Sei sulla pagina 1di 11

JOURNAL OF CATALYSIS 116, 119-129 (1989)

Kinetic Simulation of Ammonia Synthesis Catalysis


J. A. DUMESIC*-l AND A. A. TREVIAO~
*Department of Chemical Engineering, University of Wisconsin, 1415 Johnson Drive, Madison, Wisconsin
53706. and TShanahan Valley Associates, 5118 Sherwood Road, Madison, Wisconsin 53711

Received September 2, 1988; revised October 20, 1988

Kinetic simulation studies were carried out and showed that kinetic and spectroscopic data
collected on iron single crystals under ultrahigh vacuum conditions can be used to predict the rate
of ammonia synthesis over a commercial iron catalyst in a plug-flow reactor operating under
industrial reaction conditions (i.e., 720 K, 107atm). The results agree with earlier calculations by P.
Stoltze and .I. K. Norskov [Surf. Sci. Lett. 197, L230 (1988);Phys. Rev. Letf. 55,2502 (1985);Phys.
Ser. 36, 824 (1987); J. Catal. 110, 1 (1988)l. Furthermore, the kinetic simulations confirm the fact
that the dissociation of molecular dinitrogen precursor species is a slow step in the mechanism;
however, other steps may also become slow depending on reaction conditions. Atomic nitrogen
appears to be the most abundant reactive intermediate on the iron surface, except at the reactor
inlet, where a variety of adsorbed species may compete for surface sites. Finally, the kinetic
simulation results are also shown to describe the performance of iron catalysts in laboratory
reactors operating at pressures from 1 to 20 atm. 0 1989 Academic Press, Inc.

INTRODUCTION abundant surface intermediates. The use of


such kinetic simulations in directing experi-
An understanding of the kinetics of mental research in catalysis is discussed
chemical processes on solid surfaces has elsewhere in a preliminary fashion (I). The
been, and remains today, one of the most objective of the present paper is to use
important areas of heterogeneous catalysis. these numerical techniques to probe the ki-
In general, the information obtained by sur- netics of ammonia synthesis over iron cata-
face scientists on single crystal surfaces un- lysts.
der ultrahigh vacuum conditions should be There has been some disagreement (2, 3)
combined with studies of high-surface-area as to whether rates of surface reactions on
catalysts under reaction conditions to pro- single crystal iron surfaces under ultrahigh
vide the knowledge needed to postulate vacuum conditions [e. g., (4-16)] can be
realistic reaction mechanisms and to esti- used to predict rates of ammonia synthesis
mate the corresponding rate constants. over high-surface-area iron catalysts under
Progress in this direction has been slow, industrial reaction conditions. According to
at least in part due to the difficulty in ana- Stoltze and Norskov (3, 17-19), it appears
lyzing quantitatively complex reaction possible to extrapolate, with considerable
mechanisms. During the past several years, success, the results from single crystal
we have been developing numerical tech- studies to predict the performance of an in-
niques at the Unversity of Wisconsin and dustrial, potassium-promoted ammonia
Shanahan Valley Associates to explore the synthesis catalyst. In contrast, Bowker et
quantitative aspects of reaction mecha- al. (2, 20) report that this extrapolation pre-
nisms that do not make a priori assump- dicts rates of ammonia synthesis that are 40
tions about rate-determining steps or most times lower than those measured under in-
dustrial conditions.
’ Author to whom correspondence should be ad- Regardless of which of the two research
dressed. groups is correct, it should be noted that it
119
0021-9517/89$3.00
Copyright 0 19R9 by Academic Press, Inc.
All rights of reproduction in any form reserved.
120 DUMESIC AND TREVIRO

is rather impressive to be able to extrapo- used a value for the active site density that
late over 11 orders of magnitude in pressure was too low by a factor of about 50.
(from 10P6Tot-r to 100 atm) to predict the
rate of ammonia synthesis within a factor of KINETIC MODEL FORMULATION
40. Despite this fact, the origin of the differ- The kinetic mechanism employed by
ence between extrapolations of these two Stoltze and Norskov is depicted below:
groups remains unclear. A significant dif-
N2 + * = N2* (1)
ference between the approach of Stoltze
and Norskov and the method of Bowker et Nz* + * = 2N* (2)
al. is that the former group assumed that
H* + N* = NH* + * (3)
the formation of atomic, chemisorbed nitro-
gen from a molecular precursor state was NH* + H* = NH** + * (4)
rate determining, whereas Bowker et al. NH2* + H* = NH+ + *
made no assumptions regarding the nature (5)
of rate-determining steps. It is possible that NHJ* = NH3 + * (6)
this assumption and differences in the nu- H2 + 2* = 2H* (7)
merical methods used by these authors are
responsible for the different results re- These authors assumed that step (2) was
ported. rate determining and that all species were
It will be shown that the kinetic simula- competively adsorbed. They used the data
tions of the present work are in agreement of Ertl and co-workers (4, 7, 8, 12-26) to
with the calculations of Stoltze and estimate the rate constants for this step,
Norskov. Specifically, the ammonia con- and spectroscopic data to estimate the equi-
centration in the effluent of an industrial re- librium constants for the remaining steps.
actor operating at 107 atm can be predicted Although Stoltze and Norskov assumed in
with a relative uncertainty of less than 15%. their calculations that step (2) was rate de-
Moreover, we will show that the same ki- termining, they also estimated the rate con-
netic model predicts well the rate of ammo- stants for step (3). We have used these lat-
nia synthesis over iron catalysts in labora- ter estimates in our kinetic simulations of
tory reactors operating at pressures from 1 the complete reaction network.
to 20 atm. Regarding the conclusions of Table 1 summarizes our estimates of the
Bowker et al., it appears that these authors rate constants from the work of Stoltze and

TABLE 1
Rate Constants for Model of Stoltze and Norskov

Forward rate constant Reverse rate constant

Preexponential Activation Preexponential Activation


factor energy factor energy
Step (kJ/mol) (kJ/mol)

1 3.33 x lo3 Torr-’ s-’ 0.0 1.87 x 10’4s-1 43.1


2 4.29 x lo9 ss’ 28.5 1.32 x lo9 s-’ 155.0
3 1.83 x lo9 SK’ 81.3 1.15 x lO’s-’ 23.2
4 1.31 x 10’3 s-1 36.4 1.38 x lo’* s-1 0.0
5 3.88 x 10’3 s-1 38.7 2.33 x 10” s-’ 0.0
6 3.67 x lOI s-’ 39.2 2.38 x lo5 Torrr’ s-l 0.0
7 9.23 x lo3 Torx-’ s-’ 0.0 3.24 x lOi s-’ 93.8
AMMONIA SYNTHESIS KINETICS 121

TABLE 2
Rate Constants for Model of Bowker et al. (Case 1)

Forward rate constant Reverse rate constant

Preexponential Activation Preexponential Activation


factor energy factor energy
Step (kJ/mol) (kJ/mol)

3.85 x lo3 Torr-I ss’ 0.0 1.58 x 10’0s-1 46.0


2.63 x lo6 s-’ 31.3 2.63 x lOI s-’ 199.0
1.05 x 10’2 s-1 64.9 1.66 x 10’2s-1 18.8
8.32 x 10” ss’ 64.9 1.66 x 10’2s-1 18.8
6.61 x 10’2 s-1 64.9 1.66 x 10’2s-1 18.8
1.00 x 10” s-1 52.7 2.43 x lo4 Torr-’ s-’ 0.0
3.85 x 10sTorr-I ss’ 0.0 1.58 x lOI s-’ 0.0
2.63 x lo’* s-’ 0.0 2.63 x lOI ss’ 92.0

Norskov. We have combined their reported H* + N* = NH* + * (3)


values of the rate constants for steps (2) and
NH* + H* = NH2* + * (4)
(3) and the equilibrium constants for steps
(1) and (4)-(7) at 723 K with the reported NH** + H+ = NH3* + * (5)
standard enthalpy changes for each step, to
NHJ* = NH3 + * (6)
estimate the preexponential factors and ac-
tivation energies given in Table 1. In this Hz + * = HZ* (8)
estimate, we have assumed for steps (1) and
HZ* + * = 2H* (9)
(4)-(7) that the activation energy is equal to
zero in the exothermic direction. It should Two sets of activation energies and preex-
be noted that the true temperature depen- ponential factors for these steps were re-
dence of the rate and equilibrium constants ported by Bowker et al., depending on
is slightly more complex than the Arrhenius how one interprets the temperature-pro-
form used in the present paper, due to the grammed desorption data of Ertl and co-
temperature dependence of the partition workers. The two sets of rate constants are
functions. Stoltze and Norskov have used summarized in Tables 2 and 3.
this more complete temperature depen-
dence, and we have simply matched an Ar- CONSISTENCY OF KINETIC RATE
rhenius form of the rate constants to their CONSTANTS
values at 723 K (a typical ammonia synthe- As noted above, three sets of kinetic rate
sis reaction temperature). constants (Tables 1-3) have been reported
The kinetic mechanism employed by in the literature to be consistent with stud-
Bowker et al. was essentially the same as ies on iron single crystal surfaces. Two spe-
that given above, with the exception that cific sets of experiments are particularly
they allowed the dissociative adsorption of relevant in this respect: (i) sticking coeffi-
hydrogen to proceed via a precursor state cient measurements and (ii) temperature-
(i.e., step 7 was replaced by steps 8 and 9). programmed desorption studies.
This mechanism is summarized below: Ertl et al. (16) report that the initial stick-
N2 + * = NZ* ing coefficient for dissociative adsorption of
(1) nitrogen on Fe(ll1) (the most active sur-
N2* + * = 2N* (2) face plane for ammonia synthesis) at 430 K
122 DUMESIC AND TREVIAO

TABLE 3
Rate Constants for Model of Bowker et nf. (Case 2)

Forward rate constant Reverse rate constant

Preexponential Activation Preexponential Activation


factor energy factor energy
Step (kJ/mol) (kJ/mol)

3.85 x lo3 Torr-’ s-l 0.0 1.58 x lOto s-’ 46.0


1.66 x 109s-1 31.4 1.66 x 109s-1 144.0
1.05 x 10” s-1 53.5 1.66 x 10’2s-1 16.7
8.32 x 10”’ s-’ 53.5 1.66 x 10’2 s-1 16.7
6.61 x 10” s-1 53.5 1.66 x 10’2 s-1 16.7
1.00 x 10’3 s-1 52.7 2.43 x lo4 Torr-’ s-’ 0.0
3.85 x 1WTorr-’ s-’ 0.0 1.58 x lOI s-’ 0.0
2.63 x lOI s-’ 0.0 2.63 x lOI s-’ 92.0

ranges from 5 X 10e6on the clean surface to ture dependence of the sticking coefficient,
4 x 10m5on an optimally promoted potas- while the rate constants of Bowker et al.
sium surface. When the aforementioned (Case 2) are less successful in this respect.
three sets of rate constants for steps (1) and Boszo et al. (13) have reported tempera-
(2) are extrapolated to 430 K, the following ture-programmed desorption (TPD) spectra
initial sticking coefficients are calculated for nitrogen on Fe( 111). The temperature at
(assuming steady state for adsorbed dinitro- which the rate of desorption reaches a max-
gen, as suggested by Ertl et al.): 1.4 x 10es imum (for a heating rate of 10 K/s) was de-
for the model of Stoltze and Norskov, 1.2 x termined to be in the range 850 to 890 K,
10e4for the model of Bowker et al. (Case with the higher temperatures being ob-
I), and 1.0 x 10e2for Bowker et al. (Case served for lower intial surface coverages,
2). It is clear that the values for the models as expected for second-order desorption. It
of Stoltze and Norskov and Bowker et al. is possible to simulate such a TPD experi-
(Case 1) are in general agreement with the ment by solving simultaneously the two dif-
reported low value of the sticking coeffi- ferential equations associated with time de-
cient. The sticking coefficient predicted for pendencies of the surface coverages by
Case 2 of Bowker et al. appears to be too molecular and atomic nitrogen for a linear
high. rate of temperature increase. The initial
The apparent activation energy of the state for this problem is a surface at 77 K
sticking coefficient for dissociative adsorp- with a specific coverage by atomic nitrogen
tion of nitrogen on potassium-promoted (ranging from 0.2 to 0.95). When the preex-
Fe( 111) has been reported by Ertl et al. (16) ponential factors and activation energies re-
to be - 12 kJ/mol at temperatures near 400 ported by Stoltze and Norskov for steps (1)
K. The following values of this apparent and (2) are used in such a dynamic simula-
activation energy are predicted by the three tion, the predicted peak temperature for the
sets of rate constants for steps (1) and (2): TPD experiment of Boszo et al. ranges
-13 kJ/mol for the model of Stoltze and from 820 to 870 K. The predicted peak tem-
Norskov, -13 kJ/mol for the model of peratures for the rate constants of Bowker
Bowker et al. (Case l), and -0.5 kJ/mol for et al. range from 790 to 830 K for Case 1,
Bowker et al. (Case 2). The rate constants and from 760 to 820 K for Case 2. It appears
given by Stoltze and Norskov and Bowker that the rate constants of Stoltze and
et al. (Case 1) predict the proper tempera- Norskov and Bowker et al. (Case 1) are in
AMMONIA SYNTHESIS KINETICS 123

general agreement with the experimental TABLE 4


TPD spectra, while the predicted peak tem- Reactor Data for Ammonia Synthesis (21)
peratures for Bowker et al. (Case 2) seem
to be too low. Reactor type Plug flow
Reactor volume 5 cm3
Finally, the product of the equilibrium
Reactor pressure 107 atm
constants for the elementary steps of the Hz-to-N2 ratio 3:l
reaction mechanism must give the overall Space velocity 16,000 h-’
equilibrium constant for the stoichiometric Ammonia (% in effluent) 13.2
reaction (the equilibrium constant for each
step appearing in this product raised to a
power equal to the number of times that the
step occurs in the overall reaction). At a lar experimental point has been chosen by
temperature of 723 K, the overall equilib- Bowker et al. as a test case. The reactor
rium constant for ammonia synthesis (form- data for this point are summarized in Ta-
ing 2 mol of ammonia) is approximately 8 X ble 4.
10-i’ Tori-*. The values predicted by the Other relevant data are needed to simu-
three sets of rate constants in Tables l-3 late the performance of the reactor: (1) the
are 9 x lo-“, 5 X lo-“, and 7 X lo-“, reactor was shown to operate as an isother-
respectively. At a total pressure of 107 atm mal plug-flow reactor; (2) the catalyst vol-
and for a stoichiometric H2/N2 gas mixture, ume in the reactor was 2.5 cm3 (the remain-
these equilibrium constants correspond to ing volume was filled with a solid diluent);
equilibrium ammonia concentrations of 17, (3) the catalyst bed density was 2.5 g/cm3;
14, and 15%, respectively. These values are (4) the catalyst pellet size was 0.5 mm; (5)
in general agreement with the ammonia the space velocity was measured at normal
concentration of 17% given by the true temperature and pressure (NTP), and was
equilibrium constant for this reaction (with- based on the volume of the catalyst; and (6)
out making corrections for nonideal gas be- the iron surface site density of the catalyst
havior) . corresponded to about 6 X 10m5mollg [as
In summary, it appears that the rate con- estimated by CO chemisorption (21)]. Im-
stants estimated by Stoltze and Norskov portantly, the ammonia synthesis kinetics
and Bowker et al. (Case 1) are in general under these conditions were not limited by
agreement with sticking coefficient mea- transport phenomena (21).
surements for dissociative nitrogen adsorp- We have modeled the above plug-flow re-
tion on potassium-promoted Fe( 111) sur- actor by a series combination of 10,000
faces, with TPD spectra of nitrogen on well-mixed reactors. Actually, this simula-
Fe(ll1) crystals, and with the overall ther- tion can be carried out without significant
modynamics for ammonia synthesis. The loss of precision using only 1000well-mixed
rate constants of Bowker et al. (Case 2) do reactors. In each well-mixed reactor, we
not appear to be as successful in reproduc- solve simultaneously the steady-state equa-
ing the data obtained on single crystal sur- tions for all surface intermediates, the site
faces. The next sections of this paper ex- balance on the catalyst surface, and the ma-
plore the utility of the above rate constants terial balances for the gaseous species.
in explaining ammonia synthesis kinetics Since there is a change in moles during re-
over a wide range of experimental condi- action, it is perhaps useful to note the three
tions. material balances that we have used in each
REACTOR SIMULATION
of the well-mixed reactors:
The experimental data that both of the PS
-= @.5)(pN,n - PN2,0ut)
above groups have attempted to explain F (l - 2pNm lp)(l - 2pN,,in/p)(RN,)
were reported by Nielsen (21). One particu- (1)
FIG. 3. Calculations, according to the model of
FIG. 1. Calculations of the ammonia concentration Bowker et al. (Case l), of fractional surface coverages
versus dimensionless longitudinal distance from reac- by predominant adsorbed species versus dimension-
tor inlet: (0) Stoltze and Norskov, (A) Bowker ef al. less longitudinal distance from reactor inlet: (0) H*,
(Case l), (W) Bowker et al. (Case 2). (A) N*, (W N2*.

partial pressures at the outlet and inlet of


each well-mixed reactor.

RESULTS OF KINETIC SIMULATIONS UNDER


where P is the total pressure, S is the num- HIGH-PRESSURE CONDITIONS
ber of sites in each well-mixed reactor, F is The results of our kinetic simulations for
the total molar flow rate into the first reac- the experimental condition of Nielsen are
tor of the series combination, RN~is the rate summarized in Figs. 1-4, in which we re-
of consumption of gaseous N2 in step (2) port the ammonia concentration (in %) as
(turnover frequency units, s-l), and the well as the fractional surface coverages by
designations “out” and “in” refer to the the predominant adsorbed species versus

FIG. 2. Calculations, according to the model of FIG. 4. Calculations, according to the model of
Stohze and Norskov, of fractional surface coverages Bowker er al. (Case 2), of fractional surface coverages
by predominant adsorbed species versus dimension- by predominant adsorbed species versus dimension-
less longitudinal distance from reactor inlet: (@) H*, less longitudinal distance from reactor inlet: (0) H*,
(A) N*. (4 N*, (W N2*.
AMMONIA SYNTHESIS KINETICS 125

TABLE 5
Comparison of Experimental Observations with Predictions from Kinetic Simulations for Ammonia Synthesis

Experiment: Experimental Predicted value


Observation Value
Stoltze Bowker et al. Bowker et al.
and Norskov (Case 1) (Case 2)

Nielsen (21)
Ammonia concentration (%) 13.2 11.1 10.2 14.5
Dumesic et al. (22)
Turnover frequency (s-l) 0.04 0.19 0.16 0.13
Topsoe et al. (23)
Turnover frequency (s-l) 0.50 0.28 0.20 0.14
Spencer et al. (9)
Turnover frequency (s-r) 12.7 60.0 25.0 40.0

longitudinal distance from the inlet of the respectively (see Fig. 1). This comparison
plug-flow reactor. We have included these is summarized in Table 5.
results versus reactor position to determine In Fig. 5, we report the departures from
whether the nature of the most abundant equilibrium for the slowest steps of the re-
reactive intermediate on the surface de- action mechanism versus longitudinal dis-
pends on the position of the catalyst in the tance through the plug-flow reactor. The
reactor, as will be discussed later. The ef- departure from equilibrium for a given step
fluent ammonia concentrations predicted is defined as the ratio of the net rate of the
by the models of (1) Stoltze and Norskov, step to the forward rate. This value ap-
(2) Bowker et al. (Case l), and (3) Bowker proaches zero for an equilibrated step.
et al. (Case 2) are 11.1, 10.2, and 14.5%, It can be seen that all three of the above

A c

Dimensionless Reactor Distance

FIG. 5. Departure from equilibrium for slowest elementary steps in ammonia synthesis: (A) Stoltze
and Norskov, (B) Bowker et al. (Case l), (C) Bowker et al. (Case 2), (W) Step 2, (0) Step 3 in (A) and
Step 5 in (B) and (C), A) Step 6.
126 DUMESIC AND TREVIfiO

kinetic simulations predict ammonia con- thesis reactor operating at 107 atm. While
centrations in the reactor effluent near the this is an important result and shows the
experimental value of 13.2%. Therefore, utility of the kinetic models under industrial
we conclude that it is possible to extrapo- reaction conditions, this experimental point
late successfully the results from single does not provide a particularly sensitive
crystal studies to predict the performance test of the kinetic models, since the reactor
of an industrial ammonia synthesis catalyst. is operating near equilibrium (i.e., the am-
According to Bowker et al. the use of the monia concentration at the effluent is 78%
rate constants in Tables 2 and 3 should give of the equilibrium value). Thus, a model
ammonia concentrations of 3.0 and 5.4%, that is consistent with the overall thermo-
respectively. Indeed, we can obtain these dynamics of ammonia synthesis may give a
values if we decrease the value of (PSIF) by good prediction of the ammonia concentra-
a factor of 50. (This would correspond, for tion leaving the reactor under the afore-
example, to a lower value of the site density mentioned conditions; however, this model
or a lower catalyst density used by these may fail to provide a good estimate of the
authors.) Furthermore, at this lower value ammonia synthesis turnover at conditions
of PSIF, we find the same values of the sur- far from equilibrium. We now address this
face coverages by adsorbed species as problem in the present section with respect
those reported by Bowker et al. We thus to the models of Stoltze and Norskov and
conclude that the difference betweeen the Bowker et al. The comparisons that are
results of Stoltze and Norskov and those of presented below are summarized in Table
Bowker et al. is due to different values of 5.
PSIF used by these authors. Our analysis of Ammonia synthesis turnover frequencies
the reactor data of Nielsen agrees with the were measured by Dumesic et al. (22) on
calculations of Stoltze and Norskov. various magnesia-supported iron catalysts
It is useful to comment on the results in at atmospheric pressure and conversions
Figs. 2-5. All three kinetic simulations far from equilibrium. This reaction was
agree that N* is the most abundant surface found to be structure sensitive, with higher
intermediate in the reactor, except at the turnover frequencies observed for larger
reactor inlet where a variety of adsorbed ion particles, which presumably exposed
species compete for surface sites (see Figs. primarily (111) planes. The apparent turn-
2-4). The results of all the kinetic simula- over frequency for ammonia synthesis over
tions also agree with the well-known fact large metallic iron particles was reported to
that the dissociation of dinitrogen (step 2) is be 0.04 s-i in a stoichiometric Hz/N2 gas
slow (see Fig. 5): If the rate constants of mixture at a temperature of 678 K and at
Stoltze and Norskov are used, then the first ammonia concentration equal to 20% of the
hydrogenation of atomic nitrogen (step 3) is equilibrium value. The reactor employed in
the next slowest step at the reactor inlet. In these studies was a plug-flow reactor, and
contrast, if the rate constants of Bowker et the apparent turnover frequency was calcu-
al. are used, then steps (5) and (6) involving lated as the average number of ammonia
NH+ are the next slowest steps at the reac- molecules produced in the reactor per sec-
tor inlet. ond per iron surface site as titrated by CO
chemisorption. This value is described
AMMONIA SYNTHESIS TURNOVER more correctly as a site time yield.
FREQUENCIES IN LABORATORY REACTORS We have used the rate constants given in
It was shown above that each of three Tables l-3 to predict the site time yields of
models considered in this paper predicts ammonia at the conditions of Dumesic et
adequately the ammonia concentration at al. The following values were obtained:
the effluent of a commercial ammonia syn- 0.19 s-i for the model of Stoltze and
AMMONIA SYNTHESIS KINETICS 127

Norskov, 0.16 s-l for the model of Bowker nia synthesis reaction conditions employed
et al. (Case l), and 0.13 s-i for Bowker et by Spencer et al., we find that the turnover
al. (Case 2). These values are at most a frequency depends on the ammonia pres-
factor of 5 too high. This is considered to be sure. This is due to an inhibition of the for-
good agreement when it is remembered that ward rate by ammonia, since the rate of
(i) the catalyst studied by Dumesic et al. ammonia decomposition is negligible under
was not promoted with potassium and (ii) these experimental conditions. For an am-
CO chemisorption may overestimate the monia pressure between 5 and 10 Torr (typ-
number of active sites. ical values for the experiments reported),
Topsoe et al. (23) have used nitrogen the following values of the ammonia syn-
chemisorption to titrate the active sites for thesis turnover frequency are predicted by
ammonia synthesis on a series of metallic the kinetic models: 60 s-i for the model of
iron catalysts. Site time yields of ammonia Stoltze and Norskov, 25 s-l for the model
were determined by these authors at 673 K of Bowker et ul. (Case l), and 40 s-i for
using a stoichiometric HZ/N2 gas mixture at Bowker et al. (Case 2). The agreement of
atmospheric pressure and operating with an the predicted values with the experimental
effluent ammonia concentration equal to result is good, especially when it is noted
15% of the equilibrium value. The site time that the Fe(l11) surface studied by Spencer
yield of ammonia reported for a potassium- et al. did not contain potassium.
promoted iron catalyst under these condi- In summary, it can be seen in Table 5 that
tions was 0.5 s-l. This value agrees well the kinetic models addressed in the present
with the following predictions of the kinetic paper provide good predictions of the rate
models addressed in the present paper: 0.28 of ammonia synthesis over a range of tem-
s-i for the model of Stoltze and Norskov, peratures (678-748 K), a range of pressures
0.20 s-l for the model of Bowker et al. (I-20 atm), and a range of ammonia con-
(Case I), and 0.14 s-l for Bowker et al. centrations (1.5-20% of the equilibrium
(Case 2). Furthermore, if nitrogen chemi- concentration). The success of these ki-
sorption does not necessarily titrate all of netic models in predicting the performance
the active sites but instead this uptake is of an iron catalyst at 723 K and 107 atm was
simply proportional to the active site den- demonstrated in the previous section. Fur-
sity [as suggested by Topsoe et al. (23)], thermore, two of these kinetic models [i.e.,
then the agreement between the experimen- Stoltze and Norskov and Bowker et al.
tal value of the turnover frequency and the (Case I)] provided adequate descriptions of
predicted values becomes even better. the sticking coefficient for dissociative ni-
Spencer et al. (9) measured the turnover trogen adsorption and the temperature-pro-
frequency for ammonia synthesis over an grammed desorption spectra of nitrogen
Fe(l11) single crystal surface at a total from the Fe( Ill) surface. We thus reach the
pressure of 20 atm and a temperature of 748 remarkable conclusion that relatively sim-
K. For a stoichiometric HZ/N2 gas mixture, ple kinetic models are able to reproduce ni-
the turnover frequency was determined to trogen adsorption/desorption and ammonia
be 12.7 s-i. The ammonia concentrations synthesis behavior under conditions rang-
employed in this study were at most 10 ing from ultrahigh vacuum to 107 atm.
Torr, i.e. less than 1.5% of the equilibrium
concentration. Thus, these experiments SIGNIFICANCE OF KINETIC MODELS
provide a good test of ammonia synthesis It is perhaps surprising that studies of ad-
kinetic models under conditions far from sorption/desorption phenomena on single
equilibrium. crystal surfaces could be used to predict the
When the three kinetic models of the turnover frequency for ammonia synthesis
present study are applied under the ammo- on iron catalysts under a wide range of ex-
DUMESIC AND TREVIP;JO

perimental conditions. This is especially Stoltze and Norskov and Bowker et al.
unexpected since any possible coverage de- (Case 1) can be interpreted as being repre-
pendence of the rate constants was not sentative of surfaces covered primarily by
taken into account. Furthermore, it would atomic nitrogen. In fact, these are the con-
appear to be difficult to determine the rate ditions that were employed during the tem-
constants for the seven or eight elementary perature-programmed desorption studies of
steps of the reaction mechanism with suffi- iron single crystal surfaces. It is possible
cient accuracy to conduct ammonia synthe- that the kinetic models would be less suc-
sis kinetic simulations with precision. In ’ cessful under conditions where the nitrogen
view of these considerations, why do the coverage is very small or where atomic ni-
kinetic models addressed in this study work trogen is no longer the most abundant sur-
so well? face intermediate.
First, under all of the experimental con- In closing, we note that it will be impor-
ditions explored in the present paper, the tant in the future to continue the compari-
dissociation of adsorbed dinitrogen (step 2) son of results of single crystal studies with
is the slowest step in the mechanism. [This results of studies carried out on high-sur-
is even true for the conditions of Spencer et face-area catalysts under industrial reaction
al., where the forward rate of step (2) is at conditions. Studies of this nature are essen-
least an order of magnitude slower than the tial for kinetics-assisted catalyst design.
forward rate of the next-slowest step.] In
addition, adsorbed nitrogen is the most
ACKNOWLEDGMENTS
abundant nitrogen-containing surface inter-
mediate in all cases. Thus, the rate con- We thank Haldor Topsoe A/S for support and en-
couragement during this project. We are particularly
stants of steps (3)-(9) are kinetically insig- grateful to Per Stoltze and Jens Norskov for providing
nificant. It is sufficient for the purposes of us with their estimates of the rate and equilibrium con-
predicting ammonia synthesis turnover fre- stants for the elementary steps involved in ammonia
quencies to describe properly the kinetics synthesis. Finally, we thank Professor D. F. Rudd for
his interest, insight, and encouragement in catalyst de-
and thermodynamics of nitrogen adsorption sign.
[steps (1) and (2)], and these are the steps
that have been investigated in detail on sin-
gle crystal iron surfaces. The estimates of REFERENCES
the rate constants for steps (3)-(9) are used 1. Dumesic, J. A., Milligan, B. A., Greppi, L. A.,
only to decide that these steps are, in fact, Balse, V. R., Sarnowski, K. T., Beall, C. E. Ka-
much faster than step (2) and that the sur- taoka, T., Rudd, D. F., and Trevitio, A. A., Ind.
face coverages by NH, species (x > 0) are Eng. Chem. Res. 26, 1399 (1987).
2. Bowker, M., Parker, I., and Waugh, K. C., Surf.
small. Sci. Letf. 197, L223 (1988).
The possible effects of surface coverage 3. Stoltze, P., and Norskov, J. K., Surf. Sci. Lett.
on the rate constants of the ammonia syn- 197, L230 (1988).
thesis mechanism are not known at present. 4. Ertl, G., “Critical Reviews in Solid State and Ma-
It is important to note, however, that the terials Science,” p. 349. CRC Press, Boca Raton,
FL, 1982.
surface coverage by atomic nitrogen is pre- 5. Grunze, M., in “The Chemistry and Physics of
dicted by the kinetic models to be high un- Solid Surfaces and Heterogeneous Catalysis” (D.
der all of the experimental reaction condi- A. King and D. P. Woodruff, Eds.), Vol. 4, p. 413.
tions investigated in this paper. [For Elsevier, Amsterdam, 1982.
example, the lowest nitrogen coverage pre- 6. Grunze, M., Golze, M., Hirschwald, W., Freund,
H. J., Pulm, H., Seip, H., Kuppers, J., and Ertl,
dicted by the model of Stoltze and Norskov G., Phys. Reu. Left. 53, 850 (1984).
and Bowker et al. (Case 1) is 0.4 for the 7. Ertl, G., in “Catalysis, Science and Technology”
experimental conditions of Spencer et al.1 (J. R. Anderson and M. Boudart, Eds.), Vol. 4, p.
Thus, the rate constants reported by 208. Springer-Verlag, Berlin, 1983.
AMMONIA SYNTHESIS KINETICS 129

8. Ertl, G., Catal. Reu. Sci. Eng. 21, 201 (1980). 16. Ertl, G., Lee, S. B., and Weiss, M., Surf. Sci. 114,
9. Spencer, N. D., Schoonmaker, R. C., and Somor- 527 (1982).
jai, G. A., J. C&z/. 74, 129 (1982). 17. Stoltze, P., and Norskov, J. K., Phys. Rev. iett.
10. Spencer, N. D., Schoonmaker, R. C., and Somor- 55, 2502 (1985).
jai, G. A., Nature (London) 294, 1643 (1981). 18. Stoltze, P., Phys. Ser. 36, 824 (1987).
19. Stoltze, P., and Norskov, J. K., J. Cutul. 110, 1
Il. Grunze, M., Golze, M., Fuhler, J., Neumann, M.,
and Schwartz, E., in “Proceedings, 8th Intema- 2. (1988).
tional Congress on Catalysis, Berlin,” Vol. 4, p. Bowker, M., Parker, I., and Waugh, K., Appl.

133. Verlag Chemie, Weinheim, 1984. Cutull4, 101 (1985).
21. Nielsen, A., “An Investigation of Promoted Iron
12. Ertl, G., Grunze, M., and Weiss, M., J. Vuc. Sci.
Catalysts for the Synthesis of Ammonia.” Gjel-
Technol. 13, 314 (1976).
lerup, Copenhagen, 1968.
13. BOSZO, F., Ertl, G., Grunze, M., and Weiss, M., 22, Dumesic, J. A., Topsoe, H., Khammouma, S.,
J. Caral. 49, 18 (1977). and Boudart, M., J. Cutul. 37, 503 (1975).
14. Boszo, F., Ertl, G., Grunze, M., and Weiss, M., 23. Topsoe, H., Topsoe, N., Bohlbro, H., and Dume-
Appl. Surf. Sci. 1, 103 (1977). sic, J. A., in “Proceedings, 7th International Con-
15. Weiss, M., Ertl, G., and Nietsche, F., Appl. Surf. gress on Catalysis” (T. Seiyama and K. Tanabe,
Sci. 2, 614 (1979). Eds.), p. 247. Elsevier, New York, 1980.

Potrebbero piacerti anche