You are on page 1of 6

IV.

Catalyst

The industrial process of ammonia synthesis has been generally based on the
reaction of hydrogen and nitrogen at high pressure over a catalytic surface.
Traditionally the catalyst of choice is an iron-based catalyst with magnetite as its
major component. An attractive alternative to this catalytic system is proposed here
as a means of increasing ammonia conversion at lower pressures, and thus reducing
energy consumption in compressors at a lower capital cost.

The proposed system utilizes a promoted ruthenium catalyst deposited on thermally


modified active carbon, forming porous cylindrical pellets about 0.8 mm in diameter
and 3-5 mm long, which has been available to industry relatively recently.(4) This
catalyst is up to twenty times more active than fused iron catalyst at relatively high
conversion degrees. More importantly, although temperature variations have similar
effects on the two catalysts, the effects of ammonia concentration are significantly
different. Iron-based catalyst activity depends strongly on PNH3 (partial pressure of
ammonia). As PNH3 increases from 1 mol% to 10 mol% the rate of the process
decreases 10 to 25-fold. In contrast, the activity of ruthenium-based catalysts is only
slightly affected by changes in PNH3, as well as changes in total pressure. Promoted
ruthenium catalyst deposited on active graphite therefore has been found to have
excellent low pressure and low temperature performance.(5) This is of great
importance to industrial practice, taking into account contemporary tendencies to
lower the applied pressure and thus reduce energy consumption.

Further benefits of using the ruthenium-based catalyst are found in capital cost
savings. As lower pressures are used in the process, there is greater flexibility in
process compressor driver selection. Thinner-walled and lighter vessels, piping, and
fittings can be employed safely, all of which are equipment more commonly
fabricated world-wide and therefore cheaper. This new catalyst opens a world of
possibilities for industrial ammonia synthesis optimization, maintaining high
ammonia conversion and safety standards at significantly reduced costs and
increased profit.

V. Kellogg Advanced Ammonia Process

In 1979, British Petroleum approached M.W. Kellogg to participate in the


development of a new ruthenium-based catalyst. It was known at that time that
replacing the conventional iron-based catalyst, which was used for over 80 years,
with this new catalyst would increase productivity. Several pilot plant tests were
performed using ruthenium supported on a proprietary carbon structure with
various co-promoters.(6) It was evident that significant economic benefits would be
attained with this new catalyst, and so it was chosen to be an integral part of what is
now known as the Kellogg Advanced Ammonia Process (KAAP).

KAAP can be implemented in one of three ways: as an expansion to an already


existing plant, as a retrofit design to an already existing plant, or as a grassroots
design when building a brand new plant. In 1988, KAAP had its first commercial
exposure after Ocelot Ammonia Company in Kitimat, British Colombia (now known
as Pacific Ammonia Incorporated or PAI), contracted M.W. Kellogg to evaluate its
existing ammonia plant for potential capacity increases and to provide a suitable
retrofit design. This company's ammonia plant was ideal for KAAP's first
demonstration since its synthesis loop had separate feed streams for hydrogen and
nitrogen and allowed the KAAP reactor to be run under a variety of stream ratios.(7)

The KAAP system was successfully started up in November of 1992. It consisted of a


KAAP reactor which was installed downstream from the magnetite converter already
in existence. This KAAP reactor was a two-bed radial flow converter with a unique
proprietary sealing system which avoided hot spots within the catalyst bed. The
KAAP catalyst was loaded in its oxidized state, just as most catalysts are, although it
is only active in the reduced state. Therefore, fresh synthesis gas, which heated the
catalyst bed to ~300o, was used to for the reduction process. The synthesis loop
operated at the original design pressure of 2000 psia. Partially preheated feed exits a
2-bed Kellogg converter loaded with magnetite catalyst and enters a steam
superheater and generator, which generates a pressure of 3000 psia. This feed
stream, containing 15% ammonia, then passes into the KAAP converter through a
side inlet to the first bed. When the gas leaves the second bed, ammonia
concentration is increased to about 19%. A continuous purge is required in this
retrofit because the expanded plant contains more inerts than in the original
synloop. After the entire KAAP retrofit was completed, PAI had an energy savings of
0.6 mmBTU/mt.(8) This more efficient and highly flexible system has been very easy
to operate and has paved the way for grassroots facilities.
Grassroots designs are different from retrofit and expansion designs in that they use
3 and 4 bed intercooled reactors. The first bed uses conventional iron catalyst while
the remaining beds utilize the highly active KAAP catalyst. The reason for this is so
that the iron catalyst can take advantage of high ammonia reaction rates at low
ammonia concentrations. As the reaction progresses, however, the ammonia
concentrations increase, and the iron catalyst loses its effectiveness.(9) The KAAP
catalyst is then used to produce high exit ammonia concentrations at low pressures,
since it can be used at high ammonia concentrations.

The grassroots ammonia plants typically utilize KAAP in conjunction with KRES, the
Kellogg Reforming Exchange System. Together, these processes have a multitude of
benefits, several of which stem from the sole implementation of KAAP. The lower
pressure synthesis loop, which leads to significant capital savings, results from the
use of a single case gas compressor with thinner walled and lighter vessels, fittings,
and pipings. This synthesis loop is also advantageous in that it uses energy more
efficiently by recovering heat at a much higher temperature, yielding a 40% decrease
in energy conversion relative to conventional designs.(10) Since the synthesis loop is
less complex than in other plants, operator attention is expected to be less as well. In
addition, all of these benefits bring with them an expectation of greater reliability.

VI. Conclusion

The proposed ammonia synthesis design produces 1,016 metric tons/day of


ammonia at a feed of 5,500 kmol/hr. Although the lack of kinetic data deterred the
completion of the plug flow simulation, a Gibbs reactor successfully emulated the
desired results. The Haldor-Topsoe plant in Clear Lake, Texas, was chosen as the
model for the conventional ammonia plant due to its comparable operation to the
proposed simulation and its use of iron-based catalyst. This process could be further
optimized by lowering compression costs and utilizing a more efficient reactor.
Replacing the conventional catalyst with the new ruthenium-based catalyst in a
multi-bed reactor can achieve these goals. The industrial process may be safely and
productively operated at lower temperatures, thus reducing costs and increasing
profit.

M.W. Kellogg takes full advantage of this superior catalyst in its breakthrough
technology known as the Kellogg Advanced Ammonia Process, or KAAP. KAAP,
implemented as either a retrofit, expansion, or grassroots design, has proven to have
significant benefits, such as reduced capital costs and energy savings. Kellogg's new
ammonia synthesis configuration leads to a economically advantageous and flexible
ammonia plant.

2. Conventional Ammonia Production
The currently adopted ammonia production process basically employs the system invented by
Fritz Haber and Carl Bosch about 100 years ago [1]. Therefore, this system is well known as
Haber–Bosch process. About 85% of total production of ammonia worldwide is produced by this
process [5]. The ammonia synthesis occurs according to reaction (1).

3 H2 + N2 ⇋ 2 NH3 ΔH°27 °C = −46.35 kJ/mol (1)

Ammonia synthesis is an exothermic reaction (negative enthalpy change), and it occurs


spontaneously at low temperatures (negative entropy change). Although it is favored at room
temperature, the reaction rate at which the reaction occurs at room temperature is too slow to be
applicable for at an industrial scale. In order to increase the kinetics of the reaction to achieve the
targeted conversion rate, high pressure and temperature are required. To effectively synthesize
ammonia from its main components (hydrogen and nitrogen), the reaction should be performed
at a relatively high temperature and pressure of 400–500 °C and 10–30 MPa, respectively, with
the assistance of an iron-based catalyst. This condition is demanded due to the high dissociation
energy (941 kJ/mol) of triple-bonded nitrogen. However, to bring the reaction under this high
temperature and pressure, about 30 MJ/kg-NH3 of energy is required [6].

The production of ammonia from natural gas is conducted by reacting methane (natural gas) with
steam and air, coupled with the subsequent removal of water and CO 2. The products of this
process are hydrogen and nitrogen, which are the feedstock for the main ammonia synthesis.
During the process, the removal of sulfur and other impurities is important, because they can
reduce and damage the performance of the catalyst during synthesis. In the ammonia synthesis
process, both nitrogen and hydrogen are compressed to relatively high pressure to be fed to the
synthesis reactor, where the catalyst is immersed inside. The produced ammonia, together with
the unreacted hydrogen, argon and other impurities, is then cooled down for ammonia
condensation in order to separate the ammonia from the other gases. The unreacted hydrogen and
nitrogen are then recycled back and mixed together with the new feedstock. To avoid a build-up
of impurities, such as argon, a small part of the gases is purged from the process. Ammonia
synthesis produces a small amount of heat, which is released from the reactor; therefore, it can be
recovered and utilized for other processes, such as steam and power generation. In general, about
88% of hydrogen’s calorific value can be conserved [7].

Another challenge in the Haber–Bosch process is its low conversion rate; therefore, the process
must be recycled to achieve the expected production flow rate. However, at pressure of 30 MPa,
the conversion rate from the reaction is still low, no more than 25% [8]. This stream recirculation
causes some problems, including the need for an additional recirculation system and a larger
reactor, leading to high investment and operation costs.
When hydrogen is produced via water electrolysis, nitrogen can be supplied via air separation.
Air separations for nitrogen production can be conducted via membrane, cryogenic, absorption
and adsorption technologies [9]. For large scales, cryogenic separation is considered more
economical than other methods. In addition, cryogenic air separation could produce a high purity
of products [10].

The energy consumed during ammonia production, including conversion from primary sources,
typically ranges from about 28 to 37 GJ/t [5]. An ammonia production system from any primary
source, such as natural gas, is considered complex, as it includes many combined processes.
Figure 1 shows the schematic diagram of conventional ammonia production from natural gas.
The system consists of different processes: steam reformation, the water–gas shift reaction,
CO2 removal, syngas purification, and ammonia synthesis and separation. Therefore, efforts to
reduce the total energy consumption require the improvement of the whole process involved.
Due to high energy intensity of ammonia production, ammonia synthesis emits a total of 289.8
Mt-CO2 annually [11], which is almost 0.93% of global CO2 emissions [12].

Figure 2. Schematic diagram of ammonia production from natural gas, employing the Haber–
Bosch process.

Focusing on the Haber–Bosch process, many efforts to reduce its extreme conditions have been
carried out. They include the introduction of an extra component in order to inhibit the catalysis
and the alteration of geometry and electronic nature of the reacting components in order to
optimize the energetics of catalysis [13]. Ru-based catalysts can basically facilitate ammonia
synthesis under mild conditions (at a temperature of 300–450 °C and pressure of 4–15 MPa),
which is significantly lower than the conditions required for iron-based catalysts. However, Ru-
based catalysts are expensive and suffer from hydrogen poisoning [14][15]. Alkaline earth metal
oxides and hydroxides have been identified as promoters to improve the catalytic performance of
Ru-based catalysts [16]. Several electrides (crystals in which electrons serve as anions), such as
Ca2N:e−, which can be deposited in Ru nanoparticles have the potential to facilitate ammonia
synthesis at 200 °C [17]. Transition metals (TM) can also improve synthesis performance,
including lowering the pressure and temperature. This is due to the existence of scaling relations
between the transition-state energy required for the dissociation of nitrogen and the adsorption
energy of all the intermediates [18][19]. Furthermore, Kawamura and Taniguchi [20] have tested
sodium melt as a catalyst for ammonia synthesis. By using this type of catalyst, the synthesis
could be carried out at reaction temperatures of 500–590 °C and atmospheric pressure. However,
further analysis and experimentation are required to bring this method to the level of being
applicable.

You might also like