Ammonia - Catalysis and Manufacture PDF
Ammonia - Catalysis and Manufacture PDF
With contributions by
K. Aika, L. 1. Christiansen, I. Dybkjaer,
1. B. Hansen, P. E. H0jlund Nielsen,
A. Nielsen, P. Stoltze, K. Tamaru
Springer-Verlag
Berlin Heidelberg New York
London Paris Tokyo
Hong Kong Barcelona Budapest
Editor:
Anders Nielsen
Haldor Tops0e A/S
NYill011evej 55,2800 Lyngby/DK
This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, re-use of illustrations, recitation,
broadcasting, reproduction on microfilms or in other ways, and storage in data banks. Duplication
of this publication or parts thereof is only permitted under the provisions of the German Copyright
Law of September 9, 1965, in its current version, and a copyright fee must always be paid.
The use of registered names, trademarks, etc. in this publication does not imply, even in the absence
of a specific statement, that such names are exempt from the relevant protective laws and regulations
and therefore free for general use.
This book owes its existence to Dr. Ekkehard Fluck, Director of the Gmelin-
Institut. Dr. Fluck suggested that the individual chapters could be written by
staff members of Haldor Topsoe A/S, and that emphasis should be given to the
industrial manufacture of ammonia.
Upon careful consideration it was decided to ask two distinguished experts
in catalysis, Prof. Kenzi Tamaru of the Science University of Tokyo and Prof.
Ken-ichi Aika of the Tokyo Institute of Technology to write the chapter on
ammonia synthesis on non-iron catalysts.
When I started to work in catalysis more than 50 years ago, first as a student
in the Institute of Physical Chemistry with the late Prof. J. N. Bf0nsted and
subsequently in the laboratories of Dr. Haldor Topsoe, the use of catalytic
processes by industry and the literature on catalytic studies was still somewhat
limited and allowed a single person to reasonably acquaint himself with the field.
Today, catalysis is a step in the manufacture of most chemical products and
most refinery streams undergo catalytic reactions. The number of physical tools
applied to the study of catalysts is impressive and the literature on catalysis
overpowering. A reasonably comprehensive volume on the topic of ammonia
synthesis had to be a team effort. The first chapter entitled, "Thermodynamic
properties in Ammonia Synthesis" is written by Dr. Lars J. Christiansen. It should
be emphasized that this chapter does not contain the complete thermodynamics
on ammonia, but concentrates on the thermodynamic properties used for the
design and operation of ammonia synthesis units.
The second chapter "Structure and Surface Chemistry of Industrial Am-
monia Synthesis Catalysts" is written by Dr. Per Stoltze. This chapter deals with
the structure and surface chemistry of iron-based ammonia synthesis catalysts of
the type used by industry. Certain studies of single crystal surfaces are included
to the extent that they serve to add information to the main topic. This chapter
includes a presentation of the unreduced catalyst; the reduction process and the
bulk and surface structure of a reduced catalyst. A thorough discussion is given
of the different states of sorption of nitrogen and of chemisorption of hydrogen,
carbon oxides, ammonia and oxygen. The last part of the chapter gives a
detailed account of the mechanism of ammonia synthesis on iron.
The third chapter is written by Profs. Ken-ichi Aika and Kenzi Tamaru,
both of whom have contributed prominently to our knowledge of ammonia
synthesis. Their chapter is entitled, "Ammonia Synthesis over Non-Iron Catalysts
and Related Phenomena". It is recalled that osmium was indeed used in the early
vi Preface
Anders Nielsen
Table of Contents
Chapter 1
Thermodynamic Properties in Ammonia Synthesis
L. 1. Christiansen . . . . . . . . . . . . . . . . . . . .
Chapter 2
Structure and Surface Chemistry of Industrial Ammonia
Synthesis Catalysts
P. Stoltze. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Chapter 3
Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena
K.-i. Aika and K. Tamaru. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Chapter 4
Kinetics of Ammonia Synthesis and Decomposition
on Heterogeneous Catalysts
1. B. Hansen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Chapter 5
Poisoning of Ammonia Synthesis Catalysts
P. E. H0jlund Nielsen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
Chapter 6
Ammonia Production Processes
I. Dybkjaer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
Chapter 7
Ammonia Storage and Transportation-Safety
A. Nielsen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
Chapter 1
Thermodynamic Properties in Ammonia Synthesis
Lars J. Christiansen
Haldor Tops0e A/S Copenhagen, Denmark
Contents
1.1 Introduction . . . . . . . . . . . . . . 2
1. 7 References . . . . . . . . . . . . 13
1.1 Introduction
The pressure range utilized in ammonia plants implies that the ideal gas law
cannot be used. This is corrected for by use of a compressibility factor Z.,
whereby
pV = ZRT (1)
where p is the pressure, V the molar volume, T the temperature, and R the gas
constant. The compressibility factor can be found from generalized charts, but it
is more common to use an equation of state, where the pressure is given as
a function of temperature, volume, and molecular parameters, such as the
critical properties,
p = f(V,T, parameters) (2)
By use of adequate mixing rules for the molecular parameters, the same type
of equation can be used to calculate mixture properties. Numerous equations of
state have been proposed and a review can be found in Reid et al. [50].
For detailed information on compressibility factors reference is given to the
experimental data by Wiebe, Gaddy [56], by Bartlett et al. [3,4]. These papers
give values of the compressibility factors of hydrogen, nitrogen, and 3: 1 mix-
1 Thermodynamic Properties in Ammonia Synthesis 3
H=IYiH~-H' (3)
i
C p = I YiC~i - C~ (4)
i
where H? and C~i are the pure component ideal gas standard enthalpy and heat
capacity, respectively, an Yi the mole fraction. The enthalpy departure from the
ideal gas, H', can be calculated using rigorous thermodynamics as:
C~= [aH']
aT p
(6)
The pure component ideal gas standard enthalpies or heat capacities are
functions of the temperature only. They are normally represented as tables as
can be found in Rossini [51] and JANAF Tables [24] or fitted to polynomials as
shown in Reid et al. [50] and in Christiansen, Kjrer [7]. The last reference uses
a polynomial of fourth order.
(7)
The heat capacity is then simply represented by:
C~i = a2i + 2a3iT + 3a4iT2 + 4asiT3 (8)
The coefficient ali is determined so that the enthalpy at 25°C is equal to the
enthalpy of formation of the pure component. The enthalpy polynomial can
then, without further modifications, be used to calculate heats of reaction, and
consequently heat balances including chemical reactions.
The polynomial coefficients for hydrogen, nitrogen, and ammonia from
Christiansen, Kjrer [7] are determined from data given in Rossini [51] and
JANAF Tables [24] and are shown in Table 1.1 below:
NH3 - 1.320772 X 104 6.048322 4.125509 X 10- 3 - 3.692310 X 10- 8 - 1.802763 X 10- 10
H2 - 2.112450 X 10 3 7.209790 - 5.559028 X 10- 4 4.846263 X 10- 7 - 8.190294 X 10- 11
N2 - 1.976727 X 103 6.459189 5.182665 x 10- 4 2.032237xl0- 7 -7.654612xlO- 11
For any reaction mixture of hydrogen and nitrogen yielding ammonia, the
standard reaction enthalpy can be evaluated from the data in Table 1.1, if stoi-
chiometric are included. It is largest close to the dew point and decreases with
increasing temperatures.
1 Thermodynamic Properties in Ammonia Synthesis 5
K = aNH3
(9)
a?/22 ai./22
where aj is the activity of component i, which is given as
(10)
where ({Jj is the fugacity coefficient, which is a function of temperature, pressure
and composition. It is conveniently derived from an equation of state, such as
eq. (1) p. 2 as
In({Jj = - 1
RT
J -p )
00
v
[(a
ani T,V,nj,Fi
- -RTJ dV -
V
RTlnZ (11)
Expressions from different equations of state can be found in Reid et al. [50].
Insertion of (11) in (10) makes it possible to separate the ideal gas from the
non-ideal gas contribution. The equation is then
(12)
(14)
6 L. J. Christiansen
Insertion of the heat of reaction derived from equation (7) gives the following
expression for the thermodynamic equilibrium constant as a function of temper-
ature
at a4 z as 3
RInK = ao - T + azlnT + a3 T + 2 T + 3 T (15)
The constant ao is fixed so that equation (14) gives the Gibbs free energy of
formation at 25°C. - 3915 kcaljkmol [24].
The equilibrium composition is then calculated in a two step procedure. The
first step leads to the ideal gas composition as given by Kp and the second step is
the correction for the non-ideal gas behavior by use of Kq>' which includes the
influence of pressure and composition.
The most accurate experimental values appear to be those by Haber,
Rossignol (18), by Schultz, Schaefer [53J, by Haber et al. [21J for 30 atm
pressure, by Larson, Dodge [32J for 10, 30, 50, and 100 atm, and by Larson [30J
for pressures of 300, 600, and 1000 atm. Experimental and calculated data on
ammonia synthesis equilibrium at pressures from 1000 atm to 3500 atm have
been given by Winchester, Dodge [61].
The experimental data at high pressures have been analyzed by different
equations of state to calculate the fugacity coefficients in Kq>. Gillespie [IIJ and
Gillespie, Beattie [12J have used the Beattie-Bridgeman equation of state. The
method by Newton [44J is a generalized method for calculation of fugacity
coefficients.
In Figs. 1.1 to 1.3 the equilibrium % NH3 are shown as a function of
temperature at different pressures and at 0%, 10% and 20% inert (methane and
aTgon) content, respectively. The fugacity coefficient ratio is calculated by the
present author using the Martin-Hou equation of state, see [36]. Of course,
other equations of state could have been used for calculating this coefficient.
It is interesting to note that because of the composition dependence of the
fugacity coefficients the maximum yield of ammonia at high pressures may well
exist at a hydrogen-nitrogen ratio different from 3. At 200 atm and 500°C the
maximum yield at equilibrium is to be expected for a ratio of 2.9 although the
increase in only 0.01 % NH 3.
The content of ammonia in the vapor phase and the amount of dissolved
synthesis gases in the liquid ammonia phase are calculated from expressions for
the f3 value, f3i = Yi/Xb where Yi and Xi are the mole fractions in the gas and liquid
phases, respectively.
1 Thermodynamic Properties in Ammonia Synthesis 7
40~-----+4.--4.~---+--r------+------~
M
I
Z
Correlations for the {3 values are derived from the isofugacity criterion,
which can be written as
40~----~--+-~~~----+-~---1------~
201-------t--
10~----_+------+-~~_+~~~+_--~_i
Fig. 1.2 Percentage of ammonia in equilibrium mixture (67.5% H 2 , 22.5% N 2 , 3% Ar, 7% CH4 ).
The activity coefficient for ammonia is greater than one and the fugacity
coefficient decreases with increasing pressure thereby giving a higher concentra-
tion of ammonia in the gas phase than predicted by the Raoult's law alone.
Experimental measurements of the ammonia concentration in the vapor
phase have been reported by Larson, Black [31] for the equilibrium with
nitrogen and hydrogen from - 22.5 to + 18.7°C and from 50 to 1000 atm.
More recent data are reported by Michels et al. [40]. Their investigation
1 Thermodynamic Properties in Ammonia Synthesis 9
co>
:I:
Z
Fig. 1.3 Percentage of ammonia in equilibrium mixture (60% H 2 , 20% N 2 , 6% Ar, 14% CH 4 )·
covered the temperature interval from 0 to 121.8°C and pressures from 25 atm
to 785 atm. Michels et al. [38] later reported data in the same pressure range but
at temperatures below O°C. Lefrancois, Vaniscotte [33] report data for the
temperature interval - 70 to + 60°C at pressures 300 and 500 kg/cm 2 • These
data agree well with the data by Michels et al. [38,39]. Reddy, Husain [49] have
given data for the ammonia vapor concentration in a mixture with gas phase
mole ratios, H 2 :N 2 :Ar:CH4 , equal to 3:1:0.18:0.44.
The available data for the vapor concentration of ammonia in equilibrium
with a 3: 1 mixture of hydrogen and nitrogen are shown in Fig. 1.4 for pressures
equal to 10,20,30, and 50 MPa. The data used are those published by Michels et
al. These data are in good agreement with the data given by Lefrancois,
Vaniscotte [33], by Heise [22] and by Zeininger [14] at 25°C, whereas the data
given by Larson, Black [31] give higher concentrations of ammonia, in particu-
lar at the high pressures.
In Fig. 1.4 the predicted values ofthe ammonia concentration are also shown
assuming that Raoult's law is valid and neglecting the dissolved gases at 10 and
50 MPa, respectively. It is seen that Raoult's law is not valid.
The data presented by Reddy, Husain [49] also agree with Fig. 1.4 in spite of
the content of inerts in their data. This is also the case for the data given by
10 L. J. Christiansen
20r-----~-------+------_+----+_~
I 15r---t----t----++-
.!:
M
::c
z
~ 10r-----~-------+--~--~~~~
5r-----~--~~~~~~50M~
I...,o'\l.'?J.. ---
. ., \n'" '?:,..--
?-~\l~_'- Fig. 1.4 Equilibrium vapor concentra-
--- tion of ammonia in 3: 1 mixture of
H2 + N2•
280 300 320
Temperature in K
Zeininger [14]. There is hence some evidence that the pressure of inerts will not
change the ammonia vapor concentration significantly although it will increase.
More experimental information is needed in order to obtain more certainty
about this point.
In the model given by Alesandrini et al. [1] ammonia concentrations are
calculated which are too small at the higher pressures. The model has therefore
been revised by Reddy, Husain [49].
Solubilities of the gases hydrogen, nitrogen, argon, methane, and helium in the
form of f3 values (definition see p. 6) are shown in the extensive treatment in
Landolt-Bornstein [29]. The data cover the operating region normally used in
ammonia synthesis.
Data for the solubility of hydrogen in ammonia can be found in Reamer,
Sage [47], in Wiebe, Treamearne [57, 58]. In Heise [22], and in Zeininger [63].
Data for the solubility of nitrogen in ammonia can be found in Reamer, Sage
[48], in Heise [22], and in Zeininger [63]. Data for the solubility of argon in
1 Thermodynamic Properties in Ammonia Synthesis 11
ammonia can be found in Kaminishi [25], in Michels et al. [39], and in Heise
[22]. Data for the solubility of methane in ammonia can be found in Kaminishi
[25], and in Zeininger [63]. Data for the solubility of helium in ammonia can be
found in Heise [22].
Data for the solubility of hydrogen-nitrogen in the ratio 3: 1 in ammonia are
shown in Larson, Black [31], in Michels et al. [40], in Lefrancois, Vaniscotte
[34], in Michels et al. [38], and in Atroshchenko, Gavrya [2]. Data with
methane can also be found in Zeininger [63] and in Konoki et al. [28].
f3 values for hydrogen, nitrogen, argon, and methane evaluated by using. the
method given in Alesandrini et al. [1] for a mixture with the 10% inert gases and
Hz/N z = 3 are shown in Fig. 1.5. The parameters have been evaluated by use of
the data above. f3 values are shown for 10, 20, and 30 MPa, respectively, as
a function of temperature. The same figure also shows experimental data for the
solubility of helium in the range given in Heise [22]. The f3 values decrease in the
order helium, hydrogen, nitrogen, argon, and methane, which corresponds to an
increase in solubility. The f3 values decrease with increase in temperature. The
f3 values are not very sensitive to the content of inert gases, but they increase
slightly (not significant in Fig. 1.5) with increasing inert gas level. The reason is
that argon and methane have higher solubilities than hydrogen and nitrogen.
Very few data sets exist with all components present. Zeininger [63] gives
data for a mixture with hydrogen, nitrogen and methane. The data agree well
with those at high pressure shown in Fig. 1.5 whereas there is some divergence at
low pressure. In this region, however, the solubility is very low.
The transport properties to be discussed in the design of ammonia plants are the
viscosity, the thermal conductivity, and the diffusion coefficients, which are used
in calculation of transport rates in order to determine equipment sizes. The
prediction of the properties for the gases are based on kinetic gas theory,
whereas the properties of liquid ammonia are predicted by empirical correla-
tions based on experimental data.
The properties of liquid ammonia with small amounts of dissolved gases are
calculated by mixing the properties of the pure liquid NH3 with the correspond-
ing properties of the dissolved gases calculated by the kinetic gas theory as
explained below. The mixing rule is the logarithmic mean.
The calculation of transport properties of gases can be divided into three
parts similar to those used in the derivation of thermodynamic properties. The
first part deals with the pure component low pressure values, which are func-
tions of molecular parameters and temperature only. The second part is calcu-
lation of low pressure mixture properties using appropriate mixing rules, and
the third part is the correction for influence of presrure. A detailed discussion of
the kinetic gas theory is given in Hirschfelder et al. [23], and in Reid et al. [50].
The recommendations given in the latter reference are used closely in the
following, since they give a good agreement with the available data for ammonia
synthesis mixtures.
The pure component low pressure viscosity is predicted from the Chapman-
Enskog equation which includes the effect of intermolecular forces. The equa-
tion for the viscosity I'J is:
1'J=26.69~
v (J
(17)
viscosity YJ is known in the form of the ratio AMw/YJ which is equal to a function
of the heat capacity and the various modes of vibration. The method of Bromley
[6] gives acceptable results for the components present in ammonia synthesis
gas.
The binary diffusion coefficients are also predicted using kinetic gas theory. The
potential function used in evaluation of the collision integral is that given by
Neufeld et al. [43] with the modification for polar gases given by Brokaw [5].
The low pressure mixture properties are calculated using an appropriate mixing
rule, which also has been derived from kinetic gas theory. The formula includes
a binary interaction term. For viscosity the interaction parameter given by
Wilke [60] can be used. For thermal conductivity the form given by Lindsay,
Bromley [35] for the binary interaction parameter can be used, and for the
diffusion coefficients the formula given by Wilke [59] can be used.
The correction for the pressure dependence of transport properties of gases can
be tnade by use of various correlations. These include the gas density which in
turn is calculated from the compressibility factor. For viscosity the method
proposed by Dean, Stiel [9] can be used, and for the thermal conductivity the
method proposed by Stiel, Thodos [54] can be used. For the bulk diffusion
coefficient it is normally assumed that it is inversely proportional to the density.
Viscosities of hydrogen-nitrogen and of hydrogen-ammonia mixtures have
been reported in Pal, Barua [46]. Dembovskii [10] has reported viscosity data
for mixtures of ammonia, hydrogen, and nitrogen. Thermal conductivities of
nitrogen-hydrogen-ammonia mixtures are given by Golubev, Kiyashova [15].
Reference is also given to the work by Tsederberg [55]. Binary diffusion
coefficients for hydrogen-ammonia and nitrogen-ammonia are given in Mason,
Monchick [37]. The agreement between these experimental data and those
calculated by the above-mentioned methods is acceptable for design purposes.
1.7 References
1. Alesandrini CG, Lynn S, Prausnitz JM (1972) Ind Eng Chern Process Design Develop 11: 253
2. Atroshchenko VI, Gavrya NA (1959) Zh Prikl Khim 32: 100; J Appl Chern [USSR] 32: 100
3. Bartlett EP, Cupples HL, Tremearne TH (1928) J Am Chern Soc 50: 1275
14 L. 1. Christiansen
4. Bartlett EP, Hetherington HC, Kvalnes HM, Tremearne TH (1930) 1 Am Chern Soc 52: 1363
5. Brokaw RS (1969) Ind Eng Chern. Process Design Develop 8: 240
6. Bromley LA (1952) UCRL-1852: 1
7. Christiansen LJ, Kjrer 1 (1982) Enthalpy Tables of Ideal Gases, Haldor Tops0e A/S, Copen-
hagen pp. 1
8. U. S. Department of Commerce (1923) Bur Std [U.S.] Circ No. 142: 1
9. Dean DE, Stiel LI (1965) Am Inst Chern Eng 1 11: 526
10. Dembovskii VV (1968) Zavodsk Lab 34: 42; Ind Lab (USSR) 34: 52
11. Gillespie LJ (1925) 1 Math Phys 4: 84
12. Gillespie LJ, Beattie lA (1930) Phys Rev 36: 743
13. GillesIJie LJ, Beattie lA (1930) Phys Rev 36: 1008
15. Golubev IF, Kiyashova VP (1979) Tr GIAP No. 52: 57
16. Haar L (1968) 1 Res Natl Bur Std A 72: 207
17. Haar L, Gallagher lS (1978) 1 Phys Chern Ref Data 7: 635
18. Haber F, Le Rossignol R (1907) Ber Bunsenges Physik Chern 40: 2144
19. Haber F, Tamaru S (1915) Z Electrochem 21: 191
20. Haber F, Tamaru S, Oeholm LW (1915) Z Electrochem 21: 206
21. Haber F, Tamaru S, Ponnaz C (1915) Z Electrochem 21: 89
22. Heise F (1972) Ber Bunsenges Physik Chern 76: 938
23. Hirschfelder 10, Curtis CF, Bird RB (1954) Molecular Theory of Gases and Liquids, Wiley, New
York, pp. 1
24. Dow Chemical Corp (1971) lANAF Thermochemical Tables. 2nd Ed. NSRDS NBS-37: 1
25. Kaminishi G-I (1965) Intern Chern Eng 5: 749
26. Kazarnovskii YaS, (1945) Zh Fiz Khim 19: 392; CA. (1946) 1727
27. Kazarnovskii YaS, Karapet'yants MK (1941) Zh Fiz Khim 15: 966; CA. (1942) 6884
28. Konoki K, Takeuchi K, Kaminishi G-I, Toriumi T (1972) 1 Chern Eng 1apan 5: 103
29. Landolt-Bornstein (1980) 6th Ed. Pt. 4C 2: 189
30. Larson AT (1924) 1 Am Chern Soc 46: 367
31. Larson AT, Black CA (1925) 1 Am Chern Soc 47: 1015
32. Larson AT, Dodge RL (1923) 1 Am Chern Soc 45: 2918
33. Lefrancois B, Vaniscotte C (1960) Chaleur Ind No. 419: 183
34. Lefrancois B, Vaniscotte C (1960) Genie Chim 83: 139
35. Lindsay AL, Bromley LA (1950) Ind Eng Chern 42: 1508
36. Martin 11, Hou Y-C (1955) Am Inst Chern Eng 1. 1: 142
37. Mason EA, Monchick L (1962) 1 Chern Phys 36: 2746
38. Michels A, Dumoulin E, Th. Van Dijk 11 (1959) Physica 25: 840
39. Michels A, Dumoulin E, Th. Van Dijk 11 (1961) Physica 27: 886
40. Michels A, Skelton GF, Dumoulin E (1950) Physica 16: 831
41. Michels A, Wassenaar T, Wolkers Gl, de Graaf W, Louwerse P (1953) Appl Sci Res A 3: 1
42. Michels A, Wassenaar T, Wolkers G1, van Seventer W, Venteville Al (1954) Appl Sci Res A
4: 180
43. Neufeld PD, lanzen AR, Aziz RA (1972) 1 Chern Phys 57: 1100
44. Newton RH (1935) Ind Eng Chern 27: 302
45. Nielsen A (1968) An Investigation on Promoted Iron Catalysts for the Synthesis of Ammonia.
3rd Ed., Gjellerup, Copenhagen, pp. 1
46. Pal AK, Barua AK (1967) 1 Chern Phys 47: 216
47. Reamer HH, Sage BH (1959) 1 Chern Eng Data 4: 152
48. Reamer HH, Sage BH (1959) 1 Chern Eng Data 4: 303
49. Reddy KV, Husain A (1980) Ind Eng Chern Process Design Develop 19: 580
50. Reid RC, Prausnitz, 1M, Sherwood TK (1977) The Properties of Gases and Liquids 3rd Ed.
McGraw-Hili, New York pp. 1
51. Rossini FD (1953) Selected Values of Physical and Thermodynamic Properties of Hydrocarbons
and Related Compounds. Carnegie Press, Washington pp. 1
52. Sage BH, Olds RH, Lacey WN (1948) Ind Eng Chern 40: 1453
53. Schulz G, Schaefer H (1966) Ber Bunsenges Physik Chern 70: 21
54. Stiel LI, Thodos G (1964) Am Inst Chern Eng 1 10: 26
55. Tsederberg NV (1965) Thermal Conductivity of Gases and Liquids, MIT Press, Cambridge,
Mass., pp.l
56. Wiebe R, Gaddy VL (1938) 1 Am Chern Soc 60: 2300
1 Thermodynamic Properties in Ammonia Synthesis 15
Per Stoltze
Haldor Tops0e Research Laboratories Lyngby, Denmark
Contents
2.1 Introduction................................ 21
2.6.4 Physisorption of N2 . . . . . . . . . . 48
2.6.4.1 Structure of Physisorbed N2 49
2.6.4.2 Thermodynamics.. 49
2.6.4.3 Adsorption Kinetics . . . . . 49
2.6.4.4 Desorption Kinetics . . . . . 49
2.6.4.5 Kinetics of Conversion into aN 2 • 49
2.6.4.6 Properties of Physisorbed N z .......... . 49
2.6.5 Molecular Chemisorption of N2 . . .. . . . . . . . . . . . 50
2.6.5.1 Structure of Chemisorbed N 2 50
2.6.5.2 Thermodynamics.. 50
2.6.5.3 Adsorption Kinetics . . . . . . 50
2.6.5.4 Desorption Kinetics . . . . . . 50
2.6.5.5 Properties of Chemisorbed N z 50
2.6.5.6 Effect of Promoters . . . . . . 51
2.6.5.7 Effect of Preadsorbed Species 51
2.6.6 Dissociative Chemisorption of N .... 51
2.6.6.1 Structure of Chemisorbed N 52
2.6.6.2 Thermodynamics .. 52
2.6.6.3 Adsorption Kinetics . . . . . 53
2.6.6.4 Desorption Kinetics . . . . . 54
2.6.6.5 Hydrogenation of Chemisorbed N . 54
2.6.6.6 Properties of Chemisorbed N 54
2.6.6.7 Effect of Promoters . . . . . . 55
2.6.6.8 Effect of Preadsorbed Species 56
2.6.6.9 Isotopic Exchange . . . . . . 56
2.6.7 Kinetic Models of N2 Chemisorption . 57
2.6.8 Chemisorption of NH3 . . . . . . . . . 59
2.6.8.1 Structure of Chemisorbed NH3 59
2.6.8.2 Thermodynamics.. 59
2.6.8.3 Adsorption Kinetics 60
2.6.8.4 Desorption Kinetics 60
2.6.8.5 Dissociation..... 60
2.6.8.6 Properties of Chemisorbed NH3 60
2.6.8.7 Effect of Preadsorbed Species . . . . . . . . . . . . 61
2.6.9 Adsorption of N2H4 . . . . . . . . . . . . . . . . . . . . . . . 61
2.6.10 Chemisorption of O 2 . . . . . . . . . . . . . . . . . . . . . . . 61
2.6.10.1 Structure of Chemisorbed 0 62
2.6.10.2 Thermodynamics . . 62
2.6.10.3 Adsorption Kinetics . . . . . 63
2.6.10.4 Desorption Kinetics .... . 63
2.6.10.5 Properties of Chemisorbed 0 63
2.6.10.6 Oxygen Isotopic Exchange . 64
2.6.10.7 Effect of Promoters . . . . . . 64
2.6.10.8 Effect of Preadsorbed Species. 64
20 P. Stoltze
2.6.11 Adsorption of H 2 0. 65
2.6.12 Adsorption of H 2S 65
2.8 References . . . . 88
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 21
2.1 Introduction
This chapter deals with the structure and surface chemistry of industrial ammo-
nia synthesis catalysts. Results on catalyst models and single crystal surfaces are
included to the extent that they illuminate the behavior of industrial catalysts.
The industrial catalyst is prepared by fusion. The catalyst may be supplied in
the unreduced state after crushing and screening to the desired particle size or
the catalyst may be reduced and subsequently stabilized by controlled oxidation
in the catalyst factory. Although the reduced catalyst is pyrophoric, the
prereduced catalyst can be safely handled. In the ammonia synthesis plant the
catalyst is activated by reduction with a mixture of hydrogen and nitrogen as the
final step in the start-up procedure for the plant. The reduction of the
prereduced catalyst is faster and simpler than the start-up of the unreduced
catalyst.
A number of useful reviews on the structure and properties of ammonia
synthesis catalysts [1-18J and on ultra-high vacuum investigations related to
ammonia-synthesis [10, 19-25J have been published.
Catalyst constituents, which have little or no catalytic activity by themselves,
but which increase the catalytic activity for the catalyst are referred to as
promoters. Promoters which increase the catalytic activity primarily by increa-
sing the active area of the sample are referred to as structural or textural
promoters. Promoters which increase the activity of the catalyst primarily by
increasing the reaction rate per area are referred to as chemical or electronic
promoters. Constituents, which decrease the activity of the catalyst when pre-
sent in small amounts are referred to as poisons.
Catalysts containing Fe, one structural promoter and no electronic pro-
moter are commonly referred to as singly promoted. Catalysts containing Fe, one
structural promoter and one electronic promoter are referred to as double
promoted, while catalysts containing Fe, more than one structural promoter and
one or more electronic promoters are referred to as multiply promoted.
In the following, the notation e.g. (Fe, AI, K) will indicate a sample contain-
ing the elements Fe, AI, K and possibly non-metallic elements. The sequence
indicates the relative concentrations of the metals, the first metal being the most
abundant. The asterisk (*) represents a surface site; X * represents a species
X adsorbed .on a surface site.
The unreduced catalyst consists of oxides of iron with up to a few percent of AI,
Ca and K. Other elements may be present in small amounts.
22 P. Stoltze
2.2.1 Structure
2.2.1.1 Magnetite
For an industrial catalyst, energy dispersive X-ray analysis, X-ray powder
diffraction, optical microscopy and Mossbauer spectroscopy show that part of
the Al and Ca atoms are dissolved in the magnetite lattice [15,27,28,38-43].
The lattice constant of the magnetite phase of an industrial catalyst is 8.377 kX
[15]. The lines in the X-ray powder diffractions diagram are broadened [15]; the
broadening is independent of particle size [15].
From the Mossbauer spectrum [27,44,45] and the X-ray power diffraction
diagram [39] of the unreduced catalyst, it has been estimated that 85% of the Al
in the unreduced catalyst is dissolved in the magnetite. Evidence for the dissolu-
tion of K[41,46], Mg[41,46], V[41], Si[41], W[46], and Mo[46] in the
magnetite has been reported. However, due to the large size of K +, only a small
amount of K is found in the magnetite phase of the industrial catalyst.
Additional information on the structure of the magnetite phase comes from
the study of catalyst models, in particular of (Fe, AI) solid solutions. Mossbauer
spectroscopic studies of Fe304 [47] and of unstochiometric Fe-spinels [44]
have been reported.
For unreduced precipitated (Fe,AI) samples, solid solutions of Al in Fe-
oxides and of Fe in AI-oxides may be observed [48, 49] depending on composi-
tion [48] and the preparation method [49]. In experimental (Fe,AI)-oxide
samples, dissolution of Al in the magnetite has been shown by X-ray powder
diffraction [50]. The lattice constant decreases from 8.413 A to 8.365 A for
Fe304-Ah03 up to 13 atom % Al 2 0 3 and is then constant [51] indicating the
formation of a saturated solid solution with segregation of excess Ah03 as
a separate phase.
aries are small amounts of wustite [14, 15,28,40], calcium ferrites with dis-
solved promoters and a glass phase rich in silicon [26,40,41]).
2.2.1.3 Wustite
Based on X-ray powder diffraction studies, some authors have concluded that
the structure of the wustite is that of natural wustite [15, 45]. From energy
dispersive X-ray analysis and X-ray powder diffraction studies, others have
found evidence for the dissolution of AI[28,40], Ca[28], Mg[41], K[40],
V[41], or Si[41] in the wustite. In fused (Fe,AI)-oxide samples the grain
boundaries contain small but significant amounts of wustite [40].
2.2.1.4 Ferrites
Two calcium ferrites differing in their content of dissolved promoters may
coexist in the grain boundaries [28,41]. The dissolution of AI[28], K[28],
Mg[41], V[41], or Si[41] in the calcium ferrites has been reported.
While the composition of the solid solutions which constitute the unreduced
catalyst is not known in detail, an extensive knowledge exists on the location of
various additives both for catalyst models and for an industrial catalyst.
2.2.2.1 Aluminum
Evidence for the presence of Al dissolved in the magnetite phase has been found
by X-ray powder diffraction [27,28, 38-42, 53] and by chemical analysis of
powders of varying particle size [54]. The solubility of Al in the (Fe,Alh
04-phase has been determined to be 30 atom% AI[55] from measurement ofthe
Curie temperature and 50 atom % AI[55], or 67 atom % AI[39] from measure-
ment of the X-ray powder diffraction lattice constant. Other studies have
indicated homogeneous solution of Al in magnetite, at least for small amounts of
Al [56] and not too high temperatures [57].
In (Fe, AI) oxide catalyst models, some or all of the Al is dissolved in the
Fe-oxide phases, as solid solutions between Fe-oxide and AI-oxide are readily
24 P. Stoltze
2.2.2.2 Calcium
For an industrial catalyst, energy dispersive X-ray analysis, X-ray powder
diffraction and optical microscopy indicate that Ca is found dissolved in the
magnetite [28], in calcium ferrites [28,41], possibly in the glass phase [40], and
in the wustite [28]. Ca has been found in grain boundaries in a sintered (Fe,Ca)
oxide catalyst [61].
2.2.2.3 Potassium
In the unreduced catalyst [11, 14,40,61,62] and in (Fe,Mg) catalyst models
[63], K is found in the grain boundaries. Additional amounts of K may be
present as K-ferrites [64]. K has been reported to be associated with Si [54].
X-ray powder diffraction indicates that K is insoluble in magnetite [56].
The addition of Al makes the distribution ofK more homogeneous [62], and
. the addition of Al or Zr decreases the volatility of K during preparation [65].
Both observations indicate the potential for K to react with acid oxides in the
catalyst. CaO and Si0 2 decrease the water solubility of K [66]. Annealing
increases the water solubility of K [66].
The amount of K which may be extracted by H 20 has been reported to
increase [67,68] or to decrease [66] by the addition of Si; the amount of
K which may be extracted by H 20 is decreased [66] by the addition ofCa, and
decreases [67,68] or increases [66] by heating the unreduced catalyst. By
scanning electron microscopy and energy dispersive X-ray analysis it was found
that K segregates to the outer part of the catalyst particles with storage and
prolonged use [69].
tated (Fe,Co,AI) the Fe (II) : Fe (III) ratio increases [73J with increasing Co
content.
In the unreduced catalyst, Mg is present dissolved in magnetite [63J and as
an unidentified phase in grain boundaries [63].
The effect of Mn on the precursor has been studied [74J,
In Mo containing samples, Mo has been detected as grains ofK 2 Mo0 4 [75J,
CaMo0 4 [75J, FeMo0 4 [75J Fe3(Mo04h [76J, or Mo0 3 [76J depending on
the composition and preparation procedure. Mo is soluble in magnetite [77].
For precipitated (Fe,Ni,AI), Ni increases the Fe (II) : Fe(III) ratio [73].
In the unreduced catalyst, Si is found in the calcium ferrites [41J and in the
glass phase [41, 26]. Smaller amounts may be present dissolved in the magnetite
[41J or in the wustite [41]. Evidence for the association of Si with K has been
reported [54]. Si is insoluble in magnetite in the absence of other promoters
[56]. The addition of Si decreases the solubility of basic oxides in magnetite
[56].
For W containing catalysts, W is found dissolved in the magnetite [46].
2.2.3 Texture
The size of the magnetite grains in the unreduced catalyst is rather variable [78].
The cross sectional area of magnetite in KM1 is (1.28 ± O.l4)·1O- 2 mm 2 deter-
mined from planimetry [79]. The density is 4.8 gjcm 3 [14]. The porosity is
negligible [14].
For the prereduced catalyst the density is 3.73 gjcm 3 [14]. The porosity is
0.11 cm 3 jg, i.e., 41% [14J, and electron microscopy [11, 14,80J shows the
presence of a well developed pore system.
Precipitated (Fe,AI) catalyst models in the unreduced state contain pores
with a 19-20 A radius and approximately 70 A radius [48J, while in the
prereduced state these models have maxima in pore volume distribution at 20 A
and at 140 A [48]. The 20 A peak is thus unaffected by the reduction process.
The BET area is smaller for a catalyst prepared by sintering than for
a similar catalyst prepared by precipitation [81]. The BET area decreases with
increasing calcination temperature [58J, and increases with increasing Al con-
centration [61J for precipitated catalysts. For samples calcined at 600 DC, the
BET-area increases from 13 m 2 jg at 0% Al 2 0 3 to 215 m 2 jg at 88% Al 2 0 3 [48].
From a X-ray photoelectron spectroscopy study, the surface composition
3.2% Fe, 33.2% K, 8.4% AI, 3.9% Ca and 51.3% 0 (atomic %) was found [82J
for the unreduced catalyst. The surface is enriched in K and Al compared to the
bulk [52]. The prereduced catalyst shows more Fe in the surface by X-ray
photoelectron spectroscopy than the unreduced catalyst [52]. Energy dispersive
X-ray analysis of prereduced catalyst shows that the surface consists mainly of
iron oxide [83J; AI, Ca and Si are inhomogeneously distributed in the surface
[83].
26 P. Stoltze
The reduction of the industrial catalyst has been extensively studied, [89-92].
The reduced catalyst mainly consists of metallic Fe while the promoters
remain in their oxidic state. The reduction process serves two purposes, firstly
the surface of metallic iron is the active structure, and secondly the removal of
the oxygen makes the material porous and increases the surface area by a large
amount.
2.3.4.1 Aluminum
Al decreases the rate of reduction for (Fe,AI)-oxide samples [130], in particular
in wet atmospheres [135, 136]. For Al 20 3 supported catalyst models, the partial
dissolution of Al during impregnation leads to a more difficult reduction [137].
2.3.4.2 Alkali
M6ssbauer spectroscopic studies showed that alkali promotes the reduction of
Fe203 to Fe at 300°C [111, 138, 139]. Yet more direct measurements by
temperature programmed reactions demonstrate that K decreases the rate of
reduction for (Fe,K)-oxide samples [121] and for (Fe,AI,K)-oxide samples
[121]. The rate of reduction increases through the sequence (Fe,AI,M), M = Li,
Na, K, Rb, Cs for Fe203 based catalysts [139].
30 P. Stoltze
Some information on the structure and texture of the active catalyst may be
inferred from studies of the spent catalyst [153], although the details of com-
position and structure may differ due to the violence of the reaction of the
reduced catalyst with air and due to structural changes during the oxidation.
2.4.1 Iron
In the catalyst [14, 15,27,41, 112, 154] Fe is present in the reduced state mainly
as the metal. The lattice constant of the iron is 2.8601 kX [15]. On the lean side
of the gas phase equilibrium of the synthesis gas mixture Fe films do not form
bulk nitrides [155]. On the rich side of the equilibrium Fe4N may be formed
[155].
In an industrial catalyst traces of unreduced Fe are detected by Mossbauer
spectroscopy [27,42]. These traces of Fe may be present in Ca ferrites with
dissolved promoters [28,41,156] or in the glass phase [41,156]. It has been
suggested that the glass phase is inactive in the formation of the active catalyst
[157].
For unpromoted Fe oxides, both Fe304 and FeO are completely reduced
[40] at 550°C in H 2 •
For a number of (Fe,AI,K) and (Fe,AI,Cs) catalysts, in situ EXAFS and
XANES [158, 159] indicate complete reduction of Fe in the reduced state of the
catalyst. For an (Fe,AI)-oxide catalyst model containing 3% ASI 2 0 3, Mos-
sbauer spectroscopy indicates complete reduction of Fe [112] while for an
(Fe,AI)-oxide catalyst model containing 10.2% A1 2 0 3, traces of Fe (II) have been
detected in the reduced state [154].
The degree of reduction for 0.05~ 15% Fe on Al 2 0 3 is 77~97% after reduc-
tion in 3H 2 + N2 at 1 atm, 673 K [160]. This indicates that Fe supported on
Al 2 0 3 is more difficult to reduce that Fe promoted with Al 2 0 3 and that (Fe,AI)
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 31
oxide samples of high Ah03 concentration may be poor models for the indus-
trial catalyst.
The X-ray powder diffraction diagrams show line broadening for Fe due to
particle size effects [14,15,27, 154]. The particl~ size for Fe is 300 A [15].
Others have interpreted the in situ X-ray powder diffraction diagram as evi-
dence for Fe being present as a metallic glass [161].
2.4.2 Aluminum
2.4.3 Calcium
In the reduced catalyst CaO has been shown by Auger electron spectroscopy
[162] and X-ray photoelectron spectroscopy [52] to remain in its oxidic state.
Chemisorption measurements [163] and X-ray powder diffraction studies [164]
show that during the reduction CaO segregate to the space between the Fe
crystallites.
The outer shape of the grains is conserved during reduction [28, 110,
113,156,168,169], expands by up to 0.6% [168] or contracts by up to 0.5%
[168] during reduction depending on the composition and reduction temper-
ature.
32 P. Stoltze
The density of the reduced catalyst is 2.7-3.7 gjcm 3 [14]. The pore volume is
0.15 cm 3 jg [67, 6SJ, independent of the K concentration [[67,6S].
For the reduced catalyst, electron microscopy shows the presence of a well
developed pore system [92, 170, 171]. An industrial catalyst with pores of 100
and 300 A radius [14, 115J in the reduced state has been found to have 177 A
pores [l72J after passivation. The reduction of promoted magnetite creates
a system of pores with < 400 A diameter [SO]. The average pore radius
increases with increasing K concentration from 50-200 A (0% K) to 200-S00 A
(0.7S% K) after reducing at 600°C [67, 6S].
Precipitated (Fe,AI) catalyst models have pores with 19-20 A radius and
larger pores of 50-250 A radius [4S, 167].
The particle size has been studied by X-ray powder diffraction [173]. The cause
of the line broadening in the X-ray powder diffraction diagram has been
assigned to particle size effects [l64J, to the combined effect of particle size,
defects and strain [174J, or to the presence of paracrystalline defects
[166, 173, 175].
If the line broadening is assigned to particle size effects alone, the calculated
particle radius is 100-1000 A [53J, 175-250 A [174J, or ISO A [164]. For
catalyst models the particle radius is 305 ± 15 A for (Fe, AI) [154J 125-155 A for
(Fe,Mg) and 305 A for (Fe,Mg,K) [176].
If the line broadening in the X-ray powder diffraction diagram of an
industrial catalyst in the reduced state is assigned to the presence of paracrystal-
line defects, the calculated particle radius is 200 A [173, 175]. The calculated size
of the paracrystalline defects is the same in both the unreduced and in the
reduced states [173].
Physisorption measurements using N z , CO, Ar, O 2 or COz give the same area
[177J; the areas determined by this method are 0.44-10.4 m 2 jg [17S]. The BET
area for the reduced catalyst depends on the composition and structure prior to
reduction and on the conditions during the reduction. Consequently very
different values have been reported: 8 m 2 jg [l64J, 11.6 m 2 jg [179J, 15.S m 2 jg
[179J, 15 mZjg [93J, or 20.9 m 2jg [ISO]. After passivation an area of 13.1 m 2jg
found [172].
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 33
The cause of the structural promotion has been assigned to the segregation of
refractory oxides to the surface [53, 154, 195, 196] or to the formation of
paracrystalline defects [154, 166]. Theoretical considerations show that the
structural promoters must be located near or on the surface to have any effect
[197].
2.5.4.1 Aluminium
Al is a structural promoter [67,68, 198-204]. The distribution of Al in the
sample changes during reduction [40] and, in particular, the segregation of Al to
the surface has been the subject of a large number of studies.
From studies by chemisorption measurements [163, 198, 199,205-207], by
X-ray powder diffraction studies [164], by electron microscopy [92, 170, 171],
34 P. Stoltze
2.5.4.2 Calcium
Ca is a structural promoter [198,199,201]. The effect ofCa on the properties of
the catalyst has been much less studied than the effect of AI.
Ca has been shown by Auger electron spectroscopy [162] and X-ray photo-
electron spectroscopy [52] to remain in the oxidized state. Chemisorption
measurements [163] and X-ray powder diffraction studies [164] show that CaO
is segregated to the space between the Fe crystallites during the reduction. The
segregation of Ca to the surface has been demonstrated by scanning Auger
electron spectroscopy[52, 209-211].
Ca has been reported to increase the activity [218], to increase [198, 199] the
surface area, and to increase the resistance toward impurities in the gas [67,68].
The optimum Ca concentration in a (Fe,AI,K,Ca)-oxide catalyst is 2-2.5 %
[201].
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 35
2.5.4.3 Potassium
K is an electronic promoter. K acts as an promoter both by impregnation and
by addition to the melt [219].
The segregation ofK to the surface has been demonstrated by chemisorption
measurements [207, 220J, by scanning Auger electron spectroscopy
[52, 209-211J, by X-ray photoelectron spectroscopy [208J, and by electron
microscopy [92]. Single crystal studies of K overlayers on Fe(110) demonstrate
that K is not a structural promoter [221J and that K may even reduce the ability
of Al to disperse Fe [221].
The migration of K to the surface of the reduced catalyst [40J has been
demonstrated by energy dispersive X-ray analysis [28J, by field iron mass
spectroscopy [222, 223J, by chemisorption of CO, CO 2, N 2and H2 [207, 220J
by scanning Auger electron spectroscopy [209-211J, and by high-voltage elec-
tron microscopy [224].
For the catalyst it has been concluded that KH, KNH2 and K 20 are less
stable than KOH under NH3 synthesis conditions [225,226]. These consider-
ations were based on the bulk phase thermodynamics.
For K adsorbed on an Al 20 3 overlayer on Fe(100), the desorption tempera-
ture is increased to 600°C [227J, well above the typical reaction temperature
[227].
On clean single crystal surfaces X-ray photoelectron spectroscopy and
ultraviolet photoelectron spectroscopy show that K is present as the metal
[228]. On Fe(100) [229, 230J chemisorbed K is disordered. On Fe(llO) a hexag-
onal close packed structure is formed [228]. On Fe(1l1) K is disordered [230J
or forms a (3 x 3) structure [231]. At room temperature it is unlikely that
multilayer adsorption will occur [228J due to the low heat of vaporization for K.
K chemisorbed on Fe(100) desorbs under NH3 synthesis at 20 atm. [232].
This is consistent with typical temperatures for NH3 synthesis being above the
desorption temperature for K. K*jFe(lOO) is stabilized by the presence of 0*
[232].
The presence of a K or a K + Al overlayer does not cause a recrystallization
or an increase in catalytic activity after steaming and reduction in the NH3
synthesis gas [233].
The migration of K on FejAl 20 3 has been studied by Auger electron
spectroscopy [234]. The migration is faster in H2 than in O 2, and faster in moist
gas than in dry. The kinetics of surface migration was found to be consistent
with a surface diffusion mechanism [234].
Addition of small amount of K results in an increase in average pore
diameter [67, 68, 235J, no changes in the pore volume [67, 68J, an decrease in
the BET area [67,68,176, 236-241J a decrease of the active area [67, 68J, an
increase in average particle diameter [176J, a decrease [176,217,237, 242-244J
or an increase [243J in the work function. For single crystal surfaces at small
K-coverages, the work function decreases with increasing K coverage
[228,229,245].
36 P. Stoltze
A promising technique for the study of the catalytic activity of alloys is the
study of chemisorption on iron overlayers on single crystals of other metals such
as FejRu [375], FejRe [376], and FejW [377].
2.6.1 Chemisorption of H
2.6.1.2 Thermodynamics
The initial enthalpy of chemisorption has been determined for single crystal
surfaces of Fe. For H/Fe(1 00) - 86 kJ/mole was found by temperature pro-
grammed desorption [383]; for H/Fe(11 0) - 109 kJ/mole [378] by tempera-
ture programmed desorption or - 24.2 kcaljmole [384] by He-scattering; for
HjFe(111) - 88 kJ/mole by temperature programmed desorption [378]. For
D/Fe(110) - 24.2 kcaljmole has been determined from He-scattering [384]
and for D/Fe(111) - 104 kJ/mole by calorimetry [385].
Adsorbate-adsorbate repulsion at high coverage is reflected in the low-
energy electron diffractions patterns, in the occurrence of temperature pro-
grammed desorption peaks, and in the enthalpy of chemisorption.
For hydrogen adsorption on polycrystalline Fe, the enthalpy of chemisorp-
tion has been determined as - 98 kJ/mole by calorimetry [385], - 96 kJ/mole
by calorimetry [386], - 81.0 kJ/mole by electrical conductivity [387],
- 85.0 kJ/mole by volumetric chemisorption [388], - 36 kcaljmole at zero
coverage by volumetric chemisorption [389], - 15 kcaljmole at 90% saturation
[389], - 20 kcaljmole at 140 K, zero coverage [390], - 17.5 kcaljmole at 1%
of saturation calculated from the equilibrium pressure [391], and
- 5 kcaljmole at 10% of saturation calculated from the equilibrium pressure
[391].
For Fe/MgO the saturation coverage depends on particle diameter [328],
i.e., the chemisorption appears to be structure sensitive. Structural sensitivity of
NH3 synthesis will be discussed further in Sect. 7.1.1.
The studies of the chemisorption of H2 on the catalyst has been comp-
lemented by studies at low [391, 392] and at high pressures [393].
40 P. Stoltze
AI, Cal samples [404J; and 120-170,280-380 and 480-540°C for (Fe, AI, Ca, K)
samples [404].
Temperature programmed desorption of H2 from the reduced catalyst has
been reported to proceed at 90-240°C [405J or to show 4 peaks [406]. The
kinetics of thermal desorption of H2 from the catalyst is second order [407].
Evidence that chemisorbed hydrogen remains on the surface after evacuation at
550°C for 24 h has been reported [408].
2.6.1.9 H2 + D2 Exchange
The H2 + D2 isotopic exchange is interesting as one of the simplest chemical
reactions of H2 on the catalyst surface.
The activity of H2 + D z exchange increases as Fe(pc) > Fe(ll 0) >
Fe(l 00) > Fe(lll) [426].
Fast Hz + D2 isotopic scrambling has been found for an (Fe, AI, Si, Zr)
sample at - 195°C [424], while the scrambling was slow on an (Fe, AI, K)
sample at the same conditions [424].
2.6.2 Chemisorption of CO
molecular species at low temperature has been found from X-ray photoelectron
spectroscopy and ultra violet photoelectron spectroscopy [245, 427].
On single crystals a c(2 x 2) structure of CO/Fe (1 00) is formed at 373-400 K
[427,383]. For CO/Fe(11 0) at 300 K, c(2 x 4) is formed at low coverage [428,
429], p(l x 2) is formed at high coverages [428]. By NEXAFS it has been found
that for CO/Fe(100) the CO molecule is tilted 45 ± 10° [430].
The disorder observed below room temperature [383, 427] [429] is caused
by the lack of mobility [427]. Just below the dissociation temperature ordering
is observed.
While molecular and dissociated CO have similar X-ray photoelectron
spectroscopy spectra, O(ls) at 531 eV and C(ls) at 285 eV [245,427,431], the
ultra violet photoelectron spectroscopy [245, 427] and laser Raman [432]
spectra are different.
During sequential adsorption of 12CO and 13CO on Fe (1 00), isotopic
exchange is observed only among the two strongest bound molecular states
[433]. For the catalyst, isotopic scrambling between 13C160 and 12C180 is
observed at - 33°C indicating the existence of dissociated CO at these tempera-
tures. There is partial isotropic scrambling between the first and second exposure
at -195°C [434] and at -78°C [435].
2.6.2.2 Thermodynamics
For single crystal surfaces the initial enthalpy of chemisorption for CO
is -105 kJ/mole on Fe(100) by temperature programmed desorption
[383], - 96 kJ/mole on Fe(ll 0) by temperature programmed desorption [245,
436], - 91 kJ/mole by temperature programmed desorption on Fe(lll) [437],
and -155 kJ/mole by calorimetry on Fe(pc) [438].
The enthalpy of chemisorption for CO on Fe surfaces is - 9.3 to
- 4.2 kcaljmole at O°C [439] on (Fe, AI), - 23.1 to - 16.9 kcaljmole at O°C
[439] on (Fe, AI, K), - 8.0 to - 4.2 kJ/mole at - 183°C [439] on (Fe, AI, K), and
- 32 kcaljmole at 22°C [389] on (Fe, AI, K). For Fe supported on Alz0 3
a Freundlich isotherm has been found [160]. The enthalpy of adsorption is
12-30 kJ/mole at 37% of saturation [160].
For Fe/MgO the amount of CO chemisorption depends on particle diameter
[328]. The ratio of the area determined from CO chemisorption to the area
determined by the BET method decreases through the sequence Mg > Be > AI,
Si, Cr > Mn > Ca > Ti [198, 199]. Infra-red spectroscopy and microcalorimetry
show that the amount of weakly bound CO increases with the dispersion of
Fe/MgO [328].
For multiply promoted samples the CO area is 5.6 m 2/g [180].
For the catalyst, pulse chemisorption and volumetric chemisorption give
identical results at - 78 °C and - 196°C [440]. For an industrial catalyst the CO
chemisorption was initially found to be 0.7-1.4 m 2/g [14,164], i.e., 9-18 flmole/g;
in later studies the value 28-39 flmole/g was found [179]. Increa-sing the wustite
content in the unreduced sample was reported to increase the CO area [181] or to
decrease the CO area [183] of the reduced catalyst.
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 45
2.6.3 Chemisorption of CO 2
2.6.3.2 Thermodynamics
For Fe(11 0) no adsorption of CO 2 is detectable in the range 77-340 K [450].
For a (Fe, AI) catalyst model the enthalpy of chemisorption is ~ 8.7 to
- 6.5 kcaljmole at - 78°C [439] and - 17.7 to - 8.9 kcaljmole at - 78°C
[439] for (Fe, AI, K).
Pulse chemisorption and volumetric chemisorption give identical results for
the CO 2 chemisorption [440]. The enthalpy is difficult to determine because of
the tendency to dissociate at higher temperatures.
2.6.4 Physisorption of N2
2.6.4.2 Thermodynamics
The saturation coverage is 5.8 ± 0.7 {lmole/m 2 at 91 K for Fe(111) [462,463].
The heat of physisorption is 5.2-3.1 kcal/mole [390] for Fe powder at
78-90 K.
For single crystal surfaces the heat of adsorption is 25-37 kllmole depending
on the coverages by molecular and physisorbed N2 * [22, 462].
Even if CO is isoelectronic with N z, the bonding to the surface is different for the
two molecules. CO adsorbs with a high sticking probability and high binding
energy in end-on geometry, while N z adsorbs side-on in a more weakly adsor-
bed species. The dissociation of N z * is much easier than the dissociation of
CO*.
2.6.5.2 Thermodynamics
The saturation coverage for N z * on Fe(lll) is 1.16 pmole/m z [462,463]. This
is significantly less than for the physisorbed state.
The enthalpy of chemisorption for N z * is - 31 kJ/mole for N z /Fe(1 00)
[466],[229], -31 kJ/moleforN z/Fe(110)[467],and -21 kJ/moleforFe(pc)
[469] from temperature programmed desorption studies.
Laser Raman during NH3 synthesis shows peaks at 2040, 1940, 423,
443 cm- 1 [473]. These peaks have been assigned to N z * [473].
The adsorption of N z * increases the work function on Fe [22, 474]. The
dipole moment of N z * is 0.4 D [467].
2.6.6.2 Thermodynamics
From a study of N2 adsorption by volumetric chemisorption it was concluded
that the chemisorption of N2 on a (Fe, AI, K, Si) catalyst follows the Freundlich
isotherm [491, 492]. The enthalpy of chemisorption is - 38 kcaljmole 10g(fJ)
[492].
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 53
scrambling. These results give evidence for the dissociative mechanism of am-
monia synthesis [520].
Addition of Hz to N z has been reported to increase the rate of 14N + 15N
isotopic scrambling [288,481,488, 520, 522, 532, 533], to have no effect of the
rate [534] or to decrease the rate [288]. The observation of an increase in the
rate of isotopic exchange in the presence of H has been interpreted as H remov-
ing an oxide from the surface [481, 533].
The presence of K has been found to be important for the effect of H [288].
N 2 isotopic exchange over fused Fe is inhibited by Oz [490]. The activation
energy for nitrogen isotope exchange is higher in the presence of chemisorbed
oxygen [534].
The numerical models of the kinetics of NH3 synthesis will be discussed in Sect.
2.7.3 All contain a model of the kinetics of N z adsorption
(4)
(5)
as a special case. From the kinetic models the coverage of N * may be calculated
as a function of exposure. These calculations, shown in Figure 2.1, are in good
agreement with experiments at low and moderate exposures [396]. The cal-
culated peak temperatures, see Figure 2.2, as well as the peak shapes are in
reasonable agreement with experiments [396].
0.8
Q.o 0.6
2
~
u 0.4
Exposure (Po' s)
58 P. Stoltze
e
OJ
c
o
~
I-
o
Vl
OJ
o
This resulted in a somewhat too high binding energy for N * and led to the
conclusions that the enthaply measured at high coverages by N * might be more
appropriate [536]. Later the pre factors were adjusted [537] resulting is a better
value for the stability of N *. Using the data of Stoltze and N0rskov, Bowker,
Parker and Waugh [537] found a broad peak for the thermal desorption ofN 2 .
This discrepancy was resolved by Trivino and Dumesic, who concluded that
both the data-set of Stoltze and N0rskov and the revised data set of Bowker,
Parker and Waugh reproduce the experimental data [474].
2.6.8.2 Thermodynamics
The enthalpy of chemisorption is - 71 kJ/mole [541] for NH3/Fe(11 0) and
- 84 kJ/mole [397] for NH3/Fe(lll) from temperature programmed desorp-
tion.
60 P. Stoltze
2.6.8.5 Dissociation
For NH3/Fe(100) [542] and NH3/Fe(111) [397, 542] the dissociation be-
comes observable at 160 K and is complete at 320 K. For NH3/Fe(11 0) the
dissociation temperature is 260-290 K [540].
The dissociation of NH3 * on Fe(11 0) was reported to give N * and H * as
the only reaction products [540]. Others have found evidence for the intermedi-
ate formation of NH * [545].
By field ion mass spectroscopy of NH3 on Fe at room temperature,
4.10- 4 torr, NxHy species and FeNxHy species are detected [510]. Secondary
ion mass-spectroscopy of NH3/Fe(11 0) at 130 K shows FenNHW+ with
n, m = 1,2, NH n+ with n = 0,1,2,3,4, H+, Hi and Fe+ [545]. The spectrum is
interpreted as fragments of molecularily adsorbed NH 3* [545]. When the
temperature is increased, FeNHt and FeNH~ + decrease smoothly, while Fe +
increases smoothly, illustrating the decrease in surface coverage and the reaction
of NH3 (g) with the surface at higher temperatures [545]. The intensity of NHt
with NHi decreases steeply above 300 K consistent with the temperature
programmed desorption studies [545]. Both ions are probably formed from
adsorbed species [545]. The intensity ofNH+ decreases more slowly than NHi
and NHj; the reason is probably that NH+ is formed from both NH3(g) and
NH* [545].
2.6.10 Chemisorption of O 2
2.6.10.2 Thermodynamics
The amount of O 2 taken up under passivation in 0.75-1.0% O 2 in N2 is 3-4%
by weight [571]. The oxide film formed is approximately 30 A [122J or 6-12
atomic layers of oxide [572].
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 63
The enthalpy of chemisorption is 0-120 kcaljmole for (Fe, AI) samples [439J
at -183°e and 0-100 kcaljmole for (Fe, AI, K) samples [439J at the same
conditions.
The enthalpy of chemisorption for an industrial catalyst is 420 kJjmole up to
0.25 monolayer of O 2 , then the enthalpy drops to 34 kJjmole [573]. The
strength of adsorption and the adsorption capacity for O 2 on the catalyst
increases with temperature [574]. The optimum temperature range for passiva-
tion is 673-773 K [575].
By pulse chemisorption of O 2 on the catalyst, the O 2 chemisorption area is
1.7 m 2 jg at -78°e and 1.9 m 2 jg at 20 e for a sample with a BET area of
0
1Om 2 jg [576].
In the later stages of reduction the amount of oxygen adsorption is propor-
tional to the degree of reduction [122J.
photoelectron spectroscopy of Fe (2P3/2) shows that for Fe: 0 < 1: 1.5 the oxi-
dation state for Fe is + 3 [578] and not + 2 which has been suggested earlier.
At small coverages the work function increases with the coverage until
a maximum is reached at the c(2 x 2) structure [552]. At higher coverage the
work function decreases and goes through a minimum [552, 581].
The complicated behavior of the work function is not reflected in the bulk
electronic structure. The electrical resistance of a Fe film increases smoothly
with O 2 exposure [582].
The key step in the synthesis of NH3 form N z + 3 Hz is to dissociate the N-N
triple bond in the N z molecule. The direct gas phase reaction would involve
extremely endothermic and exothermic reactions. The resulting activation ener-
gies would be prohibitively high according to the principle of Sabatier.
All imaginable reaction mechanisms can be separated into two main cases.
In the associative mechanism H is added to the N z molecule before dissociation
of the N-N bond, e.g.
(6)
N z*+ 2H *::::::;;:NzH z * + 2* (7)
NzH z * + *::::::;;:2NH* (8)
66 P. Stoltze
Very endothermic or exothermic reaction steps are avoided if the N-N bond is
broken synchronously with the addition of H to the N-atoms.
In the dissociative mechanism the N-N bond is dissociated before any N-H
bond is formed, e.g.
(9)
N 2 * + *~2N* (10)
N* + H*~NH* (11)
Very endothermic or exothermic reaction steps are avoided if the N-N bond is
broken synchronously with the formation of the N * surface bond.
The number of distinct mechanisms is further varied by the number of
intermediate steps and the number of sites involved in the bonding of each
intermediate. A large number of mechanisms result in kinetics expressions which
are indistinguishable from an experimental point of view.
Catalysis may be understood at several levels. In recent years the under-
standing of ammonia synthesis has been taken to the level where the high
pressure reaction has been treated in terms of numerical models based on
a description of the reactants at the atomic level. In the present section we will
first address the questions on the nature of the catalytically active structure in
the catalysts and the information on the mechanism of ammonia synthesis
available from chemisorption studies. We will then describe some models of
ammonia synthesis in some detail and then proceed to discuss the remaining
aspects of the mechanism of ammonia synthesis based on these models.
Experimentally it has been verified that the synthesis of ammonia takes place on
the surface of the catalyst [592] rather than in the bulk. However, isotope
labeling experiments seems to indicate that small amounts of ammonia may be
formed in the bulk at very high temperatures [593]. X-ray photoelectron
spectroscopy demonstrates that Fe is present in the surface of the reduced
catalyst as Feo and not as Fe oxide [52, 208].
Under ammonia synthesis conditions, ultra-high vacuum data demonstrate
that a bulk nitride is not formed [594] whereas during NH3 decomposition the
kinetics indicate that the reaction may proceed on a completely nitrided surface
[595]. The increase in reaction rate observed by the addition ofH 2 is interpreted
as the reaction being faster on Fe metal than on Fe-nitride [595].
The coverage of the total surface by catalytically inactive structural pro-
moters has been determined by a number of techniques. The coverage by
catalytically inactive material is 60%, calculated from kinetic data for NH3
synthesis [596], 55% [191,207],60% [207], or 45% [195] from chemisorption
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 67
measurements, 45% from oxygen exchange with l80-labeled water [195], and
26% from DzjH z isotopic exchange [191]. The active fraction of the total area
generally decreases with increased promoter content [198, 199]. The active area
decreases through the sequence Mg > Be > AI, Si, Cr > Mn > Ca > Ti [198,
199] for a fixed promoter concentration. Reduction at a very high temperature
(1073 K) increases the NH3 synthesis and Nz-chemisorption rate; this has been
interpreted as reduction of a Fe(II)-spinel [123].
Calculated from the activation entropy, the density of active sites is 10 12 per
cm z for a porous Fe-catalyst and 2.5' 10 16 per cm z for a Fe-film [597].
For (Fe, AI) samples large N z chemisorption correlates with high catalyst
activity [524, 525]. In some cases no correlation between N z chemisorption and
catalytic activity has been found. This lack of correlation has been ascribed to
heterogeneity of the surface [598] or to structural sensitivity for FejMgO [328].
While the catalytic activity for the catalyst does not correlate with BET-area
[599], the catalytic activity has been found to be proportional with the Fe area
as determined by CO-chemisorption [53, 600]. Others have found no correla-
tions or a complex relationship [164, 601], probably because promotion is
important.
AI) catalyst models [608] and for Fe single crystals with evaporated overlayers
of Al and K [232].
The ultra-high vacuum studies show that during NH3 synthesis N * is formed
[414]. As the desorption temperature for N* is well above normal synthesis
temperature, the desorption of N * as N z (g) will be slow at synthesis conditions
[414]. If N * were not consumed by the reaction, it would soon inhibit the
synthesis due to blockage of active sites [414]. These observations are consistent
with Nz(g) + 2* = 2N* being the rate limiting step for NH3 synthesis [414].
Temperature programmed desorption of N z from the industrial catalyst
resembles temperature programmed desorption ofN z in the range 85 to 220 K
70 P. Stoltze
Data for the kinetics measured on single crystal surfaces are important for the
study of catalytic reactions as the surface of single crystals are approximations
to the more complicated surface of catalysts [642]. This is supported by the
agreement between rate measurements for Fe single crystals [603, 604J and the
rate measured for an industrial catalyst at 1 atm [643].
The purpose of developing models of the kinetics and mechanism of catalytic
reactions starting with a description of the reactants at the atomic level is to
understand the kinetic phenomena, rather than to give a very accurate descrip-
tion of a few phenomena. While a model of a single phenomenon may give an
accurate description of this phenomenon, this description may not be unique
and extrapolation may lead to ambiguities for situations where no experimental
data exist. However, if all aspects of interest are described using one model and
this model is in reasonable agreement with available data, this model may also
be used with some confidence in situations where experimental data are sparse
or ambiguous.
NH z * + H*::::;;;NH 3* + * (18)
NH3 *::::;;;NH 3 (g) +* (19)
Bowker, Parker, and Waugh proceed by using Arrhenius expressions with
known or estimated values for all prefactors and activations energies.
The model by Stoltze and N0rskov [396, 625, 644, 645J is based on the
reaction sequence
with the explicit assumption that the rate limiting step is the dissociation of N z *
(27)
Stoltze and N0rskov proceed by applying statistical mechanical methods to this
sequence, essentially expressing the thermodynamic properties of reactants and
intermediates in terms of spectroscopic properties. Further they treat within the
same model, the kinetics of adsorption of N z as well as the thermodynamics of
adsorption for Hz and NH 3.
Trivino and Dumesic have considered both of these reaction sequences and
compared the results and the differences in approach.
While the models of Bowker, Parker, and Waugh and by Trivino and
Dumesic do not make a priori assumptions on the nature of the rate limiting
step, Stoltze and N0rskov make the explicit assumption that the dissociation of
N z * is rate limiting. The assumption leads to a considerable simplification in the
further treatment of their model. While Bowker, Parker, and Waugh, and
Trivino and Dumesic must calculate reaction rates iteratively, the model by
Stoltze and N0rskov allows the derivation of explicit solutions for the coverages
and reaction rates. Further, a number of aspects of the kinetics of ammonia
synthesis, such as the activation enthalpy and the reaction orders may be
investigated analytically in the model by Stoltze and N0rskov.
The solution of the models involves a number of approximations.
The reaction sites are treated as identical [396, 625, 645J; the reactants,
intermediates and products chemisorb competitively on these sites. The com-
petition for the sites important kinetic consequences. The identity of the sites is
justified by the results from temperature programmed desorption from single
crystals and by the result from quantum mechanical calculations.
72 P. Stoltze
2.7.3.3 Test
By construction the model reproduces the thermodynamics of adsorption for
the intermediates and the kinetics of adsorption ad desorption for N *. As no
measured data for the catalytic activity have been used in the determination of
the input parameters, the model may be tested by a comparison between the
calculated and experimental rates of ammonia synthesis. The test is made by
calculating the exit ammonia concentration from the input composition and
operating conditions for the reactor.
Bowker, Parker, and Waugh in their first paper [535J did not appreciate that
the uniquely low sticking coefficient for N2 must also be reflected in a uniquely
low prefactor for the recombination ofN *. The somewhat too high value for the
binding energy of N * resulted in a reaction rate for NH3 synthesis too low by
a factor of 10 5 [535]. Bowker, Parker, and Waugh suggested the use of binding
energies appropriate for high coverages by N * available from gravimetric
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 73
Experimental output
isotherms [191]. However, it has been pointed out that these data are not
consistent with measurements for the sticking coefficient over single crystals.
Bowker, Parker, and Waugh later adjusted their data and obtained agreement
with experimental rates within an order of magnitude.
Using a number of data sets at 1, 150 and 300 atm, Stoltze and N0rskov
found that the calculated exit concentrations are in good agreement with the
experimental results [396, 625, 645J, see Fig. 2.3. The differences between
calculation and experiment corresponds to an error in the rate of less than
a factor of 1.5 [396, 645]. For (Fe, AI) the calculated rates are too high by
a factor of about 3. This is clearly a consequence of assuming that all sites have
the activity of Fe(lll).
Trivino and Dumesic concluded [330J that the remaining discrepancy be-
tween the results by Bowker, Parker, and Waugh and by Stoltze and N0rskov
area caused by an unusual small number of active sites assumed in the model by
Bowker, Parker, and Waugh.
Not all the input parameters are equally important for the success of the
model. Stoltze and N0rskov found that the critical parameters are the prefactor
and the activation energy for the sticking coefficient, the ground state energy for
N 2 * and the ground state energy for N * [396, 645]. These parameters are all
rather accurately known from experimental data. The reason why these are the
critical parameters is that the first three parameters determine the rate constant,
while the groundstate energy for N * determines the number of free sites [396,
645].
There is no gas reaction since no difference is seen between quenching the gas to
o°C or flow through a hot quartz tube [592J.
74 P. Stoltze
Nz(g) .3 Hz(g)
Nz*+3Hz(g)
0r 1 2NH3(g)
0 2N *.3H z(g)
--..E
...., 2NH3*
~
>. -200 I- '\Nz(g)·6H*
2NHz*.2H* -
a.
"0 2NH*+ 4H*
.c
C
w
2N*.6H*
-400 f- -
Fig. 2.4. Calculated enthalpy of the intermediates at 673 K for K-promoted Fe. Reproduced from
[396]
-100
0
E
--..
....,
-" -300
>.
01
Q;
C Nz(g)·6H*
CII
CII
~ -500
III
.0
.0
(5
-700
Fig. 2.5. Calculated Gibbs free energy off the intermediates at 673 K for K-promoted Fe. Repro-
duced from [396]
0.30
0.24
'<1
~
C
0
0.18
~
c:
QI
u
c 0.12
0
u Fig. 2.6. Trends in the ammonia produc-
'"
J: tion from 1 m 2 of catalyst as a function of
z 0.06 the number of d-electrons in the substrate.
The reaction conditions are kept constant
I atm, 400 DC, stoichiometric gas). Repro-
0 duced from [724]
4 6 8
Nli
10- 2 ~ H* -
III
e
Ol
~
0
U
f'-- II
Fig. 2.7. Coverages of intermediates on
10- 4 - - the surface of a catalyst operating at
10.1 MPa, 673 K, The concentration of
NH3 is 0 at the inlet and about 70% of
the thermodynamic equilibrium concen-
tration at the outlet. Reproduced from
-61~2* [645]
10 inlet oullet
Length through catalyst bed
by free sites, and since the calculated rate appears to be correct, it is unlikely that
the calculated coverages should be significantly in error [396].
At conditions where the concentrations of NH3 in the gas phase is signifi-
cant, the coverages by NHx* are high. At these conditions the sequence of
coverages is N* > NH* > NH2* > NH 3*. [396]. This sequence is determined
by the difference in entropy between the intermediates [396].
Evidence for the existence of significant amounts of N * has been found from
interpretation of reaction orders [607,614,617, 652-654J, from a comparison of
work function, electrical resistance and catalyst activity [655J, from laser flu-
orescence [656J, from the observation of N* by electron spectroscopy on the
catalyst [52J or a Fe single crystal [414, 603, 604J after exposure to NH3*
synthesis conditions, and from thermodynamic estimates based on data mea-
sured for the intermediates on single crystal surfaces [550].
Gravimetric measurements of adsorbed N* during NH3* synthesis have
been performed [191, 494, 657]. For a (Fe, AI) catalyst model the coverage by
N* is 0.52-0.69 [191J or 0.11-0.14 [378]. The coverage depends on the operat-
ing conditions [378].
Evidence for the existence of significant amounts of NHx* has been deduced
from interpretations of the reaction orders [512, 614, 658, 659J, laser fluore-
scence [656, 660J, and from the study of the dissociation of NH 3*/Fe(1l0) by
electron spectroscopy [661]. It is a complication in the deduction of the surface
coverages by intermediates that the coverages depend on the operating condi-
tions of the catalyst [396]. These variations may be sufficient to change the
nature of the most abundant intermediate [396]. Experimental evidence has
been found for changes in the nature of the most abundant reaction intermedi-
ates with temperature or promoter concentration [614].
The numerical models of NH3 synthesis predict that the coverages of H* is
quite large if no NH3 is present in the gas phase [396]. At low temperatures and
low conversions a dramatic increase inactivation energy has been observed. The
78 P. Stoltze
variations of the activation enthalpy with the partial pressure of NH3 will be
discussed in details in Sect. 7.6.2.
Evidence for the hydrogenation of NHx* being rate limiting has been found
from theoretical considerations [633], from interpretation of reaction orders for
NH3 synthesis [423] and for NH3 decomposition [620, 664, 678-681].
",--------------- -----
2{ -
I
~ r'--------------------------~
o
j Or -
)
-2~'····· .......... ··.......... ·............ ·· .... ·...... ··· .... ··· ...... ···· ..- Fig. 2.9. Calculated reaction orders for
N2 (solid curve), H2 (dashed curve),
and NH3 (dotted curve) for NH3 syn-
thesis at 10.1 MPa, N: H ratio 1: 3,
-4~~~~~~L-1~~1~~~1~~~ 673 K. Reproduced from [396]
o 0.2 0.4 0.6 0.8 1.0
Conversion
82 P. Stoltze
reaction enthalpy
Nz(g) + *~Nz* Hl
NH3 + *~N* + ~Hz H3
NH3 + *~NH* + Hz H4
NH3 + *~NHz* + 1Hz Hs
NH3 + *~NH3* + H6
H2 + 2*~2H* H7
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 83
and if; is the activation enthalpy for the rate limiting step.
N z* + *~2N* (33)
This expression may be interpreted by noting that Hl + H; is the activation
enthalpy for the dissociative adsorption of N*. The combinations of coverages
and enthalpies'in the expression are all enthalpies for desorption reactions. The
equation thus says that the activation energy for ammonia synthesis equals the
activation enthalpy for the rate limiting step plus the averaged cost of creating
more free sites [396, 625, 645].
The contribution from the rate limiting step is small [396, 625, 645]. At
typical conditions the dominating term originates from creation of free sites by
desorption of N *
(34)
The activation energy is not quite constant, see Fig. 2.10. The calculated
values are in agreement with experimental values, provided the experimental
values used do not span too large a range of operating conditions [396, 625,
645].
At conditions of a vanishing partial pressure of NH 3 , (}N., (}NH. (}NH2" and
(}NH3' will all vanish and the contribution from the reaction
(35)
becomes detectable [396, 625, 645]. This results in a higher activation energy
than under more usual conditions [396, 625, 645].
The reported values for the activation energy for NH3 synthesis are
14 kcaljmole [671], 48 kcaljmole [676], 43.3 kcaljmole [617], 23 kcaljmole
[670], 11.5 kcaljmole [677], 27 kcaljmole [708], 16 kcaljmole [524, 525],
50 kcaljmol [520] and 200 kllmole [696].
The rate ofNH 3 synthesis has been measured at 20 atm on single crystals of
Fe [603, 604J in a high pressure microreactor. The activation enthalpy for NH3
synthesis was 19.4 kcaljmole in the absence of K [603, 604J and
18.8 ± 0.5 kcaljmole for KjFe(100) [232J.
The reported values for the activation energy for NH3 decomposition are
0-16.6 kcaljmole [662J, 18.7-26.1 kcaljmol [662J, 31.9 kcaljmole [679J,
44 kcaljmole [236J, and 46 kcaljmole [670]. The value depends on the tempera-
ture [236, 662J and on the bulk composition of the catalyst.
The NH3 synthesis is 2.5 [709J, or 3.45 [710J times faster in D2 than in H 2.
The D-isotope effect is identical to the thermodynamic isotope effect if the
most abundant surface intermediate is N* or NH* [693]. This is consistent with
the dissociative mechanisms where H * is not involved in the rate-limiting step.
Based on Laser Raman spectroscopy failing to detect N* or NH*, the
interpretation of the inverse D2 isotope effect as a thermodynamic isotope effect
has been questioned [473].
The stochiometric number [657, 680, 699, 711J for NH3 synthesis is defined as
the number of turnovers for the rate limiting step, necessary for one turnover of
the total reaction written as N2 + 3H 2 = 2NH 3. If chain reactions are neglect-
ed, the stochiometric number is 1 if the rate limiting step is N z chemisorption
and 2 if the rate limiting step is hydrogenation.
The stochiometric number can be determined experimentally from detailed
rate measurements using a reaction mixture ofN z H2 and NH 3, which is not in
nitrogen isotopic equilibrium.
The stochiometric number for NH3 synthesis is 1 [526,527,615,712, 713J or
2 [672-676]. For NH3 decomposition, the stochiometric number is 1 [527,528J
or 2 [673].
The interpretation of the measurements has caused some polemic [672,674,
677]. The adsorption of NH3 has been found to be important [527, 528].
Neglecting the adsorption of NH3 at high partial pressures could give an
apparent value of 0 for the stochiometric number [527J; this may have been
a problem in some measurements [526].
It has been concluded that some of the earlier measurements were made at
low conversion, where the rate is not inhibited by NH3 and the stochiometric
number is undefined [677]. Other complications are the suggestions that the
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 85
2.7.9 Poisoning
:g, 0.6
e
~
u 0.4 \
\
\
\ Fig. 2.11. Calculated coverages by N *
\ (solid curve), 0* (dashed curve) and H*
"" (dotted curve) for a catalyst operating at
" ......
......
.... _- 10.1 MPa, N: H ratio 1: 3, 28% conver-
sion, 10 ppm H 2 0. Reproduced from
[396]
900 1000
Temperature (K)
4------------------~----~----~
...................................................
......
......... .
2
----------
'" 0 toward '" - 2 [396J, see Fig. 12. Experimentally, the reaction order for
H 2 0 is - 1.0 [670, 720].
The inclusion of Eq. 36 in the reaction scheme causes an additional term in
the expressions for the activation energy [396, 625].
Ht = Hl + H~ + 2H l 8N2 * + 2H 3 8NH *
(41)
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 87
reaction enthalpy
N 2(g) + *;;::=N 2* Hl
NH3 + *;;::=N* + iH2 H3
NH3 + *;;::=NH* + H2 H4
NH3 + *;;::=NH2* + 1H2 Hs
NH3 + *;;::=NH3* + H6
H2 + 2*;;::=2H* H7
H 20 + *;;::=0* + H2 Hs
(42)
250.----,----,----,----,----,
(5
.§
-,
200
/
,-
,- .... - --
.>t! I
I
I
11: 150 I
"5
.l:
I
/
1:
(l> I
c: 100 I
o /
/
Fig.2.13. Calculated activation enthalpy
~ 50 ----- .....
"."'/ for NH3 synthesis for a partially poi-
soned catalyst operating at 10.1 MPli
~ (solid curve) resp. 101 kPa (dashed
curve), N: H ratio 1 : 3, 673 K, 28 % con-
version. Reproduced from [396]
H20 concentration
88 P. Stoltze
2.8 References
45. Peev T, Krylova AV, Bozhinova A (1981) Radiochern Radioanal Lett 47: 307
46. Bleskin 01, Lachinov SS, Mirkin AE (1984) Kinet Katal25: 702
47. Topsoe H, Durnesic JA, Boudart M (1974) J Phys (Paris) 35: C6 411
48. Lachinov SS, Rubinshtein AM, Akirnov VM, Klyachov-Gurvich AL, Konyukhova IN,
Kuznetsov LD, Levitskaya TT, Pribytkova NA, Slink in AA, Chesnokova RA (1964) Kinet
Katal 5: 478
49. Lachinov SS, Torocheshnikov NS, Sirnulina lA, Lyudkovskaya BG, Rakrnat-Zade AG,
Ugnachev VI, Alipur G, Sushcheva AE (1975) Tr Mosk Khirn-Tekhnol Inst 85: 20
50. Ludwiczek H, Preizinger A, Fischer A, Hosernann R, Schoenfeld A, Vogel W (1978) J Catal51:
326
51. Michel A, Pouillard E (1948) Cornptes Rend 227: 194
52. Ertl G, Thiele N (1919) Appl Surf Sci 3: 99
53. Peters C, Schafer K, Krabetz R (1960) Z Elektrochern 64: 1194
54. Wi1chinsky ZW (1949) Anal Chern 21: 1188
55. Westrik R (1953) J Chern Phys 21: 2094
56. Dry ME, Ferreira (1967) J Catal 7: 352
57. Drnitrov M (1979) Geterog Katal 4: 349
58. Klisurski DG, Mitov IG, Petrov KP (1980) Therrnochirn Acta 41: 181
59. Tornov T, Klissurski D, Mitov I (1982) Phys Stat Sol (a) 73: 249
60. Tricker MJ, Vaishnava PP, Whan DA (1982) Appl Catal 3: 283
61. Glodeanu F, Spinzi A, Nicolaescu IV, Galatchi G, Spinzi M (1979) Rev Rourn Phys 24: 161
62. Uchida H, Todo N (1955) Repts Govt Chern Ind Research Inst Tokyo 50: 23
63. Chen HC, Anderson RB (1972) J Coll Interfacial Sci 38: 535
64. Spinzi A, Galatchi G, Spinzi M (1979) Geterog Katal 4: 439
65. Rozin AT, Kornarov VS, Efros MD, Lerneshonok GS (1982) Vestsi Akad Navuk BSSR Ser
Khirn Navuk 31
66. Uchida H, Todo N (1956) Bull Chern Soc Japan 29: 20
67. Krabetz R, Peters C (1965) Angew Chern 77: 333
68. Krabetz R, Peters C (1963) Ber Bunsen-Ges Phys Chern 67: 390
69. Egyhazi T, Scholtz J Beskov VS (1984) React Kinet Catal Lett 24: 1
70. Kornarov VS, Efros MD, Rozin AT, Lerneshonok GS, Erernenko SI (1980) Dokl Akad Nauk
BSSR 24: 1098
71. Rajararn RR, Sermon PA (1985) J Chern Soc Faraday Trans 1 81: 2577
72. Rajararn RR, Sermon PA (1985) J Chern Soc Faraday Trans 1 81: 2593
73. Brown R, Cooper ME, Whan DA (1982) Appl Catal3: 177
74. Youssef AM, Ami NM (1978) Surface Technology 7: 469
75. Marakhovets LN, Sirnulina lA, Lachinov SS, Sobolevskii VS, Lytkin VP, Christozvonov DB
(1972) ObI Neorg Tekhnol 3
76. Chernysheva LA, Tovbin MV, Zabuga VYa, Efirnova NI (1980) Kinet Katal 18: 25
77. Tirnofeev VA, Lachinov SS, Torocheshnikov NS, Sirnulina lA, Lyudkovshaya BG, Rudnitskii
LA, Chudinov MG (1973) Tr Mosk Khirn-tekhnol Inst 738: 33
78. Baranski A, Bielanski A, Blasiak E, Dulski R, Rokosz A (1968) Chern Stosow Ser A 12: 45
79. Bielanski A, Baranski A, Musial U, Rokosz A (1967) Chern Stosow Ser A 11: 365
80. Jensen EJ, Topsoe H, Sorensen 0, Kragh F, Candia R, Clausen BS, Morup S (1977) Sc J Metall
6
81. Lachinov SS, Torocheshnikov NS, Sirnulina lA, Rakhrnat-Zade AG, Ugnachev VI, Alipur G
Garanina EF (1975) Tr Mosk Khirn- Tekhnol Inst 85: 22
82. Ert! G, Prigge D, Schlagl R, Weiss M (1983) J Catal 79: 359
83. Patyi L, Tsarev VI, Krylova AV, Oravets D, Farkas Z, Torocheshnikov NS (1979) React Kine!
Catal Lett 12: 165
84. Maxwell LR, Smart JS, Brunauer S (1951) J Chern Phys 19: 30
85. Maxwell LR, Brunauer S (1949) Phys Rev 76: 175
86. Pererrnan NI, Zubova IE, Rabina PD, Kuznetsov LD, Pavlov a NZ (1978) Tr Mosk Khirn-
Tekhnol Inst irn.D I Mendeleeva 99: 50
87. Pele L (1983) Rev Chirn (Bucharest) 34: 168
88. Aleksic BD, Terelecki-Baricevic A (1973) Bull Chern Soc Beograd 38: 447
89. Medeleanu V, Doca N, Stefanescu M, Bibolaru A (1979) Rev Chirn (Bucharest), 30: 751
90. Tovbin MV, Kuznetsov VA, Popovich ZP (1973) Kinet Katal 10: 55
91. Tovbin MV, Popovich ZP (1969) Ukr Khirn Zh 35: 787
90 P. Stoltze
144. Lemeshko ND, Zabuga VYa, Chernysheva LA (1983) Ukr Khim Zh (Russ Ed) 49: 122
145. Tikhonova ON, Zubova IE, Pavlova NZ, Ivanova RF, Lyubchenko YuA, Dmitrenko LM
(1973) Tr Mask Khim -Tekhnol Inst 73: 126
146. Aleksic BD, Klisurski D, Mitov I (1981) Izu Khim 13: 660
147. Aleksic B, Bogdanov S (1985) Thermochim Acta 93
148. Mitov I, Klisurski D, Aleksic B, Gyurova L, Nikolov 0 (1984) Izu Khim 17: 305
149. Wachs IE, Dwyer DJ, Iglesia E (1984) Appl Catal 12: 201
150. Lachinov SS, Krylova AV, Pavlova NZ, Zubova IE (1978) Tr Mask Khim Tekhnol Inst im D I
Mendeleeva 99: 47
151. Lachinov SS, Simulina lA, Tikhonova ON,. Torocheshnikov NS, Alipur G, Strekalova NM
(1975) Deposited Doc 3770
152. Yatsimirskii VK, Kovalenko VN, Ishchenko EV (1982) Ukr Khim Zh (Russ Ed.) 48: 614
153. Peev T (1985) Khim Ind (Sofia) 57: 250
154. Borghard WS, Boudart M (1983) J Catal 80: 194
155. Logan SR, Moss RL, Kemball C (1958) Trans Faraday Soc 54: 922
156. Pavlova NZ, Rogozhina SA, Kuznetsov DA, Zubova IE, Malysheva TYa (1968) Kinet Katal
9: 1390
157. Rogozhina SA, Lachinov SS, Pavlova NZ, Zubova IE, Kuznetsov DA (1966) Tr Mask Khim-
Tekhnol Inst 51: 173
158. Niemann W, Clausen BS, Topsoe H (1987) NATO ASI Ser B 158: 909
159. Niemann W, Clausen BS, Topsoe H (1987) Ber Bunsen-Ges Phys Chern 91: 1292
160. Fierro JLG, Horns N, Ramirez P della Piscina, Sueiras JE (1983) Z Phys Chern (Wiesbaden)
135: 235
161. Rayment T, Schlagl R, Thomas JM, Ert! G (1985) Nature (London) 315: 311
162. Silverman DC, Boudart M (1982) J Catal 77: 208
163. Emmett PH, Brunauer S (1934) J Am Chern Soc 56: 35
164. Nielsen A, Bohlbro H (1952) J Am Chern Soc 74: 963
165. Fagherazzi G, Galante F, Garbassi F, Pernicone N (1972) J Catal 26: 344
166. Hosemann R, Hentschel MP (1986) Yak -Tech, 35: 3
167. Uchida H, Terao I, Ogawa K (1964) Bull Chern Soc (Japan) 37: 653
168. Rudnitskii LA, Demina GN, Makurina NA, Dmitrenko LM, Mazus EI, Alekseev AM (1984)
React Kinet Catal Lett 25: 297
169, Westrik R, Zwietering P (1953) Proc K Ned Akad Wet 56: 492
170. Schafer K (1960) Z Elektrochem 64: 1191
171. Kolbl H, Schottle E (1961) Z Elecktrochem 65: 91
172. Zwietering P, Koks HTL (1954) Nature, 173: 683
173. Fischer A, Hosemann R, Vogel W, Kouteckly J, Pohl J, Ralek M (1981) Stud Surf Sci Catal 7:
341
174. Herbstein FH, Smuts J (1963) J Catal 2: 69
175. Hosemann R, Preizinger A, Vogel W (1966) Ber Bunsen-Ges Phys Chern 70: 796
176. Tamaru K (1964) Bull Chern Soc Japan 39: 771
177. Brunauer S, Emmett PH (1935) J Am Chern Soc 57: 1754
178. Emmett PH, Brunauer S (1937) Trans Electrochem Soc 71: 383
179. Topsoe H, Topsoe N, Bohlbro H, Dumesic JA (1980) Proc 7th Intern Congr Catal (Tokyo) 247
180. Artyukh YuN, Fedun OS, Zyuzya LA (1980) Kinet Katal18: 20
181. Dvornik OS, Kozub GM, StreJ'tsov OA (1977) Teor Eksp Khim 13: 546
182. Dvornik OS, StreJ'stov OA, Lyubchenko YuA (1975) Dokl Akad Nauk SSSR 221: 648
183. Rabina PD, PereJ'man ND, Znbova IE, Kuznetsov LD, Maravskaya GK (1977) Ukr Khim Zh
(Russ ed) 47: 83
184. Dvornik OS, Strel'tsov OA, Lytkin VP (1975) Ukr Khim Zh (Russ ed) 41: 428
185. Dvornik OS, Strel'tsov OA, Maiboroda VP (1975) Ukr Khim Zh (Russ ed) 41: 1329
186. Dvornik OS, Strel'tsov OA, Chernobrivets VL (1975) Ukr Khim Zh (Russ ed) 41: 544
187. Limin Li, Wenxiang Wang (1983) Gaodeng Xuexiao Huaxue Xuebao 4: 763
188. Strel'tsov OA, Fedun OS, Artyukh YuN, Lytkin VP (1977) 12th Tezisy Dokl Ukr Resp Konf
Fiz Khim 159
189. Dvornik OS, Strel'tsov OA, Lytkin VP (1975) Dopov Akad Nauk Ukr RSR Ser B 623
190. Bridger, Pole, Beinlich, Tomson (1947) Chern Eng Progr 43: 291
191. Scholten JJF, Konvalinka JA, Zwietering P (1960) Trans Faraday Soc 56: 262
192. Yatsimirskii VK, Vyaz'mitina OM, Kozlova TP (1971) Teor Eksp Khim 7: 645
193. Podgurski HH, Emmett PH (1953) J Phys Chern 57: 159
92 P. Stoltze
194. Artyukh YuN, Yas'mo VI, Loza AN (1973) Kinet Katal 14: 1599
195. Solbakken V, Solbakken A, Emmett PH (1969) J Catal 15: 90
196. Kuznetsov LD, Rabina PD, Dmitrenko LM, Mischenko ShSh, Zozulya VYu (1979) Khim
Prom-st Ser Azotn Prom-st 19
197. Schultz JM (1972) J Catal 27: 64
198. Emmett PH, Brunauer S (1937) J Am Chern Soc 59: 1553
199. Dry ME, du Pleissis JAK, Leuteritz GM (1966) J Catal 6: 194
200. Dimitrov M, Rusev R (1980) God Vissh Khim -Tekhnol Inst Sofia 26: 120
201. Rabina PD, Kuznetsov LD, Anisimova MI (1969) Khim Prom 45: 350
202. Rozin AT, Komarov VS, Efros MD, Lemeshonok GS (1980) Vestsi Akad Navuk BSSR Ser
Khim Navuk 27
203. Lachinov SS, Simulina lA, Alipur GA, Nefedova NV (1978) Tr Mosk Khim-Tekhnol Inst im D
I Mendeleeva 99: 36
204. Kharchenko EV, Tovbin MV (1970) Khim Prom Ukr 15
205. Yatsimirskii VK, Girenkova NI (1979) Kinet Katal 20: 168
206. Yatsimirskii VK, Tovbin MV, Girenkova NI (1977) 2nd Tezisy Dokl Vses Simp Akt Poverkhn
Tverd Tel 22
207. Brunauer S, Emmett PH (1940) J Am Chern Soc 62: 1732
208. Chudinov MG, Perov VM, Ksenzenko VI, Alekseev AM (1986) Khim Prom-st (Moscow) 92
209. Weiss M, Ert! G (1982) Stud Surf Sci Catal 11: 277
210. Hanji K, Shimizu H, Shindo H, Onishi T, Tamaru K (1980) J Res Inst Catal Hokkaido Univ
28: 175
211. Hanji K, Shimizu H, Shindo H, Onishi T, Tamaru K (1980) J Res Inst Catal Hokkaido Univ 28:
175
212. Pernicone N, Fagherazzi G, Galante F, Garbassi F, Lazzerin F, Mattera A (1972) Proc 5th
Intern Congr Catal 2: 1241
213. Artyukh YuN, Boldyreva NA, Rusov MT (1970) Kinet Katal 11: 1531
214. Sasa Y, Uda M, Toyoshima I (1986) J Mater Sci Lett 5: 470
215. Dmitrov M (1980) Khim Ind (Sofia) 449
216. Chinh Chuong Zang, Krylova AV, Klimova GN, Lachinov SS, Torocheshnikov NS (1975) Tr
Mosk Khim-Tekhnol Inst 85: 24
217. Ivanov MM, Rudnitskii LA, Rabina PD, Kuznetsov LD (1968) Kinet Katal 9: 1239
218. Poniewierski Z, Tchorzewski T (1975) Chern Stosow 19: 161
219. Alekseeva MP, Lachinov SS, Predoiu C, Totocheshnikov NS (1968) Probl Kinet Katal Akad
Nauk SSSR 12: 175
220. Emmett PH, Brunauer S (1937) J Am Chern Soc 59: 310
221. Strongin DR, Somorjai GA (1988) Catal Lett 1: 61
222. Rudnitskii LA, Ivanov MM (1970) Kinet Katal 11: 207
223. Artyukh YuN, Bondarenko RN, Golovatyio VG, Kozub GM (1975) Dopov Akad Nauk Urk
RSR Ser B 121
224. Pennock GM, Flower HM (1984) Conf Ser - Inst Phys 68: 263
225. Van Ommen JG, Bolink WJ, Prasad J, Mars P (1975) J Catal 38: 120
226. Bonzel HP, Broden G, Krebs HJ (1983) Appl Surf Sci 16: 373
227. Bare SR, Strongin DR, Somorjai GA (1986) J Phys Chern 90: 4726
228. Broden G, Bonzel HP (1979) Surf Sci 84: 106
229. Ertl G, Weiss M, Lee SB (1979) Chern Phys Lett 60: 391
230. Seip U, Bassignana IC, Kiippers J, Ert! G (1960) Surf Sci 160: 400
231. Whitman LJ, Bartosch CE, Ho W (1986) J Chern Phys 85: 3688
232. Strongin DR, Somorjai GA (1988) J Catal 109: 51
233. Strongin DR, Bare SR, Somorjai GA (1987) J Catal 103: 289
234. Connell G, Dumesic JA (1985) J Catal 92: 17
235. Berengarten MG, Abdukadyrova SA, Zuvoba IE, Rabina PD, Rudnitskii LA, Pavlova NZ
(1973) Tr Mosk Khim-Tekhnol Inst 72: 14
236. Love K, Brunauer S (1942) J Am Chern Soc 64: 745
237. Rudnitskii LA, Ivanov MM (1969) Dokl Akad Nauk SSSR 104: 139
238. Rudnitskii LA, Ivanov MM (1969) Dokl Akad Nauk SSSR 184: 886
239. Abdukadyrova SA, Rabina PD, Zubova IE, Lyudkovskaya B (1970) Tr Mosk Khim- Teknol
Inst 67: 148
240. Kuznetsov LD, Rabina PD, Lopukhov GA (1971) Tr Nauch-Issled Proekt Inst Azoth Prom
Prod Org Sin 11: 92
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 93
433. Lu J-P, Albert MR, Bemasek SL, Dwyer DJ (1988) Surf Sci 199: L406
434. Kummer JT, Emmett PH (1951) J Am Soc 73: 2886
435. Eischens RP (1952) J Am Chern Soc 74: 6167
436. Wedler G, Ruhmann R (1982) Appl Surf Sci 14: 137
437. Seip U, Tsai C, Christmann, Kuppers J, Ert! G (1984) Surf Sci 139: 29
438. Wedler G, Colb KG, McElhiney G, Heinreich W (1978) Appl Surf Sci 2: 30
439. Beebe, Stevens (1940) J Am Chern Soc 62: 2134
440. Krylova AV, Chechulina GN, Koroleva TL, Lachinov SS, Torocheshnikov NS (1973) Depos-
ited Doc. VINITI 7698
441. Ueda K, Enatsu M (1985) Surf Sci 159: L421
442. Ueda K, Enatsu M (1985) Surf Sci 159: L421
443. Benndorf C, Kruger B, Thieme F (1985) Surf Sci 163: L675
444. Cameron SD, Dwyer DJ (1988) Surf Sci 198: 315
445. Alshorachi G, Wedler G (1985) Appl Surf Sci 20: 279
446. Moon DW, Dwyer DJ, Bemasek SL (1985) Surf Sci 163: 215
447. Huang V-V, Emmett PH (1972) J Catal 24: 101
448. Gafner G (1979) S Afrl J Phys 2: 129
449. Rao CNR, Ranga G (1988) Chern Phys Lett 146: 557
450. Behner H, Spiess W, Wedler G, Borgmann D (1986) Surf Sci 175: 276
451. Pimer M, Bauer R, Borgmann D, Wedler G (1987) Surf Sci 189/190: 147
452. Peev T, Krylova A, Nefedova N, Stoilova T (1986) J Radioanal Nucl Chern 105: 157
453. Krylova AV, Nefedova NV, Peev TM, Stoilova TI (1986) Kinet Katal 27: 520
454. Krylova AV, Ustimenko GA, Nefedova NV, Peev TM, Torocheshnikov NS (1986) Appl Catal
20: 205
455. Krylova AV, Nefedova NV, Torocheshnikov NS (1988) Appl Catal 39: 325
456. Krylova AV, Nefedova NV, Peev TM (1987) Stud Surf Sci Catal 34: 625
457. Peev T, Krilova A, Nefedova N, Stoilova T (1986) Khim Ind (Sofia) 58: 2
458. Ponec V, Knor Z (1968) J Catal 10: 73
459. Hall PG, Hope CJ (1970) J Chern Soc A 1970: 2003
460. Tomanek D, Bennemann KH (1985) Phys Rev B 31: 2488
461. Freund HJ, Bartos B, Messmer RP, Grunze M, Kuhlenbeck H, Neumann M (1987) Surf Sci
185: 187
462. Grunze M, Golze M, Fuhler J, Neumann M, Schwarz E (1985) In Proc 8th Int Congr Catal4:
133
463. Grunze M, Golze M, Hirschwald W, Freund H.-J, Pulm H, Seip U, Tsai MC, Ert! G, Kuppers J
(1984) Phys Rev Lett 53: 850
464. Grunze M, Strasser G, Golze M (1987) Appl Phys A A44: 19
465. Borgmann D, Kroninger F, Steidl P, Wedler G (1987) Surf Sci 189-190: 485
466. Ert! G, Lee SB, Weiss M (1982) Surf Sci 114: 527
467. Ert! G, Lee SB, Weiss M (1982) Surf Sci 114: 515
468. Bozso F, Ert! G, Weiss M (1977) J Catal 50: 519
469. Wedler G, Borgmann D, Geuss J (1975) Surf Sci 47: 592
470. Tsai MC, Seip U, Bassignana IC, Kuppers J, Ertl G (1985) Surf Sci 155: 387
471. Kishi K, Roberts MW (1977) Surf Sci 62: 252
472. Baro AM, Eley W (1980) Surf Sci 112: L 759
473. Liao D, Zhang H, Wang A, Cai Q (1987) Sci Sin Ser B (Eng\. Ed.) 30: 246
474. Bozso F, Ert! G, Grunze M, Weiss M (1977) J Catal 49: 18
475. Whitman LJ, Bartosch CE, Ho W, Strasser G, Grunze M (1986) Phys Rev Lett 56: 1984
476. Toyoshima 1(1981) Shinku, 24: 216
477. Ert! G, Huber M (1980) Z Phys Chern (Wiesbaden), 119: 97
478. Grunze M, Kind In D, Woodruff DP (1982) eds, The Chemistry and Physics of Solid Surfaces
and Heterogeneous Catalysis, Vol 4, p 413
479. Wedler G, Steidl G, Borgmann D (1980) Surf Sci 100: 507
480. Wedler G, Borgmann D, Alter W, Witan K (1986) Ber Bunsen-Ges Phys Chern 90: 235
481. Kummer JT, Emmett PH (1951) J Chern Phys 19: 289
482. Magomedbekov EP, Kasatkina LA (1976) Zhur Fiz Khim 50: 2414
483. Morikawa Y, Ozaki A (1968) J Catal 12: 145
484. Ert! G, Grunze M, Weiss M (1976) J Vac Sci Technol 13: 314
485. Imbihl R, Behm RJ, Ert! G, Moritz W (1982) Surf Sci 123: 129
486. Arabczyk W, Muessig HJ (1987) Vacuum, 37: 137
98 P. Stoltze
487. Dowben PA, Grunze M, Jones RG (1981) Surf Sci 109: L519
488. Wedler G, Borgmann D, Forchhungsber Wehrtech (Bundesrnisist. Verteidigung)., No. BMVg-
FBWT 76-20 Luft-Raurnfahrt Teil. 2: 151
489. Evdokirnova ZhA, Valitov NKh (1985) Zh Prikl Khirn (Leningrad) 58: 2121
490. Vito I EN, Orlova KB (1973) Kinet Katal 14: 1514
491. Kwan T (1953) J Res lust Catal Hokkaido Univ 3: 16
492. Kwan T (1955) J Res Inst Catal Hokkaido Univ 3: 109
493. Brito V, Ralek M (1978) React Kinet Catal Lett 9: 15
494. Zwietering EP, Roukens JJ (1954) Trans Faraday Soc 50: 178
495. Scholten JJF, Zwietering P, Konvalinka JA, de. Boer HJ (1959) Trans Faraday Soc
55: 2166
496. Boreskova EG, Kuchaev VL, Temkin MI (1979) Kinet Katal 20: 147
497. Minachev KhM, Evdokirnova ZhA, Valitov NKh (1982) Dokl Akad Nauk SSSR
266: 296
498. Grabke HJ (1976) Z Phys Chern (Frankfurt am Main), 100: 185
499. Ostrovskii VE, Igranova EG (1978) Kinet Katal 19: 538
500. Paal Z, Ert! G, Lee SB (1981) Appl Surf Sci 8: 231
501. Boheirn J, Brenig W, Engel T, Leuthauser U (1983) Surf Sci 131: 298
502. Rettner CT, Stein H (1987) J Chern Phys 87: 770
503. Rettner CT, Stein H (1987) Phys Rev Lett 59: 2768
504. Rettner CT, Pfnur HE, Stein H, Auerbach DJ (1988) J Vac Sci Technol A 6: 899
505. Durnesic JA, Boudart M (1975) Proc Syrnp Catal Chern Nitrogen Oxides p 95
506. Egawa C, Iwasawa Y Chern Lett 1987: 959
507. Nakata T, Matsushita S (1982) J Chern Phys 76: 6335
508. Tarnaru K (1965) In: Sacht!er WMH, Schuit GCA, Zwietering P (eds), Proc 3,d Intern Conf
Catal 1: 664
509. Takezawa N (1972) J Catal 24: 417
510. Schmidt WA (1968) Angew Chern Int Ed Engl 7: 139
511. Artyukh YuN, Golovatyi VG, Korol EN (1980) Teor Eksp Khirn 16: 714
512. Okawa T, Onishi T, Tarnaru K (1977) Z Phys Chern (Wiesbaden), 107: 239
513. Karpinski W, Palczewska W (1974) Bulll'Acad Polon Sci Chirn 22: 159
514. Kozub GM, Strel'tsov OA, Rusov MT (1973) Kinet Katal 10: 52
515. Paal Z, Ert! G (1982) Kern Kozl 57: 51
516. Arthyukh YuN, Yas'rno VI, Rusov MT (1971) Kinet Katal 8: 51
517. Gay ID, Textor M, Mason R, Iwasawa Y (1977) Proc R Soc London A356: 25
518. Tarnarau K Prep. 2nd Intern Congr Catal (Paris), 1: 325
519. Tarnaru GK (1963) Trans Faraday Soc 59: 979
520. Joris GG, Taylor HS (1939) J Chern Phys 7: 893
521. Boreskov GK, Kolchanova VM, Rachkovskii EE, Filirnonova SN, Khasin AV (1975) Kinet
Katal 16: 1218
522. Takezawa N, Toyoshirna I (1970) J Catal 19: 271
523. Kazusaka A, Toyoshirna I (1981) Z Phys Chern (Wiesbaden) 128: 111
524. Yatsirnirskii VK, Girenkova NI, Maksirnov YuV (1976) Teor Eksp Khirn 12: 263
525. Urabe K, Ozaki A (1978) J Catal 52: 542
526. Tanaka K, Matsuyarna A (1971) J Res lust Catal Hokkaido Univ 19: 63
527. Tanaka K (1966) J Res Inst Catal Hokkaido Univ 14: 153
528. Tanaka K (1965) J Res lust Catal Hokkaido Univ 13: 119
529. Napol'skikh GA, Kasatkina LA (1973) Kinet Katal 14: 653
530. Miyahara K (1956) J Res lust Catal Hokkaido Univ 4: 193
531. Vol'pin ME, Novikov YuN, Postnikov VA, Schur VB, Bayerl B, Kaden L, Wahren M,
Drnitrienko LM. Stukan RA, Nefed'ev AV (1977) Z Anorg AlIg Chern 428: 231
532. Morikawa Y, Ozaki A (1971) J Catal 23: 97
533. Boreskova EG, Kuchaev VL, Temkin MI (1984) Kinet Katal 25: 116
534. Schulz G, Schaefer H (1969) Z Phys Chern (Frankfurt am Main) 64: 333
535. Bowker M, Parker IB, Waugh KC (1985) Appl Catal 14: 101
536. Arnariglio H, Rambeau G (1976) Proc 6th Int Congr Catal 2: 1113
537. Bowker M, Parker I, Waugh KC (1988) Surf Sci 197: L223
538. Hinda F, Hirokawa K (1977) J Electron Spectrosc Related Phenorn 12: 313
539. Benndorf C, Madey TE, Johnson AL (1987) Surf Sci 187: 434
540. Erley W, Ibach H (1983) J Electron Spectrosc Related Phenorn 31: 61
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 99
588. Baer DR, Thomas MT (1986) Appl Surf Sci 26: 150
589. Nakanishi S, Sasaki K (1988) Surf Sci 194: 245
590. Brill R, Schafer H, Zimmermann G (1968) Ber Bunsen Ges Phys Chern 72: 1218
591. Zubova IE, Rabina PD, Pavlova NZ, Kuznetsov LD, Chudinov MG, Le. Congo Shang (1974)
Kinet Katal 15: 1261
592. Tovbin MV, Psheichnaya OV (1966) Kinet Katal 2: 149
593. Nwalor JU, Goodwin JG Jr, Biloen P (1989) J Catal117: 121
594. Ertl G, Huber M, Thiele N (1979) Z Naturforsch 34A: 30
595. LoeftIer DG, Schmidt LD (1976) J Catal 44: 244
596. Brill R (1970) J Catal 19: 236
597. Yatsimirskii VK (1978) Zh Fiz Khim 52: 584
598. Kuznetsova EP, Samchenko NP, Rusov MT (1974) Kinet Katal 11: 92
599. Mahapatra H, Chhabra DS, Puri VK, Sen SP (1981) Fert Technol 18: 160
600. Samchenko NP, Rusov MT, Strel'tsov OA (1966) Kinet Katal2: 96
601. Vasilevich AA, Rabina PD, Alekseev AM, Dmitrenko LM, Mosolova E, Lyubchenko YuA,
Muravskaya GK, Kuznetsov LD (1973) Izu Otd Khim Mauki Bulg Akad Nauk 6: 225
602. Brill R, Kurzidim J (1970) Colloq Int Cent Nat Rech Sci 187: 99
603. Spencer ND, Schoonmaker RC, Somorjai GA (1982) J Catal 74: 129
604. Spencer ND, Schoonmaker RC, Somarjai GA (1981) Nature (London) 294: 643
605. Strongin DR, Carrazza J, Bare SR, Somorjai GA (1987) J Catal 103: 213
606. Yatsimirskii VK (1982) Izv Sib Otd Akad Nauk SSSR Ser Khim Nauk 1982: 131
607. Rambeau G (1987) Bull Soc Chim Fr 1987: 815
608. Artyukh YuN, Zyuzya LA (1981) Kinet Katal 19: 17
609. Tornqvist E, Chen AA (1991) Catal Lett 8: 359
610. Fierro JLG, Horns N, Ramirez P, Sueiras J (1984) React Kinet Catal Lett 24: 179
611. Magomedbekov EP, Kasatkina LA (1978) Tr Mosk Khim-Tekhnol Inst im DI Mendeleeva
99: 60
612. Chirstozvonov DB, Lytkin VP, Sobolevskii VS, Kozlov LI, Markovets LN, Kirillov IP,
Korbutova ZV (1969) Izv Vyssh Ucheb Zaved Khim Khim Teknol12: 1388
613. Miki Y, Terao I, Uchida H (1967) Tokyo Kogyo Shikensho Hokoku 62: 64
614. Altenburg K, Bosch H, Van. Ommen JG, Gellings PJ (1980) J Catal 66: 326
615. Scholten JJF, Zwietering P (1957) Trans Faraday Soc 10: 1363
616. Parker IB, Waugh KC, Bowker M (1988) J Catal 114: 457
617. Nielsen A, Kjaer J, Hansen B (1964) J Catal 3: 68
618. Krabetz R, Peters C (1963) Ber Bunsen-Ges Phys Chern 67: 381
619. Bosch H, van Ommen JG, Gellings PJ (1985) Appl Catal 18: 405
620. Takezawa N (1966) Shokubai, 8: 390
621. Rudnitskii LA, Berengarten MG, Alekseev AM (1973) J Catal 30: 444
622. Rudnitskii LA, Berengarten MG (1972) Kinet Katal 13: 115
623. Samchenko NP, Golodets GI (1986) Kinet Katal 27: 378 (russ) 324 (eng)
624. Samchenko NP, Golodets GI (1986) Kinet Katal 27: 384 (russ) 329 (eng)
625. Stoltze P, Nl'Jrskov JK (1987) J Vac Sci Technol A 5: 581
626. Nl'Jrskov JK (1981) J Vac Sci Technol 18: 420
627. Holloway S, Lundqvist BI, Nl'Jrskov JK (1984) Proc 81h Intern Congr Catal (Berlin) 4: 85
628. Nl'Jrskov J, Holloway S, Lang ND (1984) Surf Sci 137: 65
629. Rudnitskii LA, Berengarten MG (1971) Dokl Akad Nauk SSSR 201: 396
630. Markert K, Wandelt K (1985) Surf Sci 159: 24
631. Schlagl R, Schoonmaker RC, Muhler M, Ertl G (1988) Catal Lett 1: 237
632. Konoplya MM, Gorlov Yul, Yatsimirskii VK (1982) Teor Eksp Khim 18: 398
633. Kai-Hui.Huang. (1981) Sci Sin (Engl. ed) 24: 800
634. Kai Huei Huang (1981) Stud Surf Sci Catal 7 A: 554
635. Tomanek D, Kreuzer HJ, Block JH (1986) J Phys Colloq 1986: 139
636. Tomanek D, Kreuzer HJ, Block JH (1985) Surf Sci 157: L315
637. Ortoleva E, Simonetta M (1985) Croat Chern Acta 57: 1387
638. Gagarin SG, Chuvylkin ND (1981) Zh Fiz Khim 55: 3094
639. Guzikevich AG, Chuiko AA, Yatsimirskii VK (1988) Teor Eksp Khim 24: 263
640. Tomanek D, Kreuzer HJ, Block JH (1985) Surf Sci 157: L315
641. Holloway S, Hodgson A, Halstead D (1988) Chern Phys Lett 147: 425
642. Boudart M (1988) Catal Lett 1: 21
643. Boudart M, LoftIer DG (1984) J Phys Chern 88: 5763
2 Structure and Surface Chemistry of Industrial Ammonia Synthesis Catalysts 101
Contents
3.2 Ammonia Synthesis Activity of Elements and Promoter Effects ... 105
3.2.1 Properties of the Elements in the Activation of Dinitrogen . . 105
3.2.2 Properties of the Elements in Ammonia Synthesis . . . . . . . 109
3.2.3 Alloying Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
3.2.4 Support and Promoter Effects . . . . . . . . . . . . . . . . . . . 114
3.2.4.1 Electron Donation to the Active Center . . . . . . . . 115
3.2.4.2 Structure Sensitivity . . . . . . . . . . . . . . . . . . . . 118
3.2.5 Preparation, Activation and Nitridation . . . . . . . . . . . . . 119
3.1 Introduction
Table 3.1. Reactivity of the elements with N2 and properties of their nitrides
IA
Li + + Li3N - 197 (S)
Na Na3N - 151 150
K K3N + 84 L
Rb Rb 3N + 180 L
Cs Cs 3N + 314 L
IIA
Be + Be 3N2 - 285 > 220
Mg + Mg3N 2 230 700
Ca +.+ Ca 3N2 - 213 H
Sr + + Sr3N2 - 197 H
Ba + + Ba3N2 - 184 H
iliA
Sc + ScN -285 H
Y + YN - 301 H
La + LaN - 301 H
IVA
Ti + TiN - 305 H
Zr + ZrN - 343 > 3000
Hf + HfN - 326 H
VA
V + VN -172 > 2300
Nb + NbN - 247 > 2300
Ta + TaN - 243 > 3000
VIA
Cr + CrN - 121 H
Mo + Mo 2N -71 H
W + WN - 71 H
U + UN - 335 H
VilA
Mn + MnsN2 -117 > 1200
Tc TcN
Re Re2N
VIII
Fe Fe4 N -12 440
Co Co 3N
Ni Ni3N +0
Ru
Rh
Pd
Os
Ir
Pt
IB
Cu CU3N + 75 450
Ag Ag3 N + 285 Ex
Au AU3N Ex
liB
Zn Zn3N2 -12 H
Cd Cd 3N2 + 79
Hg Hg3N 2 +8
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 107
II1B
B + BN - 134 > 3000
Al + Al - 243 2000
Ga GaN - 105 H
In InN - 21 H
TI Ti3N + 84
IVB
C + (CN), + 155
Si + Si3 N4 - 188 H
Ge Ge3N4 - 17 450
Sn Sn3N4 < 360
Pb
VB
P + PN - 84 750
As
Sb
Bi BiN
VIB
0 + NO +92
S S4N4 + 134 178 Ex
Se Se4N4 + 176 U
Te Te3N4
VIIB
F NF3 -109 S
Cl NCI 3 + 230 Ex
Bi NBr3 + 335 U
I NI3 +272 Ex
a + +: Reacts with N 2 directly below 300 'C, +: Reacts with N 2 directly with N 2 above 300°C,
- : Made from nitrogen compounds, - -: nitride unknown
b L: Low temperature, H: High temperature, S: Stable, U: Unstable, Ex: Explosive
chemisorb N z at room temperature are: IlA (Ca, Sr and Ba), IVA (Ti, Zr and
Hf), VA (V, Nb and Ta), VIA (Cr, Mo and W), VIlA (Re), and VIII (Fe) [30, 31].
Note that they are mostly found in Group IV A through VIII, which are known
to form interstitial nitrides. It is interesting to note that IIA metals chemisorb
N z and these metal "nitrides" are quite effective catalysts for the isotopic
equilibration of N z. However, they are inactive for ammonia synthesis because
they react to form hydrides in an Nz-H z mixture [4, 32, 33J, which will be
shown in Sects. 3.3.2 and 3.3.3.4.
A similar chemisorption can take place on other metals which do not form
a nitride from N z. The much lower ability of other metals to chemisorb
N z seems to come primarily from the difficulty in activating the N z molecule.
Even a copper surface can chemisorb N z when the copper surface is activated by
ion bombardment [34J, even though copper nitride, CU3N, is unstable.
Chemisorption ofN z was found on reduced cobalt oxide with a potassium oxide
promoter at room temperature [35, 36] and even on noble metals (Ru, Rh,
108 K. Aika and K. Tamaru
Os and Ir) promoted with alkali metals. Promoter action will be shown in
Sect. 3.2.4.
The second reason for an inability for N z chemisorption may come from an
insufficient metal-nitrogen bond energy. The heat of chemisorption of N z has
been measured over metals and is shown in Table 3.2. The differential heat of
chemisorption generally decreases with an increase in the surface coverage [37].
Although the initial heat of chemisorption is much larger on vapor deposited
films than on powder or supported metal, the value on the vapor deposited films
decreases more rapidly, presumably because of a larger degree of disorder in
crystallinity [38]. The data shown in Table 3.2 are the initial heat of adsorption
and the desorption energy.
IVA
Ti 481 610
Zr 657 686
Hf 816 652
VA
V 469 344
Nb 582 > 502' [39] 494
Ta 732 590 b [40] 486
VIA
Cr 410 439 b [38] 242
Mo 335 263' [41], 289' [42], 259" [43] 142
W 536 397 b [38], 385' [44], 314' [45], 142
389' [46], 334-372c [47]
VIlA
Mn 465 234
Tc 126
Re 167 284-313 c [48]
VIII
Fe 205 293 b [49] 24
Co 134
Ni 138 0
Ru -117 92-167" [50]
Rh -146
Pd - 209 - 159' [51]
Os - 67
Ir -109 242d [52]
Pt -142 25 c [53], 92 d [53]
"Ref. [54]
b Initial heat of adsorption on film at room temperature
c Initial heat of adsorption on filament at room temperature.
d Desorption energy on filament
, Estimation from the heat of dissociative adsorption of NO and O 2
, From Table 3.1
g Ru or RujAl 20 3 with K
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 109
Systematic studies of the catalytic activity of single metals ofNH 3 synthesis were
first made by Haber as described in the preceeding paragraphs. In these studies
not only readily reducible metals were tested but also less reducible ones as well,
110 K. Aika and K. Tamaru
Sr Mo Te Ru Rh
Sa Ce W Re Os Ir
such as cerium which was reduced with magnesium. The results of these early
studies are summarized in Fig. 3.2 [65]. In addition to the metals shown, Re
[66], Cr [67], V [68], Rh [69], Ir [69] and Tc [70] act as NH3 catalysts.
Platinum was also used [69, 71], but has a poor activity. Some of the above
metals, such as Mo, V and U are transformed into nitrides during the reaction.
The reverse reaction, NH 3 decomposition, has been studied over a series
of vapor deposited metal films [72]. The catalytic activities are also shown in
Fig. 3.2.
It is obvious that osmium and iron are the most effective elements under the
conditions studied by Haber. On the other hand, ruthenium is the most active
metal in ammonia decomposition. Since the two reactions are forward and
backward steps of the same reaction, the most active metals should be the same,
at least near equilibrium. The difference disclosed above is probably caused by
some discrepancy in the reaction conditions. It is obvious that the H2/NH3 ratio
is much larger in the synthesis reaction than in the decomposition reaction.
Adsorbed nitrogen and hydrogen often become a retarding species, which
depends upon the H2/NH3 ratio. The effect of such variables will be discussed
later.
In the 1970s, a catalyst system promoted by metallic potassium [73, 74] was
studied. The ammonia synthesis rates at 80 kPa and 588 K over transition
metals supported on active carbon and promoted by metallic potassium are
given in Fig. 3.2 [69]. The activity of isotopic equilibration ofN 2 over the same
series of catalysts at 30 kPa of N2 and at 588 K are shown in Fig. 3.3 [75]. The
same reaction over Raney metals are also shown in this figure [76]. In these
cases ruthenium is the most active metal. There is a common belief that Fe, Ru
and Os are the most active elements in ammonia synthesis, ammonia decompo-
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 111
.
l' ~ ___
15
'Ill
N 14
\
decomposition
'E
u
u
13- ~~I~\'
Q.o
12
'1\
(5
\
E
11 ,, 3
a: ,,
~ 10 ,
9 synthesis 2
5%M - KIA. C. 588K ~
'01
"-o-.. ~ ..
0;-
~
(5
synthesis \
E
2 M powder 773 K ::1.
"0 0 a::
a; 01
':;' 9- Fig. 3.2. Rate of ammonia syn-
M
J: thesis and decomposition over
Z -1
0 various metal catalysts: De-
~
composition at 0.2-0.8 kPa [72],
0 synthesis at 80 kPa (5%
M-K/AC) [69,74] and 5 M Pa
(M powder) [65]
equilibration at 588 K on
Raney M(o). 5% M-K/AC (.)
and SrNo.s(o)
o
4
3
'01
~ 2
(5
E
::1.
a:
01
9-
0
-1
Fig. 3.3. Rate of isotopic equilibration of
dinitrogen e
S N 2 + 30N2 = 2 29 N 2) over
• Space velocity 15000 h - 1, catalyst volume 0.5 ml, catalyst weight 0.5-1 g,
Tc; radioactive element
b Rate constant from Temkin equation (ex = 0.5, k 400 = 15000x2(1 - X2)-1).
Activation energy of k400 = 40-50 kcal/mol
C Relative concentration of ammonia compared to the value at equilibrium
(CeH2 +x) and transition metal phases during the synthesis condition
(N2 + 3H 2 = 50 bar, 450 to 550 QC). These phases are considered to be the active
component. However, these states are quite sensitive to oxygen containing
compounds (air, H 20 and CO) forming an inactive phase Ce02/transition metal
[272].
TiFe alloy turned out to be an active ammonia catalyst which is composed of
the mixture of Fe, TiN and TiO x surface phases mounted on a TiFe bulk phase
[30]. However, TiRu has no activity because TiN and Ti0 2 cover the surface of
TiRu and Ru bulk phase [90]. Fe91Zr9 alloy is found to be an active catalyst,
Fe-ZrOx, under the reaction conditions [91]. A Re-Pt bimetallic cluster is
thought to be formed on an Al 20 3 support [92]. It does not seem that alloying
induces drastic effects.
Precious metals which can be easily reduced are usually supported on non-
reducible oxides. Starting metal compounds may be reduced, forming metal
cluster particles. The activity is a function of the number of exposed metal
atoms. The particles have various crystal planes on the surface, each of which
may have different activities. The chemical and physical properties of the
support, concentration of metal compounds, and the reduction temperature
may influence the cluster forming process, which in turn influences the size of
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 115
metal particles (structural effect of the support). Supports and the third additives
(promoters) may have some electronic interaction with the metal clusters, which
may influence the activity (chemical effect).
Common metals often form mixed oxides with the support compounds. For
that reason common metals are usually used as massive metal catalysts. In the
case of massive metal catalysts, a few weight percent of a promoter is added.
Some promoters make mixed oxides with the active element and influence the
reduction process and the surface area (structural promoter). Others are depo-
sited on the metal surface and have an electronic interaction with the surface
(chemical effect). Ammonia activity on Fe is known to be enhanced by adding
Al 20 3 and K 20. It is believed that Al 20 3 stabilizes the high surface area of Fe
(structural effect) and K 20 promotes the ammonia activity per Fe surface area
(chemical effect). The structural effect is well studied on Fe single crystal
surfaces, where Fe(lll) is the most active plane and Fe(110) is the next and
Fe(lOO) is the least active plane [93]. Such studies have been expanded to other
catalysts such as Re, and will be reviewed in Section 3.2.4.2.
Various alloy catalysts have been studied as was shown in the previous
section. Raney metal catalyst which has small amounts of aluminum, is classified
as a massive catalyst. However, some intermetallic are transformed to the
supported metal catalysts when one of the components is separated and oxi-
dized or nitrided.
:.:: AI (15)
co
co
III
15 100
!:
o
15
E
~ Ru powder
J:"" (2.2)
Z 10
Fig. 3.4. Rate of ammonia synthesis
at 588 K and N2 + 3H 2 = 80 kPa on
Ru (2 or 5 wt %) catalysts with vari-
ous supports and promoters as a
function of average e1ectronegativi-
ties of compounds [94, 96]
Electronegativityof compounds
70>
~
3
-0
E
E
:.::
co
co
III
15 2
!:
.Q
15
-
E
5
J:
Z
'"
'5
<lI
15 Fig. 3.5. Effect of alkali metal addi-
0::
tion on the rate of ammonia synthesis
No at N2 + 3H 2 = 80 kPa on 2.5%
Ru-AC (1.0 g) at 588 K [73]
4 6
Alkali metal (mmol/g)
Table 3.5. Support and promoter 'effect in ammonia synthesis (N z + 3H 2 = 80 kPa) and N z isotopic equilibration (N2 = 20 kPa) on Ru catalysts
[94,96]
Ru surface
(1121) 314
(1120) 8.5 x 10 16 98
(1010) 2.5 x 1015 6
(0001) 5 X 10 13 0.03
vapor or presulfidation has little effect on its activity. This is contrast to the
behavior of iron. Ammonia synthesis rates were measured on Re single crystals
at 870 K under 2 MPa of N2 + 3H 2 . The results are shown on Table 3.6. The
basal plane (0001) was inactive. The open and rough planes (1010), and espe-
cially (1120) or (1121) were quite active for this reaction. The activity on (1121)
was ca. 104 times greater than that on (0001). The activation energy was
19.4 kcaljmol for all of the surfaces. The structure sensitivity of Re surfaces was
verified for ammonia synthesis [105, 106].
As shown in the previous section, ammonia catalysts are classified as two types:
The supported metal catalyst and the massive metal catalyst. For the supported
120 K. Aika and K. Tamaru
catalyst, metal salts such as chlorides or nitrates are impregnated on the support
materials and usually reduced with H2 or N2/H2 mixture. Promoter salts,
usually nitrates, are also impregnated on the supports. Mixed metal compound
salts, which are composed of metal elements and promoter elements, can also be
used as the starting materials. If the chloride ion may possibly remain after the
reduction, and its poisoning is not negligible, compounds free of chlorine should
be used. The starting materials and preparation methods may influence the
activity through the degree of metal dispersion or metal-support interaction as
was shown in the previous section.
Massive metal catalysts are usually made from a mixture of both metal and
promoter oxides by reduction with H2 or N2/H2 gas. Many studies of activation
processes have been carried out on Fe catalysts which are written elsewhere
[107]. The commercial iron catalyst needs to be reduced for a long period
because it is in a massive form and the oxide is not easily reduced completely.
The iron catalyst is activated under the ammonia synthesis condition [108],
during which the surface structure probably changes to form the active sites
[109]. On the other hand, supported precious metal catalysts such as Ru catalysts
need a short reduction time because their starting compounds are easily reduced.
Generally, they have no induction period with respect to the activity.
Alkali metals, especially potassium, were proved to be excellent promoters
for Ru catalysts. However, since metallic potassium is very reactive to water
vapor or oxygen-containing gases, the catalyst system has several problems,
such as a handling problem or deactivation. Several other preparation methods
were developed. K4Ru(CN)6 was prepared by alkali fusion of Ru metal and
KN0 3 or by adding KCN to RuCl 3 [110]. Although the hydrogenation of
K4Ru(CN)6 alone was very slow even at 773 K, alumina supported K 4Ru(CN)6
was readily hydrogenated at 653 K. The reaction stoichiometry was suggested
as follows:
F', ,
I
,\
-;- \\
0>
It
~ \
\
"0
E
\
::l.
....
:>::
a)
a)
Ln
"0 400
~
0:: o
Fig. 3.7. Effect of CsN0 3 addition on am-
200 o monia synthesis at 588 K on 2 wt % Ru cata-
lysts; Ru3(CO)12/support is treated with Hz
at 623 K for 4 h. Ru-CsOH/MgO (0 and e);
Ru-CsOH/AI 2 0 3 (0)
12
Cs/Ru mol ratio
122 K. Aika and K. Tamaru
Rate at 573K(mmol ' h-I .g-I)
....
o
Fe
Z
N
Fe-AI 2 O:3-K2 0 •
eN
Ru N
:I:
....ZI
4.7"10 Ru-A.C.- K
Raney Ru
:I:
Raney Ru-K Co>
Raney Ru-NaNO:J
Raney Ru- KN0:3
Raney Ru-Rbt)lO:J
Raney Ru-CsN0 3
5.0"10 Ru-Cs+'MgO
Fig. 3.8. Rate of ammonia synthesis at 573 K and 80 kPa on various Fe and Ru catalysts. Fe [123J,
Fe-AI 2 0 3 -K 2 0 [108, 124J, Ru and 4.7% Ru-A.C.-K [73J, Raney Ru and Raney Ru- K [84J,
Raney Ru-alkali nitrate [85, 86J, 5.0% Ru-Cs + j MgO [117J
CsN0 3 -Raney Ru is more active than the commercial iron catalyst and is one
of the most active ammonia catalysts under atmospheric pressure conditions
[85, 86J (see Fig. 3.8). However, the Ru catalysts might not be highly active at
high pressure because of hydrogen inhibition. Nevertheless, a recent work
suggests that hydrogen promotes the activation of N2 on Ru catalysts if
a lanthanide promoter is present [122].
Elements between groups lIlA and VIlA usually become nitrides under
ammonia synthesis conditions. Nitrides of uranium and molybdenum are
known to be active catalysts. When Mo0 3 is reduced with H 2- CH 4 mixture,
a high surface area M0 2C is formed. The oxide is first heated to a temperature
75 K below the temperature at which reduction of the oxide and reaction with
CH 4 occur. The temperature was then raised at a rate of 0.1- 5 K/min until the
reaction ceased (1041 to 1273 K). The quenched sample (M0 2C) had a surface
area of 51 m 2/g, which was higher than the sample (13 m2/g) made by the usual
method. The activity of the M0 2C was higher than 5% Ru/ Al z0 3 , but was lower
than Fe- Al z0 3-K 20 catalyst [125, 126]. Oxycarbonitride of molybdenum
(MoxCyN z ) was also active for ammonia synthesis [127]. The rate of ammonia
synthesis on reduced molybdenum (Mo), its carbide (MozC), and the oxycarbide
(MoOxCy) was measured with simultaneous gravimetric determination of the
nitrogen uptake of the catalysts. On prereduced M0 2C and MoOxCy the activity
increased with time as reduction proceeded and reached a constant value when
the catalyst had taken up about one monolayer of nitrogen, corresponding to
a surface stoichiometry of MoN. On Mo the steady state was reached after
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 123
taking up the same monolayer of MoN plus two layers of M0 2N. In all three
cases the steady-state rates were found to be very similar. The small amounts of
nitrogen taken up to reach similar rates suggested that the bulk plays no role in
determining the catalytic activity [128].
With respect to ammonia synthesis at 773 K, tungsten carbide (WC, 2m 2/g)
also seems to behave like tungsten metal (W), although their physical properties
differ considerably. 390 Pa of ammonia was produced at a contact time of 0.10 s
under N2 + 3H 2 = 101 kPa and at 773 K. H 2S adsorption on only a fraction of
the surface of WC inhibited the ammonia synthesis. This suggested that the
three side planes (l010), (1100) and (0110) of hexagonal WC crystal exposed
tungsten atoms in the uppermost layer. These faces would not hold carbon
atmos strongly and the "metallic" face was considered to be the active sites for
ammonia synthesis [129].
Atomically dispersed low-valency Mo, W, V and Zr ions can be deposited on
surface oxides with the use of n-allyl (Mo, W, and Zr) and cr-benzyl (V)
compounds and subsequent hydrogen treatment. Mo(C 3Hs)4 reacted with
surface hydroxyl ions and formed Mo(II). Such heterogenized metal ion cata-
lysts showed some ammonia activity, although the levels were low. Some
examples are listed in Table 3.7. Among four ions, Mo(II) ions were the most
active. TOF (NH3 molec./Mo ions/s) did not change when the Mo ion concen-
tration was increased from 10 17 to 10 18 ions/m 2-Si0 2. The activation energy
(25 kcal/mol) was also constant. The activity of the deposited Mo ion was shown
to depend on the nature of the oxide carrier as is shown in Table 3.7 [130]. The
reaction mechanism on transition metal ions might be different from that on
reduced metals [5].
Table 3.7. Activity of atomically dispersed transition metal ions (Me) in ammonia synthesis
N2 + 3H 2 = 5 MPa [130]
H2 ~ 2H(a) (2)
Vs = k[N2](AK[NH 3]2j[H2]3)-g/f
= k s [N 2] ([H2rj[NH 3]2)g/f (6)
The Temkin and Pyzhev equation expressed above was in agreement with
a number of kinetic measurement made on various catalysts as summarized in
Table 3.8 for the synthesis and in Table 3.9 for the decomposition. One
characteristic feature of the ammonia synthesis rate is the retardation by the
product ammonia, and this is reasonably explained by the Temkin theory. The
basic assumption of the rate-determining step, N 2 chemisorption was also
supported.
Although Eq. (4) was successful in described a number of experimental data,
the experimental values of gjf are not always the same. Sometimes it ranges from
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 125
Table 3.8. Kinetic parameters of NH3 synthesis according to the temkin-phzhev equation
VIA
Mo 720-823 0.1 0.5 178 [132]
W 855-951 0.1 0.5 190 [133]
W 673-773 0.1 0.5 (203) [129]
VII
Tc 623-698 0.1 0.5 190 ±20 [70]
VIII
Fe( - AI 2 0 3-K 2 O) 673-723 0.1 0.5 167 [131]
723-773 30 0.5 176 [134]
Fe( - A1 2 0 3) 645-651 0.1 0.5-0.6 [135]
Fe( - MgO) 570-680 0.1 0.5 (156) [136]
Ru 881-1013 0.1 0.5 249 [137]
Os 673-723 0.1 0 164 [138]
823-873 0.1 0.5 174 [138]
0.5 to 0.8 depending on temperature. This is partly natural because Eq. (4) is
valid only for an intermediate range of coverage (eN = 0.2-0.8). Thus, when
[NH3J in the gas mixture is very low or very high, the results tend to give lower
or higher value of g/f [124, 163]. Even with moderate coverage, the rate
constants given by the Temkin equation are dependent on the H2/N2 ratio,
and/or total pressure [21, 135, 164, 165]. Thus, the value of g/f obtained from
rate data can be an average value under the experimental conditions, and is not
a constant as postulated in the original theory. However, this rate expression is
widely used in commercial operations at high pressure with an Fe catalyst.
If the nitrogen adsorption is described by the Langmuir equation:
V = k [N2J (1 - e N )2, the rate of ammonia synthesis is described as follows:
(8)
This equation can be applied even under low NH3 pressure. When various
catalysts are examined below a total pressure of 1 atm, the activity (or NH3
pressure on the catalyst) is sometimes quite low. Under these conditions, the
Langmuir type equation is quite often used.
Equation (8) is based on the mechanism (1 to 3), where N(a) is the main
retarding species. If NH(a) is the main retarding species, the equation can be
written as follows:
(9)
In the case of Temkin expression, the equation is written as follows:
Vs = ks[N2J([H2J 2/[NH 3J 2)g/f (10)
......
tv
a-.
?"<:
~
::s
''""
Q..
?"<:
--l
Table 3.9. Kinetic parameters of NH3 decomposition according to the equation V = k(PNH3)m/(PHZ)" 8'"
....
'"s::
Metal Temperature (K) Pressure m n n/m E..I Ref.
(kPa) (kJ/mol)
VA
V 673-753 ca. 100 (Low P HZ ) 0.5 0 0 138 [139]
673-753 High P HZ 1.0 1.5 1.5 138 [139]
VIA
Mo 1027-1273 13.3 0 + 134-180 [140]
W 548-873 2-20 0-1 205 [141]
903-1023 4.7-20 0 146 [142]
953-1153 0.1 0 0 176 [143]
904-1214 7-26 0 0 163 [144]
1073 - 1523 2-35 0 + 146-197 [140]
950-1150 1-5 0 0 113-130 [145]
1073 10- 9 -10- 6 0.8 [146]
VIlA
Re 653-713 ca. 100 0.53 0.89 1.68 134 [147]
728-843 1.3-7 0.7 1.4 2.0 205 [72]
VIII
Fe 608-743 1-4.4 0.5 0.7 1.4 163 [148]
653-773 1.3-10.6 0.9-1.0· 1.4-1.5 1.5 176-209 [149]
765-797 ca. 100 0.9 1.5 1.67 226 [150]
Fe( - Al 2 0 3 608-703 100 0.6 0.85 1.42 192 [151]
-K 2 O) 693 13 0.48 0.72 1.5 [152]
752 13 0.75 0.38 0.5 [152J
CO 643-753 1.3-7 0.85 1.42 1.67 188 [72J
Ni 573 0.2-5.3 1.0 1.5 1.5 [153]
663-773 1.3-7 0.96 1.53 1.59 180 [72]
623-723 4xlO- 3-O.17 1 [154] w
Ru 543-738 1.3-7 1.2 2.0 1.67 188 [72] >
623-673 80--106 0.6 0.9 1.5 130 [155] 3
825-1009 1.0 1.75 1.75 247 [137]
30
::l
Rh 693-773 1.3-7 1.35 2.45 1.81 239 [72] ~.
Practically, the power rate expression is convenient. This is evident when the
mechanism is complicated.
Ru catalysts for ammonia synthesis were found to be retarded by hydrogen
adsorption. In this case both N(a) (or NH(a)) and H(a) may retard the reaction.
Vs = k[NzJ (1 - eN - eH)Z
(11)
Since this expression is complicated to use in analysis, it can be changed to
the power rate form although it is not correct algebraically.
Vs = k[NzJ(K 1 [NH3J [Hzr 1.5)-ZY (K z [H zJO. 5 )-2z
(12)
Table 3.10 shows the kinetic parameters of ammonia synthesis over Ru
catalysts. On Ru powder and Ru-alkali/support, hydrogen has a negative order
and the ammonia order is close to zero. This suggests strong hydrogen inhibi-
tion and weak nitrogen inhibition. On the other hand, it is suggested that the
nitrogen inhibition is extensive and hydrogen inhibition is weak on the support
Ru catalysts [122, 166]. The extent of ammonia inhibition also depends on the
kind of promoter as is shown in Table 3.10 [92, 167, 170].
When the synthesis is carried out far from equilibrium as happens by
increasing the space velocity or by decreasing the reaction temperature, two
steps can be the rate-determining. Under such a condition where the reverse rate
is negligible, the overall rate (Vs) in the steady state is equal to the rate of
* V = k[N2]n[H2]h[NH3J"
** High flow rate (100-1000 mlfmin) condition
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 129
and ifVs = Vn = Vh
Vs is given by eliminating eN
3.3.1.2 Decomposition
The rate equation of ammonia decomposition was expressed in Eq. (7) accord-
ing to the Temkin-Pyzhev mechanism (Eq. 1-3). The observed parameters in
Table 3.9 show that n/m values are mostly 3/2, which indicates that the
Temkin-Pyzhev mechanism is applicable on most catalysts (Re, Fe, Co, Ni, Ru
and Rh) under normal ammonia decomposition conditions.
However, the hydrogen order (n) is often zero for the reaction on Wand Pt
catalysts, for which a different mechanism should be applied [173].
Ammonia decomposition on a clean tungsten surface was studied by direct
and simultaneous measurements of the reaction at 773-1473 K under an ammo-
nia pressure of 10- 6 -10- 3 Pa [174]. The order of the reaction rate with respect
to ammonia pressure changed with temperature from first order at 1478 to 2/3
order at 773 K. No hydrogen was adsorbed on the surface in any form above
973 K and an increase in the hydrogen partial pressure (P H2 ) during the reaction
had no effect on either the reaction rate or on the amount of surface nitrogen
(eN)·
Under higher ammonia pressures, up to 100 Pa, thick surface nitride layers
were formed during the decomposition, which were decomposed at the same
rate in vacuo as in the steady state NH3 decomposition, provided that the
uptake of nitrogen was the same.
130 K. Aika and K. Tamaru
The rate of nitrogen desorption from the surface depended only on eN,
being faster with increasing eN.
(16)
whereas the rate of nitrogen uptake decreased with increasing eN and increased
with ammonia pressure (P NH3 ).
(17)
On the basis of these data, it was concluded that the overall reaction proceeds
through a reaction mechanism of "dynamic balance" between two consecutive
"rate-determining steps: 2NH3 --+ 2N(a) + 3H 2(V a ) and 2N(a) --+ N2 (V d). The
rate of the overall reaction (Vs) is obtained as follows:
(18)
(26)
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 133
nitrogen have been observed in TPD on W. The typical feature of the three types
of nitrogen on a W surface are summarized in Table 3.11. The first order kinetics
with respect to the nitrogen coverage, as well as the lower energy of desorption
observed with y and a types, suggest a molecular or dinitrogen nature of the
adsorbed site. This was confirmed by the absence of isotopic mixing in dini-
trogen desorbed from the surface on which a mixture of 28N 2 and 30N 2 was
preadsorbed [200,201]. On the other hand, the f3 type was regarded as an
atomic state on the basis of the second order desorption kinetics, the higher
energy of desorption, and the occurrence of isotopic mixing [42, 47, 201-204].
The TPD spectra were studied on single crystal surfaces ofW, and the results
are summarized on Table 3.11. As expected from the weak interaction, the y
state is found on any of three planes, W(100) [203], (110) [205,206], and (111)
[207]; and a state is only found on W(lll) as confirmed by FEM [203] or
probe-hole emission [207]; and the f3 state is predominant on (100) as demon-
strated by the large differences in sticking probability of nitrogen between
W(100) (0.3-0.6) [208-210] and W(1lO) (0.01-0.005) [203,210, 211J or W(111)
(0.08) [209,21OJ (see also Table 3.12). The desorption energies from single
crystalline planes have been measured and are shown in Table 3.11 [59]. Adams
and Germer demonstrated by LEED techniques that the f3 nitrogen atoms on
W(100) are located in an interstitial position surrounded by 4-square W atoms
as is shown in Fig. 3.9 [208,212].
The sticking probability s (rate of adsorption divided by the collision flux)
was studied as well on the clean W surface as a function of the coverage e. The
initial values of s (on a bare site) are as high as 0.3 to 0.4 on four-fold symmetry
sites such as (100) (see Table 3.12 and Fig. 3.9). This situation is contrary to the
case of Fe single crystals. Molecular beam experiments suggest that the dissocia-
tion preferably takes place at a vacant pair of (100) sites [73,210]. On W(100)
polycrystalline foil, s decreases gradually at lower coverage and then more
steeply with increasing e showing a convex curve. A constant value of s,
independent of coverage, indicates the presence of a precursor state of suffi-
Plane Reference
0.00.0
00.00
0.00.0
0000 Fig. 3.9. Surface Structure of W(l00) with
nitrogen atoms
ciently long life such that there is a high probability of transformation to the
chemisorbed state before desorption occurs. For the chemisorption of nitrogen
on W(100), the problem precursor is the y-state with a heat of adsorption of
43 kJ/mol (see Table 3.11). The value of s as a function of 8, s = aZ/(1
+ kd/kf(8)), indicates the following scheme:
N2 (g) ~Nia) kf(8) '2N(a)
kd
where a is a condensation coefficient, Z is the collision flux, kd is a rate of
desorption, k is a conversion rate constant. The experimental value of s is about
0.4 between 190 and 600 K, and the activation energy of k is 39 kJ/mol [213].
On the other hand, for N 2 on W(111), s decreases linearly with 8, indicating the
absence of a precursor state and direct molecular chemisorption (a-state) on
specific sites [73]. The adsorption and decomposition of ammonia on a clean
W(100) surface has been studied by photoemission spectroscopy (XPS and
UPS), and the results have been compared with those on c(2 x 2) -N ordered
W(lOO) [216].
Although the fresh surfaces of W have a strong affinity with N 2, the
ammonia activity is not high. The W surface is partly turned to nitride (WN)
136 K. Aika and K. Tamaru
interesting to point out that the mobility is not directly related to the equilibra-
tion activity. The activity seems, rather, to be related to the surface M-N bond
strength, that is, the heat of adsorption. Since the isotopic equilibration reaction
generally takes place through the dissociative adsorption and desorption, the
activation energy is equal to that of the adsorption step when the coverage is low
and is equal to the desorption step when the coverage is high. In the case of
nitrides, the coverage is high. Thus, the activation energies in Table 3.13
represent the activation energies of desorption, which may be comparable to the
heat of adsorption. Nitrides with low heats of activation (or low heat of
adsorption) are active for isotopic equilibration as is seen in Table 3.13.
It should be pointed out that nitrides of Sr and Ba are quite active. The
activity per unit surface may be the highest among those reported. However, the
estimated weight-base activity of SrN o.s is lower than Raney Ru and is higher
than 5% Ru-KjAC (see Fig. 3.3 and Table 3.13). The high catalytic activity ofBa
nitride is suggested as being due to the presence of various nitrogen-containing
complexes (Ba 2N, BaN 2, Ba 3N 4 ) on the Ba3N2 surface, which may give the best
route of bond expansion into the dissociation [32].
The rates of ammonia synthesis have been measured on several nitrides, and
the results are shown on Table 3.14 [87]. If there is no hydrogen effect, ammonia
synthesis rate can be estimated from the rate data of isotopic equilibration (see
Table 3.14 foot note). The experimental values are mostly lower than the
calculated values. In case of Ba, Sr, Ca, and Li, the nitride is changed to the
hydride and the activity decreases. Sr, the most active equilibration element,
a Ref. [87], LaN, PrN, and ErN have very low activities, and MgN o.6 is inactive
b Rate under 1.3 kPa of N2 at 573 K
'Estimated rate under 20kPa (150 Torr) ofN 2 at 588 K
d Ref. [236]
e Ref. [189]
f Ref. [32]
g Ref.
h Ref. [237]
; Ref.
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 139
• N2 + 3H 2 = 98 kPa [87]
b calculated as V([N 2]/[N 2]e - 1) where [N 2] = 25 kPa, [N 2]e is a ficticious N2 pressure under the
equilibrium with NH3 produced, and V is isotopic equilibration rate at [NZ]e
cPromoted Fe catalyst
turns out to be inactive, and Ba does not activate N 2 in the presence of H 2. The
activity of Ba comes only from the hydrogenation of the nitride which is only
formed by N2 without H 2. For UN1.7o, the rate of isotopic equilibration has a
reaction order of - 0.2 with respect to H2 [235]. Recalculation of the ammonia
activity using this data gives values closer to the experimental value of the
synthesis rate [87,235]. For CeN, hydrogen has no effect on the rate of isotopic
equilibration of N2 [237].
N2 is inferred to be adsorbed dissociatively on Dysprosium at 115 K,
yielding two XPS peaks in the N is region at 396.2 and 398.2 eV, corresponding
to a nitride and chemisorbed N(a), respectively. No peaks corresponding to
molecularly adsorbed N2 (400.2 eV) were observed. Upon heating the sample,
the N(a) is converted into nitride species. At a warm-up temperature of 300 K,
the N(a) species accounts for only ca. 10% of the total surface [238].
The rate of nitrogen sorption was studied at 10- 4 to 10- 6 Torr and 298 to
473 K on continuously renewed scandium films at a condensation rate of
4.8 x 10 13 atom cm - 2 S - 1. The rate of adsorption is limited by the rate of
physical adsorption [239].
(Ru-N == N). The wave numbers (cm -1) are 1910 for Ru-CsOHjMgO, 2268 for
RujMgO, 2214 for RujAI 20 3 , and 2240 for RujSi0 2 (Table 3.15). The deviation
from the gas phase N2 molecule is 421,163,117 and 111 cm - \ respectively. This
number is a measure of N2 activation, N-N bond looseness, and the extent of
electron donation by the Ru atom which is interacting with the support or the
promoter. Interestingly, the ammonia synthesis activity is ranged in this order
(Fig. 3.4). Although the life time of these species might be short at the synthesis
reaction conditions, the species might be a precursor of the dissociated state. A
mechanism of ammonia synthesis and the role of the support and promoter is
shown in Fig. 3.10.
142 K. Aika and K. Tamaru
acidic support
basic support
397.2 eV for N(a) on Fe at 290 K [269] and 397.5 eV for N(a) on promoted iron
at 673 K [270]. Surface reconstruction by nitrogen adsorption has been ob-
served on various surfaces [219]. Surface science reviews have been published
[3,4,271].
The valuable comments of Professor Emeritus A. Ozaki are acknowledged
by one of the authors.
3.4 References
1. Timm B (1984) The Ammonia Synthesis and Heterogeneous Catalysts, A Historical Review in
"Proc. 8th Intern. Congr. Catal.," DECHEMA ed., Verlag Chemie. Weinheim, Vol 1 p 7
2. Mittasch A (1951) Geschichte der Ammoniak Synthese, Verlag Chemie, Weinheim
3. Grunze M (1982) In: King DA, Woodruff DP (eds) Chern Phys Solid Surf Heterog Catal,
Elsevier, Amsterdam, Vol 4 p 143
4. ErtI G (1983) In: Anderson JR, Boudart M (eds) Catalysis, Science and Technology, Springer-
Verlag, Berlin, Vol 4 p 209
5. Shilov AE (1989) In: Bottomley F (ed) A Treatise on Dinitrogen Fixation, John Wiley & Sons,
New York, Sec. 1, p 31
6. Bottomley F (1989) In: Bottomley F (ed) A Treatise on Dinitrogen Fixation, John Wiley & Sons,
New York, Sec. 1, p 109
7. Khan F, Yue P-L, Rizzuti L, Augugliaro V, Schiavello M (1981) J Chern Soc Chern Commun
1049
8. Endoh E, Leland JK, Bard AJ (1986) J Phys Chern 90: 6223
9. Khader MM, Lichtin NN, Vurens GH, Salmeron M, Somorjai GA (1987) Langmuir 3: 303
10. Uyama H, Uchikura T, Niijima H, Matsumoto 0 (1987) Chern Lett 555
11: Kalman J, Varga TA, Hajos R (1983) Proc 6th Intern Symp Plasma Chern Vol 3: 686
12. Sugiyama K, Akazawa K, Oshima M, Miura H, Matsuda T, Nomura 0 (1986) Plasma Chern
Plasma Process 6: 179
13. Miyahara K (1983) Chern Lett 1971
14. Nielsen A (1970) Review of Ammonia Catalysis, In: Heinemann H (eds) Catalysis Rev, Marcel
Dekker, New York., Vol 4 p 1
15. Nielsen A (1977) Fert Sci Tehcnol Ser 2: 87
16. Ammoniak NH3-Bildung und Zerfall (1935) In: Gme1in Handbook of Inorganic Chemistry,
Vol 4 Nitrogen p 320
17. Mittasch A (1949) Early Studies of Mulicomponent Catalysts, In: Frannkenburg WG et al. Adv
Catalysis, Academic Press, New York, Vol 2 p 81
18. Ammonia In: (1968) Encyclopedia of Chemical Technology (R.E. Kirk, D.F. Othmer), Hark
HF et al. (eds) John Wiley, New York, p 258
19. Ammoniak, Synthetisches, In: Ullmann, Vol 3 p 544
20. Vancini CA, (1971) Synthesis of Ammonia, Macmillan, London
21. Nielsen A (1968) An Investigation on Promoted Iron Catalysts for the Synthesis of Ammonia,
Jul Giellerups Forlag, Copenhagen
22. Nielsen A (1981) Catal Rev -Sci Eng 23: 17
23. Boudart M (1981) Catal Rev -Sci Eng 23: 1
24. Jennings JR (1991) Catalytic Ammonia Synthesis, Fundamentals and Practice Plenum Press,
New York
25. Ozaki A, Aika K (1979) The Synthesis of Ammonia by Heterogeneous Catalysis, In: Hardy
RWF et. al. (eds) A treatise on Dinitrogen Fixation Sec. I and II, John Wiley, New York, p 169
26. Ozaki A, Aika K (1981) Catalytic Activation of Dinitrogen, In: Anderson JR, Boudart M (eds)
Catalysis-Science and Technology, Springer-Verlag, Berlin, Vol 1 p 87
27. Emmett PH, Love KS (1933) J Am Chern Soc 55: 4043
28. Aika K, Ozaki A (1968) Bull Chern Soc Jpn 41: 2818
144 K. Aika and K. Tamaru
29. Jolly WL (1964) The Inorganic Chemistry of Nitrogen, Benjamin, New York, p 36
30. Trapnell BMW (1953) Proc Roy Soc A218: 566
31. Bond GC (1962) Catalysis by Metals, Academic Press, London New York
32. Panov GI, Boreskov GK, Kharitonov AS, Moroz EM (1981) React Kinet Catal Lett 16: 247
33. Panov GI, Boreskov GK, Kharitonov AS (1982) Kinet Ketal 23: 438
34. Schlier E, Farnsworth HE (1950) Phys Rev 78: 316
35. Toyoshima I, Takezawa N, Suzuki H (1973) J Chern Soc Chern Commun 270
36. Toyoshima I, Takezawa N, Suzuki H (1977) Proc 6th Intern Congr Catal, Bond GC et al. (eds)
Chern Soc London, p 708
37. Scholten JJF, Zwietering P (1957) Trans Farad Soc 53: 1363
38. Beeck 0 (1950) Advan Catal 2: 151
39. Ko SM, Schmidt LD (1974) Surf Sci 42: 508
40. Beeck 0, Colle WA, Wheeler A (1950) Discuss Faraday Soc 8: 314
41. Oguri T (1964) J Phys Soc Jpn 19: 77
42. Oguri T (1963) J Phys Soc Jpn 18: 1280
43. Parry AA, Pryde JA (1967) Br J Appl Phys 18: 329
44. Kisliuk P (1959) J Chern Phys 31: 1605
45. Hickmott TW, Ehrlich G (1958) J Phys Chern Solids 5: 47
46. Hill MP, Lecchini SMA, Pethica BA (1966) Trans Farad Soc 62: 229
47. Madey TE, Yates JT Jr (1966) J Chern Phys 44: 1675
48. Yates JT Jr, Madey TE (1969) J Chern Phys 51: 334
49. Bagg J, Tomkins FC (1955) Trans Faraday Soc 51, 1071
50. Urabe K, Aika K, Ozaki A (1975) J Catal 38: 430
51. Obuchi A, Naito S, Onishi T, Tamaru K (1982) Surf Sci 122: 235
52. Mimeault VJ, Hansen RS (1966) J Phys Chern 70: 3001
53. Vajo JJ, Tsai W, Weinberg WH (1985) J Phys Chern 89: 3243
54. Miyazaki E, Yasumori I (1976) Surf Sci 55: 747
55. Sachtler WMH, Van Reijen LL (1962) J Res Inst Catal Hokkaido Univ 10: 87
56. Brenann D, Hayward DO, Trapnell BMW (1960) Proc Roy Soc A256: 81
57. Roberts MW (1960) Nature 188: 1020
58. Tanaka K, Tamaru K (1963) J Catal 2: 366
59. Toyoshima I, Somorjai GA (1979) Catal Rev -Sci Eng 19: 105
60. Frankenburg WG (1955) The Catalytic Synthesis of Ammonia from Nitrogen and Hydrogen,
.. In: Emmett PH (ed) Catalysis, Reinhold Pub, New York, 1955, Vol 3 p 171
61. Kunimori K, Kawai T, Kondow T, Onishi T, Tamaru K (1976) Surf Sic 59: 302
62. Davies PW, Lambert RM (1981) Surf Sci 110: 227
63. Kuwahara Y, Fujisawa M, Jo M, Onchi M, Nishijima M (1987) Surf Sci 180: 421, ibid 188: 490
64. Matsuo I, Nakamura J, Hirano H, Yamada T, Tanaka K, Tamaru K (1989) J Phys Chern
93: 7747
65. Mittasch A (1950) Adv Catal 2: 81
66. Stathism E (1937) Osterr Chemiker Ztg 40: 80
67. Mittasch A, Keunecke E (1931) Z Phys Chern 574
68. King DA, Sebba F (1965) J Catal 4: 253
69. Aika K, Yamaguchi J, Ozaki A (1973) Chern Lett 161
70. Spitsyn VI, Mikhailenko IE, Pokrovskaya OV (1982) Dokl Akad Nauk SSSR, 263: 656
71. Jost F (1908) Z Anorg Chern 57: 414
72. Logan SR, Kemball C (1960) Trans Farad Soc 56: 144
73. Aika K, Hori H, Ozaki A (1972) J Catal 27: 424
74. Ozaki A, Aika K, Hori H (1971) Bull Chern Soc Jpn 44: 3216
75. Urabe K, Oh-ya A, Ozaki A (1978) J Catal 54: 436
76. Ogata Y, Aika K, Onishi T (1989) Bull Chern Soc Jpn 62: 642
77. Keunecke E (1930) Z Elektochem 36: 690
78. Brill R, Osumi Y (1966) Bull Chern Soc Jpn 39: 1678
79. Zabuga V Ya, Markova GP (1965) Kataliz i Katalizatory, Akad Nauk Ukr SSR, Resp Mezhved
sb 110 (CA64:4312d)
80. Chalenko VG, Tovbin MV (1964) Ukr Khim Zh 30: 1128
81. Artyukh Yu. N, Rusov MT, Boldyreva NA (1967) Kinet Katal 8: 1319
82. Smith PI, Taylor DW, Dowden DA, Kemball C, Taylor D (1982) Appl Catal 3: 303
83. Komarov VS, Efros MD, Rabina PD, Dmitrenko LM, Rozin AT, Kuznetsov LD, Lemeshonok
GS, Khachaturyan IG, Mantseva GM, Eremenko SI (1983) SU 988327 (CA98:1l4516b)
3 Ammonia Synthesis over Non-Iron Catalysts and Related Phenomena 145
189. Boreskov GK, Kolchanova VM, Rachkovskii EE, Filimonova SN, Khasin AV (1975) Kinet
Katal 16: 1218
190. Moore GE, Unterwald FC (1968) J Chern Phys 48: 5393
191. Aika K, Ozaki A (1969) J Catal 14: 311
192. Rambeau G, Amariglio H (1981) J Catal 72: 1
193. Rambeau G, Amariglio H (1981) Appl Catal 1: 291
194. Rambeau G, Jorti A, Amariglio H (1982) Appl Catal 3: 273
195. Rambeau G, Jorti A, Amariglio H (1982) J Catal 74: 110
196. Danielson LR, Dresser MJ, Donaldson EE, Dickinson JT (1978) Surf Sci 71: 599
197. Danielson LR, Dresser MJ, Donaldson EE, Sandstrom DR (1978) Surf Sci 71: 615
198. Feulner P, Menzel D (1982) Phys Rev B25: 4295
199. Menzel D, Pfnur H, Feulner P (1983) Surf Sci 126: 374
200. Rigby LJ (1965) Can J Phys 43: 532
201. Yates JT Jr, Madey TE (1965) J Chern Phys 43: 1055
202. Joyner RW, Rickman J, Roberts MW (1974) J Chern Soc Faraday Trans 1 70: 1825
203. Delchar TA, Ehrlich G (1965) J Chern Phys 42: 2686
204. Goymour CG, King DA (1973) J Chern Soc Faraday Trans 1 69:749
205. Yates JT Jr, Klein R, Madey TE (1976) Surf Sci 58: 469
206. Fuggle JC, Menzel D (1978) Yak Tech 27: 130
207. Wilf M, Folman M (1975) Surf Sci 52: 10
208. Adams DL, Germer LH (1971) Surf Sci 26: 109
209. King DA, Wells MG (1974) Proc Roy Soc (London) Ser A 339: 245
210. Singh-Boparai SP, Bowker M, King DA (1975) Surf Sci 53: 55
211. Tamm PW, Schmidt LD (1971) Surf Sci 26: 286
212. Adams DL, Germer LH (1971) Surf Sci 27: 21
213. Clavenna LR, Schmidt LD (1970) Surf Sci 22: 365
214. Han HR, Schmidt LD (1971) J Phys Chern 75: 227
215. Tompkins FC, (1978) Chemisorption of Gases on Metals, Academic Press, London, p 26.
216. Egawa C, Naito S, Tamaru K (19!i3) Surf Sci 131: 49
217. Pasternak RA, Endow N, Bergsnov-Hansen B (1966) J Phys Chern 70: 1304
218. Mahnig M, Schmidt LD (1972) Z Phys Chern (N. F.) 80: 71
219. Kunimori K, Kawai T, Kondow T, Onishi T, Tamaru K (1976) Surf Sci 54: 525
220. Egawa C, Naito S, Tamaru K (1983) Surf Sci 125: 605
221. Haase G, Asscher M (1987) Surf Sci 191: 75
222. Grunze M, Golze M, Fuhler J, Neumann M, Schwarz E, Proc 8th Intern Congr Catal 1984,
DECHEMA (ed) Verlag Chemie, Weinheim Vol 4 p 133
223. Haase G, Asscher M (1987) Chern Phys Lett 142: 241
224. Khrizman lA, Korneyiuchuk G (1943) Acta Physicochim (URSS) 18: 420
225. Kazusaka A (1971) J Res Inst Catal Hokkaido Univ 19: 42
226. Kokes RJ, Emmett PH (1958) J Am Chern Soc 80: 2082
227. Honda F, Hirokawa K (1977) J Electron Spectrosc Relat Phenom 10: 125
228. Grunze M, Driscoll RK, Burland GN, Cornish JCL, Pritchard J (1979) Surf Sci 89: 381
229. Wilf M, Dawson PT (1976) Surf Sci 60: 561
230. Schwaha K, Bechtold E (1977) Surf Sci 66: 383
231. Kiss J, Berko A, Solymosi F (1981) Magy Kern Foly 87: 566
232. Hendrickx HACM, Hoek A, Nieuwenhuys BE (1983) Surf Sci 135: 81
233. Gorodetskii VV, Sobyanin VA (1980) Proc 7th Intern Congr on Cat ai, Seiyama T, Tanabe K,
(eds) Elsevier, Amsterdam, 1981, p. 566.
234. Kharitonov AS, Boreskov GK, Panov GI, Pankrotiev Yu D (1983) React Kinet Catal Lett 22:
309
235. Panov GI, Boreskov GK, Kharitonov AS, Moroz EM, Sobolev VI (1984) Kinet Katal 25: 123
236. Magomedkov EP, Kasatkina LA (1978) Trudy MKhTl im Meneleeva DI No 99: 60
237. Panov GI, Boreskov GK, Kharitonov AS (1980) Dokl Akad Nawk, SSSR 252: 646
238. Schreifels JA, Deffeyes JE, Neff LD, White JM (1982) J Electron Spectrosc Relat Phenom 25:
191
239. Varnakova RG, Morozova LV, Khamidova Kh Kh (1982) Zh Fiz Khim 56: 1533
240. Eischens RP, Jacknow J (1965) Proc 3rd Intern Congr Cat ai, Sachtler WHM et al (ed)
Amsterdam, North Holland Pub pp 627
148 K. Aika and K. Tarnaru
Contents
(1)
Here P NH3 and P H2 are the partial pressures of NH3 and H 2, t is the contact
time.
To rationalize his findings, Winter assumed that the steps leading to adsor-
bed nitrogen in the atomic state on the catalyst surface were so fast that the
equilibrium
NH3(gas) <=± N(ads) + 1.5 H 2(gas) (2)
Here PN, the pressure of nitrogen atoms in the gas, is a virtual pressure and
serves as an auxiliary quantity only.
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 151
Assuming low coverage and Henry's law to be valid one obtains for the
coverage eN of nitrogen far from equilibrium
eN = aPN (6)
Inserting P N from Eq. (5) one arrives at
P NH3
eN = a K p1.5 (7)
H2
with the following hydrogenation steps being fast. On this basis they determined
the activation energy from earlier data to be between 18.6 and 22 kcal/mol.
Roiter [15] proposed empirical rate equations for the synthesis rate, being
different for subatmospheric pressures and pressures between 50-200 atm. He
did not discuss the reaction mechanism.
Temkin and Pyzhev used the data of Winter as the starting point for their
development of the first successful rate equation for ammonia synthesis named
after them [16-18]. They demonstrated, that using Winter's assumptions leads
to an apparent activation energy for decomposition Ed larger than
97000 cal/mol, clearly disagreeing with the experimentally found energy of
51 000 cal/mol. Furthermore, they pointed out that the desorption of nitrogen in
the atomic form is extremely unlikely due to the large amount of energy needed.
If, however, the predominant view that the rate of ammonia decomposition
is determined by the rate of
2 N(ads) +=± N 2 (gas) (11)
is accepted, then the rate should be proportional to e~ again in conflict with the
experimental findings.
In order to overcome this difficulty Temkin and Pyzhev proposed to aban-
don the Langmuir adsorption concept, which basically assumes the surface of
the catalyst to be energetically uniform. Instead they chose to use for the
adsorption equilibrium the isotherm proposed by Frumkin and Slygin [19],
1
e = f lnaoP (12)
where e is the coverage, P the equilibrium pressure, and ao and f are constants,
and for the rate of adsorption,
(13)
an equation developed by Zeldowitsch and Roginsky [20]. Here P is the gas
pressure and Ka and g are constants. For the rate of desorption they used the
equation proposed by Langmuir [21]
(14)
where kd and h are constants.
Temkin and Pyzhev pointed out that choosing the three equations above is
equivalent to assuming a linear dependence of the heat of adsorption with the
coverage. In accordance with Polanyi [22] the change in energy of activation
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 153
{J=~ (16)
f
rx+{J=l (17)
Another assumption for the Temkin-Pyzhev equation is that the nitrogen
adsorption is not influenced by hydrogen and ammonia, an assumption which
seems to be corroborated by the experimental data of Emmett and Brunauer
[7].
Following Winter, Temkin and Pyzhev assumed that the amount of adsor-
bed nitrogen is determined by the equilibrium with hydrogen and ammonia in
the gas phase, i.e., the nitrogen equilibrium pressure is given by
In order to get Eq. (20) to agree with Winter's experimental findings, Temkin
and Pyzhev had to choose
rx = {J = 0.5
Using data from Emmett and Brunauer [7], Temkin and Pyzhev were able
to demonstrate that the activation energy for decomposition on the average
should be 46500 cal/mol, in good agreement with Winter's results.
To arrive at an expression for the synthesis rate equation, Eq. (19) is inserted
in Eq. (13) to get the net rate as the difference between the forward and reverse
processes,
dP NH3
dt
= k P
1 N2
(p~2)a
p2
_k 2
(P~H3)P
p3
(21)
NH3 H2
154 J. B. Hansen
(22)
(23)
dPNH3 = 0
dt
and therefore
(24)
kl = K2 (25)
k2 P
(26)
(27)
(28)
Here V is the space velocity, the ratio of the volume of gas at standard
conditions to the volume of catalyst per hour, and}' characterizes the deviation
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 155
k = 273¢kl (30)
0.751.ST
where ¢ is the catalyst bed void volume ratio.
Temkin and Pyzhev tested the rate equation by applying Eq. (28) to a series
of experiments at atmospheric pressure with a doubly promoted catalyst, first by
varying the H1iNl ratio at 400°C using a space velocity of 30000h- 1 . As
predicted by Eq. (22), the H1iNz ratio giving the highest yield at low conversion
was 1.5.
A second set of experiments was carried out at 400 and 450°C. They showed
a satisfactory constancy of k at each temperature as expected if Eq. (28) were
correct. The apparent activation energy for the decomposition rate constant kl
was calculated to be 40000 cal/mol.
The decomposition data of Chrisman [13] were evaluated with Eq. (20), and
an activation energy of 39700 cal/mol was found.
Temkin and Pyzhev derived the following integrated form of their rate
equation without making any simplifying assumptions, except assuming H 2iNl
to be 3.
(31)
(32)
(33)
Temkin and Pyzhev made an error in deriving Eq. (31), as pointed out first by
Emmett and Kummer [23 J. Emmett and Kummer introduced a factor of
(1 + Z)3 instead.
The correct first denominator term, (1 + Z)l was first worked out by
Kodama et al. [24]. Temkin also arrived at (1 + Z)2 in [25]. The error introduc-
ed by a wrong denominator is normally small as Z is small.
Temkin and Pyzhev used Eq. (31) to test their rate equation with the data of
Larson and Tour [2] at 10, 31.6 and 200 atm and at 420 and 450°C. They
concluded that the data agreed well with Eq. (21), however, the rate constants
156 J. B. Hansen
calculated for the 200 atm data were lower than the ones calculated for the lower
pressures. Temkin and Pyzhev attributed this difference to the fact that the
measurements at high pressures were carried out with different equipment than
those at low pressures. Temkin and Pyzhev later pointed out that at pressures
above 300 atm fugacities should be used instead of pressures and the constants
k2 and kl should be assumed to depend on the pressure.
By differentiating Eq. (21) with respect to temperature, the optimum syn-
thesis conditions (maximum rate) is found at a temperature at which the
equilibrium constant is equal to
P NH3
p O. 5 p1.5 =
~d
If (34)
Nz H2 S
where Ed and Es are the apparent activation energies for decomposition and
synthesis, respectively.
Love and Emmett [26] studied the decomposition of ammonian over both
singly (10.2% A1 20 3) and doubly promoted (10.2% Al 20 3 1.59% K 20) catalysts
at atmospheric pressure.
The doubly promoted catalyst showed the rate equation
dP NH3 _ k P~it (35)
-~- p O. 85
H2
which is very close to Temkin and Pyzhev's Eq. (20) if f3 is set at 0.29. The
apparent activation energy for decomposition was found to be
45600 ± 2000 cal/mol.
The data obtained on singly promoted catalysts, however, could not to be
fitted to the Temkin-Pyzhev equation. Their performance depended in a highly
complex manner on the temperature and the partial pressures and the activation
energy depended on the temperature and even displayed hysteresis effects. These
deviations from Temkin-Pyzhev kinetics could be removed by impregnating the
catalyst with a KOH solution although the activity dropped considerably.
The decomposition rate found was approximately independent of the am-
monia and hydrogen partial pressures at 390 and 450°C. At 429 DC, it was
proportional to P~·22 /P~J: i.e., showing a negative f3 value in the Temkin-
Pyzhevequation.
Love and Emmett suggested that this abnormal behaviour could be due to
the presence of substantial amounts of.NH and NH2 groups on the surface on
singly promoted catalysts, whereas on doubly promoted catalysts, such groups
would be "driven off" by the alkali.
This interpretation was supported in some chemisorption experiments with
H2 at 100°C on singly and doubly promoted catalysts previously exposed to N2
at 450°C (Brunauer and Emmett [27]).
Frankenburger later explained the phenomena by NH, NH2 and NH3 being
adsorbed on the acidic Al 20 3 groups in singly promoted catalysts. These acidic
functions could be neutralized by K 20 [28].
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 157
Brunauer et al. [29J extended the range of the Eqs. (12), (13), (14), which are the
basis for the Temkin-Pyzhev equation. In order to develop an adsorption
isotherm, Brunauer et al. subdivided the surface into elements each following
(the original) Langmuir isotherm, so that
ad1 - 8)p = b 1 8e- q / RT (36)
where q is the heat of adsorption, and ai and b i are constants. Assuming q to be
linear function of the surface area covered,
q = qo - As (37)
where qo is the heat of adsorption at 8 = 0, one finds
a e-As/RTp ds
8-
-
J_o
1
0
_ _~o=-
1 + aoe-BS/RTp
(38)
Integration gives
8 = RT In 1 + aoP (39)
A 1 + aoe- a/RT
where
a i qo RT
ao = _e / (40)
bi
ifao»l»aoe-A/RT, then Eq. (39) gives the Frumkin and Slygin isotherm (Eq.
(12)) used by Temkin and Pyzhev, if
1 RT
(41)
f A
The Temkin-Pyzhev equation is thus only valid at medium surface coverage as
already pointed out by the authors.
Brunauer et al. also developed equations for the rates of desorption and
adsorption using as starting points the equations proposed by Taylor [5J,
w = b8e - Ed/RT (42)
z = aP(1 - 8)e- E ./ RT (43)
where wand z are rates of desorption and adsorption, respectively, and band
a are constants.
They assumed the activation energies to vary linearly with coverage:
Ed = E~ - Bs (44)
Ea=E~-Js (45)
and, furthermore, divided the surface into two parts, an almost completely
158 J. B. Hansen
e
Equation (46) gives w = 0 for = 0 and the maximum value of w for = 1. e
e e
Equation (47) gives z = 0 for = 1 and a maximum value of z for = O. For
e,
large values of Eq. (46) reduces to the Langmuir desorption rate equation (14)
e,
and for small values of Eq. (47) gives the Zeldotwitsch equation with
h = B/RT (48)
g = J/RT (49)
Brunauer et al. also deduced rates for adsorption and desorption if the surface is
homogeneous, but strong attractive and repulsive forces exist between the
adsorbed molecules or atoms. They assumed the activation energy to be linear
with coverage
w = be-E~/RT e B8 /RT (50)
z = aPe-E~/RT(l - e)e- J8 /RT (51)
and found the adsorption isotherm formula by equating Eqs. (50) and (51)
At intermediate surface coverage Brunauer et al. used Eqs. (46) and (47)
(which then reduces to the Temkin Eqs. (13) and (14)) to calculate the adsorption
rate of nitrogen
in good agreement with what Love and Emmett had with the same catalyst
referred to before. An activation energy of 49000 cal/mol could also be esti-
mated in good agreement with the experimental value.
In 1943, Emmett and Kummer [23] presented the results of high pressure
experiments on a doubly promoted catalyst (3.02% Al 2 0 3 , 0.94% K 2 0) at 33.3,
66.6 and 100 atm, H2/N2 ratios: 3/1,1/1 and 1/3 and space velocities from 25 000
to 125000 h -1 at 370, 400 and 450°C were used. The data were analysed with
the Temkin-Pyzhev equation using (J. = f3 = 0.5. The rate constant k was con-
stant with variations in space velocity, except at 370°C where it decreased with
an increase in space velocity of 5. Apparent activation energy for decomposition
was found to be from 45000 to 53000 cal/mol.
Although the Temkin-Pyzhev equation seemed to fit the data reasonably
well with respect to variations in space velocity and temperature, all data sets
exhibited a clear decrease in the rate constant with increasing pressure.
Using a f3 value of approx. 0.3 as found by Love and Emmett [26] an
attempt was made to calculate the data by the Temkin-Pyzhev equation with
(J. = 0.67, but this gave a poorer agreement. With respect to changes in gas
composition the agreement with (J. = 0.67 was fair at 370 and 400°C, but the rate
constant varied by a factor 2 at 450°C between the 3: 1 and 1: 3 H2/N2 mixtures.
Emmett [30] presented a review on the ammonia decomposition reaction in
1946.
In 1947, Temkin and Kiperman [31] gave a general discussion of the
Temkin-Pyzhev equation and its usefulness for both the synthesis reaction and
the decomposition. They pointed out that it would not be valid at low ammonia
partial pressures because the equilibrium
(59)
would be displaced to the right leading to low surface coverages, where the
Temkin-Pyzhev equation is no longer valid. The rate r should be independent of
coverage and given by
(60)
Sidorov and Livshits [32] carried out experiments at 300 atm in the temperature
160 1. B. Hansen
range 450-500°C. They obtained a good fit of the data by using the Tem-
kin-Pyzhev equation with CI. = 0.5 and an activation energy of 176 kJ/mol. They
still used (1 + Z)3 in the denominator in the integration of the Temkin-Pyzhev
equation introduced by Emmett and Kummer [23].
Some data on the synthesis rate were given by Kobayashi and Kubota in
[33].
The less-than-satisfactory agreement with experimental data at higher pres-
sures ofthe Temkin-Pyzhev rate equation has been mentioned. In 1950, Temkin
modified the original equation by introducing fugacities instead of partial
pressures [34]. Furthermore, the original equation, Eq. (21), was multiplied by
a term
as the gas. The variance between the actual and calculated ammonia outlet
concentrations was between 0 and -7%. The accuracy of k2 was estimated
to be 10% for the adiabatic reactor and 16% for the quench reactor. The tem-
perature covered the range 374-555°C and the efficiencies the range 0.09 to
0.96. The results in the form of k2 as a function of temperature were plotted as
Arrhenius plot. Although the experimental scatter was considerable
(around 30%), the Temkin-Pyzhev equation simulated the plant performance
well.
Activation energies for decomposition reaction of 45800 cal/mol and
47900 cal/mol were found at 245 atm and 300 atm, respectively, in good agree-
ment with Emmett and Kummer's results [23]. A trend in k2 with pressure could
be observed. Whereas Emmett and Kummer found that k2 was proportional to
p-O.6\ Annabel found k2 to be proportional to p-o.s.
The Temkin-Pyzhev equation was also used to analyze the performance of
a catalyst charge with a feed gas containing larger amount of poisons than the
experiments above. The activity was only one third of the activity of a catalyst
operating with a relatively pure gas. A bend in the Arrhenius plot could be
observed around 450°C. The Temkin-Pyzhev equation with the as-found con-
stants was eventually used to find the optimum temperature curve in an ideal as
well as a practical reactor.
Kiperman and Granovskaya [39] studied the kinetics at low ammonia
partial pressures and efficiencies below 1%. They found the rate to be propor-
tional to the nitrogen partial pressure in agreement with the Temkin equation
(Eq. (60)).
Sidorov and Livshits used the special high-pressure version of the Temkin
equation (Eq. (61)) to study the kinetics on doubly promoted catalyst at 500°C
[40]. They found the rate constant to be independent of pressure between 10 and
400 atm with 0( = 0.5. The total pressure range studied was 10-500 atm.
Adams and Comings made an extensive study on the ammonia synthesis
rate on an industrial ammonia catalyst (2.84% Ah03, 1.04% K 20, 0.26% Si0 2)
[41]. They used an isothermal reactor (within 1 to 3°C) holding 1 ml of catalyst,
in a tubular catalyst chamber measuring 0.95 cm (inside diameter) by 1.25 cm
(length).
A total of 300 runs were made at 350,400,450,475 and 500°C and 100, 200
and 300 atm. The H2/N2 ratios studied were 3/1, 1/1 and 1/3 and the space
velocities ranged from 10000 to 125000h- 1 .
They used the original Temkin-Pyzhev equation with 0( = 0.5 to correlate
the data, but agreement between theory and experiment was even less satisfac-
tory than that obtained by Emmett and Kummer [23].
Partial pressures were replaced with fugacities calculated according to Lewis
and Randall rules, and the data sets were recalculated. The rate constant k' still
showed considerable variations with operating conditions. Below 400 °C it
decreased with space velocity and above 400 °C it increased. It decreased with
both pressure and H2/N2 ratio. In view of this, Adam and Comings resorted to
a graphical approach to reactor design.
162 J. B. Hansen
Bokhoven et al. later pointed out that the experimental data might have
been influenced by back diffusion in the reactor due to the shallow bed used
[42].
Frankenburger in a comprehensive review article [28] on the ammonia
synthesis extensively discussed the physical meaning of ex and f3 in the Tem-
kin-Pyzhevequation.
Furthermore, he discussed the unresolved problems in interpreting the
kinetics of ammonia synthesis. In principle, it should be possible by theoretical
means to calculate, for instance, the rate constant k t . Emmett and Brunauer [7]
had already pointed out that only approx. lout of 106 N2 molecules striking the
catalyst surface and possessing the necessary activation energy are actually
chemisorbed on the surface. Frankenburger pointed out that such calculations
ignore the entropy changes associated with the chemisorption step. The entropy
effects are linked to the probability that a molecule will actually be chemisorbed.
The entropy of the chemisorbed state is mostly lower than that of the gaseous
state.
Frankenburger suggested a two-step mechanism for N2 adsorption involv-
ing a short lived "physical" adsorption on the surface followed by nitrogen
dissociation. Another reason for the heterogeneous behaviour of the ammonia
catalyst surface, in addition to the two already mentioned, was suggested by
Frankenburger: "previously adsorbed particles influence the entire catalyst and
particularly its surface in such a way that it exerts forces of attraction or
repulsion toward N 2 molecules from the gas phase different from those exerted
by a completely bare surface."
Bokhoven et al. in a review article [42] critically discussed the application of
lhe Temkin-Pyzhev equation to fit the data published before 1955. They quote
their own unpublished results of synthesis experiments at 1 atm on a doubly
promoted catalyst at 350°C with H2/N2 = 3.0. The best fit is obtained with
ex = 0.6.
Concerning high pressure experiments, Bokhoven does not discuss the data
of Larson and Tour [1], because these were influenced by diffusion effects. The
data of Adams and Comings [41] were not used either because of the possible
influence of back diffusion in these experiments.
Bokhoven et al. used data from Emmett and Kummer [23], Sidorov and
Livshits [32], [42] and Nielsen [35]. They pointed out that the data of Emmett
and Kummer are restricted to a narrow efficiency range. Regarding the data of
Nielsen, they only compare those at an intermediate space velocity, because
non-isothermal conditions could be more or less pronounced with higher or
lower space velocity.
The calculated rate constants from the data of Nielsen and Sidorov and
Livshits are influenced by the H2/N2 ratio and are decreasing with pressure; in
the latter case there is only a small decrease, however, k2 is independent of
efficiency.
Bokhoven et al. cited some of their own, unpublished results obtained with
a differential reactor at 350°C whereby the exponents of the partial pressures
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 163
dP P O.9-1.0
-~=k NH3 (63)
dt p 1 .4-1.5
H2
i.e., f3 ~ 0.25 in the Temkin-Pyzhev equation. The activation energy was found
to be 38.8 kcal/mol.
Mills and Bennett made an extensive kinetic study at 100-1000 atm and 400
and 450°C on doubly promoted catalysts [48]. H2/N2 ratios of 3/1, 1/1 and 2/1
were used and the space velocities ranged from 2000 to 230000 h - 1. The reactor
used held 1.625 g or 0.602 cm 3 of catalyst and had the dimensions 4.76 cm in
diameter and 5 cm long. The maximum temperature difference recorded was
3°C and in most cases it was less than 1 °C. Approximate calculations demon-
strated that the experimental data should be free of axial or radial dispersion of
heat or mass. The size of the catalyst used was 2 mm diameter, and intraparticle
diffusion effects were thought to be absent. The lowest effectiveness factor
estimated was 0.88. Reduction of the catalyst was done according to the
procedures of Nielsen [35].
The experimental data were presented in the form of curves showing the
space time yield (cm 3 NH3 produced per cm 3 per hour) as a function of space
velocity. All data were evaluated with the Temkin-Pyzhev equation modified for
high pressure (Eq. (61)) according to Temkin [34]. The ideal solution method
was selected for calculating the fugacities.
The Temkin-Pyzhev equation in form of Eq. (61) was found to describe the
variations with space velocity not too close to equilibrium fairly well, but k2 was
clearly decreasing with pressure. To fit the data, the partial molar volumes of
nitrogen would have to be in the order of 200 or 500 cm 3/mol instead of the
more probable 20-50 cm 3 /mol. The activation energies for decomposition were
calculated at the different pressure levels.
100 36000
400 44000
1000 53000
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 165
K(~~~;)
8= H2 (67)
1+ K(~~7)
and the rate then becomes
(68)
r = 1 + K (P~H3) (70)
p1.5
H2
Kubota and Shindo [49] carried out the analysis above and reevaluated the
data given in Refs. [23, 32, 33, 41,50,51]. They used fugacities instead of partial
166 1. B. Hansen
pressures and included terms for the reverse reaction. The data were well fitted.
The results obtained on doubly promoted catalysts were best fitted by Eq. (68),
whereas Eq. (69) worked best for singly promoted catalysts. The following
activation energies were found:
Kubota and Shindo pointed out that the data could also be correlated by
Eq. (70); thus, kinetics alone cannot be proof of a mechanism. Boudart [52] as
early as 1956, had noted that it should be possible to describe the ammonia
synthesis kinetics by the Langmuir isotherm. This was also discussed by Stelling
and von Krusenstierna [53].
Both Temkin [54] and Kwan [55] showed that assuming a simple Freun-
dlich isotherm can also lead to the Temkin-Pyzhev equation.
Ozaki et al. [56] carried out experiments at and below atmospheric pressure
with two types of doubly and triply promoted catalysts (A = 0.85% Al z0 3 ,
0.27% KzO and B = 2.6% Al z0 3 , 1.6 KzO and 1.6% SiOz). Amounts of ap-
prox. 0.8 g were used. Reduction was carried out in pure H2 for 100 hours at
400°C. The experiments were carried out at 218, 251, 278, and 302°C and the
pressure was kept at 1/3, 1/2 and 1 atm. Stoichiometric gas was obtained by
cracking NH 3 and ND 3'
The experimental data were first fitted to the Temkin-Pyzhev equation. It
was noted that the value of IX decreased from 0.8 to 0.4 with increasing efficiency.
This can be attributed to low coverage by nitrogen, thus making the assump-
tions for the Temkin-Pyzhev equations invalid as noted by Brunauer [27] and
also Temkin [17], [31]. At a fixed temperature and at relative high efficiencies,
k still depended on pressure. Ozaki et al. correlated the data with Eq. (68) and
found k independent of pressure but K dependent on pressure. A successful fit,
however, was obtained using Eq. (69) with both k' and K~ independent of
pressure.
The isotope effect KH/Ko also was different from the one calculated from
partition functions if Eq. (68) was used. Good agreement between the theoretical
and experimental isotope effect was obtained with Eq. (69). The rate constants
k are identical for H2 and D2 strongly indicating that nitrogen adsorption is the
rate determining step.
Ozaki et al. also extended the Temkin-Pyzhev equation by using the iso-
therm developed by Brunauer et al. [27].
(71)
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 167
. P
and noting that if 1 »e-o.sfK o P~~;
H2
then
e = ~ In {I + Ko (~t:; )} (72)
Using the Zeldowitsch equation (Eq. (13)), one arrives at the rate equation
r = {I + Ko r
(~t::) a
(73)
.h g
WIt 0: = f.
The different possible rate equations with the underlying assumptions are
summarized for the main surface species as follows:
(74) (75)
(76) (77)
(78) (79)
For Eqs. (74) and (75) a uniform surface is assumed, whereas for Eqs. (76) and
(77) a heterogeneous surface according to Brunauer is considered. Equations
(78) and (79) are associated with a heterogeneous surface with the Temkin-
Pyzhev approximation. Ozaki et al. pointed out that, at sufficiently high efficien-
cies, Eqs. (76) and (77) reduce to Eqs. (78) and (79). If 0: = 1, Eqs. (76) and (77) are
identical with Eqs. (74) and (75). It is noted that Eqs. (78) and (79) can be
regarded as approximations to Eqs. (74) and (75), in which case IY. will be without
any physical meaning.
As to the reason for the presence of NH instead of N on surface, Ozaki et al.
pointed out that the catalysts may not have been completely reduced. Since N. ds
is more thermodynamically stable than NH. ds the reason for the dominance of
NH. ds must be a kinetic one.
Brill and Tauster [57] carried out experiments on doubly promoted cata-
lysts. No experimental details about catalysts or reduction procedures were
given. The constant 0: in the Temkin-Pyzhev equation was increased from 0.11
to 0.7 for doubly promoted catalyst between 182 and 369°C and from 0.47 to 0.7
for singly promoted catalyst between 242 and 286 0c. There was no change in
168 1. B. Hansen
ex with efficiency. Brill pointed out that the change in (J. observed by Ozaki et al.
[56J could be due to temperature changes rather than efficiency changes. He
also noted that it is impossible to calculate the activation energy from the
Temkin-Pyzhev equation, when (J. is changing with temperature, because k is
a function of temperature according to
k=kt- a (k 2K p )a e -E/RT (80)
Using a Langmuir equation Eq. (74), i.e. Nads being the most abundant
reaction intermediate, Brill obtained a good fit to the data and calculated the
activation energy on the doubly promoted catalyst to be 17.4 kcal/mol.
Krabetz and Peters [58] used both the Temkin-Pyzhev equation and the
Langmuir formulation to test data from synthesis experiments on singly and
doubly promoted catalysts at atmospheric pressure and temperatures between
280 and 370 DC. For the doubly promoted catalyst, the Temkin-Pyzhev equation
fitted the data well if (J. = 0.8 at 333 DC and (J. = 0.7 at 372 DC. The Langmuir
equation (Eq. (80)) with N as most abundant reaction intermediate, also gave
a satisfactory fit yielding an activation energy of 84-105 kJ/mol. For the singly
promoted catalyst the best fit was obtained with the Ozaki-Taylor-Boudart
equation Eq. (75), i.e. NH as most abundant reaction intermediate. The activa-
tion energy was 59 kJ/mol.
Nielsen et al. [59J carried out an extensive kinetic study on a commercial
triply promoted KMIR (K 20, CaO, A1 20 3 ) catalyst. It had been prereduced in
the size range 3-6 mm but was tested in the size range 0.3-0.7 mm. The sample
was reduced again at 150 atm up to 400 DC after which the pressure was
increased to 300 atm and the temperature to 480 DC. The reactor had an internal
. diameter of 5 mm and was equipped with three thermocouples of 1 mm in outer
diameter in the catalyst bed. Total catalyst volume was 2.5 cm 3 .
The operating conditions studied were as follows:
Pressure: 149-309 atm abs
Temperature: 330-495 DC
Space velocity: 13200-105600 vOl/h
H2/N2 ratio: 6.23-1.15
The inlet gas did not contain any inerts or ammonia. A total of 35 runs at
different conditions were carried out. Throughout the experiment the stability of
the catalyst activity was checked by measurements at standard conditions. The
influence of different departures from ideal conditions on the reliability of the
measurements is discussed in the article. In most of the experiments, the
temperature was held within 5 DC, but in runs at high pressures and space
velocities, temperature differences up to 17 DC were observed. A weighed mean
temperature was used in the analysis of the experiments. Computer calculations
using a stepwise kinetic integration, simulating the actual temperature vari-
ations, showed that using such a weighed average temperature as the assumed
isothermal operating temperature removed any error from the runs where the
temperatures differences were below 5 DC and the difference between calculated
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 169
and measured ammonia outlet percentages were small even with the 17 °c
temperature difference.
The effect of diffusion was also investigated by means of computer calcu-
lations. A run at rather extreme conditions, 450°C, 317 atm, and space velo-
city = 32100 vol/h, had an effectiveness factor of 0.5 and 0.8 at 10 and 20% into
the catalyst bed respectively. In the main part of the bed, the effectiveness factor
was very close to 1.0 as in the majority of the runs. Flow conditions in the
experiment were in the viscous region or in the transition region between
viscous and turbulent flow. The non-ideal conditions due to flow and wall effects
were believed to have minor effects. The fact that the catalyst was reduced in the
3-6 mm size range affect the intrinsic activity due to the self-poisoning by water
diffusing out of the pores during the reduction process.
Nielsen et al. analyzed their data using the rate Eqs. ( 74), (75), (76) and (77)
developed by Ozaki et al. [56]. They combined the four equations into
k 1 PNz
(81)
K2 = aN
2
H3,eq (82)
a aHz,eq
3 a Nz,eq
and taking the reciprocal reaction (ammonia decomposition) into account and
substituting activities for partial pressures, they arrived at:
r= k~{aNZK; - (~)}
{1 + K3 (a:;'3 ) f"
(83)
Expressing
k~ = k20 e - Ez/RT (84)
K3 = K30e-E3/RT (85)
it can be seen that Eq. (83) contains six unknown coefficients k20' E 2, K30' E 3,
wand a. The best values for the constants were found by computer calculations
to be
K 30 = 3.07 X 10- 2 E3 = - 19361 cal/mol
lk20 = 1.06 X 1013 E2 = 17429 cal/mol
w = 1.564 a = 0.640
1 A printing error in [59] giving k 20 as 2.12 x 10 13 or twice the correct value was later corrected by
Nielsen in [60].
170 J. B. Hansen
Equation (83) fitted the data at 370, 410, 450 and 490°C quite well when
these values are used, whereas it predicted too high reaction rates at 330 °C
compared to the measured ones. In this connection it should be remembered
that even minute amounts of oxygen-containing compounds will severely retard
the reaction rate, especially at low temperatures, as will be discussed later.
The authors pointed out that the second term in the denominator bracket far
exceeds 1 under most circumstances. In fact, only when the ammonia content is
down to around 0.1 %, is it comparable to 1. This means that Eq. (83) is reduced
to the original Temkin-Pyzhev equation. For the Temkin-Pyzhev equation,
Nielsen et al. derived an apparent activation energy of 42300 kcaljkmol with
IX = 0.64. The rate constant was found to be independent of pressure, which was
ascribed to the use of fugacities instead of partial pressures.
Temkin et al. [61] studied the rate of ammonia synthesis far from and near
the equilibrium on a doubly promoted catalyst at and below atmospheric
pressure. The catalyst, of 0.1 to 0.2 mm in size, was reduced in the synthesis gas
at a space velocity of 100000 h - 1 and a final temperature of 550°C. Batches of
0.02 to 0.05 g catalyst were used in the experiments. In one series of experiments
the partial pressure of nitrogen was fixed at 100 mm Hg and the hydrogen
pressure varied from 100 to 600 mm Hg. In another series the opposite was
done. Temperatures were 450°C and 350 °C in both series and the space
velocities used ranged from 138000 h - 1 to 2640000 h - 1, which ensured an
ammonia percentage less than 0.01 mol %, i.e. far from equilibrium.
The data could be reproduced with a kinetic equation of the form
r = k' p Hz
O. 5 p O. 5
Nz (86)
Another series of experiments were carried out at 350 and 450°C at atmospheric
pressure, but varying the H2/N2 ratio from 8.5 to 0.15. The data from this
experiment could also be fitted to Eq. (86). The apparent activation energy for
k was found to be 11.5 kcaljmol.
The authors interpreted the applicability of Eq. (92) as an indication that, at
low ammonia concentrations, the equilibrium between adsorbed nitrogen and.
ammonia plus hydrogen in the gas phase is no longer established.
Another set of experiments was carried out with stoichiometric gas near or at
an intermediate distance from equilibrium. The pressure was fixed at 1 atm and
the temperature varied in steps of 25°C from 325 to 550°C. The space velocity
ranged from 61000 h -1 to 542000 h -1.
The original Temkin-Pyzhev equation (Eq. (21)) with IX = 0.5 reproduced the
data above 400°C reasonably well. The activation energy for k2 was found to be
constant at 37.0 kcaljmol above 400°C. This corresponds to an activation
energy for k1 in Eq. (21) of 12.0 kcaljmol which is close to the 11.5 kcaljmol
found for k' in Eq. (86). At the temperatures of 325 and 350°C, k2 in Eq. (22)
declined at lower ammonia concentrations indicating that Eq. (22) is no longer
valid because the conditions are far from equilibrium.
In a later communication [62], Temkin et al. developed a rate equation
covering the complete range of operating conditions including the transition
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 171
region between Eq. (21) and Eq. (86). They assumed an associative mechanism
for the synthesis as follows:
"'1
"'2_
Step 2a N 2* + H2 <=± N 2H2* (88)
"'-1
0p 0 X-2. P~H3
op
1 II X2. H2 - X-I KPl
XI N2
W =____ 2b H2 (90)
f sin exII (op X-2. p~H3)a(,0 0 p )l-a
Xl N2+K -p2 X-l+X2. H2
2b H2
where
f=~
RT
and X?, X?. I and xg., X?. 2. are the rate constants at zero surface coverage for the
reactions described by step 1 and 2a. These are related to the equilibria
constants
o
XI
K oI _- -o- (91)
X-I
o
K O _ X2. (92)
2. - 0
X-2.
and
K?Kg.K 2b = K (93)
Introducing the substitutions
(94)
and
X?. I
1 =-0- (95)
X2.
172 1. B. Hansen
k P 1- a (1 _
± Nz
~ P~H3
K P p3
)
N2 Hz (96)
ill = (_1 + ~ P~H3 )a(_1 + 1)1 a
P Hz K P Nz P~z P H2
It can be seen that
(97)
(98)
which obviously requires that the hydrogen partial pressure is not negligible.
Close to equilibrium, the term
will be close to 1, and Eq. (96) approaches the normal Temkin-Pyzhev equation
(Eq. (21)). Ifthe system is very far from equilibrium, then
1 P~H 1
---~-« (99)
K PNzP Hz P Hz
Equation (96) then converts into Eq. (86) deduced in Ref. [61]. Calculations
showed, however, that under industrial ammonia synthesis conditions the orig-
inal Temkin-Pyzhev equation is invariably applicable. However, it was neces-
sary to apply Eq. (96) in order to get constant values of Kad at 350°C and 325 °C
for the data reported in [61]. Additional data are given at 450°C in [62]. When
the total pressure was varied from 0.5-1.0 atm, the hydrogen/nitrogen ratio
from 0.5-5, and the space velocity from 200000 to 526000 h - \ the relative
yields Z/Zeq were in the range 0.13 to 0.35. The calculated L values were
satisfactorily constant, although a tendency towards lower k+ values with
higher yields could be observed.
Bridger and Snowdon [63] cited tests from both differential and integral
reactors without giving any experimental details. They found the extended
Temkin-Pyzhev equation, Eq. (96), to describe the experimental data best with
a = 0.465 and E2 = 40.4 kcaljmol.
Aika and Ozaki [97] did experiments on 10.4 g of unpromoted catalyst at
305°C at and below atmospheric pressure. The catalyst had been reduced at
400°C for 200 h and stabilized for approx. two weeks. They tested three kinetic
equations for the best fit to the experimental data.
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 173
(100)
(101)
where Ya and y denote the measured and calculated NH3 mole fractions
respectively. Ozaki selected the best linear log-log plot ofYa vs l/v, where v is the
flow rate.
However, Logan and Philip did not as Ozaki et al. limit the value of g/f = IX
to 1.0 and indeed found a best fit with IX around 0.8. Furthermore, they found
a more normal isotope effect of Kh/Kd = 3.8 at room temperature consistent
with N being the most abundant species on the catalyst surface and adsorbed
H participating in the slow step of nitrogen dissociation.
Based on certain circumstantial evidence, Carra and Ugo [66] proposed
a mechanism where the rate-determining step is the hydrogenation of a half-
hydrogenated surface nitrogen, dehydroimide (N2H). This leads to the following
rate equation using the Langmuir adsorption isotherm
KP N2 (b h p H ,)O.5
r = 1 + (b P )05 (102)
H H2 · + b NH3 + PNH3
(104)
V = k2 ( p[; )0.75
p
(105)
(106)
(107)
with
(108)
(i.e. assuming a uniform surface with N adsorbed as the most abundant reaction
intermediate) to experimental data at low temperatures (approx. 265-340oq,
presumably atmospheric pressure, and a stoichiometric gas mixture using pure
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 175
with
cK (Pp1.5
-1 NH3 )
8 = (110)
1+CK- (P
Hz
1 NH3 )
p1.S
H3
2k = 1.7698 x 10lse-4o:T65
Dyson and Simon found, as Nielsen et al. did, the rate constant k to be
independent of pressure.
176 J. B. Hansen
Ferraris and Donati [73J reevaluated the data of Nielsen et al. using both the
Temkin equation Eq. (112) modified by Dyson and Simon and Eq. (83) used by
Nielsen et al. They used the same thermodynamic data as Dyson and Simon.
A non-linear parameter estimation program was used to calculate the constants
in the two equations minimizing the sum of squares defined by:
N
SSQ = L: (Zeale - Zmeas)2 (113)
i= 1
where Zeale and Zmeas are the calculated and measured ammonia mole fractions.
For the Dyson and Simon equation, Ferraris and Donati found
SSQ = 9.30 x lO-3 with oc = 0.5. The best fit was obtained with oc = 0.73 and
2k = 1.30 x 10 15 e -4~:30. This gave SSQ = 6.33 x 10- 3.
Using the values of Nielsen et al. Ferraris and Donati found
SSQ = 6.87 x lO-3. Reestimation of the parameters gave:
K30 = 2.1496 E3 = - 18458 caljmol
K 20 = 1.349 X 1015 E2 = 16480 caljmol
W = 1.574 oc = 0.6923
The SSQ became 6.26 x lO - 3.
In a later article [74J, Ferraris et al. again used the data by Nielsen et al. to
test 23 different rate equations for ammonia synthesis. Some of the rate equa-
tions considered both associative and dissociative mechanisms. Others were
purely empirical. They concluded that a large number of the models represented
the data equally well, in fact better than the Temkin-Pyzhev equation. This
again proves that it is extremely difficult, and in fact usually impossible, to draw
definite conclusions about the mechanism from kinetic data alone.
Cappeli and Collina [75J investigated the kinetics of five different commer-
cial ammonia synthesis catalysts. The following rate equations were tested:
1) The Temkin-Pyzhev equation modified by Dyson and Simon (112)
2) the equation proposed by Nielsen et al. (83)
and 3) a new equation:
(114)
The new equation was claimed to give the best description of the experi-
ments. The k4aNH3 term was found to be insignificant compared to the other
terms in the denominator. Industrial data from a third bed (adiabatic) with
intermediate cooling were also well described by Eq. (114).
Guacci et al. [76J tested eight different commercial ammonia catalysts under
industrial conditions. They used the Temkin-Pyzhev equation modified by
Dyson and Simon to analyze their results. The oc values found for the different
catalysts ranged from 0.426 to 0.687
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 177
(115)
SSQ = 2:(1 _ kl
kl
meas)2
calc
(116)
at each temperature level. The Q( values varied from 0.48 to 0.79 in a nonsyste-
matic manner, not correlated with the potassium content. The reaction order of
hydrogen through the value of ill did, however, increase systematically from 0.71
to 1.55 with increasing potassium content.
Altenburg et al. speculated that this trend could be explained by assuming
a fast reaction to NH, and other partly hydrogenated species between adsorbed
Nand H from polarized hydroxylic groups on the oxidic part of the catalyst,
followed by a reaction between hydrogen and the 0 groups. This pathway is
blocked by potassium either by reaction with the acidic OH groups or by
screening of adsorbed N from these groups. They noted that the promoting
effect of potassium was most pronounced at higher pressures.
tU Zhenming [78] tested two different commercial ammonia catalysts using
the Temkin-Pyzhev equation for analysing the results. The pressure was fixed at
around 148 atm and the temperature was varied from 338-550°C. The catalyst
size was 0.2-0.3 mm, and a stoichiometric feed gas was used. The optimum
Q( value was 0.5, and activation energies of 42.1 and 44.8 kcaljmol were found.
Bowker et al. [82] used computer calculations to extrapolate the results ob-
tained by Ertl et al. [83] at low pressures and coverages to industrially relevant
conditions around 450°C and above 100 atm. They used the reaction mecha-
178 J. B. Hansen
and assumed iron surface area for a commercial ammonia synthesis catalyst of
1.5 m Z/g and an energetic homogeneous surface where single-site Langmuir
isotherms describe the adsorption processes. To each of the elementary reac-
tions a preexponential factor and an activation energy were assigned. The
preexponential factors were estimated from transition state theory. All adsorp-
tion and desorption reactions of molecular Hz, N z and NH3 were assigned
a preexponential factor of 10 13 , all other reactions a factor of about lO z1 . The
experimentally available activation energies at low surface coverages were taken
from the work of Ertl et al. Knowing the overall heat of reaction (50 kJ mol- 1 )
and assigning an activation energy of 21 kJ mol- 1 to all the dehydrogenation
steps, the complete potential energy diagram could be constructed.
Using these parameters the performance of a back mix reactor was cal-
culated at conditions similar to the experimental data of Nielsen et al. [59]. The
model predicted ammonia percentages which were 1.5 x 10 5 lower than those
experimentally found.
Lowering the preexponential factors for both the forward and reverse of step
2 by a factor of 10 7 only improved the fit slightly. However, the model predicted
that the surface was totally hydrated instead of being totally nitrided as in the
case of "normal" preexponential factors of lO z1 . Using the energies determined
at high surface coverages had virtually no effect on the model predictions. The
authors attributed the very poor agreement with the experimental findings to
an energy well in the model too deep for adsorbed nitrogen atoms in the
model.
Reasonable fits could be obtained (within 10% of experimental values) if
activated dissociation of nitrogen was invoked as, found by Scholten et al. [85]
on singly promoted catalysts. The authors point out, however, that there is no
surface science experimental evidence for activated nitrogen dissociation on
potassium promoted iron.
Stoltze and N0rskov [85-88] used almost the same reaction scheme as
Bowker et al. i.e.:
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 179
the only difference being that a molecularly held precursor state of H2 is not
considered for the dissociation of hydrogen. The following assumptions were
made: the gas is considered to be ideal and all surface sites are assumed to be
equal (homogeneous surface). Adsorbate-adsorbate interaction is thus con-
sidered to be absent until the coverage reaches one adsorbate atom per two iron
atoms. Above this coverage adsorbate-adsorbate repulsion excludes further
adsorption. Hydrogen and nitrogen as well as reaction intermediates are as-
sumed to chemisorb competitively. The active area of the catalyst is taken to be
that of the CO chemisorption area. Finally, all the reaction steps are assumed to
be in equilibrium except step 2 i.e., this step is assumed to be the sole rate-
determining step.
By generalizing Fowler and Guggenheim's [89] statistical mechanical de-
scription of adsorption that covers competitive gas adsorption, the equilibrium
constants were calculated. For instance, for step 1:
K 1 -_ ZN2'
0 (119)
ZN2
where ZN2' is the partition function for N2(adsorbed) and Z~2 the partition
function for N2 (gas) at the thermodynamic reference pressure Po taken to be
101.325 kPa. The partition functions were calculated from
(120)
Ztran" Zvib, Zrot are the partition functions for translational, vibrational, and
rotational degrees of freedom, whereas Ex is the ground-state energy for the
species. The equilibria can be specified as follows:
Po
N2
Kl ( P ) 8* = 8N2* (121)
P NH3
K 6 0NH3 * = To 0* (125)
where Ox is the surface coverage by species X, and 0* is the fraction of the surface
not covered by any surface intermediate. For step 2 the rate can be calculated
from
(127)
where
(128)
Noting that
(129)
(130)
(131)
[92, 93J in their work did not vary the gas temperature, but only the temper-
ature of the solid. The three different kinetics proposed are summarised in table
4.1, where the different activation energies and preexponential factors for the
critical steps 1 and 2 in the reaction scheme (118), i.e.
Step 1 N 2 (g) + * -+ N 2 -*
Step 2 N 2 -* + * -+ 2N-*
used in the kinetics, are given together with the yield calculated from the models
(experimentally found 13.2%) and the surface coverages by the predominant
speCIes.
Bowker concluded that it is necessary to invoke an activated nitrogen
dissociation step in order to reproduce the experimental findings on industrial
catalysts as well as the measured desorption spectra widths. It is not clear from
the article, whether the surface coverages given in the table are average or
typical for the exit of the reactor. Bowker discussed the differences between the
different crystal planes of iron, both promoted and non promoted, but he still
assumed in his model the surface to be homogeneous.
This review has illustrated, that the large amount of work done on the kinetics of
ammonia synthesis and decomposition has stimulated new concepts of reaction
kinetics catalyzed by heterogeneous catalysts. The concept of a non-uniform
surface vis-a-vis a homogeneous surface is just one example, which is still
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 183
debated today. Ammonia synthesis also is one of the first reactions, where micro
kinetics based on modern surface science techniques has been used with a fair
degree of success. However, there is no doubt that the finer details of the
ammonia synthesis reaction mechanism and kinetics still will be a controversial
topic for years to come.
In summary, it is noteworthy that the first workable rate equation for
ammonia synthesis was proposed before the second world war by people with
very limited experimental resources. The Temkin-Pyzhev equation still is the
rate equation of choice for engineering purposes. Further progress in better
understanding of the fundamental rate-determining process(es) is still, however,
needed in order to develop better catalysts.
The kinetics discussed in the preceding section has been (or should have been)
derived under conditions where transport processes do not restrict the measured
reaction rate. In ammonia synthesis as with many other catalytic reactions it
may be necessary to consider also the mass and heat transfer to and inside the
catalyst particle, depending on operating conditions and catalyst particle size.
If mass and/or heat transfer is slow compared to the reaction rate, temper-
ature and concentration gradients between bulk gas and particle center will
result, thus affecting the overall reaction rate of the particle. In industrial
reactors the diffusion inside the catalyst pore system is the most important
transfer process. In laboratory reactors with low linear velocity and/or large
particles, there may also be significant concentration and temperature gradients
between particle surface and bulk gas [95, 96]. Due to the high heat conductivity
of the iron catalyst, a catalyst pellet will be almost isothermal [96].
The intrinsic activity of an ammonia synthesis catalyst is also known to
depend on the particle size during reduction [60,97,43]. This effect is due to an
irreversible poisoning of the outer reduced part of the particle by water from the
inner, still reducing part. The water causes the iron particles to recrystallize,
resulting in a lower surface area and hence lower activity. This effect can be
circumvented by a very careful reduction at high space velocities; this is possible
in laboratory reactors, but normally not feasible in industrial converters.
Thus the lower activity of large particles does not necessarily mean that the
reaction is diffusion limited at the operating conditions; diffusion retardation
can be recognized, however, by a lower activation energy for the larger particles.
The activity of a catalyst particle is often expressed by the effectiveness factor
11 defined as the ratio of the actual reaction rate R to the reaction rate Rkin at
conditions on the outer surface of the particle.
(132)
184 J. B. Hansen
0.8
Qld..
1::.
b) e 425°C
e" • 450°C
o 4750C
0.6
A t:. 5000C
° 525°C
t= 0.4
Fig 4.1. Dependence of the effectiveness
factor, on the ammonia concentration
0.2 z at different temperatures and two differ-
ent sizes of the catalyst pellets, a) 4.5 mm
b) 10 mm. SA-1 catalyst, P = 300 atm
0 (from [98])
0 0.1 0.2 0 0.1 0.2
z z
The effect of the operating conditions and particle size on 1'/ is illustrated in
Fig. 4.1 [98]. It is mainly the concentration of ammonia that varies throughout
the pellet, whereas the relative changes in the N z and Hz concentrations are
much smaller. Therefore the effectiveness factor depends mainly on the diffusion
rate of ammonia.
If the diffusion coefficient in a straight cylindrical pore is denoted by D, the
effective diffusion coefficient Deff in the catalyst is given by
Deff =-D
e (133)
T
where e is the porosity of the catalyst particle and T the so-called tortuosity
factor depending on the geometry of the pores.
D can be calculated by the methods outlined by Satterfield [99]. Under
industrial conditions, i.e. at high pressures, diffusion occurs mainly in the bulk
and D is thus the bulk diffusion coefficient of the components in the mixture. At
low pressures Knudsen diffusion will also playa role, and the pore volume
distribution must be known in order to calculate D.
r z -d [Di
- dCiJ
- +v· R -- 0 (134)
dr rZ dr 1
and·
(136)
where D j and Vj are the effective diffusion coefficient and stoichiometric coeffic-
ient of component i. R is the reaction rate for a component with Vj = 1, and rp is
the particle radius. If the particle is not spherical, the proper particle dimension
is
(137)
Vg/Ag is the ratio of surface to volume of the catalyst grain. Values of the
tortuosity factor reported in different investigations are shown in Table 4.2.
In solving (134) D j is usually considered constant, although D j depends on
the gas composition. This approximation seems reasonable considering the
uncertainty in r. For simple reaction kinetics, the solution to (134) is given in
terms of the Thiele modulus [102]. The more complicated kinetics for ammonia
synthesis does not give a simple analytical solution. Generally, a numerical
integration has to be carried out.
Bokhoven and Rayen [97J measured reaction rates on 0.5-0.7 mm and
2.4-2.8 mm particles at 1, and 30 atm and 325-550°C. The effective diffusion
coefficient of NH3 was calculated from the results of O 2 diffusion measurements
on a catalyst with the same surface area and porosity as the catalyst used in the
activity measurements. The authors approximated the reaction rate by a pseudo
first-order reaction suggested by Wagner [103J to calculate the effectiveness
factor. Good agreement between measured and calculated data was obtained.
However, the approximation above is only good, if the reaction is near equilib-
rium.
Peters and Krabetz [43J measured reaction rates at 1 atm on Al 2 0 3 pro-
moted iron catalysts with varying particle size (0.5-5.2 mm) and mean pore
radius (110-310 A). The effective diffusion coefficient ofNH 3 was also measured
in a Wicke-Kallenbach experiment. They also found good agreement between
the measured and calculated data using a pseudo first-order approach.
Dyson and Simon [72J calculated 1] for a 6-10 mm size industrial catalyst
(using rate data reported by Nielsen [59J) by numerical integration of the
Table 4.2. Tortuosity factors for different catalysts and the method by which they
were determined
differential equation. Their model also includes bulk flow due to volume change
during the reaction. An expression for the effectiveness factor as a function of
conversion and temperature was given at 150, 225, 300 atm.
Nielsen [100] also solved the simultaneous differential equations numer-
ically and reported the results for a hypothetical reactor operating at 450°C,
214 atm, and with a 3: 1 H2/N2 feed gas containing 3% ammonia and 12%
inerts. The concentration profiles calculated at the bed inlet and outlet are
shown in Figs. 4.2 and 4.3. Nielsen used r = 2.2, based on a detailed, elec-
tron-microscopic study of the pore system. Figs 4.4 and 4.5 show the calculated
ammonia concentration in the bulk at half pellet radius and the effectiveness
factor for particle diameters of 5.7 mm and 1.5 mm as a function of axial
distance. For the 1.5 mm particles, the effectiveness factor very quickly ap-
proaches 1. The higher reaction rates of the 1.5 mm particles are utilized in
modern radial flow converters.
Rusov et al. [101] investigated a complicated model that also included molar
flux due to the volume change during the reaction and different diffusibilities of
the various components. They concluded, that using a simple model with
constant D j = DNH3 for all components was almost as accurate as the complic-
ated model. The reaction rates for five different catalysts were measured in an
ideal back-mix reactor operating at 400-525°C and P = 300, 450 atm. The
kinetic parameters in the Temkin-Pyzhev equation were estimated on crushed
particles. DNH3 and r could be calculated from rate measurements on large
particles. The tortuosity of the investigated catalysts varied between 3 and 15. It
should be noted that the porosity of these investigated catalysts was rather low.
In the above-mentioned investigation (J. changed from 0.35 to 0.9 upon
. increasing the temperature. Beskov et al. [98] showed that the reaction rates on
~ 454~ ~
450'---- - - - - - - - - - - '
24 72
16 48 N
Z0
0-
•
M
J:
Z N
o J:
0- 0
0-
8 24
Fig 4.2. Radial concentration profiles in 5.7 mm
nitrogen
ammonia catalyst particles located at the bed inlet.
Pressure 214atm, temperature 450°C, inerts 12%
(from [100])
O~----------------~O
Center Surface bulk
Kinetics of Ammonia Synthesis and Decomposition on Heterogeneous Catalysts 187
~ 454~ j
450~t:-============::::1-
--
24 72
ammonia
16 - hydrogen ............." 48 -£
;/!
+
N
l:
;e
8 r- - 24
Fig 4.3. Radial concentration profiles in 5.7 mm
nitrogen ammonia catalyst particles located at the bed out-
let, Pressure 214 atm, temperature 450°C, inerts
13.7% (from [100])
o o
Center Surface bulk
0.8
large particles could also be fitted to the Temkin-Pyzhev equation, but with
lower values of IX. In the temperature range 450-525 °e, IX changed only slightly
(IX = 0.4-0.5) for the particle sizes 4-5 mm and 10 mm.
Kazarnovskaya et al. [71] reported a lot of data from an experiment in
a recycle reactor with four different particle sizes. The reactor operated at
188 J. B. Hansen
400-550 DC, P = 200, 250, 300 atm and four different space velocities. For the
largest particle size (7-10 mm) they found an apparent activation energy of
20 kcaljmol. This was half the activation energy for the small particle size
(0.25-0.5 mm).
Sokolinskii et al. [104] found the following approximate expression for 1'/ in
reasonable agreement with the experimental data reported in [71] for high
values of 1'/:
1+ ( -C)2
1] - 1 _ ~ r~Robs Ceq
(138)
- 15 DNH3C
1 _ (~)2
Ceq
where Robs is the observed reaction rate, and c and Ceq are the concentration and
equilibrium concentration of ammonia, respectively. This expression can be
used to determine whether the measured rate data are diffusion-limited.
4.8 References
31. Temkin MI, Kiperman S (1947) Zhur Fiz Khim 21: 927
32. Sidorov IP, Livshits VD (1947) Zhur Fiz Khim 21: 117
33. Kobayashi H, Kubota H (1949) Rep Faculty Eng, Hokkaido Univ 3: 136
34. Temkin MI (1950) Zhur Fiz Khim 24: 1312
35. Nielsen A (1950) An Investigation on Promoted Iron Catalysis for the Synthesis of Ammonia,
Ed I, Gjellerup, Copenhagen
36. Temkin MI, Kiperman SL, Lukjanova 11 (1951) Doklady Akad Nauk (USSR) 74: 763
37. Brill R (1951) J Chern Phys 19: 1047
38. Annable D (1952) Chern Eng Sci Vol 1,4: 145
39. Kipermann S, Granovskaya VS (1952) Zhur Fiz Khim 26: 1615
40. Sidorov IP, Livshits VD (1952) Zhur Fiz Khim 26: 538
41. Adams RM, Comings EW (1953) Chern Eng Prog 49: 359
42. Bokhoven C, van Heerden C, Westrik R, Zwietering P (1955) Catalysis vol III, Reinhold, NY
43. Peters C, Krabetz R (1956) Z Elektrochem 60: 859
44. Scwab GM, Krabetz R (1956) Z Elektrochem 60: 855
45. Kawamura M, Irie T, Yokota H (1957) 60: 1239
46. Shiskova VV, Sid oro v IP, Temkin MI (1957) Tr GIAP 7: 62
47. Logan SR, Moss RL, Kemball RL (1958) Trans Faraday Soc 54: 922
48. Mills AK, Bennett CO (1959) Jour AIChE 5: 539
49. Kubota H, Shindo M (1959) Kagaku Kogaku 23: 242
50. Emmett PH (( [8J) Fixed Nitrogen, Reinhold, NY
51. Uchida H, Kuraishi M (1955) Bull Chem Soc Japan 28: 106
52. Boudart M (1956) Jour AIChE 2: 62
53. Stelling PO, Krusenstierna 0 (1958) Acta Chern Scand 12: 1095
54. Temkin MI (1949) Probl Kinet Katal 6: 54
55. Kwan T (1956) J Phys Chem 60: 1033
56. Ozaki A, Taylor HS, Boudart M (1960) Proc Roy Soc (London) A 258: 47
57. Brill R, Tauster S (1962) J Chern Phys 36: 2100
58. Krabetz R, Peters C (1963) Ber Bunsen-Ges Phys Chern 67: 381
59. Nielsen A, Kjrer J, Hansen B (1964) J. Catal 3: 68
60. Nielsen (1968) An Investigation on Promoted Iron Catalysts for the Synthesis of Ammonia, Ed
3, Gjellerup, Copenhagen
61. Temkin MI, Morozov NM, Shapatina EN (1963) Kinetics Katal 4: 260
62. Temkin MI, Morozov NM, Shapatina EN (1963) Kinetics Katal. 4: 565
63. Bridger GW, Snowdon (1970) Catalyst Handbook, Chapter 7, Wolfe Scientific, London
64. Aika K, Ozaki A (1969) J Catal 13: 232
65. Logan SR, Philip J (1968) J Catal 11: 1
66. Carra S, Ugo R (1969) J Catal 15: 435
67. Takezawa N, Toyoshima I (1966) J Phys Chern 70: 594
68. Takezawa N, Toyoshima I (1966) J Res Inst Catalysis, Hokkaido Univ 14: 41
69. Takezawa N, Toyoshima I (1966) J Catal 6: 145
70. Brill R (1970) J Catal 16: 16
71. Kazarnovskaya DB, Atamonovskaya RM, Bomshtein EI (1973) Zh Fiz Khim 47: 1445
72. Dyson DG, Simon JM (1968) Ind Eng Chem Fundam 7:605
73. Buzzi Ferraris G, Donati G (1970) Ing Chim Italiano 6: 11
74. Buzzi Ferraris G, Donati G, Rejna F, Carra S (1974) Chern Eng Sci 29: 1621
75. Cappelli A, Collina A (1972) I Chern E Symp Series 35, 5: 10
76. Guacci U, Traina F, Buzzi Ferraris G, Barisone R (1977) Ind Eng Chern 16, No 2: 166
77. Altenburger K, Bosch H, van Ommen JG, Gellings PJ (1980) J Catal 66: 326
78. Zhenming 1. (1981) J Chern Ind Eng (China) 1: 39
79. Huang KH (1981) Stud Surf Sci 7: 554
80. Huang KH (1980) Scientia Sinica 24, NO.6: 800
81. De-ming L, Shi-liang Z, Zi-xing H, Xi-jun W (1979) J Chern Ind Eng (China) 2: 133
82. Bowker M, Parker IB, Waugh KC (1982) Appl Catal 14: 101
83. Ert! G (1983) Catalysis, Science and Technology, Vol 4, Anderson JR, Boudart M (Ed)
Springer Berlin Heidelberg New York, p 273
84. Scholten JJF, Zwietering P, Konvalinka JA, de Boer JH (1959) Trans Faraday Soc 55: 2166
85. Stolze P, Norskov JK (1985) Phys Rev Lett 55, No 22: 2501
86. Stolze P, Norskov JK (1987) J Vac Sci Technol A 5(4), 581
87. Stolze P (1987) Phys Scripta 36: 824
190 J. B. Hansen
P. E. H0jlund Nielsen
Haldor Topsoe A/S Copenhagen, Denmark
Contents
5.1 Introduction
An ammonia catalyst is poisoned when its catalytic activity is reduced due to the
presence of certain elements. These elements may be present in certain gaseous
compounds contained in the synthesis gas or solid and as such introduced to the
catalyst during the manufacturing process. With respect to the latter, reference is
made to Chap 2 by P. Stoltze.
Concerning gaseous catalyst poisons, a distinction should be made between
permanent poisons causing an irreversible loss of catalytic activity and tempor-
ary poisons which lower the activity while present in the synthesis gas. For a
review of poisoning of synthesis catalysts, see [1,2]. Permanent poisons accu-
mulate on the surface and may be detected by chemical analysis of the poisoned
catalysts, whereas temporary poisons cause a partial coverage of the catalyst
surface. Since oxygen is the most common temporary poison, it is difficult to
detect the amount on the spent catalyst by analysis since the promoter phases
are difficulty reducible metal oxides like alumina, magnesia, silica, and pot-
assium oxide.
heating in H2 did not lower the work function to its unpoisoned level, indicating
an irreversible poisoning. Experiments on iron and single (AI 20 3 ) promoted
catalysts were carried out [6-8]. They found that about 0.2 mg S/m2 catalyst
surface completely deactivated the catalyst. The maximum uptake of S was
found to be 0.4-0.5 mg/m 2 catalyst surface. Chemisorbed sulfur could not be
removed by heating the synthesis gas up to 900 K. Catalysts to which sulfur was
added during activation were just as poisoned as the ones where sulfur was
added along with the synthesis gas. A very high level of sulfur was measured, and
bulk FeS was found on the spent catalyst. An amount of 6300 ppm of sulfur
completely deactivated the catalyst, whereas amounts as low as 200 ppm gave a
partial deactivation to the order of 15% [8]. Analysis of spent industrial catalyst
from the inlet layer reveals from 1000 to 5000 ppm of sulfur on catalysts KMI
and KMII [1]. Considerable poisoning by exposure to 20-30 ppm(vol) H 2S for
24 h at a space velocity of 10 000 h -1 was observed [1].
In general, any compound that can lower the surface tension of iron may be
considered as a poison [9]. Phosphorus and arsenic compounds are also known
as permanent poisons [1]. However, in the natural gas-based plants they are
rare and only found in exceptional cases. A catalyst prepared with 1-3 wt%
PiOs was examined and found to be severely poisoned [10]. Chlorine [1] is a
serious poison that reacts with the potassium promoter and forms potassium
chloride which is slightly volatile and consequently removed from the catalyst
surface. It is expected that other halogens behave similarly [1]. Calculations
show that ppb amounts in the synthesis gas should result in a marked
deactivation [2].
High pressure studies carried out using O 2 , CO, CO 2 as poisoning agents [1]
demonstrated that 50 ppm Oz is just as poisonous as 100 ppm CO, and that a
6-day exposure to 100 ppm of CO at 450°C did not harm the catalyst.
Experiments were conducted using radioactive tracer methods with CO and
COz as poisons [17]. A slight amount of COz was found to be irreversibly
absorbed; the potassium promoter was thought to be responsible for that.
Studies were carried out with a number of promoted catalysts at temperatures
between 300 and 450°C, a space velocity of 5600, and a pressure of 100 kg/m 2
(9.8 MPa) [18]. At lower temperatures, some of the CO was not converted. At
higher temperatures, the poisoning was found to be reversible. The experiments
at lower temperatures were not continued long enough to determine if the
poisoning was reversible.
Changes in work function originating from the chemisorption of oxygenic
compounds and the HzS upon the catalyst were studied [5]. HzS was found to
cause a permanent increase in the work function, whereas the oxygenic com-
pounds caused a temporary increase. Experiments were conducted at ambient
pressure using a singly promoted catalyst with water as poisoning agent [19].
Between 1.2-32.3 mg of oxygen was found chemisorbed onto the catalyst which
has a total surface area of 250 m Z• However, the very low space velocities
employed and the duration of the experiment were not sufficient to reach steady
state. Experiments were conducted at ambient pressure [20-21] which sugges-
ted a kinetic expression, something which had been done earlier based on
theoretical consideration [22]. In [21] the rate expression was applied to data
obtained at elevated pressure. The expression being
k P k P~H3
+ N2 - -~
w = H2
[ P~H3 + C' P Ja
H20
P~2 P H2
where k+ is the rate constant for the synthesis reaction; k is the rate constant for
the decomposition reaction and (X is a constant, see also the chapter by B0gild
Hansen.
The expression assumes a displacement reaction
196 P. E. H0jlund Nielsen
8= a + bT + cTln[H~~J
By varying the H2/N2 ratio, it was found that the expression
8= a + bT + c 1TlnXH20
where XH20 is the molar fraction of H 20, gave a better overall description [27].
This apparent independence of the hydrogen pressure was explained in Refs.
[28] and [29]. By using surface science data, kinetics valid at high pressure were
established. From these data it was established that during synthesis the surface
was covered by N-species; see also the old result in Ref. [15]. By assuming that a
chemisorbed 0 also influenced chemisorption upon neighbouring sites, an
explanation of the results of Ref. [15] could be established. Accordingly, Stoltze
and N0rskov [28, 29] simply added an equilibrium reaction to their set of
equations, see the chapter by B0gild Hansen upon kinetics, the equation being
H 20 (g) + 3*f:+2H - * + 0 - *
where * denotes a surface site.
By doing so, a satisfactory description of the results reported in Refs. [26, 27]
could be obtained.
Poisoning of Ammonia Synthesis Catalysts 197
----::..::--..:::..,--
----,
1.0
'"'''. ,
"'-. ", SObar-400°C
O.B '\/-
.'( \SObar-4S00C
\. \\
Z.
:2: 0.6 \. \
\
tI
Cl \. \\
<II
:g> \ \
q; 0.4
a:
\. \\
\. \
\
\ \
0.2 \ \
\ \
\, \
0 ''''''-.. ' .... -
ru 100
ppm "0"
5.7 References
1. Nielsen A (1968) An Investigation on Promoted Catalysts for the Synthesis of Ammonia, Jui.
Gjellerup, Copenhagen, pp 126
2. H0jlund Nielsen PE, (1991) Catalytic Ammonia Synthesis, ed. by J. R. Jennings, Plenum
New York
3. Benard J, Oudar J, Barbouth N, Margot E, Berthier Y (1979) Surf Sci 88: L35
198 P. E. Hojlund Nielsen
4. Grabke HJ, Paulitschke W, Tauber G, Vielhaus H (1977) Surf Sci 63: 377
5. Enikeev E, Krylova AV (1962) Kinet Katal 3: 139
6. Brill R, Tauster S (1963) Ber Bunsen-Ges phys Chern 67: 390
7. Tauster S, PhD Thesis, Polytechnic Institute of Brooklyn, June 1964, available university
microfilms, 64-10, 728
8. Brill R, Schaefer H, Zimmerman G (1968) Ber Bunsen-Ges Phys Chern 72: 1218
9. Rostrup-Nielsen JR, Hojlund Nielsen PE (1985) In: Oudar J, Wise H (eds) Deactivation and
Poisoning of Catalyst Dekker, New York, p 259
10. Marakhovets et al. (1972) Izv Vyssh Ucheb Zaved Khim Tekhnol p 735
11. Catalyst Handbook, ed. M. V. Twigg, Wolfe London 1989, p 407
12. Larson AT, Tour RS (1922) Chern Met Eng 26: 647
13. Almquist JA, Black CA (1926) J Am Chern Soc 48: 2814
13. Almquist JA (1926) J Am Chern Soc 48: 2820
14. Emmett PH and Brunauer S (1930) J Am Chern Soc 52: 2682
15. Ussatschew PW, Tarakanowa WJ, Komarov WA (1934) Z Elektrochem 40: 647
16. Brunauer S, Emmett PH (1940) J Am Chern Soc 62: 1732
17. Bokhoven C (1954) Proceedings of the Second Radiosotope Conference, Oxford Butterworths,
London, pp 53
18. Uchida H, Todo N (1951) Report of the Tokyo Industrial Research Institute Laboratory 46: 213
19. Royen P, Langhans GH (1962) Z Anorg Allg Chern 315: 1
20. Smirnov lA, Morozov NM, Temkin MI (1965) Kinet Katal 6: 351
21. Smirnov IA (1966) Kinet Katal 7: 107
22. Kiperman SL (1954) Zhur Fiz Khim 28: 389
23. Brill R. Hensel J, Schaefer H (1969) Ber Bunsen-Ges Phys Chern 73: 1003
24. Ozaki A, Taylor H, Boudart M (1960) Proc Roy Soc (London) Ser A258, 47
25. Boreskova EG, Kuchaev VL, Temkin MI (1984) Kinet Katal 25: 112
26. Hojlund Nielsen PE; The Poisoning of Ammonia Synthesis Catalysts by Oxygenic Compounds,
lecture given at the ACS Meeting, Washington, D.C. Sept. 1983.
27. Andersen S, Haldor Topsoe AjS, Private Communication
28. Stoltze P (1987) Phys Scr 36: 824
29. Stoltze P, Norskov JK (1987) J Vac Sci Technol A5, 581
30. Kirkerod T, Skaugset P (1991) Abstracts of IV Nordic Symposium on Catalysis, Trondheim
Chapter 6
Ammonia Production Processes
Ib Dybkjaer
Haldor TopS0e A/S Lyngby, Denmark
Contents
6.1 Introduction
The present chapter describes process technology for the production of am-
monia. The first part gives a short review of early developments in ammonia
technology. The second part contains a brief description of processes used in the
production of ammonia synthesis gas, i.e. a mixture of hydrogen and nitrogen
with or without minor amounts of impurities such as methane, argon, etc. A
third part describes the conversion of synthesis gas to ammonia. This part
includes discussion of the design of ammonia synthesis converters. A fourth part
describes integrated processes for production of ammonia from primary raw
materials such as hydrocarbons and coal. A final part deals with storage of
ammonia.
Several publications are available which give an integrated treatment of
important aspects of ammonia production. Some examples are [1-9, 876, 949].
Relevant patents are reviewed in [10, 11]. Historical aspects are treated
especially in [12-17, 877]. Reviews of the technological status at various times
are given in [1-9J and in [18-42,949]. Catalytic reactions and catalysts relevant
for ammonia production are treated in [43-46].
6.2 Historical
The main purpose of the present chapter is to describe modern technology for
ammonia production. For the sake of completeness a brief description of the
early development of ammonia synthesis technology will be given in the
following. More details can be found in the relevant literature.
The history of modern ammonia production started in Germany just after
1900, when Fritz Haber and his assistants developed the process concept which
is the basis for all ammonia production even today. Haber patented his work in
two famous patents, the "circulation patent" [47J and the "high pressure patent"
[48]. The claims of the patents read:
"Circulation patent": "Verfahren zur synthetischen Darstellung von Am-
moniak aus den Elementen, wobei ein geeignetes Gemenge von Stickstoff und
Wasserstoff kontinuerlich der Ammoniakbildung mittels erhitzter Katalysa-
toren und nachofolgender Ammoniakentziehung unterworfen wird, dadurch
gekennzeichnet, daB hierbei unter dauerndem Druck gearbeitet und dafiir
gesorgt wird, daB die Warme der ammoniakhaltigen Reaktionsgase auf das von
neuem der Reaktion zu unterwerfende ammoniakfreie Gasgemisch iibertragen
wird".
"High Pressure patent": "Verfahren zur Darstellung von Ammoniak aus den
Elementen durch Katalyse unter Druck bei erh6hter Temperatur, dadurch
Ammonia Production Processes 203
gekennzeichnet, daB die Vereinigung unter sehr hohen Drucken von etwa 100
Atmospharen, zweckmassig aber von 150 bis 250 Atm. und mehr vorgenommen
wird."
The German company BASF had been following Haber's work through
their two employees, Carl Bosch and Alvin Mittasch and eventually bought
Haber's patents. Development work was continued on catalyst, process, and
equipment design, and in 1913 the first commercial plant with a capacity of 30
tons/day was commissioned at Oppau/Ludwigshafen near Mannheim in Ger-
many.
By 1916 the capacity had been increased to 250 tons/day. In 1917 a second
plant was started at Leuna near Leipzig. At the end of the First World War this
plant produced 240000 tons/year [14]. In 1937, the annual world capacity was
755000 tons/year, of which 72% was still concentrated in Oppau and Leuna
[16].
Soon after the First World War, others started development work, partly on
the basis of the German pioneering effort. A plant was started at Terni, Italy in
1920 based on technology developed by Luigi Casale. In France, a plant was
constructed in 1921 at Grand Paroisse near Montereau, based on developments
by M. G. Claude. Both the Casale and the Claude technologies operated at
extremely high pressures. In further early developments in Europe, technology
developed by Giacomo Fauser was adopted by the Italian company Monteca-
tini, and Friederich Uhde GmbH constructed a plant to operate on coke oven
gas from the Mont Cenis company. This plant used a special process later
known as the Mont Cenis process. Its features were a very low operating
temperature and pressure and a special catalyst based on iron cyanide.
Meanwhile, an independent development took place in the USA. As early as
1918 a plant called "US Nitrate Plant No.1" was constructed in Muscle Shoals,
Alabama. This plant was, however, not very successful, but the American
development continued, and during the 1920s several successful plants were
constructed, some based on American, some on European technology. In 1928,
the Nitrogen Engineering Corporation (NEC) was commissioned to construct a
plant in Europe [49], thereby entering the worldwide competition.
During the 1930s and during the Second World War production capacity
was increasing rapidly using now well-established technologies. In 1945 about
125 plants existed with a total capacity of 4.5 million tons [49]. The most
important processes were: Haber-Bosch, Casale, Claude, Fauser, NEC, and
Mont Cenis. For descriptions of the technical features of these processes, see
[50].
Just before the second world war it was a general belief [18] that ammonia
synthesis technology was mature, and that no significant further developments
could be expected. This was to some extent true, since most of the features which
characterize modern synthesis technology - including catalyst type and most
major converter designs - were already well proven in industrial operation.
However, since that time, there has been a tremendous development in the scale
204 Ib Dybkjaer
of operation, and at the same time, the main type of feedstock has shifted from
coke to natural gas, naphtha or fuel oil. New technology for the production of
synthesis gas has been developed, and integrated, efficient processes for large
scale production of ammonia from the different feedstocks have emerged.
The ultimate raw material for ammonia production is synthesis gas, a mixture of
hydrogen and nitrogen with or without minor quantities of impurities such as
methane, argon, etc. Normally, the concept "Production of Ammonia" is,
however, understood to cover the complete transformation of primary raw
materials into ammonia [8]. The nitrogen part of the synthesis gas is always in
some way derived from air, either in a separate process step, most often by
cryogenic air separation, or in integrated processes, where air is used as a
reactant in such a way that oxygen is consumed in the gas preparation process,
while nitrogen remains as constituent in the synthesis gas (see below). When
nitrogen is produced in a separate process by air separation, the oxygen part of
the air is most often used in the process for production of the hydrogen part of
the synthesis gas.
The most important primary raw materials for the hydrogen part of the
synthesis gas are natural gas and other light hydrocarbons. Other raw materials
include heavy hydrocarbons, solid materials such as lignite and coal, and
hydrogen recovered from off-gases from other processes. Production of syn-
thesis gas can be subdivided into the following process steps (see Fig. 6.1).
-Gas Preparation
From natural gas and light hydrocarbons by steam reforming, catalytic
autothermal reforming, or partial oxidation.
From heavy oil or coal by partial oxidation or gasification.
From water by electrolysis.
From hydrogen-rich off-gases by separation of hydrogen.
-Gas Purification
In most cases, the gas purification train contains the following process steps:
- Carbon monoxide conversion
- Carbon dioxide removal
- Final purification
Final purification can be by:
- Methanation or Methanolation
- Cryogenic purification ("Nitrogen wash")
- Copper liquor wash
Ammonia Production Processes 205
Natural Gas
LPG, Naphtha
Enriched
Air Air
Autothermal
Reforming
,--- ---,
I Cryogenic I
I Purifier, I
L~r_J
reactor tubes which are heated from the outside by burning fuel. The reactor
tubes are made of high alloy steel and filled with a nickel-based catalyst. A
comprehensive treatise on steam reforming reactions and steam reforming
catalysts is given in [64, 65]. General literature on reforming: [66-72, 880, 881];
208 Ib Dybkjaer
dominant in the US ammonia industry, when natural gas became the preferred
feedstock. In early plants using steam reforming, the operating pressure was low
[71, 139-141]. As developments in metallurgy allowed, the pressure was
gradually increased up to the present levels of about 35 bar [142-145].
6.3.2.2 Gasification
In certain situations coal and lignite are attractive as feedstocks for the
production of ammonia, and plants using these raw materials have been
constructed in many parts of the world. It should also be noted that coke was
the most important feed- stock in the early days of the ammonia industry in
plants using the water gas process (see Sect. 6.3.6). Natural gas or heavier
hydrocarbons replaced coke during and after the Second World War except in
special circumstances. During the oil crises of the 1970s, coal appeared to be the
future feedstock for ammonia production, and many studies and much develop-
ment work were carried out on processes for the production of ammonia from
coal. However, later developments in oil prices made coal less competitive as a
feedstock, and the expected large scale change in the industry from natural gas
and heavier hydrocarbons to coal never did materialize.
Ammonia Production Processes 211
The product gas from steam reforming, catalytic partial oxidation, partial
oxidation, or gasification processes contains in all cases significant amounts of
carbon monoxide and carbon dioxide. Since the feed gas for the ammonia
synthesis loop must be completely free of these compounds, they must be
removed in the gas preparation part of the plant. Carbon monoxide is removed
in a conversion step by the so-called "water-gas shift" reaction or just the "shift"
reaction, in some cases followed by further conversion by selective oxidation.
In plants based on steam reforming of light hydrocarbons, the product gas from
the shift conversion contains about 18 vol% (dry basis) carbon dioxide. Gas
originating from partial oxidation of heavy hydrocarbons or from gasification of
solid feedstocks contains even higher concentrations of carbon dioxide, and in
addition hydrogen sulfide may also be present in significant quantities depend-
ing on the sulfur content in the feed and on the type of shift conversion
technology employed. In plants using standard shift systems, sulfur is most often
removed from the raw gas upstream of the shift section in selective sulfur
removal units, whereas the sulfur is left in the gas in plants that use special sulfur
tolerant shift catalysts. In every case all sulfur compounds and all carbon oxides
must be removed from the gas before it can be used for ammonia synthesis.
Essentially all sulfur compounds (if present) and the bulk of the carbon dioxide
are removed in so-called acid gas removal units or carbon dioxide removal
units, which in all cases feature absorption of the acid gases in a suitable solvent.
214 Ib Dybkjaer
After bulk removal of carbon monoxide in the shift conversion section and of
carbon dioxide in the carbon dioxide removal section, the synthesis gas still
contains typically 0.2-0.5 vol % carbon monoxide and 0.01-0.2 vol % carbon
dioxide. These compounds must, together with any water present, be removed
quantitatively, i.e. to low ppm levels, before the gas can be admitted to the
synthesis converter, because all oxygen containing compounds are poisons to
the ammonia synthesis catalysts [238]. The most important technologies for this
final purification are discussed in the following.
process removes carbon monoxide and carbon dioxide to the required low levels
by absorption in a solution containing cuprammonium salts of acetic, formic, or
carbonic acid. The chemistry involved is rather complex; both the Cu + jCu + + -
ratio, the pH, the ammonia concentration, the total concentration in water, the
temperature and pressure etc. must be controlled to obtain proper absorption
and to avoid precipitation of copper salts. Reference is made to [321, 323]. The
loaded solution is regenerated by pressure release and heating, and the composi-
tion is adjusted by necessary addition of ammonia, adjustment of the
Cu + jCu + + -ratio (by addition of air if required) etc., before the solution is
recycled to the absorber.
The copper liquor washing process has the advantage that the carbon oxides
are removed by absorption rather than by a hydrogen-consuming reaction as in
the more commonly used methanation reaction. It is, however, difficult to
operate; maintenance coats are high; it has a high energy consumption; it tends
to pollute the environment; and it therefore has largely been replaced by other
purification processes. It is not used in any modern ammonia plants.
igatic;ms on the risk imposed by the simultaneous presence of nitric oxides and
olefins in cryogenic units are reported in [334].
After pretreatment the gas is cooled by heat exchange with the cold product
gas. Additional refrigeration is obtained by expansion of the gas and the
nitrogen, or by addition ofliquid nitrogen. The gas is cooled to about - 190°C
~t which temperature partial condensation of carbon monoxide and methane
.::urs. The gas is contacted countercurrently in a washing column with liquid
ltrogen. The gas leaving the top of the washing column is hydrogen containing
.tbout 10 vol% nitrogen and only traces (about 100 ppm) of argon and methane.
After addition of nitrogen to adjust the hydrogen to nitrogen ratio to 3.0 this
is an excellent feed gas to a so-called "inert free" synthesis loop which may often
be operated without withdrawal of a purge gas stream. The liquid nitrogen
leaving the bottom of the washing column contains essentially all the carbon
monoxide and methane present in the original feed gas. These compounds can
be used as fuel or returned to the feed gas preparation.
The nitrogen required for the nitrogen wash is normally produced in the air
separation plant which produces oxygen for the gasification or partial oxidation
process. High purity nitrogen is required with less than 10 ppm oxygen.
Descriptions of the nitrogen wash technology may be found in [335-340].
6.3.5.4 Methanation
By far the most important process for removal of the last traces of carbon
monoxide and carbon dioxide from ammonia synthesis gas in methanation. In
this process, carbon oxides are reacted to extinction (less than 10 ppm) at
250-350°C over a nickel-containing catalyst according to the following ex-
othermic reactions:
co + 3H 2 -+ CH 4 + H 20( - LlHg 98 = 206 kJ/mol) (6)
CO 2 + 4H2 -+ CH 4 + H 20 ( - LlHg 98 = 165 kJ/mol) (7)
218 Ib Dybkjaer
6.3.5.5 Methanolation
Methanolation [907, 908] may be considered an alternative to methanation.
The residual carbon oxides are converted in the methanolation process to
methanol, preferably at high pressure, i.e. after the synthesis gas compressor.
The methanol formed is removed by washing with water and may be returned to
the feed stream to the reformer or separated as a product. Full conversion of
.carbon oxides is not possible, and a clean up methanation unit must be installed
after the methanolation unit.
By far the largest part of the world's ammonia production is made from
synthesis gas which is produced by one of the processes described above, i.e. by
steam reforming oflight hydrocarbons, by partial oxidation of hydrocarbons, or
by gasification of solid feedstocks, in all cases followed by proper conversion,
purification, and adjustment of composition.
There are, however, also other sources of synthesis gas, although their
importance is limited.
In the early days of the ammonia industry the predominant feedstock was
coke. Synthesis gas was produced in so-called water gas units, which were cyclic
units. Coke was heated by partial combustion with air, whereby a mixture of
essentially nitrogen and carbon dioxide, so-called generator gas, was produced.
The hot coke was thereafter contacted with steam whereby a mixture of
hydrogen and carbon oxides,so-called water gas, was produced. The two gas
Ammonia Production Processes 219
streams were purified and mixed, and carbon monoxide was catalytically
converted to carbon dioxide in a high temperature shift unit operating at low
pressure. The converted gas was compressed to, say, 25 bar, carbon dioxide was
removed by washing with water, and the product gas was further compressed to,
say, 200 bar, whereafter residual carbon oxides were removed in a copper liquor
wash before the gas was sent to the synthesis unit. The process, which was used
in the first ammonia plants in Oppau and Leuna, is described in [50, 366]. An
alternative process for synthesis gas preparation, which was used mainly in
Europe and Asia, was purification of coke oven gas [50, 367]. In this process,
coke oven gas was first purified by the removal of tar, benzene, ammonia, sulfur-
compounds etc., after which hydrogen was recovered in a cryogenic unit.
Nitrogen was obtained from an air separation plant, and the pure gases were
mixed and sent to synthesis.
There is still a limited production of ammonia based on coke oven gas,
although its importance is dwindling. The prospects for the future are reviewed
in [368, 369].
Hydrogen for ammonia production may also be produced by hydrolysis
[370-374] or it may be obtained from refinery off gas [375]. Plants have been
constructed from such feedstocks in the past [376, 377] and it may still be
economical in special circumstances. In modern times plants have been con-
structed on the basis of off gases from petrochemical plants such as ethylene
plants and methanol plants [378, 379], but such cases will most likely remain
rare and be economical only in special circumstances. Production of ammonia
from gases produced in steel plants is suggested in [380-383].
Whatever the production process for synthesis gas has been, the gas must be
compressed before it is added to the ammonia synthesis loop. Up to around
1950, synthesis gas was generally produced at atmospheric pressure. The
pressure has since then been gradually increased, and today the synthesis gas is
normally available for compression at 25-30 bar in plants using steam refor-
ming and at even higher pressure, up to about 70 bar [166] in plants based on
partial oxidation of heavy oil. Synthesis pressure was typically 200-400 bar in
early ammonia plants with extremes of 1000 bar in the original Claude process
[384] and around 80 bar in the Mont Cenis process (385). Capacities were low
compared to today's standard - increasing through the years to a maximum of
300-400 MTPD around 1960 - in part because of capacity limitations in the
reciprocating compressors used in the early plants. In the 1960s the capacity of
ammonia plants increased to about 1000 MTPD (metric tons per day) when
centrifugal compressors became available; synthesis pressures were around 150
bar in the first ammonia plants due to limitations in pressure available from the
centrifugal compressors.
220 Ib Dybkjaer
The technology has since developed further, and today plants with a
capacity of 1500 MTPD and a synthesis pressure of 220 bar are operating. In
fact the availability of compressors is no longer imposing any limit either on
capacity or on synthesis pressure in ammonia plants.
As indicated above, two types of compressors are important for the com-
pression of ammonia synthesis gas. In old plants - and in some small modern
plants - reciprocating compressors dominate, whereas centrifugal compressors
are used exclusively in all large modern plants. Reciprocating compressors are
described in [386, 387] and centrifugal compressors in [388-397]. Extensive
discussions on the use of centrifugal compressors are given in [392-397].
Reciprocating compressors are generally driven by synchronous electric motors
or in a few cases by steam turbines with special speed reduction gears. They are
more expensive than centrifugal compressors, require more maintenance, and it
is normal practice to install two machines in parallel so that excessive down time
caused by the compressor is avoided. They are lubricated so that there is a risk
for contamination of the synthesis loop with lubricating oil, they require much
space and heavy foundations for installation; and their capacity limitation
makes multi train installations necessary above a certain plant size. On the other
hand they are highly efficient, and the efficiency is rather independent of pressure
and load.
Centrifugal compressors can handle very large gas volumes; they are reliable
so that only one machine is normally installed even in very large ammonia
plants; and they have the advantage of being oil free. There is a practical limit to
the minimum clearance between an impeller and the stator and therefore to the
minimum flow (in actual volume) which can be handled by a centrifugal
compressor. Technological developments are constantly pushing this limit, but
there will always be a certain capacity limit below which centrifugal compressors
can not be used. Present technology (early 1990s) probably allows a flow
corresponding to around 350 actual cubic meters per hour at the compressor
outlet to be economically compressed in a centrifugal compressor at ammonia
plant conditions. This corresponds to an upper limit in synthesis pressure of
around 100 bar in a plant producing 300 MTPD ammonia.
Centrifugal compressors operate with very high rotational speed and are
most often driven directly by steam turbines, although motor drivers are also
used in some cases, especially for relatively small capacities. The centrifugal
compressors are inherently less efficient than reciprocating compressors, al-
though technology development may be closing the gap.
Direct steam turbine driver also means that generator loss, distribution loss,
and motor and gear loss are avoided, so that the overall efficiency of a steam
turbine driven centrifugal compressor may well compare favourably with the
efficiency of a motor driven reciprocating compressor, at least in the upper end
of the capacity ranges. Motor drives and steam turbine drivers for the large
power consumers in ammonia plants are compared in [398].
Gas turbines have also been considered as drivers for compressors in
ammonia plants. General discussions are given in [399, 400]. The exhaust gas
Ammonia Production Processes 221
from the gas turbine can be used for steam production, for preheat duties, or as
combustion air in the primary reformer [401,402]. Experience from an actual
installation is reported in [909].
The synthesis of ammonia takes place in a recycle loop which always contains
the following elements:
- A reactor system comprising one or more catalytic synthesis reactors with
associated temperature control and heat recovery equipment. The design of
the reactors and of reactor systems is discussed in Sect 6.4.2 below. The
catalysts employed are discussed in Chapter 2. Thermodynamics of the
synthesis reaction are dealt with in Chapter 1. Reaction kinetics are discussed
in Chapter 4.
- Unit(s) for cooling the product gas to recover heat and to condense the
product ammonia. The layout of these units, which are integrated with other
parts of the synthesis loop, is dealt with in the discussion of individual
processes in Sect 6.4.6.
- Unit(s) for the separation of product ammonia from unreacted gas and for the
adjustment of product properties (degassing, adjustment of temperature and
pressure). The gasjliquid equilibrium in the separation stage is dealt with in
Chapter 1.
- Units for preheating the reactor feed. These units are most often integrated
with the cooling/condensing units mentioned above.
222 Ib Dybkjaer
- Equipment for the addition of make-up gas and, if the make-up gas contains
inerts, for removal of purge gas from the recycling gas to prevent the build up
of inerts in the loop. The addition of make-up gas and removal of purge gas
may be done at various points in the recycle loop as discussed in Sect. 6.4.1.1
below.
- Equipment for the recirculation of non converted gas and make-up gas to the
synthesis reactor. The recirculation equipment is in many cases integrated
with equipment for compression of make-up gas to synthesis pressure as
discussed in Sect 6.4.1.1 below.
.-----4
L-_-IIIoL-_. . 6
recycle of unconverted gas). In modern processes the concept with the addition of
make-up gas after the separator has been used or proposed in a number of
processes where the synthesis gas is especially purified before addition to the
synthesis loop. Examples are the MDF process by Humphries and Glasgow
[341-343], the Linde process for small ammonia plants [344] the PARC process
by KTI [345-349], and other processes using pressure swing absorption for
purification of synthesis gas. Additional examples are the C. F. Braun process
[326-331,409-412], the ICI AMV process [412-418] and other processes using
cryogenic units for gas purification, and the new M. W. Kellogg process [419],
the LEAD process by Humphries and Glasgow (420), and other processes where
the gas is dried in molecular sieve units before being added to the loop.
Descriptions of all these processes are given in Sect 6.5.3.2.3.
In cases where the synthesis gas is not especially purified, and where it may
therefore contain traces of water vapour and carbon dioxide, the make-up gas is
added before the separator. In this way the make-up gas is in contact with liquid
ammonia, which absorbes the traces of water vapour and carbon dioxide so that
these poisons do not reach the synthesis converter.
This latter layout means that the ammonia is no longer condensed and
separated at the highest possible concentration, and it further means that the
inlet concentration of ammonia to the synthesis converter is equal to the outlet
concentration from the separator, since the gas after the separator is not diluted
with fresh synthesis gas. As a consequence, there is a certain increase in the
energy consumption in the loop because of a slightly increased recirculation rate
and/or a certain increase in the energy required for condensation of ammonia.
Synthesis loops with make-up gas addition upstream of the separator are often
referred to as the "3 nozzle design" and the "4 nozzle design" depending on the
location of the recirculator [404, 405, 421, 422]. These designations refer to the
layout ofthe compressor section. In the "4 nozzle design" illustrated in Fig. 6.2B,
the circulator is located after the separator. This means that make-up gas and
recycle gas cannot be mixed in the compressor, so that 4 nozzles are required,
inlet and outlet for the make-up gas and inlet and outlet for the recycle gas. In
the "3 nozzle design" illustrated in Fig. 6.2 C, the recycle compressor is located
before the separator. In this way the make-up gas can be mixed with the recycle
gas in the compressor, and therefore only 3 nozzles are required, inlet for make-
up gas and recycle gas and outlet for the combined stream.
The "3 nozzle design" has a disadvantage compared to the "4 nozzle design"
in that the volume of gas passing through the recycle compressor is increased
because the product ammonia also has to be compressed here. The difference
between the energy consumption in the two concepts is discussed in [404, 405].
Arrangements corresponding to the "4 nozzle design" have been used in
many commercial processes including the Grande Paroisse process [423] and
processes by Uhde [424,425] and Topsoe [426,427]. The "3-nozzle design" has
e.g. been used in the Casale process (with recycle done by an ejector) [384, 408,
428] and notably by M.W. Kellogg [429-432].
Ammonia Production Processes 225
In loops operating at high pressure (above about 250 bar) it has been found
advantageous to install two ammonia separators (404), (405). Fig. 6.2D and 6.2E
show two ways of doing this. Fig. 6.2 D is equivalent to the "3 Nozzle Design"
shown in Fig. 6.2 C, whereas Fig. 6.2 E represents the corresponding "4 Nozzle
Design". The characteristics of the designs with one separator are maintained in
each case, and in addition energy is saved, because part of the product ammonia
bypasses the last, energy consuming refrigeration stage. The "3 nozzle design
with 2 separators" (Fig. 6.2 D) has been used by M. W. Kellogg [407, 433J,
Chemico [434J, BASF [435J, Fauser-Montecatini [436J, Pritchard [437].
Topsoe [438J, and Esso [439]. The "4 nozzle design with 2 separators"
(Fig. 6.2 E) has been used in the Haber-Bosch-Mittach process [440J and the
equivalent NEC-process [441J, by Uhde [442, 443J, and by ICI [444].
In a few cases, loop designs have been suggested which cannot be described
by any of the types represented by Fig. 6.2 A-E. This is notably so for certain
complicated loops with double conversion and double separation, e.g., in late
versions of the Claude process [445-447]. A similar concept has been proposed
by PDIL [448]. It has also been proposed [449, 450J to place two synthesis
loops in series to improve overall efficiency, and to install a "preconverter" - i.e.
a synthesis converter in the make-up gas stream - in order to improve
conversion and to protect the synthesis loop against poisons [451]. Pre-
conversion has also been used in revamp of existing synthesis loops to increase
capacity [452].
catalyst activity (or volume for constant activity) very significantly since the
ammonia synthesis reaction on existing types of catalyst is strongly inhibited by
the product ammonia (see Chapter 2).
A further point which favours an elevated operating temperature in the
catalyst bed, and thereby - through the equilibrium relations - an elevated
operating pressure, is the risk of poisoning. Ammonia synthesis catalysts are
extremely sensitive to poisoning by oxygen-containing compounds such as H 2 0
and CO. The poisoning is caused by a reversible adsorption of oxygen species on
the active sites, and the equilibrium is such that at temperatures below about
350 DC, almost complete coverage - and therefore almost complete deactivation
of the catalyst - is obtained even at concentrations of oxygen-containing
compounds about or below 1 ppm [238]. Such low concentrations are very
difficult (and expensive) to achieve, and as a consequence, the risk of poisoning
sets a practical lower limit to the catalyst temperature and thereby the operating
pressure.
The power consumption for synthesis gas compression, for recirculation,
and for refrigeration depends critically on the operating pressure. An extensive
study on the influence of operating pressure (in the range 140-315 bar
(2000-4500 psig)) on total power consumption is reported in [404, 405]. Loops
with single and double condensation and with 3- and 4-nozzle compressor
configuration (see Sect 6.3.1) are considered, and the power consumption is
given for a number of cases. The main conclusions are that when realistic
compressor efficiencies are used, then the total power consumption for a
constant loop pressure drop depends only marginally on operating pressure in
the range considered.
There seems to be a flat minimum at about 155 bar (2200 psig); a lower loop
pressure drop means, of course, lower power consumption; the 4-nozzle design is
slightly more energy efficient than the 3-nozzle design over the entire pressure
range; and the design with double condensation (considered only at high
pressure) is more efficient than single condensation. Some of the results are
shown graphically in Fig. 6.3 (from [405J).
32 --------=
total recycle refrigerant
and syn. gas compressors 0
!-3stagedesign
a abQ\le3200 p.s.i.g.
'- 24
~0.
Q.I
III
'-
0
/.
.r;
"6 16
.!:!
~ /' refrigerant
0
Q.I
.r;
l-
--- compressor
S ,,
,
Fig. 6.3. Theoretical horsepower versus syn-
thesis pressure for base case design (early M. W.
Kellogg process) for plant capacity of 1500 short
tons of ammonia per day (from [405])
M
1.5
J:
Z
I-
~
...,
<!)
III 1.0
g'
.~
>-
.n~ 0.5
promise between energy consumption and capital costs in various parts of the
synthesis unit. For an existing synthesis loop, a reduction in separator temper-
ature will mean an increase in production capacity [449, 462] since the
converter with its given catalyst volume will be able to make more ammonia due
to the lower inlet concentration.
The inert level in the synthesis loop (most often measured at converter inlet)
depends on the inert level in the make-up gas, the production of ammonia per
unit make-up gas (the loop efficiency), and the purge rate. It may easily be seen
that when two of these four parameters are given, then the two others can be
calculated, if steady state (including constant pressure) is maintained in the loop.
The inert level in the make-up gas is, of course, determined by the conditions in
the synthesis gas preparation unit. The ammonia production is determined by
conditions around the converter, the gas flow (which may be expressed by the
recycle ratio) the inlet temperature and pressure, catalyst volume and activity,
and converter configuration (see Sect. 6.4.3).
This means that for a given situation neither the purge rate nor the inert level
can be changed without affecting operating pressure. It also means that if the
operating pressure or the catalyst activity (by change of catalyst) is changed,
then the other parameters will change until a new steady state is obtained.
6.4.2.1 General
Ideally, the design of a chemical reactor is a strictly logical process, in which the
optimum solution to the problem is determined from experimental and other
data by well-defined procedures. However, the real world does not work in this
way. This is clearly shown, if not by other evidence, then by the multitude of
different solutions which have been found in reactor designs for the same
problem. All the solutions are normally technically and economically very
reasonable - and none of them can claim to be clearly superior in every case.
The term "reactor design" is often interpreted as meaning "calculation of
necessary reactor (catalyst) volume". There are, however, [457] several other
factors to consider in reactor design - at least when this is interpreted as all the
information required as an instruction to a workshop to manufacture a reactor
for a specific practical application. Some of the more important points are listed
below (from [459]).
Chemistry and thermodynamics
- the reaction including possible side reactions
- thermodynamics
- equilibrium
- heat of reaction
- properties of reactants and products
230 Ib Dybkjaer
Reaction kinetics
- intrinsic + diffusion parameters
- pellet kinetics
Properties of catalyst
- thermal stability
- particle size and shape
- pore system
- mechanical strength
- change in properties during activation and
operation
Process optimization
- energy recovery
- pressure drop
limitations on volumetric flow
Process control
- temperature control
- flow distribution
- safety aspects, start-up/shut-down
Mechanical design
- materials of construction
- workshop manufacture
- sea and land transport
- erection
The basic data - the chemistry and the thermodynamics for the process - will
define limits for a broad range of conditions at which the reaction is possible.
The kinetics of the reaction and the properties of the catalyst, especially its
thermal stability, will further narrow the range of possible reaction conditions
and define a "window" of possible operating parameters. Process optimization,
energy efficiency, and safety aspects will then determine at what conditions
within the "window" the reactor should operate to give the optimum result. And
then mathematical models are used to determine how big the reactor must be to
obtain the performance (conversion and pressure drop) determined by the
process optimization.
- internal cooling with cooling tubes in the catalyst bed or with catalyst in tubes
surrounded by a cooling medium. The internal cooling can be effected either
with gas flow in the same direction in the catalyst bed and cooling channels
(co current flow), with gas flow in opposite directions in the catalyst bed and
cooling channels (countercurrent flow), or with gas flow in the catalyst bed
perpendicular to the flow in cooling channels (cross flow). The cooling
medium can be either synthesis gas or some other medium, for example
boiling water.
- quench cooling by injection of cold gas. The injection of quench gas can be
either between adiabatic beds or into a catalyst bed at different locations.
Flow in the catalyst beds can be either axial or radial in vertical converters or
downwards in horizontal converters.
- external cooling by heat exchange between catalyst beds. The cooling medium
can be either synthesis gas or some other medium, for example boiling water.
Flow in the catalyst beds can be either axial or radial in vertical converters or
downwards in horizontal converters.
It is of course possible to combine several cooling methods in the same converter
~ystem. Furthermore, the catalyst beds and/or heat exchangers including feed
effluent heat exchanger may be arranged in one pressure shell or in individual
pressure shells. It is clear that this leads to a very significant number of possible
converter configurations.
The various converter types may be characterized by the temperature profile
through the catalyst bed(s) or by the temperature/concentration profile (plots of
temperature vs ammonia concentration for the gas passing the converter) (see
Fig. 6.6a-d below). Such profiles are often compared to maximum reaction rate
profiles, see Fig. 6.5 (from [460]). It is seen from this figure that when the
temperature is increased (at otherwise constant conditions, including constant
ammonia concentration), then the reaction rate will increase up to a maximum
value; when the temperature is further increased, the rate decreases until it
becomes zero at the equilibrium temperature. The temperature/concentration
points where maximum rate is achieved describe a curve, the maximum rate
curve, which will normally be roughly parallel to the equilibrium curve, but at
30-50 °C lower temperature. It is clear that the minimum catalyst volume would
be obtained in a converter where this maximum rate curve were followed. In the
early days of ammonia production, available technology limited the obtainable
size of the converter pressure shell, and the physical dimensions of the converter
Ammonia Production Processes 233
optimum operating
:z:'" line
z lines with constant
reaction rate
(5
>
Temperature
thus limited the achievable production capacity. Great emphasis was therefore
given to the ammonia production per unit converter volume. Quite complicated
mechanical constructions were used to maximize the production capacity of a
given volume, and the converters were compared to the "ideal" converter where
the temperature/concentration plot follows the maximum rate curve (see e.g.
469)). As technology developed, other considerations such as optimum heat
recovery, reliable mechanical construction, easy catalyst loading and unloading,
etc. became more important than maximum production per unit volume, and
different - mechanically simpler - converter configurations have, therefore,
become more popular.
In connection with "revamps", i.e. modification of existing plants for in-
creased capacity and/or increased energy efficiency (see Sect. 6.4.3.4), where the
pressure shell of existing converters is reused, it has, however, again become a
major consideration to maximize the production capacity of the volume avail-
able inside the pressure shell.
Ammonia converters most often consist of two separate parts, an outer
pressure vessel and an internal "basket" containing the catalyst bed(s), internal
piping for gas distribution, heat exchangers (when applicable) for control of
catalyst temperatures, and in some cases a feed-effluent heat exchanger, so that
all high temperatures are contained inside the pressure shell.
Early types of ammonia synthesis converters are described in [4, 50, 385].
More recent developments are discussed in [23, 490-492]. Good overall reviews
of different converter designs may be found in [493,494].
A discussion of the types of heaters used for start-up of ammonia synthesis
converters (fired vs. electrical heaters) is given in [910].
234 Ib Dybkjaer
60)
0
U 550
~
E
~
::I:
c5
Qj 450
0-
E
~
120 3500
0.1 0.2
.,.
200 400 600
Temperature (Oe)
6b)
0
4 U 550
~
g ~
"
"
::I:
Qj 450
0-
E
{!!.
3500
0.2
.,.
200 400 600 0.1
Temperatu re (Oe)
6e) I I
2 U 550
~
~
E e" 450
::I:
6 ~
E
{!!.
3500
.,.
400 600 0.1· 0.2
Temperature (Oe)
6d)
U 550
:§ ~
~
:x:
"
"Qj 450
0-
E
{!!.
120 3000
.,.
200 400 600 0.1 0.2
Temperature (Oe)
Ammonia Production Processes 235
Fig. 6.6. Schematic drawing, typical temperature profile, and operating curve (temperature/am-
monia concentration plot) for four important converter types. (from [471])
a Internal cooling, countercurrent flow (TV A-converter)
b Internal cooling, cocurrent flow (NEe-converter)
c Quench cooling
d Indirect cooling (heat exchange)
236 Ib Dybkjaer
the cross flow configuration are significantly higher than in axial flow. A
converter with radial flow and with two zones in each of a series of catalyst beds,
an adiabatic zone and a zone with cooling tubes, is described in [521].
converter is shown in Fig. 6.7 (from [491]). The feed gas enters the converter at
the bottom and passes as shell cooling gas in the annulus between catalyst
basket and pressure shell to the feed-effluent heat exchanger mounted in the top
of the converter.
The feed gas passes on the shell side of the exchanger to the first catalyst bed.
Between the exchanger and the first catalyst bed and between the following
catalyst beds cold gas is added via quench gas distributors for temperature
control. The converted gas passes from the last catalyst bed through a center
pipe to the tube side of the feed-effluent heat exchanger and leaves the converter
at the top.
This converter type has given excellent service in industry. It is, however, due
to the inherent weaknesses of the quench cooling system, not very efficient, and
the performance has in many cases been improved by revamping, see Sect.
6.4.3.4.
Designs similar to the Kellogg design have been used by others, e.g. Grande
Paroisse [385,492,529], Kubec [530, 531] and Casale [490]. A design in which
quench cooling is applied between catalyst beds contained in separate vessels
has been used by Pritchard [437].
A later development is the M. W. Kellogg horizontal quench converter, see
Fig. 6.8 (from [492]). This converter design is described in [533-536]. The
catalyst beds are arranged in a basket which fits into a horizontal pressure shell;
outlet
catalyst bed
catalyst
catalyst
shell
inlet
manhole
Fig. 6.8. Quench cooled horizontal converter. M. W. Kellogg design (from [492])
in this way catalyst loading and unloading are facilitated and there is no need for
overhead structure or crane. The basket may be removed from the pressure shell
simply by drawing it out on tracks.
In operation, gas flow is downwards through the catalyst beds. Because of
the large area, shallow beds, a low pressure drop is obtained even with small
catalyst particles. This means that the horizontal design offers some of the same
advantages offered by radial flow converters (see Sect. 6.4.3.2.3.).
6.4.3.2.2 Axial Flow, Quench Cooling by Injection of Gas into the Catalyst Bed
This converter type has mainly been used by ICI; it is described in [23,490,491].
A schematic drawing is shown in Fig. 6.9 (from [491J). Synthesis gas enters at
the bottom of the converter, flows as shell cooling gas in the annulus between the
catalyst basket and pressure shell to the top of the converter, down on the shell
side of a centrally mounted feed-effluent heat exchanger, upwards in an annulus
between heat exchanger and catalyst bed, down through the catalyst bed, and up
on the tube side of the feed effluent exchanger to the outlet in the top of the
converter. Quench gas is added at various levels in the catalyst bed through gas
distributors. A special version of this converter type - the opposed flow
converter - was developed for very large capacities [537, 538]. In this design the
converted gas is collected in the middle of the catalyst bed, and there is down
flow in the upper half and up flow in the lower half of the bed. The special feature
of the ICI design - one uninterrupted catalyst bed which can be emptied
through an opening in the bottom of the converter - is maintained in the
opposed-flow converter.
A design similar to the ICI-design has been suggested by Chemico [539].
gas outlet
quench gas inlet quench gas inlet
G
quench gas
KD»+++-- distributors
heat exchanger
diameter, but above a certain size this becomes technically and economically
unacceptable. In order to partially overcome the pressure drop problem it is
normal in axial flow converters to use relatively large catalyst particles.
However, these have lower activity than small particles due to diffusion
restrictions, and a larger catalyst volume is therefore required [459, 540].
In converters with radial flow the above mentioned disadvantages do not
exist. Radial flow converters can be designed for very large capacities without
excessive reactor diameter, and a low pressure drop can be maintained even with
very small catalyst particles. The advantages of radial flow reactors for ammonia
synthesis is discussed in [459, 502, 541, 542]. The pressure drop as function of
catalyst particle size is discussed in [488]. Radial flow has been used in
converters with cooling tubes [520], in quench cooled converters (see below)
and in converters with indirect cooling (see Sect. 6.4.3.3.2).
The most widely used quench cooled radial flow converter has been the
Tops0e designed S-100 converter, which is a two-bed quench cooled reactor. A
schematic drawing is shown in Fig. 6.10 (from [544]). The converter is described
in [23, 427, 490, 491, 534, 535, 543-547]. The main part of the synthesis gas
enters the converter at the top and passes downwards as shell cooling gas in the
annulus between the pressure shell and the catalyst basket. It then passes
242 Ib Dybkjaer
pressure sheil
~,,*-- refractory cement Fig. 6.10. Two bed radial flow quench
cooled converter. Haldor Tops0e S-100
Converter (from [543])
I- outlet main gas
inlet cold by-pass gas
and start -up heater
through the feed/effluent heat exchanger ("lower heat exchanger") on the shell
side and up through a central pipe to the first, upper catalyst bed. After passing
the catalyst bed in radial flow in the outwards direction, the gas is cooled by
quenching with fresh synthesis gas before it passes through the second, lower
catalyst bed in inward direction, downwards in an annulus between the catalyst
bed and the center pipe to the tube side of the lower heat exchanger and out
through the bottom of the converter. Cold gas is added through the bottom and
mixed with the inlet gas to the first catalyst bed for temperature control.
Radial flow quench converters have also been used by Chemoproject [548J,
Osterreichische Stickstoffwerke [549J, and Lummus [550]. In these converters,
as well as in the Top0e converter, the flow direction is restricted to be essentially
radial.
An axial-radial flow converter has been introduced by Ammonia Casale
[490,492,551-553]. The special feature of this design is that gas can enter each
catalyst bed both from the top (in axial direction) and from the side through
perforations (in radial direction), see Fig. 6.11 (from [492J). The gas leaves the
catalyst bed through perforations in the inner wall. There are no perforations in
an upper part of the inner wall, and the gas entering at the top of the bed is
therefore forced to flow through part of the catalyst in partially axial flow before
it can leave the catalyst bed. This flow principle can be used in quench cooled
Ammonia Production Processes 243
radial
gas flow zone
converters and in indirectly cooled converters (see Sect. 6.4.3.3.2). The design has
found widespread use in modification of existing converters, see Sect. 6.4.3.4.
Fig. 6.12. Converter with indirect cooling, axial flow. OSW design
(from [568])
Ammonia Production Processes 245
inlet
---=----f----=----
bed 2
outlet
interbed heat exchanger
Fig. 6.13. Horizontal converter with indirect cooling between catalyst beds. M. W. Kellogg design
(from [492])
in the heat exchanger and mixing with cold gas ("cold shot") for temperature
control, the gas passes downwards through the first catalysed bed, through the
interbed heat exchanger on the tube side, downwards through the second
catalyst bed (which is divided into two sections) and out in the same end of the
pressure shell as the main inlet. This design offers the same advantages as a
horizontal quench cooled converter - a low pressure drop with small catalyst
particles and the easy removal of converter internals without requiring a crane
or overhead structure.
A design with indirect cooling between catalyst beds is also used by C. F.
Braun. In this case the catalyst beds and the heat exchangers are each in their
separate pressure shell. In an early design [401, 411, 572, 573J two adiabatic
converters were used, and the cooling between the converters was by heat
exchange between feed and product gas to/from the first converter. The
arrangement and the converter design is shown in Fig. 6.14, and Fig. 6.15 (from
[572J). In a later design [573- 576, 952J three reactors are used, and the cooling
between the reactors is provided partly by gas/gas heat exchange, partly by
external cooling, for example by raising or superheating steam. The C. F. Braun
converter (see Fig. 6.15) contains one adiabatic catalyst bed with axial flow. Gas
enters at the bottom, flows as shell cooling gas to the top, down through the
catalyst bed and out at the bottom. The design of the bottom part of the
converter is critical due to the relatively high temperatures prevailing there.
leI has proposed a split flow converter with indirect cooling between beds
[577]. Feed gas is introduced in the top of the converter and flows to the shell
side of an interbed heat exchanger, where it is preheated to the inlet temperature
of the first conversion stage. It is then passed to at least two parallel axial flow
beds installed above the heat exchanger, where the first conversion takes place.
After these "upstream" beds the gas is cooled by passing through the interbed
heat exchanger on the tube side. It then reacts further in at least two parallel
axial flow beds installed below the heat exchanger. A somewhat similar system
has been proposed by M. W. Kellogg [578]. In this case the beds are arranged in
separate vessels, the heat exchanger is external, and the beds are with opposed
flow.
246 Ib Dybkjaer
intercooler'-++-ttlil
gas
Fig. 6.18. Four bed Kellogg quench cooled converter after modifica-
tion to Haldor Tops0e's 8-200 two bed radial flow concept (from
gas inlet [591])
gas outlet
Fig. 6.19. Four bed Kellogg quench cooled converter after modifica-
tion to Ammonia Casale's four bed quench cooled axial-radical flow
/' concept (from [591])
gas inlet
Ammonia Production Processes 251
M. W. Kellogg and the Uhde revamp technologies have only been applied to
converters originally designed by M. W. Kellogg and Uhde respectively, where-
as the Haldor Tops0e and the Casale technologies have been used to modify a
number of different converter designs (see [591]).
Revamps consisting of the replacement of old converters with completely
new converters are described in [409], (c. F. Braun: in one case a complete new
synthesis loop replaced 5 identical old loops; studies are reported on the
addition of a third reactor to an existing two reactor configuration) and in [603]
(Haldor Tops0e: an S-200 converter replaced an existing M. W. Kellogg 4-bed
quench converter). Revamp by addition of an extra converter in series with the
existing converter has been proposed by Tops0e [955], and by M. W. Kellogg
using their new KAAP technology (see Sect. 6.5.3.2.3) [949].
Nitriding is caused by ammonia reacting with the surface of the steel forming
a hard and brittle surface layer of nitride [626,627]. It has been found by both
experience and experiments that low alloy steel can be safely used in gas
containing ammonia at temperatures below approximately 400°C, whereas
stainless steel, Inconel and Incoloy can be used at any temperature relevant for
ammonia synthesis loop equipment. Formulas for prediction of the rate of
nitriding of stainless steels in ammonia synthesis loops are given in [628].
Stress corrosion cracking may be caused by a number of reasons including
presence of trace compounds, especially chlorine, poor heat treatment after
welding, etc. Some examples of stress corrosion on loop equipment are given in
[629-633]. Stress corrosion cracking of ammonia storage tanks is dealt with in
Sect. 6.5.6.1.
The above indicates that low alloy steel, for example 21/2 Cr 1/2 Mo, can be
used for most loop equipment including the converter pressure shell. Critical
parts are the reactor internals (the "basket"), which are most often made from
stainless stell (SS 304), and the inlet part of the waste heat boiler which must
either be made from or cladded by corrosion resistant material, most often
Incoloy.
washing of the purge gas. The ammonia vapour from the last flash is normally
sent to the refrigeration compressor.
Other fluids than ammonia have been used in the refrigeration circuit (see
[636J), and absorption refrigeration (normally based on ammonia/water) [637,
638J has been used instead of mechanical refrigeration. It has been proposed to
remove the ammonia from the partly cooled reactor effluent by absorption in
water. This procedure makes it necessary to dry the recycle gas before it reenters
the converter. The drying may be done with molecular sieves or by injection of
liquid ammonia into the gas stream and separation of the resulting ammonia -
water mixture [32, 639].
Ammonia recovery in part by adsorption of ammonia vapour on a solid
adsorbent has also been proposed [640-642]. The warm gas from heat recovery
passes through a cold adsorption vessel in which the adsorbent is saturated with
ammonia. The ammonia is des orbed and goes with the product gas to further
refrigeration cooling and product separation. The cold gas from the separator
passes through another adsorption vessel which has been heated and depleted
from ammonia by warm converter effluent. This vessel is now cooled and
ammonia is adsorbed from the separator effluent. In this way the adsorption
beds act as both heat and mass exchangers, transferring heat from the reactor
effluent to the reactor feed and at the same time ammonia from reactor feed to
reactor effluent. Both energy savings and increased production capacity are
claimed.
Separation of ammonia through semipermeable membranes has been sug-
gested. Use of a polyethylene membrane was proposed in 1954 [643J, whereas a
more recent development based on anion-exchange membranes is described in
[644,645J.
The synthesis gas produced by any of the methods described in Sect. 6.3. will
contain certain concentrations of "inerts", i.e. compounds which are not con-
sumed by the ammonia synthesis reaction and do not interfere with the catalyst
performance. The inerts are typically methane, argon, and traces of other rare
gases. In some cases significant amounts of helium may be present originating
from He-containing natural gas.
In plants using a liquid nitrogen wash for final purification (see Sect. 6.3.5.2)
the concentration of inerts may be so low - about 100 vol ppm or less - that they
are dissolved in the liquid ammonia leaving the separator. In all other cases it is
necessary to purge a gas stream from the synthesis loop in order to prevent
excessive build up of the inerts.
The purge gas is ideally extracted at a point, where the inerts concentration
is highest and the ammonia concentration lowest, i.e. after product separation
and before make-up gas addition. However, for various reasons this is not
Ammonia Production Processes 255
always possible, as discussed in Sect. 6.4.1.1. The purge gas will, wherever it is
taken from the loop, contain~a few percent ammonia, 10-30% inerts, mainly
methane and argon, and 60-85% of a 3 to 1 mixture of hydrogen and nitrogen.
Ammonia is normally recovered by condensation after cooling to a very low
temperature in a purge gas chiller and/or by washing with water. The remaining
gas may be used as fuel in the reformer or in auxilliary installations.
The hydrogen and nitrogen content of the purge gas represents, however,
potential raw materials for ammonia production, and energy has been invested
in the production of especially the hydrogen and in bringing the gases to the
synthesis pressure. As energy cost increased, it became increasingly attractive to
separate at least the hydrogen from the purge gas and return it to the synthesis
loop. In this way the yield of ammonia from hydrogen in the make-up gas in a
typical synthesis loop may be increased from about 94% to above 98%.
Recovery of nitrogen is less attractive, especially in plants based on steam
reforming, because preferential recovery of hydrogen will shift the required
hydrogen/nitrogen ratio in the synthesis gas to lower values, from 3.0 to in some
cases as low as 2.8. This makes it possible to transfer some of the reforming duty
from the primary to the secondary reformer (see Sect. 6.3.1 and [646, 647]),
which means a saving in the capital cost of new plants and possibilities for a
capacity increase in existing plants. The addition of purge gas recovery is an
attractive revamp option, partly because it can result in both a reduction in the
specific energy consumption and in an increased capacity, partly because in
most cases it can be done as a separate undertaking with minimum interference
to other parts of the plant.
Several types of processes have been developed for purge gas recovery. The
most important are discussed in the following. Reviews and comparisons may be
found in [648-650].
whereas the off gas becomes available at low pressure; typical recoveries (the
fraction of a compound in the incoming gas which is found in the hydrogen rich
gas) are about 90-95% for hydrogen, 25% for nitrogen, 25% for argon, and 4%
for methane. Cryogenic purge gas recovery is discussed in [650-659]. Results
obtained in industrial plants are described in [660-662]. Argon recovery in a
cryogenic purge gas recovery unit is discussed in [663].
6.5.1 General
In the foregoing sections the individual process steps involved in the production
of ammonia from various feedstocks have been described. However, it is very
important how these "building blocks" are combined with each other, and with
the steam and power systems, to form a complete facility for the production of
ammonia. The way this is accomplished has a major impact on plant efficiency
and reliability, and much of the difference between the several ammonia
processes and much of the development in ammonia production technology
may today be found in these areas. It may be said that while "Ammonia
Technology" was in the early days of the industry most often understood as
"Ammonia Synthesis Technology" or even "Ammonia Converter and Catalyst
Technology", it is today interpreted as the complete technology involved in
transformation of the primary feedstock to the final product ammonia.
The most important factor for the choice of process layout is, of course, the
type of feedstock. Before the Second World War coke dominated (see Sect.
6.3.6). During the war several plants based on natural gas were constructed in
the USA [522], and natural gas has since then been the preferred feedstock in
the USA as well as in other parts of the world. There has, however, also been a
significant production based on partial oxidation of heavy fuel oil or gasification
of coal, especially in Europe and in countries like India and China; also naphtha
has been a preferred feedstock in some areas. During the 1970s there was, due to
the oil crises, a renewed interest, especially in the USA, in coal as feedstock for
ammonia production, but an expected major change to coal-based production
of ammonia did not materialize. Comparisons of the economics of ammonia
production from different feedstocks may be found in [160, 676-691].
Another important factor is plant capacity. Until 1960 the capacity of
individual installations had increased to a maximum of about 400 MTPD. The
units were often multi-train units meaning that several parallel trains were
installed in synthesis gas preparation and the synthesis loop not necessarily the
258 Ib Dybkjaer
same number of trains in each part of the plant. In the early 1960s technological
developments made it possible to construct large capacity single train plants,
[692-695J, and since then this has been the dominating concept. In special
situations small plants may, of course, be of interest [696-700J; but in "normal"
conditions plants with capacities below about 1000 MTPD (metric tons per day)
are rarely considered [958]. On the other hand, it seems that logistics and the
dwindling economy of scale have limited the maximum capacity to below 2000
MTPD. Today there are no technical limitations which prevent the construction
of single train ammonia plants with a larger capacity.
In the following section, energy balances in ammonia production will be
discussed. The difficulties related to the comparison of energy consumption in
different process schemes will be dealt with, and finally a description will be
given of the main concepts in the integration of ammonia production units and
steam and power systems.
= CpdT - T:Cp/TdT
= d(~H)* (1 - Te/T) (14)
260 Ib Dybkjaer
In this case the exergy change is consequently equal to the change in enthalpy
multiplied with the ideal Carnot efficiency of a heat engine. The exergy change is,
therefore, small when all heat is transferred to or from the surroundings at a
temperature close to Te.
The maximum work a system can perform is found by combining Eqs. (11)
and (13).
Wide a] = - ~Ex (15)
i.e. the maximum work is equal to the negative value of the exergy change, which
is the sum of the exergies of the leaving streams minus the sum of the exergies of
the entering streams. The lost work is then found as the difference between the
ideal work the system can perform and the actual work it delivers. In the actual
work should also be included work which could be extracted from useful heat
transferred to or from the system by using an ideal Carnot engine. The lost work
is then:
(16)
where Ts is the temperature of the source delivering the heat Qs to the system.
In processes with chemical reactions it is necessary, if one is interested in
absolute values of the exergy, to define a convenient standard state where, in
principle, the compound can do no more work. Different standard states have
been defined in the literature depending on how many 'spheres' (atmosphere,
hydrosphere, lithosphere) are included in the analysis.
In gas processing systems such as the production of ammonia from natural
gas, liquid water and air at 25°C and atmospheric pressure are normally
selected as standard states for the environment. This implies that these raw
-materials are assigned zero exergies. This may not always be realistic, in
particular in places where clean water is expensive to provide.
There is also the problem of carbon dioxide which is only present in a very
small amount in the atmosphere. It can, therefore, be argued whether it is correct
to assign a value of the exergy of carbon dioxide equal to the theoretical
minimum work to extract it from the atmosphere.
For this reason, some authors prefer to define the standard state for each
pure component at atmospheric pressure, and some prefer to assign a partial
pressure to carbon dioxide as it exists in a typical flue gas. The difference
between different standard states is mainly the value assigned to air. For other
gas processing streams the difference in exergy is less than 1-2%.
In the following, the standard state selected for each component in air is the
partial pressure in dry air at atmospheric pressure. The standard state for water
is liquid at 1 atmosphere. The temperature of the surroundings is 25°C. The
exergy of a stream of arbitrary composition can now be calculated by use of
Eq. (12).
This is done in two steps. The first contribution EX 25 is the Gibbs energy of
formation of the mixture from the compounds in the surroundings and consider-
Ammonia Production Processes 261
ing the work required to transfer the individual reactants and products to 1 atm
partial pressure, which is the standard state for the Gibbs energy.
The second contribution Ex is the transfer of the compounds in the mixture
from 1 atm partial pressure at 25°C to the actual temperature and the actual
partial pressures in the mixture. This is done by use of Eq. (13), where the
enthalpy and entropy differences are calculated by standard methods. An
example of results obtained by exergy analysis of an ammonia production
process is given in Sect. 6.5.3.1.2.
c
o
'5. M 0.3r--r-----r-------,-------,
EJ:
::Jz
Ill ....
C :::E
8 ... 0.2
618
liil!>
C -
=c c3\
<II
u
0.1
The above points indicate that the steam and power balance of an ammoma
plant is of major importance for the overall efficiency.
Ammonia Production Processes 265
The concept of recovering heat for the production of high pressure steam
and using the steam for driving compressors via steam turbines in plants based
on steam reforming was developed during tne 1960s [426, 712-714]. Descrip-
tions of systems for heat recovery by high pressure steam production may be
found in [709, 715-718]. See also Sect. 6.4.4. A discussion of the energy and
money value of waste heat and steam is given in [719].
An example of a steam generation system in a large, modern, natural gas
based ammonia plant is shown in Fig. 21 and 22 (from (720}). Fig. 6.21 shows
how heat is recovered for preheating boiler feed water and for generating and
superheating high pressure (110 bar, 510°C) steam. The total heat recovery in
the steam production system is 3.00 Gcal/MT of ammonia corresponding to
4.52 tons of high pressure superheated steam per ton ammonia. It could be
possible by adjusting the design - for example by introducing combustion air
preheat - to reduce the heat available for steam production to about 2.4 Gcal
corresponding to about 3.6 tons of steam per ton ammonia. (It is of course
possible by extra firing to increase the steam production to any desired level).
Fig 6.22 shows how low grade heat is recovered in the same plant; as much
as possible - in this case slightly above 0.4 Gcal/MT of ammonia - is used to
preheat demineralized water before deaeration. Other uses of low grade heat
include regeneration of solvent in the carbon dioxide removal unit and produc-
tion of low pressure steam which is used in deaeration and in the carbon dioxide
removal unit. It may also be noted that condensate stripping is in this plant
integrated with deaeration and carbon dioxide removal. General descriptions of
process condensate treatment are given in [721, 722, 920].
Fig 6.23 shows schematically how steam is used in the same plant. Super-
heated high pressure steam from the process units generated as shown in
Fig. 6.21 is passed through the back pressure part of the turbine driving the
synthesis gas compressor; part of the steam is condensed in the condensing part
to provide sufficient power, while the main part is extracted as medium pressure
steam. The medium pressure steam is partly used for process steam, partly for
other turbines driving the air compressor, refrigeration compressor, and flue gas
fan. Low pressure steam to balance the needs in carbon dioxide removal, boiler
feed water deaeration, and other uses is generated in the back pressure turbine
driving the flue gas fan and extracted from the turbine driving the air com-
pressor. Total power production in the four large turbines is about 23 MW in
this 1000 MTPD ammonia plant.
Figs. 6.21, 6.22, and 6.23 show only one possible layout of the steam system
in large ammonia plants. Other similar systems are described in [31, 615, 718,
723, 724, 949]. Other types of drivers for ammonia plant compressors than steam
turbines are discussed in [398] (electrical motors) and [399, 400] (gas turbines).
Steam and power production for ammonia plants in cogeneration units is
described in [725-727,921]. Steam and power balances in ammonia plants and
ammonia-urea complexes are discussed in [959] with special reference to the use
of gas turbines.
t..)
superheater superheater a,
a,
BFW preheater downstream in reformer
downstream HTS sec. ref. WHS
superheated HP steam ~
I •
at 110 at. 510°C U
'<
r::r
.E.
~
>-;
BFW preheater
in reformer WH5
boiler in loop
Fig. 6.21. Steam production in a large, modern natural gas based ammonia plant (from [720])
DMW from polisher condenser
2nd separator
condensate from
3 rd separator
excess condensate
from Benfield
N
~
.....
MP steam grid IV
HP steam from process 0-
00
110 at. 510°C 38at.375°C process steam
MP steam
to consumers 0=
MP steam t:I
'<
export cr"
2r.
~
....
CW LP steam
to consumers
turbine condensate
Fig. 6.23. Use of high pressure steam in a large modern natural gas based ammonia plant (same plant as Figs. 21 and 22)
Ammonia Production Processes 269
Table 6.3. Thermodynamic values of components in ammonia production (Gcal per ton
Ammonia).
~ Synthesis at 140 kg/cm 2 g using a two bed radial flow converter with indirect
cooling.
Use of centrifugal compressors driven by steam turbines.
The energy balance was based on the following assumptions:
~ Feed and fuel: 100% methane available at the required pressure.
Cooling water available at 30 DC at the required pressure.
Ambient conditions: 30 DC, 65% relative humidity, 1 atm.
Product ammonia supplied at - 33°C to atmospheric storage.
No import or export of steam or power across ammonia plant battery limits.
The results of the analysis are summarized in Table 4.
Losses are calculated as the energy content (heating value or exergy) of the
ingoing streams minus the content in the outgoing streams including in- or
outgoing work. The total loss, therefore, includes the heat loss to the surro-
undings and the heat loss to the cooling water. For the exergy, these two losses
are in most cases much smaller than the losses due to the irreversibilities
Table 6.4. Energy analysis, low energy ammonia plant (Gcal/MT Ammonia)
produced in the process. The reason is, of course, that the heat loss is multiplied
by the Carnot efficiency.
The largest transfer of heat to the cooling water occurs in the steam system,
where the cooling needed to condense the exit streams from the turbines and for
cooling in compressor interstage coolers is equal to 1.56 GcaljMT. But since
most of the heat is transferred at about 50 DC, the Carnot efficiency is only 8%,
and the exergy loss to the cooling water is only 0.13 Gcal/MT.
The heat loss from the primary reformer occurs at a high temperature, hence
giving a high Carnot efficiency. The exergy value is best evaluated by consider-
ing the extra fuel necessary to compensate for the heat loss. This is equivalent to
about 5% of the total exergy loss from the reforming section or about 3% of the
heating value of the total fuel to the reformer.
Almost 70% of the total exergy loss is lost in reforming section and the
associated steam generation, mainly due to irreversibilities present in the
combustion.
This result can be compared with an ordinary energy or first law analysis,
which shows that almost all energy is transferred to the cooling water.
The exergy analysis can be used to calculate a thermodynamic efficiency for
production of ammonia. It is simply the ratio between the exergy of product and
the total feed which is:
Efficiency = exergy ratio = 4.81/7.33 = 0.66 or 66%
Feed
St -I p.
nm. ......
cr
= ~ 0
'<
Pressure, barg ~
Temp., ·C 'j;;.
Carbon Oxides 1.5-2.0% CO 0.Q1-1% COz 5 ppm ~
Compressor Stages 2 2 1
Air
Feed
2 Steam : Prim. Sec. r--1J~~v.
Ref. Ref. (HTS)
1 H
Pressure, barg 4.2 3.2
Temp., ·C 770 930
Carbon Oxides 1.7% CO 0.01% COz 5 ppm
Compressor Stages 1 2 1
Air
4 Feed __ , . I I Final
Steam Pnm. ~ Sec. Purif. Lr...-I NH3
(Meth.) 1..:r11 Synt.
Pressure, barg 14.5-18.4
Temp., ·C 800-820
Carbon Oxides 0.8-1.0% CO 0.01% COz 5 ppm
Compressor Stages 3
Air
Feed
5
Steam
Feed
6
Steam
Pressure, barg
Temp., ·C
Carbon Oxides 0.5% CO 0.01% CO2 5 ppm
Compressor Stages 4 +2
Air
Feed
7
Steam
Pressure, barg
i
o
i:l
Temp., ·C 0;'
Carbon Oxides 0.5% CO 0.01-0.1% CO2 5 ppm '"c:I
....
Com pressor Stages 2 to 220 bar o
0-
3 to 330 bar
ao·
Fig. 6.24. Some steps in the development of ammonia process layout between 1940 and 1972 (efr. also Table 5) (adapted from [25]). i:l
'"c:I
....
o
g
'"~
'"
tv
-.l
W
N
....,
""'"
5i
t:l
'<
cr'
5.
%
....
Table 6.5. Comments to Fig. 6.24
No. Year Capacity COr Removal No.ofNH 3 Process Features and Improvements
short tons/day Converters
tl
'-"
276 Ib Dybkjaer
fuel
I A ..... 1\ 1\ J\ .... I
VVv""v'4
steam "
a i r - - - -........- - - - - j
heat recovery
heat recovery
water
NH3 converter
----.-----.
:>
§o
::>
0.;'
'"0
...,
o
0-
~
o·
::>
~g
'"
ri
Fig. 6.25. The classical single-train M. W. Kellogg ammonia process. Simplified process flow diagram (from [430])
'"
N
...,
...,
278 Ib Dybkjaer
- Final purification was in the early plants most often done with a copper
liquor wash. With the advent of the low temperature shift, it became feasible
to do the final purification by methanation, and since then this has been the
preferred method. See also Sect. 6.3.5.
Ammonia Production Processes 279
In Gases where excess air is used in the secondary reformer, it is necessary to adjust
the composition of the synthesis gas by the removal of excess nitrogen. This has
been done in a cryogenic purification unit or in a Pressure Swing Adsorption
unit. See also Sect. 6.3.5.
In the ammonia synthesis, improved reactor design, increased converter
volume, and to some extent improved catalysts have improved the conversion
efficiency. Improved heat recovery has been introduced with the production
of high pressure steam instead of preheating of boiler feed water, and single
train capacity has been dramatically increased. Drying of make-up synthesis
gas and recovery of hydrogen from the purge gas have been introduced to
improve performance. See also Sects. 6.4.3,6.4.4 and 6.4.7, where the improve-
ments in loop and converter design have been extensively reviewed.
In the steam and power system, developments in compressor and turbine
technology have led to significant energy savings. The use of gas turbines as
direct drivers for compressors or as part of cogeneration units producing
steam and power has also been introduced. Heat recovery has been improved
by increasing the steam pressure, by introducing high pressure steam produc-
tion in the synthesis loop, and by combining steam production and steam
superheating downstream of the secondary reformer. Production of process
steam from low level heat by feed gas saturation has been used in some cases,
and stripping of process condensate with low or medium pressure steam for
reuse as boiler feed water has been introduced. See also Sects. 6.3.7,6.4.4, and
6.5.2.3.
Developments in instrumentation and computer science have led to the
increased use of advanced control systems and computerized optimization of
plant operations. Advanced control systems may be used to keep key
operating parameters (e.g. steam to carbon ratio at inlet reformer, reformer
exit temperature, hydrogen to nitrogen ratio of synthesis gas, synthesis loop
pressure, etc.) constant in spite of variations in feedstock properties or other
external changes [742-750, 922, 965]. The control systems can be in open
loop, where the computer system evaluates data collected by the plant
instrumentation and gives new set points to the plant operator who then
manually changes the controls. The system can also be in closed loop where
the set points are directly controlled and changed by the computer to
optimum values calculated in a computer model on the basis of operating
data and external (economic) data.
Computer simulation of plant operation may also be made off line real time using
computer models of the complete installation, which are capable of simulating
dynamic plant response to changes in operating parameters, plant upsets, etc.
Such systems may be used for off line optimization studies and for operator
training in handling emegencies, start-up- and shut-down situations, etc. Without
risk to plant or personnel. Simulators are described in [751-758, 923].
280 Ib Dybkjaer
The Braun Purifier process [28, 32, 33, 37, 324-330,409-412, 759, 952] has
been available for more than 20 years. A simplified process flow scheme of the
updated version is shown in Fig. 6.26 (from [330]). Key features of the process
are:
- Reforming with a large ( ~ 50%) excess of air. Steam to carbon ratio at inlet
of the primary reformer approximately 2.8. Relatively high methane leakage
( ~ 1% dry) from the secondary reformer.
gas turbine
, . primary reformer C02 CO 2 removal
, aIr compressOr r.:-:l secondary
alr--, .--. ~ reformer
I I
steam I
LL::::r:J • I
water
methanator
•• cryogenic purifier
:>
ammonia product
~
::s
iii'
..,'"0
o
Po
§
g.
Fig. 6.26. The Braun purifier process, simplified process flow diagram (from [330]) ::s
~
~
~
'"
to..>
00
......
282 Ib Dybkjaer
-. Air compressor driven directly by gas turbine. Exhaust from the gas turbine is
used as combustion air in the primary reformer.
- Conventional high temperature and low temperature shift conversion.
- Carbon dioxide removal, normally by BASF's MDEA process.
- Final purification by methanation followed by adjustment of synthesis gas
composition (removal of excess nitrogen and part of the inerts) in a cryogenic
unit (referred to as the Braun purifier). Gas drying upstream of the purifier.
- Synthesis of ammonia in a "dry" synthesis loop (with addition of make-up gas
after the ammonia separator) at approximately 180 kg/cm 2 g with the Braun
converter system consisting of two or three single-bed, adiabatic, axial flow
converters. Cooling between the converters is by synthesis gas preheating
and/or by steam production.
Proprietary items are the purifier and the converters. The process scheme as
such is unique and is not used by other licensors. The use of excess air in the
secondary reformer and the high methane leakage both contribute to a signific-
ant reduction in the size of the primary reformer compared to more conven-
tional process schemes and thereby to savings in fuel consumption. On the other
hand, the power consumption in the air compressor is higher than in conven-
tional plants; the whole front end up to the purifier must handle more gas, and
the purifier adds pressure drop. Energy consumption in the synthesis loop is low
due to the dry and relatively pure synthesis gas. The use of a gas turbine as the
driver for the air compressor adds extra firing and thereby increases the amount
of waste heat generated in the plant. The extra heat can be recovered for steam
production, but the plants will by necessity generate a surplus of steam for
export.
About fifteen plants have been designed using the Braun process. Experience
from operating plants using the Braun technology is reported in [329, 330, 332,
333, 760-763, 952]. A net energy consumption of 6.7 Gcal/MT of ammonia has
been reported [330] for a specific case, operating in a cold climate with partial
production of gaseous ammonia and with full credit for a large export of
18 kg/cm 2 g steam. In [952] energy consumptions of 7.18 and 6.92 Gcal/MT of
ammonia are reported for two plants, also in a cold climate.
I CI has been active in the design and operation of ammonia plants since
before the Second World War. In recent years, they have commercialized two
processes namely the AMV process, and the LCA process.
The AMV process[37, 413-418, 764] was introduced in the early 1980s. A
simplified process flowsheet is shown in Fig. 6.27 (from [416]). Key features of
the process are:
- Reforming with an excess of process air. Steam to carbon ratio at inlet
reformer approximately 2.8. Relatively high methane leakage ( ~ 1% dry)
from the secondary reformer. Addition of part of the process steam via a
saturator. Combustion air preheat.
- Process air compressor driven by a steam turbine (all steam produced in the
.-------<.~ CO 2 product
.s
~
CII
2G>
c: .0
CII
~ Ja
C02removat
methanation
>
§o
t:l
Pi·
'"C
...o
0-
§.
o·
t:l
...o'"C
~
1 • I: :: 1and drying I.. ~ '"
r.l
rec ered ammonia '"
ammonia svnl SIS NH3 product
tv
00
....,
Fig. 6.27. The lei AMV process. Simplified process flow diagram (from [416])
284 Ib Dybkjaer
. plant goes to this turbine) with a turbo-alternator on the same shaft. All other
power consumers, including synthesis gas compressor, use electric motors.
- Conventional high temperature and low temperature shift conversion.
- Carbon dioxide removal by the Selexol process; compression to synthesis
pressure before final purification.
- Final purification by methanation and drying at synthesis pressure. After
drying, the gas is mixed with recycle gas and off-gas from a cryogenic purge
gas recovery unit. The mixed gas goes via a recirculation compressor and a
gas-gas heat exchanger, the hot exchanger, to the synthesis converter. Be-
tween the recirculator and the exchanger, a part of the gas is taken to a
cryogenic purge gas recovery, where inerts are removed and the gas composi-
tion adjusted by removal of excess nitrogen.
- Synthesis at about 90 kg/cmZg in a three-bed converter with quench between
the two first beds and indirect cooling (converter feed preheat) between the
second and third bed.
As in the Braun process, the size of the primary reformer is reduced and the size
of the air compressor increased in the AMV process compared to more
conventional process schemes. This is due to the operation with excess process
air and with high methane leakage. Power consumption in the synthetic gas
compressor is low because of the low synthesis pressure and the low suction
temperature (gas direct from the low temperature Selexol COz removal unit),
but this is compensated by increased power consumption for the compression of
excess nitrogen and high power consumption in the refrigeration section.
The AMV process scheme may be considered as a variation of the classical
scheme with some unique features. It has been used as described in the design of
{)ne plant and- with some modifications to accommodate integration with a
urea plant-in another plant. Experience from operation of a plant using the
process has been reported in [417]. The expected energy consumption has been
given as less than 7.0 Gcal/MT of ammonia (for a plant located in a cold
climate), but actual data are not available.
The LeA process was announced by ICI in the late 1980s [350-357,
924-927, 951]. A simplified ftowsheet is shown in Fig. 6.28 (from [350J). Key
features of this process are:
- Reforming in gas-heated reformer (unfired, heat supplied by heat exchange
with product gas from the secondary reformer) at a steam to carbon ratio of
2.5. Process steam supplied via a saturator. Large excess of process air; high
methane leakage (1 %). Process air preheat in fired heater.
- One-stage shift conversion in a cooled reactor; exit temperature approxim-
ately 265°C.
- Gas purification: PSA with adjustment of the hydrogen to nitrogen ratio.
Carbon dioxide removal (optional) in a low pressure MDEA wash on the off-
gas from PSA unit. Methanation upstream of the synthesis gas compressor.
- Synthesis in a low pressure loop (approx. 80 kg/cmZg) with a tubular conver-
ter (TV A type).
air
30 De
gas driers
2582 kmol/h
Proprietary items are the gas-heated reformer [117, 889], the PSA unit [765]
and the synthesis converter. The process scheme as such is unique. It is,
according to literature, suitable only for small-capacity plants. The process has
so far only been used in one installation comprising two units, each with a
capacity of 450 MTPD.
The amount of process air is adjusted so that sufficient heat is available in
the effluent from the secondary reformer to supply all heat required for the
primary reformer; the excess of process air is approximately 75% compared to
the conventional process. The extra power for compression ofthis excess air and
the loss of hydrogen in the PSA unit increase the energy consumption. On the
other hand, firing in the plant is greatly reduced, and the natural gas consump-
tion is low, only 6.44 Gcal/MT ammonia. The process produces very little steam
(the only boiler is a 60 kg/cm 2 g boiler in the synthesis loop), and extra power
must be supplied from outside battery limits for compressor drivers etc. It is
suggested to raise the required power in a gas-fired, combined cycle power plant
with high efficiency. If this system-which is not incorporated in the operating
plants-is considered, a net energy consumption of 6.95 Gcal/MT ammonia is
claimed. The energy consumption in the actual plants is reported to be
7.6 Gcal/MT [952].
M. W. Kellogg has played a major role in the development of the modern
ammonia industry (see Sect. 6.5.3.2.1). Their most recent Low Energy Process
[37,419,766-768,960] represents a further development of the classical process
maintaining basically the same process scheme, but introducing updated tech-
nology and energy saving features in the various process steps. A simplified
process flow sheet is shown in Fig. 6.29 from ([419]). Key features of the process
are:
- Reforming with the stoichiometric amount of process air (i.e. the amount of
air required to give a hydrogen to nitrogen ratio of 3.0 at the ammonia
synthesis converter inlet without dedicated installations for adjustment of the
gas composition). Steam to carbon ratio about 3.3; low methane leakage
(approximately 0.3% dry).
- Air compressor driven by a steam turbine; preheat of process air to a very
high temperature in the reformer waste heat section.
- Gas purification by conventional shift; carbon dioxide removal normally by
Selexol or Benfield; methanation; drying of synthesis gas with molecular
sieves.
- Synthesis at 140 kg/cm 2 g or 180 kg/cm 2 g (depending on plant capacity) in a
"dry" synthesis loop (make-up gas added after ammonia separator) using a
two-bed horizontal converter with indirect cooling by gas/gas exchange.
Refrigeration in so-called unitized heat exchanger. Purge gas recovery.
Proprietary items are the primary reformer, the synthesis converter, and the
unitized heat exchanger. About 6 plants have been designed using various
versions of this technology. (The total number of ammonia plants designed by
M. W. Kellogg is much higher). Operating experience from a plant using the
to process steam
high temperature
shift converter
heat
recovery
condensate to
cooling boiler feedwater
process~ ;, system
steam
ammonia
to I ~t
'atmosphere to c;:ondensate
stripper
i
o
2.
po
cooling purge gas
to fuel ~
c.
~
'"
o·
:;
::,0
o
@
lean pump ~
Fig. 6.29. M. W. Kellogg's low energy process. Simplified process flow diagram (from [419])
N
00
-..l
288 Ib Dybkjaer
Selexol process for carbon dioxide removal and located in a cold climate is
reported in (766). Energy consumption is reported to be below 29GJ (LHV)jMT
of ammonia.
M. W. Kellogg has also suggested a process (the KAAP process) based on a
new catalyst [928, 949]. This technology has been used in a revamp project
where a new reactor was installed downstream of the existing reactor in an
ammonia synthesis loop (see Sect. 6.4.3.4). In addition, more radically new
process schemes deviating from the traditional route have been described. An
example is a scheme based on so-called parallel reforming [30, 32, 120, 121J and
a low pressure loop with ammonia recovery by water absorption [769]. None of
these new developments have been implemented in practice.
Haldor Tops0e AjS has supplied technology for ammonia production for
more than 30 years. Unlike the other process licensors, Tops0e is also a catalyst
supplier, and the development of catalysts and process has gone hand in hand.
The most modern, the Tops0e low energy ammonia process is, as is the M.W.
Kellogg low energy process, based on the traditional process layout with new
energy saving developments introduced in the individual process steps. Descrip-
tions are given in [28, 29, 32, 33, 37, 246, 595, 770, 929]. A simplified flow sheet is
shown in Fig. 6.30 (from [246J). The process is available in two main versions,
the main difference being the steam to carbon ratio at the reformer inlet. The
first version (with a steam to carbon ratio about 3.3) has been used in about 15
industrial plants. The second version (with a steam to carbon ratio about 2.5 and
primary shift
desulphurization reformer secondary conversion CO 2 removal methanation
reformer
air
steam
natural gas
fuel
product ammonia
main
compressor '--------....=.:.:.=:..::=='-'
Fig. 6.30. Haldor Tops0e's low energy process. Simplified process flow diagram (from [246])
Ammonia Production Processes 289
other energy saving features} is not yet proven in any industrial plant. But each
process step is individually proven, and two plans based on the concept are
under construction.
Key points in the two versions of the Tops0e process are:
First version:
Reforming with stoichiometric air and low methane leakage. Steam to carbon
ratio about 3.3.
- Process air compressor driven by a stream turbine (or gas turbine).
- Gas purification: conventional shift; carbon dioxide removal by Benfield or
Vetrocoke; methanation.
- Synthesis at 220 or 140 kgjcm 2 g with Tops0e 2-bed radial flow converter (S-
200).
Second version:
- Reforming with stoichiometric air and low methane leakage. Steam to carbon
ratio 2.5.
- Process air compressor driven by a steam turbine (or gas turbine).
- Gas purification: shift conversion with medium and low temperature cata-
lysts, both copper-based; carbon dioxide removal by MDEA or Selexol;
methanation.
- Synthesis at 140 kgjcm 2 g with Tops0e 2-bed and I-bed radial flow converters
(S-250).
Proprietary items are the primary reformer, the ammonia synthesis converter,
and certain catalysts. Operating experience in plants designed by Tops0e has
been described in [720, 771-774, 956, 966, 967]. In [967] a net energy consump-
tion of 6.97 GcaljMT ammonia was reported for a plant located in a warm
climate. For the most energy efficient process concept, which has so far not been
demonstrated in an industrial installation, a net consumption of 6.67 GcaljMT
ammonia is claimed for a stand alone plant [920].
Uhde GmbH is one of the pioneers in the ammonia industry. Even before the
Second World War they constructed ammonia plants (based on the Mont Cenis
process). Later they built plants in cooperation with many other companies,
such as 6sw, Tops0e, and ICI.
Uhde has recently commercialized a new low energy process scheme [41,
585, 775, 930]. A simplified process flow is shown in Fig. 6.31 (from [41]). The
process scheme is, like the Tops0e and Kellogg process schemes, a further
development of classical process schemes. Key features of the process are:
- Reforming at a slightly increased pressure with stoichiometric air and high
methane leakage ( :::::; 0.6%). Steam to carbon ratio about 3.0.
- Process air compressor driven by a steam turbine (or gas turbine).
- Gas purification: conventional shift; carbon dioxide removal by MDEA;
methanation.
- Synthesis at approximately 180 kgjcm 2 g in two converters, a two-bed and a
one-bed radial flow converter with steam production between the beds.
290 Ib Dybkjaer
combustion
aIr
process air
HP-steam
superheated
BFW
C02
s ngas compressor
HP-steam to
superheater
product NH3
Fig. 6.31. Uhde's low energy
fuel process. Block diagram (from
L!!1C!!~~J---"""'"""- [585])
Proprietary items are the primary reformer and the synthesis converters. Two
plants are operating with a design according to this process scheme, while a
third is under construction. An expected net energy consumption of 6.67-7.20
GcaljMT ammonia is claimed, depending on the conditions.
As mentioned above, several other companies have proposed low energy
process schemes, some based on the conventional process route, some represen-
ting a more innovative approach. In the following the most important of these
process schemes will be briefly discussed.
The Exxon Chemical Low Energy Ammonia Process has been used in one
plant. Key features of the process are:
Reforming with stoichiometric air and low methane leakage. Steam to carbon
ratio about 3.3.
- Air compressor driven by a gas turbine. Exhaust from the gas turbine to the
reformer as hot combustion air.
Conventional shift conversion (with low temperature shift guard vessel);
carbon dioxide removal with the Catacarb process; methanation.
Drying of synthesis gas with molecular sieves installed between the first and
second compressor stages.
Ammonia Production Processes 291
- Ammonia synthesis at 140 bar with a two-bed radial flow converter with
indirect cooling. Make-up gas addition after the separator.
- Cryogenic purge gas recovery.
The energy consumption of the process is given as 29 GJ/MT (6.93 GcaljMT)
ammoma.
The Exxon Chemical process is a good example of processes which are
designed for a specific case in cooperation between the plant owner (Exxon
Chemical), the contractor (Bechtel), and process licensors (Haldor Tops0e for
the synthesis loop and converter). The process and experience from operation of
the plant is described in [776, 777].
The Fluor Ammonia Process [28, 30, 32, 298, 778] is based on a conventional
process scheme. The main features are:
- Use of Fluor's proprietary polypropylene carbonate process (see Sect. 6.3.4)
for carbon dioxide removal. It is suggested [778] that carbon dioxide removal
and methanation be installed after synthesis gas compression.
- Use of heat downstream of the low temperature shift converter for absorption
refrigeration which is used in the ammonia recovery section. In conventional
plants this low level heat is used for reboiling in the carbon dioxide removal
process.
- Air compressor driven by a gas turbine; exhaust from the gas turbine to the
reformer as hot combustion air.
- Primary reforming in a special combination of adiabatic reactors and rehea-
ting elements [125, 126].
It is claimed that the above modifications will reduce the energy consumption
from 32.0 x 199 Btu/short ton (8.89 GcaljMT) to 29.5 x 109 Btu/short ton
(8.19 Gcal/MT) of ammonia [778].
Foster Wheeler's AM 2 Process [779-781] belongs to the group of processes
using an excess of process air to the secondary reformer, so that a specific
process step is required to separate excess nitrogen from the synthesis gas. Key
features of the process are:
- Only 20-50% of the total hydrocarbon feedstock is treated in the primary
reformer which operates at high pressure and a relatively low temperature.
- The remaining feedstock is, together with product gas from the primary
reformer, reacted in the secondary reformer with a large excess of process air.
Methane leakage from the secondary reformer is high, 2.5 vol %.
- Gas purification is by conventional shift conversion followed by a physical
carbon dioxide removal (e.g. Selexol) and final purification in a cryogenic
unit. The hydrogen to nitrogen ratio of the synthesis gas at the cryogenic unit
inlet is 1.0-1.8.
- Synthesis loop conditions have not been specified.
It is claimed that the energy consumption in the process may be as low as 25.1
x 109 Btu/short ton of ammonia (7.0 GcaljMT).
292 Ib Dybkjaer
F oster Wheeler has also suggested a process where the excess of process air is
so high that the primary reformer is deleted. This process (see Sect. 6.5.3.3) can
operate on all types of hydrocarbon feedstock including feedstocks which can
not be used as feed for a primary reformer.
Humphreys & Glasgow have suggested a number of process schemes i.e. the
MDF-process, the BYAS-process, and the LEAD-process.
The MDF-process [28, 30, 32, 341-343,782] consists basically of a hydrogen
plant with PSA purification, an air separation plant for production of nitrogen,
and a synthesis loop converting the hydrogen and nitrogen to ammonia. Key
features of the process are:
- Steam reforming (primary reforming only) at a pressure of about 25 bar.
- Purification of hydrogen by high temperature shift conversion and Pressure
Swing Absorption. Optional carbon dioxide-removal located upstream of the
PSA unit in case carbon dioxide is required.
- Mixing of the pure hydrogen with pure nitrogen which is imported or
produced in an air separation plant.
- Compression and synthesis in an inert-free, dry synthesis loop.
Energy consumption including energy for air separation is given as 32.8
GJjMT of ammonia (7.84 GcaljMT). Significant savings in capital costs are
claimed [782].
The BYAS-process [783-785] differs from the conventional process scheme
mainly in the reformer section and in the final gas purification. Key features of
the process are:
- A significant part of the process feed is sent direct to the secondary reformer
bypassing the primary reformer.
- An excess of air is used in the secondary reformer; the amount of excess
depends on the accepted methane leakage, which may be relatively high.
- Conventional shift conversion and carbon dioxide removal.
- Final purification and adjustment of the hydrogen to nitrogen ratio in a
cryogenic unit.
- Synthesis in an inert-free, dry synthesis loop.
Energy consumption values from 8.0 Gcal/MT [32] to 7.06 Gcal/MT [28]
are quoted.
Projects and Development India Limited (PDIL) has proposed a scheme using
enriched air as process air [789,790]. The scheme is very similar to other modern
low energy concepts with the exception of the use of enriched air. Key features
are:
- Air separation to produce pure oxygen and nitrogen; nitrogen is used as plant
inert gas or exported.
- Primary reforming at elevated pressure, about 45 kg/cm 2 g, and low exit
temperature, about 725°C.
- Secondary reforming using enriched air with 24-30 vol% oxygen. Low
methane leakage.
- Air compressor driven by a gas turbine; compressed air for oxygen produc-
tion may be drawn from an inter mediate stage of the air compressor. Exhaust
from the gas turbine used as hot combustion air for the primary reformer.
- Conventional high and low temperature shift conversion; carbon dioxide
removal preferably by a physical absorption process; methanation.
- Compression and synthesis at 220 kg/cm 2 g using a two-bed radial flow
converter.
Energy consumption for the process scheme is given [789] as 7.40 Gcal/MT of
ammonia.
PDIL has also proposed a process [791] where so much oxygen is added to
the process air that the primary reformer duty is reduced to zero, i.e. the process
uses an auto thermal reformer only in the reforming step.
Ammonia Production Processes 295
. of steam required for the process. The reduced flow inside the reformer tubes
will result in a reduced firing required for the process. The methane leakage
will increase slightly due to the reduced steam to carbon ratio, but the overall
results is typically an energy saving of about 0.20 GcaljMT ammonia when
the steam to carbon ratio is reduced from 4 to 3.
Operation of the reformer at a reduced steam to carbon ratio requires
modification of the downstream shift (if the steam to carbon ratio goes below
the critical level for by-product formation) and carbon dioxide removal
sections in order to fully obtain the potential energy savings. The reduced
steam to carbon ratio will also require a highly active reforming catalyst in
order to eliminate the risk of carbon formation and resulting hot bands on the
reformer tubes.
~ Introduction of combustion air preheat will reduce the amount of fuel
required in the reformer by using the low-level energy in the fuel gas. When
the stack temperature is reduced from about 200°C to 100 °C, the corres-
ponding energy saving is about 0.10 GcaljMT ammonia. The lower temper-
ature of the fuel gas may cause corrosion problems in the form of condensing
sulfuric acid, if the sulfur content of the fuel is high. This problem may be
eliminated by desulfurizing the fuel gas together with the process feed gas. The
energy available in the preheated, pressurized fuel gas may be recovered in an
expansion turbine, which will reduce the pressure to that required in the fuel
gas header system.
~ The feed gas saturator is used to transfer the low-level energy in the fuel gas
into process steam. Hot boiler feed water or process condensate is contacted
with cold natural gas in the saturator, whereby the natural gas is preheated,
and, at the same time, saturated with water. The water leaves the saturator,
make-up water or condensate is added and the water is then reheated in a coil
in the reformer waste heat section. About 20% of the required process steam
may be taken up by the natural gas in the saturator. The energy saving is
typically around 0.10 GcaljMT ammonia. The saturator also serves as a
process condensate stripper whereby overall water consumption may be
reduced, and a pollution problem may be solved. The choice between
combustion air preheat and saturator depends on local conditions. Com-
bustion air preheat saves fuel directly whereas the saturator saves steam.
~ An adiabatic pre-reformer is an attractive revamp option, especially when the
aim is to minimize steam export and/or to increase plant capacity in cases
where the primary reformer is the bottleneck.
In the typical reformer waste heat section, the process feed is preheated to
about 520°C before entering the reformer tubes, where reforming takes place
using the additional heat supplied by the burners. The burner fuel could be
reduced, if the process feed were preheated to higher temperatures, but this
may not be feasible due to the increased potential for carbon formation by the
cracking of hydrocarbons. By passing the preheated process feed through an
adiabatic pre-reformer, the higher hydrocarbons in the feed will react
Ammonia Production Processes 297
When the steam to carbon ratio in the reforming section is reduced, the
conditions in the shift section must be carefully evaluated. If the steam to dry
gas ratio becomes too low, severe problems may arise due to conversion of the
iron oxide in the high temperature shift catalyst to iron carbide, which will
promote formation of undesirable by-products (hydrocarbons and oxy-
genates) (see Sect. 6.3.3.1).
When the steam to carbon ratio in the reforming section is not reduced
below the critical limit for carbide formation, it is possible, by adding an extra
shift catalyst bed after the conventional shift unit, to reduce the carbon
monoxide leakage to around 0.05 dry vol%. This means that less hydrogen
will be lost by reaction with carbon monoxide in the methanator, and that the
synthesis loop will become more efficient due to the lower inert content in the
make-up gas.
Several revamp options are available for modification of the carbon dioxide
removal section depending on the type of carbon dioxide removal process.
The processes mostly used in ammonia plants are chemical absorption
processes based on either hot potassium carbonate (HPC) such as Benfield, or
Vetrocoke, or amine solutions such as MEA. The chemical carbon dioxide
removal processes may be improved or replaced with a physical process in
which the absorbent is regenerated by simply flashing off carbon dioxide. In
this way the need for regeneration heat may be reduced or eliminated. A
physical carbon dioxide removal system may result in energy savings of
0.01-0.35 GcaljMT ammonia.
In certain cases methanolation may be an attractive revamp option
[907,908]. In this process, the process gas after carbon dioxide removal is fed
to a methanol synthesis reactor instead of a methanation reactor. Residual
carbon oxides are converted to methanol, which can be recycled to the
reformer on the process side, whereby hydrogen loss is avoided, and the inert
content of the make-up gas is reduced. An overall reduction in reformer feed
consumption of 0.15-0.30 Gcal/MT NH3 can be obtained depending on
plant layout. There will, however, be an increased consumption of reformer
fuel to compensate for the reduced heating value of the loop purge gas caused
by the reduced methane content. The total saving in energy consumption is
therefore somewhat smaller, in the range of 0.05-0.10 GcaljMT NH 3 . As an
alternative to recycling of methanol, it can be recovered as a product which is
exported or converted to formaldehyde or ureaform, which can be used for
coating of urea to prevent caking.
298 Ib Dybkjaer
A B
H2 S
fuel gas
Fig. 6.32. Two classical process schemes for production of ammonia from fuel oil (or coal). (from
[818J)
a Scheme with Shell gasification (or other process with cooling of raw gas to near ambient
temperature)
b Scheme with Texaco gasification (or other process with soot removal at elevated temperature)
two stages, using a special high activity sulfur tolerant shift catalyst. After the
shift conversion, hydrogen sulfide and carbon dioxide are removed in an acid gas
removal unit; the remaining traces of carbon oxides are removed by methan-
ation; and ammonia is synthesized in a normal synthesis loop using the Topsoe
S-200 two bed radial flow converter. A synthesis gas generation process similar
to this scheme and using the same type of catalysts-but based on gasification
with oxygen-is used in a large ammonia plant in Japan [213,214].
Case a above corresponds to Eq. (17). It is seen that in this case, which is
based on the use of pure methane as feed, on 100% conversion of the feed and
100% yield of ammonia from hydrogen and nitrogen and of carbon dioxide
from carbon and oxygen, there is a deficit of carbon dioxide of 13 % compared to
the amount required for full conversion of ammonia to urea.
_Case b illustrates that if the feed contains heavy hydrocarbons correspond-
ing to an overall hydrogen to carbon atomic ratio slightly above 3.0, then the
stoichiometric production of carbon dioxide may automatically be achieved.
Case c shows a solution where the extra carbon dioxide is obtained by
production of excess synthesis gas. The excess synthesis gas can be extracted as a
side stream before the synthesis gas compressor and used as fuel in the reformer
or exported as such or as hydrogen after purification. Alternatively, it can be
compressed and passed through the synthesis loop which will then operate with
a high purge rate and correspondingly low inert level. The actual solution will
depend on the specific case, especially on the composition of the feedstock.
Solutions of this type are used by the process schemes (see Sect. 6.5.3.2.3) using a
stoichiometric amount of process air, such as the Kellogg, Topsoe and Uhde
processes.
Case d shows that stoichiometric carbon dioxide production can also be
achieved by using an excess of process air. However, in this case it becomes
necessary to introduce an extra gas separation process step in order to remove
the excess nitrogen which is introduced with the excess air. As explained in
Sect. 6.5.3.2.3, several process schemes exist where excess process air is used, and
the excess nitrogen is removed in dedicated units. Processes using excess process
air are the Braun process, the ICI processes, and several process schemes of
minor importance.
302 Ib Dybkjaer
without partial by-pass of the shift unit and/or the carbon dioxide removal unit
to adjust the ammonia to methanol production ratio. Conversion of existing
ammonia plants to combined production of ammonia and methanol is de-
scribed in [935]. Integration of ammonia production with the production of
steel has also been described [380-384]. Production of ammonia from off gases
from ethylene plants or methanol plants has also been used [378,379]. Integ-
rated production of ammonia and methanol by modifying the methanol process
by incorporating an oxygen-fired secondary reformer and producing ammonia
from the off-gas and nitrogen from the air separation unit is discussed in [968].
A plant producing both methanol and ammonia from synthesis gas produced by
partial oxidation of heavy oil is described in [166-168].
Facilities for storage of ammonia are required at the production facilities either
as buffer capacity to smooth out variations in production and demand between
the ammonia plant and downstream units using ammonia as raw material or to
hold product between shipments from plants exporting ammonia. Required
storage capacity may range from less than one to a few days' production
capacity for intermediate storage and up to one month's production capacity, or
even more, for exporting plants located in remote areas. This means that
ammonia storage tanks in ammonia plants may range in size from less than
100 t to above 40000 t.
In addition to this, very large ammonia storage facilities have in recent years
been constructed at terminals served by ammonia pipelines. Pipeline transport
of ammonia and related storage facilities are discussed in [830,831]. In the
construction and operation of ammonia storage facilities the special properties
of ammonia-an inflammable and toxic compound-must be taken into account.
The properties of ammonia and the hazards involved in handling and storage of
this product are dealt with in chapter 7. General descriptions of the design of
storage facilities are given in [832,833]. Surveys of actual installations may be
found in [834-840, 936]. Safety recommendations are given in [841-843] (see
also chapter 7).
Depending on the capacity, ammonia may be stored by three different
methods:
- In non-refrigerated pressure vessels-perhaps cooled by water spray and/or
painted with reflecting paint. The storage pressure corresponds to ammonia
vapour pressure at ambient temperature. The storage vessels, often referred to
as "bullets", are usually cylinders; capacity is generally low, below about
200 t. Non-refrigerated spherical pressure vessels with capacities up to 1500 t
ammonia have been used [833].
304 Ib Dybkjaer
r---'I
I
•
I
ammonia
manufacturing
plant
;:
o
.~
-oE
Sa 1----\:><1-- pipe line fill
1----\:><1-- rail tank car
I----,../r-----t><l-- fill
L.o----J~-i><3-- highway
f---t:<1--- truck fill
r----'
I I ship or
river barge
dock
M
Fig. 6.33. Generalized scheme for supply of ammonia from manufacturing plant to various
transport facilities (from [831])
This type of storage, as mentioned above, is most often used for storage of
relatively small quantities of ammonia as intermediate storage between the
ammonia plant and ammonia consumers or transport systems accepting warm
ammonia. Pressurized tanks are also used extensively in the distribution system
for storage of small quantities of ammonia, for example in field tanks in
connection with direct application of liquid ammonia as a fertilizer. Non-
refrigerated pressurized storage tanks may be protected against overheating in
hot weather by insulation, by water spray, or by circulating ammonia through a
water cooler. The design pressure is typically around 18 bar.
The design of pressurized storage tanks and related safety aspects are
reviewed in [845]. Stress corrosion cracking is a much debated risk in pre-
ssurized storage tanks. The mechanism of stress corrosion cracking (including
the effect of oxygen and water) is discussed in [846-850, 937] and, in an
authoritative review, in [938]. It has been generally accepted that the addition of
water to ammonia inhibits stress corrosion cracking [832,846], and in certain
situations (transport of ammonia in containers made from certain specified
306 Ib Dybkjaer
purger
condensate
receiver
storage
f
~.
vapor return line
tank
I
§
heater r
'"
Fig. 6.34. Typical fully refrigerated ammonia storage unit. Simplified process flow diagram (from [832])
~
308 Ib Dybkjaer
Models for tank behaviour and potential hazards during tank loading are
described in [945].
Retrofit of existing tanks for added safety is discussed in [870]. Special design
and operating considerations relating to the influence of climatic conditions are
reported in [871]. Safety aspects including maintenance and inspection proced-
ures are reviewed with reference to specific cases in [872-875, 946-948]. For
risk of stress corrosion cracking in atmospheric ammonia storage tanks see
Sect. 6.6.1.
6.7 References
1. Harding AJ (1959) Ammonia manufacture and uses, Oxford University Press, London
2. Vancini CA (1961) La Sintesi dell'Ammoniaca. Ulrico Hoepli, Milano
3. Sauchelli V, Stevenson FJ, Timm B, Danz W, Axelrod LC, O'Hare TE, Hansen HJ (1964) In:
Sauchelli V (ed) Fertilizer nitrogen its chemistry and technology. Reinhold, New York
4. Vancini CA (1971) Synthesis of ammonia. Macmillan, London
5. LeBlanc JR Jr, Madhavan S, Porter RE (1978) In: Kirk-Othmer Encyclopedia of Chemical
Technology, 3rd ed. Wiley, New York, vol 2, p 470
6. Slack AV, James GR (eds) Ammonia. Marcel Dekker, New York, vol. 1 (1973) vol 2 (1974) vol 3
(1977) vol 4 (1979)
7. Strelzoff S (1981) Technology and manufacture of ammonia. Wiley, New York
8. Bakemeier H, Huberich T, Krabetz R, Liebe W, Schunck M, Mayer D (1985) In: Ullman's
Encyclopedia of Industrial Chemistry, 5th edn, vol A2, VCH, Weinheim, p 143
9. Noyes R (1964) Ammonia and synthesis gas 1964. Noyes Development Coporation, New York
10. Noyes R (1967) Ammonia and synthesis gas 1967. Noyes Development Coporation, New York
11. Brykowski FJ (ed) (1981) Ammonia and synthesis gas recent and energy-saving processes,
Noyes Data Corporation, Park Ridge, p 1
12. Mittasch A (1951) Geschichte der Ammoniaksynthese. Verlag Chemie, Weinheim
13. Timm B (1960) Chern Ind (London) 12: 274
14. Timm B (1963) Chern Ing Tech 35: 817
15. Timm B, Danz W (1964) In: Sauchelli V (ed) Fertilizer nitrogen its chemistry and technology.
Reinhold, New York, p 40
16. Appl M (1976) Nitrogen 100: 47
17. Slack AV (1973) In: Slack AV, James GR (ed) Ammonia, Marcel Dekker, New York 1: 5
18. Waeser B (1939) Chern Fabr 12(31, 32): 370
19. Brown CO (1954) Chern Eng Prog 50(11): 556
20. Nitrogen (1962) 16: 35
21. Strelzoff S, Vasan S, (1963) Chern Eng Prog 59(11): 60
22. Bresler SA, James GR (1965) Chern Eng (NY) 72(13): 109
23. Allen JB (1965) Chern Process Eng (London) 46(9): 473
24. Mayo HC, Finneran JA (1968) Erdol Kohle, Erdgas, Petrochem 21(7): 404
25. Quartulli J, Wagener D (1973) Erdol Kohle, Erdgas, Petrochem Brenns Chern 26(4): 192
26. Lyon SD (1975) Chern Ind (London) 17: 731
27. Quartulli OJ , Buividas LJ (1976) Nitrogen 100: 60
28. Pachaiyappan V (1979) Fert News (August) 41
29. CEER (1979) Chern Econ Eng Rev 11(5): 24
30. Zardi U, Antonini A (1979) Nitrogen 122: 33
31. Honti GD (1981) In: More AI (ed) Fert nitrogen: Proc Br Sulphus Corp Int ConfFert Technol,
4th, part 1, 1
32. Saviano F, Lagana V, Bisi P (1981) Hydrocarbon Process 60(7): 99
33. Rai HS (1982) CEER Chern Eng World 17(1): 69
34. Brown FC (1983) Proc Fert Soc London 218: 1
35. Dybkjaer I, (1983) ECN Eur Chern News Fertilizers '83 Supplement (February 21): 15
36. Picciotti M, Pocini CA (1984) Chim Ind (Milan) 66(2): 97
Ammonia Production Processes 309
88. Rostrup-Nielsen JR (1982) In: Figueiredo JL (ed) Progress in catalyst deactivation. Martinus
Nijhoff Publishers, The Hague, p 209
89. Davies J, Lihou DA (1971) Chern Process Eng (London) 52(4): 71
90. Marsch HD, Herbort HJ (1982) Hydrocarbon Process 61(6): 101
91. Nitrogen (1987) 166: 24
92. Nitrogen (1987) 167: 31
93. Leyel CP (1975) Ammonia Plant Saf 17: 41
94. Thuillier J, Pons F (1978) Ammonia Plant Saf 20: 89
95. Kawai T, Takemura K, Zaghloul MB (1984) Paper presented at International Plant Engr Conf,
Bombay
96. Kawai T, Takemura K, Shibaksi T, Rump L, Danielsen B, Wrisberg J (1982) Paper presented at
AIChE Ammonia Saf Symp. San Francisco
97. Muhlenforth CJ (1989) Finds 4(2): 14
98. Salot WJ (1972) Ammonia Plant Saf 14: 119
99. Salot WJ (1973) Ammonia Plant Saf 15: 1
100. Konoki K, Shinohara T, Shibata K, (1982) Plant/Oper Prog 1(2): 122
101. Imoto Y, Terada S, Maki K (1982) Plant/Oper Prog 1(2): 127
102. Kawai T, Takemura K, Shibasaki T, Mohri T (1982) Plant/Oper Prog 1(3): 181
103. Kawai T, Takemura K, Shibasaki T, Mohri T (1980) Ammonia Plant Saf 22: 119
104. Kawai T, Mohri T, Takemura K, Shibasaki T (1984) Ammonia Plant Saf 24: 131
105. Cockerham RG, Percival G (1957/58) Trans Inst Gas Eng 107: 390
106. Jockel H, Triebskorn BE (1973) Hydrocarbon Process 52(1): 93
107. Rail W, (1967) Erd61 Kohle, Erdgas, Petrochem 20(5): 351
108. Davies HS, Humphries KJ, Hebden D, Percy DA (1967) IGE J (October): 708
109. Jockel H (1969) GWF Gas Wasserfach 110(21): 561
110. Ishiguro T (1968) Hydrocarbon Process 47(2): 87
111. Rostrup-Nielsen JR, Tottrup PB (1979) In: Proc Symp Sci Catal its Appl Ind, Sindri, p 379
112. Clark DN, Henson WGS (1988) Ammonia Plant Saf 28: 99
113. Mosley T, Stephens RW, Stweart KD, Wood J (1972) J Catal 24: 18
114. Stahl H, Rostrup-Nielsen J, Udengaard NR (1985) In: Fuel cell seminar 1985. Tuscon, Arizona,
p 83
115. Udengaard NR, Christiansen LJ, Summers WA (1988) Endurance testing of a high-efficiency
steam reformer for fuel cell power plants. EPRI AP-6071, Project 2192-1. Electric Power
Research Institute, California.
116. Sederquist RA (1978) US 4071330; (1978) CA 89: No. 8711
117. Smith ASP, Doy RJ, Limbach APJ (1986) EP 194067, CA (1986) 105: No. 229260.
118. Ruziska PA (1984) EP 113198; CA 101: No. 113278
119. Pinto A (1988) US 4750986; see EP 124226 CA (1985) 102: No. 27497
120. Crawford DB, Becker CL, LeBlanc JR (1978) US 4079017; CA (1979) 91: No. 213715
121. Crawford DB, Becker CL, LeBlanc JR (1979) US 416229. see BE 856919 CA (1978) 89:
No. 179566
122. Miyasugi T, Kosaka S, Kawai T, Suzuki A (1984) Ammoonia Plant Saf 24: 64
123. Pagani G, Brusasco G, Gramatica G (1981) In: More AI (ed) Fert nitrogen: Proc Br Sulphur
Corp Int Conf Fert Technol, 4th, part 1, p 195
124. Nitrogen (1989) 179: 16
125. Bogart MJP (1973) US 3743488; CA (1973) 79: No. 106547
126. Bogart MJP (1974) US 3795485; CA (1974) 81: No. 15210
127. Nobles EJ (1973) In: Slack AV, James GR (eds) Ammonia, vol 1. Marcel Dekker, New York,
p275
128. Fuderer A (1987) US 4650651; CA (1987) 106: No. 179566
129. Herbort H-J, Marsch H-D (1987) (GB 2181740; see DE 3532413 CA (1987) 106: No. 158813
130. Marsch H-D, Thiagarajan N (1989) Ammonia Plant Saf 29: 195
131. Noyes R (1967) Ammonia and synthesis gas 1967. Noyes Development Corporation, Park
Ridge, p 78
132. Nitrogen (1962) 17: 35
133. Chern Eng Int Ed (1962) 69(14): 89
134. Chern Eng Int Ed (1966) 73(1): 24
135. Hydrocarbon Process (1984) 63(4): 103
136. Flytzani-Stephanopoulos M, Voecks GE (1981) Energy Prog 1(1-4): 52
137. Reidel JC (1954) Oil Gas J 52(40): 60
Ammonia Production Processes 311
236. Campbell JS (1970) Ind Eng Chern Process Des Dev 9(4): 588
237. Young PW, Clark CB (1973) Chern Eng Prog 69(5): 69 and Ammonia Plant Saf 15: 18
238. Rostrup-Nielsen JR, Hejlund Nielsen PE (1985) In: Diidar J, Wise H (eds.) Deactivation and
poisoning of catalysts. Marcel Dekker, New York, p 259
239. Kitchen D (1988) In: Nitrogen 88 Br Sulphur's 12th Int Conf, Geneva, p 127
240. Pedersen PS (1988) In: Nitrogen 88 Br Sulphur's 12th Int Conf Geneva, p 111
241. Lundberg WC (1979) Chern Eng Prog 75(6): 81 and Ammonia Plant Saf 21: 105
242. Lorenz E (1967) Ind Chim Beige 32: (Spec No.) (Pt. 2) 377
243. Auer W, Lorenz E, Griindler KH (1971) Paper presented at The 68th National Meeting of the
American Institute of Chemical Engineers, Houston, USA
244. Dybkjaer I (1979) In: Ammonia Coal Symp. Tennessee Valley Authority, Muscle Shoals, AL
USA, p 133
245. Dybkjaer I, Bohlbro H, Aldridge CL, Riley KL (1979) Ammonia Plant Saf 21: 145
246. Dybkjaer I (1981) In: More AI (ed) Fert nitrogen: Proc Br Sulphur Corp Int ConfFert Technol,
4th, part II, p 503
247. Hejlund-Nielsen PE, Begild-Hansen J (1982) J Mol Catal17: 183
248. Hejlund-Nielsen PE, Begild-Hansen J (1981) Paper presented at Communicacao ao Coloquio
Nacional de Catalise Industrial, Lisboa
249. Hansen JB, Carstensen JH, Pedersen PS (1989) Ammonia Plant Saf 29: 204
250. Kitchen D, Pinto A, Praag H van (1989) Ammonia Plant sar 29: 212
251. Carstensen JH, Begild-Hansen J, Pedersen PS (1990) Ammonia Plant Saf 30: 139
252. Buckthorp CM (1978) Nitrogen 113: 34
253. Bonacci JC, Otchy TG (1978) Ammonia Plant Saf 20: 165
254. Colby JH, White GA, Notwick PN Jr (1979) Ammonia Plant Saf 21: 138
255. Krishnaswami KR, Neelakantan PS (1989) Fert. News (December) 31
256. Cover AE, Hubbard DA, Jain SK, Shah KV, Koneru PB, Won EW (1985) Review of selected
gas removal processes for SNG production. Gas Research Institute, Contract No. 5082-222-
0754, p 293
257. Christensen KG, Stupin WJ (1978) FE-2240-49, Report to the United States Department of
Energy and Gas Research Institute. Contract No. EX-76-C-OI-2240
258. Kohl AL, Riesenfeld FC (1979) Gas Purification, 3rd ed. Gulf Publishing Company, Houston
259. Schmidt HW, Henrici H-J (1972) Chern Ztg 96(3): 154
260. Goar BG (1971) Oil Gas J 69(28): 75, 69(29): 84
261: Hochgesand G (1968) Chern Ing Tech 40(9/10): 432
262. Werner D (1981) Chern Ing Tech 53(2): 73
263. Christensen KG, Stupin WJ (1978) Hydrocarbon Process 57(2): 125
264. Tennyson RN, Schaaf RP (1977) Oil Gas J 75(2): 78
265. Schmidt HW (1968) Chern Ing Tech 40(9/10): 425
266. Thirkell H (1974) In: Slack AV, James GR (eds) Ammonia, vol 2. Marcel Dekker, New York,
p 117
267. Strelzoff S (1981) Technology and manufacture of ammonia. Wiley, New York, p 193
268. Strelzoff S (1975) Chern Eng (NY) 82(19): 115
269. Stokes KJ (1981) In: More AI (ed) Fert nitrogen: Proc Br Sulphur Corp Int Conf Fert Technol
4th, part II, 525
270. Stokes KJ (1980) Ammonia Plant Saf 22: 178
271. Stokes KJ (1981) Nitrogen 131: 35
272. Brown FC, Leci CL (1982) Proc Fert Soc London 210: 1
273. Jackson JM (1974) In: Slack AV, James GR (eds) Ammonia, vol 2. Marcel Dekker, New York,
p 183
274. Field JH (1974) In: Slack AV, James GR (eds) Ammonia, vol 2. Marcel Dekker, New York,
p 153
275. Butwell KF, Kubek DJ (1977) Hydrocarbon Process 56(10): 173
276. Nitrogen (1975) 96: 33
277. Nitrogen (1976) 102: 40
278. Butwell KF, Kubek DJ, Sigmund PW (1979) Chern Eng Prog 75(2): 75; Ammonia Plant Saf21:
156
279. Pierce JD, Chao EI (1985) Paper presented at Ammonia Saf Symp, Seattle paper 30e, p 1
280. Linsmayer S (1972) Chern Tech 24(2): 74
281. Elberling K, Gabriel W (1977) Chern Tech 29(1): 43
314 Ib Dybkjaer
423. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p324
424. Vancini CA (1971) Synthesis of ammonia. The Macmillan Press Ltd, London and Basingstoke,
p 241
425. Fertilizer Focus (1987) 4(10): 36
426. Tops0e HFA, Poulsen HF, Nielsen A (1967) Chern Eng Prog 63(10): 67
427. Nitrogen (1964) 31: 22
428. Vancini CA (1971) Synthesis of ammonia. The Macmillan Press Ltd, London and Basingstoke,
p 245
429. Finneran JA, Mayo HC (1967) US 3350170; CA (1968) 68: No. 4558
430. Chern Eng (NY) (1967) 74(24): 112
431. Chakley DE, Turner W (1967) Oil Gas Int 7(10): 49
432. Axelrod LC, Quartulli OJ, Turner W (1966) Europe Oil 5(3): 58
433. Pet Process (1956) 10: 167
434. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p303
435. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p 310
436. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p 319
437. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p 341
438. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p 350
439. Spielman M, Baumann GP, Hering B (1968) US 3388968; CA (1968) 69: No. 44970
440. Vancini CA (1971) Synthesis of ammonia. The Macmillan Press Ltd, London and Basingstoke,
p234
440. Vancini CA (1971) Synthesis of ammonia. The Macmillan Press Ltd, London and Basingstoke,
p 237
442. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p 313
443. Mundo K-J, (1972) Chern Anlagen Verfahren 6: 49
444. Slack AV (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
. p 331
445. Vancini CA (1971) Synthesis of ammonia. The Macmillan Press Ltd, London and Basingstoke,
p248
446. Shearon WH Jr, Thompson HL (1952) Ind Eng Chern 44(2): 254
447. Thompson HL, Guillaumeron P, Updegraff NC (1952) Chern Eng Prog 48(9): 468
448. Nitrogen (1987) 169: 31
449. Neth N, Puhl H, Liebe W (1982) Chern Eng Prog 78(7): 69
450. Landrum LH (1984) Ammonia Plant Saf 24: 22
451. Vancini CA (1971) Synthesis of ammonia. The Macmillan Press Ltd, London and Basingstoke,
p 122
452. Bendix H, Lenz L (1989) Ammonia Plant Saf 29: 221
453. Ripps DL (1977) In: Slack AV, James GR (eds) Ammonia, vol 3. Marcel Dekker, New York,
p 235
454. Kjaer J (1966) Chern Tech (Berlin) 18(3): 138
455. Almasy GA, Hay 11, Pallai 1M (1966) Br Chern Eng 11(3): 188
456. Almasy G, Hay J, Jedlovszky P, Pallai, I (1966) Int Chern Eng 6(2): 233
457. Rase HF (1977) Chemical reactor design for process plants, vol 2. Wiley, New York, p 61
458. Kjaer J (1985) Comput Chern Eng 9(2): 153
459. Dybkjaer I (1986) In: De Lasa HI (ed) Chemical Reactor Design and Technology, Nasa ASI
Series, Series E: Applied Sciences No. 110. Martinus Nijhoff Publishers, Dordrecht, p 795
460. Dybkjaer I, Gam EA (1985) Ammonia Plant Sat 25: 15
461. Dybkjaer I, Gam EA (1984) CEER Chern Econ Eng Rev 16(9): 29
462. Jarvan JE (1978) Oil Gas J 76(5): 178
463. Kjaer J (1979) Measurement and calculation of temperature and conversion in fixed-bed
catalytic reactors, 2nd ed Haldor Tops0e A/S, Vedbaek, Denmark
464. Hlavacek V (1970) Ind Eng Chern 62(7): 8
318 Ib Dybkjaer
611. MacLean DI, Prince CE, Chae YC (1980) Ammonia Plant Saf 22: I
672. Schendel RL, Mariz CL, Mak JY (1983) Hydrocarbon Process 62(8): 58
673. Chern Eng (London) (1979) 345: 395
674. Sheridan III JJ, Eisenberg FG, Greskovich EJ, Sandrock GD, Huston EL (1983) J Less
Common Met 89: 447
675. Santangelo JG, Chen GY (1983) CHEMTECH 13(10): 621
676. Duff BS (1955) Chern Eng Prog 51(1): 12-J
677. Mundo K-J (1973) Chern Eng Tech 45(lOa): 632
678. Strelzoff S (1974) Hydrocarbon Processes 53(10): 133
679. Chern Eng (NY) (1974) 81(23): 52
680. Buividas LJ, Finneran JA, Quartulli OJ (1974) Chern Eng Prog 70(10): 21
681. Rothman SN, Frank ME (1975) Ammonia Plant Saf 17: 19
682. Hess M (1976) Hydrocarbon Process 55(11): 97
683. Netzer D, Moe J (1977) Chern Eng (NY) 84(23): 129
684. Waitzman DA (1978) Chern Eng (NY) 85(3): 69
685. Czuppon TA, Buividas LJ (1979) Hydrocarbon Process 58(9): 197
686. Nichols D, Blouin GM (1979) CHEMTECH 9(9): 512
687. Buividas LJ (1981) Chern Eng Prog 77(5): 44; Ammonia Plant Saf (1981) 23: 67
688. Axelrod L (1981) Catal Rev Sci Eng 23(1,2): 53
689. Omori T (1982) CEER Chern Econ Eng Rev 14(9): 7
690. Staege H (1982) Tech Mitt Krupp Werksber 40(1): 1
691. Stratton A, Teper M (1984) The economics of producing ammonia and hydrogen. lEA Coal
Research, London, p 1
692. Holroyd R (1967) Chern Ind (August 5): 1310
693. Bogner H (1965) Nitrogen 34: 32
694. Axelrod L, Daze RE, Wickham HP (1968) Chern Eng Prog 64(7): 17
695. Quartulli OJ, Turner W (1972) Nitrogen 80: 28; (1973) 81: 32
696. Nitrogen (1976) 100: 77
697. Ennis R, Lesur PF (1977) Hydrocarbon Process 56(12): 121
698. Arkley KI, Pinto A (1987) In: FAI Seminar 1986. The Fertiliser Society ofIndia, New Delhi,
p SI/2
699. Pinto A, Trotter SG (1987) Proc Fert Soc London 260
700. Arkley KI, Halstead JM, Pinto A (1988) Ammonia Plant Saf 28: 58
701. Dybkjaer I (1984) In: IFA Technical Conf, Paris, 13-1
702. Denbigh KG (1956) Chern Eng Sci 6(1): 1
703. Riekert L (1974) Chern Eng Sci 29: 1613
704. Gaggioli RA, Petit PJ (1977) CHEMTECH 7(7): 496
705. Gaggioli RA (ed) (1980) Thermodynamics second law analysis ACS Symp Ser 122
706. Sussman MV (1980) Chern Eng Prog 76(1): 37
707. Christiansen LJ (1987) In: CEF 87 Proc XVIII Congo EFCE, Sicily, P 217
708. Vaclavek V, Duy Phi D, Kleinova N (1987) In: CEF 87 Proc XVIII Congr EFCE, Sicily, p 721
709. Silberring L (1971) Chern Ing Tech 43(12): 711
710. Pinto A, Rogerson PL (1977) Chern Eng Prog 73(7): 95
711. Cremer H (1980) In: Gaggioli RA (ed) Thermodynamics second law analysis ACS Symp Ser 122:
111
712. Nimmo NM (1967) Chern Eng (London) 210: CE 156
713. Tops0e H (1966) In: World Power Conf. Tokyo, paper 135 IlIA
714. Finneran JA, Mayo HC, Multharp RH, Smith RB (1969) US 3441393; CA (1969)71: No. 23310
715. Silberring L (1969) Chern Process Eng Heat Transfer Surv (August) Suppl: 74
716. Slack JB (1972) Chern Eng (NY) 79(2): 107
717. Renker W (1972) Chern Tech 24(10): 596
718. Silberring L (1979) Nitrogen 120: 35
719. van Campagne WL (1981) Hydrocarbon Process 60(8): 117; (1982) 61(7): 145; (1983) 67(7): 57
720. Tasrif A, Madsen J (1989) In: Proc Fertilizer Asia Conf and Exhibition, Manila 1989, Br
Sulphur Corp, pISS
721. Quartulli OJ (1975) Hydrocarbon Process 54(10): 94
722. Buividas LJ (1980) Proc Environ Symp, p 365
723. Bignon M (1970) In: Proc Fert Soc London 117: 98
724. Hinchley P (1979) Chern Eng (NY) 86(17): 120
725. Sault RA (1968) Chern Eng Prog 64(3): 57
Ammonia Production Processes 323
890. Schneider RV III: (1991) FAI Seminar 1990, The Fertiliser Association ofIndia, New Delhi, p
SII-l
891. Shiris PJ, Cassata JR, Mandelik BG, van Dijk CP (1984) US 4479925
892. Degand P, Schurmans JP, Julemont V (1992) Ammonia Plant Saf 32: 129
893. Quintana ME, Skinner GF (1991) FAI Seminar 1990. The Fertiliser Association ofIndia, New
Delhi, p SII-2
894. Kumar R, Srinivasan V (1991) Indian Fertilizer Annual Scene p 17
895. Hauser N, Bayens CA (1992) Ammonia Plant Saf 32: 268
896. Nitrogen (1990) 186: 21
897. Roos H, Wanjek H, Sprague M (1990) Ammonia Plant Saf 30: 187
898. Kitchen D, Henson WGS, Madsen JK (1990) Ammonia Plant Saf 30: 105
899. Carstensen JH, B0gild-Hansen J, Pedersen PS (1991) Ammonia Plant Saf 31: 113
900. Bartoo RK, Gemborys TM, Wolf CW (1991) In: Nitrogen 91 Preprints. Br Sulphur Corp
London, p 127
901. Nitrogen (1992) 198: 26
902. Song CC, Wohlgeschalfen KR, Verduijn WD, Qadir A (1992) Ammonia Plant Saf 32: 187
903. Shiyong Z (1991) In: Nitrogen 91 Preprints, Br Sulphur Corp London, p 141
904. Shah VA, Huurdeman TL (1990) Ammmonia Plant Saf 30: 216
905. Hefner W, Herion C, Meissner H, Appl M (1991) FAI Seminar, 1990. The Fertiliser Association
of India, New Delhi, p SII-3
906. Meissner H, Wammes W, Hefner W (1991) In: Nitrogen 91 Preprints. Br Sulphur Corp
London, p 151
907. Nitrogen (1992) 197: 18
908. S0gaard-Andersen P, Hansen 0 (1992) Ammonia Plant Saf 32: 177
909. Rail W (1991) Ammonia Plant Saf 31: 247
910. Rasmussen CV (1991) Nitrogen 194: 30
911. Peterson RB, Finello R, Denavit GA (1984) US 4452760; CA (1985) 102: No. 187456t
912. Grotz BJ (1986) WO Patent Application 86/06058; CA (1987) 106: No. 69646q
913. Clayton KA, Shannahan N, Wallace B (1991) Ammonia Plant Saf 31: 212
914. Campbell JD, Rawlinson RL, WIlson KC (1992) Ammonia Plant Saf 32: 1
915. Huurdeman TL, Schrijen HC (1992) Ammonia Plant Saf 32: 10
916. Prescott GR (1992) Ammonia Plant Saf 32: 217
917. Heuser A (1992) Ammonia Plant Saf 32: 243
918. Wagner GH, Heuser A, Heinke G (1992) Ammonia Plant Saf 32: 252
919. Timbres DH, van Moorsel WH, Johnson DL (1990) Ammonia Plant Saf 30: 200
920. Madsen J (1991) Ammonia Plant Saf 31: 227
921. Patel NM, Kittelstad KJ (1990) Ammonia Plant Saf 30: 208
922. Allen RL Jr, Moser GA (1992) Ammonia Plant Saf 32: 170
923. Grossmann G, Dejaeger J (1992) Ammonia Plant Saf 32: 164
924. Kitchen D, Pinto A (1991) FAI Seminar 1990. The Fertiliser Association ofIndia, New Delhi, p
SIII-l
925. Hicks TC, Pinto A, Moss MJS (1990) Ammonia Plant Saf 31: 219
926. Kitchen D, Pinto A (1991) Ammonia Plant Saf 31: 152
927. Armitage PM, Elkins KJ, Kitchen D, Pinto A (1992) Ammonia Plant Saf 32: 111
928. LeBlanc JR, Shires PJ (1992) Hydrocarbon Technology Int 3: 141
929. Dybkjrer I (1991) FAI Seminar 1990. The fertiliser Assoication ofIndia, New Delhi, p S III-2
930. Hakmann R (1991) FAI Seminar 1990. The Fertiliser Assoication ofIndia, New Delhi, p SIll 3
931. Jungr M (1991) Nitrogen 191: 42
932. Sharma SP, Bhaduri AK (1992) FAI Seminar 1991. The Fertiliser Assoication of India, New
Delhi, p SII/l-l
933. Tsujimoto M, Bendix H, Lenz L, Johannes D (1990) Ammonia Plant Saf 30: 167
934. Tusjimoto M, Johannes D (1991) In: Nitrogen 91 Preprints. Br Sulphur Corp London, p 69
935. Brown F (1992) Finds, First Quarter, p 32
936. Hale CC, Lichtenberg WH (1990) Ammonia Plant Saf 30: 225
937. Lunde L, Nyborg R (1990) Ammonia Plant Saf 30: 60
938. Lunde L, Nyborg R (1991) Proc No. 307. The Fertiliser Society, London, p 1
939. Hewerdine S (1991) Proc No. 308. The Fertiliser Society, London, p 1
940. Appl M, Fassler K, Fromm D, Gebhard H, Porte H (1990) Ammonia Plant Saf. 30: 22
941. Selva RA, Heuser AH, (1990) Ammonia Plant Saf 30: 39
942. Burke BG, Moore DE (1990) Ammonia Plant Saf 30: 91
Ammonia Production Prucesses 327
943. Conley MJ, Angelsen S, WIlliams D (1991) Ammonia Plant Saf. 31: 159
944. Thompson JH (1990) Ammonia Plant Saf 30: 241
945. Tilton IN, Squire RH, Saffle CS, Atkins CR (1992) Ammonia Plant Saf 32: 63
946. Wiltzen RCA (1991) Ammonia Plant Saf 31: 28
947. Squire RH (1991) Ammonia Plant Saf 31: 131
948. Ali SB, Smallwood RE (1991) Ammonia Plant Saf 31: 142
949. Appl M (1992) Nitrogen 199: 46; (1992) 200: 27; (1992) 202: 44
950. Tomasi L (1992) Nitrogen 199: 35
951. Elkins KJ, Gow AJ, Kitchen D, Pinto A (1992) Proc No. 319. The Fertiliser Society, London, p
1
952. Grotz BJ, Grisolia L (1992) Nitrogen 199: 39
953. Nitrogen (1989) 180: 20
954. Nitrogen (189) 182: 25
955. Karthikeyan S, Neelakantan PS (1991) IFA-FADINAP Regional Fertilizer Conf for Asia and
the Pacific. New Delhi, December 1991, p 3
956. Simanjuntak R, Eko NB (1992) Asiafab 1: 18
957. Asiafab (1992) 1: 15
958. Grotz BJ, Stupin WJ (1990) Fertilizer Latin American International Conf (British Sulphur).
Caracas, Venezuela (October 1990) p 1
959. Dybkjrer I. (1990) IFA Technical Conference, October 1-5, 1990. Venice, Italy, p 1
960. LeBlanc JR (1986) CEER Chern Econ Eng Rev 18(5): 22
961. Rostrup-Nielsen JR, Nielsen PEH, S0rensen NK, Carstensen JH (1993) Ammonia Plant Saf. 33:
184
962. Mohri T, Takemura K, Shibasaki T (1993) Ammonia Plant Saf 33: 86
963. Verduijn WD (1993) Ammonia Plant Saf 33: 165
964. March HD, Thiagarajan N (1993) Ammonia Plant Saf 33: 108
965. Deshmukh D, Raagaard S, Chawes L, Olesen L (1993) Ammonia Plant Saf 33: 278
966. Dybkjrer I (1992) IFA Technical Conference, October 5-8, 1992; The Hague, The Netherlands,
pi
967. Dybkjrer I (1992) IFA-FADINAP Regional Conference for Asia and the Pacific, Bali,
Indonesia, November 30-December 2, 1992, p 1
968. Lee JM, Cialkowski EJ (1993) Ammonia Plant Saf 33: 53
Chapter 7
Ammonia Storage and Transportation-Safety
Anders Nielsen
Haldor Tops0e A/S Copenhagen, Denmark
Contents
This section deals with certain physical and chemical data, which have particu-
lar reference to the safety in storage, transportation and use of ammonia.
At atmospheric temperature and pressure, ammonia is a colorless gas with a
sharp and pungent odor. Ammonia also exists as a colorless liquid. Its vapor
pressure at - 33.6 DC is 1 atm, at + 4.7 DC it is 5 atm, at 20 DC it is 8 atm, and at
25.7 DC it is 10 atm. At 50.1 DC, the vapor pressure is 20 atm. Ammonia vapor at
the boiling point of - 33 DC will have a vapor density of approximately 70% of
the density of ambient air. However, ammonia and air can, under certain
conditions, form mixtures that are denser than ambient air [1], as a result ofthe
mixture being at a lower temperature caused by the evaporation of cold
ammonia. An ammonia cloud formed by an accidental release of ammonia may
contain a mist ofliquid ammonia and its density may be greater than that of air.
For the density of an ammonia cloud containing liquid ammonia as it is diluted
with dry air, see [2].
Several literature sources list different values for the flammability limits of
ammonia mixtures with air, but 16-27% ammonia by volume appears typical.
The flammability limits of ammonia in oxygen are 15 to 79% [3]. Limits of
flammability for various oxygen/nitrogen ammonia mixtures of various temper-
atures as well as the flammability characteristics of ammonia/water vapor/air
mixtures are given in [4]. Flammability limit measurements were carried out at
three temperatures with a supporting atmosphere containing from 21 % oxygen
(air) down to 11 % oxygen. All experiments were carried out at atmospheric
pressure. The data show that the flammability limits widen as temperature
increases. At 400 DC the flammability limits have widened to 11 to 37% in air
[4].
The high solubility ofNH 3 in H 2 0 and the high heat of mixing must be fully
considered in the estimation of its behavior in accidents.
The threshold of perception of ammonia varies with the individual and may also
depend on atmospheric conditions. A lower limit of perception in the range of
0.4 to 2 wt ppm is given in [5]. In this report it is also concluded that levels of
ammonia at the perception threshold may cause changes in the biopotentials of
the brain.
Plant surveys [6] report concentrations of 9 to 45 ppm for various areas.
Such ammonia concentrations initially are irritating to the eyes or nose, but the
Ammonia Storage and Transportation-Safety 331
Storage facilities for ammonia are discussed in this volume in the chapter by
Dybkjaer. Here we will discuss storage facilities with reference to associated
safety aspects. Three principally different types of ammonia storage tanks are
used in producing plants and distribution terminals. Pressurized (unrefrigerated)
storage tanks operate at ambient temperature and are typically designed for a
pressure of 18.25 bar. Storage spheres are refrigerated typically by a single-stage
refrigeration compressor, operated at temperatures from - 1 to + 2°C, and
are designed for pressure of 3.8 to 5.15 bar. The fully-refrigerated storage tank is
serviced by two-stage refrigeration, operated at a temperature of - 33°C, and is
typically designed for a pressure of 1.117 bar [16]. It is important to also design
the tank for a certain underpressure. A capacity of pressurized storage tanks up
to 270 t, refrigerated spheres from 450 to 2750 t, and fully-refrigerated storage
tanks up to 45000 t is reported in [16]. Agricultural-use tanks range from less
than one ton to a few tons in size, the larger tanks being used as applicator tanks
as well as nurse tanks for smaller tanks.
Ammonia Storage and Transportation-Safety 333
Liquid ammonia has a high thermal coefficient of expansion, and rules and
regulations exist in various countries governing the maximum filling of station-
ary and transportation tanks. A guide for developing a training program for
anhydrous ammonia workers has been issued by the U.S. National Institute for
Occupational Safety and Health, Cincinnati [18].
A further reference to actual design, practice and experience of refrigerated
ammonia storage in North America is reported in [16], while an ammonia
storage terminal safety program discusses, among other subjects, safety training
and accident prevention [13].
While the many thousands of small tanks used in distribution and local
storage present their own safety problems, of which examples will be listed
below, very large storage facilities represent a potential for very large ammonia
spills. One effort to reduce the potential hazard is to use a double-walled tank or
to protect the tank by dikes, retention pits or concrete walls. Dikes, and
particularly concrete walls, will limit the surface of the pool of liquid ammonia
which could be formed in the case of the failure of a large storage tank. In the
case of a tank surrounded by a dike, an initial high rate of vaporization would be
controlled primarily by heat transfer from the ground and therefore the initial
flash depends upon the condition of the ground within the dike [19].
The initial vaporization can create toxic and fire hazards a considerable
distance downwind from the tank for a period of time, which is determined
largely by atmospheric conditions. Subsequent steady state vaporization pre-
sents a smaller fire hazard and a more moderate toxicity hazard, and is
controlled by the surface area of the spill and atmospheric conditions.
Stress corrosion cracking has been a serious problem in Europe and also in
the U.S. Early observations came from Denmark [20]. Quenched and tempered
steels appear particularly vulnerable. Inhibition with a minimum of 0.2 % water
is used to reduce the risk of cracking. Oxygen (air) plays an important role in the
corrosion. Since stress corrosion cracking is the result of chemical processes, it is
much less likely and would occur much slower in atmospheric storage tanks
containing liquid ammonia at - 33°C [20]. According to T. Hallan [94] crack
initiation is very infrequent at - 33°C and crack growth rate is approximately
three times slower at - 33 °C than at 18°C. Cf. also Nyborg, Lunde and Conley
[95] concerning ammonia storage vessel life prediction and Nyborg and Lunde
concerning measures for reducing stress, corrosion cracking [96].
For a long time it was believed in the industry that fully-refrigerated
ammonia storage tanks would not suffer stress corrosion cracking. In 1987 [80]
in the U.K., a fully-refrigerated ammonia storage tank was decommissioned for
inspection after almost ten years of service. In the inspection, stress corrosion
cracking was found. The water content of ammonia storaged in this tank had
been typically 0.02 wt %, which had been considered sufficient to inhibit stress
corrosion cracking. The oxygen content of the liquid or vapor had not been
determined, and there was a possibility, particularly during the commissioning
ofthe tank, that some oxygen could have been present and played a role in crack
initiation. The structural integrity of the tank was analyzed and verified by
334 A. Nielsen
fracture mechanics [83]. Certain precautions were taken during the recommis-
sioning of the tank.
A new disclosure of stress corrosion cracking in a fully-refrigerated ammonia
storage tank was reported a year later [89]. Two identical tanks were inspected
of which one showed a significantly higher level of attack by SCC than the other
which could not be correlated with differences in the oxygen or water content of
the ammonia.
A detailed study of stress corrosion cracking has been carried out at
"Institutt for Energiteknikk" Kjeller, Norway financed by ammonia producers
and safety authorities. The experimental results and conclusions as to the
influence of tank environment, electrochemical potential, temperature, steel
properties and types of welding electrodes are reported and recommendations
given [92]. Application of cathodic protection in a large atmospheric ammonia
storage tank is described in [97].
A description of leakage in the floor of an ammonia storage tank (a
particularly critical point of damage) and the repair operations are given in [21].
7.5.1 Pipelines
There are three major ammonia pipelines; two in the U.S. and one in the USSR.
One of the American pipelines (MidAmerica Pipeline System) transports am-
monia from the Texas Panhandle to points in Kansas, Nebraska and Iowa. Its
total peak capacity delivered to a number of points is 8000 t/day [22]. The other
U.S. pipeline, the Gulf Central Pipeline, transports anhydrous ammonia from
major producers along the Texas and Louisiana Gulf Coast to points in Iowa,
Illinois, Nebraska, Indiana and Missouri [23]. Further information on the
maintenance and repair of the MidAmerica Pipeline System is given in [24]. The
total weight of ammonia in the MidAmerica Pipeline System, when full, is
approximately 20,000 tons. Lock valves are 10 miles apart so it would be
possible that 400 tons could get out between lock valves [22]. It is known from
[12] that in one accident 700 tons ofliquid ammonia leaked from the pipeline in
the USSR. In the USSR pipeline accident, the aerosol cloud of ammonia covered
a forested territory about 40 km 2 in area; however, all residents in the region
were evacuated in time. Pipeline operations in the U.S. are governed by the
national Gas Pipeline Safety Act of 1968, as amended in 1979 [17].
Ammonia Storage and Transportation-Safety 335
7.5.3 By Rail
Ammonia is extensively transported by rail car. The type oftank cars used in the
U.S. for transportation of anhydrous ammonia as well as their safety aspects,
inspections and safety record are discussed in [28]. Here also two accidents
involving ammonia tank cars are dealt with. The detailed description of the
accident which occurred in 1976 at Glen Ellyn, Illinois, is described in [29,49].
This accident was due to a derailment as two trains passed each other. The tank
head of an ammonia car was punctured by a coupler of an adjacent car. One of
the conclusions of the analysis is that if the tank car had been provided with
head shields, the tank head would not have been punctured during the derail-
ment. Another accident in which a railroad tank car carrying anhydrous
ammonia suffered a puncture is described in [30]. This tank car contained about
75 t of anhydrous ammonia at a pressure of 4 to 5 atm. The handling of the
railroad derailments involving anhydrous ammonia and other hazardous ma-
terials is discussed in [31]. Computer-aided ammonia rail car loading is
described in [32].
A first-hand account of the efforts of an emergency response team at the
Pensacola ammonia accident, which was due to a train derailment, is given in
[48]. A special feature in some ammonia tank car accidents is the appearance of
336 A. Nielsen
delayed tank car failures. This is reviewed in [50]. In this article, delayed tank
car ruptures which occurred in accidents at Cummings, Iowa and Crestview,
Fla. are discussed. A detailed survey of damages from ammonia spills has been
given by Markham [79].
For transportation accidents by railroad, see also [47]. This report gives a
survey of damage during 177 anhydrous ammonia highway incidents and 570
rail incidents in the U.S. for the years 1971 to 1982. It lists the identification
number for anhydrous ammonia as UNlO05 and the required labeling as a
poisonous gas.
7.5.4 By Road
Various road vehicles are used to transport ammonia. The U.S. Department of
Transportation has limited the use of certain high stress steels only for ammonia
which has a minimum content of water of 0.2 wt %, or a purity at least 99.995%
[33]. A more recent reference [17] mentions that when ammonia is shipped in
containers constructed of quenched and tempered steel it must contain a
minimum of 0.2 wt % water. Various countries specify maximum percentage
filling of tanks to provide a cushion. The European Council of Chemical
Manufacturer's Federation (CEFIC) Zurich has issued a safety card for road
transportation of anhydrous ammonia, containing certain information and
recommendations which have been used in some European countries in connec-
tion with local recommendations and lists of laws and rules relevant to
transportation and storage of anhydrous ammonia [34].
Two road vehicle accidents are particularly well known. In one, a semi-
trailer truck ruptured suddenly in the yard of a factory at Lievin (France). The
accident and the analysis of the cause of the accident is reported in [35]. The
almost instantaneous escape of 19 tons of ammonia caused severe injuries to 20
persons, 5 of whom died. Another serious accident occurred in Houston, Texas,
when a tractor semi-trailer tank transporting approximately 27 m 3 of anhyd-
rous ammonia struck and then penetrated a bridge rail on a ramp connecting
two highways. The tractor and trailer left the ramp and fell on to a highway. The
anhydrous ammonia was released and 78 persons were hospitalized. Six persons
died as a result of the accident. Since the accident happened on a highway, most
of the victims were motorists [36, 37].
the tank, due to instrument failure as a result of freezing up, have been reported
[45].
A general survey of refrigerated ammonia storage in North America, treating
design practices and experience, tank locations relative to neighborhoods, and
various technical questions, is given in [16].
Upon release into the atmosphere of an ammonia spill, a portion of it will flash
off as a vapor. This portion is the initial flash followed by a period of
evaporation until all of the ammonia has evaporated, been dissolved in water or
otherwise disposed of. Some of the fundamentals of the behavior of ammonia
released are discussed in [1,51].
There is a significant difference in the release of ammonia from a refrigerated
tank operating slightly above atmospheric pressure, where the amount of the
initial flash is only a few tenths of a percent of the total, and the release from a
pressurized storage system under a pressure of about 9 atm and 24 cC, in which
case approximately 20% of the spilled ammonia would be flashed. Figures are
given showing the amount of flash depending upon the initial pressure and
temperature. Figures are also shown about the behavior of small and large pools
regarding temperature development and evaporation rates at certain wind and
temperature conditions. The importance of meteorological conditions is pointed
out. There is a difference between the situation at night with an overcast
condition, and the situation on a hot sunny day, where as much as 80% of the
radiation may be absorbed and re-radiation is negligible due to the low
temperature of the ammonia pool. In addition, the question of vapor dispersion
in dealt with [51].
In [1], the various ways of releasing ammonia are discussed, and spills on
land and on water are discussed in a qualitative manner. The· question of liquid
ammonia jets, their direction, the fraction of ammonia remaining airborne and
falling to the ground, the formation of dense air ammonia mixtures, and the
formation of fogs are discussed in [1].
A simple model has been established [52] for the release of anhydrous
ammonia from pressurized containers. The flash off and the entrainment of air in
ammonia vapor is discussed. It is pointed out that the density of the mixture
depends on the airborne fraction of the liquid, the ambient temperature, the
relative humidity and the mass of air entrained in a given sized ammonia release.
Based on the model, the calculation of hazard ranges is discussed and two
instances of releases from pressurized containers, that are fairly well described,
are used for comparison. It is concluded that the mixture of ammonia and air
resulting from a certain release from a pressurized tank is likely to be denser
than air. A detailed account of the Houston, Texas anhydrous ammonia release
Ammonia Storage and Transportation-Safety 339
closing doors and windows survive, and that they can help themselves by
stuffing towels in openings to make the house airtight [28, 55]. (The chance of
survival in a house will depend on the time the house is engulfed in an ammonia
cloud, the conditions of the cloud and the tightness of the house). In the
Pensacola accident the cloud travelled almost 15 miles in a period of an hour
before dissipating [55]. Fire fighters directed a hose into the liquid ammonia
stream and backed this up with a water fog to knock down the vapors which at
the time had reached about 40 m upwind from the spill [48]. A two-phase cloud
that formed during release from a pressure tank is discussed in [56] as well as
the atmospheric dispersion of ammonia.
It is reported that much of the information concerning ammonia aerosol
clouds comes from investigators at the scene of the accidents rather than from
controlled experiments. Risk models for the prediction of risk distances during
release of toxic industrial chemicals, including ammonia, are discussed in [57].
Two sizes ofreleases are considered: one 15 to 25 t and one 100 to 150 t. The risk
distances are given for two wind velocities.
41 tons of ammonia and was conducted under the most stable atmospheric
conditions. An extensive instrumentation system was set up to measure width
and height of the ammonia cloud as it moved in the wind direction, particularly
at 100 m and 800 m from the spill point. Further measurements were carried out
by portable ground level stations at 1.4, 2.8 and 5.5 km downwind. Preliminary
results of maximum ammonia concentrations measured in test four were as
follows: 100 m 6.5%, 800 m 2.1 % and 2800 m 0.5%. It is noteworthy that a fatal
concentration of 0.5% ammonia was registered at a distance of 2800 m.
Reference [93] includes a simulation of Test No. 1.
A study on the predicted hazards from ammonia spills on and under water
has been reported [61-63]. An important aspect is the partition between
ammonia dissolving in water and that being released to the atmosphere. This
partition function is higher, close to 0.9, when liquid ammonia is released
underwater, and lower, depending on release conditions and atmospheric
conditions, when a large quantity is released instantaneously on the water
surface where liquid ammonia will boil to produce saturated ammonia vapor,
and the reaction between ammonia and water also contributes to the vapor
formation. Another phenomenon is the aerosol formation as fine drops of
ammonia are thrown into the air. Partition function ratios of approximately
0.56 have been found for such releases. It must be pointed out that this is an
extrapolation from smaller spills of approximately 10 and 200 kg of ammonia,
the first in a swimming pool and the second on the surface of a lake. Models
have been established. The situation for release on a river such as from a barge
carrying liquid ammonia are studied in detail. Two cases treated are a 200t
surface release and a 3000 t surface spill. It is pointed out that a liquid ammonia
barge that sinks in water may eventually release liquid or gaseous ammonia. The
partition between ammonia picked up by water and ammonia released to the
atmosphere depends on many factors; an important one for underwater releases
is the ratio of the water depth to the pipe diameter of release. The model predicts
that if the depth is on the order of 20 outlet pipe diameters, the partition ratio
may be about 0.9, and it may approach 1.0 for greater depths. One of the
conclusions of the report is that vapor clouds formed from spills on water,
although containing a fraction of aerosols, will have buoyant behavior depend-
ing on wind velocity.
The model has been used to describe the simulation of a number of
situations of transport of hazardous cargo by ship [64]. One of the situations
covered is a fictitious anhydrous ammonia casualty at Louisville, Kentucky. The
greatest potential hazards in ammonia spills, on or below the surface of water,
are obviously those associated with barge transport on rivers and in narrow
coastal waters, and those associated with tanker transport in narrow streams
and harbors.
342 A. Nielsen
The largest ammonia storage facilities are located as part of ammonia producing
plants, or at large distribution centers or terminals. A large number of smaller
storage tanks are typically operated by ammonia distribution companies and
thousands of small tanks are used in distribution and local storage. In Denmark
where the use of liquid ammonia as a direct application fertilizer is wide spread,
there are on the order of 10000 small tanks [20].
A large amount of ammonia is stored in the transportation system in vessels,
barges, tank cars and in ammonia pipelines. According to [65] the Gulf Central
Pipeline holds about 70000 tons of ammonia.
For transfer of - 33°C ammonia, e.g., from an ocean going tanker, into a
pressure or semi-refrigerated storage installation, an ammonia heating install-
ation is required. The actual temperature and pressure in refrigerated storage
tanks depends upon the location of the storage, whether at sea level or at a
higher elevation.
The pressurized storage tanks are typically made for capacities of up to
270 tons, but are found in all sizes down to a few cubic meters in size and even
smaller pressurized containers are widely used in industry down to sizes of one
to a few kg capacities.
For maintaining safety around ammonia installations rules, and regulations
exist in the various countries and even the smallest containers are subject to
rigid and detailed specifications [17]. A detailed survey of refrigerated ammonia
storage tanks in the U.S. and Canada, with information on location and design
features of such tanks, is given by Hale [16]. A HAZOP study of a fully-
refrigerated storage installation is described in [90]. Cf. also [91] on topics of
risk analysis and emergency management.
Two underground ammonia facilities are described. One storage cavern of a
capacity of 20000 tons of ammonia has been operated by Dupont, and Norsk
Hydro, is operating underground storage in Norway of approximately 50000
tons in size [66].
Modern ammonia plants typically in the size of 1000-2000 tid capacity, in
most cases include one or more atmospheric storage tanks as part of the facility.
Details of the design and construction of a high~integrity ammonia storage tank
is given in [84]. Comeau [67] discussed precautions for ammonia storage tanks
and particularly the design of surrounding protection dikes. The function of
such dikes is to keep evaporation to a minimum in the event of a leak and to
protect the tanks against external physical damage. Many consider the best
protection of an atmospheric storage tank to be a high wall of prestressed
concrete around the tank which would further limit evaporation if the tank fails.
One of the technical and safety aspects of storing anhydrous ammonia in
atmospheric tanks is to protect the soil below the tank from freezing. This
problem and experience with tank foundation heaters are discussed by Comeau
and Weber [43]. Special problems arise in the shutdown and startup of
Ammonia Storage and Transportation-Safety 343
Grigor'evskii. The USSR pipeline has a throughput of 2.5 million t/year and
crosses several major rivers.
According to [65] ammonia transportation by pipelines requires the am-
monia to be heated up, at least, to 2°C, which in most cases means that it must
be warmed at the supply terminal and cooled again to - 33°C at the receiving
terminal. For transportation of ammonia, particularly by pipeline, the PVT
properties of ammonia with a small content of water are important. Such data
are found in [73]. The Interstate Commerce Commission requires that all
anhydrous ammonia transported interstate by pipeline must contain a min-
imum of 0.2 wt % of water [73]. The water is added as a corrosion inhibitor and
will, according to experience gained by a survey of ammonia storage spheres,
limit the risk of stress corrosion cracking in surfaces exposed to ammonia liquid
[74, 75]. It is also stated in [75] that contamination with air is the primary cause
of stress corrosion cracking in ammonia; and [75] also discusses the phenomena
that the oxygen content is higher in the gas phase of the tank while water is
contained predominantly in the liquid, and thus a fairly complex picture for
stress corrosion cracking is developed which is important for transportation of
ammonia in any tank. In [17] is given a survey of the various codes and
regulations, covering storage and handling of ammonia in the USA as regulated
through OSHA. It is emphasized that additional municipal, country and state
regulations exist in many locations. It is pointed out that shipping containers
constructed of quenched and tempered steel must contain, at least, 0.2 wt %
water, which is also a requirement for ammonia transported by pipeline [65].
Rail cars can typically hold 26 tons of ammonia, while larger jumbo cars
may hold up to 80 tons and road tankers are, in the U.S., limited to 25 tons of
ammonia. One of the minimum requirements in the U.S. is that containers such
as road and rail tankers used for ammonia shipments must not be filled to more
than a certain percentage by volume to provide a cushion for thermal expansion
should its temperature increase [17]. The tanks are typically designed for 15.5
bar (vapor pressure of ammonia at 43°C) and provided with pressure relief
devices [17]. Detailed description of a tanker loading station for liquid am-
monia is given in [76]. The facility described has the capacity to handle the
loading of 12 tankers per day and to unload 15 tankers per day. Details of a
computer-aided ammonia rail car station are given in [32] including a descrip-
tion of the computer hardware and software used in the facility.
7.11 References
6. National Institute for Occupational Safety and Health (1974) Rockville, MD, PB-246669, p.l
7. National Research Council (1972) Committee on toxicology, Washington, DC, PB-244336, P 1
8. Legters L (1980) AD-A094501 p 1
9. Lessenger JE (1985) Plant Oper Progr 4: 20
10. Barber JC (1978) Ammonia Plant Saf 20: 5
11. Morgan GO, Reed JD (1965) Ammonia Plant Saf 7: 38
12. Zakaznov VF, Kursheva LA, Upadyshev KL (1980) Khim Prom (Moscow) 12: 361; Soviet Chern
Ind (1980) 12: 728
13. Hale CC, Lichtenberg WH (1980) Ammonia Plant Saf 22: 35
14. Michels AMJF, Dumoulin EM, Gerver JH, (1957) Rec Trav Chim 76: 5
15. Leleu J (1976) Inst NatI Rech Secur Paris Cah Notes Doc No. 1024-8476,427
16. Hale CC (1984) Ammonia Plant Saf 24: 181
17. Brenchley DL, Athey GF, Bomelburg HJ (1981) PNL-4006: 1
18. Karches GJ, Froehlich PA, Bicknell RJ (1978) PB 80-189475: 1
19. Husa WH, Bulkley WL (1965) Ammonia Plant Saf7:41
20. Arup H (1977) Ammonia Plant Saf 19:73
21. Lichtenberg WH (1972) Ammonia Plant Saf 14: 24
22. Rohleder GV (1969) Ammonia Plant Saf 11 : 35
23. Inkofer WA (1969) Ammonia Plant Saf 11 :40
24. Ludddeke DE (1975) Ammonia Plant Saf 17:99
25. Briley GG (1967) Ammonia Plant Saf 9: 10
26. Hakansson R (1977) Ammonia Plant Saf 19: 119
27. Caserta LV (1972) Ammonia Plant Saf 14:31
28. Heller FJ (1981) Ammonia Plant Saf 23: 132
29. National Transportation Safety Board (1976) Washington, DC PB 267939, P 1
30. Cato GA. Dobbs WF (1971) Ammonia Plant Saf 13: 1
31. O'Driscoll JJ (1974) Abtr Papers 78th Nat Meeting AIChE Salt Lake City Paper 35D, p 1
32. Arseneaux AA (1985) Ammonia Plant Saf 25: 150
33. Olsen EA (1969) Ammonia Plant Saf 11 :46
34. Conseil Europeen des Federations de l'Industrie Chimique (1979) Eur Council Chern ManufFed
Safety Card-CEFIC TEC-T-I-Rev 3
35. Medard L (1970) Ammonia Plant Saf 12: 17
36. National Transportation Safety Board (1977) Washington, DC PB-268251, p 1
37. Nfitional Transportation Safety Board (1979) Washington, DC PB-80-144942, p 1
38. Nielsen A (1971) Ammonia Plant Saf 13: 103
39. Hutchings J, Sanderson G, Davies DGS, Davies MAP (1972) Ammonia Plant Saf 14: 102
40. Phelps EH (1972) Ammonia Plant Saf 14: 109
41. Lonsdale H (1975) Ammonia Plant Saf 17: 126
42. Lichtenberg WH (1977) Ammonia Plant Saf 19: 59
43. Comeau ET, Weber ML (1977) Ammonia Plant Saf 19:63
44. Esrig MI, Ahmad S, Mayo HC (1975) Ammonia Plant Saf 17: 93
45. Winegar BW (1980) Ammonia Plant Saf 22: 226
46. Sterling MB (1977) Ammonia Plant Saf 19: 77
47. Transportation Research Board (1983) Washington, DC PB-84-143635, P 1
48. Stueben WJ, Ball WL (1979) Ammonia Plant Saf 21 : 76
49. Day BF (1978) Ammonia Plant Saf 20: 30
50. Eiber RJ (1981) Ammonia Plant Saf 23: 146
51. Ball WL (1970) Ammonia Plant Saf 12: 1
52. Kaiser CD, Walker BC (1978) Atoms Environ 12: 2289
53. National Transportation Safety Board (1969) Washington, DC, PB-198790, p 1
54. Greiner ML (1984) Ammonia Plant Saf 24: 109
55. National Transportation Safety Board (1978) Washington DC, PB-28325, p 1
56. Kansa EJ, Ermak DL, Chan ST, Rodean HC (1983) UCRL-88649-Rev 2, p 1
57. Andersson JO, Broxvall A, Karlsson E, Karlsson N, Nyren K, Rejnus L, Winter S (1983) FOA-C-
40183-C2
58. Resplandy A (1969) Chim Ind Genie Chim 102: 691
59. Goldwire HC (1986) Chern Eng Prog 82:35
60. Goldwire HC et al. (1985) UCID-20562
61. Raj PK, Hagopian JH, Kale1kar AS (1974) CG-D-74-74, AD-779400, pI
62. Raj PK, Hagopian JH, Kale1kiar AS, Cece J (1975) Ammonia Plant Saf 17: 102
346 A. Nielsen