a new catalyst for a known reaction is desired, or even
when an unknown reaction to obtain a given product is
needed. This would be the beginning of the era of Catalysis
Informatics,113 and there are strong indications that ammonia
synthesis will continue to be the bellwether reaction in this
development.

References
1. Smil, V. Nature 1999, 400, 415.
2. Ullmann, F. Ullmanns Encyclopedia of Industrial Chemistry; WILEY-VCH:
Weinheim, 2010.
3. Schuth, F.; Palkovits, R.; Schlogl, R.; Su, D. S. Energ. Environ. Sci. 2012, 5,
62786289.
4. Tamaru, K. In Catalytic Ammonia Synthesis Fundamentals and Practise; Jennings,
J. P., Ed.; Plenum Press: New York, 1991.
5. Schrock, R. R. Proc. Natl. Acad. Sci. 2006, 103, 17087.
6. Hinnemann, B.; Nrskov, J. K. Top. Catal. 2006, 37, 5570.
7. Ammonia: Catalysis and Manufacture; Nielesen, A., Ed.; Springer-Verlag:
New York, 1995.
8. Hopper, C. W. In Catalytic Ammonia Synthesis Fundamentals and Practise;
Jennings, J. P., Ed.; Plenum Press: New York, 1991.
9. Haber, F. Nobel Prize Lectures 1918; Elsevier: Amsterdam, The Netherlands,
1966.
10. Bosch, C. Nobel Prize Lecture 1931; Elsevier: Amsterdam, The Netherlands, 1966.
11. Haber, F.; van Oordt, G. Z. Fur Anorg. Chem. 1905, 44, 341378.
12. Mittasch, A. Adv. Catal. 1950, 2, 81.
13. Mittasch, A. U.S. Patent 1,173,532, 1913.
14. Aika, K.-I.; Hori, H.; Ozaki, A. J. Catal. 1972, 27, 424431.
15. Ozaki, A.; Aika, K.; Furuta, A.; Okagami, A. U.S. Patent 3,770,658, 1973.
16. Ozaki, A.; Aika, K.-I. In Catalysis, Science and Technology; Boudart, M., Ed.;
Springer: Berlin, 1981.
17. Aika, K.; Ozaki, A. In Ammonia: Catalysis and Manufacture; Nielesen, A., Ed.;
Springer-Verlag: New York, 1995.
18. Dahl, S.; Logadottir, A.; Jacobsen, C. J. H.; Nrskov, J. K. Appl. Catal. Gen. 2001,
222, 1929.
19. Logadottir, A.; Rod, T. H.; Nrskov, J. K.; Hammer, B.; Dahl, S.; Jacobsen, C. J. H.
J. Catal. 2001, 197, 229231.
20. Jacobsen, C. J. H.; Dahl, S.; Clausen, B. S.; Bahn, S.; Logadottir, A.;
Nrskov, J. K. J. Am. Chem. Soc. 2001, 123, 84048405.
21. Boisen, A.; Dahl, S.; Nrskov, J. K.; Christensen, C. H. J. Catal. 2005, 230,
309312.
22. Foster, A.; James, P. G.; McCarroll, J. J.; Tennison, S. R. U.S. Patent 4,163,775,
1979.
23. Chementator, Chem. Eng. 1993, 100, 19.
24. McCarroll, J. J.; Tennison, S. R. U.S. Patent 4,600,571, 1986.
25. Boudart, M. Top. Catal. 1994, 1, 405414.
26. Nielsen, A. An Investigation on Promoted Iron Catalysts for the Synthesis of
Ammonia; Gjellerup: Copenhagen, 1968.
27. Jennings, J. R. Catalytic Ammonia Synthesis: Fundamentals and Practice; Plenum
Press: New York, 1994.
28. Lundqvist, B. I.; Bligaard, T.; Nrskov, J. K. In Handbook in Surface Science vol 3:
Dynamics; Hasselbrink, E., Lundqvist, B. I., Eds.; 2008; p 269; Series editors
Richardson N. V.; Holloway, S.
29. Grunze, M. In The Chemistry and Physics of Solid Surfaces and Heterogeneous
Catalysis; King, D. A., Woodruff, D. P., Eds.; Elsevier: Amsterdam, The
Netherlands, 1982; Vol 4.

474

Ammonia Synthesis: State of the Bellwether Reaction

30. Tamaru, K. Catalytic Ammonia Synthesis, Fundamentals and Practice; Plenum


Press: New York, 1991.
31. Honkala, K.; Hellman, A.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.;
Christensen, C. H.; Nrskov, J. K. Science 2005, 307, 555558.
32. Dybkjaer, I. In Ammonia: Catalysis and Manufacture; Nielesen, A., Ed.; SpringerVerlag: New York, 1995; p 199.
33. Jacobsen, C. J. H.; Dahl, S.; Boisen, A.; Clausen, B. S.; Topse, H.; Logadottir, A.;
Nrskov, J. K. J. Catal. 2002, 205, 382387.
34. Gramatica, G.; Pernicone, N. In Catalytic Ammonia Synthesis; Jennings, J. R., Ed.;
Plenum: New York, 1991; p 211.
35. Somorjai, G. A.; Li, Y. Proc. Natl. Acad. Sci. 2011, 108, 917924.
36. Somorjai, G. A. J. Phys. Chem. B 2000, 104, 2969.
37. Somorjai, G. A.; Park, J. Y. Surf. Sci. 2009, 603.
38. Hammer, B.; Nrskov, J. K. Adv. Catal. 2000, 45, 1253.
39. Nrskov, J. K.; Scheffler, M.; Toulhoat, H. MRS Bull. 2006, 31, 669.
40. Hammer, B.; Nrskov, J. K. Nature 1995, 376, 238.
41. Lopez, N.; Nrskov, J. K. J. Am. Chem. Soc. 2002, 124, 11262.
42. Nrskov, J. K.; et al. Chem. Soc. Rev. 2008, 37, 2163.
43. Nakata, T.; Matsushita, S. J. Phys. Chem. 1968, 72, 458.
44. Shvachko, V. I.; Fogel, Y. M.; Kolot, Y. Kinet. Katal. 1966, 7, 834.
45. Ertl, G. J. Vac. Sci. Technol. A 1983, 1, 1247.
46. Ertl, G.; Huber, M.; Lee, S. B.; Paal, Z.; Weiss, M. Appl. Surf. Sci. 1981, 8, 373.
47. Ertl, G.; Lee, S. B.; Weiss, M. Surf. Sci. 1982, 114, 527545.
48. Ertl, G.; Weiss, M.; Lee, S. B. Chem. Phys. Lett. 1979, 60, 391394.
49. Emmett, P. H.; Brunauer, S. J. Am. Chem. Soc. 1933, 55, 17381739.
50. Emmett, P. H.; Brunauer, S. J. Am. Chem. Soc. 1934, 56, 3541.
51. Boszo, F.; Ertl, G.; Grunze, M.; Weiss, M. J. Catal. 1977, 49, 18.
52. Paal, Z.; Ertl, G.; Lee, S. B. Appl. Surf. Sci. 1981, 8, 231.
53. Spencer, N. B.; Schoonmaker, R. C.; Somorjai, G. A. J. Catal. 1982, 74, 129.
54. Strongin, D. R.; Carrazza, J.; Bare, S. R.; Somorjai, G. A. J. Catal. 1987, 103,
213.
55. Strongin, D. R.; Somorjai, G. A. J. Catal. 1988, 109, 51.
56. Stoltze, P.; Nrskov, J. K. Phys. Rev. Lett. 1985, 55, 2502.
57. Dumesic, J. A.; Trivino, A. A. J. Catal. 1989, 116, 119.
58. Dumesic, J. A.; Trevino, A. A. J. Catal. 1989, 116, 119129.
59. Stoltze, P.; Nrskov, J. K. J. Catal. 1988, 110, 110.
60. Hinrichsen, O. Cat. Today 1999, 53, 177188.
61. Ertl, G. Catal. Rev. Sci. Eng. 1980, 21, 201.
62. Strongin, D. R.; Bare, S. R.; Somorjai, G. A. J. Catal. 1988, 109, 51.
63. Mortensen, J. J.; Ganduglia-Pirovano, M. V.; Hansen, L. B.; Hammer, B.;
Stoltze, P.; Nrskov, J. K. Surf. Sci. 1999, 422, 816.
64. Mortensen, J. J.; Hansen, L. B.; Hammer, B.; Nrskov, J. K. J. Catal. 1999, 182,
479.
65. Dietrich, H.; Geng, P.; Jacobi, K.; Ertl, G. Chem. Eng. Sci. 1996, 51, 1683.
66. Mortensen, Y. M.; Hammer, B.; Norskov, J. K. J. Catal. 1997, 169, 85.
67. Dahl, S.; Logadottir, A.; Egeberg, R. C.; Larsen, J. H.; Chorkendorff, I.;
Tornqvist, E.; Nrskov, J. K. Phys. Rev. Lett. 1999, 83, 1814.
68. Diekhoner, L. M.; Mortensen, H.; Baurichter, A.; Luntz, A. C.; Hammer, B. Phys.
Rev. Lett. 2000, 84, 4906.
69. Shetty, S.; Jansen, A. P. J.; van Santen, R. A. J. Phys. Chem. C 2008, 112,
1776817771.
70. Gavnholt, J.; Schiotz, J. Phys. Rev. B 2008, 77, 035404.
71. Van Hardeveld, R.; Van Montfoort, A. Surf. Sci. 1966, 4, 396.
72. Logadottir, A.; Nrskov, J. K. J. Catal. 2003, 220, 273279.
73. Jacobsen, C. J. H.; Dahl, S.; Hansen, P. L.; Tornqvist, E.; Jensen, L.; Topsoe, H.;
Prip, D. V.; Moenshaug, P. B.; Chorkendorff, I. J. Mol. Catal. A Chem. 2000, 163,
1926.
74. Aika, K.; Kumasaka, M.; Oma, T.; Matsuda, H.; Watanabe, N.; Yamazaki, K.;
Ozaki, A.; Onishi, T. Appl. Catal. 1986, 28, 57.
75. Bossi, A.; Garbassi, F.; Petrini, G.; Zanderighi, L. J. Chem. Soc. Faraday Trans.
1982, 78, 1029.
76. Stoltze, P. Prog. Surf. Sci. 2000, 65, 65.
77. Dumesic, J. A.; Rudd, D. F.; Aparicio, L. M.; Rekoske, J. E.; Trivino, A. A. The
Microkinetics of Heterogeneous Catalysis; ACS: Washington, 1993.

78. Askgaard, T.; Nrskov, J. K.; Ovesen, C. V.; Stoltze, P. J. Catal. 1995, 156,
229.
79. Greeley, J.; Mavrikakis, M. J. Am. Chem. Soc. 2004, 126, 3910.
80. Kandoi, S.; Greeley, J.; Sanchez-Castillo, M. A.; Evans, S. T.; Gokhale, A. A.;
Dumesic, J. A.; Mavrikakis, M. Top. Catal. 2006, 37, 17.
81. Ovesen, C. V. C.; Hammershi, B. S.; Steffensen, G.; Askgaard, T.;
Chorkendorff, I.; Nrskov, J. K.; Rasmussen, J. P.; Stoltze, P.; Taylor, P. J. Catal.
1996, 158, 170.
82. Linic, S.; Barteau, M. A. J. Am. Chem. Soc. 2003, 125, 4034.
83. Stegelmann, C.; Schidt, N. C.; Campbell, C. T.; Stoltze, P. J. Catal. 2004, 221,
630.
84. Nrskov, J. K. Prog. Surf. Sci. 1991, 38, 103144.
85. Dahl, S.; Sehested, J.; Jacobsen, C. J. H.; Tornqvist, E.; Chorkendorff, I. J. Catal.
2000, 192, 391399.
86. Stoltze, P. Phys. Scripta 1987, 36, 824.
87. Bligaard, T.; Nrskov, J. K.; Dahl, S.; Matthiesen, J.; Christensen, C. H.;
Sehested, J. J. Catal. 2004, 224, 206217.
88. Hellman, A.; Honkala, K. J. Chem. Phys. 2007, 127, 194704.
89. Sehested, J.; Jacobsen, C. J. H.; Tornqvist, E.; Rokni, S.; Stoltze, P. J. Catal.
1999, 188, 8389.
90. Fastrup, B. J. Catal. 1997, 168, 235244.
91. Hellman, A.; Honkala, K.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.;
Christensen, C. H.; Nrskov, J. K. Surf. Sci. 2009, 603, 17311739.
92. Hinrichsen, O.; Rosowski, E.; Hornung, A.; Muhler, M.; Ertl, G. J. Catal. 1997,
165, 33.
93. Hellman, A.; Honkala, K.; Remediakis, I. N.; Logadottir, A.; Carlsson, A.; Dahl, S.;
Christensen, C. H.; Nrskov, J. K. Surf. Sci. 2006, 600, 42644268.
94. Sabatier, P. Ber. Deutsch. Chem. Ges. 1911, 44, 1984.
95. Balandin, A. A. Adv. Catal. 1969, 19, 1.
96. Aika, K.; Yamazaki, K.; Ozaki, A. Chem. Lett. (Jap.) 1973, 2, 161.
97. Wang, S.; Petzold, V.; Tripkovic, V.; Kleis, J.; Howalt, J. G.; Skulason, E.;
Fernandez, E. M.; Hvolbk, B.; Jones, G.; Toftelund, A.; Falsig, H.; Bjorketun, M.;
Studt, F.; Abild-Pedersen, F.; Rossmeisl, J.; Nrskov, J. K.; Bligaard, T. Phys.
Chem. Chem. Phys. 2011, 13, 20760.
98. Nrskov, J. K.; Bligaard, T.; Logadottir, A.; Bahn, S. R.; Hansen, L. B.;
Bollinger, M. V.; Bengaard, H. S.; Hammer, B.; Sljivancanin, Z.; Mavrikakis, M.;
Xu, Y.; Dahl, S.; Jacobson, C. J. H. J. Catal. 2002, 209, 275.
99. Brnsted, N. Chem. Rev. 1928, 5, 231.
100. Evans, M. G.; Polanyi, M. Trans. Faraday Soc. 1937, 33, 448.
101. Boudart, M. In Handbook of Heterogeneous Catalysis; Ertl, G., Knozinger, H.,
Weitkamp, J., Eds.; Wiley-VCH: Weinheim, 1997.
102. Munter, T. R.; Bligaard, T.; Christensen, C. H.; Norskov, J. K. Phys. Chem. Chem.
Phys. 2008, 10, 5202.
103. Nrskov, J. K.; Abild-Pedersen, F.; Studt, F.; Bligaard, T. Proc. Natl. Acad. Sci.
2011, 108, 937943.
104. Abild-Pedersen, F.; Greeley, J.; Studt, F.; Rossmeisl, J.; Munter, T. R.;
Moses, P. G.; Skullason, E.; Bligaard, T.; Nrskov, J. K. Phys. Rev. Lett. 2007, 99,
016105.
105. Nrskov, J. K.; Bligaard, T.; Rossmeisl, J.; Christensen, C. H. Nat. Chem. 2009,
1, 37.
106. Studt, F.; Abild-Pedersen, F.; Bligaard, T.; Srensen, R. Z.; Christensen, C. H.;
Nrskov, J. K. Science 2008, 320, 13201322.
107. Czuppon, T. A.; Knez, S. A.; Schneider, R. W.; Worobets, G. Ammonia Plant Safety
Relat. Facil. 1994, 34, 236.
108. Strait, R. Nitrog. Methan. 1999, 238, 37.
109. Schuth, F. Chem. Eng. Technol. 2011, 83, 1984.
110. Srensen, R. Z.; Hummelshj, J. S.; Klerke, A.; Reves, J. B.; Vegge, T.;
Nrskov, J. K.; Christensen, C. H. J. Am. Chem. Soc. 2008, 130, 86608668.
111. Hansgen, D. A.; Vlachos, D. G.; Chen, J. G. Nat. Chem. 2010, 2, 484489.
112. Skulason, E.; Bligaard, T.; Gudmundsdottir, S.; Studt, F.; Jan Rossmeisl, F. A.-P.;
Vegge, T.; Jonsson, H.; Nrskov, J. K. Phys. Chem. Chem. Phys. 2012, 14,
1235.
113. Hummelshj, J. S.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nrskov, J. K.
Angew. Chem. Int. Ed. 2012, 51, 272274.

Você também pode gostar