Sie sind auf Seite 1von 10

Topics in Catalysis 1 (1994) 405-414

405

Ammonia synthesis:
the bellwether reaction in heterogeneous catalysis
M. Boudart
Department of Chemical Engineering, Stanford University, Stanford,
CA 94305-5025, USA

This is a personal account of the Havreholm Conference with a choice of t o p i s dealing


with catalytic ammonia synthesis. Among the general concepts that were retained are: the rate
determining step with rates of adsorption and desorption of nitrogen; the direct activated dissociative adsorption of N2; the surface crystalline anisotropy of iron; the role of promoters in
industrial iron based catalysts; and the atomic structure of the metallic surface on the industrial
multiply promoted catalyst. Finally, a new isotope jump technique to measure an upper limit
to the real turnover frequency is discussed.
Keywords: ammonia synthesis; mierokinetic modeling; nitrogen dissociation; surface
structure; rate determining step; promotion; isotope tracing

1. Introduction

It is a fitting tribute to the work and life of Mikhail I. Temkin (1908-1991) to


start this personal report of the Havreholm Conference on Ammonia Synthesis and
Beyond with a few remarks on the pivotal role of Temkin in the development of
modern catalytic synthesis of ammonia. Since both Mikhail Temkin and Paul
Emmett had a profound and parallel influence on the theory and practice of ammonia synthesis, it is of great significance to quote Paul Emmett in his comments on
the seminal paper of M.I. Temkin and V. Pyzhev, an English version of which
appeared in 1940 [1].
The quotation by Emmett is borrowed from a volume dedicated to Paul Emmett
[2]: "With a few exceptions, work has tended to show that the slow step in the synthesis of ammonia is the chemisorption of nitrogen and the slow step for the decomposition is the desorption of nitrogen. Furthermore, it turns out that the
decomposition and synthesis of ammonia usually involve in the rate expression a
term

where y / x is close to 1.5. In 1940, Temkin and Pyzhev derived an equation consistent with both of these observations. It has formed the basis for most of the kinetic
J.C. Baltzer AG, Science Publishers

406

M. Boudart / The bellwether reaction

treatments of ammonia synthesis and decomposition in recent years .... The


authors assumed that the adsorption of nitrogen on the iron catalyst in the presence
of an ammonia-hydrogen mixture is the same as it would be when at a nitrogen
pressure equivalent to the existing partial pressure of ammonia and hydrogen in the
gas mixture."
The three seminal ideas in this early work of Temkin are powerful because of
their generality. The first one is that adsorption of nitrogen is rate determining,
with a clear notion of the now accepted meaning of the rate determining step. The
second one is the virtual pressure or fugacity of adsorbed nitrogen, a concept of
great importance to the understanding of catalytic cycles at the steady-state.
Indeed, it implies that the active adsorbed intermediates are not necessarily in
equilibrium with fluid phase species, as assumed in conventional LangrnuirHinshelwood kinetics. The third idea is the kinetic description of the catalytic surface as a non-uniform one. The last was systematized later by Temkin's school both
in theory and application to a large number of important catalytic reactions. The
importance of Temkin's theory of kinetics on non-uniform surfaces is not so much
in its formalism to fit kinetic data, but in the deeper kinetic understanding of how
any catalyst works and how to select the catalyst with the fastest turnover rate [3].
Like the kinetic concepts of Christiansen and Horiuti, those of Temkin were far
ahead of their common acceptance by the catalytic community. Even today, more
than fifty years after the Temkin-Pyzhev paper, the idea of fugacity of adsorbed
species is not appreciated by the majority of workers in catalytic kinetics.
The life work of Temkin is inseparably linked to ammonia synthesis. It illustrates vividly how so many of the general concepts in heterogeneous catalysis have
originated and still are tested in the study of ammonia synthesis.

2. The rate determining step (RDS)


As it appears in the above quotation from Paul Emmett, it was recognized
rather early, but in a vague and intuitive sort of way, that adsorption of dinitrogen
was the "slow" step in ammonia synthesis. The concept was correctly used by
Temkin and Pyzhev, but it took the theoretical and experimental work of Horiuti
and co-workers applied to ammonia synthesis to define the concept in its rigorous
form: for details, see ref. [4].
In a catalytic cycle, there may exist one elementary step (called step for short)
that is the only one that is not in quasi-equilibrium. That step, if it exists, is called
the rate determining step (RDS): it is the only step that enters through its rate constants into the rate equation for the turnover frequency. Thus the RDS is the only
kinetically significant step. After some experimental difficulties, Horiuti, and subsequently others, identified the RDS for ammonia synthesis on iron at medium
pressures and temperatures. The technique invented by Horiuti was an isotope
technique in which a system at chemical equilibrium containing N2, H2, and NH3 is

M. Boudart / The bellwetherreaction

407

perturbed by introduction of N2 tagged with ISN and the rate of appearance of


ISN in a m m o n i a ismeasured: thisis the rate of reaction at equilibrium ifthe chosen
isotopic equilibrationreaction goes through the RDS. Then by also measuring the
rate of reaction near chemical equilibrium, it is possible to determine h o w m a n y
times the suspected R D S must take place for the catalyticcycle to turnover once.
This number was found ultimately to be unity, and this can bc understood only if
the adsorption of nitrogen is the RDS. A related new isotope jump technique consisting of perturbing the ammonia synthesis occurring at the steady state was presented at the Conference [5]and willbc discussed later.
At the Conference, it was noted by Tamaru [6] that the R D S in the catalytic
cycle for the ammonia decomposition is not always the dcsorption of nitrogen. In
fact, as pressure and temperature arc varied, Horiuti's concept of the R D S ceases
to apply as more than one step in the cycle is not equilibrated. To describe such
situations,itisuseful to extend Horiuti's concept of the R D S [7]:ifthere exists one
step that is the only one to be kinetically significant in the sense defined above, it
can also be called RDS even though there may be other steps that also are not in
quasi-equilibrium. This matter was brought up in a slightly different way at the
Conference by Campbell [8].
Another general concept that first appeared in connection with a m m o n i a synthesis is that of most abundant reactive intermediate (MARI). In fact, the concept
was assumed implicitly by Temkin and Pyzhev in their original paper [2]. Indeed,
without the assumption that chemisorbed nitrogen was the M A R I , they would not
have been able to obtain their famous rate equation.
To illustrate the points made in this section, the shift in R D S for the decomposition of ammonia is noted in table 1 [9]. The shift depends on temperature and pressure. The R D S in both Horiuti's sense and the extended sense can shift or cease to
exist, even though there is no apparent change in mechanism defined as the set of
elementary steps in the catalytic cycle and adsorbed nitrogen seems to remain the
M A R I in all cases.

3. Rates ofdesorption of nitrogen


If the two-way, i.e.,reversible,adsorption of nitrogen is the R D S for a m m o n i a
Table I
CatalyticdecompositionofNH3. M A R I isadsorbedN inallcases
p

Orderwithrespect
toNI-I3

Metal

RDS

Ref.

H
M
L
LL

L
M
H
HH

fractional
zero
fractional
one

Fe
Ru
Mo
Pt

desorptionofN2
same
none
adsorptionofNH3

[2]
[I0]
[II]
[I2]

M. Boudart/ Thebellwetherreaction

408

synthesis on iron as it is believed to be under normal conditions of T and p, it is


obvious that the data required to calculate the turnover rate of ammonia synthesis
are quite limited. Indeed, as stressed by Aparicio and Dumesic [13, eq. 5], the turnover frequency for ammonia synthesis can be written as
TOF=

kdO2(K [N2][H213
[NH312

l)

where kd is the rate constant for desorption of N . , 0 the surface coverage by N


atoms, K the known equilibrium constant for the synthesis reaction, and the quantities between brackets the partial pressures expressed in units of the standard state
pressure. Note that knowledge of the rate constant for adsorption of nitrogen is
not necessary provided that 0 can be determined.
So, what do we need to calculate the TOF? We need only values ofkd and 0 and
the number of sites, assumed to be identical and non-interacting, in Langmuir fashion. To the extent that this Langmuirian behavior is realistic, then both kd and 0
can be determined from temperature programmed desorption (TPD), the identical
active sites being counted by selective chemisorption. If 0 is not known but estimated to lie between 0.5 and ,-~ 1, as appears to be the case under conditions of
ammonia synthesis, it is easy to obtain a corresponding estimate of TOF within a
factor of two. The success of such a simplistic approach is documented in the paper
of Aparicio and Dumesic, who also point out that a TPD experiment probes the surface with a high surface coverage, corresponding to values under synthesis conditions. Then, interactions between adsorbed N atoms, important as they may be,
will not change appreciably with coverage. Finally, as will be discussed later,
at high values of 0, the surface may have been reconstructed so as to appear uniform, not only in terms of lateral interactions between adatoms, but also with
respect to the nature of active sites. The relevance of nitrogen TPD experiments performed on real ammonia synthesis conditions is beautifully and convincingly
demonstrated by Fastrup [14]. In that work, Fastrup also demonstrates the need
for strict criteria of reproducibility that should always be observed in future work
on ammonia synthesis with iron based catalysts. She also concludes that Langmuir
assumptions are not strictly applicable. While this conclusion should always be
kept in mind, what remains astonishing is how good Langmuir assumptions can be
under real conditions of ammonia synthesis on a multicomponent iron based catalyst. This paradox will be examined again later.

4. Rate of adsorption of nitrogen


While the rate of adsorption of nitrogen is not necessary in the calculation of
TOF for ammonia synthesis as just discussed, it is still a most important quantity to
know. Indeed, so-called nitrogen activation on iron surfaces has long been consid-

M. Boudart / The bellwether reaction

409

ered as the best example of Taylor's activated adsorption. Early work reviewed in
this volume [15] deals with single crystal surfaces of iron at very low pressures of
N2. The work did not reveal the expected activation barrier for chemisorption of
N2, as only surface temperature was varied. But more recent work with seeded
molecular beams of N2 shows that the sticking coefficient of N2 on iron increases
with increasing kinetic energy of the incident molecules [16]. This is due to "hot"
molecules of N2 surmounting an activation barrier for adsorption. Taylor's concept of activated adsorption first introduced in 1932 [17] has been vindicated 55
years later in connection with the rate determining step for ammonia synthesis.
This most important theoretical and experimental result was discussed in this Conference by Ertl [15] and Bowker [18], as well as by Stoltze and Norskov [19].
Thus we must retain both indirect activation, mediated through a molecular
adsorbed precursor and direct dissociation by surmounting an activation barrier,
as discussed by Butler and Hayden in the case of H2 on W(100) [20]. Both modes of
activation have been discussed in the case of N2 on Fe(111) [15,19]. As a noteworthy result of the Havreholm Conference, it was agreed upon that the direct
mode of dissociation by activated adsorption is the one of importance under the
conditions of ammonia synthesis [18].

5. Surface crystalline anisotropy of iron in a m m o n i a synthesis


The suggestion that the (111) face of iron is preferentially exposed following
topotactic reduction of magnetite, Fe304, in hydrogen was made first by Westrik
and Zwietering [21]. Since industrial catalysts are made by careful reduction of
magnetite fused with non-reducible oxide promoters, the role of (111 ) facets of iron
at the surface of ammonia catalysts appears to be important. It was confirmed by
the recent work of Schl6gl et al. as summarized in this Volume [22]. These observations must be viewed in the light of one of the most important advances in the study
of single crystal model catalysts, namely the striking work of Spencer, Schoonmaker, and Somorjai [23], as amplified by the work of the Somorjai group reviewed
in this Volume [24]. These studies of the areal time yield of ammonia at 20 bar on
five different faces of iron single crystals established the enormous structure sensitivity of ammonia synthesis. The (111) face of iron turned out to be the most active
face investigated. Faces that do not expose C7 sites exhibit yields that are one or
two orders of magnitude smaller than the (111) and (211) faces that do. The superior activity of C7 sites in ammonia synthesis on supported clusters of iron had previously been emphasized by Dumesic et al. [25].
Additionally, it was shown by Dumesic et al. [26] that ammonia pretreatment
of iron clusters reconstructs their surface with appearance of C7 sites. This surface
reconstruction is reversibly erased by reduction in H2. Thus, surface nitrogen
reconstructs the iron surface as first observed by field emission microscopy of an
iron tip [27].

410

M. Boudart / The bellwether reaction

When a commercial iron catalyst is examined under ammonia synthesis conditions by temperature programmed reaction, first at increasing temperatures, then
at decreasing temperatures, the yield of ammonia versus temperature exhibits an
interesting hysteresis loop first reported by Richard and Vanderspurt [28]. This
hysteresis was interpreted in terms of a beneficial reconstruction of the surface
under increasing amounts of NH3 followed by a return to the original surface structure at lower yields of ammonia. Similar, more extensive observations were
reported at the Havreholm Conference by Waugh et al. [29].
Two related questions remain to be answered. What is the role of the alumina
and potassium oxide promoters in the commercial catalysts? First, do they contribute to the restructuring of the iron surface? Second, does the iron surface of the
commercial catalyst behave in synthesis like the best face of crystalline iron, i.e.,
the (111) face? Let us address these two questions in turn.

6. The role of promoters


With the developments of the BET method and of the selective chemisorption
techniques by Emmett and co-workers [2], the total specific surface of iron catalysts
promoted with A1203 and K20 can be determined together with the fraction of
the total surface occupied by A1203. Also, the fraction of the Al203 surface associated with K20 can be measured. The emerging picture of a fully reduced, doubly
promoted catalyst, as supported by Auger electron spectroscopy [30] is the following: free iron patches covering about half of the total surface are surrounded by
islands of promoters ca. 2 nm in diameter and 1 nm thick. These promoter islands
consist of Al203 decorated with KzO. Although the role of Al203 may be in part
to stabilize (111) facets [29], its primary role seems to prevent sintering of the metal
by acting as a spacer, as documented once more at the Havreholm Conference
[22]. The role of K20 remains in doubt especially because its beneficial effect
becomes apparent only at high pressure, and not at atmospheric pressure [21,29].
One explanation of the effect of K20 is that the latter is associated with Fe, perhaps at the periphery of promoted islands and, as a result, nitrogen is more weakly
held on the surrounding iron [14,24]. At high pressures, with high partial pressures
of NH3 and thus higher fugacity of surface nitrogen, the fraction of the Fe surface
covered with N atoms increases from ca. 1/2 to almost unity [13,14]. Thus the effect
of K20 would be to make more Fe sites available, effectively decreasing the apparent inhibition at higher partial pressures of ammonia.
But the situation is further complicated by evidence [24] suggesting that metallic
iron creeps over A1203 islands that stabilize the Fe (111) facets responsible for
superior activity. This new phenomenon might also account for the hysteresis phenomena in TPR as reported in this volume [29]. This an area where much more
work remains to be done.

M. Boudart / The bellwether reaction

411

7. T h e a t o m i c structure o f i r o n o n industrial catalysts

Another unresolved question is whether the working industrial iron catalyst


exposes mostly (111) facets or not. If not, further improvements of catalyst performance are possible, because of the demonstrated decisive superiority of the (111)
orientation over those that do not expose C7 sites. The remaining doubt is largely
due to the uncertainty of extrapolations of TOF values as reported in the literature,
as well as to the reluctance of many workers in catalysis to report their rate data
expressed in values of TOF.
Admittedly, any value of TOF remains nominal unless the number of active
sites is known and all active sites have the same value of TOF. The ultimate goal is
to determine TOF without having to count the number of active sites. This question will be examined at the end of this review. Yet, if the method of counting active
sites is spelled out, comparisons between TOF values obtained in different l~boratories can be made unequivocally. Certainly, the use of conversion as a measure
of activity should be banned. Even more objectionable is the use of relative conversion, e,g., calculated conversion versus experimental conversion: this is unacceptable as it is tantamount to activity expressed in the arbitrary units that have
retarded the development of scientific catalysis until relatively recently.
A key paper illustrating the quantitative value of TOF is that of Topsoe et al.
[31]. They reported values of TOF for ammonia synthesis from stoichiometric mixtures of N2 and H2 at atmospheric pressure, 673 K and at an efficiency of0.15, efficiency being defined as %NH3 in outlet divided by %NH3 at equilibrium. They
counted active sites by chemisorption of N2 at 573 K. Values of TOF are essentially
the same, namely 0.5 s -1, for three samples of MgO supported Fe catalysts with
Fe particle size between 5 and 10 urn, as well as for a commercial multipromoted
catalyst with 60 um Fe particles. But when the active sites were counted by CO chemisorption at 195 K, the Emmett technique for counting all surface Fe atoms, the
values of TOF for the three supported catalysts were only about 10-2 s -1, but
remained at about the same value, namely 0.5 s -I, for the commercial catalyst.
Therefore it appears that N2 chemisorption counts only the most active sites, and
that the commercial catalyst surface contains mostly these most active sites. The
question is: what is the nature of these active sites?
A tentative answer is that the active sites on the commercial catalyst surface are
C7 sites as found on (111) or perhaps also (211) facets. This answer came for a comparison of the value of TOF = 0.5 s -1, as cited above with a value calculated by
extrapolating the data of Spencer et al. [23] obtained for ammonia synthesis on
Fe(111) from stoichiometric mixtures at 20 bar, 798 K and an efficiency of 0.015.
This extrapolation to the conditions of Topsoe et al. [31] yields a TOF value differing only by a factor of 2 from the value ofTopsoe et al. The details of the extrapolation have been fully published [32] and, as all extrapolations of rates, are subject
to error. Yet, the result of the extrapolation seems reasonable when considered
together with the above results of Topsoe et al. [31]. By contrast, we note a state-

412

M. Boudart / The bellwether reaction

ment in the paper of Waugh et al. in this Volume [29]: "The surface restructuring
of the industrial catalysts does not produce dominantly the (111) surface, the
observed turnover numbers being 103 times lower than that of the (111) surface."
Unfortunately, what the authors call turnover number is actually an areal yield:
NH3 molecules cm -2 s -1. Until a TOF value is available, the nature and magnitude
of the discrepancy noted by Waugh et al. remains unclear.
In the meantime, it is useful to note that Dumesic et al. [33] used successfully a
microkinetic analysis to calculate values of site time yield for commercial Fe catalysts over a wide range of temperatures and pressures. They note that the same
microkinetic models were successful "within a factor of about 3" to describe both
the results of Spencer et al. [23] on Fe(111) and those ofTopsoe et al. [31] on a commercial catalyst. Again, it appears that the active sites on the commercial catalyst
are (111) facets, but this conclusion remains tentative.
At this moment what is clear from the work of Topsoe et al. [31] is that most of
the surface of a commercial Fe catalyst seems to be covered with the most active
sites capable of chemisorbing nitrogen at high temperature. This, together with the
high surface coverage by nitrogen at high values of nitrogen fugacity, helps in
understanding why the Langrnuir model is so successful, though not perfect, in
describing the kinetics of ammonia synthesis under commercial conditions.

8. T h e H o l y Grail o f catalysis: counting the active sites or determining the


turnover frequency

The problems of identifying and counting the active sites on a catalytic surface
have just been discussed. A way to bypass this difficult task is to measure directly
the turnover frequency. As of now, the best technique to approach this elusive goal
is the steady state isotope technique for kinetic analysis, or SSITKA, which consists of perturbing a catalytic reaction taking place at the steady state in a stirred
tank reactor by an isotope jump. The method is similar to the isotope jump technique of Horiuti, reviewed above, except that the Horiuti method perturbs the equilibrium and SSITKA perturbs the steady state. In the case of ammonia synthesis,
the perturbation is effected by replacing suddenly 14N14N in the feed by 15N tagged
nitrogen and measuring the relaxation time r of the decay of the 14N containing
ammonia. The technique is described by Nwalor and Goodwin [5] and applied by
them to ammonia synthesis on a silica supported ruthenium without and with
potassium promoter. As the authors note, the turnover frequency TOF is given by
TOF --- O/r,
where 0 is the fraction of the active sites covered with adsorbed N atoms which
are the most abundant surface intermediates passing through the rate determining
step. If we assume that 0 is equal to unity, we obtain an upper bound value for
TOF:

M. Boudart / The bellwether reaction

413

TOFmax = 1 / r .
On the other hand, we may obtain a lower bound value for the T O F by referring
the rate to the total number of sites, active or not, as counted by a chemisorption
technique such as hydrogen chemisorption on supported metal clusters. This value
of T O F will be called TOFmin.
Now, if I combine the last columns of tables 2 and 3 in the paper of Nwalor
and Goodwin, I obtain with the above symbols the following results. For unpromoted ruthenium:
TOFmin = 0.6 x 10 -4 s -1

and

TOFmax = 5.6 x 10 -4 s -1 .

But for potassium promoted ruthenium, I obtain:


TOFmin = 62 x 10 -4 s -I

and

TOFraax -- 88 x 10 -4 s -I .

What I note is that for the u n p r o m o t e d catalyst, the lower bound and upper bound
values of T O F differ by a factor of 10. But for the promoted catalyst, they differ
by a factor of only 1.4. Thus, as noted by Nwalor and Goodwin [5], the role of the
promoter is to create sites of superior activity that actually cover most of the metal
surface [5]. I believe that this notion of m i n i m u m and m a x i m u m values of T O F
can be of general use whenever SSITKA is applicable: i.e., it gives an upper bound
value for T O F and kinetic guidance in the search for more active catalysts. In conclusion, it must be noted that the first application of SSITKA was for the study of
a commercial iron catalyst for ammonia synthesis, the trustworthy bellwether of
heterogeneous catalysis [34].

9. C o n c l u s i o n
Twenty years ago, a Battelle Colloquium took place in Gstaad, Switzerland, on
"The Physical Basis of Heterogeneous Catalysis" [2]. As noted in the Introduction,
it was dedicated to Paul H. Emmett "for his contributions to heterogeneous catalysis over fifty years". The prefatory chapter by Emmett in the Battelle Colloquium
Volume was a summary of "Fifty Years of Progress in the Study of the Catalytic
Synthesis of A m m o n i a " . This chapter is well worth reading today. Twenty years
later, the Havreholm Conference has looked again at ammonia synthesis with an
emphasis on current developments, especially as they impact on the future of heterogeneous catalysis. The Havreholm Conference was dedicated to Haldor Topsoe
and Anders Nielsen for their fifty years of leadership in heterogeneous catalysis
and a m m o n i a synthesis. After reviewing the proceedings of the Conference, I have
every reason to believe that they will be eminently worth reading twenty years
from now.

414

M. Boudart / The bellwether reaction

References
[I] M.I. Temkin andV. Pyzhev, Acta Physicochim. URSS 12 (1940) 327.
[2] P.H. Emmett, in: The Physical Basis for Heterogeneous Catalysis, eds. E. Grauglis and
R.I. Jaffee (Plenum Press, New York, 1975).
[3] M. Boudart and G. Dj~ga-Mariadassou, Kinetics of Heterogeneous Catalytic Reactions
(Princeton University Press, Princeton, 1984).
[4] M. Boudart, Kinetics of Chemical Processes (Butterworth-Heinemann, Stoneham, 1991).
[5] J.U. Nwalor and J.G. Goodwin Jr., this volume.
[6] K. Tamaru, communication at the conference.
[7] M. Boudart and K. Tamaru, Catal. Lett. 9 (I 991) 15.
[8] C.T. Campbell, this volume.
[9] M. Boudart, Ind. & Eng. Chem. Res., submitted.
[10] W. Tsai and W.H. Weinberg, J. Phys. Chem. 91 (1987) 5302.
[11] M. Boudart, C. Egawa, S.T. Oyama and K. Tamaru, J. Chim. Phys. 78 (1981) 987.
[12] D.G. L6ffler and L.D. Schmidt, I. Catal. 41 (1976) 440.
[13] L.M. Aparicio and J.A. Dumesic, this volume.
[14] B. Fastrup, this volume.
[15] G. Ertl, this volume.
[16] C.T. Rettner and H. Stein, Phys. Rev. Lett. 59 (1987) 2768.
[17] H.S. Taylor, Trans. Faraday Soc. 28 (I 932) 137.
[18] M. Bowker, this volume.
[19] P. Stoltze and J.K. N~rskov, this volume.
[20] D.A. Butler and B.E. Hayden, this volume.
[21] R. Westrik and P. Zwietering, Proc. Koninkl. Ned. Akad. Wetenschap. B56 (1953) 492.
[22] J. Schfitze, W. Mahdi, B. Herzog and R. Schl6gl, this volume.
[23] N.D. Spencer, R.C. Schooumaker and G.A. Somorjai, J. Catal. 74 (1982) 129.
[24] G.A. Somorjai and N. Materer, this volume.
[25] J.A. Dumesic, H. Topsee and M. Boudart, Proc. Natl. Acad. Sci. US 74 (1977) 806.
[26] J.A. Dumesic, H. Topsoe and M. Boudart, J. Catal. 37 (1975) 513.
[27] R. Brill, E.L. Richter and E. Ruth, Angew. Chem. 6 (1967) 882.
[28] M.A. Richard and R.H. Vanderspurt, J. Catal. 94 (1985) 563.
[29] K.C. Waugh, D.A. Butler and B.E. Hayden, this volume.
[30] D.C. Silverman and M. Boudart, J. Catal. 77 (1982) 208.
[31] H. Tops~e, N. Topsee, H. Bohlbro and J.A. Dumesic in: Proc. 7th Int. Congr. on Catalysis,
Part A, eds. T. Seiyama and K. Tanabe (Kodansha, Tokyo, 1981) p. 247.
[32] M. Boudart and D.G. L6ffler, J. Phys. Chem. 88 (1984) 5763.
[33] J.A. Dumesic, D.F. Rudd, L.M. Aparicio, J.E. Rekoske and A.A. Trevifio, The Microkinetics
of Heterogeneous Catalysis (American Chemical Society, Washington, 1993) p. 155.
[34] J.U. Nwalor, J.G. Goodwin Jr. and P. Biloen, J. Catal. 117 (1989) 121.

Das könnte Ihnen auch gefallen