You are on page 1of 37

Michael Charles Grovesnitrclar.a01.pub3−aff −0001nitrclar.a01.

pub3−aff −0001

NITRIC ACID
1. Introduction

Nitric acid [7697-37-2], HNO3 , also known as aqua fortis, azotic acid, hydrogen
nitrate, or nitryl hydroxide, is a chemical of major industrial importance. Because
of its properties as a very strong acid and a powerful oxidizing agent, as well as
its ability to nitrate organic compounds, nitric acid is essential in the produc-
tion of many chemicals (eg, pharmaceuticals, dyes, synthetic fibers and plastics,
insecticides, fungicides, and explosives) but is used mostly in the production of
ammonium nitrate for the fertilizer industry (see also FERTILIZERS). By the end
of the nineteenth century, its industrial importance had already become estab-
lished in the production of explosives and dyestuffs. After World War II, nitric
acid production grew rapidly with the expanding use of synthetic fertilizers. Since
the year 2000, nitric acid consumption has been increasing steadily by about 2%
pa. Around 80% of nitric acid is for ammonium nitrate manufacture, of which
four-fifths are used for fertilizer production, most of the rest being for ammonium
nitrate-based explosives. Reflecting geographical fertilizer preferences, nitric acid
production is concentrated in Europe (53%), China (15%), and the United States
(13%), with the rest of the world accounting for less than 20%. Only about 10%
of global production is traded (1). Recent years have seen a growth in demand for
nitric acid in the production of polyurethanes via toluene diisocyanate (TDI) and
methylene diphenyl diisocyanate (MDI). Other uses for nitric acid are in the man-
ufacture of polyamides, high explosives (trinitrotoluene (TNT), nitroglycerin, etc),
metal nitrates, dyestuffs (via nitrobenzene and nitrochlorobenzene), and nitrocel-
lulose; in the treatment of metals (eg, the pickling of stainless steels and metal
etching); as a rocket propellant; and for nuclear fuel processing.
The early history of nitric acid is clouded by the lack of standardized
nomenclature. Scholars have identified possible references in a seventh century
Chinese text and eighth century Arabic works. In Europe, the earliest unambigu-
ous description of the preparation of nitric acid is found in Liber de inventione
veritatis written around 1300 (2). In the Early Modern Period it was referred to
as aqua fortis (strong water) or spiritus nitri (spirit of nitre) (3). From that time
onward, nitric acid was produced primarily from saltpeter [7757-79-1] (potassium
nitrate) and sulfuric acid. In the nineteenth century, Chilean saltpeter [7631-99-4]
(sodium nitrate) from South America largely replaced potassium nitrate. Toward
the end of the century, increasing worries about a perceived impending shortage
of Chilean saltpeter for industry, agriculture, and the military led to an upsurge
in research into new technologies. In 1903, the first commercially successful
plant for the oxidation of atmospheric nitrogen (to nitrogen oxides) in an electric
arc was started up in Norway. The nitrogen oxides were absorbed in water
to make dilute nitric acid. Capacity was rapidly expanded at various, mainly
Norwegian, locations. Although the process was energy intensive, consuming
about 16 MWh/ton 100% nitric acid (4), electric arc plants were in operation
in Norway as late as 1940, taking advantage of the inexpensive hydroelectric
power. In Germany, insights gained by W. Ostwald into the catalytic oxidation
of ammonia on platinum led to a first commercial 3 tonnes/day plant in Gerthe
in 1908 (5). The earliest plant to be erected outside Germany was built by a
1

Kirk-Othmer Encyclopedia of Chemical Technology. Copyright © 2020 John Wiley & Sons, Inc. All rights reserved.
DOI: 10.1002/0471238961.1409201803120118.a01.pub3
2 NITRIC ACID

British company in Vilvoorde, Belgium, under an Ostwald license and went


into operation in 1913. In this year, the synthesis of ammonia from atmospheric
nitrogen and hydrogen from water gas produced in coke ovens was successfully
demonstrated using the Haber–Bosch process. During the World War I, all major
belligerents built nitric acid plants based on catalytic ammonia oxidation, none
more so than Germany, which was producing in excess of 900 tonnes/day of nitric
acid when the war ended (estimated from Ref. 6).
A 1920 comparison of production costs of concentrated nitric acid showed
that the cheapest route was via Haber–Bosch ammonia synthesis followed by
Ostwald catalytic oxidation and concentration of the resulting acid. Nitric acid
from Chilean nitrate was twice as expensive (4), p. 160. Around 1926, in the
United States the raw materials costs for making nitric acid from Chile nitrate or
ammonia changed sharply in favor of the latter, as Haber–Bosch ammonia became
locally available (7). From then on, nitric acid manufacture by ammonia oxida-
tion came to eclipse all other processes. In 1927, the first plants operating under
pressure between 0.7 and 0.9 MPa abs came on stream in Germany, Italy, and
the United States, exploiting the new, corrosion-resistant chrome-nickel steels
(8). The pressure plants were much more compact than the previous entirely
atmospheric plants and permitted higher nitric acid concentrations to be reached.
Other milestones followed. In 1929, DuPont was able to increase the strength and
conversion efficiency of the platinum ammonia oxidation catalyst by adding up to
10% rhodium as an alloying material (9). Toward the end of the 1960s the eco-
nomics of the process were improved by the introduction of effective platinum
recovery systems based on palladium, initially alloyed with 20% gold (10), and
later in the mid-1980s with 5% nickel and no gold. In response to increasingly
strict regulation, NOx emission reduction via selective catalytic processes became
more widespread from the early 1980s (11), p. 79. (NOx encompasses all the oxides
of nitrogen which can potentially form nitric acid: NO, NO2 , N2 O4 , and N2 O3 .
Nitrous oxide N2 O is not considered to be NOx .) The early to mid-1990s saw
the replacement of standard woven ammonia oxidation catalyst gauzes by man-
ufacturers’ proprietary knitted gauzes, which had higher conversion efficiency
and allowed longer campaign lengths or a reduction in precious metal inventory
(12–14). The most recent significant advance in the first two decades of the present
century has been the development of nitrous oxide (N2 O) abatement technologies
and their increasing adoption, nitrous oxide being a potent greenhouse gas.
Today, ammonia oxidation is the basis of all commercial nitric acid produc-
tion. Plant capacities range between about 100 and close to 3000 tonnes/day.
Monopressure processes operating at 0.7–1.4 MPa abs minimize capital invest-
ment, whereas dual-pressure processes with ammonia oxidation at 0.3–0.6 MPa
abs and absorption at 1.0–1.4 MPa abs optimize ammonia conversion efficiency
and catalyst use. These process variants produce “weak” nitric acid (55–68 wt%).
Acid concentrations in the range 50–60 wt% are typically used to manufacture
ammonium nitrate for fertilizer and mining explosive production. Higher concen-
trations up to 68 wt% are employed in organic nitrations. “Strong” nitric acid (up
to 99 wt%) is required for many organic reactions of industrial importance. Nitric
acid concentration (NAC) processes use extractive distillation to concentrate the
weak acid. The most commonly used dehydrating agent is sulfuric acid followed
by magnesium nitrate. Both enhance the volatility of HNO3 relative to H2 O
NITRIC ACID 3

so that highly concentrated nitric acid can be distilled off from the maximum
boiling azeotropic mixture. Direct strong nitric (DSN) processes make the acid
directly from nitrogen oxides obtained by the oxidation of ammonia but are of
little importance nowadays.

2. Physical Properties

Crystals of pure nitric acid are colorless and quite stable. Above the melting point
of −41.6∘ C, nitric acid is a colorless liquid that gives off white fumes in moist
air and, unless kept below 0∘ C, has a tendency to decompose, forming nitrogen
dioxide. The rate of decomposition is accelerated by exposure to light and increase
in temperature. Depending on the concentration of dissolved nitrogen dioxide,
the color may range from yellow to red. The normal boiling point of nitric acid
is 83.4∘ C, but when heated the liquid gradually decomposes to form a maximum
boiling azeotrope at 120∘ C with a concentration variously given as 67.4 (15), 68.5
(16), or 69 wt% HNO3 (17).
Nitric acid is completely miscible with water. The freezing point curve for
aqueous solutions of nitric acid has two maxima corresponding to melting points
for two hydrates of nitric acid: the monohydrate (77.77 wt% acid) at −37.62∘ C and
the trihydrate (53.83 wt% acid) at −18.47∘ C (18). More recent work has confirmed
the existence of the dihydrate (63.62 wt% acid), which may play a significant role
in atmospheric chemistry (19). This latter work locates the freezing point minima
at about 32, 73, and 90 wt% acid. There is some variability in the data available
on partial pressures of acid and water vapor over solutions of nitric acid. This is
most evident in the vapor pressure measurements of one component when the
solution concentration of that component is very low. The decomposition of nitric
acid at high temperatures and concentrations can also lead to some doubt about
the accuracy of vapor pressure measurements. The vapor pressure and density
of such acid containing nitrogen dioxide (ie, fuming nitric acid) increase with the
percentage of dioxide present. Data that have been subjected to thermodynamic
tests and corroborated by the measurements of other studies are therefore the
most reliable; see, for example (20). The density, viscosity, and thermal conductiv-
ity of nitric acid solutions are given in Table 1. A detailed tabulation of densities
is available (23). For all temperatures, density increases with acid concentration;
viscosity reaches a maximum at ca 60–70 wt% acid (24).
Thermodynamic data for nitric acid are given in Table 2. Properties for
the ternary systems sulfuric acid–nitric acid–water (17,25) and magnesium
nitrate–nitric acid–water (26–29) used in processes for concentrating nitric acid
are available.

3. Chemical Properties

Nitric acid is a strong monobasic acid, a powerful oxidizing agent, and nitrates
many organic compounds
3.1. Acidic Properties. As a typical acid, it reacts readily with alkalis,
basic oxides, and carbonates to form salts. The largest industrial application of
Table 1. Physical Properties of Nitric Acid Solutions at 20∘ C
Partial
c
Bp at 101 pressures,a kPab Specific heat,e Viscosity,e Thermal conductivity,e
d mPa s (=cP) W/(m K)
HNO3 , wt% kPa,b ∘ C HNO3 H2 O Density, g/mL J/(g K)f
0.0 100.0 2.27 0.9982 4.19 1.0 0.60
10.0 102.2 2.12 1.0543 3.73 1.1 0.57
20.0 104.4 1.91 1.1150 3.39 1.2 0.53
30.0 107.3 1.61 1.1800 3.18 1.4 0.50
40.0 110.8 0.01 1.28 1.2463 3.01 1.6 0.47
50.0 114.7 0.04 0.94 1.3100 2.85 1.9 0.43
60.0 118.2 0.12 0.61 1.3667 2.64 2.0 0.40

4
70.0 119.3 0.37 0.33 1.4134 2.43 2.0 0.37
80.0 112.1 1.22 0.12 1.4521 2.22 1.9 0.34
90.0 96.0 3.57 0.02 1.4826 1.97 1.4 0.31
100.0 83.4 6.20 0 1.5129 1.76 0.9 0.28
a Refs. 21 and 22.
b To convert kPa to mm Hg, multiply by 7.5.
c Ref. 17.
d Ref. 23.
e Ref. 24.
f To convert J/(g K) to cal/(g K), divide by 4.184.
NITRIC ACID 5

Table 2. Thermodynamic Properties of Nitric Acid and Its Main Hydrates


Propertya HNO3 HNO3 ⋅H2 O HNO3 ⋅3 H2 O
nitric acid, wt% 100.0 77.77 53.83
freezing point, ∘ Cb –41.59 –37.62 –18.47
heat of fusion, kJ/molb 10.48 17.5 29.12
heat of formation at 25∘ C, kJ/molc –174.10 –473.46 –1056.04
free energy of formation at 25∘ C, kJ/molc –80.71 –328.77 –811.89
entropy at 25∘ C, J/(mol K)c 155.60 216.90 346.98
heat of vaporization at 25∘ C, kJ/molc 39.04
a To convert kJ to kcal, divide by 4.184.
b Ref. 18.
c Ref. 45.

nitric acid is the reaction with ammonia to produce ammonium nitrate. However,
because of its oxidizing nature, nitric acid does not always behave as a typical
acid. Bases having metallic radicals in a reduced state are oxidized by nitric acid
(eg, ferrous and stannous hydroxide become ferric and stannic salts). Except for
magnesium and manganese in very dilute acid, nitric acid does not liberate hydro-
gen upon reaction with metals.
3.2. Oxidizing Properties. Nitric acid is a powerful oxidizing agent (elec-
tron acceptor) that reacts violently with many organic materials (eg, turpentine,
charcoal, and charred sawdust) (30,31). The concentrated acid may react explo-
sively with Ethanol (qv). Such oxidizing properties have had military application:
nitric acid is used with certain organics, for example, furfuryl alcohol and aniline,
as rocket propellant (see also EXPLOSIVES AND PROPELLANTS).
Depending on acid concentration, temperature, and the reducing agent
involved, any of the following oxidations may occur:

4 HNO3 + 2 e− → 2 NO3− + 2 H2 O + 2 NO2 (1)

8 HNO3 + 6 e− → 6 NO3− + 4 H2 O + 2 NO (2)

10 HNO3 + 8 e− → 8 NO3− + 5 H2 O + N2 O (3)

10 HNO3 + 8 e− → 9 NO3− + 3 H2 O + NH4+ (4)

16 HNO3 + 12 e− → 14 NO3− + 4 H2 O + 2 NH3 OH+ (5)

Low-strength nitric acid favors the generation of nitric oxide, whereas con-
centrated acid favors the formation of nitrogen peroxide, the equilibrium mix-
ture of nitrogen dioxide and dinitrogen tetroxide. Concentrated acid converts the
oxides, sulfides, and so on of most elements in a low oxidation state to a higher
level, for example, sulfur dioxide is oxidized to sulfuric acid. As a general rule,
metals below hydrogen in the electrochemical series, such as copper, yield nitro-
gen peroxide and nitric oxide. Those above hydrogen such as tin react to produce
nitrogen, ammonia, hydroxylamine, or nitric oxide when treated with nitric acid.
Most nonmetallic elements (except nitrogen, oxygen, chlorine, and bromine)
are oxidized to their highest state as acids. Heated with concentrated acid, some-
times in the presence of a catalyst, sulfur, phosphorus, arsenic, and iodine form
6 NITRIC ACID

sulfuric, orthophosphoric, orthoarsenic, and iodic acid, respectively. Silicon and


carbon react to produce their dioxides.
Nitric acid reacts with all metals except gold, iridium, platinum, rhodium,
tantalum, and certain alloys. It reacts violently with sodium and potassium to
produce nitrogen. Most metals are converted into nitrates; arsenic, antimony, and
tin form oxides. Chromium, iron, and aluminum readily dissolve in dilute nitric
acid but with concentrated acid form a metal oxide layer that passivates the metal,
that is, prevents further reaction.
3.3. Organic Reactions. Nitric acid is used extensively in industry to
nitrate aliphatic and aromatic compounds (32). In many instances nitration
requires the use of sulfuric acid as a dehydrating agent or catalyst; the extent
of nitration achieved depends on the concentration of nitric and sulfuric acids
used. This is of industrial importance in the manufacture of nitrobenzene and
dinitrotoluene, which are intermediates in the manufacture of polyurethanes.
TNT is an explosive. Various isomers of mononitrotoluene are used to make opti-
cal brighteners, herbicides (qv), and insecticides. Such nitrations are generally
attributed to the presence of the nitronium ion, NO2 + , the concentration of which
increases with acid strength (see also NITRATION).

HNO3 + H+ ⇌ NO2 + + H2 O (6)

ArH + NO2 + → ArNO2 + H+ (7)

Alcohols and glycerols are nitrated by esterification in a mixture of concen-


trated nitric and sulfuric acids. This reaction is of importance in the production
of nitroglycerin from glycerol and nitrocellulose from cellulose.

ROH + HONO2 → RONO2 + H2 O (8)

Dilute nitric acid can be used to oxidize an aliphatic hydrocarbon. For


example, a significant use for nitric acid is the oxidation of cyclohexanol and
cyclohexanone (qv) to produce adipic acid (qv). Most adipic acid is used for the
production of nylon-6,6.

3 C6 H11 OH + 3 C6 H10 O + 14 HNO3 → 6 HOOC(CH2 )4 COOH + 14 NO + 10 H2 O


(9)

4. Manufacture and Processing

Until the end of the nineteenth century, nitric acid was made by heating a metallic
nitrate salt with less-volatile concentrated sulfuric acid. Removal of the volatile
nitric acid permitted the reaction to go to completion. This method is still used for
laboratory preparation of the acid.
Nowadays, almost all commercial quantities of nitric acid are manufactured
by the oxidation of ammonia with air to form nitrogen oxides that are absorbed
in water to form nitric acid. Because nitric acid has a maximum boiling azeotrope
at about 68.5 wt%, the processes are usually categorized as either weak (sub-
azeotropic) or direct strong (superazeotropic). Typically, weak processes make
NITRIC ACID 7

55–68 wt% acid, and direct strong processes make up to 99 wt% acid. Strong acid
may also be made indirectly from the weak acid using extractive distillation with
a dehydrating agent such as sulfuric acid or magnesium nitrate to increase the
volatility of nitric acid above that of water.
4.1. Weak Acid Process. Two main processes variants are in use nowa-
days, the monopressure process and the dual-pressure process. In the monopres-
sure process, ammonia oxidation and absorption of nitrogen oxides take place at
essentially the same pressure, whereas in the dual-pressure process, ammonia
oxidation takes place at a lower pressure than the absorption. The monopressure
process has the lower capital cost, but the dual-pressure process has a lower over-
all ammonia consumption, since higher ammonia oxidation efficiency is obtained
at lower pressures. In areas of the world with traditionally low energy and there-
fore ammonia prices such as the United States and the Soviet Union, the mono
high-pressure process is favored. In areas of high energy costs such as Western
Europe, the dual-pressure process is preferred for plant capacities above about
500 ton HNO3 /day. Even in low energy cost areas, the dual-pressure process is
preferred for the highest capacities, and many licensors now offer both process
variants. Figures 1 and 2 show examples of the two kinds of processes.
Monopressure Process. Monopressure processes are either medium
pressure, 0.4–0.6 MPa abs or high pressure, 0.7–1.4 MPa abs. The high-pressure

To Air
atmosphere

F
III

EX AC AC D

Water
TGH
CW

CW
F

ST Super-
heater NH3 Absorber
TGH
NOx (& N2O) Burner ST
abatement)
I
III ST NH3 ST
Evaporator CW
Boiler
II
F
NH3

TGH ECO CC

Liq. Bleacher
NH3 CW
BFW
Product
acid

Fig. 1. Monopressure process using catalytic NOx reduction and with tertiary (tail gas)
N2 O abatement, where AC = air compressor, BFW = boiler feed water, CC = cooler con-
denser, CW = cooling water, D = makeup driver, ECO = economizer, EX = expander,
F = filter, ST = steam, and TGH = tail gas heater. The Roman numerals refer to the possible
locations of N2 O abatement: I = primary, II = secondary, and III = tertiary.
8 NITRIC ACID

To Air
atmosphere

EX AC NC D

Water
TGH

CW
F
ST Super-
heater NH3 Absorber
NOx (& N2O) Burner
abatement)
CW NH3 ST
Evaporator CW
Boiler

NH3 F

TGH ECO TGH CC TGH CC

Liq. CW Bleacher
NH3 CW
BFW
Product
acid

Fig. 2. Dual-pressure process using catalytic NOx reduction and optional tertiary N2 O
abatement; NC = NOx compressor. See Figure 1 for other abbreviations.

process has been the most prevalent design. Higher operating pressures reduce
equipment size and capital cost, but the medium-pressure process has a higher
energy efficiency as less compression energy is consumed. The capital cost of the
mono high-pressure process is about 10–15% lower than that of the dual-pressure
process for the same capacity. Higher gauze temperatures and operating pres-
sures accommodate a more efficient recovery of process energy, either as steam or
as reheated tail gas, providing power for air compression. However, the specific
consumption of ammonia in the oxidation step, which is responsible for about
90% of the operating costs, is greater at higher pressure.
Dual-Pressure Process. Dual-pressure processes have a medium-pressure
(0.3–0.6 MPa abs) front end for ammonia oxidation and a high-pressure
(1.0–1.4 MPa abs) tail end for absorption. Some older plants still use atmospheric
pressure for ammonia oxidation with absorption at 0.4–0.6 MPa abs.
Compared to mono high-pressure plants, the lower oxidation pressure
improves ammonia yield and catalyst performance. Platinum losses are signifi-
cantly lower, and production runs are extended by a longer catalyst life. Due to
the split in operating pressures, the dual-pressure process requires a specialized
stainless steel NOx compressor.
Process Description. Air is supplied to the process from a compressor that
is powered by an expander and a makeup driver, all having a common shaft. The
expander is a turbine that recovers energy from the tail gas remaining after the
NOx absorption step as it is reduced to atmospheric pressure. The makeup driver,
usually a steam turbine or more rarely an electric motor, meets the balance of
any power requirement for air compression (and in the case of the dual-pressure
process, NOx gas compression). Ammonia and air are filtered and mixed such that
there is an excess of oxygen and are passed over a platinum/rhodium catalyst in
NITRIC ACID 9

the ammonia burner to produce nitric oxide, water vapor, and much heat. The
resulting gases are cooled, heating tail gas and generating steam that can be
exported or used to power the steam turbine. As the process gases cool, nitric
oxide is further oxidized to form nitrogen peroxide.
The cooled process gases are then cooled in a cooler condenser, and nitric
acid condensate forms. As hot, concentrated liquid nitric acid is very corrosive,
either the cooler condenser is made of expensive high alloy chrome-nickel steels
or even more expensive materials such as zirconium, or the condensation process
is designed to avoid situations in which the nitric acid can come into contact with
very much hotter process gas and become hot and concentrated. The latter pro-
cess usually results in a cooling medium heated to a higher temperature than
a conventional cooling water return, allowing some of the heat of condensation
to be usefully recovered, and permits a greater use of less-exotic chrome-nickel
stainless steel (33,34).
In the dual-pressure process, the gases from the first condensation step
are further compressed and then cooled and partly condensed in a second cooler
condenser.
The process gases then enter an absorber, a tower in which nitrogen perox-
ide reacts with water, producing liquid nitric acid. Nitric oxide, released by the
formation of the nitric acid, must be oxidized to complete the conversion of nitro-
gen oxides to nitric acid. Additional oxygen is required for this reaction and is
supplied to the absorber as secondary air. The nitric acid from the absorber con-
tains quantities of dissolved nitrogen oxides, most of which are stripped from the
acid in a bleacher with all or part of the secondary air, before the latter flows to
the absorber. In a monopressure plant, the bleacher can be internal, installed
inside the bottom of the absorption tower, or a separate, external tower. In a
dual-pressure plant, the bleacher is a separate tower, due to the differing pressure
levels of the absorber and the bleacher. The purified nitric acid from the bleacher
is sent to storage or downstream process units. The tail gas from the absorber is
heated using thermal energy from the process. Residual levels of NOx and N2 O in
the tail gas, which are too high for release into the environment, are removed, and
the hot tail gas discharged to atmosphere via an expander, to recover mechanical
energy.
Compression and Expansion. For many reasons the compressor–
expander set may be considered the heart of a nitric acid process. To a large
extent it sets the energy efficiency and operating flexibility of the process and
represents between 35% and 45% of the total equipment cost. In small monopres-
sure plants, the air compressor is typically a multistage centrifugal unit with
intercoolers, while for large-capacity plants, axial flow compressors are often
used for the initial stages of compression. A dual-pressure plant usually has an
axial air compressor. Sometimes integrally geared centrifugal units with air com-
pressor and expander pinions running off the bull gear and NOx compressor and
steam turbine as inline machines are used for low capacities around 500 ton/day.
In either case, an intercooler for the air compressor is unnecessary because of the
low pressure ratio.
The multistage centrifugal NOx compressor of the dual-pressure process
compresses low to medium-pressure gases (0.3–0.6 MPa abs) containing nitrogen
oxides to 1.0–1.4 MPa abs. The corrosive environment demands the use of
10 NITRIC ACID

stainless steels resistant to nitric acid. Unconverted ammonia from the ammo-
nia burner reacts with nitrogen dioxide and water in the gas stream to form
solid ammonium nitrate deposits, especially in the NOx compressor. If left to
accumulate, the ammonium nitrate presents an explosive hazard. To prevent the
buildup of such deposits, steam or water is sprayed into the suction of the NOx
compressor, with water in some cases also being sprayed into the compressor at
different locations (35). The ammonium-nitrate-containing wash liquid is drained
from the compressor casing at a number of points. The washing procedure is
conducted continuously in a few plants, but more usually once per shift and
during light off of the ammonia burner.
The tail gas expander supplies the lion’s share of the mechanical energy
required to drive the compressor or compressors, with the steam turbine or
electric motor driver contributing only about 10–30%. Expanders for nitric
acid plants are classified according to their inlet temperature because this has
a strong influence on the mechanical design. “Cool” expanders operate below
260∘ C. In inline trains they have single or multiple axial stages. Cool expanders
are no longer used for new plants, as the low temperature is disadvantageous for
overall plant energy efficiency and problematic for SCR NOx abatement systems
using ammonia, because the expander discharge temperature is typically so low
that ammonium nitrate formation here is a possibility if there is ammonia slip
from the SCR. Cool expanders are designed according to the principles used
for steam turbines. “Warm” expanders operate at inlet temperatures between
260 and 510∘ C. High-temperature steam turbines provide the design principles.
“Hot” expanders run at between 510 and around 760∘ C. Their design is based on
industrial gas turbines.
Hot gas expanders incorporate special design features to take account of
the high temperatures, the higher potential for corrosion, and the greater ther-
mal expansion. Creep must be considered for materials in the hottest parts of
the expander, leading to a finite life for these components. This also applies to
the hot-running upstream equipment. In many hot gas expanders, the casing, the
blade carrier, rotor disk, rotor, and labyrinths are cooled by a stream of air, which
mixes with the process medium after use (36), Chapt. 4.
For a given tail gas flow, the amount of recoverable mechanical energy is
proportional to the absolute inlet temperature of the expander and is higher, the
greater the pressure expansion ratio and isentropic efficiency. As the amount of
mechanical energy that can be recovered from a given quantity of heat is typically
greater with an expander than with a condensing steam turbine, net steam export
from a nitric acid plant can be increased by designing for a high expander inlet
temperature. For example, changing the expander inlet temperature from 450 to
650∘ C can increase the net export of steam by about 0.35 t/t of nitric acid product,
the precise figure depending on the steam conditions and the expander and steam
turbine efficiencies. Another way to increase the net steam export is by using an
electric motor driver instead of a steam turbine. The steam export is raised by
about 4 tonnes for every MWh consumed by the electric motor driver.
The compressor or compressors, the expander, and the driver are all carefully
matched to the process and its design capacity. The driver must be designed both
for normal operation at the design and turndown capacities and also for start-up.
The latter is normally the defining case, as the expander provides little or no
NITRIC ACID 11

power during the initial start-up phase of the compressor train since the tail gas
temperature is still low, whereas the operating conditions for the air compressor
(and NOx compressor) correspond roughly to operation at 65–70% capacity.
The characteristics of the compressor train and the process limit the efficient
turndown range of most nitric acid plants to around 65–70% of the design capacity.
The performance of the absorber is very sensitive to the pressure, so it is necessary
to maintain the absorption pressure at a high level across the production capacity
range. An expander behaves as a simple orifice, with the plant pressure varying
proportionally to the plant capacity. To maintain the plant pressure while reduc-
ing the flow, it is therefore necessary to provide additional throttling upstream,
which reduces energy efficiency, or employ an expander with an adjustable first
stator row (37). While this is a practicable solution for cool and warm expanders,
adjustable inlet stators are not usually used in hot expanders because of the
difficulties occasioned by the high temperature. Some hot gas expander manufac-
turers provide a design that allows the stator setting to be changed at shutdown
without changing the assembly (38), Chapt. 6. A reduction in flow through the
compressor or compressors while maintaining the plant pressure moves the com-
pressors’ operating point toward the surge line, limiting the turndown. Having
the air compressor also fitted with adjustable guide vanes can ease this problem.
Due to the corrosive nature of NOx gas it is not feasible to equip the NOx com-
pressor of a dual-pressure plant with adjustable guide vanes. The specific driver
power demand increases as the expander and compressors are operated at lower
capacities away from their point of maximum thermodynamic efficiency, while the
tail gas temperature at the expander inlet often falls at lower capacities, reducing
energy recovery. The plant may therefore generate insufficient steam at low plant
capacity to run the compressor train, if it has a steam turbine drive.
Additional measures to obtain turndowns lower than the typical 65–70%
include venting a portion of the compressed process air to atmosphere to move the
operating point of the compressor away from the surge line. In a monopressure
process, the flow rate through the air compressor can be increased by bypass-
ing some of the air around part of the process and into the expander (39). In the
dual-pressure process, the pressure of the process air is lower than the expander
inlet pressure, so the bypass air must be supplied to an intermediate section of
the expander, or the expander casing divided at a suitable pressure level. To keep
the NOx compressor in its operating range, a part of the discharged NOx gas can
be recycled to the suction via the first cooler condenser (40). Turndowns as low as
35% are said to be achievable.
Most commonly it is required to increase the capacity. Since the air compres-
sor is customarily designed to deliver 100% capacity at unfavorable high ambient
temperature and relative humidity conditions, during cooler, drier parts of the
year and often at nighttime, a plant will usually be able to operate at higher
capacities of up to 110–115% of the design capacity. To obtain still more produc-
tion requires additional oxygen. Since the secondary air is roughly one-fifth of the
total air, a 20% increase in capacity is obtainable by diverting all the air from the
original air compressor to the ammonia burner and installing an additional new
secondary air compressor. Oxygen or oxygen-enriched air can be used instead of,
or in addition to, the secondary air. Liquid or gaseous oxygen may be supplied to
the absorber or the bleaching tower (41–43), and liquid oxygen may also be fed to
12 NITRIC ACID

the weak acid streams flowing to the absorber (44). For the given new capacity, the
total gas flow is less than if air had been used, so that the load on the expander
(and NOx compressor) does not increase in proportion to the capacity increase.
When the plant throughput is increased beyond the original design capacity, the
effect on the plant components must be thoroughly investigated. The heat recov-
ery boiler may need additional heat exchange area; the steam drum may need
modification to give the required steam quality at the higher steam flow. Safety
valves may require rerating. The changed hydraulic conditions in the absorption
and bleaching towers may require trays to be modified or packings exchanged. As
the temperature profile changes through the plant, the acid dew point may occur
in a new location, and the corrosive effects must be guarded against by installing
additional heating area or employing more corrosion-resistant materials for the
affected equipment. Modifications to the expander and the NOx compressor may
be necessary for the higher gas flows.
Conversion of Ammonia. A mixture of air and ammonia [7664-41-7] with
an excess of oxygen is passed over a catalyst to form nitric oxide [10102-43-9] and
water vapor (eq. 10):

NH3 + 1.25 O2 → NO + 1.5 H2 O ΔHr⊖ = −225.5 kJ∕mol (10)

A small amount of nitrogen [7727-37-9] is also formed:

NH3 + 0.75 O2 → 0.5 N2 + 1.5 H2 O ΔHr⊖ = −316.8 kJ∕mol (11)

and, especially on platinum alloy catalysts, some nitrous oxide [10024-97-2], too:

NH3 + O2 → 0.5 N2 O + 1.5 H2 O ΔHr⊖ = −276.0 kJ∕mol (12)

The standard enthalpies of reaction ΔHr⊖ have been calculated from stan-
dard molar enthalpies of formation at 298.15 K (45). To convert kJ to kcal, multiply
by 4.184.
With platinum-based gauze catalysts, the yield of the desired product,
NO, is higher at low pressure. In atmospheric pressure burners, for economic
reasons now obsolete for new plants, nitric oxide yields of up to 98% are achieved.
At medium pressure (0.3–0.6 MPa), the average NO yield is 93–97%, and at
high pressure (0.7–1.4 MPa) around 90–93% (46). Nitrogen and nitrous oxide
are formed in roughly equal proportions so that the typical tail gas N2 O con-
centrations are 650–800 ppmv for atmospheric burners, 900–1100 ppmv for
medium-pressure burners, and 1500–3000 ppm for high-pressure burners (47).
Although the great majority of nitric acid plants worldwide use platinum
group metal-based wire gauze catalysts, metal oxide materials are also employed.
Plants in Australia (48) and Canada (49) have been using cobalt oxide at least
since the end of the past century. An advantage of metal oxide catalysts is that
they produce less of the potent greenhouse gas N2 O (50). In the Soviet Union,
a combination of conventional platinum alloy gauzes installed upstream of an
iron oxide catalyst as pellets or later as a honeycomb has been intensively inves-
tigated and tested in several commercial mono high-pressure (0.7 MPa) plants.
The platinum inventory could be reduced by 20–25%, and the platinum losses by
15% (51). Other much-investigated materials include cobalt oxide doped with rare
NITRIC ACID 13

earth metals, and mixed metal oxides with perovskite structure, frequently con-
taining additional metal oxides in small quantities (52). None of these has so far
been commercialized.
Platinum alloy-based catalysts are in the form of fine wire mesh or gauze
and are used in multiple layers to form a gauze pack. Until the development of
knitted gauzes in the early 1990s, all gauzes were woven. Knitted gauzes can be
produced with less waste and also have advantages for nitric acid manufacturers
including higher ammonia conversion efficiency, greater mechanical strength and
flexibility, reduced metal losses, and longer campaign length before gauze replace-
ment is necessary (13). Nowadays at least 95% of all ammonia oxidation gauzes
are knitted (53), p. 94.
At its most basic, a gauze pack consists of a uniform set of single-layer knit-
ted gauzes of identical structure and composition, typically 95 wt% platinum with
5 wt% rhodium for increased strength and improved reaction yield. At the next
level of sophistication, different wire diameters may be used within the gauze
pack. Yet more complex gauze packs have increasing concentrations of palladium
in the direction of gas flow. Platinum lost from the first gauzes is partly recov-
ered on the downstream gauzes with higher palladium concentration. The ternary,
platinum-enriched alloy thus formed is able to catalyst ammonia oxidation to
nitric oxide; also, N2 O is adsorbed on the palladium surface where it can react with
ammonia to form nitrogen and water vapor, lowering greenhouse gas emissions
(54). Further advantages claimed for such tailored gauze packs are a reduction in
platinum group metal inventory and lower precious metal losses. The campaign
length can be extended by 50–100% without a concomitant reduction in ammo-
nia conversion efficiency (55). Varieties of structured or tailored gauze packs have
been developed by various suppliers ((55–57)), each supplier having their own
proprietary forms.
The performance of a gauze pack varies with age due to changes in its
structure and composition. A new gauze wire is smooth and round. After a short
period of operation, conversion efficiency increases as the surface roughens and
metal migrates to the surface to form small metal growths or “cauliflowers”
that increase the surface area of the catalyst. Cauliflower formation is most
pronounced on the top gauzes, where most reaction takes place, see Figure 3. As
these cauliflowers increase in size, the catalyst structure weakens and catalyst is
lost as volatile platinum and rhodium oxides, PtO2 and RhO2 . Platinum is lost at
a higher rate than rhodium. The catalyst surface becomes enriched with inactive
and nonvolatile rhodium oxide, Rh2 O3 , and ammonia conversion efficiency begins
a gradual decline. The rate of metal loss is a strong function of temperature; as
much as a 10-fold increase has been reported for a temperature change from 820
to 920∘ C (58). The amount of precious metal loss is financially significant, and
various methods such as tailored gauze packs, downstream catchment gauzes,
and glass fiber filters are used to reduce or recover catalyst losses.
The platinum gauze catalysts are poisoned by a large number of substances
including iron, carbon, sulfur, chlorine, nickel, cobalt, chromium, magnesium, cal-
cium, aluminum, silicon, and their compounds (59). For this reason, the air and
ammonia feeds are filtered to a very high standard, and equipment and piping
are of stainless steel rather than carbon steel, to prevent rust formation. If the
gauzes are operated at too low a temperature, irreversible deactivation due to the
14 NITRIC ACID

(a) (b) (c)


Fig. 3. Scanning electron microscope images taken from used gauze pack from ammonia
burner of nitric acid plant. (a) First (top) gauze in gas flow direction, complete coverage of
surface with facets and “cauliflower” structures. (b) Middle layer gauze, some facets and
hollow structures projecting from wire. (c) Last gauze in pack, structure changes caused
by temperature effects only. Source: By kind permission of Umicore AG & Co. KG., Hanau,
Germany.

Fig. 4. Scanning electron microscope image of characteristic rhodium(III) oxide Rh2 O3


needles on the top gauze of a deactivated gauze pack. Source: By kind permission of Her-
aeus Deutschland GmbH & Co. KG, Hanau, Germany.

formation of characteristic needles of rhodium(III) oxide, Rh2 O3 , is the result, see


Figure 4. It is therefore important to ensure that the gauzes rapidly reach their
design-operating temperature during a plant start-up.
To start the ammonia oxidation reaction, the gauze pack must be heated to
raise the temperature to the point at which a self-sustaining reaction between
ammonia and oxygen is ignited. Ignition is usually with the help of hydrogen
flames at start-up, but electrical heating is also used (60). As the smooth wires
of new gauzes are not as active as they will later become, a used gauze with its
rougher surface may be laid on top to facilitate the first light-off. New gauze packs
are also available which have had their surfaces roughened during the production
process (55,61).
Volatilized platinum and rhodium oxides can be recovered on “catchment”
or “getter” gauzes placed immediately below the ammonia oxidation gauzes.
NITRIC ACID 15

Catchment gauzes are made from an alloy of palladium with 5% nickel or


tungsten for strength (61). Platinum recovery rates are between 60% and 85%
(62). Up to around 30% of the lost rhodium can be recovered. The recovery of
platinum and rhodium is at the cost of a loss of palladium from the catchment
gauzes, but overall the economics of the process are improved by installing a
catchment.
A combination of theoretical molecular modeling techniques and experimen-
tal investigations mainly at low pressure (up to about 6 kPa abs) and ultrahigh
vacuum (10−4 –1.0 Pa abs) to reduce or eliminate mass transfer limitations contin-
ues to deepen the understanding of the reaction mechanism of ammonia oxidation
(63–67).
In the first step, ammonia and oxygen are adsorbed at specific sites on the
platinum catalyst. Adsorbed ammonia is then successively stripped of its hydro-
gen atoms to form adsorbed NH2 , NH, and finally nitrogen atoms by interac-
tion with adsorbed oxygen atoms. These interactions also result in the formation
of hydroxyl groups (OH) on the catalyst surface, which in their turn also strip
hydrogen atoms from ammonia, forming water in the process. Adsorbed nitric
oxide, the favored product at high temperatures, is produced by the combina-
tion of an oxygen and a nitrogen atom. As TAP (temporal analysis of products)
experiments using isotopically labeled reactants have shown, NO is the primary
product, with the unwanted N2 O and N2 forming via secondary reactions involv-
ing NO, and pairs of adsorbed nitrogen atoms originating from ammonia, respec-
tively (68).
To bridge the gap between reaction mechanisms on platinum surfaces and
the behavior of industrial ammonia oxidation reactors requires a consideration
not only of the rates of the elementary surface reactions but also of the mass
transfer of the reactants and products to and from the surface, respectively. Such
studies (69), including ones employing CFD (computational fluid dynamics) in
combination with chemical kinetics (70,71), have improved our knowledge of the
process. Almost all (90%) of the ammonia conversion takes place on the top gauze,
which tallies with the observation that the cauliflower surface restructuring
takes place at the top of the gauze pack. The temperature of the top gauze is
several tens of degrees (∘ C) hotter than the adiabatic temperature of the oxidation
products. It was found experimentally, by supplying nitric oxide in the burner
feed, that nitric oxide does not decompose on the gauzes even at long contact
times (69). The view that short contact times between NO and the catalyst are
important to maximize yield is a widely held one (72), p. 601, and goes back to
Wilhelm Ostwald himself (73).
The operating parameters of an industrial ammonia burner depend on the
process pressure for which the unit has been designed. Thus, the higher the pro-
cess pressure, the higher are the gauze temperature, the superficial burner load,
the gauze count, and the specific primary precious metal losses per ton of nitric
acid product. Because of the higher primary metal losses the gauzes need to be
changed more frequently in plants with high-pressure burners.
Thus, atmospheric pressure burners operate at 820–860∘ C with a superficial
burner load of 3–5 tonnes of NH3 nitrogen per m2 of burner cross-sectional area
per day (tNH3 -N/m2 /d) and have primary precious metal losses between 30 and
60 mg/t HNO3 and campaigns lasting 10–15 months. At medium pressure, the
16 NITRIC ACID

corresponding figures are 860–890∘ C, 6–17 tNH3 -N/m2 /d, 80–200 mg/t HNO3 , and
4–9 months, and for high-pressure burners 890–930∘ C, 35 tNH3 -N/m2 /d or more,
200–350 mg/t HNO3 , and 3–4 months (46).
While some older plants with burners at atmospheric pressure are still in
operation, no such new plants are being built. The higher capital cost and space
requirement compared with mono high-pressure or modern dual-pressure plants
of the same capacity make them unattractive. It is also more difficult to meet cur-
rent NOx emission limits with the medium-pressure absorption always associated
with atmospheric ammonia oxidation.
Industrial ammonia burners at atmospheric pressure typically have an
ammonia feed concentration of 12 vol%, while medium- to high-pressure burners
operate at around 10–10.8 vol%. Higher concentrations are restricted by the
lower explosive limit for ammonia in air, which is 12.4 vol% at 0.8 MPa abs,
13 vol% at 0.5 MPa abs, and 13.8 vol% at atmospheric pressure (74).
With advances in ammonia burner design, single-train dual-pressure plants
are now feasible up to about 1600 mtpd with a burner diameter of up to 6 m
(75). Larger capacities require two parallel burners. Based on the capacity of the
available compressor trains, the current limit on capacity is “more than 3000”
mtpd for dual-pressure plants (76). The largest monopressure plants are about
1000 mtpd, for economic rather than technological reasons.
Oxidation of Nitric Oxide. Nitric oxide reacts with oxygen to yield nitrogen
dioxide [10102-44-0] according to the reversible reaction (eq. 13)

NO + 0.5 O2 ⇌ NO2 ΔHr⊖ = −58.1 kJ∕mol (13)

Low temperatures strongly favor the formation of nitrogen dioxide. Below


150∘ C, the equilibrium composition is almost 100% NO2 . The oxidation of nitric
oxide is a slow reaction and is a rare example of one that is faster the lower the
temperature, the rate increasing by about a quarter between 100 and 50∘ C. The
rate expression is
d[NO]
= −2k[NO]2 [O2 ] (14)
dt
The recently reassessed rate constant, valid in the range 0–507∘ C, with T the
temperature in K, is (77)
{ }
44,700 10.2
− T +9.685
k = 10 T2 cm6 ∕mol2 ∕s (15)

In almost all situations encountered in a modern nitric acid plant either


the NO2 concentration or the process temperature is low enough for the reverse
reaction to be neglected. The rate of NO oxidation is proportional to the cube of
the total pressure; thus, the time needed to obtain a given extent of oxidation is
inversely proportional to the cube of the pressure. This is of great significance for
plant design and economics, since the volume of equipment needed to oxidize the
nitric oxide is inversely proportional to the cube of the pressure.
Nitrogen dioxide rapidly forms an equilibrium mixture with its dimer, dini-
trogen tetroxide [10544-72-6 ] (ΔHr⊖ = −27.7 kJ/mol of NO2 consumed). Low
temperature and high pressure favor the formation of dinitrogen tetroxide.
NITRIC ACID 17

Absorption of Nitrogen Oxides. The “absorption of NOx gases is probably


the most complex case in all absorption operations” with “more than 40”
equilibrium reactions to consider (78). Due to its industrial significance this
system has been intensively investigated (79–83).
The overall equilibrium involved in NOx absorption in water is usually sum-
marized as

1.5 N2 O4 (g) + H2 O(l) ⇌ 2 HNO3 (aq) + NO(g) ΔHr⊖ = −52.4 kJ∕mol (16)

The standard enthalpy of reaction is for nitric acid at infinite dilution. The
nitric oxide released according to reaction 16 oxidizes in the gas phase to NO2
according to equation 13, and the next absorption step follows via the dimer.
It has been empirically determined that the equilibrium constant derived from
equation 16
PNO
K = 3∕2 (17)
PN O
2 4

does not depend on the temperature but is a function only of equilibrium nitric
acid concentration, falling with increasing acid strength (80). A high gas-phase
N2 O4 concentration combined with a low NO concentration at the inlet of an
absorber is a prerequisite for a high product acid concentration. High total
pressure and low temperature increase the proportion of N2 O4 and therefore the
equilibrium nitric acid strength.
At high nitrogen peroxide concentrations, the main route to nitric acid
is via absorption of N2 O4 with rapid hydrolysis to nitric acid and nitrous acid
[7782-77-6], HNO2 . Nitrous acid decomposition to nitric oxide, nitric acid, and
water follows. At very low nitrogen peroxide concentrations (<∼100 ppmv at
1.0 MPa total pressure, calculated from Figure 4 in Ref. 78), NO2 absorption
dominates. Low temperatures and high nitric acid concentrations (40–60 wt%)
enhance the solubility of nitric oxide, and liquid-phase oxidation of nitric oxide
by nitric acid yielding nitrogen dioxide and nitrous acid takes place (84). Nitrous
acid plays a significant role in absorption (84). The nitrous acid concentration
increases with rising nitric acid strength, particularly above about 40 wt% HNO3 .
The nitrous acid concentration that is stable increases with falling temperature,
too, so that excessive cooling in the bottom section of an absorber may be coun-
terproductive, leading to formation of HNO2 rather than to increased HNO3 . The
resulting high level of nitrous acid in the nitric acid leaving the absorber must be
stripped, increasing the load on the bleacher. If additional cooling is available, it
is more usefully employed in the upper section of the absorber to reduce the NOx
concentration in the tail gas.
Because of the complexity of the phenomena governing NOx gas absorption,
the early equilibrium tray efficiency approach (85,86), p. 178, has been superseded
by rigorous methods of increasing sophistication (79,87–93). Rigorous models con-
sider the equilibria in the gas and liquid phases, the kinetics of the chemical
reactions involved, thermal effects, and the heat and mass transfer between the
phases, taking account of absorption tray hydraulics and internals, such as cool-
ing coils. Good to very good agreement with laboratory or industrial plant data is
claimed by the references.
18 NITRIC ACID

New plants are equipped exclusively with trayed absorbers with cooling coils
on the trays. Sieve trays are employed predominantly, due to the lower capital cost
and superior performance (83) compared with bubble cap trays, which were pop-
ular in the past. The largest absorbers are around 80 m high with a diameter of
6 m and some 50 trays. Packed absorbers with external cooling and acid circula-
tion systems, known from the infancy of the Ostwald technology, fell out of favor
toward the end of the 1960s (94), p. 664, because of the greater complexity and
higher pumping power consumption.
The nitrous gases from the ammonia burner are cooled in several heat
exchangers to recover useful heat energy before entering a cooler condenser in
which nitric acid forms. In a monopressure plant, the process gas remaining after
the condenser flows to the bottom of the absorber, whereas in a dual-pressure
plant, the process gas from the first cooler condenser is compressed to the
absorption pressure and cooled again, passing through a second cooler condenser
before entering the absorber. The concentration of the nitric acid formed in
the first cooler condenser is greater at higher pressures. At low pressure, the
concentration is between 1 and 10 wt%; at medium pressure (0.4–0.6 MPa abs),
the concentration is in the range 37–40 wt%, and in a mono high-pressure unit
(0.7–1.4 MPa abs), condensate strengths of 45–50 wt% are attained. The acid
condensate is sent to a tray in the absorber selected to cause the least disturbance
to the smooth acid concentration profile over the height of the absorber. Failure
to do this results in a higher than necessary NOx concentration in the tail gas, so
normally the absorber is configured to enable one of several trays to be supplied
with acid condensate.
By the overall reaction 16, it is apparent that nitric oxide is produced when
nitric acid is formed, lowering the degree of oxidation of the nitrous gases. To make
product acid of high strength, the nitrogen oxides must be highly oxidized before
entering the absorber. With medium-pressure absorbers it may be necessary to
provide additional oxidation volume, but in most plants the space below the first
tray is volume enough. Air, used to strip dissolved NOx from the product acid, is
mixed with the process gas to speed oxidation.
The deeper understanding of the mechanisms of absorption described
above led to several retrofit processes which could enhance the performance of
existing absorbers. The high efficiency absorption (HEA) technology developed
by Rhône-Poulenc (95) involved increasing substantially the liquid residence
time on trays higher up in the tower. The higher solubility of nitric oxide in more
concentrated nitric acid and the oxidative power of nitric acid were exploited
in the Bolme process (96,97) in which the tail gas leaving the existing absorber
with a low concentration of NOx was scrubbed with 25–30 wt% nitric acid.
The dissolved NOx was recovered in a second step by heating, stripping, and
cooling the scrubber liquor. These processes had their heyday in the 1980s, being
superseded by SCR (selective catalytic reduction of NOx with ammonia) and the
almost universal use of high-pressure absorption for new plants.
Another class of technique employs additional oxygen or hydrogen peroxide
to enhance the conversion of nitric oxide and nitrous acid into nitric acid.
Gaseous oxygen can be used as bleaching air, or added upstream of the cooler
condenser. The increased oxygen partial pressure speeds the rate of nitric oxide
oxidation to nitrogen peroxide, and hence of nitric acid formation. The net result is
NITRIC ACID 19

a reduction of the NOx concentration in the tail gas (98,99). Liquid oxygen injected
directly into acid condensate streams flowing to the absorption system improves
absorption performance by oxidizing nitrous acid to nitric acid that would other-
wise decompose to nitric oxide. The technology comes into its own in older plants
using absorbers with packing, as each of the typically three or four towers has
a number of recirculation cooling and transfer streams of weak acid containing
nitrous acid (44).
Hydrogen peroxide solution supplied to the top of the absorber lowers the
NOx concentration in the tail gas exiting the absorber. The technology is well
established, being listed as the BAT (best available technique) for nitric acid man-
ufacture in the European Union (100), pp. 135–136. A 2012 investigation in the
United States found hydrogen peroxide-assisted NOx abatement to have a similar
cost per ton of nitric acid product to the more popular SCR (101). The main appli-
cation is in lowering visible NOx emissions at plant start-up, but the technology
is also used in nitric acid plants that have such a low tail gas temperature down-
stream of the tail gas turbine that there is risk of ammonium nitrate formation
if SCR were used (100). A solution of nitric acid with hydrogen peroxide performs
better than hydrogen peroxide solution alone (102–104).
Some promising studies into the kinetics of liquid-phase oxidation of NOx
and nitrous acid with ozone have been carried out (105,106); however, ozone
appears not yet to have been used for emission reduction in commercial scale
nitric acid plants.
NOx Abatement. The practical lower limit for the NOx concentration in the
tail gas leaving a high-pressure absorber is about 150 ppmv. Adding further trays
reduces the NOx concentration very slowly. Until the beginning of the present cen-
tury, an absorber could realistically be sized to meet the then legal emission limits,
for example, US EPA 3 lb NOx /ton of product nitric acid or 1.5 kg NOx /metric ton,
equivalent to 200–230 ppmv, or German TA Luft of 1986 0.45 g/m3 = 219 ppmv,
with m3 being the gas volume at 0∘ C and 0.1013 MPa abs. (In each case, the mass
unit is defined as though the NOx were present entirely as NO2 .) Current limits
are much more stringent. The German TA Luft of 2002 (107) demands 0.20 g/m3
or 97 ppmv NOx . In 2012, the US EPA imposed a limit of 0.5 lb NOx /ton nitric acid
product or 0.25 kg/metric ton as a 30-day rolling average (108) on new or modi-
fied nitric acid plants, corresponding to an average tail gas NOx concentration of
30–40 ppmv. However, the revised limit applies to all periods of operation includ-
ing start-up, shutdown, and malfunction (SSM), during which NOx emissions are
normally much higher. It is not feasible to build an absorber to achieve these lim-
its by itself, and nowadays high-pressure absorbers are typically designed with
an outlet NOx concentration of 200–400 ppmv at normal operating conditions.
During start-up and part load, the plant pressure is reduced, and the NOx outlet
concentration is therefore higher.
Several methods of reducing NOx emissions below what an absorber can
achieve are therefore in use, from hydrogen peroxide in the absorber, described
above, to nonselective catalytic reduction (NSCR) and SCR. The latter is the pre-
dominant technology. Having considered all these technologies (101), the US EPA
determined in 2012 (109) that SCR is the BSER (“best system of emission reduc-
tion”) for NOx .
20 NITRIC ACID

NSCR uses a precious metal catalyst and fuel (hydrogen or gaseous hydro-
carbon such as propane or natural gas) to reduce nitrogen oxides to nitrogen and,
in the process, combust any remaining free oxygen in the tail gas. The reductant
consumption is therefore substantially larger than in an SCR system. There is a
large temperature rise due to combustion, about 130∘ C for every volume percent
of oxygen. NSCR is unsuitable as a retrofit option, as the downstream equip-
ment must be designed for the high operating temperatures. No new plants are
being designed with the economically unattractive NSCR technology, and only a
few plants still in operation use it. On the positive side, NSCR removes not only
NOx but also around 80% of the greenhouse gas nitrous oxide present in the tail
gas (110).
SCR uses a catalyst and ammonia to abate NOx . In contrast to the fate of
the reductants used in NSCR, the ammonia in SCR reacts practically only with
the NOx and not to any extent with the oxygen in the tail gas. The amount of
ammonia required is about 1–1.2 mol of NH3 /mol NOx destroyed.
Oxides of vanadium V2 O5 , tungsten WO3 , and molybdenum MoO3 are good
SCR materials (111). The most commonly used is V2 O5 on a titanium dioxide TiO2
carrier. The catalysts are available as pellets and honeycomb. Typically, pellets
will be used for reactors under pressure. Honeycomb catalyst is used where very
little pressure drop can be permitted, such as downstream of the tail gas expander.
Iron zeolites have established themselves as an alternative for several rea-
sons: V2 O5 catalysts are less active than iron zeolites at temperatures above 400∘ C
(112,113), and form increasing amounts of N2 O above 350∘ C (114,115), whereas
iron zeolites actually decompose nitrous oxide; furthermore, V2 O5 is carcinogenic,
whereas iron zeolites pose no serious health problem.
Either catalyst type can meet current NOx emission limits. The ammonia
slip is below the EU BAT level of 5 ppmv (100), p. 141. Even this level of ammonia
presents a potential safety hazard, since ammonia with NOx at low temperatures
can form ammonium nitrate and nitrite, which could accumulate in downstream
equipment and pose an explosion hazard upon being heated. To avoid this risk,
the ammonia supply is interlocked to prevent ammonia from entering the NOx
abatement reactor when the temperature in the relevant parts of the plant is so
low that ammonium nitrate/nitrite could form. In many plants this is the case at
start-up, so that the NOx abatement reactor cannot be put into operation early
enough to prevent visible NOx emissions, unless additional special measures are
adopted.
N2 O Abatement. At the beginning of this century, nitric acid plants repre-
sented the single largest industrial process source of the potent greenhouse gas
nitrous oxide, N2 O, with annual emissions estimated at 0.4 million metric tons
(47) equivalent to 120 million tons of CO2 . Especially since the ratification of the
Kyoto Protocol in 2005, some but not all nitric acid plants have been equipped
with N2 O abatement technology. Approaches to reducing nitrous oxide emissions
can be categorized as (1) primary, designed to prevent N2 O from forming in the
first place, (2) secondary, the destruction of N2 O between the ammonia burner
and the absorber, and (3) tertiary, removal of N2 O in the tail gas downstream of
the absorber, see the Roman numerals I, II, and III in Figure 1.
Primary measures: Platinum gauze vendors have achieved some success
in modifying the catalyst pack to reduce N2 O production “significantly” (54).
NITRIC ACID 21

The greatest potential for primary N2 O emission reduction is the use of


nonplatinum-based catalysts, which, however, are still very uncommon.
Secondary measures: Because N2 O decomposes at high temperatures,
providing an empty chamber between the ammonia burner and the first heat
exchanger will reduce N2 O emissions (116). A 3.5–m-long cylindrical space below
a typical medium-pressure burner would be sufficient to destroy 90% of the N2 O.
This technology is not easily realized as a retrofit, however. The most common
secondary method is to install a nitrous oxide decomposition catalyst underneath
the gauze and getter pack. The catalyst is usually in the form of loose pellets, and
the burner design must be such as to prevent gas from bypassing the catalyst
bed. Modifications may be required for retrofits. Catalyst materials known to
be in commercial use from literature include copper-zinc spinel (117), precious
metal-impregnated alumina (118), and cobalt aluminate-supported ceria (119).
Tertiary methods: N2 O is catalytically decomposed or reduced in a reactor
usually located immediately upstream of the tail gas expander where the tem-
perature is highest. Many candidate catalyst materials have been tested, includ-
ing zeolites, supported precious metals, and pure and mixed oxides (120). Of the
zeolites, various iron- and copper-exchanged zeolites have been intensively inves-
tigated (121–123). Iron zeolites are the main tertiary catalyst materials in use.
NOx increases the rate of catalytic decomposition of N2 O (124) on iron zeolites, so
N2 O decomposition is carried out on a first catalyst bed, NOx SCR, with ammonia
on the second one. Falling activity with temperature places the practical lower
limit of N2 O decomposition on iron zeolites at about 425∘ C (125). The temper-
ature range can be extended downward by reducing N2 O as well as NOx with
ammonia. The ammonia consumption is correspondingly higher, between 0.7 and
1.4 mol/mol nitrogen oxides, in the gas (126). A further extension down to about
330∘ C is possible using a hydrocarbon such as natural gas as the reducing agent
for N2 O (127). As the hydrocarbon reacts only with the N2 O, its consumption
is very much lower than in NSCR. Since NOx strongly inhibits the reduction
of N2 O with hydrocarbons, NOx must first be reduced completely with ammo-
nia. The two reductions are carried out sequentially in one catalyst bed (125).
For high-temperature tail gas, above about 550∘ C, alternative N2 O decomposition
catalysts such as temperature-stable pure or mixed oxides are suitable. The inde-
pendently assessed performance of secondary and tertiary technologies at many
installations is published at the website of the Clean Development Mechanism in
the form of monitoring reports for nitric acid plant CDM projects (128).
4.2. Nitric Acid Concentration. Nitric acid has a maximum boiling
azeotrope at about 68.5 wt%, so it is not possible to obtain greater concentrations
by a simple distillation of the weak acid. The solution is extractive distillation
of weak nitric acid with a third component which increases the volatility of
nitric acid relative to water. Commonly, sulfuric acid or magnesium nitrate is
used. Nitric acid concentrations of up to 99 wt% can be obtained. To reduce the
size of the extractive distillation equipment, nitric acid of lower than azeotropic
concentration may be distilled up to the azeotrope in a pretreatment step.
In one type of nitric acid concentration process, sulfuric acid, of up to 85
wt% typically, and the nitric acid to be concentrated are supplied to a distilla-
tion column, the sulfuric acid higher up than the nitric acid. The bottom of the
distillation column is heated either by live steam or via a steam-heated reboiler.
22 NITRIC ACID

Indirect heating permits more dilute sulfuric acid to be used. The bottom stream is
weak sulfuric acid, which must be reconcentrated in a separate process step. The
strong nitric vapors leaving the top of the column are condensed and a portion
returned to the column as reflux. The condensed strong nitric acid has a brown
color due to dissolved nitrogen oxides formed from the decomposition of part of
the nitric acid. Unless they can be tolerated in the final strong acid product, the
nitrogen oxides are stripped from the acid with air in a bleaching column. The
stripped nitrogen oxides are recovered from the air as subazeotropic nitric acid by
absorption in water in a further column. The final product, strong nitric acid, has
a concentration of up to 99 wt%.
When magnesium nitrate is used as the third component, magnesium
nitrate solution of between 60 and 80 wt% is mixed with the weak nitric
acid before being fed to the distillation column. The bottom product is 40–65
wt% magnesium nitrate solution, which is reconcentrated in a single-stage,
steam-heated vacuum flash drum before being recycled to the column with
fresh weak nitric acid. The vapor from the vacuum evaporator contains some
nitric acid, which is recovered by absorption in water. For the strong acid vapors
leaving the distillation column the process resembles the sulfuric acid process,
with condensation, reflux, bleaching, and recovery of nitrogen oxides from the
bleaching air by absorption in water (29,129).
The ease of concentrating aqueous magnesium nitrate solution and the use of
less-expensive materials of construction make magnesium nitrate NAC technol-
ogy less capital intensive than the sulfuric acid method. Sulfuric acid processes
are usually preferred when reconcentration of the sulfuric acid is not required,
that is, when the dilute sulfuric can be used to make another product.
4.3. Direct Strong Nitric Acid Processes. DSN processes use many of
the production steps described for weak acid production. Ammonia and air are
combusted over a platinum alloy catalyst, and the process gas stream is cooled
to the point at which weak acid condensate is formed. Tail gases are reheated
by hot process gas and provide energy recovery in a gas expander. DSN pro-
cesses differ from weak acid plants in the additional processing steps required
to achieve super-azeotropic strengths of nitric acid. The combustion of ammonia
and cooling of process gas are often carried out at atmospheric pressure. Excess
water can then be directly removed as process condensate containing only 1–3
wt% nitric acid. Processes using medium- or high-pressure ammonia combustion
have a 45–50 wt% nitric condensate, requiring either a rectification step to remove
excess water from the plant or the coproduction of weak acid. The dehydrated pro-
cess gases are usually compressed, mixed with air and nitrogen peroxide obtained
from the bleaching of product acid, and then cooled to promote the full oxidation
of nitric oxide to nitrogen peroxide. The gases are sometimes contacted with weak
to azeotropic strength acid, so that the nitric oxide can react with nitric acid per
equation 16 to restore chemical equilibrium by forming the dioxide and water.
The key to producing strong acid is obtaining high partial pressures of
nitrogen peroxide. In older processes such as HOKO (130), nitrogen peroxide is
absorbed in chilled 98 wt% acid, distilled off, and condensed. The peroxide is
mixed with oxygen and weak nitric acid in an autoclave at 5 MPa to produce
strong acid. Processes such as SABAR generate strong acid by azeotropic recti-
fication (131). Nitrogen peroxide in the process gas is chemically absorbed into
NITRIC ACID 23

azeotropic or greater strength acid. Air stripping and rectification of the result-
ing acid stream produces strong acid of 98–99 wt% strength. Various process
schemes (132–134) fall between these two examples, some of which coproduce
both strong and weak acids. Flow schemes for many of the DSN processes, as
well as NAC plants, are available (135). DSN plants are uncommon nowadays.
The last DSN plant in the United States stopped production in 2012 and was
replaced by a weak nitric acid plant and an extractive distillation unit (136).
4.4. Other Processes. Most other routes for making nitric acid involve
the formation of nitrogen oxide directly from the air using various energy sources,
such as electric arcs, fossil fuels, and shock waves. Air heated to extremely high
temperatures forms nitric oxide (ca 8 vol% at 4000 K), which must then be rapidly
quenched to avoid decomposition of the nitric oxide on cooling. The electric arc
process developed at the beginning of the twentieth century is described above.
In the 1960s it was proposed to carry out the electric arc heating at elevated pres-
sure, rapidly cooling the air and nitric oxide mixture gas by expansion through a
nozzle followed by electrical power production in a magnetohydrodynamic gener-
ator. Shock tube experiments in the 1950s yielded promisingly high nitric oxide
concentrations of 7% due to the very rapid cooling, but the high energy input made
the process uneconomic (137).
The Wisconsin process, developed during World War II, heated air to
>2000∘ C by contact with a bed of fuel-heated magnesia pebbles. Rapid quench-
ing by contact with a second cold bed yielded 1–2 wt% nitric oxide. Later, a
method using three beds of silica gel in series was developed to recover the
nitric oxide. The first bed dehumidified the gas, while the second catalytically
oxidized the nitric oxide to nitrogen peroxide, and the third adsorbed the
nitrogen dioxide, which then could be stripped in a highly concentrated form
(138,139). A 40 ton/day demonstration plant making 55–60 wt% nitric acid was
built in Lawrence, Kansas, in 1952 and operated until 1954. About 2400 tons
were produced in total, but the process did not prove its economic viability
(137,140).
Alternative approaches to nitric oxide formation include irradiation of
air in nuclear reactors (141) and the oxidation of ammonia to nitric oxide in
a fuel cell generating electricity (142). Neither method appears to have been
further developed in the recent past. Current research centers on nitrogen
fixation in nonthermal plasmas, which are cooler and less energy intensive than
thermal plasmas, such as those generated in the electric arc furnaces of the
past century. Nitrogen fixation via the Haber–Bosch ammonia process has a
theoretical minimum energy consumption 2.5 times greater than the route via
nonthermal plasma, making the latter a potentially very interesting alternative
in the long term (143,144). Promising results have been obtained in a test
reactor (145).

5. Materials of Construction

Early nitric acid plants were severely restricted in design by the available
materials of construction. Until high chromium content iron was developed
24 NITRIC ACID

in the 1920s, commonly used construction materials were acid brick, earth-
enware, and Glass (qv). As better materials of construction came along, nitric
acid process design could be improved. Higher operating pressures resulted
in lower plant capital costs. The following is an overview of the materials of
construction used in modern processes. More detailed information concerning
the application and corrosion resistance of specific materials can be found
(146,147).
5.1. Weak Acid. Stainless steels are the standard material of construc-
tion for most of the piping and equipment in weak nitric acid plants. Different
grades are used, depending on the specific demands placed upon them. Stabilized
austenitic stainless, steels such as 1.4541/UNS S 32100, are the most common
materials for gaseous media in which condensation of nitric acid is not expected.
Such media are the tail gas, the process air and ammonia. For the latter two, the
use of cheaper carbon steel would be unacceptable due to the risk of formation of
rust and contamination of the noble metal ammonia oxidation catalyst.
Where low-temperature, noncondensing nitric is handled, as in the absorber,
low carbon austenitic steels such as 1.4306/UNS S30403 are a common choice.
More severe conditions of nitric acid condensation in hot gas streams, such in
gas cooler condensers, demand stainless steels with higher chromium content; a
typical alloy for this duty is 1.4335/UNS S31002. Pitting corrosion caused by high
levels of chloride in the cooling water can be a problem, especially for UNS S30403.
A good alternative for the absorber cooling coils is 1.4362/UNS S32304, a duplex
stainless steel with good resistance to chloride pitting and nitric acid at low tem-
peratures. Even more highly alloyed stainless steels such as 1.4563/UNS N08028
or 1.4477/UNS S32906 are available for the more severe process conditions in
cooler condensers with high chloride levels in the cooling water (148,149).
Both the microstructure and the composition are important for corrosion
resistance. The intergranular corrosion or “Huey” test according to ASTM
A262-Practice C is therefore used to evaluate candidate steels for nitric acid
service. It involves boiling heat-treated specimens in nitric acid and measuring
the weight loss.
Although a duplex stainless steel such as UNS S32304 is more expensive
than UNS S30403, its higher strength and better corrosion resistance make it a
cost-saving candidate for large equipment items such as absorbers (150). Because
of their lower nickel and higher chromium content than typical austenitic steels,
they possess the ductability of austenitic stainless and the stress–corrosion crack-
ing resistance of ferritic stainless steel.
Ammonia burner catalyst baskets and gauze supports must be resistant
to oxidation, nitriding, and distortion from high temperatures. Typical mate-
rials of construction are high strength alloys made of iron–nickel–chromium,
nickel–chromium, and nickel–chromium–tungsten–molybdenum.
Zirconium is used in nitric acid service for especially severe conditions in
cooler condensers, tail gas preheaters, and reboilers. The excellent corrosion resis-
tance of zirconium is a result of the formation of a strongly adhering inert oxide
film. The film is diffusion bonded to the metal and has the ability to repair itself,
should it be damaged. Zirconium rivals tantalum in its corrosion resistance to
nitric acid (151), having low corrosion rates at all concentrations up to the boiling
point. Its resistance extends up to 230∘ C and 65 wt%. It is susceptible to stress
NITRIC ACID 25

corrosion cracking, particularly at concentrations above 70 wt%, which can be


prevented by avoiding sustained high tensile stresses (152).
5.2. Strong Nitric Acid. Materials of construction commonly used in the
production of strong acid (98–99 wt%) are aluminum, tantalum, borosilicate glass,
glass-lined steel, high silica cast iron, and high silica stainless steels. Stainless
steel is often used for the storage and shipment of up to 95 wt% nitric acid. At
higher concentrations, corrosion rates for stainless steel become excessive, and
aluminum is the commonly accepted material of construction. The use of alu-
minum alloys is restricted to high acid concentrations of 80–100 wt% and mod-
erate temperatures of less than 38∘ C. High silica stainless steels (ca 4–7 wt%
silicon) have demonstrated good corrosion resistance to concentrated nitric acid.
However, corrosion resistance does not significantly extend to weaker strengths
of acid. These steels are commonly used in direct strong acid processes and appar-
ently played a key role in their development. High silica cast iron also displays
good corrosion resistance to strong acid but is more commonly used in columns
and pump casings.
Titanium is resistant to nitric acid from 65 to 90 wt% and dilute acid below
10 wt%. It is subject to stress–corrosion cracking for concentrations above 90 wt%
and, because of the potential for a pyrophoric reaction, is not used in red fuming
acid service. Tantalum exhibits good corrosion resistance to nitric acid over a wide
range of concentrations and temperatures. It is expensive and typically not used
in conditions where other materials provide acceptable service. Tantalum is most
commonly used in applications where the nitric acid is close to or above its normal
boiling point.
Glass offers good resistance to strong acid at high temperatures. However,
it is subject to thermal shock and a gradual loss in integrity as materials such as
iron and silica are leached out into the acid. Nonmetallic materials such as PTFE
and PVDF can be used for nitric acid to a limited degree but are mainly restricted
to weak acid service at ambient to moderate temperatures.

6. Economic Aspects

World nitric acid production in 2016 was about 62 × 106 t (153). The merchant
portion is under 10% of the total.
About 80% of world nitric acid production is converted to nitrates, pri-
marily ammonium nitrate and calcium ammonium nitrate. Some 35% (154), or
13.5 × 106 t HNO3 /a (estimated from Ref. 155, p. 89), of the ammonium nitrate
is employed in explosives, mainly for mining. The rest of the ammonium nitrate
and all of the calcium ammonium nitrate are used as fertilizers. Of the other 20%
of global nitric acid production, about 13% is consumed in the manufacture of the
organic chemicals nitrobenzene, adipic acid, TDI, and nitrochlorobenzenes (1).
Europe, the Soviet Union, the United States, and China dominate nitric
acid use statistics, together accounting for 80% of world consumption. Since 2000,
world production has seen steady growth at an average compound annual growth
rate (CAGR) of 2% (1). Future global market growth rates are variously estimated
at between 1.8% (156) and over 3.3% CAGR (157).
26 NITRIC ACID

7. Specifications and Standards

7.1. Shipment. The United States Department of Transportation (DOT)


classifies nitric acid as a hazardous material requiring proper packaging, labeling,
and shipping documentation for transportation. The permitted forms of packag-
ing for bulk or nonbulk shipping as well as the authorized modes of transport
depend on the DOT categories, of which there are five. Four categories are non-
fuming nitric acid of various concentrations, CHNO3 in wt%: (1) CHNO3 ≤ 20%; (2)
20% < CHNO3 ≤ 65%; (3) 65% < CHNO3 ≤ 70%; and (4) 70% < CHNO3 ≤ 100%. For
unambiguous identification, these are allocated the internationally recognized ID
number UN 2301. The fifth category is red fuming nitric acid, UN 2302.
All categories are in Hazard Class 8, meaning corrosive material, and must
be labeled “corrosive.” Red fuming nitric acid must also be labeled “oxidizer” and
“poison.” Depending on the category, nitric acid may be shipped in either stainless
steel or aluminum. Bulk nitric acid can be shipped by railcar, tank truck, portable
tank, or seagoing vessel. Nonbulk packaging includes drums. For other than red
fuming acid, a variety of smaller containers are permitted, all of which require
various forms of individual packaging, depending on the acid strength and mode
of transportation. Packages are mostly glass, earthenware, or fluorinated plastic
containers ranging in size from 0.5 to 2.5 L. Transportation on passenger aircraft
or railcar is forbidden for all categories of nitric acid except for containers up to
1.0 L and concentrations up to 20 wt%. Red fuming acid is forbidden on cargo
aircraft. The full DOT requirements for packaging and shipping nitric acid are
available (158).
Increasingly, the Globally Harmonized System of Classification and
Labelling of Chemicals (GHS) is being adopted around the world, including
in the United States. In the European Union, the GHS has been adopted as
the Chemical Labelling and Packaging Regulation (CLP). Details of the CLP
requirements for nitric acid are available (159). The International Maritime Dan-
gerous Goods Code contains specific regulations for international sea transport
(160). The European fertilizer industry has published a guideline on the bulk
transportation of nitric acid (161).
7.2. Standards. The products of Ostwald nitric acid plants or NAC plants
are suitable for most uses without further treatment. Some applications demand
high and reproducible levels of purity, however. These include semiconductor
and photovoltaic cell manufacture, the production of pharmaceuticals, chemical
analysis, and laboratory work in general. The American Chemical Society (ACS)
defines three grades of reagent acid (162): “nitric acid,” having a concentration
of 68.0–70.0 wt% HNO3 , “nitric acid, 90%,” having a concentration of ≥90 wt%
HNO3 , and “nitric acid, ultratrace,” with a concentration in the range 65–70
wt% HNO3 and maximum permitted concentrations of many metals in the low
single-digit ppb range. The first two grades have maximum allowable levels of
chlorides, sulfates, arsenic, heavy metals, iron, and residue after burning in the
low ppm or subppm range, the limits for the 90% grade, except for the residue,
being less stringent. The 90% grade has, in addition, a maximum content for
dissolved oxides as N2 O3 . The ultratrace grade is intended for use in analyses
of trace metal content, and the other grades for more general use. Standards for
high-purity nitric acid for use in the semiconductor and photovoltaic industries
NITRIC ACID 27

are available from the umbrella organization SEMI (163,164). Repeated distil-
lation is employed to purify nitric acid. For the highest purity grades, used in
trace element analysis, the method of choice is subcooled distillation, in which an
infrared radiation source is used to heat the acid to a temperature just below the
boiling point, thus avoiding the formation of bursting bubbles and aerosols which
could spread contaminants through the equipment (165).

8. Analytical and Test Methods

8.1. Qualitative Analysis. Nitric acid may be detected by the classical


brown-ring test, the copper-turnings test, the reduction of nitrate to ammonia by
active metal or alloy, or the nitrogen precipitation test. Nitrous acid or nitrites
interfere with most of these tests, but such interference may be eliminated by
acidifying with sulfuric acid, adding ammonium sulfate crystals, and evaporating
to a low volume.
In the brown-ring test, concentrated sulfuric acid is carefully poured down
the side of an inclined test tube to form a separate layer beneath the solution to be
tested. After the two layers have been cooled without mixing, a few drops of a fer-
rous sulfate solution are placed inside the inclined tube and drained down to the
interface. The appearance within a few minutes of a brown layer of an iron nitrosyl
complex, [Fe(H2 O)5 (NO)]2+ , at the junction of the two liquids shows the presence
of nitrates. Nitric acid or nitrates may be detected after the removal of nitrous
acid by warming a mixture of the material to be tested, a 1:1 solution of sulfuric
acid, and copper turnings in a test tube. The evolution of brown fumes indicates
the presence of nitrates. Nitrates are easily reduced by active metals or alloys in
alkaline solution to give ammonia. Suitable reducing agents are aluminum, zinc,
or Devarda’s alloy (50 wt% Cu, 45 wt% Al, and 5 wt% Zn).
Nitric acid may be precipitated by nitron [2218-94-2] (4,5-dihydro-2,4-diph-
enyl-5-4,5-dihydro-2,4-diphenyl-5-phenylimino-1H-1,2,4-triazolium inner salt).
The yellow precipitate may be seen at dilutions as low as 1:60,000 at 25∘ C or
1:80,000 at 0∘ C. To prevent nitrous acid from interfering with the test results, it
may be removed by treating the solution with hydrazine sulfate, sodium azide, or
sulfamic acid.
8.2. Quantitative Analysis. The total acidity of nitric acid solution may
be determined by conventional titration using phenolphthalein as the indicator.
Other Acidic Impurities.
Sulfuric Acid. The sample is evaporated to dryness on a steam bath. The
residue is removed with water, and evaporation is repeated until the sample is
free from nitric acid fumes. The residue is diluted with water and titrated with
standardized sodium hydroxide solution using phenolphthalein as an indicator.
Hydrochloric Acid. Hydrochloric acid is determined gravimetrically as sil-
ver chloride.
Nitrous Acid. Lower oxides of nitrogen and nitrous acid generally are
determined and reported as NO2 . In the absence of organic matter or other
reducing agents, the determination may be made by titrating with potassium
permanganate solution. A sharp end point is achieved by rapidly adding the
permanganate solution to the solution containing the oxide of nitrogen or nitrous
28 NITRIC ACID

acid. If the addition of permanganate solution is slow, some oxidation of the


sample by dissolved air occurs, resulting from the dilution of the solution with
water.
Other Methods. Ion chromatography using conductance detection can be
used to measure low (<1%) levels of nitrite, chloride, sulfate, and other ions in
nitric acid. Techniques for ion chromatographic analysis are available (166). Trace
heavy metals may be determined by inductively coupled plasma–optical emission
spectroscopy (ICP–OES).
Ferrous Sulfate Titration. For determination of nitric acid in mixed
acid or for nitrates that are free from interferences, ferrous sulfate titration,
the nitrometer method, and Devarda’s method give excellent results. The deter-
mination of nitric acid and nitrates in mixed acid is based on the oxidation of
ferrous sulfate [7720-78-7] by nitric acid and may be subject to interference by
other materials that reduce nitric acid or oxidize ferrous sulfate. Small amounts
of sodium chloride, potassium bromide, or potassium iodide may be tolerated
without serious interference, as can nitrous acid up to 50% of the total amount of
nitric acid present. Strong oxidizing agents, for example, chlorates, iodates, and
bromates, interfere by oxidizing the standardized ferrous sulfate.
Possible interferences and variation of results from modified techniques can
be avoided by titrating the sample in exactly the same way and by employing
approximately the same amounts of materials as in the initial standardization
of the ferrous sulfate against a known quantity of nitric acid. The ferrous sul-
fate solution is added in a thin stream until the initially yellowish solution turns
brown. The titration is complete when the faint brownish-tinged end point is
reached.
Nitrometer Method. The nitrometer method is also used to determine
nitric acid or nitrates in mixed acid or oleum. It involves the measurement of the
volume of NO gas that is liberated when mercury is oxidized by nitric acid. The
method is based on the following reaction:

2 HNO3 + 3 H2 SO4 + 3 Hg → 3 HgSO4 + 4 H2 O + 2 NO (18)

Devarda’s Method. Nitrogen in nitrates or nitric acid also may be deter-


mined by the Kjeldahl method or by Devarda’s method. The latter is both conve-
nient and accurate when no organic nitrogen is present. The nitrate is reduced by
Devarda’s alloy to ammonia in an alkaline solution. The ammonia is distilled and
titrated with standard acid.

9. Health and Safety Factors

Nitric acid and the oxides of nitrogen found in its fumes are highly toxic and capa-
ble of causing severe injury and death. In assessing the hazard due to nitric acid
vapor, the likelihood of nitrogen peroxide also being present should not be forgot-
ten, as NO2 is even more toxic than nitric acid.
Nitric acid is corrosive and can destroy human tissue. The US EPA lists
nitric acid as a toxic substance, making it subject to a broad set of laws and reg-
ulations designed to protect the public and the environment. It is also classed as
NITRIC ACID 29

an Extremely Hazardous Substance in any concentration above 1 wt%, requir-


ing facilities that hold more than the threshold quantity of 1000 lb to comply
with emergency planning requirements (167). OSHA categorizes nitric acid with
a concentration of 94.5 wt% or above as a highly hazardous chemical, mandat-
ing process safety management at facilities with more than the 500 lb threshold
inventory (168). Nitric acid is also classed as an air contaminant by OSHA. Per-
mitted exposure limits are derived from medical data, for example (169), and
subject to periodic review. The current (2019) US permitted occupational exposure
limits for nitric acid in the air are 5 mg/m3 (∼2 ppmv) for an 8-h time-weighted
average (TWA) and 10 mg/m3 (∼4 ppmv) for a 15-min short-term exposure (STEL),
the m3 volume being at 0.101325 MPa abs and 25∘ C (170). The NIOSH immedi-
ately dangerous to life or health (IDLH) concentration of nitric acid is 25 ppmv. At
or above the IDLH concentration, a worker must wear a highly reliable breathing
apparatus. For judging the effect and the actions required in the case of a release
of nitric acid vapor which affects the general public, the acute exposure guide-
line levels (AEGL) apply. There are three levels of severity: AEGL-1, no lasting
or disabling adverse effects; AEGL-2, irreversible, or serious long-lasting effects;
AEGL-3, life-threatening effects or death. Concentration limits are tabulated for
five exposure periods. For the shortest period, 10 min, the nitric acid vapor concen-
trations are AEGL-1 0.16 ppmv, AEGL-2 43 ppmv, and AEGL-3 170 ppmv (171).
In the European Union, the recommended STEL is 1 ppmv (2.6 mg/m3 ). A TWA
recommendation does not exist. It is the member states’ responsibility to imple-
ment exposure limits in national law. Many such as the United Kingdom, France,
Germany, Italy, and Spain follow the EU STEL recommendation. The Netherlands
is more restrictive with 0.5 ppmv (1.3 mg/m3 ), and some other states have higher
STELs.
Symptoms of nitric acid vapor inhalation may take up to 48 h to appear
and can prove fatal. They include irritation of the throat and nose, intractable
coughing, chest pain, difficulty in breathing, giddiness, nausea, ulceration of the
nasal mucous membranes, pulmonary edema, and chemical pneumonia. After a
recovery that can last several weeks, a relapse may occur, resulting in death by
pulmonary fibrosis or bronchopneumonia. Asthmatic or allergic persons are the
most sensitive to low concentrations of nitric acid fumes.
The symptoms resulting from skin contact vary from moderate irritation to
severe burn, depending on the contact time and the strength of the nitric acid.
Signs of contact may include a yellow discoloration of the skin; severe burns may
penetrate deeply, causing ulceration and the scarring of tissue. Contact with the
eyes can cause serious damage and result in blindness.
First aid practices for the treatment of exposure to nitric acid should be
obtained from a current version of the Safety Data Sheet or other appropriate
safety literature. Acid in contact with the skin must be removed immediately
by washing with large amounts of a proprietary liquid such as DIPHOTERINE
(Prevor, France), or failing that with water. Continuous flushing with large quan-
tities of liquid is required for acid burns to the eye; the eyelids should be gently
lifted to ensure adequate washing. For cases of ingestion, a poison control center
should be contacted for advice. A conscious and alert victim who is not convulsing
should drink several glasses of water for dilution and follow with lime milk or
milk of magnesia. Vomiting should not be induced, nor should attempts be made
30 NITRIC ACID

to neutralize the acid with sodium bicarbonate. If exposure occurs by inhalation,


the person should be moved to fresh air and given support in breathing as needed.
In all cases, it is important to seek medical attention immediately.

10. Future Trends

Within the next 5–10 years it is to be expected that all nitric acid plants worldwide
will be equipped with effective NOx and nitrous oxide emission reduction systems.
Emission limits will continue to become more stringent, and the focus of regula-
tory authorities will turn increasingly to reducing emissions of N2 O and especially
of visible NOx during start-up, shutdown, and malfunction, as is already the case
in the United States.
The future seems bright for nitric acid manufacture. As the world’s popula-
tion expands from its current 7.6 billion (end of 2019) to 9.7 billion in 2050, the
demand for fertilizers and for plastics requiring nitric acid for their production
will rise.
With the heightening interest in green, sustainable, climate-friendly tech-
nologies and the decarbonization of fertilizer production, especially in Europe
(172), the demand for nitrate-based fertilizers will increase. In contrast to urea,
nitrate fertilizers deliver nutrient that is immediately available to plants, and
they carry no carbon dioxide burden. Urea also has a problem with ammonia
emissions to the environment. “Carbon-free” ammonia will increasingly be
manufactured from hydrogen obtained from the electrolysis of water instead of
natural gas; a rising number of demonstration and even commercial projects
are currently being realized (173,174). As the main use of the ammonia is likely
to be as an energy carrier, considerable quantities of by-product oxygen from
the water electrolysis will become available for the manufacture of nitric acid.
The use of oxygen in the nitric acid process can increase the capacity of an
existing plant or permit a reduction in the capital cost of a new plant. Further-
more, the ready availability of oxygen will give stimulus to the development of
novel nitric acid plant flowsheets with reduced capital cost and higher energy
efficiency (175).
Finally, ongoing nonprecious metal catalyst developments may end the
century-long reign of platinum/rhodium as the catalyst material of choice for
ammonia oxidation. New optimum operating conditions may then result in new
process designs and more efficient nitric acid production with long periods of
uninterrupted operation, as in ammonia plants.

BIBLIOGRAPHY
“Nitric Acid” in ECT 3rd ed., Vol. 15, pp. 853–871, by D. J. Newman, Barnard and Burk,
Inc.; in ECT 4th ed., Vol. 17, pp. 80–107, by S. I. Clarke and W. J. Mazzafro, Air Products and
Chemicals, Inc.; “Nitric Acid” in ECT (online), posting date: 4 December 2000, by Stephen I.
Clarke and William J. Mazzafro, Air Products and Chemicals, Inc.; Published online: 13
May 2005, updated by Wiley Staff.
NITRIC ACID 31

CITED PUBLICATIONS
1. IHS Markit. https://ihsmarkit.com/products/nitric-acid-chemical-economics-hand
book.html (2017). (accessed 22 September 2019).
2. V. Karpenko, Bull. Hist. Chem. 34, 105–116 (2009).
3. W. A. Hamor, in F. Rolt-Wheeler, ed. The Science-History of the Universe, 1st ed.,
New York, The Current Literature Publishing Company, 1909.
4. J. N. Pring, in E. Thorpe, ed. The Electric Furnace – Monographs on Industrial Chem-
istry, London, Longmans, Green and Co., 1921.
5. B. Waeser, Die Luftstickstoff-Industrie mit Berücksichtigung der chilenischen Indus-
trie und des Kokereistickstoffs (The Atmospheric Nitrogen Industry taking into account
the Chilean Industry and Coke Oven Nitrogen), Otto Spamer, Leipzig, 1932.
6. B. Waeser, Die Luftstickstoffindustrie (The Atmospheric Nitrogen Industry), 1st ed.,
Otto Spamer, Leipzig, 1922.
7. T. H. Chilton, Chem. Eng. Progr. Monogr. Ser. 56, 1–21 (1960).
8. P. Osswal and K. H. Schäfer, in W. Foerst, ed. Ullmanns Encyklopädie der technischen
Chemie (Ullmann’s Encyclopedia of Industrial Chemistry), 3rd ed., Munich, Urban &
Schwarzenberg, 1964, pp. 3–40.
9. U.S. Pat. 1,706,055 (1929), C. W. Davis (to E.I. Du Pont De Nemours & Co.).
10. H. Holzmann, Platin. Met. Rev. 13, 2–8 (1969).
11. G. D. Honti, in C. Keleti, ed. Nitric Acid and Fertilizer Nitrates, New York, Marcel
Dekker, Inc., 1985, pp. 61–98.
12. U.S. Pat. 6,089,051 (2000), M. Gorywoda, M. Hörmann, G. Lindenmayer, D. F. Lupton
and B. Streb (to Heraeus, W.C.).
13. B. T. Horner, Platin. Met. Rev. 37, 76–85 (1993).
14. U.S. Pat. 6,073,467 (2000), S. Blass, H. Dübler, D. Königs, T. Stoll and H. Voss (to
Degussa AG).
15. T. Boublík and K. Kuchynka, Collect. Czech Chem. C 25, 579–582 (1960). DOI:
10.1135/cccc19600579
16. R. M. Counce and co-workers, Extractive Distillation of Nitric Acid Using the Two-Pot
Concept, Report No.: ORNL/TM-7982, Oak Ridge National Laboratory, Oak Ridge,
1982.
17. S. R. M. Ellis and J. M. Thwaites, J. Appl. Chem. 7, 152–160 (1957). DOI:
10.1002/jctb.5010070402
18. W. R. Forsythe and W. F. Giauque, J. Am. Chem. Soc. 64, 48–61 (1942). DOI:
10.1021/ja01253a014
19. K. D. Beyer and A. R. Hansen, J. Phys. Chem. A 106, 10275–10284 (2002). DOI:
10.1021/jp025535o
20. S. L. Clegg and P. Brimblecombe, J. Phys. Chem. 94, 5369–5380 (1990). DOI:
10.1021/j100376a038
21. R. Vandoni and M. Laudy, J. Phys. Chim. 49, 99–102 (1952).
22. G. Aunis, J. Phys. Chim. 49, 103–108 (1952).
23. D. W. Green and M. Z. Southard, eds., Perry’s Chemical Engineers’ Handbook, 9th ed.,
McGraw-Hill Book Co., Inc., New York, 2018.
24. T. R. Bump and W. L. Sibbitt, Ind. Eng. Chem. 47, 1665–1670 (1955). DOI:
10.1021/ie50548a055
25. C. McKinley and G. G. Brown, Chem. Met. Eng. 49, 142–144 (1942).
26. B. E. Thomson, J. J. Derby, and E. H. Stalzer, Vapor–Liquid Equilibrium of
the Mg(NO3 )2 –HNO3 –H2 O System, Report No.: ORNL/MIT-360, Massachusetts
Institute of Technology, Oak Ridge, TN, 1983.
27. S. Takeshi and co-workers, Kagaku Kogaku Ronbunshu 11, 267–271 (1985). DOI:
10.1252/kakoronbunshu.11.267
32 NITRIC ACID

28. R. T. Jubin, J. L. Marley, and R. M. Counce, J. Chem. Eng. Data 31, 86–88 (1986).
DOI: 10.1021/je00043a025
29. J. G. Sloan, in W. F. Furter, ed. Thermodynamic Behavior of Electrolytes in Mixed Sol-
vents, Advances in Chemistry, Vol. 155, Washington, DC, American Chemical Society,
1976, pp. 128–142.
30. R. J. Sr. Lewis, Dangerous Properties of Industrial Materials, 11th ed., John Wiley &
Sons, Inc., Hoboken, NJ, 2004.
31. L. Bretherick, Handbook of Reactive Chemical Hazards, Butterworths Publishers,
Stoneham, MA, 1990.
32. L. F. Albright and C. Hanson, eds., Industrial and Laboratory Nitrations, 22nd ed.,
American Chemical Society, Washington, DC, 1976.
33. Anon, Nitrogen and Syngas 315 (2012).
34. M. Groves, World Fertilizer, 57–60, July–August (2017).
35. U.S. Pat. 4,400,891 (1983), G. Kongshaug (to Norsk Hydro).
36. H. Bloch and C. Soares, Turboexpanders and Process Applications, Gulf Professional
Publishing, Woburn, MA, 2001.
37. W. Hanggeli, Turbomach. Int., March (1982).
38. M. Singh, M. J. Drosjack, and D. H. Linden, Expanders for Oil and Gas Operation,
1st ed., McGraw-Hill Education, New York, 2014.
39. J. S. Cichowski, Chem. Eng., 149–150, March (1982).
40. WO Pat. 2011/054928 (2011), H. J. Kneuper, J. T. Nickel, B. Menning, T. M. Scholbrock,
H. Joost, M. Holzbecher and D. Baumann (to BASF SE).
41. EP Pat. 0004746 (1979), P. G. Blakey and B. K. Smith (to BOC Ltd.).
42. U.S. Pat. 4,183,906 (1980), R. W. Watson and P. G. Blakey (to BOC Ltd.).
43. U.S. Pat. 6,165,435 (2000), D. F. Echegaray, A. A. Velloso and M. L. Wagner (to Praxair
Technology, Inc.).
44. W. Bachleitner, L. Popovici and C. Boldizar, ‘Oxyboost technology’ – increasing effi-
ciency of nitric acid plants through introduction of oxygen. Proceedings of 30th Inter-
national Nitrogen and Syngas Conference, London, pp. 1–8, 2017.
45. D. D. Wagman and co-workers, J. Phys. Chem. Ref. Data 11, 2-1–2-392 (1982).
46. Anon, Nitrogen and Methanol 245, 38–47 (2000).
47. M. Schwefer, R. Maurer and M. Groves, Reduction of nitrous oxide emissions from
nitric acid plants. Nitrogen 2000 International Conference, Vienna, pp. 61–81, 2000.
48. Environment & Safety – Orica Kooragang Island, 2019. https://www.orica.com/
ArticleDocuments/485/Information%20for%20the%20local%20community%20-
%20booklet.pdf (accessed 19 October 2019).
49. Anon, Nitrogen and Methanol 241 (1999). September–October
50. G. Biausque and Y. Schuurman, J. Catal. 276, 306–313 (2010). DOI: 10.1016/j.jcat.
2010.09.022
51. V. A. Sadykov and co-workers, Appl. Catal. A-Gen. 204, 59–87 (2000). DOI:
10.1016/S0926-860X(00)00506-8
52. U.S. Pat. 10,414,654 (2019), M. Schwefer, R. Siefert, K. Ruthardt, A. Cremona and E.
Vogna (to thyssenkrupp Industrial Solutions).
53. K. Kermeli, E. Worrell, W. Graus, and M. Corsten, Energy Efficiency and Cost Sav-
ing Opportunities for Ammonia and Nitrogenous Fertilizer Production, Energy Star®
Guide, U.S. Environmental Protection Agency, 2017.
54. Anon, Nitrogen and Syngas 344, 18–21 (2016).
55. H. Frankland and co-workers, Johnson Matthey Tech. 61, 183–189 (2017). DOI:
10.1595/205651317X695640
56. U.S. Pat. 9,056,307 (2015), T. Keller, U. Jantsch and G. J. Lambert (to Heraeus Mate-
rials Technology).
57. U.S. Pat. 10,258,966 (2019), D. Born, D. Königs and C. Goerens (to Umicore AG).
58. H. Connor, Platin. Met. Rev. 11, 60–69 (1967).
NITRIC ACID 33

59. O. Henkes, Contaminants and their effect on catalytic gauze performance in nitric
acid plants. Proceedings of 32nd International Conference Nitrogen and Syngas,
Berlin, pp. 1–9, 2019.
60. V. I. Chernyshev and S. V. Zjuzin, Platin. Met. Rev. 45, 34–40 (2001).
61. A. Bazhenov and G. Gushin, Nitrogen and Syngas 349, 1–3 (2017).
62. L. du Chatelier, Nitrogen and Methanol, , March–April (1996).
63. R. Imbihl and co-workers, Phys. Chem. Chem. Phys. 9, 3522–3540 (2007). DOI:
10.1039/B700866J
64. R. Kraehnert and M. Baerns, Chem. Eng. J. 137, 361–375 (2008). DOI: 10.1016/j.cej.
2007.05.005
65. Y. F. Zeng and R. Imbihl, J. Catal. 261, 129–136 (2009). DOI: 10.1016/j.jcat.2008.
05.032
66. J. D. Gonzalez and co-workers, Proc. Combust. Inst. 36, 637–644 (2017). DOI:
10.1016/j.proci.2016.04.004
67. H. Ma and W. F. Schneider, ACS Catal. 9, 2407–2414 (2019). DOI: 10.1021/acscatal.
8b04251
68. J. Pérez-Ramírez and co-workers, J. Catal. 261, 217–223 (2009). DOI: 10.1016/j.jcat.
2008.11.018
69. M. Warner, The Kinetics of Industrial Ammonia Combustion. PhD Thesis. University
of Sydney, 2013.
70. T. Heydt, Untersuchung der Pt-katalysierten NH3 -Oxidation unter operando Bedin-
gungen und CFD-Simulation (Investigation of Pt-catalyzed NH3 Oxidation under
Operating Conditions and CFD Simulation). PhD Thesis. Technical University of
Darmstadt, 2018.
71. A. F. Eulefi, Study of Nitrous Oxide Production in the Nitric Acid Process. PhD Thesis.
University of Alberta, 2018.
72. A. F. Hollemann, Lehrbuch der anorganischen Chemie/Hollemann-Wiberg, 33rd ed.,
Walter de Gruyter & Co., Berlin, 1985.
73. U.S. Pat. 858,904 (1907), W. Ostwald.
74. H. Connor, Platin. Met. Rev. 11, 2–9 (1967).
75. Y. C. Choi and E. Weiland, World Fertilizer Magazine, 49–56, July–August (2018).
76. S. Ubben, Nitrogen and Syngas 357, 50–55 (2019).
77. IUPAC Task Group on Atmospheric Chemical Kinetic Data Evaluation, 2017. http://
iupac.pole-ether.fr/htdocs/supp_info/Guide_to_Gas-Phase_Datasheets_Final_Oct_
2017.pdf (accessed 22 September 2019).
78. J. B. Joshi, V. V. Mahajani, and V. A. Juvekar, Chem. Eng. Commun. 33, 1–92 (1985).
79. P. J. Hoftyzer and F. J. G. Kwanten, in G. Nonhebel, ed. Gas Purification Processes for
Air Pollution Control, London, Newnes-Butterworths, 1972, pp. 164–187.
80. J. J. Carberry, Chem. Eng. Sci. 9, 189–194 (1959). DOI: 10.1016/0009-2509(59)85001-6
81. S. P. S. Andrew and D. Hanson, Chem. Eng. Sci. 14, 105–113 (1961). DOI:
10.1016/0009-2509(61)85060-4
82. J. B. Lefers, Absorption of Nitrogen Oxides into Diluted and Concentrated Nitric Acid.
PhD Thesis. TH Delft, 1980.
83. D. N. Miller, AIChe J. 33, 1351–1358 (1987). DOI: 10.1002/aic.690330812
84. N. D. Ingale, I. B. Chatterjee, and J. B. Joshi, Chem. Eng. J. 155, 851–858 (2009). DOI:
10.1016/j.cej.2009.09.017
85. M. Koukolik and J. Marek, Mathematical model of HNO3 oxidation-absorption equip-
ment. Proceedings of 4th European Symposium on Chemical Reaction Engineering,
Brussels, pp. 347–359, 1968.
86. W. I. Atroschtschenko and S. I. Kargin, Die Technologie der Salpetersäure (The Tech-
nology of Nitric Acid), 1st ed., Staatsverlag für wissenschaftlich-technische Literatur
der Chemie ‘Goschimisdat’, Moscow, 1962.
34 NITRIC ACID

87. H. Holma and J. Sohlo, Comput. Chem. Eng. 3, 135–141 (1979). DOI: 10.1016/0098-
1354(79)80024-1
88. G. Emig, K. Wohlfahrt, and U. Hoffmann, Comput. Chem. Eng. 3, 143–150 (1979).
DOI: 10.1016/0098-1354(79)80025-3
89. W. Wiegand, E. Scheibler, and M. Thiemann, Chem. Eng. Technol. 13, 289–297 (1990).
DOI: 10.1002/ceat.270130139
90. N. J. Suchak and J. B. Joshi, AIChE J. 40, 944–956 (1994). DOI: 10.1002/aic.690400605
91. B. Hüpen and E. Y. Kenig, Chem. Eng. Sci. 60, 6462–6471 (2005). DOI:
10.1016/j.ces.2005.04.060
92. E. Kenig and P. Seferlis, Chem. Eng. Prog., 65–73, January (2009).
93. I. L. S. B. Vilarinho and co-workers, Comp. Aid. Ch. 38, 1479–1484 (2016). DOI:
10.1016/B978-0-444-63428-3.50251-4
94. K. J. Mundo, in K. Winnacker and L. Küchler, eds. Chemische Technologie, Munich,
Carl Hanser Verlag, 1969, pp. 595–712.
95. U.S. Pat. 4,372,935 (1981), R. Botton and D. Cosserat (to Rhône-Poulenc Industries).
96. D. W. Bolme and A. Horton, Chem. Eng. Prog., 95–98, March (1979).
97. GB Pat. 1,587,456 (1981), D. W. Bolme.
98. A. Fareid, L. Hjørnevik and G. Kongshaug, Boosting existing nitric acid production.
Proceedings No. 246 of the The Fertiliser Society, London, pp. 1–20, 1986.
99. European Pat. 0808,797 (2003), M. Wagner (to Praxair Technology Inc.).
100. Integrated Pollution Prevention and Control Reference Document on Best Available
Techniques for the Manufacture of Large Volume Inorganic Chemicals – Ammonia,
Acids and Fertilisers, European Commission, 2007.
101. K. Heller, B. Depro, S. Norris, and T. Lentz, Economic Impact Analysis for the Nitric
Acid Manufacturing NSPS, Report No.: EPA-452/R-12-004, U.S. Environmental Pro-
tection Agency, Research Triangle Park, NC, 2012.
102. D. Thomas and J. Vanderschuren, Chem. Eng. Sci. 51, 2649–2654 (1996). DOI:
10.1016/0009-2509(96)00131-5
103. USP Technologies, Technical Bulletin, Nitrogen Oxides (NOx) Abatement with
Hydrogen Peroxide (2015). http://www.h2o2.com/files/NOx-Tech-Bulletin-15-HR.pdf
(accessed 28 July 2019).
104. O. Ghriss and co-workers, Atmos. Pollut. Res. 10, 180–186 (2019). DOI: 10.1016/j.apr.
2018.07.007
105. A. Chacuk and co-workers, Chem. Eng. Sci. 62, 7446–7453 (2007). DOI:
10.1016/j.ces.2007.08.023
106. K. Skalska, J. S. Miller, and S. Ledakowicz, Chem. Pap. 62, 193–197 (2011). DOI:
10.2478/s11696-010-0082-y
107. (German) Federal Ministry for Environment, Nature Conservation and Nuclear
Safety, TA Luft: First General Administrative Regulation Pertaining the Fed-
eral Immission Control Act, 2002. https://www.bmu.de/fileadmin/Daten_BMU/
Download_PDF/Luft/taluft_engl.pdf (accessed 29 July 2019).
108. U.S. Environmental Protection Agency, Standards of Performance for Nitric Acid
Plants for Which Construction, Reconstruction, or Modification Commenced After
October 14, 2011, U.S. Code of Federal Regulations, 40 CFR 60.72a, US EPA, 2012.
109. U.S. Environmental Protection Agency, Federal Register, 157, 48433-48448, 2012.
110. U.S. Environmental Protection Agency, Available And Emerging Technologies For
Reducing Greenhouse Gas Emissions From The Nitric Acid Production Industry, US
EPA, Research Triangle Park, NC, 2010.
111. G. Busca and co-workers, Appl. Catal. B-Environ. 18, 1–36 (1998). DOI:
10.1016/S0926-3373(98)00040-X
112. R. M. Heck, Catal. Today 53, 519–523 (1999). DOI: 10.1016/S0920-5861(99)00139-X
113. R. Q. Long and R. T. Yang, J. Catal. 207, 224–231 (2002). DOI: 10.1006/jcat.2002.3528
NITRIC ACID 35

114. M. Jouannic, G. Blanchard and J. Boisne, Most recent industrial experience and the
development of the Rhône-Poulenc DeNOx process. Proceedings of Asia Nitrogen
International Conference, Bali, Indonesia, pp. 139–152, 1994.
115. Y. H. Lee, M. H. Kim and S. W. Ham, The formation of N2 O from NH3 -SCR reac-
tion over commercial V2 O5 /TiO2 -based catalysts. Proceedings of 21st North American
Catalysis Society Meeting, San Francisco, pp. 438–439, 2009.
116. U.S. Pat. 4,973,457 (1990), G. Kongshaug, L. Hjørnevik and N. Øystein (to Norsk
Hydro).
117. BASF, Catalysts, 2012. http://catalysts.motumdev.com/files/pdf/BF-9834_US_N2O_
and_DeNOX_Technote.pdf (accessed 19 October 2019).
118. W. Boll, Prolonged campaign length of high pressure nitric acid plant by optimiza-
tion of secondary catalyst design for N2 O abatement. Proceedings of International
Nitrogen and Syngas Conference, London, 2017.
119. K. M. Åbø, Investigation of the Yara 58-Y1 nitrous oxide decomposition catalyst. MSc
Thesis. Norwegian Institute of Science and Technology, Trondheim, 2014.
120. M. Jabłońska and co-workers, RSC Adv. 9, 3979–3986 (2019). DOI: 10.1039/
c8ra10509j
121. M. Mauvezin and co-workers, Catal. Lett. 62, 41–44 (1999). DOI: 10.1023/A:
1019078401694
122. J. Pérez-Ramírez and co-workers, Catal. Commun. 3, 19–23 (2002). DOI:
10.1016/S1566-7367(01)00072-3
123. K. Jíša and co-workers, J. Catal. 262, 27–34 (2009). DOI: 10.1016/j.jcat.2008.11.025
124. M. Kögel and co-workers, Catal. Commun. 2, 273–276 (2001). DOI: 10.1016/S1566-
7367(01)00046-2
125. M. C. E. Groves and A. Sasonow, J. Integr. Environ. Sci. 7, 211–222 (2010). DOI:
10.1080/19438151003621334
126. France Pat. 2,789,911 (2000), G. Delahay, M. Mauvezin, B. Coq and N. Bernard (to
Grande Paroisse SA).
127. U.S. Pat. 7,393,512 (2008), M. Schwefer and M. Groves (to Uhde GmbH).
128. UNFCCC. https://cdm.unfccc.int/Projects/projsearch.html (accessed 31 July 2019).
129. H. Winterbauer, Chem. Eng. Technol. 28 (2005). DOI: 10.1002/ceat.200500026
130. G. Fauser, Chem. Met. Eng. 39, 430 (1932).
131. L. Hellmer, Chem. Eng. Prog. 68, 67–71 (1972).
132. T. Ohrul, M. Okubu, and O. Imai, Hydrocarbon Proc. 57, 163 (1978).
133. L. M. Marzo and J. M. Marzo, Chem. Eng. 87, 54–55 (1980).
134. R. J. Hanbury, Chem. Eng., 50–51, December (1972).
135. G. D. Honti, in C. Keleti, ed. Nitric Acid and Fertilizer Nitrates, New York, 129, Marcel
Dekker, Inc., 1985, p. 99.
136. LBS Industries Inc., Reports for Q3 2012. http://investors.lsbindustries.com/
news-releases/news-release-details/lsb-industries-inc-reports-results-2012-third-
quarter-and (accessed 27 August 2019).
®
137. F. Maslan, Process for thermal fixation of atmospheric nitrogen. The Space Congress
Proceedings, pp. 1–13, 1969.
138. N. Gilbert and F. Daniels, Ind. Eng. Chem. 40, 1719 (1948). DOI: 10.1021/ie50465a026
139. E. G. Foster and F. Daniels, Ind. Eng. Chem. 43, 986–992 (1951). DOI: 10.1021/
ie50496a054
140. F. D. Miles, Nitric Acid, Manufacture and Uses, Oxford University Press, London,
1961.
141. P. Harteck and S. Dondes, Science 146, 30–35 (1964). DOI: 10.1126/science.146.
3640.30
142. C. T. Sigal and C. G. Vayenas, Solid State Ion. 5, 567–570 (1981). DOI:
10.1016/0167-2738(81)90318-0
36 NITRIC ACID

143. W. Wang and co-workers, ChemSusChem 10, 2145–2157 (2017). DOI: 10.1002/
cssc.201700095
144. P. Peng and co-workers, ChemSusChem 12, 3702–3712 (2019). DOI: 10.1002/cssc.
201901211
145. B. S. Patil and co-workers, AIChE J. 64, 526–537 (2017). DOI: 10.1002/aic.15922
146. J. Eimers, in C. Keleti, ed. Nitric Acid and Fertilizer Nitrates, New York, Marcel
Dekker, Inc., 1985, pp. 149–157.
147. C. M. Schillmoller, Selection and Use of Stainless Steels and Nickel-Bearing Alloys in
Nitric Acid; NiDI Technical Series No. 10 075, Nickel Development Institute, 1995.
148. Anon, Nitrogen and Methanol 239, 55–60 (1999).
149. D. Gullberg, Nitrogen and Syngas 327, 51–55 (2014).
150. Anon, Nitrogen 203, May–June (1993).
151. R. D. Cooks, in B. J. Moniz and W. I. Pollock, eds. Process Industries Corrosion – The
Theory and Practice, Houston, TX, National Association of Corrosion Engineers, 1986,
pp. 259–263.
152. T. L. Yau, Corrosion 39, 167–174 (1983). DOI: 10.5006/1.3580832
153. Trammo, Inc. https://www.trammo.com/nitric-acid (accessed 18 January 2019).
154. IHS Markit. https://ihsmarkit.com/products/ammonium-nitrate-chemical-economics-
handbook.html (accessed 22 September 2019).
155. Yara International, Yara Fertilizer Industry Handbook 2018 (2018). https://www
.yara.com/siteassets/investors/057-reports-and-presentations/other/2018/fertilizer-
industry-handbook-2018-with-notes.pdf/ (accessed 22 September 2019).
156. Grand View Research, Inc. https://www.grandviewresearch.com/industry-analysis/
nitric-acid-market (accessed 22 September 2019).
157. Global Info Research. https://www.globalinforesearch.com/global-nitric-acid-
market_p171438.html (accessed 22 September 2019).
158. U.S. Dept. of Trade (DOT), Transportation, U.S. Code of Federal Regulations, 49 CFR
Parts 172, 173 and 176, US DOT, 2018.
159. ECHA – European Chemical Agency. https://echa.europa.eu/information-on-
chemicals/cl-inventory-database/-/discli/details/75872 (accessed 27 August 2019).
160. International Maritime Organization, International Maritime Dangerous Goods Code
(IMDG Code), IMO, 2018.
161. Fertilizers Europe, Guidance for Transporting Nitric Acid in Tanks, 2014. http://
productstewardship.eu/fileadmin/user_upload/user_upload_prodstew/documents/
Transporting_Nitric_Acid_in_Tanks.pdf (accessed 14 April 2019).
162. ACS, in T. Tyner and J. Francis, eds. ACS Reagent Chemicals, Online Edition ed.,
Washington, DC, ACS Publications, 2017 https://pubs.acs.org/doi/book/10.1021/
acsreagents (accessed 4 June 2019).
163. Semiconductor Equipment and Materials International (SEMI), SEMI C35-0118 –
Specification and Guide for Nitric Acid, SEMI, Milpitas, CA, 2018.
164. Semiconductor Equipment and Materials International (SEMI), SEMI PV16-0316 –
Specification for Nitric Acid Used in Photovoltaic Applications, SEMI, Milpitas, CA,
2016.
165. U.S. States Environmental Protection Agency, Inorganic Analytes, Revision 6, Decem-
ber, 2018.
166. W. R. Jones and P. Jandik, Am. Lab., June (1990).
167. U.S. Environmental Protection Agency, Emergency Planning and Notification, U.S.
Code of Federal Regulations, 40 CFR Ch. 1, Part 355, US EPA, 2018.
168. U.S. Occupational Safety and Health Administration, Process Safety Management of
Highly Hazardous Chemicals, U.S. Code of Federal Regulations, 29 CFR 1910.119,
US OSHA, 2018.
NITRIC ACID 37

169. National Research Council of the National Academies, Acute Exposure Guideline Lev-
els for Selected Airborne Chemicals, The National Academies Press, Washington, DC,
2014, pp. 139–175.
170. Centers for Disease Control and Prevention – NIOSH. https://www.cdc.gov/niosh/
idlh/7697372.html (accessed 27 August 2019).
171. U.S. Environmental Protection Agency. https://www.epa.gov/aegl/nitric-acid-results-
aegl-program (accessed 27 August 2019).
172. Fertilizers Europe, Feeding Life 2030 Report – The European Fertilizer Industry at
the Crossroads between Nutrition and Energy, Fertilizers Europe, Brussels, 2018.
173. Demonstration Project Archives – Ammonia Energy Association. https://www
.ammoniaenergy.org/category/demonstration-project/ (accessed 4 October 2019).
174. Commercial Technologies Archives – Ammonia Energy Association. https://www
.ammoniaenergy.org/category/commercial-technologies/ (accessed 4 October 2019).
175. Germany Pat. 10,2017,119,471 (2019), J. Dammeier and R. Schallert (to thyssenkrupp
AG).

MICHAEL C. E. GROVES
ThyssenKrupp Industrial Solutions AG,
Dortmund, Germany

You might also like